+ All Categories
Home > Documents > Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons....

Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons....

Date post: 13-Mar-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
6
Engineering Cu surfaces for the electrocatalytic conversion of CO 2 : Controlling selectivity toward oxygenates and hydrocarbons Christopher Hahn a,b,1 , Toru Hatsukade a,1 , Youn-Geun Kim c , Arturas Vailionis d , Jack H. Baricuatro c , Drew C. Higgins a , Stephanie A. Nitopi a , Manuel P. Soriaga c , and Thomas F. Jaramillo a,b,2 a Department of Chemical Engineering, Stanford University, Stanford, CA 94305; b SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, Menlo Park, CA 94025; c Joint Center for Artificial Photosynthesis, Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, CA 91125; and d Geballe Laboratory for Advanced Materials, Stanford University, CA 94305 Edited by Jean-Michel Savéant, Université Paris Diderot, Paris, France, and approved April 10, 2017 (received for review November 16, 2016) In this study we control the surface structure of Cu thin-film catalysts to probe the relationship between active sites and catalytic activity for the electroreduction of CO 2 to fuels and chemicals. Here, we report physical vapor deposition of Cu thin films on large-format (6 cm 2 ) single-crystal substrates, and confirm epitaxial growth in the <100>, <111>, and <751> orientations using X-ray pole fig- ures. To understand the relationship between the bulk and surface structures, in situ electrochemical scanning tunneling microscopy was conducted on Cu(100), (111), and (751) thin films. The studies revealed that Cu(100) and (111) have surface adlattices that are identical to the bulk structure, and that Cu(751) has a heteroge- neous kinked surface with (110) terraces that is closely related to the bulk structure. Electrochemical CO 2 reduction testing showed that whereas both Cu(100) and (751) thin films are more active and selective for CC coupling than Cu(111), Cu(751) is the most selective for >2e oxygenate formation at low overpotentials. Our results demonstrate that epitaxy can be used to grow single- crystal analogous materials as large-format electrodes that provide insights on controlling electrocatalytic activity and selectivity for this reaction. carbon dioxide reduction | epitaxy | electrocatalysis | copper T he electrochemical reduction of CO 2 (CO 2 R) is a process that could couple to renewable energy from wind and solar to directly produce fuels and chemicals in a sustainable manner. However, developing catalysts is a major challenge for this re- action, and significant advances are needed to overcome the issues of poor energy efficiency and product selectivity. One reason for these issues is that there are a limited number of catalysts that can effectively convert CO 2 to products that require more than two electrons (>2e products), e.g., methane, methanol, ethylene, etc. (1, 2). Therefore, developing catalysts that are effective for CO 2 R to >2e products would greatly improve prospects for utilization, and such an endeavor requires a deeper understanding of the relevant surface chemistry. Out of the polycrystalline metals, Cu is the only one that has shown a propensity for CO 2 R to >2e products at considerable rates and selectivity (2, 3). To date, its uniqueness is reflected by how nearly all work on catalysts with improved activity and se- lectivity for >2e products is based on Cu (46). However, poly- crystalline Cu is not particularly selective toward any one >2e reduction product (7). Thus, it is critical to understand what active site motifs lead to this unique selectivity for further reduced products and to apply this knowledge to develop new materials with this electrocatalytic behavior. Single-crystal studies on Cu have shown that CO 2 R activity and selectivity are extremely sensitive to surface structure. In particu- lar, facet sensitivities for CC coupling are the most widely stud- ied, with experimental reports concluding that Cu(100) terraces and any orientation of step sites are more active and selective for CC coupling than Cu(111) (8). Similarly, single-crystal investigations of CO reduction have demonstrated low over- potential production of ethylene, the simplest CC coupled hy- drocarbon, on Cu(100), while observing no early onset of ethylene on Cu(111) (9). Theoretical studies have shown that this difference in activity could be due to lower kinetic barriers for CC coupling on Cu(100) vs. Cu(111) (10, 11). Whereas the aforementioned single-crystal studies have elucidated design motifs for the synthesis of surface-structure-engineered electrocatalysts with higher CC coupling selectivity and activity, much remains to be learned re- garding structure sensitivity for this reaction. Electrodes consisting of single crystals are notably small (<0.1 cm 2 ) to achieve a high degree of uniformity across the surface. Such small electrodes, however, pose challenges to identifying and quantifying all prod- ucts of the reaction. Also, single crystals that exhibit interesting catalytic behavior are difficult to integrate into devices. These challenges motivate implementing synthetic routes to larger- format electrodes for CO 2 R with surfaces analogous to that of single crystals and the compatibility to be incorporated into a device architecture. One method for growing single-crystal analogous materials is epitaxy, where an underlying single-crystal substrate is used to control the growth orientation of an overlayer via interfacial en- ergetics. For transition metals such as Cu, vacuum growth studies Significance Anthropogenic global warming necessitates the development of renewable carbon-free and carbon-neutral technologies for the future. Electrochemical CO 2 reduction is one such technol- ogy that has the potential to impact climate change by en- abling sustainable routes for the production of fuels and chemicals. Whereas the field of CO 2 reduction has attracted great interest, current state-of-the-art electrocatalysts must be improved in product selectivity and energy efficiency to make this pathway viable for the future. Here, we investigate how controlling the surface structure of copper electrocatalysts can guide CO 2 reduction activity and selectivity. We show how the coordination environment of Cu surfaces influences oxygenate vs. hydrocarbon formation, providing insights on how to im- prove selectivity and energy efficiency toward more valuable CO 2 reduction products. Author contributions: C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., S.A.N., M.P.S., and T.F.J. designed research; C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., and S.A.N. performed research; C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., S.A.N., M.P.S., and T.F.J. analyzed data; C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., S.A.N., M.P.S., and T.F.J. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. 1 C.H. and T.H. contributed equally to this work. 2 To whom correspondence should be addressed. Email: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1618935114/-/DCSupplemental. 59185923 | PNAS | June 6, 2017 | vol. 114 | no. 23 www.pnas.org/cgi/doi/10.1073/pnas.1618935114 Downloaded by guest on March 14, 2020
Transcript
Page 1: Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons. Results and Discussion Out-of-Plane Thin-Film Texture. Thin films of Cu were synthesized

Engineering Cu surfaces for the electrocatalyticconversion of CO2: Controlling selectivitytoward oxygenates and hydrocarbonsChristopher Hahna,b,1, Toru Hatsukadea,1, Youn-Geun Kimc, Arturas Vailionisd, Jack H. Baricuatroc, Drew C. Higginsa,Stephanie A. Nitopia, Manuel P. Soriagac, and Thomas F. Jaramilloa,b,2

aDepartment of Chemical Engineering, Stanford University, Stanford, CA 94305; bSUNCAT Center for Interface Science and Catalysis, SLAC NationalAccelerator Laboratory, Menlo Park, CA 94025; cJoint Center for Artificial Photosynthesis, Division of Chemistry and Chemical Engineering, CaliforniaInstitute of Technology, Pasadena, CA 91125; and dGeballe Laboratory for Advanced Materials, Stanford University, CA 94305

Edited by Jean-Michel Savéant, Université Paris Diderot, Paris, France, and approved April 10, 2017 (received for review November 16, 2016)

In this study we control the surface structure of Cu thin-filmcatalysts to probe the relationship between active sites and catalyticactivity for the electroreduction of CO2 to fuels and chemicals. Here,we report physical vapor deposition of Cu thin films on large-format(∼6 cm2) single-crystal substrates, and confirm epitaxial growth inthe <100>, <111>, and <751> orientations using X-ray pole fig-ures. To understand the relationship between the bulk and surfacestructures, in situ electrochemical scanning tunneling microscopywas conducted on Cu(100), (111), and (751) thin films. The studiesrevealed that Cu(100) and (111) have surface adlattices that areidentical to the bulk structure, and that Cu(751) has a heteroge-neous kinked surface with (110) terraces that is closely related tothe bulk structure. Electrochemical CO2 reduction testing showedthat whereas both Cu(100) and (751) thin films are more activeand selective for C–C coupling than Cu(111), Cu(751) is the mostselective for >2e− oxygenate formation at low overpotentials.Our results demonstrate that epitaxy can be used to grow single-crystal analogous materials as large-format electrodes that provideinsights on controlling electrocatalytic activity and selectivity forthis reaction.

carbon dioxide reduction | epitaxy | electrocatalysis | copper

The electrochemical reduction of CO2 (CO2R) is a processthat could couple to renewable energy from wind and solar to

directly produce fuels and chemicals in a sustainable manner.However, developing catalysts is a major challenge for this re-action, and significant advances are needed to overcome the issuesof poor energy efficiency and product selectivity. One reason forthese issues is that there are a limited number of catalysts that caneffectively convert CO2 to products that require more than twoelectrons (>2e− products), e.g., methane, methanol, ethylene, etc.(1, 2). Therefore, developing catalysts that are effective for CO2Rto >2e− products would greatly improve prospects for utilization,and such an endeavor requires a deeper understanding of therelevant surface chemistry.Out of the polycrystalline metals, Cu is the only one that has

shown a propensity for CO2R to >2e− products at considerablerates and selectivity (2, 3). To date, its uniqueness is reflected byhow nearly all work on catalysts with improved activity and se-lectivity for >2e− products is based on Cu (4–6). However, poly-crystalline Cu is not particularly selective toward any one >2e−

reduction product (7). Thus, it is critical to understand what activesite motifs lead to this unique selectivity for further reducedproducts and to apply this knowledge to develop new materialswith this electrocatalytic behavior.Single-crystal studies on Cu have shown that CO2R activity and

selectivity are extremely sensitive to surface structure. In particu-lar, facet sensitivities for C–C coupling are the most widely stud-ied, with experimental reports concluding that Cu(100) terracesand any orientation of step sites are more active and selectivefor C–C coupling than Cu(111) (8). Similarly, single-crystal

investigations of CO reduction have demonstrated low over-potential production of ethylene, the simplest C–C coupled hy-drocarbon, on Cu(100), while observing no early onset of ethyleneon Cu(111) (9). Theoretical studies have shown that this differencein activity could be due to lower kinetic barriers for C–C couplingon Cu(100) vs. Cu(111) (10, 11). Whereas the aforementionedsingle-crystal studies have elucidated design motifs for the synthesisof surface-structure-engineered electrocatalysts with higher C–Ccoupling selectivity and activity, much remains to be learned re-garding structure sensitivity for this reaction. Electrodes consistingof single crystals are notably small (<0.1 cm2) to achieve a highdegree of uniformity across the surface. Such small electrodes,however, pose challenges to identifying and quantifying all prod-ucts of the reaction. Also, single crystals that exhibit interestingcatalytic behavior are difficult to integrate into devices. Thesechallenges motivate implementing synthetic routes to larger-format electrodes for CO2R with surfaces analogous to that ofsingle crystals and the compatibility to be incorporated into adevice architecture.One method for growing single-crystal analogous materials is

epitaxy, where an underlying single-crystal substrate is used tocontrol the growth orientation of an overlayer via interfacial en-ergetics. For transition metals such as Cu, vacuum growth studies

Significance

Anthropogenic global warming necessitates the developmentof renewable carbon-free and carbon-neutral technologies forthe future. Electrochemical CO2 reduction is one such technol-ogy that has the potential to impact climate change by en-abling sustainable routes for the production of fuels andchemicals. Whereas the field of CO2 reduction has attractedgreat interest, current state-of-the-art electrocatalysts must beimproved in product selectivity and energy efficiency to makethis pathway viable for the future. Here, we investigate howcontrolling the surface structure of copper electrocatalysts canguide CO2 reduction activity and selectivity. We show how thecoordination environment of Cu surfaces influences oxygenatevs. hydrocarbon formation, providing insights on how to im-prove selectivity and energy efficiency toward more valuableCO2 reduction products.

Author contributions: C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., S.A.N., M.P.S., and T.F.J.designed research; C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., and S.A.N. performed research;C.H., T.H., Y.-G.K., A.V., J.H.B., D.C.H., S.A.N., M.P.S., and T.F.J. analyzed data; C.H., T.H.,Y.-G.K., A.V., J.H.B., D.C.H., S.A.N., M.P.S., and T.F.J. wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.1C.H. and T.H. contributed equally to this work.2To whom correspondence should be addressed. Email: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1618935114/-/DCSupplemental.

5918–5923 | PNAS | June 6, 2017 | vol. 114 | no. 23 www.pnas.org/cgi/doi/10.1073/pnas.1618935114

Dow

nloa

ded

by g

uest

on

Mar

ch 1

4, 2

020

Page 2: Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons. Results and Discussion Out-of-Plane Thin-Film Texture. Thin films of Cu were synthesized

have shown that the surface structure is often related to the bulkcrystallographic orientation (12–14), indicating that epitaxialgrowth can be used for engineering the surface structure of elec-trocatalysts. In particular, researchers have used physical vapordeposition (PVD) and molecular beam epitaxy to successfullygrow Cu epitaxially on Si and Al2O3 (12, 13, 15–17). It is worthnoting that Si has garnered interest as a cathode in photo-electrochemical (PEC) cells (18, 19), and developing syntheticmethods to engineer the surface structure of Cu on Si could beadvantageous for controlling the performance and selectivity ofPEC CO2R devices. In this study, we use electron-beam (e-beam)deposition to epitaxially grow large-format single-crystal analo-gous Cu thin films on Si and Al2O3 single crystals. After growth, acombination of X-ray pole figures and electrochemical scanningtunneling microscopy are used to correlate the bulk and in situsurface structures. After physical characterization, we use ourpreviously reported electrochemical flow cell design with highproduct detection sensitivity to examine the dependence of CO2Ractivity and selectivity on surface structure (7). Using these results,we confirm the dependence of C–C coupling on surface structure,and provide insights on surface motifs that govern selectivitybetween >2e− oxygenates and hydrocarbons.

Results and DiscussionOut-of-Plane Thin-Film Texture. Thin films of Cu were synthesizedon large-format (27 mm × 42 mm) Al2O3(0001), Si(100), andSi(111) substrates in a three-source PVD chamber according to theprocedures outlined in the Supporting Information. On Al2O3(0001),a Ti layer was used to increase adhesion of Cu to the oxide sub-strate. These procedures allow for the synthesis of Cu thin-filmelectrodes that are ∼2–3 orders of magnitude larger in geometricarea than those used in typical single-crystal electrochemistrystudies. X-ray diffraction (XRD) symmetrical scans were usedto determine whether the single-crystal substrates impact the out-of-plane growth orientation of Cu thin films. As shown in Fig. 1,samples grown on Ti/Al2O3(0001) show only the face-centeredcubic (fcc) Cu(111) peak (black), indicating that the Cu thin filmsare strongly textured in the <111> orientation out-of-plane with therelationship Cu{111} k Al2O3(0001). To confirm the role of thesingle-crystal substrate interface, Cu thin films were also growndirectly on Al2O3(0001). The almost identical X-ray diffractograms

(Fig. S1) show that the Ti binding layer does not influence theout-of-plane texture induced by the Al2O3(0001) substrate.These similarities in out-of-plane growth on Ti and Al2O3(0001)are analogous to a previous report showing that hexagonal close-packed (hcp) Ti grows in the <001> orientation on Al2O3(001),allowing for growth of fcc Cu(111) on the similarly close-packedTi(001) (12).In contrast, samples grown on Si(100) show only the fcc Cu(200)

peak, indicating that the Cu thin films are textured in the <100>orientation out-of-plane with the relationship Cu{100} k Si(100).The difference in growth direction on Si(100) compared withgrowth on Al2O3(0001) clearly demonstrates that the single-crystalsubstrates guide the growth direction, likely due to differences ininterfacial energy. The results on Si(100) are similar to previousreports that have shown Cu grows in the <100> orientation fromhydrogen-terminated Si(100) surfaces (13, 20, 21). These reportsshow that Cu thin-film growth on Si(100) proceeds through theformation of a silicide due to diffusion at the interface (21). UnlikeCu thin films on Al2O3(0001) and Si(100), samples grown onSi(111) show no peaks within the range of the X-ray diffractogram,indicating that samples are not oriented out-of-plane in a lowMiller index direction. Although it might be expected that Cugrows in the <111> direction on Si(111) due to similarities in thefcc and diamond (111) plane, researchers have demonstratedradically different textures for growth of Cu on Si(111) dependingon the growth conditions (13, 15, 21). A previous report showedthat the strain from the large lattice mismatch at the interfacebetween Si(111) and Cu can cause the Cu thin film to grow in thehigh Miller index <531> orientation (15). Therefore, whereasXRD symmetrical scans show that low Miller index films are tex-tured out-of-plane with the relationships Cu{111} k Al2O3(0001)and Cu{100} k Si(100), a different XRD analysis is necessary todetermine this relationship for Cu thin films on Si(111).

In-Plane Thin Film Texture. Whereas XRD symmetrical scans es-tablish the out-of-plane texture relationships, both out-of-planeand in-plane texture analyses are necessary to determine whetherthe Cu thin films grow epitaxially on Al2O3 and Si. To this end,X-ray pole figure analysis was conducted on Cu thin films onTi/Al2O3(0001), Si(100), and Si(111) (Fig. 2). An X-ray polefigure for Cu(111) on Ti/Al2O3(0001) shows sixfold symmetryfor the Cu(200) Bragg reflections, indicating both strong out-of-plane and in-plane texture and thus epitaxial growth on theTi/Al2O3(0001) substrate (Fig. 2A). There are six diffraction spotsin the pole figure instead of three because there are two discretesets of crystallites from twinning with an azimuthal angle of 60°apart. Twinning defects are common in the epitaxial growth offcc metals due to both growth accidents and grain encounters(16, 17). Similar to the Cu(200) pole figure for Cu(111) growthon Ti/Al2O3(0001), the Cu(111) pole figure for Cu(100) growthon Si(100) shows discrete Bragg reflections, indicating cube-on-cube epitaxial growth of Cu on the Si(100) substrate (Fig. 2B).Fourfold symmetry is observed for the Cu(111) Bragg reflectionswith an azimuthal angle of 90° apart, which is expected for aCu(100) single crystal. Whereas Cu thin films on Si(111) exhibitno diffraction intensity in symmetric scans, a Cu(111) pole figureclearly shows strong out-of-plane and in-plane texture indicatingepitaxial growth of Cu on Si(111) (Fig. 2C). The Cu(111) Braggpeaks have threefold symmetry with an azimuthal angle of 120°apart. To better understand the out-of-plane growth orientationof Cu thin films on Si(111), we constructed an orientation distri-bution function (ODF) with three X-ray pole figure scans se-quentially collected on the same sample for the Cu(111), (200),and (022) Bragg reflections (Fig. S2). Using the ODF, an inversepole figure was calculated, demonstrating that the majority ofdiffraction intensity parallel to Si(111) comes from the (751) plane(Fig. 2D). Therefore, Cu films on Si(111) grow predominantly inthe <751> direction with the growth relationship Cu{751} k Si(111).

Fig. 1. XRD symmetrical scans of Cu thin films on Al2O3(0001) and Si(100)compared with the Joint Committee on Powder Diffraction Standards Cureference pattern #00–004-0836.

Hahn et al. PNAS | June 6, 2017 | vol. 114 | no. 23 | 5919

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Mar

ch 1

4, 2

020

Page 3: Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons. Results and Discussion Out-of-Plane Thin-Film Texture. Thin films of Cu were synthesized

These results demonstrate that PVD can be used to epitaxiallygrow large-format electrodes in both low and high Millerindex orientations.

Comparison of Ideal Surface Structures. To investigate structure–activity relationships for CO2R, it is important to correlate thegrowth orientation of Cu thin films with the expected surfacestructure. According to XRD results, Cu thin films are epitaxi-ally grown in the <111>, <100>, and <751> orientations onAl2O3(0001), Si(100), and Si(111), respectively. These samples willhenceforth be discussed as Cu(111), (100), and (751) for sim-plicity. As shown in Fig. 3, the epitaxial Cu films are expected tohave either a flat (Fig. 3 A and B) or kinked (Fig. 3C) surfacestructure depending on the growth direction of the film. The flatCu(111) and Cu(100) surfaces (Fig. 3 A and B) are expected tohave ninefold and eightfold coordination for all atoms on theirsurfaces, respectively. Unlike Cu(111) and Cu(100), Cu(751) iskinked, so microfacet notation is used to determine the ratios ofthe low Miller index facets (111), (110), and (100) present in a(751) unit cell (Fig. S3). With this notation, Cu(751) can be de-scribed as Cu S-[11(111) + 42(110) + 21(100)], indicating that thereare one (111), two (110), and one (100) unit cells in a (751) unitcell (22). Using these ratios, Cu(751) can be visualized as akinked surface with narrow (110) terraces and a high density of(100) and (111) step sites. The presence of all fcc low Millerindex facets suggests a high degree of site heterogeneity on theCu(751) surface. This heterogeneity is clearly observed in Fig.3C, where the kinks in Cu(751) are shown to have sites with six-,seven-, and eightfold coordination. Although there are addi-tional sites with 10- and 11-fold coordination, these atoms areexpected to be subsurface and have a minimal impact onelectrocatalysis. In summary, the different copper surfaces canbe placed in the order Cu(751) ≤ Cu(100) < Cu(111) in termsof coordination number.

Electrochemical Scanning Tunneling Microscopy. To understand thecorrelation between bulk orientation and surface structure in

electrolyte solution, electrochemical scanning tunneling micros-copy (ECSTM) was used to examine the in situ surface structureof Cu(111), (100), and (751) thin films. For the ECSTM study, theCu(111) thin film was grown on Si(110) instead of Al2O3(0001) toallow for electrical back-contact to the sample. The differentsingle-crystal substrates yielded Cu films that are identical in tex-ture (Fig. S4). Each sample was immersed in 0.1 M HClO4 andscanned from the open-circuit potential to −0.76 V vs. reversiblehydrogen electrode (RHE) at 50 mV s−1 to reduce the surfaceoxide into Cu metal. Afterward, ECSTM images were collectedat −0.24 V vs. RHE. Additional details about these experiments canbe found in Supporting Information. Low-magnification ECSTMimages show the morphological similarities of the three Cu thin-film orientations (Fig. S5). A large-scale view of the Cu(111)surface (Fig. S5A) depicts multiple stacks of Cu layers that arerandomly interspersed from each other. The hexagonal topo-graphic unit in each stack shares similar corner internal angles,although a few edges appear irregular and almost rounded. Asimilar magnification view of the Cu(100) surface (Fig. S5B) showsstacks that are more interconnected than those of Cu(111). Alarge number of Cu layers still appear stacked on top of each otherbut, unlike their Cu(111) counterpart, the constituent units form aspiral ramp that implies the formation of screw dislocations duringgrowth. For Cu(751), only a nondescript film terrain is observedwith grains of various sizes that range from 20 to 60 nm (Fig. S5C).Similarly, all samples show evidence that epitaxy likely proceedsthrough either a Volmer–Weber or Stranski–Krastanov growthmechanism. This is consistent with previous studies that showepitaxy of Cu on Si proceeds through 3D growth (13, 16).At higher magnification, atomically resolved ECSTM images

(Fig. 4) show the in situ surface structures of Cu(111), (100), and(751) thin films. A high-resolution image of the Cu(111) surface(Fig. 4A) shows a threefold rotational symmetry that is indicativeof a well-ordered close-packed Cu(111) structure. Imaging ofCu(100) at various points in the spirals unveiled a square Cu(100)net with an interatomic distance of 0.27 ± 0.01 nm (Fig. 4B).These ECSTM results demonstrate that for Cu(111) and (100)thin films, the majority of the in situ surface structure is identicalto the bulk orientation (Fig. 2B and Fig. S4) and the ideal atomicmodels (Fig. 3 A and B). Throughout the several-hour duration ofthe measurements, the surface structure of the epitaxial Cu thin

Fig. 2. X-ray pole figures for (A) Cu on Ti/Al2O3(0001); Cu(200) intensitiesare shown, (B) Cu on Si(100); Cu(111) intensities are shown, and (C) Cu onSi(111); Cu(111) intensities are shown. (D) An inverse pole figure showshighest intensities for the (751) plane, indicating that Cu on Si(111) is pre-dominantly oriented in the <751> direction out-of-plane.

Fig. 3. Color-coded atomic models showing coordination numbers for the(A) Cu(111), (B) (100), and (C) (751) surfaces.

5920 | www.pnas.org/cgi/doi/10.1073/pnas.1618935114 Hahn et al.

Dow

nloa

ded

by g

uest

on

Mar

ch 1

4, 2

020

Page 4: Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons. Results and Discussion Out-of-Plane Thin-Film Texture. Thin films of Cu were synthesized

films remains stable. Cu(111) and (100) electrodes have beenshown to be impervious to surface reconstruction in alkaline solu-tion (23); polycrystalline Cu surfaces, however, tend to restructure,initially into (111) and later into (100) (24). This highlights theimportance of using surface-sensitive in situ tools to investigate therelationship between the surface and bulk structures.In contrast to Cu(111) and (100), an atomically resolved

ECSTM image of the surface of the Cu(751) thin film shows moreheterogeneity in structure (Fig. 4C). This image is marked by color-coded stippled circles to serve as a visual guide for the geometricrelationship of the Cu surface atoms. The terraces delimited byblue and white circles are relatively wide and narrow, respectively.Noteworthy features include: (i) the presence of bright and darkregions that are indicative of a highly stepped surface, (ii) theprevalence of (110) terraces, defined by interatomic distances of0.36 nm and 0.26 nm ± 0.01 nm along the [001] and [110] direc-tions, respectively, (iii) the variable width of the (110) terraces; interms of the number of atoms n, the span ranges from 2 (e.g., whitecircles at the bottom left of Fig. 4) to 7 (e.g., blue circles found atthe top), and (iv) the interruption, by a kink, of 2–4 atoms alongevery 20-atom edge; that is, about 10–20% of the steps are kinked.Whereas ex situ XRD pole figure measurements reveal the

predominance of the (751) structure in the bulk (Fig. 2D), in situECSTM data show that the (751) structure populates a minorfraction of the film surface (Fig. 4C). The step notation of Cu(751)is Cu(S)–[4(110) × (311)], a designation more instructive forstructure visualization than the condensed (hkl) notation; it ex-presses the fact that the surface is characterized by 4-atom-wide(110) terraces separated by kinked (311) steps. The vestiges of thekinked (311) steps that could be associated with the (751) adlatticeare marked by the broken-line arrow aligned with the vectordrawn on the schematic model on the right. The Cu(110) terracesdemarcated by the white circles are only either 2- or 3-atom-widealong the [110] direction. The missing atoms are most likely highlyundercoordinated kink sites that are rendered unstable in thepresence of the acidic 0.1 M HClO4 electrolyte. Because steps of(100) and (110) orientations and their respective fourfold andtwofold rotational symmetries are more easily discerned than the

(111) steps, the overall highly stepped surface is better describedas a composite of Cu(S)–[n(110) × (100)], where n varies from 2 to7. A comparison with an ideal (751) surface (Fig. 4 C and D)where n = 4 shows that this heterogeneity in terrace length leadsto a difference in the overall distribution of six-, seven-, andeightfold-coordinated sites. Nevertheless, in situ ECSTM confirmsthe existence of a kinked Cu surface with (110) terraces that isundercoordinated in comparison with Cu(111) and (100). Hence-forth, the bulk structure Cu(751) will continue to be used to ref-erence the kinked Cu sample unless details about the in situ surfacestructure are discussed.The aforementioned conclusions show that PVD can be used to

control the surface structure of large-format Cu electrodes, to gaindeeper insights into the activity and selectivity of Cu surfaces forCO2 electroreduction. Below, we describe our measurements ofcatalytic activity and selectivity for CO2 electroreduction for thesethree different surfaces with distinct coordination environmentsand site motifs.

Electrochemical CO2 Reduction Selectivity and Activity. After bulkand surface-structure characterization, Cu(111), (100), and (751)were tested for their CO2 reduction selectivity and activity withinour previously described electrochemical cell (7). The Cu thinfilms have an exposed geometric electrode area of 5.9 cm2,allowing for electrochemical testing of samples that are ∼2 ordersof magnitude larger in area than those used in typical single-crystalstudies. Each epitaxially grown Cu sample is tested using chro-noamperometry (CA) at a single potential for 1 h in CO2-purged0.1 M KHCO3 electrolyte. The gas- and liquid-phase productswere quantified using a combination of gas chromatography andNMR spectroscopy, respectively. Current efficiencies for detectedproducts (Fig. S6) indicate that Cu(111), Cu(100), and Cu(751)thin films all make >2e− reduction products in the tested potentialrange between −0.89 and −1.10 vs. RHE. These products includehydrocarbons such as CH4 and C2H4, and various oxygenates suchas carbonyls and alcohols that are typical products for both single-crystal and polycrystalline Cu within this potential range (7, 8). Aplot comparing the total CO2 reduction rate as a function of ap-plied potential confirms that CO2 mass transport is not limitedwithin this potential range for the three Cu surfaces (Fig. S7).To examine the C–C coupling selectivity of the different Cu

surfaces, the current efficiencies for >2e− products are groupedby the number of carbons within a given product into the cate-gories C1, C2, and C3 (Fig. 5). The 2e− CO2R products, CO andHCOO−, are excluded from the analysis because on Cu surfacesCO is an intermediate for all >2e− C1, C2, and C3 products, andHCOO− is considered to be a terminal 2e− pathway formedthrough a different mechanism than that of CO (8, 25). At −0.89and −0.97 V vs. RHE, Cu(100) and Cu(751) are clearly moreselective for C2 and C3 products than Cu(111), indicating that flat(100) and kinked surfaces are more selective for C–C coupling atlower overpotentials. In addition, larger partial current densitiesfor C2 and C3 products from Cu(100) and Cu(751) demonstratethat this improvement in selectivity over Cu(111) is primarily dueto an increase in the absolute rate of C–C coupling rather thansimply a decrease in C1 activity (Fig. S8). These conclusions aresimilar to those drawn from single-crystal electrochemistry ex-periments that showed higher C2H4/CH4 ratios for surfaces with(100) facets and/or step sites with any microfacet orientation (8).Also, a recent report shows the same trends in C–C coupling se-lectivity for CO electroreduction on Cu(111) and Cu(100) singlecrystals (9). Theoretical studies indicate that this difference inselectivity could be due to a lower kinetic barrier for CO di-merization on Cu(100) compared with Cu(111) (10). Our resultssuggest that there is a strong correlation between C–C couplingselectivity and the coordination number of the surface, becauseCu(111) is more coordinated than Cu(100) and Cu(751) (Figs. 3and 4). Although it is clear that C–C coupling is favored on more

Fig. 4. Atomically resolved in situ ECSTM images of (A) Cu(111), (B) Cu(100),and (C) Cu (751) thin films. (D) An ideal atomic model of the Cu(751) surfaceis used to compare step orientations.

Hahn et al. PNAS | June 6, 2017 | vol. 114 | no. 23 | 5921

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Mar

ch 1

4, 2

020

Page 5: Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons. Results and Discussion Out-of-Plane Thin-Film Texture. Thin films of Cu were synthesized

undercoordinated surfaces, further work will be necessary to elu-cidate how the geometry of kinked surfaces affects kinetic barriers.Nevertheless, these results demonstrate that PVD can be usedto surface-structure-engineer large-format electrodes that haveanalogous electrocatalytic properties to Cu single crystals.Among the >2e− products, it is useful to compare selectivity

toward oxygenates and hydrocarbons to help understand ele-ments of the reaction mechanism, because oxygenates are lessreduced than hydrocarbons with the same number of carbons.Examining the product distribution in this manner also assessespossibilities for developing catalysts with desired selectivity for anumber of important chemical products. In particular, achievinggreater selectivity toward oxygenates is of interest because hydro-carbons with the same number of carbons are typically less valu-able (26). By grouping the current efficiencies for >2e− productsinto the categories hydrocarbons and/or oxygenates (Fig. S6), it isclearly shown that Cu(111), (100), and (751) are all more selectivefor hydrocarbons than oxygenates across the measured range ofpotentials (Figs. S9 and S10). In particular, high selectivity for thehydrocarbons CH4 and C2H4 has been commonly observed for Cuof any surface orientation (8). Whereas all epitaxial Cu thin filmsshow higher selectivity toward hydrocarbons, examining the po-tential dependence of oxygenate/hydrocarbon ratios can aid indetermining which active site motifs lead to greater oxygenateproduction (Fig. 6). At the highest potential of −0.89 V vs. RHE,both Cu(751) and Cu(100) have higher oxygenate/hydrocarbonratios than that of Cu(111), which makes only hydrocarbons at thispotential. Whereas both Cu(751) and (100) are both selective andactive at this potential for C–C coupling, the higher oxygenate/hydrocarbon ratio for Cu(751) than (100) indicates a clear dis-tinction between the surfaces in oxygenate selectivity. A compar-ison of Cu(751) and (100) indicates that the primary distinctions insurface structure on the former are the step sites and narrow (110)terraces, which lead to differences in geometry and a lower av-erage coordination number (Figs. 3 and 4). Therefore, thesetrends suggest that the geometry of undercoordinated sites on Cusurfaces can be engineered to guide selectivity toward greateroxygenate production.To hypothesize why undercoordinated sites are more selective

for oxygenates, it is useful to examine how the geometry and co-ordination of atoms will impact the surface coverage of interme-diates during CO2R. Online electrochemical mass spectrometryexperiments demonstrate a strong correlation between the applied

potentials for the hydrogen evolution reaction (HER) and re-duction of possible oxygenated CO2R intermediates on Cu elec-trodes (27). Recent theoretical studies indicate that C–C coupledproduct selectivity from Cu is strongly correlated to the coverageof CO* and H* simultaneously on the surface (28, 29). Thesereports suggest that the dominant mechanisms for C–C couplingand hydrogenation at lower overpotentials are chemical (hydridetransfer) rather than electrochemical (proton-coupled electrontransfer). Therefore, at lower overpotentials the selectivity foroxygenates or hydrocarbons could be determined by the ability ofthe surface to accommodate simultaneous coverages of CO* andH*. By examining structure models and atomically resolved in situECSTM images (Figs. 3 and 4), we can postulate how CO* andH* coverages depend on the geometry and coordination of atomson Cu(111), (100), and (751). Whereas the Cu(111) and (100)surfaces have a high density of sites with eight- and ninefold co-ordination, respectively, the Cu(751) surface has a low density ofsites with six-, seven-, and eightfold coordination (Fig. 3). A closerlook at the structure models shows that Cu(111), (100), and (751)each have six, four, and two nearest surface neighbors, re-spectively. It was formerly noted that the in situ surface structureof Cu(751) is better described as Cu(S)–[n(110) × (100)], where nvaries from 2 to 7, and that the surface heterogeneity changes thedistribution of six-, seven-, and eightfold coordination sites. Thisheterogeneity does not affect the trend in nearest surface neigh-bors, because the main structural motif of (110) terraces, whereeach Cu atom will have two nearest surface neighbors, is presentin both the composite Cu(S)–[n(110) × (100)] (Fig. 4C) and theideal Cu(751) (Fig. 4D) surface structures. Since it is statisticallyless likely for a CO* dimer to be adjacent to H* atoms on asurface with fewer neighbors, Cu(751) could have greater oxy-genate selectivity because it is more difficult to hydrogenate C–Ccoupled CO2R intermediates on its surface. This trend in surfaceneighbors is corroborated by comparing the trend in oxygenate/hydrocarbon ratios at −0.89 V vs. RHE, where Cu(751) > Cu(100) > Cu(111). At more negative potentials, all surfaces havesimilar oxygenate/hydrocarbon ratios, suggesting that CO* and H*coverages may be less important at high overpotentials. Theaforementioned theory study predicts that at more negative po-tentials the barrier height for chemical hydrogenation increases,whereas the barrier height for electrochemical hydrogenation de-creases (28). Therefore, these similarities in selectivity at higher

Fig. 5. Current efficiencies for >2e− C1, C2, and C3 products as a function ofpotential for Cu(111), (751), and (100).

Fig. 6. Oxygenate/hydrocarbon ratios for >2e− reduction products as afunction of potential for Cu(111), (751), and (100).

5922 | www.pnas.org/cgi/doi/10.1073/pnas.1618935114 Hahn et al.

Dow

nloa

ded

by g

uest

on

Mar

ch 1

4, 2

020

Page 6: Engineering Cu surfaces for the electrocatalytic ... · between >2e− oxygenates and hydrocarbons. Results and Discussion Out-of-Plane Thin-Film Texture. Thin films of Cu were synthesized

overpotentials could be due to a shift in the hydrogenationmechanism from chemical to electrochemical. If the predominantmechanism for hydrogenation at these overpotentials is elec-trochemical, it is expected that the oxygenate/hydrocarbonratio would be less sensitive to geometry and coverage on thesurface because the proton and electron transfers are con-certed. For polycrystalline Cu electrodes, the selective pro-duction of ethanol from the electrochemical reduction of COin alkaline solution was achieved by atomic-level surfacemodification via mild oxidation–reduction cycles (30). Resultsfrom the present work provide impetus for the exploration ofadvanced engineering strategies to design electrocatalyticsurfaces with undercoordinated sites that can steer selectivitytoward oxygenate formation.

ConclusionsWe have investigated PVD as a growth method to surface-structure-engineer large-format Cu thin films for electrochemicalCO2 reduction. X-ray pole figures indicate that the Cu thin filmsgrow epitaxially with <111>, <100>, and <751> out-of-plane ori-entations on Al2O3(0001), Si(100), and Si(111), respectively. TheXRD results show that Cu can be epitaxially grown in both low andhigh Miller index directions using different single-crystal substrateorientations. Analysis of the structure in situ using ECSTM con-firmed the existence of three unique Cu surface structures fromepitaxial growth. Electrochemical testing of these samples for CO2reduction led to a number of important observations. Thin-filmorientations with more undercoordinated sites are more activeand selective for C–C coupling, which is consistent with previousstudies on small-format single crystals. This demonstrates that

PVD can be used to grow large-format electrodes that haveanalogous electrocatalytic properties to single crystals. Further-more, analysis of oxygenate vs. hydrocarbon selectivity reveals thatat −0.89 V vs. RHE, Cu(751) has the highest oxygenate/hydro-carbon ratio of the three Cu orientations. We suggest that thisimprovement in oxygenate selectivity is related to the fewernumber of nearest neighbors on the Cu(S)–[n(110) × (100)] sur-face, or the topmost layer of the Cu(751) film, because barriers forhydride transfer are predicted to be lower than those for proton-coupled electron transfer at lower overpotentials. Additionalmechanistic details can be unveiled from future investigations thatexamine fine surface structural nuances under operando condi-tions, akin to the protocols that led to the discovery of Cu(511) asa selective ethanol-generating surface formed from polycrystallineCu (30). Our results demonstrate that epitaxy can aid in the dis-covery of structure–activity relationships for CO2R, providing in-sights into designing more active and selective electrocatalysts.

ACKNOWLEDGMENTS. We thank Dr. Jakob Kibsgaard and Dr. Karen Chanfor their assistance in constructing the Cu surface structure models. Addi-tional thanks go to the Stanford NMR Facility. Part of this work was per-formed at the Stanford Nano Shared Facilities (SNSF) and the StanfordNanofabrication Facility (SNF), supported by the National Science Founda-tion under Award ECCS-1542152. This material is based upon work per-formed by the Joint Center for Artificial Photosynthesis, a Department ofEnergy (DOE) Innovation Hub, as follows: the development of electrochemicaltesting of Cu thin films was supported through the Office of Science of theUS DOE under Award DE-SC0004993; the development of epitaxial growthwas supported by the Global Climate Energy Project at Stanford University;the procurement of the physical vapor deposition chamber was supportedby the DOE, Laboratory Directed Research and Development funding un-der Award DE-AC02-76SF00515.

1. Kortlever R, Shen J, Schouten KJP, Calle-Vallejo F, Koper MTM (2015) Catalysts andreaction pathways for the electrochemical reduction of carbon dioxide. J Phys ChemLett 6:4073–4082.

2. Hori Y (2016) CO2 reduction using electrochemical approach. Solar to ChemicalEnergy Conversion: Theory and Application, eds Sugiyama M, Fujii K, Nakamura S(Springer International, Cham, Switzerland), pp 191–211.

3. Kuhl KP, et al. (2014) Electrocatalytic conversion of carbon dioxide to methane andmethanol on transition metal surfaces. J Am Chem Soc 136:14107–14113.

4. Manthiram K, Beberwyck BJ, Alivisatos AP (2014) Enhanced electrochemical metha-nation of carbon dioxide with a dispersible nanoscale copper catalyst. J Am Chem Soc136:13319–13325.

5. Roberts FS, Kuhl KP, Nilsson A (2015) High selectivity for ethylene from carbon dioxidereduction over copper nanocube electrocatalysts. Angew Chem Int Ed Engl 54:5179–5182.

6. Ren D, et al. (2015) Selective electrochemical reduction of carbon dioxide to ethyleneand ethanol on copper(I) oxide catalysts. ACS Catal 5:2814–2821.

7. Kuhl KP, Cave ER, Abram DN, Jaramillo TF (2012) New insights into the electro-chemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ Sci5:7050–7059.

8. Hori Y (2008) Electrochemical CO2 reduction on metal electrodes. Modern Aspects ofElectrochemistry, No. 42, eds Vayenas CG, White RE, Gamboa-Aldeco ME (Springer,New York), pp 89–189.

9. Schouten KJ, Qin Z, Pérez Gallent E, Koper MT (2012) Two pathways for the formationof ethylene in CO reduction on single-crystal copper electrodes. J Am Chem Soc 134:9864–9867.

10. Montoya JH, Shi C, Chan K, Nørskov JK (2015) Theoretical insights into a CO di-merization mechanism in CO2 electroreduction. J Phys Chem Lett 6:2032–2037.

11. Li H, Li Y, Koper MTM, Calle-Vallejo F (2014) Bond-making and breaking betweencarbon, nitrogen, and oxygen in electrocatalysis. J Am Chem Soc 136:15694–15701.

12. Dehm G, Scheu C, Rühle M, Raj R (1998) Growth and structure of internalCu/Al2O3 and Cu/Ti/Al2O3 interfaces. Acta Mater 46:759–772.

13. Demczyk BG, Naik R, Auner G, Kota C, Rao U (1994) Growth of Cu films on hydrogenterminated Si(100) and Si(111) surfaces. J Appl Phys 75:1956–1961.

14. Reckinger N, et al. (2014) Anomalous moiré pattern of graphene investigated byscanning tunneling microscopy: Evidence of graphene growth on oxidized Cu(111).Nano Res 7:154–162.

15. Knorr DB, Lu T-M (1991) Effects of deposition conditions on texture in copper thinfilms on Si (111). Textures Microstruct 13:155–164.

16. Jiang H, Klemmer TJ, Barnard JA, Doyle WD, Payzant EA (1998) Epitaxial growth ofCu(111) films on Si(110) by magnetron sputtering: Orientation and twin growth. ThinSolid Films 315:13–16.

17. Knoll E, Bialas H (1994) Growth modes of epitaxial copper films on c-sapphire. ThinSolid Films 250:42–46.

18. Kumar B, et al. (2012) Photochemical and photoelectrochemical reduction of CO2.Annu Rev Phys Chem 63:541–569.

19. Walter MG, et al. (2010) Solar water splitting cells. Chem Rev 110:6446–6473.20. Hashim I, Park B, Atwater HA (1993) Epitaxial growth of Cu (001) on Si (001):

Mechanisms of orientation development and defect morphology. Appl Phys Lett 63:2833–2835.

21. Chang CA (1990) Formation of copper silicides from Cu(100)/Si(100) and Cu(111)/Si(111) structures. J Appl Phys 67:566–569.

22. Van Hove MA, Somorjai GA (1980) A new microfacet notation for high-Miller-indexsurfaces of cubic materials with terrace, step and kink structures. Surf Sci 92:489–518.

23. Kim Y-G, et al. (2016) Surface reconstruction of pure-Cu single-crystal electrodesunder CO-reduction potentials in alkaline solutions: A study by seriatim ECSTM-DEMS.J Electroanal Chem 780:290–295.

24. Kim Y-G, Baricuatro JH, Javier A, Gregoire JM, Soriaga MP (2014) The evolution of thepolycrystalline copper surface, first to Cu(111) and then to Cu(100), at a fixed CO2RRpotential: A study by operando EC-STM. Langmuir 30:15053–15056.

25. Peterson AA, Nørskov JK (2012) Activity descriptors for CO2 electroreduction tomethane on transition-metal catalysts. J Phys Chem Lett 3:251–258.

26. Shaner MR, Atwater HA, Lewis NS, McFarland EW (2016) A comparative tech-noeconomic analysis of renewable hydrogen production using solar energy. EnergyEnviron Sci 9:2354–2371.

27. Schouten KJP, Kwon Y, van der Ham CJM, Qin Z, Koper MTM (2011) A new mecha-nism for the selectivity to C1 and C2 species in the electrochemical reduction ofcarbon dioxide on copper electrodes. Chem Sci 2:1902–1909.

28. Goodpaster JD, Bell AT, Head-Gordon M (2016) Identification of possible pathwaysfor C-C bond formation during electrochemical reduction of CO2: New theoreticalinsights from an improved electrochemical model. J Phys Chem Lett 7:1471–1477.

29. Sandberg RB, Montoya JH, Chan K, Nørskov JK (2016) CO-CO coupling on Cu facets:Coverage, strain and field effects. Surf Sci 654:56–62.

30. Kim Y-G, Javier A, Baricuatro JH, Soriaga MP (2016) Regulating the product distri-bution of CO reduction by the atomic-level structural modification of the Cuelectrode surface. Electrocatalysis 7:391–399.

Hahn et al. PNAS | June 6, 2017 | vol. 114 | no. 23 | 5923

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Mar

ch 1

4, 2

020


Recommended