Supplementary Materials for - Science...2 Supplementary Methods 1. CO 2 and CO reduction on carbon...

67
www.sciencemag.org/content/360/6390/783/suppl/DC1 Supplementary Materials for CO2 electroreduction to ethylene via hydroxide-mediated copper catalysis at an abrupt interface Cao-Thang Dinh,* Thomas Burdyny,* Md Golam Kibria,* Ali Seifitokaldani,* Christine M. Gabardo, F. Pelayo García de Arquer, Amirreza Kiani, Jonathan P. Edwards, Phil De Luna, Oleksandr S. Bushuyev, Chengqin Zou, Rafael Quintero-Bermudez, Yuanjie Pang, David Sinton, Edward H. Sargent† *These authors contributed equally to this work. †Corresponding author. Email: [email protected] Published 18 May 2018, Science 360, 783 (2018) DOI: 10.1126/science.aas9100 This PDF file includes: Materials and Methods Figs. S1 to S28 Tables S1 to S13 References

Transcript of Supplementary Materials for - Science...2 Supplementary Methods 1. CO 2 and CO reduction on carbon...

  • www.sciencemag.org/content/360/6390/783/suppl/DC1

    Supplementary Materials for

    CO2 electroreduction to ethylene via hydroxide-mediated copper catalysis at an abrupt interface

    Cao-Thang Dinh,* Thomas Burdyny,* Md Golam Kibria,* Ali Seifitokaldani,* Christine M. Gabardo, F. Pelayo García de Arquer, Amirreza Kiani,

    Jonathan P. Edwards, Phil De Luna, Oleksandr S. Bushuyev, Chengqin Zou, Rafael Quintero-Bermudez, Yuanjie Pang, David Sinton, Edward H. Sargent†

    *These authors contributed equally to this work.

    †Corresponding author. Email: [email protected]

    Published 18 May 2018, Science 360, 783 (2018) DOI: 10.1126/science.aas9100

    This PDF file includes:

    Materials and Methods Figs. S1 to S28 Tables S1 to S13 References

    mailto:[email protected]

  • 2

    Supplementary Methods

    1. CO2 and CO reduction on carbon gas-diffusion layers:

    Electroreduction for the main figure results were performed in a flow-cell configuration consisting

    of a gas-diffusion layer, anion exchange membrane (Fumasep FAB-PK-130) and nickel foam (1.6

    mm thickness, MTI Corporation) anode. All experiments in Fig. 1 and 2 were deposited on top of

    the microporous side of a Freudenberg gas-diffusion layer. The combined catalyst and diffusion

    layer, anion exchange membrane and nickel anode were then positioned and clamped together

    using polytetrafluoroethylene (PTFE) spacers such that a liquid electrolyte could be introduced

    into the chambers between the anode and membrane as well as the membrane and the cathode.

    Gaseous CO2 could then be passed behind the gas-diffusion layer and diffuse into the liquid

    electrolyte present at the catalyst. In the catholyte stream a port drilled into the PTFE spacer is

    present for a Ag/AgCl reference electrode to be positioned a specific distance from the working

    electrode.

    All CO2 and CO reduction experiments were performed using an electrochemical workstation

    (Autolab PGSTAT302N) with a Ag/AgCl reference (with 3 M KCl as the filling solution). iR

    compensation losses between the working and reference electrodes were measured using

    electrochemical impedance spectroscopy (EIS) as shown in Table S4. Electrode potentials after iR

    compensation were rescaled to the reversible hydrogen electrode (RHE) reference by:

    𝐸𝑅𝐻𝐸 = 𝐸𝐴𝑔/𝐴𝑔𝐶𝑙 + 0.197 𝑉 + 0.591 𝑥 𝑝𝐻 (1)

    where the pH in Eq. (1) is determined using a reaction-diffusion model as described in Section 6

    of the Supplementary Materials and shown in Table S4 for 1, 5 and 10 M KOH. For all reported

    data the pH value used in Eq. (1) is taken at a current of 0 mA/cm2 as this represents the worst-

    case scenarios for our reported potentials. For completeness, the predicted electrode pH as a

  • 3

    function of current density up to 25 mA/cm2 is shown in Fig. S3. The geometric area is used for

    reporting all the experimental data in the report, employing the same convention for reporting FE’s

    and current densities as previous works. All potentials reported here were obtained by averaging

    over a timespan of at least 150 s for each applied current.

    The electrolytes (KOH solution of various concentrations, 100 mL) were circulated through the

    electrochemical cell using peristaltic pumps with a silicone Shore A50 tubing. The electrolyte flow

    was keep at 10 mL min-1. The CO2 (Linde, 99.99%) flow was kept constant at 50 mL min-1 using

    a mass flow controller. The reactions were run for at least 150 s before the gas products were

    collected for analysis. KOH concentrations of 15 M could not reliable be evaluated as salt

    precipitated on the gas-diffusion electrode, preventing accurate measurement of onset potentials.

    The gas products from CO2 reduction (CO, H2, CH4 and C2H4) and CO reduction (H2, CH4 and

    C2H4) were analyzed using a gas chromatograph (PerkinElmer Clarus 680) coupled with a thermal

    conductivity detector (TCD) and a flame ionization detector (FID). The gas chromatograph was

    equipped with a Molecular Sieve 5A and Carboxen-1000 column packed columns. Argon (Linde,

    99.999%) was used as the carrier gas. All Faradaic efficiencies reported were averaged from at

    least three different runs.

    The onset potentials for CO and C2H4 are the lowest potentials at which the Faradaic efficiency

    for CO and C2H4 are respectively higher than 1 ±0.3%. To measure the onset potential, the flow

    rate of CO2 and CO feedstock was reduced to 2 – 5 mL min-1 so that the GC detector signals of

    the products are at least 5 times higher than the noise signal. 13C-labelling experiments were

    performed in the similar conditions as those of 12CO2. The ethylene product was analyzed using a

    mass spectrometer (PerkinElmer Clarus SQ 8C) coupled with a HP PLOT-Q capillary column

    (Agilent). A typical chromatograph for ethylene analysis is shown in Fig. S24.

  • 4

    The liquid products were quantified using Nuclear magnetic resonance spectroscopy (NMR). 1H

    NMR spectra of freshly acquired samples were collected on Agilent DD2 500 spectrometer in 10%

    D2O using water suppression mode, with Dimethyl sulfoxide (DMSO) as an internal standard.

    Sixteen second relaxation time between the pulses was used to allow for complete proton

    relaxation.

    The CO2 reduction experiments using Cu catalysts (10 nm, 25 nm, 1000 nm, and 1000 g) were

    also performed in an H-cell configuration as controls. Running in 1 M KOH produced only H2 as

    the concentration of CO2 is less than 10-10 M. The electrolytes for both the cathode and anode in

    were then run in 0.1 KHCO3 to compare with prior conditions where peak ethylene selectivities

    were achieved. CO2 gas was bubbled through the catholyte for at least 30 min to saturate the

    electrolyte with CO2. The CO2 flow was controlled at 30 mL min-1. The reactions were performed

    under potentiostatic mode with all samples showing peak ethylene selectivities between -0.95 V

    to -1.2 V vs. RHE (after iR correction). The gas and liquid products were analyzed after 1 hour of

    reaction using gas chromatography and NMR as described above. All samples showed

    substantially reduced ethylene FE’s in the range of 30-35%, confirming that the high C2 selectivity

    and efficiency originates from the combination of reaction interface design and the strong alkaline

    electrolyte; rather than the nature of the catalyst itself.

    2. CO2 reduction on polymer-based gas-diffusion layers:

    Electrochemical experiments performed on the polymer-based gas-diffusion layer in Fig. 3 and 4

    were performed using the same potentiostatic equipment, system orientation, flow conditions and

    product analysis procedures as the carbon gas-diffusion flow-cell experiments described above.

  • 5

    For the long-term stability tests the operating potential was fixed at -1.8V vs. Ag/AgCl. The

    electrolyte used was 7 M KOH and was changed every 12 hours during the test to maintain similar

    ion concentrations and conductivity in the catholyte and anolyte flow channels. The pH of the

    catholyte slightly changed (

  • 6

    Vienna ab initio Simulation Package (VASP) (40) was used to perform all the plane wave density

    functional theory (DFT) computations. The projected augmented wave (PAW) approach (41)

    together with the generalized gradient approximation (GGA) parametrized by Perdew, Burke and

    Ernzerhof (PBE) (44) are employed. Three different crystalline facets, (111), (100) and (110) of

    the pristine copper are approximated by a 4×4×4 slab in a 20 Å vacuum. Due to the vacuum, dipole

    corrections are implemented. To resemble the real bulk material and the surface, respectively, two

    bottom layers are fixed in their positions while the two top layers are free to move due to interaction

    with the adsorbates. A cut-off energy of 400 eV for the plane wave basis sets and a 4 × 4 × 1 Γ-

    centered Monkhorst-Pack mesh for the k-points sampling in the first Brillouin zone, with a first

    order Methfessel-Paxton smearing parameter σ of 0.1 eV ensured that the energy convergence

    criteria is better than 1 meV. The k-points grid is doubled for charge density calculations. The self-

    consistent field (SCF) convergence criterion is set to 1 × 10−4 eV for electronic iteration and the

    ionic relaxation continued until the maximum force was less than 0.02 eV/Å. This was updated by

    the conjugate gradient approach.

    Almost all previous literature indicates that the CO-CO coupling is pH independent because no

    proton participates in the coupling reaction. However, we explicitly investigate the effect of OH

    on changing the electronic structure and, in turn, the coupling reaction energy. In this work, we

    explicitly considered OH in our simulations to see how CO bonding and CO-CO coupling energies

    are affected. For Cu(100) and Cu(111), in two cases, with OH ion and without OH ion, we

    calculated the zero point energy (ZPE), heat capacity, and entropy of the adsorbate and molecular

    gases based on the vibrational frequencies within quasi-harmonic approximation framework (42).

    These values were added to the DFT ground state energy to obtain the room temperature Gibbs

  • 7

    free energies. Moreover, we considered the transition state to calculate the activation energy

    barriers of the CO-CO coupling for these facets using the nudged elastic band (NEB) method (43).

    In addition, to include the solvent effects, we considered three explicit water molecules with and

    without OH ion. Previous studies show that three water molecules are enough to stabilize OH ion,

    and we considered similar configurations in our analysis (45-47). We considered OH ion in two

    forms: adsorbed on the surface and solvated within the explicit water molecules. The same

    calculations for vibrational frequencies and transition states have been done to convert all energies

    to the Gibbs free energy at room temperature and to calculate the activation energy barriers for

    CO-CO coupling.

    For the calculation without explicit water molecules, one and two adsorbed carbon monoxides are

    simulated on all three surfaces with a varying number and proximity of hydroxides. In each case,

    different rational possible configurations (overall more than 200 configurations) are considered to

    find the global minima. The CO adsorption energy is calculated as:

    𝐸𝑎𝑑𝑠𝑜𝑟𝑝𝑡𝑖𝑜𝑛 = 𝐸∗𝐶𝑂𝑛−𝑂𝐻 − (𝐸∗

    𝑛−𝑂𝐻 + 𝐸𝐶𝑂) (2)

    Where, 𝐸∗𝐶𝑂𝑛−𝑂𝐻 is the electronic structure energy of the adsorbed CO on the catalyst with n-OH

    ions on the surface, 𝐸∗𝑛−𝑂𝐻 is the energy of the slab without the CO adsorbate but includes n-OH

    ions, and 𝐸𝐶𝑂 is the CO energy in gas phase without the catalyst. The CO-CO coupling energy

    barrier is calculated according to the following reaction and the corresponding reaction free

    energy:

    𝑂𝐶𝐶𝑂 ∗ = 𝐶𝑂

    ∗ + 𝐶𝑂 ∗ (3)

    𝐸𝐶−𝐶 𝑐𝑜𝑢𝑝𝑙𝑖𝑛𝑔 𝑏𝑎𝑟𝑟𝑖𝑒𝑟𝑛−𝑂𝐻 = 𝐸 𝑂𝐶𝐶𝑂 ∗

    𝑛−𝑂𝐻 − (𝐸 𝐶𝑂 ∗𝑛−𝑂𝐻 + 𝐸 𝐶𝑂 ∗

    𝑛−𝑂𝐻) (4)

  • 8

    In equation (4), the zero point energy (ZPE), heat capacity, and entropy are considered to convert

    all energies to the Gibbs free energy. In addition, the transition state is considered via NEB method

    to calculate the activation energy barrier.

    In our simulations without explicit water molecules, 0, 1 and 2 OH are considered on a surface of

    16 (4×4) copper atoms, corresponding to 0, 1/16 and 2/16 ML concentrations, respectively.

    However, in configurations where OH is very close to the adsorbed CO, it is assumed that the OH

    concentration is even higher and forced to be very close to the adsorbates, and the local minima is

    calculated in this case. These cases are virtually described by 3/16 and 4/16 ML in Fig. S8.

    Otherwise, in 0, 1/16 and 2/16 ML concentrations, the configuration with the global minima energy

    is the reference for our calculations. This includes all possible configurations for OH ions adsorbed

    on the surface or being present in the vacuum just above the surface. The DFT energy of the

    optimized configurations with the global minimum energies are tabulated in Table S2. CO

    adsorption energies and the CO-CO coupling energy barriers are also mentioned in Table S3.

    From Fig. S8 and the data in Table S3, related to the calculations without explicit water and without

    considering the transition state, we see that CO-CO coupling energy barrier decreases with

    increasing OH concentration. It is found that increasing the OH concentration on the slab surface

    implicitly changes the charge density of the C atoms in adsorbed CO and OCCO, such that CO

    bonding becomes weaker and CO-CO coupling becomes stronger and more favorable. On the

    (100) surface, for instance, CO without any OH is adsorbed on the hollow site with four bonds to

    the surrounding copper ions. However, increasing the OH concentration pushes CO to the bridge

    site with two folds bonding and weaker binding energy. This bonding becomes even weaker as

    OH is further added to the surface, such that OH in the vicinity pushes CO to the top site with one

  • 9

    fold bonding. The same trend is observed on two other facets. Fig. S8 demonstrates this situation

    on (100).

    Table S.4 and S.5 show the calculated ZPE, heat capacity, entropy and the Gibbs free energies at

    room temperature of the initial and final state of CO-CO coupling without OH, with 1 adsorbed

    OH, and with 2 adsorbed OH on Cu(100) and Cu(111), respectively. The calculated

    thermodynamic and activation energy barriers are mentioned in Table S6 and S7 for Cu(100) and

    Cu(111), respectively. In few of the simulations, the transition state energy is less than the final

    state energy, indicating a simple uphill energy barrier. In these cases, therefore, the final energy is

    considered as the main energy barrier.

    Including three explicit water molecules, we studied the CO-CO coupling energy barrier with and

    without OH on both Cu(100) and Cu(111). OH is considered in two forms: adsorbed and solvated.

    Surprisingly, when OH is solvated in the explicit water molecules, the activation energy barrier

    for CO-CO coupling is higher than when OH is adsorbed on the surface still with presence of the

    explicit water molecules. In general, we considered the lowest energy barrier between these two

    situations as the activation energy barrier for CO-CO coupling with OH. Table S.8 and S.9 show

    the detail thermodynamic energies of the initial and final states of CO-CO coupling on Cu(100)

    and Cu(111) facets, respectively. The calculated activation energy barriers for these two facets are

    mentioned in Table S10 and S11. Fig. S9 demonstrates these calculated activation energy barriers

    with and without explicit waters on both Cu(100) and Cu(111) facet. Fig. S10 shows the initial

    state (IS), transition state (TS) and the final state (FS) of the CO-CO coupling on Cu(100) with

    explicit water molecules, without OH ion, with one OH ion adsorbed on the surface and one OH

    ion solvated in explicit water molecules.

  • 10

    In general, we conclude that at higher OH concentrations CO bonding is weaker and this could

    lead to more CO in gaseous products and easier CO-CO coupling. Indeed, both are shown under

    experimental conditions. We calculated the electronic charge density on each ion by Bader charge

    analysis (51). As demonstrated in Fig. S8, we see that increasing the OH concentration, not only

    decreases the CO-CO coupling energy barrier but also increases the charge difference between

    two C atoms in coupled carbon monoxides. This increased charge difference, makes a stronger

    intramolecular dipole and consequently more stable coupling due to the stronger ionic adsorption.

    The electronic charge distribution and also charge difference between two C atoms in coupled

    carbon monoxides, i.e. proportional to the dipole magnitude.

    5. Material synthesis:

    The carbon-based Freudenberg (Fuel cell Store) gas-diffusion layer (GDL) was used as substrate

    for the experiments in Fig. 1 and 2. Cu (99.99%) was evaporated on the substrate using an

    Angstrom Nexdep Evaporator. The deposition was performed in ~10-5-10-6 Torr at 1.5 Å/sec. The

    thicknesses of the samples was confirmed via scanning electron microscopy (SEM). The 1000 µg

    sample was prepared by drop-casting commercial Cu nanoparticles (NPs) (Sigma-Aldrich, particle

    size

  • 11

  • 12

    fluorescence spectra were collected using a 4-element Si(Li) drift detector under ambient

    atmosphere with a custom in-situ liquid cell fabricated from Teflon. CO2 gas was continuously

    bubbled into the electrolyte to ensure saturation. The pre-edge region was scanned from 8879 eV

    to 8974 eV at a rate of 5 eV/s, the near-edge region was scanned from 8974 eV to 9019 eV at a

    rate of 0.45 eV/s and the post-edge region was canned from 9019 eV to 12 k with a rate of 0.05

    k/s. The XAS data were analyzed using the software package Athena.

    The electrochemically active surface areas of the catalysts were measured using a double layer

    capacitance method in 0.1 M HClO4 electrolyte (49). Cyclic voltammetry scans were recorded in

    the potential range from -0.05 to 0.2 V vs. RHE. The roughness factors of the catalysts were

    compared to electrochemically polished Cu foil whose surface area is defined as 1.

    7. Modeling of CO2 diffusion into the liquid electrolyte:

    As described in Section 1 of the Supplementary Materials, a reaction-diffusion model is used to

    predict the pH in the vicinity of the catalyst layer which can then be used to adjust the measured

    potentials in Eq. (1). At 0 mA/cm2 this surface pH differs from the bulk KOH concentration due

    to CO2 that diffuses across the gas-liquid interface and interacts with hydroxide, lowering the local

    pH within the catalyst layer from that of the bulk KOH electrolyte. Additionally, as current is

    applied to the catalyst layer, hydroxide will be generated causing the local pH to increase from this

    minimum case and eventually surpass the bulk pH of the electrolyte. The reaction-diffusion model

    takes into account interactions between the following species in the electrolyte (CO2, OH-, HCO3

    -

    and CO32-) as well as the consumption of CO2 and the production of OH

    - in the catalyst layer. The

    following equations governing these interactions are adopted from previous works (48, 50):

    𝜕[𝐶𝑂2]

    𝜕𝑡= 𝐷𝐶𝑂2

    𝜕2[𝐶𝑂2]

    𝜕𝑥2− [𝐶𝑂2][𝑂𝐻

    −]𝑘1𝑓 + [𝐻𝐶𝑂3−]𝑘1𝑟 − 𝑅𝐶𝑂2 (5)

  • 13

    𝜕[𝐻𝐶𝑂3−]

    𝜕𝑡= 𝐷𝐻𝐶𝑂3−

    𝜕2[𝐻𝐶𝑂3−]

    𝜕𝑥2+ [𝐶𝑂2][𝑂𝐻

    −]𝑘1𝑓 − [𝐻𝐶𝑂3−]𝑘1𝑟 − [𝐻𝐶𝑂3

    −][𝑂𝐻−]𝑘2𝑓 + [𝐶𝑂32−]𝑘2𝑟 (6)

    𝜕[𝐶𝑂32−]

    𝜕𝑡= 𝐷𝐶𝑂32−

    𝜕2[𝐶𝑂32−]

    𝜕𝑥2+ [𝐻𝐶𝑂3

    −][𝑂𝐻−]𝑘2𝑓 − [𝐶𝑂32−]𝑘2𝑟 (7)

    𝜕[𝑂𝐻−]

    𝜕𝑡= 𝐷𝑂𝐻−

    𝜕2[𝑂𝐻−]

    𝜕𝑥2− [𝐶𝑂2][𝑂𝐻

    −]𝑘1𝑓 + [𝐻𝐶𝑂3−]𝑘1𝑟 − [𝐻𝐶𝑂3

    −][𝑂𝐻−]𝑘2𝑓 + [𝐶𝑂32−]𝑘2𝑟 + 𝑅𝑂𝐻 (8)

    where RCO2 and ROH account for the consumption of CO2 in the reduction reaction and the

    production of OH-, respectively. As these reactions only occur within the catalyst layer the

    following equations are spatially dependent and are assumed to occur homogeneously throughout

    the entire catalyst layer:

    𝑅𝐶𝑂2 = {

    𝑗

    𝐹(

    𝐹𝐸𝐶𝑂

    𝑛𝑒,𝐶𝑂+

    𝐹𝐸𝐶2𝐻4

    𝑛𝑒,𝐶2𝐻4+

    𝐹𝐸𝐸𝑡𝑂𝐻

    𝑛𝑒,𝐸𝑡𝑂𝐻)

    𝜀

    𝐿𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡 , 0 ≤ 𝑥 ≤ 𝐿𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡

    0 , 𝑥 > 𝐿𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡

    (9)

    𝑅𝑂𝐻 = {

    𝑗

    𝐹

    𝜀

    𝐿𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡 , 0 ≤ 𝑥 ≤ 𝐿𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡

    0 , 𝑥 > 𝐿𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡 (10)

    In calculating Eqs. (9) and (10) the geometric current density, j, and product selectivities, FECO,

    FEC2H4, FEEtOH and FEH2, are imposed. A porosity of 60%, ε, and selectivity distribution of 10%

    H2, 10% CO and 80% (EtOH+C2H4) is assumed. Panels (f), (g) and (h) in Fig. S3 assume a catalyst

    thickness, Lcatalyst, of 100 nm, which matches the experimental catalyst used in Fig. 1.

    As shown in Fig. S3e, a gas-liquid interface is assumed at x = 0 µm. The catalyst layer then extends

    into the electrolyte a specified thickness from this boundary. A liquid diffusion thickness of x =

    500 µm was assumed in our simulations as our cathode is recessed in a cavity of a similar depth

    due to the compression of a sealing gasket. The fluid outside of this cavity is assumed to be

    continuously mixed and replaced. A smaller diffusion thickness would slightly reduce the local

    reaction pH by providing a shorter diffusion distance of hydroxide to the bulk, but the bounds of

    this value depend on flow within the fluid channel. The maximum solubility of CO2 in KOH

  • 14

    electrolyte was modeled using Henry’s constant and the Sechenov equation at 1 atm and 298 K to

    account for ‘salting out’ effects (Fig. S2b) (52).

    The catalyst layer porosity of 60% is based on a loosely packed spheres model while some fuel

    cell studies report porosities of 40% (53). The diffusion coefficient inside the simulated region was

    adjusted for porosity and viscosity. The assumption of a homogeneous reaction occurring laterally

    throughout the catalyst layer (Eq. (9)) is more accurate for lower current densities and thinner

    catalyst layers.

    The following boundary conditions were used to solve the coupled equations. The left boundary

    condition of Eq. (5) was set as the solubility limit of CO2 in a specified bulk KOH concentration,

    while a no-flux boundary condition was applied for CO2 at the right-hand boundary. For OH-,

    HCO3- and CO3

    2- no-flux boundary conditions were applied at the left boundary while the

    concentrations at the right boundary were set to the equilibrium values in the specified bulk KOH

    concentration. For all simulations we prescribe current density, selectivity, bulk KOH

    concentration and catalyst layer thickness.

    From the solved system of equations the concentration profile of CO2 and OH- in the electrolyte

    can be found for a variety of inputs as shown in Fig. 1D and Fig. S3. As seen in Fig. 1D the KOH

    media is particularly parasitic to CO2 and is neutralized by OH- well before the 500 µm liquid

    diffusion boundary, with or without a catalyst layer or current density imposed. The largest

    predicted correction from the bulk pH values occurs in 1 M KOH at 0 mA/cm2 where the potential

    is adjusted using a pH of 12.42, instead of the bulk electrolyte pH (pH = 14), equivalent to a 93 mV

    difference.

  • 15

    Fig. S1. Structural characterization of the samples before and after CO2 reduction (a, b)

    SEM images of the Cu deposited on a carbon gas-diffusion electrode with a thickness of 100 nm

    before reaction. (c) Cross-sectional SEM image of the Cu sample simultaneously deposited on flat

    silicon substrates for thickness measurement. (d, e, f) SEM images of the sample after the CO2

    reduction reaction in 1, 5, and 10 M KOH electrolytes, respectively. The CO2 reduction tests were

    performed at a current density of 50 mA/cm2 for 500 seconds.

  • 16

  • 17

    Fig. S2. Validating the local surface pH by Cr catalyst. (a, b) SEM images of Cr catalyst

    deposited on carbon gas diffusion layer. (c) Comparison of HER activity of Cr catalyst in 1 M, 5

    M and 10 M KOH electrolyte. (d-f) HER activity of Cr catalyst in Ar and CO2 using 1 M, 5 M and

    10 M KOH electrolytes, respectively.

  • 18

    Fig. S3. Modelling of the gas-liquid diffusion interface with and without a catalyst layer or

    applied current. (a) Schematic of CO2 penetration across a gas-liquid interface. (b) Maximum

    CO2 solubility in various bulk KOH concentrations due to ‘salting out’ effects. (c) CO2 diffusion

    into various KOH concentrations without a catalyst layer or CO2 reduction (j = 0 mA cm-2). (d)

    hydroxide concentration within the diffusion layer thickness (500 µm) at 0 mA/cm2 for various

    KOH concentrations, (e) Schematic of CO2 penetration across a gas-liquid interface where CO2

    reduction occurs on the red catalyst, (f) predicted electrode surface pH as a function current density

    and KOH concentration for a 100 nm thick catalyst, (g) CO2 distribution throughout a 100 nm

    thick catalyst layer at a current of 250 mA cm-2. (h) predicted hydroxide concentration inside the

    catalyst layer over a broad current density and electrolyte concentration for a 100 nm thick catalyst.

  • 19

    Fig. S4. Kinetic of CO reduction to ethylene. (a) ethylene Faradaic efficiency at low applied

    potential region showing similar ethylene onset potentials for 1 M and 10 M KOH electrolyte. (b)

    Tafel plots for CO reduction to ethylene using 1 M and 10 M KOH electrolyte.

  • 20

    Fig. S5. In-situ X-ray absorption spectroscopy (XAS) study in 5 M KOH electrolyte. The Cu

    K-edge spectra of (a) the 100 nm Cu sample under open circuit potential (OCP), at -0.16 V, and at

    -0.96 V vs. RHE. (b) Reference standards of Cu metal, Cu2O, and CuO were taken to measure the

    Cu0, Cu+, and Cu2+ oxidation states, respectively.

  • 21

    Fig. S6. Transmission electron microscopy (TEM) characterization of the samples before

    and after CO2 reduction. (a, e) TEM images of the Cu deposited on a carbon gas-diffusion

    electrode with a thickness of 100 nm before reaction. TEM images of the sample after CO2

    reduction reaction in 1 M (b, f), 5 M (c, g), and 10 M (d, h) KOH electrolytes. The CO2 reduction

    tests were performed at a current density of 50 mA/cm2 for 500 seconds.

  • 22

    Fig. S7. Effect of OH on the adsorbed OCCO intermediate. Oxidation states (charge density)

    of the adsorbed OCCO elements and the adjacent copper as calculated by Bader charge analysis.

    Oxidation states on (100), (110) and (111) facets of copper, without OH, with one and two OH

    without including explicit water molecules are shown for the most stable configurations.

    Cu (100)

    No OH One OH Two OH

    Cu (110)

    Cu (111)

  • 23

    Fig. S8. Effect of OH ion on CO binding energy and CO-CO coupling on Cu (100), (110) and

    (111) facets, without including explicit water molecules. (a) CO adsorption energy at different

    OH concentrations. (b) Charge differences of two carbons in adsorbed OCCO at different OH

  • 24

    concentrations. (c, d) Electron charge density and the optimized configuration of the adsorbed CO

    with no OH and one OH on Cu(100). (e, f) Electron charge density and the optimized configuration

    of the adsorbed OCCO with no OH and one OH on Cu(100). The charge difference between two

    carbons and the induced dipole in the presence of OH is shown in (f). (g) Density Functional

    Theory (DFT) results for CO-CO coupling on Cu (100) surface with and without the presence of

    OH. (h) DFT results showing the reduction of CO-CO coupling barrier in the presence of OH.

  • 25

    Fig. S9. Energies of the initial state (IS), transition state (TS) and the final state (FS) of CO-

    CO coupling on both Cu(100) and Cu(111) facets with and without explicit water molecules,

    with different number of OH ions.

  • 26

    Fig. S10. Initial state (IS), transition state (TS) and the final state (FS) of CO-CO coupling

    on Cu(100) with three explicit water molecules. (left) CO-CO coupling without any OH ion,

    (middle) CO-CO coupling with one OH ion adsorbed on the surface, and (right) CO-CO coupling

    with one OH solvated in explicit water molecules. Brown and black spheres demonstrate copper

    and carbon, respectively. Red spheres are for oxygen atoms in CO and CO-CO intermediates. Pink

    spheres are oxygen in the explicit water. Purple sphere demonstrates oxygen in OH ion. White

    spheres show hydrogen in explicit waters, and the light blue sphere shows hydrogen in OH ion.

  • 27

    Fig. S11. Morphology characterization of Cu deposited on carbon gas-diffusion layer. SEM

    images of the samples with deposition thickness of (a) 10 nm; (b) 25 nm; (c) 1000 nm. (d) SEM

    image of Cu nanoparticles deposited on gas-diffusion electrode with a catalyst loading of 1000

    g/cm2 using drop-casting technique.

  • 28

    Fig. S12. Cross-section characterization of the 25 nm sample Structural and compositional

    analysis technique using a focused ion beam (FIB) system and scanning transmission electron

    microscope (STEM). (a) Cross-section STEM of sample 25 nm. The fill region shows the

    roughness of the reaction interface where Cu is deposited. Elemental mapping for Carbon (c) and

    Cu (d) and overlap of all elements (b) on the cross-section mode of 25 nm. Tungsten was used to

    protect the surface during sample preparation using FIB.

  • 29

    Fig. S13. Chemical and physical characterization of Cu catalysts. (a) Cu 2p and (b) O 1s XPS

    spectra of the samples showing the presence of metallic and oxide phases on the surface of all

    thermal evaporated Cu samples and commercial Cu nanoparticles. (c) XRD diffraction patterns of

    1000 nm samples showing the presence of metallic crystalline phase.

  • 30

    Fig. S14. Performance of Cu catalysts in an H-cell configuration. Faradaic efficiencies for gas

    products in the potential range of -0.9 to -1.2 V vs. RHE (with iR correction) for 10 nm (a); 25 nm

    (b); 1000 nm (c); and 1000 g (d). The electrolyte was 0.1 M KHCO3 for all runs.

  • 31

    Fig. S15. Electrochemical characterization of Cu catalysts. (a) Linear sweep voltammetry scan

    with CO2 flow using 10 M KOH electrolyte on Cu samples of varying thickness. (b) Surface

    roughness factors of the samples normalized to polished Cu foil. The 10 nm, 25 nm, and 1000 nm

    samples are thermally deposited while the 1000 g sample uses drop-casted nanoparticles.

  • 32

    Fig. S16. Performance of Cu catalysts in 10 M KOH electrolyte. Faradaic efficiency for CO

    (a); and CH4 (b) on 10 nm, 25 nm, 1000 nm, and 1000 g catalysts using 10 M KOH electrolyte

    in the current range of 200 to 300 mA cm-2. (c) Faradaic efficiency for all products on the catalyst

    at the optimum current density for ethylene: 10 nm (250 mA cm-2); 25 nm (275 mA cm-2); 1000

    nm and 1000 g (225 mA cm-2). (d) C2H4 mass activity analysis showing exceptionally high mass

    activity of abrupt reaction interface samples.

  • 33

    Fig. S17. Representative nuclear magnetic resonance (NMR) spectra of the liquid products.

    NMR spectrum of the reaction products obtained upon CO2 reduction on 25 nm sample at current

    density of 275 mA cm-2 using 10 M KOH electrolyte. (a) full spectrum (largest singlet at 2.71 ppm

    corresponds to the reference DMSO signal, while the noisy part around 4.60 ppm is due to

    suppressed water signal); (b) magnified portion of the spectrum demonstrating ethanol and acetate

    product peaks; (c) magnified portion of the spectrum demonstrating the formation of a small

    amount of formate.

  • 34

    Fig. S18. Effect of varying CO2 concentrations on overall CO2 reduction selectivity on the 25 nm

    sample.

  • 35

    Fig. S19. CO2 reduction performance on an abrupt reaction interface sample. (a) Optimal C2

    selectivity on the 25 nm sample using 10 M KOH (at 275 mA cm-2); 5.5 M KOH (at 500 mA cm-

    2); 3.5 M KOH (at 750 mA/cm-2). (b) Applied potentials (iR corrected) at different current densities

    on the 25 nm sample using an optimized electrolyte for high C2H4 selectivity: 10 M KOH (at 275

    mA cm-2); 5.5 M KOH + 4 M KI (at 500 mA/cm-2); and 3.5 M KOH + 5 M KI (at 750 mA/cm-2).

  • 36

    Fig. S20. Stability test of the Cu catalyst deposited onto a traditional carbon gas-diffusion

    electrode. Long-term performance test of CO2 reduction to ethylene in 7 M KOH on the 1000 nm

    sample: (a) current density vs. time; and (b) Faradaic efficiency for gas products vs. time. The

    sample was tested under a constant applied voltage of -0.6 V vs. RHE (at which the sample shows

    the highest Faradaic efficiency for ethylene). Long-term performance test of CO2 reduction to

    ethylene in 1 M KOH on the 1000 nm sample: (c) current density vs. time; and (d) Faradaic

  • 37

    efficiency for gas products vs. time. The sample was tested under a constant applied voltage of -0.7

    V vs. RHE at which the current density is in the range of 100 mA/cm2. The samples were tested at

    different KOH concentration electrolytes and current densities to show that the low stability of the

    traditional gas-diffusion electrodes for CO2 reduction occurs at all testing conditions.

  • 38

    Fig. S21. Surface characterization of bare carbon gas-diffusion electrode under CO2

    reduction conditions. Contact angle measurements on (a) the starting gas-diffusion electrode

    and, (b) the electrode after applying a voltage of -0.8 V vs. RHE under CO2 reduction conditions

    for 1 hour. The contact angle reduced from 145° (starting electrode) to 117° (after operation),

    indicating that the hydrophobicity of the carbon electrode is reduced under CO2 reduction

    conditions. (c) O 1s XP spectra of the carbon gas-diffusion electrode before and after being applied

    an negative potential under CO2-RR conditions. The data show that the oxygen content in the

    carbon gas-diffusion electrode was doubled after being applied negative potentials of -0.4 and -0.8

    V vs. RHE, implying an increase in the surface OH, and COOH of the carbon electrode which

    leads to a more hydrophilic surface.

  • 39

    Fig. S22. Characterization of Graphite/Carbon NPs/Cu/PTFE electrode. Top view SEM

    image of Cu/PTFE after being coated with (a) carbon nanoparticles (NPs), and (b) by both carbon

    NPs and graphite. Cu 2p (c) and O 1s (d) XPS spectra of Cu/PTFE before being coated by Carbon

    and graphite showing a mixed oxidation state of Cu0 and Cu2+.

  • 40

    Fig. S23. Performance of the polymer-based gas-diffusion layer electrodes in 7 M KOH

    electrolyte. (a) Dependence of current density on applied potential for Graphite/Carbon

    NPs/Cu/PTFE and Graphite/Cu/PTFE electrodes. Faradaic efficiency for CO, H2 and CH4 on

    Graphite/Carbon NPs/Cu/PTFE (b) and Graphite/Cu/PTFE (c) using 7 M KOH electrolyte in the

  • 41

    applied potential range of -0.33 to -0.6 V vs. RHE. (d) Faradaic efficiency for all products on the

    Graphite/Carbon NPs/Cu/PTFE catalyst at the optimum potential for ethylene of -0.57 V vs. RHE.

  • 42

    Fig. S24. GC-MS analyses of ethylene produced from CO2-RR on Graphite/Carbon

    NPs/Cu/PTFE using 12CO2 and 13CO2 as feedstock. The GC-MS analysis of ethylene obtained

    from 12CO2 shows typical mass spectrum patterns of 12C-ethylene (m/s =28). When 13C-labeling

    CO2 was used as the reactant, the mass spectrum of ethylene shows patterns of 13C-ethylene (m/z

    =30), confirming that the ethylene is produced from CO2. The reaction was performed in 7 M KOH

    electrolyte with a constant applied potential of – 0.55 V vs. RHE.

  • 43

    Fig. S25. Stability tests of polymer-based gas-diffusion electrodes. Long-term performance

    tests of CO2 reduction to ethylene in 7 M KOH for Graphite/Carbon NPs/Cu/PTFE,

    Graphite/Cu/PTFE and Cu/Carbon gas-diffusion layers: current density vs time (a) and Faradaic

    efficiency for gas products vs. time on Graphite/Carbon NPs/Cu/PTFE (b); Graphite/Cu/PTFE (c);

  • 44

    and Cu on traditional Carbon gas-diffusion electrode (d). The samples were tested under a constant

    applied voltage of -0.55 V vs. RHE.

  • 45

    Fig. S26. Characterization of Cu/PTFE and Graphite/Cu/PTFE before and after reaction.

    Optical photographs of Cu/PTFE before (a) and after CO2-RR in 7 M KOH electrolyte at -0.55 V

    vs. RHE for 2 hours (b). Optical photographs of Graphite/Cu/PTFE before (c) and after CO2-RR

    in 7 M KOH electrolyte at -0.55 V vs. RHE for 40 hours. The arrow mark in (b) shows the

    disappearance of Cu layer which is due to the dissolution of non-electronically connected Cu area

    on the PTFE substrate. The arrow mark in (d) shows the presence of Cu islands or exposed Cu

    area which resulted from the detachment of the graphite layer from the Cu/PTFE surface. We

    rationalize that the low stability of Graphite/Cu/PTFE catalyst is because of the weak interaction

    between Cu surface and graphite which leads to the detachment of graphite from Cu surface. The

    a b

    c d

    Disappearance of Cu

    Exposed/re-deposited Cu

    300 m 300 m

    300 m300 m

  • 46

    Cu area without electrical contact would be dissolved in the electrolyte and redeposited on

    graphite, forming a catalyst for hydrogen evolution.

  • 47

    Fig. S27. Scanning electron microscopy (SEM) images of NiFeOx deposited on Ni foam using

    electrodeposition method. The sample was prepared as previously reported (39).

  • 48

    Fig. S28. (a) Oxygen evolution reaction (OER) performance of a NiFeO/Ni foam measured in a

    3-electrode setup (with iR correction); (b) Full-cell potential performance (with iR correction); (c)

    CO2-RR performance of the cathode potential vs. RHE as calculated from the OER (a) and full-

    cell performance (b) data. The additional data point (black circle) was measured in a 3-electrode

    setup and converted to a RHE scale using the predicted pH of the reaction-diffusion model. The

    overlap between both the black and red circles confirm the surface pH modelling data; (d)

  • 49

    Chronoamperometry of full-cell operation using Graphite/Carbon NP/Cu/PTFE NiFeOx (39)

    catalysts at a fixed cell-voltage of 2.4 V over 1 hour.

  • 50

    Table S1. Summary of CO2 reduction to C2H4 performance on different catalysts.

    Catalyst

    C2H4 onset

    potential

    (V vs.

    RHE)

    J(C2H4)

    mA cm-2

    C2H4

    Faradaic

    efficiency

    (%)

    C2H4

    energy

    conversion

    (%)(*)

    C2H4

    mass

    activity

    (A mg-1)

    Reference

    Plasma-

    Oxidized Cu

    -0.6 12 60 33 - (13)

    Plasma-Cu

    Nanocubes

    -0.6 16 40 25 - (54)

    N-doped

    graphene dots

    -0.45 40 33 18 0.08 (11)

    Cu

    nanostructure

    ~ -0.3 100 40 23 0.33 (9)

    Cu

    nanodendrites

    - 97 55 28 - (10)

    Cu

    nanoparticles

    -0.36 150 36 23 0.15 (12)

    Abrupt Cu

    Interface

    -0.165 473 66 40 17.8 This work

    Abrupt Cu

    Interface

    -0.165 184 66 44 6.7 This work

    (*) C2H4 energy efficiency is calculated for the half-cell (i.e. assuming the overpotential of the

    oxygen evolution reaction is zero). C2H4 energy efficiency = (1.23 + (-EC2H4))*FE(C2H4)/(1.23

    + (-E)), where E is the applied potential vs. RHE; EC2H4 = 0.08 V is thermodynamic potential

    (vs. RHE) of CO2 reduction to ethylene; FE(C2H4) is the measured C2H4 Faradaic efficiency in

    percentage.

  • 51

    Table S2. Ground state energy of the adsorbates on three different facets of the copper,

    without including explicit water.

    Adsorbates

    Ground state energy (eV)

    111 100 110

    CO -14.767 -14.767 -14.767

    Cu -247.790 -243.406 -299.709

    Cu-OH -258.531 -254.334 -310.672

    Cu-2OH -269.182 -265.208 -321.635

    Cu-CO -263.479 -259.149 -315.372

    Cu-CO-OH -274.225 -270.036 -326.168

    Cu-CO-2OH -284.884 -280.898 -336.956

    Cu-2CO -279.154 -274.791 -331.114

    Cu-2CO-OH -289.804 -285.540 -342.020

    Cu-2CO-2OH -300.390 -296.602 -352.896

    Cu-OCCO -277.486 -273.830 -329.684

    Cu-OCCO-OH -288.273 -284.673 -340.667

    Cu-OCCO-2OH -298.934 -295.757 -351.496

  • 52

    Table S3. CO adsorption energy and C-C coupling energy barrier on three different facets

    of copper at different OH concentrations, without including explicit water molecules.

    OH

    Concentration

    CO adsorption energy (eV) CO-CO coupling energy barrier

    (eV)

    111 100 110 111 100 110

    0 -0.922 -0.976 -0.897 1.668 0.961 1.430

    1/16 -0.927 -0.935 -0.730 1.530 0.867 1.353

    2/16 -0.936 -0.924 -0.555 1.456 0.845 1.400

    3/16 -0.686 -0.919 -0.726 1.371 0.822 1.304

    4/16 -0.556 -0.823 -0.495 1.234 0.763 1.382

  • 53

    Table S4. Electronic structure energy (E), zero point energy (ZPE), heat capacity (Cv) and

    entropy (S) at room temperature (T), and Gibbs free energy (G) all in eV, for CO-CO

    coupling on Cu(100) without explicit water and with different number of OH ions. IS

    mentions the initial state, i.e. 2 adsorbed CO on the surface, and FS mentions the final state,

    i.e. OCCO intermediate.

    OH Concentration State E ZPE Cv S.T G (eV)

    0 IS -274.791 0.324 0.160 0.304 -274.611

    FS -273.830 0.415 0.119 0.234 -273.530

    1-OH (1/16 ML) IS -285.540 0.354 0.159 0.332 -285.359

    FS -284.673 0.416 0.119 0.237 -284.375

    2-OH (2/16ML)

    IS -296.368 0.340 0.159 0.331 -296.200

    FS -295.757 0.423 0.119 0.238 -295.453

  • 54

    Table S5. Electronic structure energy (E), zero point energy (ZPE), heat capacity (Cv) and

    entropy (S) at room temperature (T), and Gibbs free energy (G) all in eV, for CO-CO

    coupling on Cu(111) without explicit water and with different number of OH ions. IS

    mentions the initial state, i.e. 2 adsorbed CO on the surface, and FS mentions the final state,

    i.e. OCCO intermediate.

    OH

    Concentration State E ZPE Cv S.T G (eV)

    0-OH IS -279.154 0.359 0.156 0.297 -278.936

    FS -277.486 0.398 0.115 0.234 -277.207

    1-OH (1/16ML) IS -289.804 0.352 0.152 0.287 -289.587

    FS -288.273 0.401 0.112 0.217 -287.977

    2-OH (2/16ML)

    IS -300.159 0.366 0.164 0.416 -300.045

    FS -298.934 0.387 0.111 0.216 -298.652

  • 55

    Table S6. Thermodynamic Gibbs energy differences of the initial and final states, and also

    the activation energy barrier for CO-CO coupling on Cu(100) without explicit water and

    with different number of OH ions. IS mentions the initial state, i.e. 2 adsorbed CO on the

    surface, TS mentions the transition state and FS mentions the final state, i.e. OCCO

    intermediate.

    OH

    Concentration State G (eV)

    Thermodynamic

    Energy Barrier

    (eV)

    Activation

    Energy Barrier

    (eV)

    0-OH

    IS -274.611

    TS -273.530 1.081 1.081

    FS -273.530

    1-OH (1/16ML)

    IS -285.359

    TS -284.375 0.984 0.984

    FS -284.375

    2-OH (2/16ML)

    IS -296.200

    TS -295.274 0.747 0.926

    FS -295.453

  • 56

    Table S7. Thermodynamic Gibbs energy differences of the initial and final states, and also

    the activation energy barrier for CO-CO coupling on Cu(111) without explicit water and

    with different number of OH ions. IS mentions the initial state, i.e. 2 adsorbed CO on the

    surface, TS mentions the transition state and FS mentions the final state, i.e. OCCO

    intermediate.

    OH

    Concentration State G (eV)

    Thermodynamic

    Energy Barrier (eV)

    Activation

    Energy Barrier

    (eV)

    0-OH

    IS -278.936

    TS -277.187 1.729 1.749

    FS -277.207

    1-OH

    IS -289.587

    TS -287.977 1.609 1.609

    FS -287.977

    2-OH

    IS -300.045

    TS -298.453 1.393 1.592

    FS -298.652

  • 57

    Table S8. Electronic structure energy (E), zero point energy (ZPE), heat capacity (Cv) and

    entropy (S) at room temperature (T), and Gibbs free energy (G) all in eV, for CO-CO

    coupling on Cu(100) with three explicit water and with one OH ion (i.e. 1/16ML) as adsorbed

    or solvated. IS mentions the initial state, i.e. 2 adsorbed CO on the surface, and FS mentions

    the final state, i.e. OCCO intermediate.

    OH

    Concentration State E ZPE Cv S.T G (eV)

    0-OH IS -317.742 0.328 0.156 0.292 -317.550

    FS -316.742 0.424 0.116 0.215 -316.417

    1 adsorbed OH IS -328.754 0.349 0.160 0.370 -328.615

    FS -327.809 0.425 0.115 0.213 -327.482

    1 solvated OH IS -326.356 0.305 0.126 0.247 -326.172

    FS -325.463 0.419 0.098 0.189 -325.135

  • 58

    Table S9. Electronic structure energy (E), zero point energy (ZPE), heat capacity (Cv) and

    entropy (S) at room temperature (T), and Gibbs free energy (G) all in eV, for CO-CO

    coupling on Cu(111) with three explicit water and with one OH ion (i.e. 1/16ML) as adsorbed

    or solvated. IS mentions the initial state, i.e. 2 adsorbed CO on the surface, and FS mentions

    the final state, i.e. OCCO intermediate.

    OH

    Concentration State E ZPE Cv S.T G (eV)

    0-OH

    IS -321.991 0.351 0.152 0.284 -321.772

    FS -320.306 0.392 0.132 0.267 -320.049

    1 adsorbed OH

    IS -332.896 0.353 0.149 0.278 -332.672

    FS -331.235 0.392 0.132 0.271 -330.982

    1 solvated OH

    IS -330.715 0.344 0.121 0.218 -330.468

    FS -328.678 0.387 0.121 0.249 -328.671

  • 59

    Table S10. Thermodynamic Gibbs energy differences of the initial and final states, and also

    the activation energy barrier for CO-CO coupling on Cu(100) with three explicit water

    molecules and with one OH ion (i.e. 1/16ML) as adsorbed or solvated. IS mentions the initial

    state, i.e. 2 adsorbed CO on the surface, TS mentions the transition state and FS mentions

    the final state, i.e. OCCO intermediate.

    OH

    Concentration State G (eV)

    Thermodynamic

    Energy Barrier

    (eV)

    Activation

    Energy Barrier

    (eV)

    0-OH

    IS -317.550

    TS -316.328 1.133 1.222

    FS -316.417

    1 adsorbed OH

    IS -328.615

    TS -327.474 1.132 1.140

    FS -327.482

    1 solvated OH

    IS -326.172

    TS -324.318 1.037 1.854

    FS -325.135

  • 60

    Table S11. Thermodynamic Gibbs energy differences of the initial and final states, and also

    the activation energy barrier for CO-CO coupling on Cu(111) with three explicit water

    molecules and with one OH ion (i.e. 1/16ML) as adsorbed or solvated. IS mentions the initial

    state, i.e. 2 adsorbed CO on the surface, TS mentions the transition state and FS mentions

    the final state, i.e. OCCO intermediate.

    OH

    Concentration State G (eV)

    Thermodynamic

    Energy Barrier

    (eV)

    Activation

    Energy Barrier

    (eV)

    0-OH

    IS -321.772

    TS -319.887 1.724 1.886

    FS -320.049

    1 adsorbed OH

    IS -332.672

    TS -330.847 1.690 1.825

    FS -330.982

    1 solvated OH

    IS -330.468

    TS -328.171 1.797 2.297

    FS -328.671

  • 61

    Table S12. Physical properties of KOH electrolyte and its effect on the C2H4 onset potential

    and Tafel slope

    KOH

    concentration

    (M) Resistance () (*) Surface pH (**)

    C2H4 onset

    potential (V vs.

    RHE) (†)

    C2H4 Tafel

    slope (mV

    dec-1) (‡)

    1 2.58 12.42 -0.465 135

    5 1.08 14.48 -0.285 97

    10 0.98 14.94 -0.165 65

    (*): Measured using electrochemical impedance spectroscopy (EIS); (**) Surface pH predicted

    using a 1D reaction-diffusion model at 0 mA/cm2 accounting for the effect of CO2 gas in the

    catalyst layer; (†) The potential at which the catalyst shows an ethylene faradaic efficiency of

    1%; (‡) Calculated based on the experimental C2H4 partial current densities.

  • 62

    Table S13. Table of CO2 reduction to ethylene stabilities reported in Literature

    Catalyst Reported

    Duration (h) Stable C2H4

    Selectivity

    C2H4

    Current

    Density

    (mA/cm2)

    Source

    Cu-Nanocubes 1 45 % 23 (54)

    Cu-Nanoparticles 4 18 % 27 (12)

    Plasma-Cu 5 58 % 7 (13)

    Cu-Mesocrystals 6 23 % 4 (55)

    CuZn-

    Nanoparticles 8 ~30 % 14 (56)

    Oxide-derived

    CuZn 10 8 % 3 (58)

    Cu-Nanocubes 10 32 % 7 (57)

    Graphite/Cu/PTFE 24 55 % 55 This work

    Graphite/Carbon

    NP’s/Cu/PTFE 150 70 % 55-70 This work

  • 63

    References and Notes 1. M. Schreier, F. Héroguel, L. Steier, S. Ahmad, J. S. Luterbacher, M. T. Mayer, J. Luo, M.

    Grätzel, Solar conversion of CO2 to CO using Earth-abundant electrocatalysts prepared by atomic layer modification of CuO. Nat. Energy 2, 17087 (2017). doi:10.1038/nenergy.2017.87

    2. S. Lin, C. S. Diercks, Y.-B. Zhang, N. Kornienko, E. M. Nichols, Y. Zhao, A. R. Paris, D. Kim, P. Yang, O. M. Yaghi, C. J. Chang, Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 349, 1208–1213 (2015). doi:10.1126/science.aac8343 Medline

    3. S. Gao, Y. Lin, X. Jiao, Y. Sun, Q. Luo, W. Zhang, D. Li, J. Yang, Y. Xie, Partially oxidized atomic cobalt layers for carbon dioxide electroreduction to liquid fuel. Nature 529, 68–71 (2016). doi:10.1038/nature16455 Medline

    4. R. Kas, K. K. Hummadi, R. Kortlever, P. de Wit, A. Milbrat, M. W. J. Luiten-Olieman, N. E. Benes, M. T. M. Koper, G. Mul, Three-dimensional porous hollow fibre copper electrodes for efficient and high-rate electrochemical carbon dioxide reduction. Nat. Commun. 7, 10748 (2016). doi:10.1038/ncomms10748 Medline

    5. M. Liu, Y. Pang, B. Zhang, P. De Luna, O. Voznyy, J. Xu, X. Zheng, C. T. Dinh, F. Fan, C. Cao, F. P. G. de Arquer, T. S. Safaei, A. Mepham, A. Klinkova, E. Kumacheva, T. Filleter, D. Sinton, S. O. Kelley, E. H. Sargent, Enhanced electrocatalytic CO2 reduction via field-induced reagent concentration. Nature 537, 382–386 (2016). doi:10.1038/nature19060 Medline

    6. K. D. Yang, W. R. Ko, J. H. Lee, S. J. Kim, H. Lee, M. H. Lee, K. T. Nam, Morphology-directed selective production of ethylene or ethane from CO2 on a Cu mesopore electrode. Angew. Chem. Int. Ed. 56, 796–800 (2017). doi:10.1002/anie.201610432 Medline

    7. A. Loiudice, P. Lobaccaro, E. A. Kamali, T. Thao, B. H. Huang, J. W. Ager, R. Buonsanti, Tailoring copper nanocrystals towards C2 products in electrochemical CO2 reduction. Angew. Chem. Int. Ed. 55, 5789–5792 (2016). doi:10.1002/anie.201601582 Medline

    8. F. S. Roberts, K. P. Kuhl, A. Nilsson, High selectivity for ethylene from carbon dioxide reduction over copper nanocube electrocatalysts. Angew. Chem. Int. Ed. 54, 5179–5182 (2015). doi:10.1002/anie.201412214 Medline

    9. T. T. Hoang, S. Ma, J. I. Gold, P. J. Kenis, A. A. Gewirth, Nanoporous copper films by additive-controlled electrodeposition: CO2 reduction catalysis. ACS Catal. 7, 3313–3321 (2017). doi:10.1021/acscatal.6b03613

    10. C. Reller, R. Krause, E. Volkova, B. Schmid, S. Neubauer, A. Rucki, M. Schuster, G. Schmid, Selective electroreduction of CO2 toward ethylene on nano dendritic copper catalysts at high current density. Adv. Energy Mater. 7, 1602114 (2017). doi:10.1002/aenm.201602114

    11. J. Wu, S. Ma, J. Sun, J. I. Gold, C. Tiwary, B. Kim, L. Zhu, N. Chopra, I. N. Odeh, R. Vajtai, A. Z. Yu, R. Luo, J. Lou, G. Ding, P. J. A. Kenis, P. M. Ajayan, A metal-free

    http://dx.doi.org/10.1038/nenergy.2017.87http://dx.doi.org/10.1126/science.aac8343http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=26292706&dopt=Abstracthttp://dx.doi.org/10.1038/nature16455http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=26738592&dopt=Abstracthttp://dx.doi.org/10.1038/ncomms10748http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=26888578&dopt=Abstracthttp://dx.doi.org/10.1038/nature19060http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27487220&dopt=Abstracthttp://dx.doi.org/10.1002/anie.201610432http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28000371&dopt=Abstracthttp://dx.doi.org/10.1002/anie.201601582http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27059162&dopt=Abstracthttp://dx.doi.org/10.1002/anie.201412214http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25728325&dopt=Abstracthttp://dx.doi.org/10.1021/acscatal.6b03613http://dx.doi.org/10.1002/aenm.201602114

  • 64

    electrocatalyst for carbon dioxide reduction to multi-carbon hydrocarbons and oxygenates. Nat. Commun. 7, 13869 (2016). doi:10.1038/ncomms13869 Medline

    12. S. Ma, M. Sadakiyo, R. Luo, M. Heima, M. Yamauchi, P. J. A. Kenis, One-step electrosynthesis of ethylene and ethanol from CO2 in an alkaline electrolyzer. J. Power Sources 301, 219–228 (2016). doi:10.1016/j.jpowsour.2015.09.124

    13. H. Mistry, A. S. Varela, C. S. Bonifacio, I. Zegkinoglou, I. Sinev, Y.-W. Choi, K. Kisslinger, E. A. Stach, J. C. Yang, P. Strasser, B. R. Cuenya, Highly selective plasma-activated copper catalysts for carbon dioxide reduction to ethylene. Nat. Commun. 7, 12123 (2016). doi:10.1038/ncomms12123 Medline

    14. Y. Li, F. Cui, M. B. Ross, D. Kim, Y. Sun, P. Yang, Structure-sensitive CO2 electroreduction to hydrocarbons on ultrathin 5-fold twinned copper nanowires. Nano Lett. 17, 1312–1317 (2017). doi:10.1021/acs.nanolett.6b05287 Medline

    15. P. De Luna, R. Quintero-Bermudez, C.-T. Dinh, M. B. Ross, O. S. Bushuyev, P. Todorović, T. Regier, S. O. Kelley, P. Yang, E. H. Sargent, Catalyst electro-redeposition controls morphology and oxidation state for selective carbon dioxide reduction. Nat. Catal. 1, 103–110 (2018). doi:10.1038/s41929-017-0018-9

    16. Y. Chen, C. W. Li, M. W. Kanan, Aqueous CO2 reduction at very low overpotential on oxide-derived Au nanoparticles. J. Am. Chem. Soc. 134, 19969–19972 (2012). doi:10.1021/ja309317u Medline

    17. D. Ren, Y. Deng, A. D. Handoko, C. S. Chen, S. Malkhandi, B. S. Yeo, Selective electrochemical reduction of carbon dioxide to ethylene and ethanol on copper(I) oxide catalysts. ACS Catal. 5, 2814–2821 (2015). doi:10.1021/cs502128q

    18. J. Wu, R. M. Yadav, M. Liu, P. P. Sharma, C. S. Tiwary, L. Ma, X. Zou, X.-D. Zhou, B. I. Yakobson, J. Lou, P. M. Ajayan, Achieving highly efficient, selective, and stable CO2 reduction on nitrogen-doped carbon nanotubes. ACS Nano 9, 5364–5371 (2015). doi:10.1021/acsnano.5b01079 Medline

    19. D. Kim, C. Xie, N. Becknell, Y. Yu, M. Karamad, K. Chan, E. J. Crumlin, J. K. Nørskov, P. Yang, Electrochemical activation of CO2 through atomic ordering transformations of AuCu nanoparticles. J. Am. Chem. Soc. 139, 8329–8336 (2017). doi:10.1021/jacs.7b03516 Medline

    20. S. Ma, M. Sadakiyo, M. Heima, R. Luo, R. T. Haasch, J. I. Gold, M. Yamauchi, P. J. A. Kenis, Electroreduction of carbon dioxide to hydrocarbons using bimetallic Cu–Pd catalysts with different mixing patterns. J. Am. Chem. Soc. 139, 47–50 (2017). doi:10.1021/jacs.6b10740 Medline

    21. R. Kas, R. Kortlever, H. Yilmaz, M. T. M. Koper, G. Mul, Manipulating the hydrocarbon selectivity of copper nanoparticles in CO2 electroreduction by process conditions. ChemElectroChem 2, 354–358 (2015). doi:10.1002/celc.201402373

    22. H. Xiao, T. Cheng, W. A. Goddard 3rd, R. Sundararaman, Mechanistic explanation of the pH dependence and onset potentials for hydrocarbon products from electrochemical reduction of CO on Cu (111). J. Am. Chem. Soc. 138, 483–486 (2016). doi:10.1021/jacs.5b11390 Medline

    http://dx.doi.org/10.1038/ncomms13869http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27958290&dopt=Abstracthttp://dx.doi.org/10.1016/j.jpowsour.2015.09.124http://dx.doi.org/10.1038/ncomms12123http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27356485&dopt=Abstracthttp://dx.doi.org/10.1021/acs.nanolett.6b05287http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28094953&dopt=Abstracthttp://dx.doi.org/10.1038/s41929-017-0018-9http://dx.doi.org/10.1021/ja309317uhttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=23171134&dopt=Abstracthttp://dx.doi.org/10.1021/cs502128qhttp://dx.doi.org/10.1021/acsnano.5b01079http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25897553&dopt=Abstracthttp://dx.doi.org/10.1021/jacs.7b03516http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28551991&dopt=Abstracthttp://dx.doi.org/10.1021/jacs.6b10740http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27958727&dopt=Abstracthttp://dx.doi.org/10.1002/celc.201402373http://dx.doi.org/10.1021/jacs.5b11390http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=26716884&dopt=Abstract

  • 65

    23. Y. Lum, B. Yue, P. Lobaccaro, A. T. Bell, J. W. Ager, Optimizing C-C coupling on oxide-derived copper catalysts for electrochemical CO2 reduction. J. Phys. Chem. C 121, 14191–14203 (2017). doi:10.1021/acs.jpcc.7b03673

    24. Y. Huang, A. D. Handoko, P. Hirunsit, B. S. Yeo, Electrochemical reduction of CO2 using copper single-crystal surfaces: Effects of CO* coverage on the selective formation of ethylene. ACS Catal. 7, 1749–1756 (2017). doi:10.1021/acscatal.6b03147

    25. M. Favaro, H. Xiao, T. Cheng, W. A. Goddard 3rd, J. Yano, E. J. Crumlin, Subsurface oxide plays a critical role in CO2 activation by Cu(111) surfaces to form chemisorbed CO2, the first step in reduction of CO2. Proc. Natl. Acad. Sci. U.S.A. 114, 6706–6711 (2017). Medline

    26. C. Liu, M. P. Lourenço, S. Hedström, F. Cavalca, O. Diaz-Morales, H. A. Duarte, A. Nilsson, L. G. M. Pettersson, Stability and effects of subsurface oxygen in oxide-derived Cu catalyst for CO2 reduction. J. Phys. Chem. C 121, 25010–25017 (2017). doi:10.1021/acs.jpcc.7b08269

    27. R. Subbaraman, D. Tripkovic, D. Strmcnik, K.-C. Chang, M. Uchimura, A. P. Paulikas, V. Stamenkovic, N. M. Markovic, Enhancing hydrogen evolution activity in water splitting by tailoring Li+-Ni(OH)2-Pt interfaces. Science 334, 1256–1260 (2011). doi:10.1126/science.1211934 Medline

    28. Y. Y. Birdja, M. T. M. Koper, The importance of cannizzaro-type reactions during electrocatalytic reduction of carbon dioxide. J. Am. Chem. Soc. 139, 2030–2034 (2017). doi:10.1021/jacs.6b12008 Medline

    29. V. R. Stamenkovic, D. Strmcnik, P. P. Lopes, N. M. Markovic, Energy and fuels from electrochemical interfaces. Nat. Mater. 16, 57–69 (2016). doi:10.1038/nmat4738 Medline

    30. W. Sheng, M. Myint, J. G. Chen, Y. Yan, Correlating the hydrogen evolution reaction activity in alkaline electrolytes with the hydrogen binding energy on monometallic surfaces. Energy Environ. Sci. 6, 1509–1512 (2013). doi:10.1039/c3ee00045a

    31. H. Ooka, M. C. Figueiredo, M. T. M. Koper, Competition between hydrogen evolution and carbon dioxide reduction on copper electrodes in mildly acidic media. Langmuir 33, 9307–9313 (2017). doi:10.1021/acs.langmuir.7b00696 Medline

    32. N. Gupta, M. Gattrell, B. MacDougall, Calculation for the cathode surface concentrations in the electrochemical reduction of CO2 in KHCO3 solutions. J. Appl. Electrochem. 36, 161–172 (2006). doi:10.1007/s10800-005-9058-y

    33. K. J. P. Schouten, Z. Qin, E. Pérez Gallent, M. T. M. Koper, Two pathways for the formation of ethylene in CO reduction on single-crystal copper electrodes. J. Am. Chem. Soc. 134, 9864–9867 (2012). doi:10.1021/ja302668n Medline

    34. J. H. Montoya, C. Shi, K. Chan, J. K. Nørskov, Theoretical insights into a CO dimerization mechanism in CO2 electroreduction. J. Phys. Chem. Lett. 6, 2032–2037 (2015). doi:10.1021/acs.jpclett.5b00722 Medline

    35. J. D. Goodpaster, A. T. Bell, M. Head-Gordon, Identification of possible pathways for C–C bond formation during electrochemical reduction of CO2: New theoretical insights from

    http://dx.doi.org/10.1021/acs.jpcc.7b03673http://dx.doi.org/10.1021/acscatal.6b03147http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28607092&dopt=Abstracthttp://dx.doi.org/10.1021/acs.jpcc.7b08269http://dx.doi.org/10.1126/science.1211934http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=22144621&dopt=Abstracthttp://dx.doi.org/10.1021/jacs.6b12008http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28099805&dopt=Abstracthttp://dx.doi.org/10.1038/nmat4738http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27994237&dopt=Abstracthttp://dx.doi.org/10.1039/c3ee00045ahttp://dx.doi.org/10.1021/acs.langmuir.7b00696http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28453940&dopt=Abstracthttp://dx.doi.org/10.1007/s10800-005-9058-yhttp://dx.doi.org/10.1021/ja302668nhttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=22670713&dopt=Abstracthttp://dx.doi.org/10.1021/acs.jpclett.5b00722http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=26266498&dopt=Abstract

  • 66

    an improved electrochemical model. J. Phys. Chem. Lett. 7, 1471–1477 (2016). doi:10.1021/acs.jpclett.6b00358 Medline

    36. K. J. P. Schouten, E. Pérez Gallent, M. T. M. Koper, The influence of pH on the reduction of CO and CO2 to hydrocarbons on copper electrodes. J. Electroanal. Chem. 716, 53–57 (2014). doi:10.1016/j.jelechem.2013.08.033

    37. A. S. Varela, W. Ju, T. Reier, P. Strasser, Tuning the catalytic activity and selectivity of Cu for CO2 electroreduction in the presence of halides. ACS Catal. 6, 2136–2144 (2016). doi:10.1021/acscatal.5b02550

    38. Y.-G. Kim, A. Javier, J. H. Baricuatro, D. Torelli, K. D. Cummins, C. F. Tsang, J. C. Hemminger, M. P. Soriaga, Surface reconstruction of pure-Cu single-crystal electrodes under CO-reduction potentials in alkaline solutions: A study by seriatim ECSTM-DEMS. J. Electroanal. Chem. 780, 290–295 (2016). doi:10.1016/j.jelechem.2016.09.029

    39. X. Lu, C. Zhao, Electrodeposition of hierarchically structured three-dimensional nickel-iron electrodes for efficient oxygen evolution at high current densities. Nat. Commun. 6, 6616 (2015). doi:10.1038/ncomms7616 Medline

    40. G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996). doi:10.1103/PhysRevB.54.11169 Medline

    41. G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999). doi:10.1103/PhysRevB.59.1758

    42. S. R. Bahn, K. W. Jacobsen, An object-oriented scripting interface to a legacy electronic structure code. Comput. Sci. Eng. 4, 56–66 (2002). doi:10.1109/5992.998641

    43. G. Henkelman, H. Jónsson, Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 113, 9978–9985 (2000). doi:10.1063/1.1323224

    44. J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). doi:10.1103/PhysRevLett.77.3865 Medline

    45. S. E. Hill, E. D. Glendening, D. Feller, Theoretical study of cation/ether complexes: The alkali metals and dimethyl ether. J. Phys. Chem. A 101, 6125–6131 (1997). doi:10.1021/jp971124+

    46. P. Vassilev, M. J. Louwerse, E. J. Baerends, Hydroxyl radical and hydroxide ion in liquid water: A comparative electron density functional theory study. J. Phys. Chem. B 109, 23605–23610 (2005). doi:10.1021/jp044751p Medline

    47. E. I. Wei, E. I. Proynov, A. Milet, D. R. Salahub, Solvation of the hydroxide anion: A combined DFT and molecular dynamics study. J. Phys. Chem. A 104, 2384–2395 (2000). doi:10.1021/jp992540s

    48. T. Burdyny, P. J. Graham, Y. Pang, C.-T. Dinh, M. Liu, E. H. Sargent, D. Sinton, Nanomorphology-enhanced gas-evolution intensifies CO2 reduction electrochemistry. ACS Sustain. Chem. Eng. 5, 4031–4040 (2017). doi:10.1021/acssuschemeng.7b00023

    http://dx.doi.org/10.1021/acs.jpclett.6b00358http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=27045040&dopt=Abstracthttp://dx.doi.org/10.1016/j.jelechem.2013.08.033http://dx.doi.org/10.1021/acscatal.5b02550http://dx.doi.org/10.1016/j.jelechem.2016.09.029http://dx.doi.org/10.1038/ncomms7616http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25776015&dopt=Abstracthttp://dx.doi.org/10.1103/PhysRevB.54.11169http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=9984901&dopt=Abstracthttp://dx.doi.org/10.1103/PhysRevB.59.1758http://dx.doi.org/10.1109/5992.998641http://dx.doi.org/10.1063/1.1323224http://dx.doi.org/10.1103/PhysRevLett.77.3865http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=10062328&dopt=Abstracthttp://dx.doi.org/10.1021/jp971124+http://dx.doi.org/10.1021/jp044751phttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=16375337&dopt=Abstracthttp://dx.doi.org/10.1021/jp992540shttp://dx.doi.org/10.1021/acssuschemeng.7b00023

  • 67

    49. C. W. Li, J. Ciston, M. W. Kanan, Electroreduction of carbon monoxide to liquid fuel on oxide-derived nanocrystalline copper. Nature 508, 504–507 (2014). doi:10.1038/nature13249 Medline

    50. M. R. Singh, E. L. Clark, A. T. Bell, Effects of electrolyte, catalyst, and membrane composition and operating conditions on the performance of solar-driven electrochemical reduction of carbon dioxide. Phys. Chem. Chem. Phys. 17, 18924–18936 (2015). doi:10.1039/C5CP03283K Medline

    51. W. Tang, E. Sanville, G. Henkelman, A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter 21, 084204 (2009). Medline

    52. S. Weisenberger, A. Schumpe, Estimation of gas solubilities in salt solutions at temperatures from 273 K to 363 K. AIChE J. 42, 298–300 (1996). doi:10.1002/aic.690420130

    53. K. Wu, E. Birgersson, B. Kim, P. J. A. Kenis, I. A. Karimi, Modeling and experimental validation of electrochemical reduction of CO2 to CO in a microfluidic cell. J. Electrochem. Soc. 162, F23–F32 (2015). doi:10.1149/2.1021414jes

    54. D. Gao, I. Zegkinoglou, N. J. Divins, F. Scholten, I. Sinev, P. Grosse, B. Roldan Cuenya, Plasma-activated copper nanocube catalysts for efficient carbon dioxide electroreduction to hydrocarbons and alcohols. ACS Nano 11, 4825–4831 (2017). doi:10.1021/acsnano.7b01257 Medline

    55. C. S. Chen, A. D. Handoko, J. H. Wan, L. Ma, D. Ren, B. S. Yeo, Stable and selective electrochemical reduction of carbon dioxide to ethylene on copper mesocrystals. Catal. Sci. Technol. 5, 161–168 (2015). doi:10.1039/C4CY00906A

    56. Y. Peng, T. Wu, L. Sun, J. M. V. Nsanzimana, A. C. Fisher, X. Wang, Selective electrochemical reduction of CO2 to ethylene on nanopores-modified copper electrodes in aqueous solution. ACS Appl. Mater. Interfaces 9, 32782–32789 (2017). doi:10.1021/acsami.7b10421 Medline

    57. D. Kim, C. S. Kley, Y. Li, P. Yang, Copper nanoparticle ensembles for selective electroreduction of CO2 to C2-C3 products. Proc. Natl. Acad. Sci. U.S.A. 114, 10560–10565 (2017). doi:10.1073/pnas.1711493114 Medline

    58. D. Ren, B. S.-H. Ang, B. S. Yeo, Tuning the selectivity of carbon dioxide electroreduction toward ethanol on oxide-derived CuxZn catalysts. ACS Catal. 6, 8239–8247 (2016). doi:10.1021/acscatal.6b02162

    http://dx.doi.org/10.1038/nature13249http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=24717429&dopt=Abstracthttp://dx.doi.org/10.1039/C5CP03283Khttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=26103939&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=21817356&dopt=Abstracthttp://dx.doi.org/10.1002/aic.690420130http://dx.doi.org/10.1149/2.1021414jeshttp://dx.doi.org/10.1021/acsnano.7b01257http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28441005&dopt=Abstracthttp://dx.doi.org/10.1039/C4CY00906Ahttp://dx.doi.org/10.1021/acsami.7b10421http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28880067&dopt=Abstracthttp://dx.doi.org/10.1073/pnas.1711493114http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=28923930&dopt=Abstracthttp://dx.doi.org/10.1021/acscatal.6b02162