• 검색 결과가 없습니다.

Modulating Local CO2 Concentration as a General Strategy for Enhancing C-C Coupling in CO2 Electroreduction

N/A
N/A
Protected

Academic year: 2021

Share "Modulating Local CO2 Concentration as a General Strategy for Enhancing C-C Coupling in CO2 Electroreduction"

Copied!
18
0
0

로드 중.... (전체 텍스트 보기)

전체 글

(1)

Article

Modulating Local CO

2

Concentration as a

General Strategy for Enhancing C C Coupling

in CO

2

Electroreduction

Local environment plays an important role in steering the reaction pathways in electrochemical CO2reduction reaction. Here, we present three approaches to modulate local CO2concentration in gas-diffusion electrode flow electrolyzers. Employing monodisperse Cu2O nanoparticles as the model catalysts, we demonstrate that providing a moderate local CO2concentration is effective in promoting C–C coupling. Ultimately, this study serves as a rational guide to tune CO2mass transport in gas-diffusion electrode electrolyzers for the optimal production of valuable multi-carbon molecules.

Ying Chuan Tan, Kelvin Berm Lee, Hakhyeon Song, Jihun Oh jihun.oh@kaist.ac.kr

HIGHLIGHTS

Three ways to modulate local CO2 concentration in GDE flow electrolyzer

Providing moderate local CO2 concentration enabled optimal C2+product selectivity

75.5% faradaic efficiency for C2+ products at 300 mA cm 2in KHCO3electrolyte

C2+product selectivity showed significant sensitivity toward CO2 mass transport

Tan et al., Joule4, 1104–1120 May 20, 2020ª 2020 Elsevier Inc. https://doi.org/10.1016/j.joule.2020.03.013

(2)

Article

Modulating Local CO

2

Concentration

as a General Strategy for Enhancing

C

C Coupling in CO

2

Electroreduction

Ying Chuan Tan,

1,4

Kelvin Berm Lee,

1,4

Hakhyeon Song,

1

and Jihun Oh

1,2,3,5,

*

SUMMARY

Flow electrolyzers based on gas-diffusion electrodes (GDEs) have

been increasingly employed to advance toward industry-relevant

electrochemical CO

2

reduction reaction (CO2RR) performance,

though fundamental understanding of the GDE system is still

lack-ing. Here, we propose that regulating local CO

2

concentration on

copper (Cu) surfaces is an effective and general strategy to promote

CC coupling in CO2RR. Local CO

2

concentration could influence the

surface coverage of *CO

2

, *H, and *CO, which affects the reaction

pathways toward multi-carbon (C

2+

) products. Guided by

mass-transport modeling, we have identified three approaches to

modu-late the local CO

2

concentration in GDE-based electrolyzers: (1)

cata-lyst layer structure, (2) feed CO

2

concentration, and (3) feed flow

rate. Utilizing Cu

2

O nanoparticles as the model catalysts,

modula-tion of local CO

2

concentration enabled an optimized faradaic

effi-ciency toward C

2+

products of up to 75.5% at 300 mA cm

2

and

C

2+

partial current density of up to 342 mA cm

2

in 1.0 M KHCO

3

.

INTRODUCTION

With the growing abundance of cheap renewable electricity, electrochemical CO2 reduction reaction (CO2RR) has gained increasing attention from the research community due to its potential to economically convert anthropogenic CO2to a wide variety of carbon-based chemicals that are traditionally derived from fossil fuels.1–4In particular, immense efforts have been focused on the selective produc-tion of multi-carbon (C2+) products, such as ethylene, ethanol, and 1-propanol, due to their large market size and high energy density.5–7

C2+product faradaic efficiency (FE) of more than 70% in conventional H-type or sandwich liquid cells has been achieved with a low partial current density (jC2+) of less than 20 mA cm2due to the low CO2availability in such systems.8,9In order to overcome CO2 mass transport issues, gas-diffusion electrodes (GDEs) have been employed in flow electrolyzers to achieve industrially relevant current densities (>200 mA cm2).10–12However, transferring these high C

2+product selectivity per-formances to GDE-based electrolyzers has been challenging without relying on alka-line electrolyte.13–15The use of alkaline electrolyte has raised concerns regarding the cost for electrolyte regeneration due to the continuous carbonation formation, which also causes stability issues.16Only recently, Li et al. succeeded in attaining C2+product FE of over 80% in near-neutral 1 M KHCO3electrolyte at high current densities by employing molecular additives to increase *CO adsorbate stability on Cu surfaces.17Nonetheless, there must be a scientific rationale behind the unusual difficulty in promoting C–C coupling selectively in neutral media using unmodified metallic Cu or oxide-derived Cu (OD-Cu) in GDE-based systems.

Context & Scale

Electrochemical CO2reduction powered by renewable energy is a potential approach to cut CO2 emission while creating value to the society. The electrosynthesis of multi-carbon (C2+) products, such as ethylene, ethanol, and 1-propanol, from CO2is desired because of the high energy density and large global markets of these products. In order to achieve selective conversion of CO2into C2+products, past studies have been focused on the design of catalysts and the tuning of local environment (e.g., pH, cations, and molecular additives). Here, we report that local CO2 concentration has been an overlooked factor that affects the selectivity toward C2+products. Utilizing our control catalysts and a gas-diffusion electrode flow cell operated at industrially relevant activity in near-neutral electrolyte, we show that a high local CO2 concentration leads to a suboptimal selectivity. Instead, providing a moderate local CO2 concentration is effective in promoting C–C coupling toward C2+products.

(3)

In H-type cells, CC coupling is enhanced on Cu nanostructured surfaces for two reasons: (1) presence of high local pH due to the operation at higher current den-sities and lower mass transport of OHbyproducts as compared to smooth sur-faces; (2) increased retention time of the intermediates within the nanostructures, which enables higher rate of intermediate re-adsorption.9,18,19On the other hand, recent studies have demonstrated the control of surface adsorbates coverage via varying COx(x = 1 or 2) feed concentrations.20,21These results have highlighted the importance of controlling the feed concentration, and thus local COx concen-trations, to attain an optimal surface coverage of adsorbates (*CO2, *CO, *H, and other intermediates) for CC coupling. For example, as ethylene is formed through CO dimerization via Langmuir–Hinshelwood mechanism, *CO coverage is a crucial factor in achieving selective ethylene production.20,22Conversely, the effect of *CO2(a precursor of *CO) coverage on CC coupling has not been inves-tigated in detail. A high local CO2concentration (local [CO2]), i.e., CO2 concentra-tion at or near the catalyst surface could lead to a higher populaconcentra-tion of unreacted *CO2, which may not be ideal for CO dimerization. A previous study utilizing mass-transport model coupled with micro-kinetic model has shown that under high applied potential, *CO2(including chemisorbed *CO2d) are still the major surface species at high CO2-availability conditions.23Hence, the poor C2+product selec-tivity in GDE-based systems with KHCO3electrolyte could be due to the subopti-mal adsorbate coverage from the high CO2 mass transport and, thus, high local [CO2].13,14Reducing the local [CO2] and thus *CO2coverage could increase the relative coverage of minority surface species, such as *CO that is beneficial for C–C coupling. The local [CO2] is directly related to the CO2surface coverage of the catalyst as follows:21

qCO2 = q

,½CO 2,e

ECO2

RT (Equation 1)

whereq* is the free surface site coverage; ECO2is the CO2adsorption energy;R is the

ideal gas constant; andT is the temperature.

Therefore, we propose that within the CO dimerization potential window, there is an intimate relation between local [CO2] and C–C coupling reaction. At a low local [CO2], hydrogen evolution reaction (HER) will dominate. However, at a high local [CO2], we expect to observe a suboptimal C2+product selectivity. Instead, C–C coupling is most ideal at a moderate local [CO2]. In order to prove the above hypoth-esis, we designed a systematic study that could effectively decouple local [CO2] from the current density and applied potential.

Here, we take advantage of the highly tunable system parameters within GDE-based flow electrolyzers to investigate the effect of local [CO2] on the C2+product selec-tivity. Guided by mass-transport modeling, we have identified three methods to control local [CO2] in our system (Figure 1): (1) catalyst layer structure, i.e., thickness and porosity, (2) CO2feed concentration, and (3) feed flow rate. Employing mono-dispersed Cu2O nanoparticles (NPs) as a model building block to construct the porous catalyst layers, we show that thicker catalyst layers exhibit lower local [CO2], which is beneficial for CO dimerization. Similarly, we demonstrated that feed flow conditions were also efficient parameters to modulate local [CO2] that led to effective control of reaction pathways. Therefore, these findings establish the modulation of local [CO2] as a general and convenient tool to achieve selective C2+production with GDE-based electrolyzers for CO2RR in a neutral electrolyte. In addition, our study highlights the substantial influence of CO2mass transport on the kinetics of CO2RR.

1Department of Materials Science and

Engineering, Korea Advanced Institute of Science and Technology (KAIST), 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of Korea

2Graduate School of Energy, Environment, Water

and Sustainability (EEWS), KAIST, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of Korea

3KAIST Institute for NanoCentury, Korea

Advanced Institute of Science and Technology (KAIST), Yuseong-gu, Daejeon 34141, Republic of Korea

4These authors contributed equally 5Lead Contact

*Correspondence:jihun.oh@kaist.ac.kr https://doi.org/10.1016/j.joule.2020.03.013

(4)

RESULTS AND DISCUSSION

Predicting Methods to Modulate Local CO2Concentration

Mass-transport modeling has been employed to estimate the local [CO2] and the local pH on catalyst surfaces in both H-type cells and GDE-based flow cells.9,13,15,24–26It is important to note that the one-dimensional numerical model used may not provide an accurate quantitative representation of the actual systems due to the simplistic assumptions involved. However, mass-transport models have shown to provide useful insights on the relative changes in the calculated con-centrations as a function of various parameters that match well with experimental measurements.25Based on our GDE system as shown in Figure 2A, we utilized such a mass-transport model to identify parameters that can affect the local [CO2] at constant current densities.

As the superior CO2mass transport in GDE-based systems originates from shorter diffusional lengths than that in conventional H-type cells (Figure S1), we predicted that increasing the porous catalyst layer thickness would increase diffusional mass transfer resistance and, thus, modulate local [CO2]. Note that by increasing the thick-ness while keeping catalyst loading constant, the porosity of the catalyst layer must increase, which actually provides better CO2mass transport. The effect of porosity changes relative to thickness is accounted for in our model by the inclusion of an effective diffusion coefficient term (Supplemental Experimental Proceduresand Fig-ure S2A). According to our mass-transport model, as the catalyst layer thickness varies from 1–5 mm, the local [CO2] gradient within the catalyst layer increases at all current densities, leading to an appreciable decrease in the average local [CO2] (Figures 2B and 2D). Though the changes in both catalyst layer porosity and catalyst layer thickness could affect CO2mass transport and thus local [CO2] in an inverse manner, our simulation result shows that the catalyst layer thickness plays a more significant role in controlling local [CO2].

Another intuitive way to regulate the local [CO2] is to control the gaseous CO2 con-centration at the gas-electrolyte-catalyst three-phase interface, which determines the saturated [CO2] atx = 0 and thus the maximum local [CO2] within the catalyst layer (Figure 2A and Equation S9). As expected, reducing the gaseous CO2 concen-tration from 100% to 20% caused a significant decrease in local [CO2] at various

Figure 1. Simplified Schematic Illustration of the Strategies to Modulate Local [CO2]

Proposed methods to modulate the local [CO2] within a GDE-based system: controlling catalyst

layer structure, CO2feed concentration, and feed flow rate. Dark gray, red, and blue balls represent

(5)

current densities (Figures 2C and 2E). In practice, the gaseous CO2concentration at the three-phase interface can be controlled in two ways: (1) adjusting the CO2 con-centration in the feed gas by introducing inert gas (e.g., N2), and (2) controlling the feed flow rate, which directly affects the gas boundary layer thickness and hence

Figure 2. Utilizing Mass-Transport Modeling to Predict Different Ways to Tune the Local CO2

Concentration

(A) Graphical representation of the boundaries used in our model. GDL, MPL, and CL are abbreviations for gas-diffusion layer, microporous layer, and catalyst layer, respectively. (B) Calculated local [CO2] as a function of relative position within the catalyst layer for various

catalyst layer thicknesses.

(C) Calculated local [CO2] as a function of relative position within the catalyst layer for various

gaseous CO2concentration at the three-phase interface.

(D) Calculated averaged local [CO2] as a function of catalyst layer thickness.

(E) Calculated averaged local [CO2] as a function of gaseous CO2concentration at the three-phase

interface. Note that a reasonable local [CO2] could not be obtained for gaseous CO2concentration

of 20% and 40% at 300 mA cm2due to the computed negative [CO2] values within the catalyst

layer. In a practical scenario, this suggests that CO2has been depleted partially within the catalyst

(6)

controls the diffusional transport of CO2molecules from the bulk gas to the three-phase interface (Figure S3and Equation S13). In addition, gaseous CO2 concentra-tion at the triple interfaces can be diluted if we assume that the molar product formation rates and the CO2 consumption rate remain relatively constant when the feed flow rate (i.e., CO2molar flow rate) decreases. Therefore, we would expect a lower gaseous CO2 concentration at the three-phase interface at a lower feed flow rate, and vice versa.

From the results of our mass-transport model, we predict that controlling the catalyst layer thickness and feed flow conditions can effectively modulate the local [CO2] and, thus, affect the CO2RR performance.

Controlling Catalyst Layer Structure

In this work, we demonstrate that the thickness and porosity of the catalyst layer can be controlled by regulating the evaporation rate of the catalyst ink. This can be achieved in two ways: (1) controlling the ink deposition rate using different deposition techniques and (2) adjusting the drying temperature of the ink. First, we investigated dropcasting, hand-painting, and airbrushing methods for the construction of catalyst layers on GDEs (Figure S4). Utilizing monodispersed Cu2O NPs with an average size of ~60 nm as building blocks (Figures 3A, 3B, and

S5), we constructed the catalyst layers on GDEs while keeping the catalyst loading constant at 0.6 mg cm2. The three catalyst layer deposition methods were per-formed on a hot plate at 50C as described in theExperimental Procedures. From the cross-sectional images in Figures 3C–3H, the airbrushed catalyst layer was 4 times thicker and more porous than the drop-casted and hand-painted layers (Table S1). The fast evaporation rate of solvent in the airbrushing method facilitates formation of a well-defined, thicker catalyst layer due to instantaneous fixation of scattered ink mists on the solidified catalyst layers (Figure S4C). For the dropcasted and hand-painted catalyst layers, their thickness were even lower than that of the hypothetical film of non-porous Cu2O, which implied that a portion of the catalyst ink permeated through the microporous layer (MPL) because of the slow evapora-tion rate of solvent. On the other hand, the dropcasting method resulted in a signif-icantly more uneven catalyst layer than those of hand-painting and airbrushing methods (Figures S6andS7). Importantly, after CO2RR, the OD-Cu catalyst layer thickness of the airbrushed samples was still substantially thicker than those prepared by dropcasting and hand-painting methods (Figures S6G–S6L andS8). Therefore, catalyst layer structures assembled by nanoparticulate building blocks can be controlled rationally by various deposition techniques.

Utilizing a home-built flow electrolyzer (Figure S9), the attained electrochemical measurements (Figure 4) clearly show the excellent performance of the airbrushed cathode in selective formation of C2+products compared to those of the dropcasted and hand-painted cathodes. In detail, the C2+product FE in the airbrushed cathode is 30%, 45%, and 55% at 100, 200, and 300 mA cm2, respectively (0.75, 0.82, and 0.88 V versus reversible hydrogen electrode [RHE]). On the other hand, both dropcasted and hand-painted cathodes achieved similar C2+ product FEs from 9% to 34% as the current densities were increased from 100 to 300 mA cm2, though the potential required for the dropcasted cathode was more negative than the hand-painted one. In contrast, the difference in FE for C1products formed by each cathode was less obvious (Figures 4A and 4B). With a closer inspection, we noticed a slightly higherjC1for the hand-painted cathode than that for the airbrush

(7)

cathode at the potential range, while the dropcasted cathode exhibited the lowest jC1(Figure 4C).

The above results are mostly consistent to our hypothesized expectations. Due to the thicker catalyst layer of the airbrushed cathode, the modulated local [CO2] should provide a more optimal surface concentration of *CO2, *CO, and *H for the promotion of CO dimerization. On the other hand, the thinner catalyst layers of dropcasted and hand-painted cathodes should yield less C2+products due to a higher local [CO2] and a suboptimal *CO coverage. However, this prediction was not observed for the dropcasted cathode, as it generated the least C1 and C2+products, which indicates that it was operating in a CO2-deficient environment.

Figure 3. Physical Characterization of Catalyst Layers Assembled by Cu2O NPs via Various

Deposition Techniques

(A and B) TEM image (A) and SEM image (B) of as-synthesized Cu2O nanoparticulate building

blocks.

(C–H) Cross-sectional SEM images and their energy-dispersive X-ray spectroscopy (EDS) elemental mappings of Cu for catalyst layers prepared by dropcasting ([C] and [D]), hand-painting ([E] and [F]), and airbrushing ([G] and [H]). The high porosity of the airbrushed sample is evident from the presence of macropores as denoted by the arrows in (G). Note that the lumps of Cu2O

nanoparticles present at the MPL layer in (G) were due to the partial collapsing of the Cu2O layer

during the sample preparation process for the cross-sectional characterization. The collapsing of the Cu2O layer was more evident for the airbrushed sample due to the much thicker catalyst layer,

(8)

We believe that this is attributed to the non-uniform catalyst layer, which provided a poorly defined gas-electrolyte-catalyst interface and, thus, over-limited CO2 availability.27This could also explain the higher CH4FE for the dropcasted cathode than other cathodes, which is due to the combined effect of higher applied potential

Figure 4. Effect of Catalyst Deposition Methods on CO2RR Performance

(A) Faradaic efficiencies of the major products and the applied potential at different current densities for various catalyst deposition methods.

(B) Faradaic efficiencies of C1and C2+products as a function of applied potential for various

catalyst deposition methods.

(C) Partial current densities of C1and C2+products as a function of applied potential for various

catalyst deposition methods. For clarity, formate, methane, and carbon monoxide are grouped together as C1products in red; ethanol, ethylene, and other multi-carbon products (1-propanol and

acetate) are grouped together as C2+products in blue.

When error bars are presented, data are represented as meanG standard deviation of at least 3 individual experiments.

(9)

and a higher *H coverage that promotes surface recombination of *H and *CO as reported previously.22Additionally, the intersection between jC2+profiles for the hand-painted cathode and the airbrushed cathode around 0.75 V versus RHE (Figure 4C) suggests that at the low potential, local [CO2] in a catalyst layer does not play a significant role in enhancing activity in the formation of C2+products. Local pH has been reported substantially as an important parameter in affecting the C–C coupling in CO2RR.9,15,18Here, the local pH is not a contributing factor accord-ing to our mass-transport modelaccord-ing simulation as the pH in a catalyst layer is pre-dicted to decrease slightly with increased thickness of the catalyst layer at a constant current density; more specifically, local pH decreased by approximately 0.4 when catalyst layer thickness increased from 1 to 3.5 mm at 200 mA cm2(Figure S10). This inverse relation between local pH and catalyst layer thickness is counterintuitive due to numerous reports on the presence of high local pH for catalysts with rough surfaces and nanostructures.9,18,19However, the higher pH in these reports origi-nates from the higher current density (i.e., higher production rate of OH) that the catalysts can achieve with a higher electrochemical active surface area. On the other hand, our simulation result presented in Figure S10 was calculated based on a constant current density, and thus a constant production rate of OH (mol s1). Therefore, with a higher volume of the reaction zone in the thicker catalyst layers, the production rate of OHper volume (mol m3s1) and thus local concentration of OH(mol m3) will be lower than that in the thinner catalyst layers. Hence, in our system, we have demonstrated that local [CO2] was the dominant factor in promoting C–C coupling. Nonetheless, we would expect that further enhancement of C2+production could be achieved by optimizing both local CO2concentration and local pH.

We further studied the influence of local [CO2] in a catalyst layer on C–C coupling reaction by controlling the evaporation rate of the catalyst ink with different temperature, using the airbrushing method. Cu2O catalyst ink was airbrushed on carbon papers at 25C, 50C, and 75C, which resulted in the catalyst layer thickness of 1.94, 3.59, and 4.59 mm, respectively (Figures S11andS12;Table S2). This obser-vation agrees well with the aforementioned statement that a fast evaporation rate of the catalyst ink promotes the formation of a thicker catalyst layer.

Consistent to the result discussed above, we have also observed similar patterns in selectivity for C2+products as a function of catalyst layer thickness, i.e., deposition temperature (Figure S13). The C2+product FE was the lowest for the cathode pre-pared at 25C due to the thinnest catalyst layer, thus experiencing the highest local [CO2]. In contrast, C2+product FE was the highest for the 75C cathode as it had the thickest catalyst layer, which experienced the most optimal (moderate) local [CO2]. Controlling Feed Flow Conditions

Adjusting feed flow conditions provides an avenue to tune the local [CO2]. Here, em-ploying airbrushed Cu2O GDE samples prepared at 50C, we controlled the CO2 concentration fed to our GDE-based flow electrolyzer from 25% to 100% (N2 as the inert carrier gas) at a constant feed flow rate of 20 sccm.

InFigure 5, as the CO2concentration of the feed gas reduced to 50%, the applied potentials required to achieve each current density (100 and 200 mA cm2) were similar. However, as the CO2feed concentration was reduced to 25%, the applied potential required to attain 200 mA cm2increased. At the same time, HER was acti-vated, which indicated a lack of CO2availability, i.e., low local [CO2].

(10)

Interestingly, at 200 mA cm2, as the feed CO2concentration was diluted from 100% to 50%, the C2+product FE increased substantially from 44.9% to 55.5% while the C1 product FE was reduced from 38.2% to 23.6%. This observation indicated that more intermediate *CO species were utilized to form C2+ products via dimerization instead of desorbing as CO when the feed gas CO2concentration decreased. The higher rate of C–C coupling can be attributed to the more optimal coverage of *CO, *H and other intermediate species due to the lower (more moderate) local [CO2] and thus lower *CO2surface coverage. With a closer observation, the C2+ product FE was similar for both 50% and 75% CO2 feed, but H2FE was slightly increased in the more diluted CO2feed. This indicated that the local [CO2] for the 50% CO2feed was near the borderline region between low and moderate levels. Consequently, it is important to be aware of the plateau of C2+product FE when modulating the local [CO2]. Reducing the CO2feed concentration further would decrease C2+product FE and increase H2FE as observed for 15% and 25% CO2 feed (Figures 5andS14).

However, at 100 mA cm2, the C2+ product FE remained almost constant while the C1product FE was positively correlated for all CO2feed concentration. Similar to the observations reported above, C–C coupling rate is more sensitive to the local

Figure 5. Effect of CO2Feed Concentration on CO2RR Performance at 20 sccm Feed Flow Rate

(A) Faradaic efficiencies of the major products and the applied potential at various CO2feed

concentration at different current densities.

(B and C) Partial current densities of C1products (B) and C2+products (C) as a function of CO2feed

concentration.

When error bars are presented, data are represented as meanG standard deviation of at least 3 individual experiments.

(11)

[CO2] at higher current density even though the potential is above the onset poten-tial for C2+products. This observation could be related to the sparser *CO coverage at lower potential, thus, a reduction in *CO2 coverage does not assist in CO dimerization.28

A lower gaseous CO2concentration at the gas-electrolyte interface and thus lower local [CO2] would also lead to a slower rate of neutralization between CO2 and OHproduced during CO2RR. As a result, the local pH is predicted to be slightly higher; local pH increases by approximately 0.2 as gaseous CO2concentration re-duces from 100% to 50% at 200 mA cm2(Figure S15A). The presence of a higher local [OH] is likely to play a role in promoting CC coupling.15However, we expect the contribution by the slight change in local pH to be less significant than that of local [CO2] as discussed previously (Figure S10).

Next, we investigated the effect of feed flow rate on the CO2RR performance while using a 100% CO2feed. At 200 mA cm2, as the feed flow rate decreased from 40 to 5 sccm, the C2+ product FE increased exponentially to 60.3% while the C1product FE exhibited the opposite trend (Figure 6). At the same time, the applied potential and H2FE remained relatively constant, which indicated sufficient

Figure 6. Effect of Feed Flow Rate on CO2RR Performance with 100% CO2Feed

(A) Faradaic efficiencies of the major products and the applied potential at various CO2feed flow

rates at different current densities.

(B and C) Partial current densities of C1products (B) and C2+products (C) as a function of feed flow

rate.

When error bars are presented, data are represented as meanG standard deviation of at least 3 individual experiments.

(12)

CO2availability even at the lowest feed flow rate, i.e., local [CO2] was still at the moderate level at 5 sccm feed flow rate. We expected that further reducing the feed flow rate would shift the local [CO2] to a lower level, which would result in a decrease in C2+product FE and an increase in H2FE as observed in a previous report.29

The enhanced selectivity for multi-carbon products was attributed to a lower gaseous CO2 concentration at the gas-electrolyte-catalyst three-phase interface, which was regulated by the feed flow rate in two ways. First, the feed flow rate can affect the local [CO2] due to the change in the gas boundary layer thickness (

Fig-ure S3). Second, as the feed flow rate decreases, CO2 concentration within the gas chamber becomes more diluted by the gaseous products. Based on the product FEs obtained at each flow rate, we calculated the gaseous CO2concentration at the outlet stream and thus estimated an average CO2concentration within the electro-lyzer gas chamber (Figure S15B). The average bulk CO2 concentration clearly showed an inverse-exponential relation with the feed flow rate, which could also explain why the observed rise in C2+product FE was the most significant at the lowest flow rate. Another possible contributing factor for the enhanced of C–C coupling could be ascribed to the higher probability of re-adsorption of the prema-turely desorbed CO molecules for further dimerization.30,31This higher rate of re-adsorption of CO is a result of longer retention time within the gas chamber and a higher CO concentration at the gas-electrolyte-catalyst three-phase interface. Comparing the feed flow rate normalized to the geometric electrode area, our optimal feed flow rate (100% CO2) of 2.5 sccm cm2is substantially lower than what was reported for GDE-based flow electrolyzers (7–50 sccm cm2).13,14,17 There-fore, our result suggested that the flow conditions from past literatures were mostly unoptimized for C–C coupling, thus explaining their poor C2+product FE in neutral electrolytes.

We further explored the effect of CO2feed concentration at various flow rates on CO2RR performance (Figures S16 and S17). The C2+ product FE was found to peak at 61.9% at 75% CO2and feed flow of 10 sccm at 200 mA cm2(Figure 7A), which was a significant improvement from 44.9% (or 25.4% for dropcasted samples) at the default flow conditions. In addition to providing a favorable adsorbate coverage for CO dimerization, modulating the supply of CO2is also an effective way to increase the CO2conversion rate as long as CO2availability is not compro-mised. In our system, the single-pass CO2conversion was enhanced from 7.3% (or 5.9% for the dropcasted samples) to 22.6% by reducing the pure CO2feed flow from 20 to 5 sccm (Figure 7B). Therefore, with the rational system control, it is possible to simultaneously achieve both high C2+product selectivity and CO2 con-version rate, which are critical factors to minimize operating and energy costs incurred from product purification.2,32However, achieving high CO

2conversion at industrially relevant conditions is rarely reported or discussed and thus should deserve more attention in the future.29,33,34

In general, we have demonstrated experimentally that having a moderate local [CO2] is important in enhancing C2+product FE in CO2RR (Figure S18). A low local [CO2] would reduce all CO2RR products and increase HER (Figure S18A). On the other hand, a high local [CO2] would give a suboptimal C2+product FE and, interestingly, a higher C1production. The high C1product FE has been consistently observed throughout our studies above; more specifically, when a high local [CO2] was present in the thin catalyst layer, 100% CO2feed concentration and a high feed

(13)

flow rate. We believe that in high local [CO2] environment, the surface coverage of *CO2 is high, which could physically hinder the surface recombination of C–C coupling (Figure S18C). Alternatively, the high surface coverage of *CO2 could also contribute to repulsive adsorbate-adsorbate interactions with *CO, which reduces the binding energy of *CO.36 In either case, unreacted *CO desorbs prematurely, which thus leads to a higher C1 product FE. Nonetheless, these hypotheses require rigorous in situ surface characterizations to further provide experimental evidences on the relationship between CO2mass transport and the molecular surface coverage.25,37

Apart from promoting CO dimerization, we also observed different ethylene and ethanol selectivity characteristics when we regulated the feed flow conditions (Figure S19). This phenomenon could indicate different product-specific sites on OD-Cu involved for the two C2products as reported by Lum and Ager.38Hence, GDE-based systems have the potential to provide new mechanistic insights toward achieving high product selectivity in the future.

Further Performance Optimization and Stability

With the aim of further promoting C2+product FE performance, we investigated the effect of Cu2O particle size and Nafion binder loading (Figures S20–S22). Remarkably, at the optimized conditions (airbrushed at 50C, 180 nm Cu2O particle

Figure 7. Optimizing C–C Coupling in CO2RR

(A and B) Colored contour map of C2+product faradaic efficiency (A) and single-pass CO2

conversion (B) as a function of CO2feed concentration and feed flow rate at 200 mA cm2. The peak

values are denoted by a star (white star) and the performance values at default conditions are denoted by a circle (B) in the respective contour maps. The data points used to generate the contour maps are tabulated inTable S3.

(C) Comparison of C2+product FE versus current density of our OD-Cu optimized in this work with

other reported Cu catalysts in 1 M KHCO3; porous OD-Cu,13sputtered Cu and pyridine-polymer/

Cu,17and OD-Cu gauze.35

(D) Stability test performed at a constant current density of 200 mA cm2using airbrushed and dropcasted Cu2O samples with both optimized conditions (12 sccm, 75% CO2, 2% Nafion) and

(14)

size, 2% Nafion loading), the C2+product FE was further improved to 75.5% at 300 mA cm2. At an even higher current density of 500 mA cm2, the C2+product FE could maintain at a relatively high level of 68.4% (jC2+: 342 mA cm2). These results represent one of the best C2+product FE ever reported at high current density in KHCO3electrolyte (Figure 7C;Table S4). To the best of our knowledge, the only report that exhibits better C2+product selectivity relies on pyridine-based additives to tune the *CO binding energy of Cu surfaces. Therefore, we foresee that C2+ product FE could be further improved by both utilizing appropriate additives/dop-ants and moderating an optimal local [CO2].

We evaluated the stability of our GDE-based electrolyzer equipped with the as-pre-pared Cu2O catalyst layers at a fixed current density of 200 mA cm2in various flow conditions. The airbrushed samples displayed very stable C2+product FEs over 10 h with minimal rise in H2production and applied potential (Figures 7D andS23). As expected, the airbrushed sample with optimized conditions (12 sccm, 75% CO2, 2% Nafion) exhibited a much higher C2+product FE of 67%–69% than that with the default conditions. On the other hand, dropcasted Cu2O samples did not exhibit such stability; C2+product FEs were reduced drastically while H2 FEs increased to more than 65% within 5 h of measurement. The poor stability was a result of flood-ing in the gas chamber, which significantly limited the CO2availability. We posit that the deep infiltration of Cu2O NPs into the MPL and gas diffusion layer (GDL) dur-ing the deposition procedure could be responsible for the increased hydrophilicity of the carbon paper during the electrolysis. Even though adjusting the flow condi-tions could increase the C2+product FE for the dropcasted samples, our result also highlights the importance of proper catalyst deposition techniques that directly affect both product selectivity and system stability.

CONCLUSIONS

In this work, we have employed a GDE-based flow electrolyzer to show the intimate relation between local [CO2] and C2+product selectivity during CO2RR. Guided by mass-transport modeling, we have identified three methods to control local [CO2] in our system: (1) catalyst layer structure, i.e., thickness and porosity, (2) CO2feed concentration, and (3) feed flow rate. Specifically, increasing the catalyst layer thickness and decreasing the CO2feed concentration and feed flow rate can all reduce the average local [CO2]. Contrary to common intuition, this study shows that providing a maximal CO2transport can lead to a suboptimal C2+product FE. Instead, by modulating the local [CO2], CC coupling can be significantly enhanced. Through further optimizations, we have achieved C2+product FE of up to 75.5% at 300 mA cm2and a C2+product partial current density of up to 342 mA cm2. More importantly, we have demonstrated that even with the same catalyst, the kinetics of CO2RR can be substantially influenced by the modulation of CO2mass transport. Consequently, this study presents a rational approach to evaluate the full potential of any catalyst by providing an optimal local environment in a GDE-based neutral flow electrolyzer for selective multi-carbon production.

Although our work provides a clear relationship between the local [CO2] and C2+ prod-uct FE, the quantitative effect of local [CO2] on the surface coverage of intermediate spe-cies remains elusive. We also acknowledge that the C2+product FEs could be sensitive to other experimental parameters, e.g., electrochemically active surface area, the de-gree of copper-carbon interactions (due to the different infiltration depth of catalyst par-ticles), etc. In this respect, future mechanistic and kinetic studies could be performed in a more controlled environment that can provide a large tunable range of local [CO2], such

(15)

as utilizing a pressurized H-type cell.39Future work should also focus on integratingin situ spectroscopy techniques with GDE-based systems to quantify the surface molecular species, which will provide greater insights on the fundamental effects of the actual com-plex local environment.25,37By further coupling spectroscopy data with micro-kinetic modeling,40,41one could even discover a rational strategy to progress toward the theo-retical maximum C2+product FE.

EXPERIMENTAL PROCEDURES

Mass-Transport Simulation

A mass-transport model, which takes into account the electrochemical reactions involved in CO2RR, HER, and acid-base reactions (CO2-HCO3-CO32equilibrium), is used in this work to estimate the local [CO2] and local pH. This model is based on previous CO2RR studies utilizing H-type cells and GDE-based systems.9,13,15,24–26 SeeSupplemental Informationfor more details of the simulation model.

Synthesis of 60 nm Cu2O Nanoparticles

The synthetic procedure was modified from a previous report.42 Generally, under ambient conditions, 0.80 mL of 0.50 M Cu(Cl)2and 0.40 g polyvinylpyrrolidone (PVP, MW: 40,000) were mixed well in 160 mL of deionized (DI) water. Next, 10 mL of 0.20 M NaOH was added dropwise (6.67 mL min1) to the solution under vigorous stir-ring. After stirring for 5 min, 10 mL of 0.1 M ascorbic acid was added dropwise (2 mL min1). The resultant suspension was left to react under stirring for another 5 min. Finally, the Cu2O NPs were collected, centrifuged, and washed for 4 cycles with ethanol before drying in a vacuum oven overnight at 60C. SeeSupplemental Experimental Procedures

for the synthesis procedures of 180 nm Cu2O and 1 mm Cu2O particles. Preparation of Catalyst Layer on GDEs

Cu2O/GDE catalyst electrodes were fabricated by deposition of Cu2O NPs via various techniques on the MPL of a commercial carbon paper (GDL-39BC, SIGRACET). Dropcasting method: Cu2O NPs (16 mg) were dispersed in 2 mL of isopropanol and 100 mL of Nafion ionomer solution (5 wt %), and the solution was sonicated for 30 min. The as-prepared catalyst ink was deposited onto a 2.53 2.5 cm2carbon paper on a hot plate dropwise using a pipette until catalyst loading of 0.6 mg cm2was achieved. The deposited carbon paper was further dried at room temper-ature in vacuum for 24 h.

Hand-painting method: Cu2O NPs (16 mg) were dispersed in 2 mL of isopropanol and 100 mL of Nafion ionomer solution (5 wt %) and the solution was sonicated for 30 min. The as-prepared catalyst ink was deposited onto a 2.53 2.5 cm2carbon paper on a hot plate using a paint brush until catalyst loading of 0.6 mg cm2was achieved. The di-rection of the painting strokes was changed frequently (i.e., left to right, right to left, top to bottom, and bottom to top) to ensure uniformity of the catalyst layer. The deposited carbon paper was further dried at room temperature in vacuum for 24 h. Airbrushing method: Cu2O NPs (50 mg) were dispersed in 20 mL of isopropanol and 313 mL of Nafion ionomer solution (5 wt %), and the solution was sonicated for 30 min. The as-prepared catalyst ink was deposited onto a 53 5 cm2carbon paper on a hot plate with an airbrush gun (Infinity CR Plus 0.4 mm, Harder & Steenbeck) using N2gas at 0.6 bar. For uniform deposition and consistent evaporation of solvent, the ink was sprayed 10 cm away from a carbon paper at a constant sweeping speed (one hor-izontal sweep per one second). The discharging rate of catalyst ink was qualitatively adjusted so that seven sweeps can deposit one layer of the catalyst ink on the carbon

(16)

paper (Figure S24). After spraying each layer, 20 s waiting time was carried out to ensure complete evaporation of solvent before a subsequent spray process and it was repeated until catalyst loading of 0.6 mg cm2was achieved. The deposited carbon paper was further dried at room temperature in vacuum for 24 h.

Electroreduction of CO2in a GDE-Based Flow Electrolyzer

The electroreduction of CO2was carried out in a flow electrolyzer comprised of cathode and anode compartments and an anion-exchange membrane (AEM, Selemion mem-brane) placed between the two compartments (Figure S9). The geometric area of the working and counter electrodes was 2 cm2. For 3-electrode measurements, a Ag/ AgCl reference electrode with saturated KCl (RE-1B, EC-Frontier) and a NiFeMo plate as the anode were used. All the Cu2O-deposited cathodes have a total catalyst loading of 0.6 mg cm2, unless stated otherwise. During the electrolysis, 1.0 M KHCO3 electro-lyte was continuously supplied by a peristaltic pump (Masterflex L/S pump) from a 50 mL reservoir at a flow rate of 12 mL min1. For the stability test, a 500 mL reservoir was used instead. To the cathode GDE, CO2(99.999%) (and N2for controlling CO2feed concen-tration) was fed at a flow rate of 20 sccm (or otherwise stated in the figure or main text) using a mass flow controller (MFC KOREA) while the anode compartment with an oxygen vent was filled with the same electrolyte.

Chronopotentiometry measurements were conducted with an electrochemical workstation (VSP/VMP3B-5, Biologic). Before each measurement, a pre-reduction step was carried out for 10 min at 10 mA cm2. The chronopotentiometry electrolysis experiments were conducted for 30 min for each CO2RR measurement. Note that all the cathodic current densities reported in this work are presented in their absolute values. During the electrolysis, ohmic resistance was determined through electro-chemical impedance spectroscopy at 10 min intervals. At each time interval, the ohmic potential drop was manually compensated (80%iR-correction) by subtracting the product of current and solution resistance from the applied potential. We found that periodic solution resistance measurement is important as even a slight change in solution resistance over time would result a significant error in theiR compensated potential due to the large current involved during CO2RR electrolysis. Electrode potentials measured on the Ag/AgCl scale (EAg/AgCl) were converted to the RHE scale (ERHE) using the following equation:

ERHE = EAg=AgCl + 0:197 + 0:0591 3 pH

where the pH value used was based on the local pH calculated from our mass-trans-port model.13

The gaseous products were quantified every 10 min during the analyses using an on-line gas chromatography (INFICON, 2-channel 3000 Micro GC) equipped with a Molsieve 5A column and a Plot Q column coupled with thermal conductivity detector (TCD). Meth-anol, ethMeth-anol, and 1-propanol products were quantified using a headspace gas chroma-tography equipped with flame ionization detector (FID) as described in a previous report.43Formate and acetate products were quantified by high performance liquid chro-matography (HPLC) with a SUGAR SH1101 column (Shodex).

The FE for the formation of a given product was calculated as follows: FEið%Þ =ni,zQi,F

whereniis the number of moles of the product (from GC and LC analysis) based on the outlet flow rate (measured with an electronic flow meter)44;z

(17)

electrons required to generate the product (2 for carbon monoxide, formate, and hydrogen; 6 for methanol; 8 for methane and acetate; 12 for ethylene and ethanol; and 18 for 1-propanol);F is the Faraday’s constant (96,485 C mol1);Q is the amount of charge passed during the measurement. The specific product partial current den-sity was calculated by multiplying the product’s FE with the total current denden-sity. The single-pass CO2conversion was calculated as follows:

CO2Conversion ð%Þ = P

iyi,ni nCO2; feed

3 100

wherenCO2; feedis the number of moles of CO2fed into the electrolyzer (per unit time)

based on the inlet flow rate44;y

iis the stoichiometric coefficient for a given CO2RR product (1 for CO, methane, and formate; 2 for ethylene, ethanol, and acetate; and 3 for 1-propanol), ni is the number of moles of the CO2RR product, i.e., C-based product (per unit time) as detected from GC and LC analysis, based on the outlet flow rate (measured with an electronic gas flow meter).

SUPPLEMENTAL INFORMATION

Supplemental Information can be found online athttps://doi.org/10.1016/j.joule. 2020.03.013.

ACKNOWLEDGMENTS

The authors thank Dr. Joel W. Ager (Lawrence Berkeley National Laboratory) for providing the simulation codes to model CO2mass transport in H-type cells. The au-thors gratefully acknowledge the financial support provided by the Creative Mate-rials Discovery Program (NRF-2017M3D1A1040692), funded by the Korean govern-ment (Ministry of Science and ICT).

AUTHOR CONTRIBUTIONS

Y.C.T., K.B.L., and J.O. conceived the project, contributed to the analysis and dis-cussion on the data, and wrote the manuscript. Y.C.T. performed the mass-transport simulation. Y.C.T. and K.B.L. designed the experiments, synthesized and character-ized the materials, and performed the electrochemical tests. H.S. provided help and advice on the GDE-based flow electrolyzer setup. J.O. supervised and guided the whole project.

DECLARATION OF INTERESTS

The authors declare no competing interests. Received: January 20, 2020

Revised: February 24, 2020 Accepted: March 17, 2020 Published: April 13, 2020

REFERENCES

1.Birdja, Y.Y., Pe´rez-Gallent, E., Figueiredo, M.C., Go¨ttle, A.J., Calle-Vallejo, F., and Koper, M.T.M. (2019). Advances and challenges in understanding the electrocatalytic conversion of carbon dioxide to fuels. Nat. Energy4, 732–745.

2.Verma, S., Lu, S., and Kenis, P.J.A. (2019). Co-electrolysis of CO2and glycerol as a pathway to

carbon chemicals with

improved technoeconomics due to low electricity consumption. Nat. Energy4, 466–474.

3.Martı´n, A.J., and Pe´rez-Ramı´rez, J. (2019). Heading to distributed electrocatalytic conversion of small abundant molecules into fuels, chemicals, and fertilizers. Joule3, 2602– 2621.

4.Smith, W.A., Burdyny, T., Vermaas, D.A., and Geerlings, H. (2019). Pathways to industrial-scale fuel out of thin air from CO2electrolysis.

Joule3, 1822–1834.

5.Gao, D., Ara´n-Ais, R.M., Jeon, H.S., and Roldan Cuenya, B. (2019). Rational catalyst and electrolyte design for CO2electroreduction

towards multicarbon products. Nat. Catal.2, 198–210.

(18)

6.De Luna, P., Hahn, C., Higgins, D., Jaffer, S.A., Jaramillo, T.F., and Sargent, E.H. (2019). What would it take for renewably powered electrosynthesis to displace petrochemical processes? Science364, eaav3506.

7.Bushuyev, O.S., De Luna, P., Dinh, C.T., Tao, L., Saur, G., van de Lagemaat, J., Kelley, S.O., and Sargent, E.H. (2018). What should we make with CO2and how can we make it? Joule2, 825–832.

8.Jung, H., Lee, S.Y., Lee, C.W., Cho, M.K., Won, D.H., Kim, C., Oh, H.S., Min, B.K., and Hwang, Y.J. (2019). Electrochemical fragmentation of Cu2O nanoparticles enhancing selective C–C coupling from CO2reduction reaction. J. Am.

Chem. Soc.141, 4624–4633.

9.Lum, Y., Yue, B., Lobaccaro, P., Bell, A.T., and Ager, J.W. (2017). Optimizing C–C coupling on oxide-derived copper catalysts for

electrochemical CO2reduction. J. Phys. Chem.

C121, 14191–14203.

10.Weekes, D.M., Salvatore, D.A., Reyes, A., Huang, A., and Berlinguette, C.P. (2018). Electrolytic CO2reduction in a flow cell. Acc.

Chem. Res.51, 910–918.

11.Kim, B., Seong, H., Song, J.T., Kwak, K., Song, H., Tan, Y.C., Park, G., Lee, D., and Oh, J. (2020). Over a 15.9% solar-to-CO conversion from dilute CO2streams catalyzed by gold

nanoclusters exhibiting a high CO2binding

affinity. ACS Energy Lett.5, 749–757. 12.Zheng, T., Jiang, K., Ta, N., Hu, Y., Zeng, J., Liu,

J., and Wang, H. (2019). Large-scale and highly selective CO2electrocatalytic reduction on

nickel single-atom catalyst. Joule3, 265–278. 13.Lv, J.J., Jouny, M., Luc, W., Zhu, W., Zhu, J.J., and Jiao, F. (2018). A highly porous copper electrocatalyst for carbon dioxide reduction. Adv. Mater.30, e1803111.

14.Ma, S., Sadakiyo, M., Luo, R., Heima, M., Yamauchi, M., and Kenis, P.J.A. (2016). One-step electrosynthesis of ethylene and ethanol from CO2in an alkaline electrolyzer. J. Power

Sources301, 219–228.

15.Dinh, C.T., Burdyny, T., Kibria, M.G., Seifitokaldani, A., Gabardo, C.M., Garcı´a de Arquer, F.P., Kiani, A., Edwards, J.P., De Luna, P., Bushuyev, O.S., et al. (2018). CO2

electroreduction to ethylene via hydroxide-mediated copper catalysis at an abrupt interface. Science360, 783–787.

16.Verma, S., Hamasaki, Y., Kim, C., Huang, W., Lu, S., Jhong, H.-R.M., Gewirth, A.A., Fujigaya, T., Nakashima, N., and Kenis, P.J.A. (2018). Insights into the low overpotential

electroreduction of CO2to CO on a supported

gold catalyst in an alkaline flow electrolyzer. ACS Energy Lett.3, 193–198.

17.Li, F., Thevenon, A., Rosas-Herna´ndez, A., Wang, Z., Li, Y., Gabardo, C.M., Ozden, A., Dinh, C.T., Li, J., Wang, Y., et al. (2020). Molecular tuning of CO2-to-ethylene

conversion. Nature577, 509–513. 18.Song, H., Im, M., Song, J.T., Lim, J.-A., Kim,

B.-S., Kwon, Y., Ryu, S., and Oh, J. (2018). Effect of mass transfer and kinetics in ordered Cu-mesostructures for electrochemical CO2

reduction. Appl. Catal. B232, 391–396.

19.Ren, D., Fong, J., and Yeo, B.S. (2018). The effects of currents and potentials on the selectivities of copper toward carbon dioxide electroreduction. Nat. Commun.9, 925. 20.Wang, X., de Arau´jo, J.F., Ju, W., Bagger, A.,

Schmies, H., Ku¨hl, S., Rossmeisl, J., and Strasser, P. (2019). Mechanistic reaction pathways of enhanced ethylene yields during electroreduction of CO2-CO co-feeds on Cu

and Cu-tandem electrocatalysts. Nat. Nanotechnol.14, 1063–1070.

21.Li, J., Wang, Z., McCallum, C., Xu, Y., Li, F., Wang, Y., Gabardo, C.M., Dinh, C.-T., Zhuang, T.-T., Wang, L., et al. (2019). Constraining CO coverage on copper promotes high-efficiency ethylene electroproduction. Nat. Catal.2, 1124–1131.

22.Schreier, M., Yoon, Y., Jackson, M.N., and Surendranath, Y. (2018). Competition between H and CO for active sites governs copper-mediated electrosynthesis of hydrocarbon fuels. Angew. Chem. Int. Ed. Engl.57, 10221–10225. 23.Singh, M.R., Goodpaster, J.D., Weber, A.Z.,

Head-Gordon, M., and Bell, A.T. (2017). Mechanistic insights into electrochemical reduction of CO2over Ag using density

functional theory and transport models. Proc. Natl. Acad. Sci. USA114, E8812–E8821. 24.Singh, M.R., Kwon, Y., Lum, Y., Ager, J.W., and

Bell, A.T. (2016). Hydrolysis of electrolyte cations enhances the electrochemical reduction of CO2over Ag and Cu. J. Am.

Chem. Soc.138, 13006–13012.

25.Yang, K., Kas, R., and Smith, W.A. (2019). In situ infrared spectroscopy reveals persistent alkalinity near electrode surfaces during CO2

electroreduction. J. Am. Chem. Soc.141, 15891–15900.

26.Gupta, N., Gattrell, M., and MacDougall, B. (2006). Calculation for the cathode surface concentrations in the electrochemical reduction of CO2in KHCO3solutions. J. Appl.

Electrochem.36, 161–172.

27.Jhong, H., Brushett, F.R., and Kenis, P.J.A. (2013). The effects of catalyst layer deposition methodology on electrode performance. Adv. Energy Mater.3, 589–599.

28.Huang, Y., Handoko, A.D., Hirunsit, P., and Yeo, B.S. (2017). Electrochemical reduction of CO2

using copper single-crystal surfaces: effects of CO* coverage on the selective formation of ethylene. ACS Catal.7, 1749–1756. 29.Gabardo, C.M., O’Brien, C.P., Edwards, J.P.,

McCallum, C., Xu, Y., Dinh, C.-T., Li, J., Sargent, E.H., and Sinton, D. (2019). Continuous carbon dioxide electroreduction to concentrated multi-carbon products using a membrane electrode assembly. Joule3, 2777–2791. 30.Gurudayal, Perone, D., Malani, S., Lum, Y.,

Haussener, S., and Ager, J.W. (2019). Sequential cascade electrocatalytic conversion of carbon dioxide to C–C coupled products. ACS Appl. Energy Mater.2, 4551–4559. 31.Lum, Y., and Ager, J.W. (2018). Sequential

catalysis controls selectivity in electrochemical CO2reduction on Cu. Energy Environ. Sci.11,

2935–2944.

32.Jouny, M., Luc, W., and Jiao, F. (2018). General techno-economic analysis of CO2electrolysis

systems. Ind. Eng. Chem. Res.57, 2165–2177. 33.Endr}odi, B., Kecsenovity, E., Samu, A., Darvas,

F., Jones, R.V., To¨ro¨k, V., Danyi, A., and Jana´ky, C. (2019). Multilayer electrolyzer stack converts carbon dioxide to gas products at high pressure with high efficiency. ACS Energy Lett. 4, 1770–1777.

34.Ripatti, D.S., Veltman, T.R., and Kanan, M.W. (2019). Carbon monoxide gas diffusion electrolysis that produces concentrated C2

products with high single-pass conversion. Joule3, 240–256.

35.Zhang, J., Luo, W., and Zu¨ttel, A. (2019). Self-supported copper-based gas diffusion electrodes for CO2electrochemical reduction.

J. Mater. Chem. A7, 26285–26292. 36.Sandberg, R.B., Montoya, J.H., Chan, K., and

Nørskov, J.K. (2016). CO–CO coupling on Cu facets: coverage, strain and field effects. Surf. Sci.654, 56–62.

37.Handoko, A.D., Wei, F., Jenndy, Yeo, B.S., and Seh, Z.W. (2018). Understanding

heterogeneous electrocatalytic carbon dioxide reduction through operando techniques. Nat. Catal.1, 922–934.

38.Lum, Y., and Ager, J.W. (2019). Evidence for product-specific active sites on oxide-derived Cu catalysts for electrochemical CO2

reduction. Nat. Catal.2, 86–93. 39.Lamaison, S., Wakerley, D., Blanchard, J.,

Montero, D., Rousse, G., Mercier, D., Marcus, P., Taverna, D., Giaume, D., Mougel, V., and Fontecave, M. (2020). High-current-density CO2

-to-CO electroreduction on Ag-alloyed Zn dendrites at elevated pressure. Joule4, 395–406. 40.Goodpaster, J.D., Bell, A.T., and

Head-Gordon, M. (2016). Identification of possible pathways for C–C bond formation during electrochemical reduction of CO2: new

theoretical insights from an improved electrochemical model. J. Phys. Chem. Lett.7, 1471–1477.

41.Liu, X., Schlexer, P., Xiao, J., Ji, Y., Wang, L., Sandberg, R.B., Tang, M., Brown, K.S., Peng, H., Ringe, S., et al. (2019). pH effects on the electrochemical reduction of CO(2) towards C2

products on stepped copper. Nat. Commun. 10, 32.

42.Yec, C.C., and Zeng, H.C. (2012). Synthetic architecture of multiple core–shell and yolk– shell structures of (Cu2O@)nCu2O (n = 1–4) with

centricity and eccentricity. Chem. Mater.24, 1917–1929.

43.Huang, Y., Deng, Y., Handoko, A.D., Goh, G.K.L., and Yeo, B.S. (2018). Rational design of sulfur-doped copper catalysts for the selective electroreduction of carbon dioxide to formate. ChemSusChem11, 320–326.

44.Ma, M., Clark, E.L., Dalsgaard, S., Therkildsen, K.T., Chorkendorff, I., and Seger, B.J. (2020). Insights into the carbon balance for CO2

electroreduction on Cu using gas diffusion electrode reactor designs. Energy Environ. Sci. 13, 977–985.

수치

Figure 1. Simplified Schematic Illustration of the Strategies to Modulate Local [CO 2 ]
Figure 2. Utilizing Mass-Transport Modeling to Predict Different Ways to Tune the Local CO 2
Figure 3. Physical Characterization of Catalyst Layers Assembled by Cu 2 O NPs via Various
Figure 4. Effect of Catalyst Deposition Methods on CO2RR Performance
+4

참조

관련 문서

→ For the turbulent jet motion and heat and mass transport, simulation of turbulence is important.. in the mean-flow equations in a way which close these equations by

local stationary behavior in the case of nonholonomic subsidiary conditions become identical with those for holonomic constraints.. The cable is of uniform

With these results, we can suggest the hypothesis that increased cardiac output in the patients who administered ketamine increased muscle blood flow, and this is the

For this reason, in this study, comparing the traditional manufacturing process and relatively high temperature of exhaust gas heat and CO2 laser

Many other parts of method (e.g. sample matrix, CO2 in air, source of water used for your mobile phase) can change the pH of the mobile phase causing shifts in retention,

The seven steps can be organized in order of selecting local cultural resources of Fujian Province, selecting local design prototypes for character

Just as we achieved good control with a joint-based controller that was based on a linearizing and decoupling model of the arm, we can do the same for the Cartesian case.

Since the Ministry of Gender Equality and Family designated Iksan city as a Women Friendly City first in our country in 2009, 50 local organizations have been