You are on page 1of 7

Research Article

pubs.acs.org/acscatalysis

Controlling the Product Syngas H2:CO Ratio through Pulsed-Bias


Electrochemical Reduction of CO2 on Copper
Bijandra Kumar, Joseph P. Brian, Veerendra Atla, Sudesh Kumari, Kari A. Bertram, Robert T. White,
and Joshua M. Spurgeon*
Conn Center for Renewable Energy Research, University of Louisville, Louisville, Kentucky 40292, United States
*
S Supporting Information

ABSTRACT: The electrochemical reduction of CO2 is a


promising method for sustainable, carbon-neutral chemical syn-
thesis as well as the storage of intermittent renewable energy in the
form of energy-dense fuels compatible with existing infrastructure.
In this work, we investigated a pulsed-bias technique for CO2
reduction on Cu, which led to a major shift in the product
selectivity relative to potentiostatic electrolysis conditions. With
applied voltage pulses in the millisecond time regime, syngas (CO
+ H2) became the only product and had a pulse-time-dependent
H2:CO molar ratio, ranging from ∼32:1 to 9:16 for pulse times
between 10 and 80 ms, respectively, at the same applied working
potential. X-ray photoelectron spectroscopy (XPS) and X-ray diffraction (XRD) data suggested that in situ oxidation and
reduction of the Cu partially caused the preference for CO formation over other carbon products on polycrystalline Cu.
Significant nonfaradaic current arising from electrical double layer charging and discharging was also suspected to contribute to
the desorption of key reaction intermediates and further promote CO. The results provide an electronic technique for the
electrochemical production of a controllable syngas feedstock for utilization in numerous industrial applications (e.g., Fischer−
Tropsch process and hydroformylation of alkenes to aldehydes).
KEYWORDS: copper, catalyst, CO2 reduction, syngas, selectivity, electrochemistry

1. INTRODUCTION oatomic carbon nanomaterials.10 Among these catalysts,


As societal energy demand continues to increase, the burning of relatively inexpensive and earth abundant copper (Cu) metal
fossil fuels to provide a majority of this energy will emit has the unusual property of producing hydrocarbons from
greenhouse gas CO2 at a rate that will put progressively greater CO2RR, and has therefore been widely studied for this
strain on the global environment.1 Among various emerging reaction.11 However, unlike Ag7b,12 and Au,7c,d,f Cu has
CO2 mitigation technologies2 (e.g., sequestration,3 chemical relatively poor CO2 reduction product selectivity due to the
conversion,4 and electrochemical reduction), the electro- multiple competing reaction pathways it can promote as well as
chemical CO2 reduction reaction (CO2RR) has advantages as a sensitivity to crystal orientation leading to variability from the
a means to sustainably utilize this waste stream to produce heterogeneity of active sites present on the polycrystalline
valuable chemicals and fuels.5 For instance, electrochemical surface.7e,13 As many as 16 products (e.g., methane, ethylene,
conversion is directly controllable by adjustment of the applied formic acid, carbon monoxide, etc.) have been identified from
electrode potential, and an electrochemical system can operate the electrochemical reduction of CO2 on a polycrystalline Cu
near room temperature, have a relatively small footprint, film.7e Consequently, numerous efforts have been pursued to
require minimal chemical intake, and be easily scaled up. With enhance the selectivity of polycrystalline Cu by fine-tuning the
the electricity supplied by intermittent renewable sources (i.e., structure,14 size,15 morphology,16 and composition,13a but
solar and wind), CO2 reduction is also an ideal method to store controlling the CO2RR reaction pathway remains a key
energy as chemical bonds in the form of energy-dense challenge and significantly varies with respect to the applied
molecules (e.g., syngas, ethanol, methane, etc.) for clean potential. Here, we introduce a pulsed-bias electrochemical
power utilities and transportation fuels.6 Because of the stability reduction process for the conversion of CO2 on polycrystalline
of the CO2 molecule, the cathodic reduction overpotential is Cu catalyst, displaying high selectivity for CO formation over
significant, and thus highly active and selective electrocatalysts other carbon products. A few previous studies have employed a
are necessary to efficiently promote CO2RR over hydrogen square wave potential with 3−5 s pulses on Cu in an attempt to
evolution.
Numerous heterogeneous catalysts have been employed for Received: March 23, 2016
CO2 electrochemical reduction, including transition metals7 Revised: June 5, 2016
and transition metal oxides,8 metal complexes,9 and heter- Published: June 8, 2016

© 2016 American Chemical Society 4739 DOI: 10.1021/acscatal.6b00857


ACS Catal. 2016, 6, 4739−4745
ACS Catalysis Research Article

Figure 1. CO2 reduction performance of Cu catalyst without pulsing. (a) Schematic diagram of customized electrochemical cell. (b) Current density
vs potential (J-E) curve for Cu in CO2-saturated 0.1 M KHCO3 solution. (c) Chronoamperometry (CA) analysis at different applied potentials (vs
RHE) and (d) the corresponding measured faradaic efficiency (F.E.) for each product.

prevent catalyst poisoning,17 but the product distribution was separated the cathode and anode compartments to prevent
only moderately affected, permitting a modest enhancement in oxidation of the reduction products at the anode. The cathode
the faradaic efficiency of CH4 and C2H4 at high cathodic compartment contained a gas bubbler at the bottom, a port to
potentials. These results are significantly different than position the Ag/AgCl (3.0 M KCl) reference electrode in the
investigated herein using pulses in the millisecond time regime center, and a wide window for the catalyst working electrode.
in which transient phenomena were observed to be increasingly The CO2 (99.99%, Specialty Gases) was injected into the
dominant and the CO2RR product selectivity was completely catholyte at a flow rate of 20 sccm using a calibrated mass flow
different. Interestingly, the proposed method also offers a novel controller (MKS Instruments). The anode compartment
route for tuning the ratio of H2 and CO (i.e., syngas) at a given contained a Pt mesh counter electrode and was also actively
applied potential by varying the pulse time. The selectivity bubbled with CO2 during measurements to maintain a uniform
toward syngas is attributed to the pulse-induced oxidation/ cell pH of 6.8. All CO2 reduction experiments were carried out
reduction of the Cu surface and the constantly shifting electrical in CO2-saturated 0.1 M KHCO3 (99.99%, Sigma-Aldrich)
double layer, which leads to CO formation over other carbon aqueous solution at room temperature (298 K). A potentiostat
products (e.g., methane or ethylene). The ability to translate (Biologic SP-200) with electrochemical impedance spectrosco-
CO2 into a tunable syngas (H2+CO) composition, which can py (EIS) was used for all measurements. The potentials were
be used in a variety of different subsequent catalytic recorded using an Ag/AgCl (3.0 M KCl) reference electrode
transformation processes,4b,18 could facilitate the industrializa- and reported on a reversible hydrogen electrode (RHE) scale
tion of the CO2 electrochemical reduction process catalyzed by according to VRHE = VAg/AgCl + 0.210 + 0.059*pHsoln.19 The
an earth abundant material. solution pH was measured to be 6.8. Potentiostatic EIS
measurements were performed before every CO2RR experi-
2. EXPERIMENTAL SECTION ment to determine the uncompensated solution resistance, Ru,
2.A. Electrochemical Measurements. All experiments and the potentiostat subsequently compensated for 85% of Ru
used polycrystalline Cu foil (99.99%, Alfa Aesar) as working during electrolysis. The current densities were determined
electrodes. Prior to electrochemical experiments, the Cu foil relative to geometric projected catalyst area throughout this
was chemically cleaned with 1.0 M HCl solution to remove the study.
surface oxide and adsorbed impurities. Cu samples were then 2.B. Product Analysis. CO2 reduction products were
rinsed with isopropyl alcohol and DI water. CO2 electro- measured by gas chromatography (GC, SRI 8610) and nuclear
reduction studies were performed using a custom-made, magnetic resonance (NMR, Bruker 400 MHz) for the gas and
gastight, two-compartment polycarbonate electrochemical cell liquid products, respectively. Both instruments were calibrated
(Figure 1a). The cell was designed with a large active with standard gases or liquid solutions. For real-time gas phase
electrocatalyst area (∼5 cm2) relative to a minimal electrolyte product analysis, the outlet of the cathode compartment was
volume (7 mL catholyte), with well-dispersed, uniform CO2 connected to the GC inlet. The GC system used an automatic
reactant to maximize carbon product concentration and valve injection (1 mL sample) and a thermal conductivity
faradaic efficiency, which helps in the detection of gaseous detector (TCD) and flame ionization detector (FID). Ultrahigh
and liquid products with high sensitivity as demonstrated purity nitrogen (99.99%, Specialty Gases) was used as the
previously.7e An anion exchange membrane (Selemion AMV) carrier gas for all experiments and was chosen to enable
4740 DOI: 10.1021/acscatal.6b00857
ACS Catal. 2016, 6, 4739−4745
ACS Catalysis Research Article

accurate hydrogen quantification. The first gas sample was results on Cu foil catalyst in aqueous electrolyte within the
injected after 5 min of potentiostatic CO2 reduction, followed limited cathodic potential range.7e,20
by three additional GC measurements at 18 min intervals. In 3.B. CO2 Reduction with Pulsed Bias. Figure 2a
addition, NMR spectroscopy was used to analyze and quantify generically illustrates the potential waveform applied during
the products in the liquid phase by performing 1H NMR
experiments. Samples were prepared by mixing D2O and
electrolyte aliquots in a 1:1 volume ratio. Dimethyl sulfoxide
(DMSO) was added at a known low concentration for internal
calibration. Faradaic efficiency (F.E.) was calculated for the
chronoamperometric measurements without pulsed bias by
determining the charge required to produce the measured
product concentration and dividing by the total charge passed
during the time the sample underwent electrolysis. To ensure a
steady state concentration in the output stream during product
quantification, the first GC measurement (at 5 min) was
excluded, and the following three GC measurements (at 23, 41,
and 59 min) were used to determine an average faradaic
efficiency.
C. Catalyst Characterization. A scanning electron micro-
scope (FEI NOVA nano-SEM 600) was used to study the
surface morphology of the Cu foil. The catalyst surface
properties before and after measurement were characterized
with X-ray photoelectron spectroscopy (XPS) with a VG
Scientific MultiLab 3000 custom-built UHV system (with Cu−
Kα radiation). X-ray diffraction analysis was done by using HR Figure 2. CO2 reduction performance of Cu catalyst with pulsing. (a)
XRD Bruker D8 instrument. Illustrative pulse in which the applied potential was alternated between
an anodic rest potential, Ea (0.6 V vs RHE), for a time period of ta, and
3. RESULTS the cathodic working potential, Ec, for a time period of tc. (b−d)
Reduction products in percentage (%) of charge required at different
3.A. CO2 Reduction without Pulsed Bias. The chemical applied potentials (Ec), (b) −1.05 V, (c) −1.0 V, and (d) −0.95 V vs
surface cleaning step used before every experiment was a simple RHE, as a function of pulse time, tc. For (b−d), ta = tc for each
and effective method to remove the oxide layer and other measurement, and CA denotes chronoamperometric conditions at the
impurities present on the Cu surface (Figures S1 and S2). SEM cathodic working potential without pulsed bias.
analysis of cleaned Cu films showed featureless surface
morphology and a lack of physically adsorbed particles. Further
elemental analysis by XPS also confirmed the absence of any the pulsed-bias electrolysis experiments. In this method, the
metal impurities within the XPS detection limit (Figure S2). applied bias was alternated between an anodic rest potential, Ea,
Figure 1b shows the current density vs potential (J-E) at +0.61 V vs RHE (0.0 V vs Ag/AgCl), for a time ta, and a
polarization curve for CO2RR obtained by sweeping the cathodic working potential, Ec, negative of the CO2 onset
potential between +0.6 and −1.2 V vs RHE at a scan rate of 20 potential (−0.63 V vs RHE) for a time tc. Other than the
mV s−1. A similar polarization curve was obtained in Ar- applied potential pulse mode, experimental conditions were
saturated electrolyte, in which case Cu yields only H2 product identical to previous experiments. Due to the discrete, step-like
(Figure S4). The observed CO2 reduction onset potential switch between Ea and Ec, sharp transient current spikes
(−0.63 V vs RHE) and measured current density (∼17 mA occurred immediately after the application of both Ea and Ec
cm−2) at −1.2 V vs RHE are consistent with the current density throughout the pulsed potential waveform (Figure S7).
reported previously for polycrystalline Cu in CO2-saturated 0.1 Deconvolution of the faradaic reaction current from the
M KHCO3 solution.7e Chronoamperometry (CA) experiments nonfaradaic transient current processes during pulsed elec-
were performed at potentials varying from −0.65 to −1.15 V vs trolysis was not straightforward and deemed too unreliable for
RHE at 50 mV intervals for 60 min at each potential (Figure accurately reporting faradaic efficiencies of products. CO2
1c). Figure 1d shows the average faradaic efficiencies (F.E.) for reduction product selectivity was therefore reported for each
all identified major products as a function of applied potential species as the charge required to create that amount of product
(shown in Table S1). At low potential (positive of −0.90 V vs as a percentage of the total charge required to create all
RHE) H2, CO, and HCOOH were the major products. In this detected products (rather than the total charge passed during
region, hydrogen evolution is kinetically dominant, and the the experiment). Figure 2b−d shows the variation in product
faradaic efficiency of H2 formation remained high, varying from selectivity at Ec of −1.05, −1.00, and −0.95 V vs RHE while
96.4% to 53.5% at −0.65 V and −0.90 V vs RHE, respectively. varying pulse time tc from 0 to 100 ms with equivalent time
At higher applied potential (negative of −0.90 V vs RHE), the periods of ta. Without pulsed bias, CH4, C2H4, and H2 were the
total current attributable to CO 2 reduction increased main products in this potential window (Figure 1d), but the
significantly, with the faradaic efficiency of CO2RR higher application of the pulsed potential strongly favored the
than hydrogen evolution in the studied potential range. direction of charge to CO and H2 in a ratio dependent on
The faradaic efficiency of H2 formation reached a minimum the pulse time. At 100 ms pulse times some hydrocarbon
of 26.3% at −1.05 V vs RHE, where CH4 and C2H4 were the products were still detected, but at pulse times <50 ms, CH4
major carbonaceous products. The observed values and trend in and C2H4 products were completely suppressed with CO as the
the product selectivity match well with previously reported only detectable carbonaceous product. Under these conditions,
4741 DOI: 10.1021/acscatal.6b00857
ACS Catal. 2016, 6, 4739−4745
ACS Catalysis Research Article

CO formation peaked at 50 ms pulse times with a ∼1:1 molar attained a maximum value of ∼64% at ta = 80 ms and remained
ratio of H2 to CO. Shorter pulse times led to H2 production consistent up through ta = 130 ms. By varying the pulse times
increasingly being favored over CO formation, reaching ∼97% for a given cathodic working potential (i.e., −1.0 V vs RHE),
H2 for 10 ms pulse periods. For all pulsed-bias conditions CO production was thus promoted relative to H2 production
tested, no liquid phase products were detected (Figure S8), and from a molar ratio of ∼1:10 under potentiostatic conditions to
these species are therefore omitted from the graphs for CO2 16:9 (CO:H2) with tc = 50 ms and ta = 80 ms. The CO and H2
reduction under pulsed bias. formation rates with varying ta pulse time are given in Figure
Figure 3a shows the product selectivity for pulsed-bias CO2 3b. As expected, as the rest pulse time ta became longer with
reduction at a constant tc (50 ms) while varying ta from 0 to respect to the working pulse time tc at sufficiently cathodic
potentials for reduction, the overall product formation rate
decreased. Although the total reduction rate decreased
modestly by ∼20% between potentiostatic and peak CO
pulsing conditions (tc = 50 ms, ta = 80 ms), CO formation
demonstrated the opposite trend and increased from ∼0.5
nmol s−1 cm−2 to ∼5.3 nmol s−1 cm−2 at the same conditions.

4. DISCUSSION
The chronoamperometric experiments without a pulsed bias
show the well-established CO2RR product distribution as a
Figure 3. Selectivity dependence on ta. (a) Product charge percentage
function of applied potential on a Cu foil electrocatalyst in
at −1.0 V vs RHE with tc = 50 ms and different ta pulse times. (b) aqueous electrolyte (Figure 1d).7e Theoretical and experimen-
Product formation rate with varying ta pulse times. tal studies have elucidated numerous reaction mechanisms and
pathways to account for the product selectivity behavior on Cu
surfaces.7e,13c,20,21 The CO2 electroreduction mechanism on Cu
140 ms at a cathodic working potential, Ec, of −1.0 V vs RHE. is a complex phenomenon with multiple proton-coupled
For a short rest pulse of ta = 10 ms, the GC measured ∼97% H2 electron-transfer steps and possible reaction pathways depend-
and ∼3% CO, while CH4 and C2H4 were out of the detection ent upon the formation of various intermediate species,
range. The introduction of this short pulse thus strongly involving both electrochemical as well as nonelectrochemical
reduced the rate of CO2 reduction. However, with longer pulses steps (e.g., adsorption, dimerization, etc.).21a,d It is, however,
of ta > 10 ms, CO as a percentage of the product increased and generally accepted that the unique ability of Cu electrocatalysts

Figure 4. Catalyst surface effects during pulsed-bias electrolysis. (a) SEM image of a Cu foil after pulsed-biased conditions at a cathodic working
potential of −1.0 V vs RHE. (b) XPS spectra on a Cu foil electrode that was pristine (Cu), tested under chronoamperometric conditions (Cu CA),
and tested under pulsed-bias conditions with 50 ms pulses (Cu 50 ms) and corresponding (c) XRD spectra. (d) Schematic diagram representing the
proposed mechanistic effects on CO2 reduction from the pulsed-bias method as compared to potentiostatic electrolysis. Under potentiostatic
conditions, bound CO* can desorb to form CO, further reduce to CH4, or react with an adjacent CO* and reduce to C2H4. Under pulsed-bias
conditions in this time regime, it is proposed that anodic current during the Ea pulse oxidizes the surface with water which is subsequently reduced
back to a nanoscopically roughened metal surface as in (a) from cathodic current flow during the Ec pulse. Adsorbed CO2 during the Ec pulse is
reduced to CO*, which is proposed to primarily desorb to CO owing to both the changed catalyst surface active sites relative to the pure
polycrystalline Cu as well as the changing electric field due to the double layer charging/discharging.

4742 DOI: 10.1021/acscatal.6b00857


ACS Catal. 2016, 6, 4739−4745
ACS Catalysis Research Article

to form significant amounts of hydrocarbon products is due to a after potentiostatic electrolysis, and a Cu foil after pulsed-bias
surface binding energy which leads to near-optimal stability of electrolysis. The additional peak present at 530 eV after pulsed-
the adsorbed CO intermediate species which is a precursor to bias electrolysis for 60 min, which was observed for each set of
the formation of higher carbon products (e.g., methane, pulse times tested and ending on either potential, corresponds
ethylene, and methanol). Thus, at increasingly reductive to CuOx (Figure S11).23 The results are further supported by
potentials, CO2RR on Cu promotes hydrocarbon products XRD analysis where characteristic peaks of CuOx are clearly
over CO formation (Figure 1d).7e visible (Figure 4c). The pulsed-bias condition is thus
As seen in Figure 2, the application of a pulsed bias rather interpreted to cause the in situ formation of CuO and Cu2O
than a potentiostatic bias led to a drastically altered CO2RR surface layers, perhaps due to anodic current flow during the Ea
product distribution at a Cu electrode. The pulses switched pulse. Subsequent electrochemical reduction of the oxide layer
between the reductive working potential and a rest potential during the Ec pulse likely leads to the surface roughening,
below the thermodynamic potential for CO2 reduction or similar to the intentional nanostructuring that arises in oxide-
hydrogen evolution. The introduction of the pulsed potential derived Cu catalysts for CO2RR.24 Furthermore, oxide-derived
waveform changed the reaction mechanism, directing more Cu as either foils24 or highly dense nanowires14 has shown high
charge at the tested working potentials to CO formation and CO2RR selectivity toward CO formation as compared to
suppressing other reduction products. In particular, for pulse standard polycrystalline Cu. The enhanced CO faradaic
durations of tc = ta ≥ 50 ms, the percent of charge passed to efficiency was attributed to the improved stabilization of the
produce H2 decreased and the CO2 reduction products CO2•− intermediate on the high surface area of oxide-derived
increased relative to the potentiostatic condition, peaking at a Cu nanowire arrays.14 It should also be noted that the presence
∼ 1:1 ratio at 50 ms pulses. This result is consistent with the of an oxide layer during the Ec pulse likely further contributes
pulsed-bias approach easing the effects of mass-transport to the CO selectivity. Copper oxide surfaces have been found in
limitations on the reaction. In potentiostatic mode, diffusion some cases to favor CO formation in comparison to other
of dissolved CO2 to the catalyst surface limits CO2RR and carbon products (e.g., CH4 and C2H4).14,24,25 For instance, Lan
convolutes the kinetic competition with hydrogen evolution, et al., reported CO as the main product on Cu(core)/
promoting H2. Pulsed mode, on the other hand, provides CuO(shell) structures even at high applied potential (−1.73 V
additional time for CO2 diffusion. During periods at the Ec vs Ag/AgCl).26 The morphological and chemical trans-
potential, CO2 reduction can proceed in kinetic competition formation of the Cu catalyst surface under the transient
with hydrogen evolution. During periods at the Ea potential conditions of pulsed-bias electrolysis thus leads to desorption of
below either thermodynamic reaction potential, neither the CO species before it can undergo further reduction to
reduction reaction would be predicted to occur, providing hydrocarbon products. Thus, while the kinetics for CO2
time for CO2 from the bulk electrolyte to diffuse to the catalyst conversion to CO are sufficiently fast for an appreciable
surface for reaction during the next Ec pulse. This type of pulse reaction rate during the millisecond pulse time regime, sluggish
voltammetry to mitigate mass-transport limitations in a reaction kinetics for further reduction of CO to CH4 could be inhibiting
has been successfully applied in other electrochemical systems, hydrocarbon formation under these pulse conditions.
for instance as a technique to enable conformal electro- On the millisecond time scale for pulses used in this work,
deposition on microstructured high-aspect ratio surfaces.22 transient current spikes after the step-change in potential were
Although the pulsed-bias effect on mass-transport limitations significant and prevented the establishment of a steady-state
explains the promotion of CO2 reduction over H2 formation, reduction current within a given pulse. This type of transient
this theory does not account for the observed mechanistic electrochemical current is commonly associated with a
change to favor CO over other carbonaceous products nor the nonfaradaic process, usually due to a capacitive charging
decrease in CO production at pulse times <50 ms. current rather than faradaic reaction current. The capacitive
Further investigation was necessary to elucidate the behavior primarily arises from the charging and discharging of
mechanistic shift in CO2RR under millisecond potential pulses. the electrical double layer, or Helmholtz layer, at the electrode
One hypothesized reason for the selective formation of CO surface.19 We speculate that the changing local electric field
under pulsed-bias conditions was that hydrocarbons formed at near the catalyst surface during the charging and discharging of
reductive conditions during the Ec pulse could be oxidized from the electrical double layer could have a significant effect on the
transient anodic currents during the Ea pulse. To examine this, binding energy experienced by adsorbed reaction intermediates.
we performed control experiments by introducing CH4 at 20 A nonideal binding energy during the transient charging current
sccm without CO2 under pulsed-bias reaction conditions. The could thus lead to desorption of the critical CO intermediate
resulting absence of any CO peak in the GC spectra suggested before further reduction to hydrocarbon species can occur. The
that CO formation was only due to CO2 reduction rather than changing Helmholtz layer effect is consistent with the near-total
the conversion of any hydrocarbons formed in situ. We cannot, directing of CO2 reduction current to CO product under
however, exclude the possibility of the oxidation of surface pulsed bias. Furthermore, this effect is consistent with the
adsorbed hydrocarbon intermediates during pulsed-bias elec- observed behavior at shorter pulse times (i.e., <50 ms), in
trolysis conditions. A possible change to the Cu catalyst surface which hydrogen evolution was increasingly promoted over CO
induced by pulsed-bias conditions could also be responsible for formation (Figure 2). At shorter pulse times, the nonfaradaic
the changed CO2RR product branching ratio. Figure 4a shows capacitive current became increasingly dominant, and the
the Cu catalyst surface after 60 min of electrolysis under pulsed- rapidly changing electrical double layer may inhibit the binding
bias conditions with a cathodic working potential of −1.0 V vs of CO2 before it can even be reduced to CO. Figure 4d
RHE. The catalyst surface displayed increased nanoscopic illustrates the changes to the mechanistic pathway of CO2RR
roughness compared to a freshly cleaned Cu foil electrode on Cu between potentiostatic and pulsed-bias electrolysis due
(Figures S1 and S10). Figure 4b shows the surface chemical to the combined effects of a surface oxide and a shifting
analysis by XPS of the O 1s spectra for pristine Cu, a Cu foil electrical double layer as proposed in this work.
4743 DOI: 10.1021/acscatal.6b00857
ACS Catal. 2016, 6, 4739−4745
ACS Catalysis Research Article

The dependence of the electrolysis product selectivity on the Cu catalysts. The pulsed waveform dependency of the H2 to
pulse time periods provides an electronic technique for tuning CO ratio permits the syngas product composition to be easily
the H2 to CO ratio at a specified working potential. The adjusted in real time by changing the pulse duration and/or
production of syngas of varying composition can therefore be cathodic working potential. This technique provides versatility
controlled without changing the applied working voltage, and for integrating electrochemical CO2 reduction into industrial
hence the overpotential, of the reaction. As Figures 2 and 3 processes for the synthesis of fuels or value-added chemicals at
demonstrate, syngas can be produced without significant a large scale.
byproducts with CO ranging from ∼3% to ∼64% by changing
the pulse times within a range of 10−80 ms. The syngas
production rate, however, decreased as the anodic rest potential
■ ASSOCIATED CONTENT
* Supporting Information
S
Ea time became a greater fraction of the overall potential The Supporting Information is available free of charge on the
waveform (Figure 3b). The decline in total product formation ACS Publications website at DOI: 10.1021/acscatal.6b00857.
can be attributed to the drop in the fraction of time (i.e., tc/(tc Additional information about the morphological, ele-
+ ta)) responsible for CO2 electroreduction. Because the mental, and electrochemical characterization results
applied potential Ea during ta is positive of the thermodynamic (PDF)


potential for both CO2 reduction and hydrogen evolution,
syngas production does not occur within this period. Pulsed-
bias electrochemical CO2 reduction thus holds promise for AUTHOR INFORMATION
industrial applications requiring syngas. Syngas can either be Corresponding Author
used directly as a fuel or as a feedstock for the synthesis of *E-mail: joshua.spurgeon@louisville.edu.
numerous, energy-dense chemicals including synthetic natural Notes
gas, liquid hydrocarbons via the Fischer−Tropsch process, and The authors declare no competing financial interest.
value-added chemicals through the hydroformylation process.18
In each syngas conversion process, a specific ratio of H2 to CO
is required in the process feed gas to optimize the product
■ ACKNOWLEDGMENTS
The authors acknowledge financial support from the Kentucky
yield. The pulsed-bias technique could be applied to correct Department for Energy Development and Independence grant
and optimize the syngas composition in real time without #PON2 127 1500002410. The authors also acknowledge
degrading the electrolysis efficiency. The pulsing method could support from the Conn Center for Renewable Energy Research
also provide versatility to a centralized CO2 reduction system at the University of Louisville.


meant to supply syngas to a variety of chemical processes. One
current commercial technology for adjusting the H2:CO syngas REFERENCES
ratio involves a system of specialized membranes with flow (1) (a) Davis, S. J.; Caldeira, K.; Matthews, H. D. Science 2010, 329,
control valves.27 The pulsed-bias electrolysis method, in 1330−1333. (b) Hall-Spencer, J. M.; Rodolfo-Metalpa, R.; Martin, S.;
comparison, has the advantage of tuning the syngas ratio at Ransome, E.; Fine, M.; Turner, S. M.; Rowley, S. J.; Tedesco, D.; Buia,
the source without the need for any capital-intensive hardware M.-C. Nature 2008, 454, 96−99. (c) Hoffert, M. I.; Caldeira, K.; Jain,
or additional energy expenditures. A. K.; Haites, E. F.; Harvey, L. D. D.; Potter, S. D.; Schlesinger, M. E.;
Schneider, S. H.; Watts, R. G.; Wigley, T. M. L.; Wuebbles, D. J.
5. CONCLUSION Nature 1998, 395, 881−884.
(2) Mikkelsen, M.; Jorgensen, M.; Krebs, F. C. Energy Environ. Sci.
Electrochemical CO 2 reduction at Cu electrodes was 2010, 3, 43−81.
characterized with full product quantification in aqueous (3) Sanna, A.; Uibu, M.; Caramanna, G.; Kuusik, R.; Maroto-Valer,
electrolyte under both potentiostatic and pulsed-bias con- M. M. Chem. Soc. Rev. 2014, 43, 8049−8080.
ditions. The introduction of a square-wave applied potential to (4) (a) Daza, Y. A.; Kent, R. A.; Yung, M. M.; Kuhn, J. N. Ind. Eng.
the Cu catalyst drastically altered the CO2RR product Chem. Res. 2014, 53, 5828−5837. (b) Xiaoding, X.; Moulijn, J. A.
selectivity, producing only H2 and CO in ratios strongly Energy Fuels 1996, 10, 305−325.
dependent on the pulse time. The product syngas H2:CO molar (5) (a) Kumar, B.; Brian, J. P.; Atla, V.; Kumari, S.; Bertram, K. A.;
ratio varied between ∼32:1 and 9:16 with working pulse times White, R. T.; Spurgeon, J. Catal. Today 2016, 270, 19−30. (b) Qiao, J.;
of 10 and 50 ms, respectively. Several complex phenomena are Liu, Y.; Hong, F.; Zhang, J. Chem. Soc. Rev. 2014, 43, 631−675.
(c) Costentin, C.; Robert, M.; Saveant, J.-M. Chem. Soc. Rev. 2013, 42,
likely interacting during the transient electrochemical process 2423−2436. (d) Jitaru, M.; Lowy, D. A.; Toma, M.; Toma, B. C.;
to create the observed behavior. Increased CO2 reduction Oniciu, L. J. Appl. Electrochem. 1997, 27, 875−889.
relative to hydrogen evolution at longer pulse times is evidence (6) (a) Kumar, B.; Llorente, M.; Froehlich, J.; Dang, T.; Sathrum, A.;
that the pulse waveform enhanced the mass-transport of Kubiak, C. P. Annu. Rev. Phys. Chem. 2012, 63, 541−569. (b) Martin,
dissolved CO2 molecules to the electrode surface. In situ A. J.; Larrazabal, G. O.; Perez-Ramirez, J. Green Chem. 2015, 17,
oxidation and reduction of the Cu surface during transient 5114−5130. (c) Kondratenko, E. V.; Mul, G.; Baltrusaitis, J.;
anodic/cathodic currents is suspected to create a changing Larrazabal, G. O.; Perez-Ramirez, J. Energy Environ. Sci. 2013, 6,
surface of nanostructured Cu oxide and oxide-derived Cu, both 3112−3135.
of which have been demonstrated to favor CO production over (7) (a) Azuma, M.; Hashimoto, K.; Hiramoto, M.; Watanabe, M.;
hydrocarbons. The transient, capacitive charging and discharg- Sakata, T. J. Electrochem. Soc. 1990, 137, 1772−1778. (b) Rosen, B. A.;
Salehi-Khojin, A.; Thorson, M. R.; Zhu, W.; Whipple, D. T.; Kenis, P.
ing of the electrical double layer at the electrode surface is also J. A.; Masel, R. I. Science 2011, 334, 643−644. (c) Feng, X.; Jiang, K.;
proposed to lead to an unstable local electric field at the catalyst Fan, S.; Kanan, M. W. J. Am. Chem. Soc. 2015, 137, 4606−4609.
surface on the millisecond scale that inhibits the binding of key (d) Zhu, W.; Michalsky, R.; Metin, Ö .; Lv, H.; Guo, S.; Wright, C. J.;
intermediate species and prevents the subsequent reduction Sun, X.; Peterson, A. A.; Sun, S. J. Am. Chem. Soc. 2013, 135, 16833−
steps that lead to hydrocarbon products. The results highlight a 16836. (e) Kuhl, K. P.; Cave, E. R.; Abram, D. N.; Jaramillo, T. F.
unique strategy for tunable syngas production on cost-effective Energy Environ. Sci. 2012, 5, 7050−7059. (f) Kauffman, D. R.; Alfonso,

4744 DOI: 10.1021/acscatal.6b00857


ACS Catal. 2016, 6, 4739−4745
ACS Catalysis Research Article

D.; Matranga, C.; Qian, H.; Jin, R. J. Am. Chem. Soc. 2012, 134,
10237−10243.
(8) Chen, Y.; Kanan, M. W. J. Am. Chem. Soc. 2012, 134, 1986−1989.
(9) Ramos Sende, J. A.; Arana, C. R.; Hernandez, L.; Potts, K. T.;
Keshevarzk, M.; Abruna, H. D. Inorg. Chem. 1995, 34, 3339−3348.
(10) (a) Kumar, B.; Asadi, M.; Pisasale, D.; Sinha-Ray, S.; Rosen, B.
A.; Haasch, R.; Abiade, J.; Yarin, A. L.; Salehi-Khojin, A. Nat. Commun.
2013, 4, 2819. (b) Liu, Y.; Chen, S.; Quan, X.; Yu, H. J. Am. Chem. Soc.
2015, 137, 11631−11636.
(11) Gattrell, M.; Gupta, N.; Co, A. J. Electroanal. Chem. 2006, 594,
1−19.
(12) Lu, Q.; Rosen, J.; Zhou, Y.; Hutchings, G. S.; Kimmel, Y. C.;
Chen, J. G.; Jiao, F. Nat. Commun. 2014, 5, 3242.
(13) (a) Guo, X.; Zhang, Y.; Deng, C.; Li, X.; Xue, Y.; Yan, Y.-M.;
Sun, K. Chem. Commun. 2015, 51, 1345−1348. (b) Nie, X.; Esopi, M.
R.; Janik, M. J.; Asthagiri, A. Angew. Chem., Int. Ed. 2013, 52, 2459−
2462. (c) Durand, W. J.; Peterson, A. A.; Studt, F.; Abild-Pedersen, F.;
Norskov, J. K. Surf. Sci. 2011, 605, 1354−1359.
(14) Raciti, D.; Livi, K. J.; Wang, C. Nano Lett. 2015, 15, 6829−6835.
(15) Reske, R.; Mistry, H.; Behafarid, F.; Roldan Cuenya, B.; Strasser,
P. J. Am. Chem. Soc. 2014, 136, 6978−6986.
(16) Sen, S.; Liu, D.; Palmore, G. T. R. ACS Catal. 2014, 4, 3091−
3095.
(17) (a) Yano, J.; Morita, T.; Shimano, K.; Nagami, Y.; Yamasaki, S. J.
Solid State Electrochem. 2007, 11, 554−557. (b) Ishimaru, S.;
Shiratsuchi, R.; Nogami, G. J. Electrochem. Soc. 2000, 147, 1864−
1867. (c) Yano, J.; Yamasaki, S. J. Appl. Electrochem. 2008, 38, 1721−
1726. (d) Nogami, G.; Itagaki, H.; Shiratsuchi, R. J. Electrochem. Soc.
1994, 141, 1138−1142.
(18) Wilhelm, D. J.; Simbeck, D. R.; Karp, A. D.; Dickenson, R. L.
Fuel Process. Technol. 2001, 71, 139−148.
(19) Bard, A. J.; Faulkner, L. R. Electrochemical Methods:
Fundamentals and Applications, 2nd ed. ed.; John Wiley & Sons: 2001.
(20) Hori, Y.; Murata, A.; Takahashi, R. J. Chem. Soc., Faraday Trans.
1 1989, 85, 2309−2326.
(21) (a) Kortlever, R.; Shen, J.; Schouten, K. J. P.; Calle-Vallejo, F.;
Koper, M. T. M. J. Phys. Chem. Lett. 2015, 6, 4073−4082. (b) Peterson,
A. A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Norskov, J. K. Energy
Environ. Sci. 2010, 3, 1311−1315. (c) Nie, X.; Esopi, M. R.; Janik, M.
J.; Asthagiri, A. Angew. Chem., Int. Ed. 2013, 52, 2459−2462.
(d) Schouten, K. J. P.; Kwon, Y.; van der Ham, C. J. M.; Qin, Z.;
Koper, M. T. M. Chem. Sci. 2011, 2, 1902−1909.
(22) Sun, J. J.; Kondo, K.; Okamura, T.; Oh, S. J.; Tomisaka, M.;
Yonemura, H.; Hoshino, M.; Takahashi, K. J. Electrochem. Soc. 2003,
150, G355−G358.
(23) (a) Liu, P.; Li, Z.; Cai, W.; Fang, M.; Luo, X. RSC Adv. 2011, 1,
847−851. (b) Diaz-Droguett, D. E.; Espinoza, R.; Fuenzalida, V. M.
Appl. Surf. Sci. 2011, 257, 4597−4602.
(24) Li, C. W.; Kanan, M. W. J. Am. Chem. Soc. 2012, 134, 7231−
7234.
(25) (a) Xie, J.; Huang, Y.; Yu, H. Front. Environ. Sci. Eng. 2015, 9,
861−866. (b) Kas, R.; Kortlever, R.; Milbrat, A.; Koper, M. T. M.;
Mul, G.; Baltrusaitis, J. Phys. Chem. Chem. Phys. 2014, 16, 12194−
12201.
(26) (a) Lan, Y.; Gai, C.; Kenis, P. J. A.; Lu, J. ChemElectroChem
2014, 1, 1577−1582. (b) Lan, Y.; Ma, S.; Lu, J.; Kenis, P. J. A. Int. J.
Electrochem. Sci. 2014, 9, 7300−7308.
(27) Generon Innovative Gas Systems. http://www.generon.com/
pdf/Syngas-H2-CO-Ratio-Adjustment-CO-Recovery.pdf (accessed
June 1, 2016).

4745 DOI: 10.1021/acscatal.6b00857


ACS Catal. 2016, 6, 4739−4745

You might also like