The growing interest in the growth and study of thin films of low-dimensional metallic delafossites, with the general formula ABO2, is driven by their potential to exhibit electronic and magnetic characteristics that are not accessible in bulk systems. The layered structure of these compounds introduces unique surface states as well as electronic and structural reconstructions, making the investigation of their surface behavior pivotal to understanding their intrinsic electronic structure. In this work, we study the surface phenomena of epitaxially grown PtCoO2, PdCoO2, and PdCrO2 films, utilizing a combination of molecular-beam epitaxy and angle-resolved photoemission spectroscopy. Through precise control of surface termination and treatment, we discover a pronounced 3×3 surface reconstruction in PtCoO2 films and PdCoO2 films, alongside a 2 × 2 surface reconstruction observed in PdCrO2 films. These reconstructions have not been reported in prior studies of delafossites. Furthermore, our computational investigations demonstrate the BO2 surface’s relative stability compared to the A-terminated surface and the significant reduction in surface energy facilitated by the reconstruction of the A-terminated surface. These experimental and theoretical insights illuminate the complex surface dynamics in metallic delafossites, paving the way for future explorations of their distinctive properties in low-dimensional studies.

In recent years, metallic oxides with a delafossite structure have captured significant attention due to their extraordinary structural and electronic characteristics. PtCoO2 single crystals are distinguished by their unparalleled high conductivity per carrier. Similarly, PdCoO2 single crystals possess the longest mean free path of up to 20 μm at 4 K, surpassing all other known oxide materials.1–3 The structure of these delafossites consists of layers of platinum (or palladium) that create a single, closed electron pocket at the center of the Brillouin zone at the Fermi level (EF), which is pivotal for the high conductivity in PtCoO2 (PdCoO2) single crystals.4–6 The CoO2 layers adjacent to these conductive platinum (or palladium) layers serve as insulating spacers, leading to the out-of-plane resistivity being more than 1000 times higher than the in-plane resistivity in PdCoO2 single crystals at low temperature.1,7 Notably, PtCoO2 films maintain relatively high conductivity even with reduced dimensionality, exhibiting diminished sensitivity to film thickness in comparison to copper.8 Among delafossite compounds, PdCrO2 is particularly remarkable for its unique combination of antiferromagnetic order (AFM) at around 37 K while maintaining its metallic conductivity.9,10 It exhibits a finite interlayer coupling between palladium layers and adjacent insulating CrO2 layers, making this an interesting system to study Kondo-like couplings in a hidden Mott material.9–13 

With their unique layered structures, metallic delafossites offer opportunities to create novel low-dimensional materials with distinctive properties and functionalities. The polar layers in delafossites, denoted as A1+ (A = Pt, Pd) and BO21 (B = Co, Cr) in the bulk, exhibit distinct surface states when cleaved in vacuum, effectively addressing the challenges posed by polar surfaces.14 The A polar surfaces facilitate pronounced electron doping and stabilize the surface states of platinum and palladium layers, whereas hole-doped surface states emerge at the BO2 terminations. The CoO2-terminated surface displays a substantial spin splitting of the surface states, arising due to a strong breaking of the inversion-symmetry at the surface.15 The ferromagnetism observed in the palladium surface state opens up possibilities for creating two-dimensional ferromagnets.14,16 The superlattice structures of PdCoO2 and PdCrO2, characterized by structural symmetry breaking, provide a potential avenue for studying interlayer electron interactions. In addition, the Fermi surface of PdCrO2, which has predominantly palladium character, undergoes a magnetic reconstruction into a 3×3 pattern driven by AFM order at low temperatures.9–13 Despite the significant promise held by low-dimensional delafossites, interpretation of surface reconstruction in thin films is complicated by the simultaneous presence of surface states and electronic reconstructions driven by AFM order. This complexity underscores the need for further detailed investigations to decipher the intricate phenomena underlying these observations.

Through the powerful combination of molecular-beam epitaxy (MBE) and angle-resolved photoemission spectroscopy (ARPES), we have grown films of PtCoO2, PdCoO2, and PdCrO2 with controlled terminations, enabling a detailed investigation into their electronic structures. These delafossites exhibit distinct surface reconstructions, including a new 2 × 2 surface reconstruction in PdCrO2 that has never been previously reported for any delafossite material. Utilizing first-principles density functional theory (DFT) calculations, we demonstrate that surface reconstructions induced by excess oxygen significantly reduce the surface energy for A terminations, while BO2 terminations exhibit relative stability compared to A terminations. In order to show a comprehensive comparison of the metallic delafossites – PtCoO2, PdCoO2, and PdCrO2—We present side-by-side ARPES, low-energy electron diffraction (LEED), reflection high-energy electron diffraction (RHEED), and first-principles calculations of these materials in the subsequent figures.

Metallic delafossite thin films of PtCoO2, PdCoO2, and PdCrO2 are synthesized in a Veeco Gen10 MBE system on (001) sapphire substrates. The atomic structure of delafossites, exemplified by PtCoO2, is illustrated in Fig. 1(a). The growth process is described in detail in the supplementary material. Figure 1(b) presents the results of PtCoO2 films’ growth in an adsorption-controlled regime, highlighting the narrow growth window for achieving single-phase PtCoO2 films. The atomic force microscopy image in Fig. 1(c) reveals 180° in-plane rotation twins within these films, in agreement with prior reports of the epitaxial growth of PtCoO2, PdCoO2, and PdCrO2 films on (001) Al2O3 substrates.17–25 The pure-phase nature of the PtCoO2 films is confirmed by θ–2θ x-ray diffraction scans in Fig. 1(d). Resistivity vs temperature measurements of a 13.3-nm-thick PtCoO2 film, conducted using a Quantum Design Physical Property Measurement System (PPMS) employing a four-point van der Pauw geometry, are shown in Fig. 1(e). As the 4-nm-thick PdCoO2 buffer layer is quite flat as shown in the supplementary material (Fig. S1), its resistance has been subtracted out of the resistance measurement of the PtCoO2 film (assuming a simple parallel resistance model). The resistivity of the PtCoO2 film is subsequently calculated using the average thickness of the PtCoO2 film. The residual resistivity ratio (RRR = ρ300K/ρ4K) of this PtCoO2 film is 1.91 in its as-grown state (i.e., without any ex situ post anneal). For comparison, the RRR of a PtCoO2 single crystal is 52.5.3 The resistivity of our PtCoO2 film shows a similar linear temperature dependence behavior at high temperatures to that observed in the single crystals. The resistivity comparison between the film and the single crystal is shown in the supplementary material (Fig. S1). The microstructure of the PtCoO2 film is investigated by high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images in Fig. 1(f) and electron energy-loss spectroscopy (EELS) maps found in the supplementary material (Fig. S4). Details on the growth method, as well as the structural and electrical characterization of PdCrO2 films, can be found in the supplementary material (Figs. S2 and S3). All PtCoO2 films investigated in this study exhibit a mixed termination of platinum and CoO2, a result of the adsorption-controlled method utilized to produce them.24 The palladium termination of the PdCoO2 film and the CrO2 termination of the PdCrO2 films were achieved via the shutter-controlled growth method employed for their growths, as described in prior work.22,24 Note that not all delafossite films with the described terminations exhibit surface reconstructions. For example, no reconstruction features are present in the Pd- and CoO2- terminated PdCoO2 films described in Ref. 24. This points to a multitude of routes to compensate for the polar surface charge in these materials and a delicate balance of energetics. The precise conditions for stabilizing the different reconstructions observed in these delafossite films, thus, require further study.

FIG. 1.

Structural and electrical characterizations of PtCoO2 thin films grown on (001) Al2O3 substrates by MBE. (a) Crystal structure of PtCoO2, with (100) and (110) planes illustrated on the atomic (001) plane by the black lines. (b) Diagram showing the phases obtained as a function of substrate temperature and Pt:Co ratios during film growth by co-deposition on (001) Al2O3 substrates. Red circles indicate phase-pure PtCoO2 films, blue squares indicate PtCoO2 films in which Co3O4 exists as a second phase, yellow triangles indicate conditions under which platinum exists as a second phase, and purple diamonds indicate where Co3O4 and platinum impurity phases also exist in the PtCoO2 films. All PtCoO2 films described here are deposited onto 4-nm-thick PdCoO2 films. (c) Atomic force microscopy image of a 13.3-nm-thick PtCoO2 film deposited on top of a 4.0-nm-thick PdCoO2 buffer layer. (d) X-ray diffraction of the same PtCoO2 thin film characterized in (c). * denotes the 006 peak of the (001) Al2O3 substrate. (e) Temperature-dependent resistivity measurements of the same PtCoO2 film characterized in (c). (f) STEM-HAADF image of a PtCoO2 film grown on a 4-nm-thick buffer layer of PdCoO2 grown on a (001) Al2O3 substrate, viewed along the [201] zone axis of PtCoO2. The film shows high crystalline quality with abrupt interfaces. (g) A higher magnification view of the region outlined by the orange box in (f). The inset of (g) displays an overlay with the structural model where platinum gives the strongest contrast, followed by cobalt due to its lower atomic number, whereas oxygen is invisible in the STEM-HAADF image.

FIG. 1.

Structural and electrical characterizations of PtCoO2 thin films grown on (001) Al2O3 substrates by MBE. (a) Crystal structure of PtCoO2, with (100) and (110) planes illustrated on the atomic (001) plane by the black lines. (b) Diagram showing the phases obtained as a function of substrate temperature and Pt:Co ratios during film growth by co-deposition on (001) Al2O3 substrates. Red circles indicate phase-pure PtCoO2 films, blue squares indicate PtCoO2 films in which Co3O4 exists as a second phase, yellow triangles indicate conditions under which platinum exists as a second phase, and purple diamonds indicate where Co3O4 and platinum impurity phases also exist in the PtCoO2 films. All PtCoO2 films described here are deposited onto 4-nm-thick PdCoO2 films. (c) Atomic force microscopy image of a 13.3-nm-thick PtCoO2 film deposited on top of a 4.0-nm-thick PdCoO2 buffer layer. (d) X-ray diffraction of the same PtCoO2 thin film characterized in (c). * denotes the 006 peak of the (001) Al2O3 substrate. (e) Temperature-dependent resistivity measurements of the same PtCoO2 film characterized in (c). (f) STEM-HAADF image of a PtCoO2 film grown on a 4-nm-thick buffer layer of PdCoO2 grown on a (001) Al2O3 substrate, viewed along the [201] zone axis of PtCoO2. The film shows high crystalline quality with abrupt interfaces. (g) A higher magnification view of the region outlined by the orange box in (f). The inset of (g) displays an overlay with the structural model where platinum gives the strongest contrast, followed by cobalt due to its lower atomic number, whereas oxygen is invisible in the STEM-HAADF image.

Close modal

ARPES measurements were employed to investigate the bulk states and surface reconstructions of metallic delafossite films. ARPES measurements were conducted using lab-based ARPES systems, which consist of a Scienta Omicron VUV 5000 emitting He-I photons at 21.2 eV and He-II photons at 40.8 eV, and a Fermion Instruments BL1200s discharge lamp for neon-I photons at 16.85 eV.26 The detection of emitted electrons was performed using a VG Scienta R4000 electron analyzer. Prior to ARPES measurements, all films were exposed to air and then underwent re-annealing in ozone at a background partial pressure of 5 × 10−6 Torr at 480 °C for 5 min.

In Figs. 2(a), 2(c), and 2(e), we compare the Fermi surfaces of PtCoO2, PdCoO2, and PdCrO2 films, respectively. Specifically, Fig. 2(a) highlights a hexagonal pocket, denoted as the α band. This intense band is centered at the Γ point and represents the platinum-driven band of PtCoO2, consistent with observation from previous studies on PtCoO2 single crystals.2 Interestingly, in addition to the bulk α band, additional band features labeled as α′ bands are observed both at EF and 150 meV below it, as shown in Figs. 2(a) and 2(b). These α′ bands occupy a similar momentum position to those of the reconstructed bands in PdCrO2 at low temperatures.11,13 While the known additional bands in the PdCrO2 case arise due to band folding at the magnetic Brillouin zone boundary, we will show below that these new states in PtCoO2 films reflect a new surface reconstruction. This observation is corroborated by the observation of the same α′ bands identified in another PtCoO2 film using 16.85 eV photons, whereas these bands were absent in the ARPES data obtained with 40.8 eV photon energy, as shown in Figs. S6 and S7 in the supplementary material. The absence might be related to the photon energy dependence of the relevant photoemission cross section.

FIG. 2.

Photoemission intensity maps, collected using 21.2 eV photons, at EF for (a) the 17.3-nm-thick PtCoO2 film (including 4-nm-thick PdCoO2 buffer layer), (c) a 18.3-nm-thick PdCoO2 film, and (e) a 14.1-nm-thick PdCrO2 film. (b), (d), and (f) represent the same maps as (a), (c), and (e), respectively, but are taken at 150 meV below EF. The data of the PtCoO2 and PdCoO2 films were obtained at a temperature of 6 K, while the PdCrO2 film data were taken at 50 K, i.e., above the AFM transition temperature (37 K) observed in PdCrO2 single crystals. Data of the PdCrO2 film collected at 6 K are shown in Fig. S5 in the supplementary material.

FIG. 2.

Photoemission intensity maps, collected using 21.2 eV photons, at EF for (a) the 17.3-nm-thick PtCoO2 film (including 4-nm-thick PdCoO2 buffer layer), (c) a 18.3-nm-thick PdCoO2 film, and (e) a 14.1-nm-thick PdCrO2 film. (b), (d), and (f) represent the same maps as (a), (c), and (e), respectively, but are taken at 150 meV below EF. The data of the PtCoO2 and PdCoO2 films were obtained at a temperature of 6 K, while the PdCrO2 film data were taken at 50 K, i.e., above the AFM transition temperature (37 K) observed in PdCrO2 single crystals. Data of the PdCrO2 film collected at 6 K are shown in Fig. S5 in the supplementary material.

Close modal

In a palladium-terminated PdCoO2 film, analogous new band features referred to as β′ bands were found as shown in Figs. 2(c) and 2(d). The PdCoO2 bulk state, denoted as the β band, primarily exhibits palladium character according to prior work on PdCoO2 single crystals.14,27,28 It is important to note that the β′ bands in the PdCoO2 film differ from the symmetry of surface states attributed to palladium and CoO2 layers, as discussed in previous work.14,24,29 Thus, the presence of the α′ bands in PtCoO2 films and the β′ bands in PdCoO2 films is new territory compared to prior work.

The band structure of the PdCrO2 film is shown in Figs. 2(e) and 2(f). Although the band structures of PdCoO2 and PtCoO2 are similar, that of PdCrO2 exhibits distinctive behavior. This band structure exhibits a hexagonal pocket β band centered at Γ point, representing the primary palladium character of the bulk state, in line with previous reports on PdCrO2 single crystals.11,13 The bulk bands observed in PtCoO2, PdCoO2, and PdCrO2 films are consistent with our independent DFT calculations (see Fig. S11 in the supplementary material).28 Nonetheless, the PdCrO2 film also presents additional band features, denoted as β″ bands, which have not been reported in any prior delafossite materials. Note that the data for the β″ band in our PdCrO2 film were obtained above the AFM transition temperature, suggesting that this band’s presence is not associated with the AFM order in PdCrO2. Nevertheless, the β″ bands are also present in the ARPES result of this PdCrO2 film taken at 6 K, as shown in the supplementary material (Fig. S5). Moreover, the β″ band exhibits distinct dispersion compared to the reconstructed bands driven by AFM order in PdCrO2 single crystals. In addition, the β″ bands are also present in a thinner PdCrO2 film (4 nm thickness) as shown in Fig. S8 in the supplementary material. Nonetheless, not all the PdCrO2 films with CrO2 termination exhibit the β″ bands. Figure S9 in the supplementary material shows the dispersions at the M point of the PtCoO2 and PdCrO2 films.

We further compare the dispersion of these additional bands in the metallic delafossite films along the Γ-K direction, as illustrated in Fig. 3. Figures 3(a), 3(d), and 3(g) display the momentum dispersion curves at EF of the PtCoO2, PdCoO2, and PdCrO2 films, respectively. The high-intensity α band in Fig. 3(a) and the highly dispersive β bands observed in Figs. 3(d) and 3(g) correspond to the platinum and palladium bulk state in PtCoO2, PdCoO2, and PdCrO2 films, as previously shown in Figs. 2(a), 2(c), and 2(e). The α′ band in the PtCoO2 film and the β′ band in the PdCoO2 film exhibit similar dispersion, as illustrated in Figs. 3(c) and 3(f). Conversely, the β″ band in the PdCrO2 film displays a distinct kF position compared to the α′ band in the PtCoO2 film and the β′ band in the PdCrO2 film, as shown in Fig. 3(i). In addition, Fig. 3(a) reveals the subtle presence of an additional γ band in the PtCoO2 film, a feature also shown in Fig. 2(a). This γ band could potentially be associated with a platinum surface state, similar to the palladium surface state observed in PdCoO2 single crystals.14 Given the analogous behavior of the α′ band in the PtCoO2 film and the β′ band in the PdCoO2 film, our analysis now focuses on the comparison between the PtCoO2 and PdCrO2 films.

FIG. 3.

Photoemission intensity distributions of the PtCoO2, PdCoO2, and PdCrO2 films measured at a photon energy of 21.2 eV. (a) Momentum distribution curve across Γ-K of the PtCoO2 film at EF. The platinum bulk state α band, additional α′ band, and surface-state-like γ band are indicated. (b) The photoemission intensity distribution of the PtCoO2 film between EF and 500 meV below EF. (c) Momentum distribution curves between EF and 250 meV below EF across Γ-K of the PtCoO2 film. The yellow triangles denote the additional α′ bands. (d), (e), and (f) The same as (a), (b), and (c), but for the PdCoO2 film. The palladium bulk band in the PdCoO2 film is indicated as the β band. An additional β′ band is also present and is shown by red triangles in (f). It is located at the same momentum position as the α′ band in the PtCoO2 film. (g), (h), and (i) Corresponding measurements on the PdCrO2 film, with the β band representing the palladium bulk band, and the additional β″ band indicated by red diamonds in (i), which differs from the α′ band and β′ band observed in the PtCoO2 and PdCoO2 film shown in (c) and (f), as indicated by the light yellow and red triangles in (i). The PtCoO2 and PdCoO2 film data were acquired at a temperature of 6 K, while the PdCrO2 film data were obtained at 50 K.

FIG. 3.

Photoemission intensity distributions of the PtCoO2, PdCoO2, and PdCrO2 films measured at a photon energy of 21.2 eV. (a) Momentum distribution curve across Γ-K of the PtCoO2 film at EF. The platinum bulk state α band, additional α′ band, and surface-state-like γ band are indicated. (b) The photoemission intensity distribution of the PtCoO2 film between EF and 500 meV below EF. (c) Momentum distribution curves between EF and 250 meV below EF across Γ-K of the PtCoO2 film. The yellow triangles denote the additional α′ bands. (d), (e), and (f) The same as (a), (b), and (c), but for the PdCoO2 film. The palladium bulk band in the PdCoO2 film is indicated as the β band. An additional β′ band is also present and is shown by red triangles in (f). It is located at the same momentum position as the α′ band in the PtCoO2 film. (g), (h), and (i) Corresponding measurements on the PdCrO2 film, with the β band representing the palladium bulk band, and the additional β″ band indicated by red diamonds in (i), which differs from the α′ band and β′ band observed in the PtCoO2 and PdCoO2 film shown in (c) and (f), as indicated by the light yellow and red triangles in (i). The PtCoO2 and PdCoO2 film data were acquired at a temperature of 6 K, while the PdCrO2 film data were obtained at 50 K.

Close modal

To investigate the surface reconstructions of the PtCoO2 film and the PdCrO2 films, we analyzed their LEED images displayed in Figs. 4(a) and 4(c). In the composite LEED image of the PtCoO2 film [Fig. 4(a)], the reciprocal lattice of PtCoO2 is marked by red solid diamonds, as shown by the red solid hexagonal zone in the center. A distinct 3×3 reconstruction with a 30° rotation is indicated by the green diamonds along with a less pronounced 4 × 4 reconstruction signified by the yellow hexagons. The 3×3 reconstruction with a 30° rotation is simplistically denoted as 3×3. The green zones (solid lines) and yellow zones (dashed lines) denote the pronounced 3×3 and the subtle 4 × 4 reconstructed orientations, respectively. Importantly, these reconstruction patterns observed in the PtCoO2 LEED image correspond precisely with those observed in the RHEED image along the [100] direction in reciprocal space, as evidenced by comparing the white dashed line in Fig. 4(a) with the peak positions in the RHEED image [Fig. 4(b)]. Similarly, the LEED image for the PdCrO2 film [Fig. 4(c)] displays the reciprocal lattice of PdCrO2 denoted by red solid diamonds, with a strong 2 × 2 reconstruction denoted by blue hexagons and a weak 3×3 reconstruction represented by green diamonds. The blue zones (solid lines) and green zones (dashed lines) represent the strong 2 × 2 and the weak 3×3 reconstructed orientations, respectively. Once again, the reconstructions identified in the LEED images are seen to be consistent with the peaks observed in the RHEED image along the [100] direction, as indicated in Figs. 4(c) and 4(d), ensuring a consistent interpretation of the surface reconstructions across different imaging techniques.

FIG. 4.

Surface reconstructions observed via LEED, RHEED, and ARPES. (a) LEED image of the PtCoO2 film, where red solid diamonds correspond to the 1 × 1 reciprocal lattice. The green diamonds highlight the 3×3 reconstruction, as shown in the hexagonal green zone. The yellow hexagons indicate the observed 4 × 4 reconstruction in LEED, signified by the zone covered by dashed yellow lines. The black arrow corresponds to the [100] electron-beam direction of the RHEED image shown in (b). The white dashed line denotes the direction along which the RHEED image is captured. (b) RHEED pattern of the PtCoO2 in the (100)* reciprocal lattice plane, corresponding to the white dashed line in (a). The dashed line indicates where the intensity data are extracted. The symbols on the dashed line match those on the white dashed line in (a), precisely aligning with the peaks observed in the RHEED image. (c) and (d) The same information as (a) and (b), respectively, but for the PdCrO2 film. In (c), the red solid diamonds correspond to the 1 × 1 reciprocal lattice of the PdCrO2 film, while the blue hexagons illustrate the 2 × 2 reconstruction shown in the blue hexagonal zone. A weak 3×3 reconstruction is indicated by the green diamonds in the dashed hexagonal green zones. (e) The Fermi surface of the PtCoO2 film, where the central hexagonal pocket made up of green dots is extracted from where the bulk state α band crosses EF. The surrounding hexagonal pockets are multiplied by the 3×3 folding. The red arrows indicate one of the nesting directions. (f) The same as (e), but for the PdCrO2 film with the 2 × 2 reconstruction.

FIG. 4.

Surface reconstructions observed via LEED, RHEED, and ARPES. (a) LEED image of the PtCoO2 film, where red solid diamonds correspond to the 1 × 1 reciprocal lattice. The green diamonds highlight the 3×3 reconstruction, as shown in the hexagonal green zone. The yellow hexagons indicate the observed 4 × 4 reconstruction in LEED, signified by the zone covered by dashed yellow lines. The black arrow corresponds to the [100] electron-beam direction of the RHEED image shown in (b). The white dashed line denotes the direction along which the RHEED image is captured. (b) RHEED pattern of the PtCoO2 in the (100)* reciprocal lattice plane, corresponding to the white dashed line in (a). The dashed line indicates where the intensity data are extracted. The symbols on the dashed line match those on the white dashed line in (a), precisely aligning with the peaks observed in the RHEED image. (c) and (d) The same information as (a) and (b), respectively, but for the PdCrO2 film. In (c), the red solid diamonds correspond to the 1 × 1 reciprocal lattice of the PdCrO2 film, while the blue hexagons illustrate the 2 × 2 reconstruction shown in the blue hexagonal zone. A weak 3×3 reconstruction is indicated by the green diamonds in the dashed hexagonal green zones. (e) The Fermi surface of the PtCoO2 film, where the central hexagonal pocket made up of green dots is extracted from where the bulk state α band crosses EF. The surrounding hexagonal pockets are multiplied by the 3×3 folding. The red arrows indicate one of the nesting directions. (f) The same as (e), but for the PdCrO2 film with the 2 × 2 reconstruction.

Close modal

A detailed analysis of the Fermi surface measured on the PtCoO2 film is performed by extracting the bulk α band at EF in Fig. 2(a) and folding it with the 3×3 reconstruction. The corresponding results are presented in Fig. 4(e), where we observed a striking match between the additional band features at EF (α′ band) and the folded band with the 3×3 reconstruction. Despite its presence in both the LEED and RHEED analyses of the PtCoO2 film, we did not observe any additional 4 × 4 reconstruction at EF. It is worth noting that the intensity of the 4 × 4 reconstruction is considerably weaker compared to the 3×3 reconstruction, which may account for its limited visibility. Similarly, we folded the bulk band of the PdCrO2 film at EF, which revealed an almost exact correspondence between the additional β′ band and the band folded with the 2 × 2 reconstruction, as presented in Fig. 4(f). Interestingly, the band structure of the PdCrO2 film at EF did not exhibit the 3×3 reconstruction suggested by the LEED result. This discrepancy could be attributed to the relatively weak 3×3 reconstruction compared to the 2 × 2 reconstruction, as evidenced in the LEED analysis of the PdCrO2 film presented in Fig. 4(f).

Our findings provide conclusive evidence for the primary 3×3 surface reconstruction in the PtCoO2 film, which is also observed in the PdCoO2 film. In addition, the main 2 × 2 reconstruction has been firmly identified in the PdCrO2 film. Evidence of these reconstructions is consistently detected by the ARPES, LEED, and RHEED measurements on our films. Nonetheless, such specific reconstructions have not been previously reported in delafossite materials, which are usually prepared by cleavage of single crystals, prompting further investigation into their origin. Figure 4 elucidates the consistency between the RHEED data and the reconstructions observed in LEED results. We further conduct a comparative analysis of the RHEED images of the PdCoO2 film [in Figs. 2(c) and 2(d)] at various stages of growth, enabling real-time monitoring during the growth process. These RHEED images, presented in Fig. S10 in the supplementary material, reveal the emergence of one-third-order peaks during the annealing of the as-grown PdCoO2 film in distilled ozone and persist throughout the cooling process. Remarkably, these one-third-order peaks in the PdCoO2 film mirror those seen in the PtCoO2 film [Fig. 4(b)], corresponding to the 3×3 reconstruction. This observation suggests a potential link between the 3×3 reconstruction and excess oxygen on the film surface, implying that excess oxygen on the surface may mitigate the polar surface charge by reducing the surface energy. Note that the PdCoO2 film possesses a full layer of palladium termination, introducing a positive charge. The additional oxygen helps to neutralize this polar surface charge, thus lowering the surface energy. Addressing the challenges of polar surfaces can be approached in various ways; for instance, when the PtCoO2 film is grown by co-deposition, the result is a mixed termination by platinum and CoO2. It is important to mention that all films are subjected to annealing in distilled ozone to provide fresh surfaces prior to the ARPES measurement; thus, there are multiple steps during growth and annealing where excess oxygen is supplied and could, thus, attach to the film surface should it be energetically favorable to do so.

Given the limitations of STEM to clearly resolve surface atoms, we next discuss the potential surface scenarios compatible with 3×3 and 2 × 2 reconstructions through DFT calculations. We have chosen to omit scenarios leading to the weak 4 × 4 reconstruction observed in PtCoO2 due to the complexity introduced by their large supercell sizes. For the 3×3 reconstruction, the scenarios considered include the presence of one oxygen vacancy (3×3 −1/3 oxygen), an excess of A-site atoms on a BO2-terminated surface (3×3 + n/3 A-site), and an excess of oxygen atoms on an A-terminated surface (3×3 + n/3 oxygen), where n equals 1 or 2. Similarly, for the 2 × 2 reconstruction, we explore the possibility of one oxygen vacancy (2 × 2 − 1/4 oxygen), an excess of A atoms on the BO2-terminated surface (2 × 2 + n/4 A-site), and an excess of oxygen atoms on an A-terminated surface (2 × 2 + n/4 oxygen), where n is 1, 2, or 3. In addition, we have calculated the pristine surfaces, denoted as BO2- and A-site termination in Fig. 5. For each scenario, we search for the lowest energy surface atomic configuration and then calculate its surface energy relative to the pristine BO2-terminated surface. The results are illustrated in Fig. 5, with computational details available in the supplementary material (Fig. S12).

FIG. 5.

Surface energy analysis via first-principles DFT calculations for PtCoO2, PdCoO2, and PdCrO2. The surface energies for different scenarios are compared against the BO2 terminated surface in panel 2, i.e., oxygen vacancy, A excess, and oxygen excess, which modify the pristine A or BO2 terminated surfaces and lead to 3×3 and 2 × 2 reconstruction. The vertical arrows point to the surface configurations with the lowest surface energy. “BO2-term” stands for films terminated by a CoO2 layer or CrO2 layer, and “A-term” represents films terminated by a palladium or platinum layer. We used non-spin-polarized calculations for PtCoO2 and PdCoO2, and spin-polarized calculations for PdCrO2, assuming a ferromagnetic configuration.

FIG. 5.

Surface energy analysis via first-principles DFT calculations for PtCoO2, PdCoO2, and PdCrO2. The surface energies for different scenarios are compared against the BO2 terminated surface in panel 2, i.e., oxygen vacancy, A excess, and oxygen excess, which modify the pristine A or BO2 terminated surfaces and lead to 3×3 and 2 × 2 reconstruction. The vertical arrows point to the surface configurations with the lowest surface energy. “BO2-term” stands for films terminated by a CoO2 layer or CrO2 layer, and “A-term” represents films terminated by a palladium or platinum layer. We used non-spin-polarized calculations for PtCoO2 and PdCoO2, and spin-polarized calculations for PdCrO2, assuming a ferromagnetic configuration.

Close modal

For all three compounds studied, the surface energies associated with creating an oxygen vacancy or adding excess A-site atom(s) on BO2-terminated surfaces are found to be positive, often exceeding 0.5 eV per unit cell area (panels 1 and 3 in Fig. 5). This indicates that both removing oxygen from and adding A atom(s) onto the BO2-terminated surface are energetically costly processes. Conversely, the introduction of excess oxygen atoms on A-terminated surfaces significantly lowers the surface energy, as shown in panel 5 in Fig. 5. In certain cases, this adjustment leads to negative surface energies when compared to the BO2-terminated surface, suggesting that adding oxygen to A-terminated surfaces is energetically favorable and could lead to more stable surface configurations.

The above results further support the hypothesis that the reconstructions observed in PtCoO2 (mixed termination) and PdCoO2 (palladium termination) films could be attributed to excess oxygen, which effectively reduces the surface energy. In our ARPES and LEED results, the 3×3 reconstruction is manifest on both PtCoO2 and PdCoO2 films. Nonetheless, our calculations indicate that scenarios leading to the 3×3 reconstruction do not consistently exhibit the lowest surface energy. This discrepancy might originate from the limitation of DFT in accurately determining surface energies or could be due to additional complexities present on the experimental film surfaces that are not fully accounted for in the DFT models. For instance, the mixed termination observed in PtCoO2 could alter the energy landscape in ways not captured by our DFT simulations. Therefore, while our DFT calculations align with our ARPES and LEED findings in suggesting that excess oxygen contributes to the observed reconstructions, they do not conclusively determine the size of the reconstructed supercell. In the case of PdCrO2 films grown with CrO2 terminations, scenarios involving excess oxygen are not applicable. This leaves the 2 × 2 reconstruction with 1/4 A-site addition as a scenario with relatively low surface energy, aligning with our experimental observations and presenting a possible explanation for the surface structure of PdCrO2 films.

In conclusion, we have successfully synthesized high-quality PtCoO2, PdCoO2, and PdCrO2 films by MBE, followed by a comprehensive investigation of their surface reconstructions using a combination of ARPES, LEED, and RHEED. Our investigations have unveiled the presence of surface reconstructions in these films, including a 3×3 reconstruction in PtCoO2 films with mixed platinum–CoO2 terminations and in palladium-terminated PdCoO2 film, as well as a distinctive 2 × 2 reconstruction is seen in CrO2-terminated PdCrO2 films. These findings, which are corroborated by ARPES, LEED, and RHEED analyses, highlight surface reconstructions in these films that have not been previously reported in delafossite single crystals. These new reconstructions are likely closely linked to the polar terminations of the as-grown thin films, where the additional atoms on pristine surfaces could potentially lower their surface energy. DFT calculations shed light on the relative instability of surfaces terminated with A (A = Pt, Pd) atoms compared to those terminated with BO2 (B = Co, Cr), as evidenced by their higher formation energies. Nonetheless, the presence of excess oxygen can significantly mitigate these energies, leading to reconstructed configurations. Our results show general consistency (with some discrepancies) between the calculated and experimentally observed sizes of the reconstructed supercells. By elucidating the distinct surface reconstructions from the surface states associated with different terminations in metallic delafossites, as well as from the electronic reconstructions driven by magnetic order, our work not only contributes to the understanding of surface phenomena in delafossite materials but also sets the stage for further exploration of highly two-dimensional studies in this intriguing class of materials.

See the supplementary material for a description of how the PtCoO2 and PdCrO2 films are grown and analyzed, additional characterization by TEM, RHEED, and ARPES, and details on the method of DFT calculations.

This paper was primarily supported by the U.S. Department of Energy, Office of Basic Sciences, Division of Materials Science and Engineering under Award No. DE-SC0002334. This research was funded in part by the Gordon and Betty Moore Foundation’s EPiQS Initiative (Grant Nos. GBMF3850 and GBMF9073 to Cornell University). This paper made use of the Cornell Center for Materials Research shared facilities, which are supported through the NSF Materials Research Science and Engineering Centers Program (Grant No. DMR-1719875). B.D.F., M.R.B., and B.P. acknowledged support from the National Science Foundation Platform for the Accelerated Realization, Analysis, and Discovery of Interface Materials (PARADIM) under Cooperative Agreement No. DMR-2039380. This paper also made use of the Cornell Energy Systems Institute Shared Facilities partly sponsored by the NSF (Grant No. MRI DMR-1338010) and the Kavli Institute at Cornell. Substrate preparation was performed, in part, at the Cornell NanoScale Facility, a member of the National Nanotechnology Coordinated Infrastructure, which is supported by the NSF (Grant No. NNCI-2025233). P.K. acknowledged support from the European Research Council (through the QUESTDO Project, 714193) and The Leverhulme Trust (Grant No. RPG-2023-256).

The authors have no conflicts to disclose.

Qi Song: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Writing – original draft (equal). Zhiren He: Formal analysis (equal); Methodology (equal); Writing – review & editing (supporting). Brendan D. Faeth: Data curation (equal); Formal analysis (equal); Writing – review & editing (equal). Christopher T. Parzyck: Data curation (equal); Formal analysis (supporting); Writing – review & editing (equal). Anna Scheid: Data curation (equal); Formal analysis (equal). Chad J. Mowers: Data curation (supporting); Writing – review & editing (supporting). Yufan Feng: Data curation (supporting). Qing Xu: Data curation (equal); Formal analysis (equal). Sonia Hasko: Data curation (supporting). Jisung Park: Resources (supporting). Matthew R. Barone: Resources (supporting). Y. Eren Suyolcu: Data curation (equal). Peter A. van Aken: Resources (supporting). Betül Pamuk: Formal analysis (supporting); Writing – review & editing (supporting). Craig J. Fennie: Resources (supporting). Phil D. C. King: Investigation (equal); Writing – review & editing (supporting). Kyle M. Shen: Resources (supporting). Darrell G. Schlom: Funding acquisition (equal); Investigation (equal); Resources (equal); Supervision (equal); Writing – review & editing (equal).

Datasets generated and analyzed during the course of this study are available at https://doi.org/10.34863/0ngf-h790.

1.
C. W.
Hicks
,
A. S.
Gibbs
,
A. P.
Mackenzie
,
H.
Takatsu
,
Y.
Maeno
, and
E. A.
Yelland
, “
Quantum oscillations and high carrier mobility in the delafossite PdCoO2
,”
Phys. Rev. Lett.
109
,
116401
(
2012
).
2.
P.
Kushwaha
,
V.
Sunko
,
P. J.
Moll
,
L.
Bawden
,
J. M.
Riley
,
N.
Nandi
,
H.
Rosner
,
M. P.
Schmidt
,
F.
Arnold
,
E.
Hassinger
,
T. K.
Kim
,
M.
Hoesch
,
A. P.
MacKenzie
, and
P. D. C.
King
, “
Nearly free electrons in a 5d delafossite oxide metal
,”
Sci. Adv.
1
,
e1500692
(
2015
).
3.
A. P.
Mackenzie
, “
The properties of ultrapure delafossite metals
,”
Rep. Prog. Phys.
80
,
032501
(
2017
).
4.
V.
Eyert
,
R.
Frésard
, and
A.
Maignan
, “
On the metallic conductivity of the delafossites PdCoO2 and PtCoO2
,”
Chem. Mater.
20
,
2370
2373
(
2008
).
5.
K.
Kim
,
H. C.
Choi
, and
B. I.
Min
, “
Fermi surface and surface electronic structure of delafossite PdCoO2
,”
Phys. Rev. B
80
,
035116
(
2009
).
6.
K. P.
Ong
,
D. J.
Singh
, and
P.
Wu
, “
Unusual transport and strongly anisotropic thermopower in PtCoO2 and PdCoO2
,”
Phys. Rev. Lett.
104
,
176601
(
2010
).
7.
Y.
Zhang
,
F.
Tutt
,
G. N.
Evans
,
P.
Sharma
,
G.
Haugstad
,
B.
Kaiser
,
J.
Ramberger
,
S.
Bayliff
,
Y.
Tao
,
M.
Manno
,
J.
Garcia-Barriocanal
,
V.
Chaturvedi
,
R. M.
Fernandes
,
T.
Birol
,
W. E.
Seyfried
, Jr.
, and
C.
Leighton
, “
Crystal-chemical origins of the ultrahigh conductivity of metallic delafossites
,”
Nat. Commun.
15
,
1399
(
2024
).
8.
S.
Kumar
,
C.
Multunas
,
B.
Defay
,
D.
Gall
, and
R.
Sundararaman
, “
Ultralow electron-surface scattering in nanoscale metals leveraging Fermi-surface anisotropy
,”
Phys. Rev. Mater.
6
,
085002
(
2022
).
9.
H.
Takatsu
,
H.
Yoshizawa
, and
Y.
Maeno
, “
Comparative study of conductive delafossites with and without frustrated spins on a triangular lattice, PdMO2 (M = Cr; Co)
,”
J. Phys.: Conf. Ser.
145
,
012046
(
2009
).
10.
H.
Takatsu
and
Y.
Maeno
, “
Single crystal growth of the metallic triangular-lattice antiferromagnet PdCrO2
,”
J. Cryst. Growth
312
,
3461
3465
(
2010
).
11.
H. J.
Noh
,
J.
Jeong
,
B.
Chang
,
D.
Jeong
,
H. S.
Moon
,
E. J.
Cho
,
J. M.
Ok
,
J. S.
Kim
,
K.
Kim
,
B. I.
Min
,
H. K.
Lee
,
J. Y.
Kim
,
B. G.
Park
,
H. D.
Kim
, and
S.
Lee
, “
Direct observation of localized spin antiferromagnetic transition in PdCrO2 by angle-resolved photoemission spectroscopy
,”
Sci. Rep.
4
,
3680
(
2014
).
12.
M. D.
Le
,
S.
Jeon
,
A. I.
Kolesnikov
,
D. J.
Voneshen
,
A. S.
Gibbs
,
J. S.
Kim
,
J.
Jeong
,
H. J.
Noh
,
C.
Park
,
J.
Yu
,
T. G.
Perring
, and
J. G.
Park
, “
Magnetic interactions in PdCrO2 and their effects on its magnetic structure
,”
Phys. Rev. B
98
,
024429
(
2018
).
13.
V.
Sunko
,
F.
Mazzola
,
S.
Kitamura
,
S.
Khim
,
P.
Kushwaha
,
O. J.
Clark
,
M. D.
Watson
,
I.
Marković
,
D.
Biswas
,
L.
Pourovskii
,
T. K.
Kim
,
T. L.
Lee
,
P. K.
Thakur
,
H.
Rosner
,
A.
Georges
,
R.
Moessner
,
T.
Oka
,
A. P.
Mackenzie
, and
P. D.
King
, “
Probing spin correlations using angle-resolved photoemission in a coupled metallic/Mott insulator system
,”
Sci. Adv.
6
,
eaaz0611
(
2020
).
14.
F.
Mazzola
,
V.
Sunko
,
S.
Khim
,
H.
Rosner
,
P.
Kushwaha
,
O. J.
Clark
,
L.
Bawden
,
I.
Marković
,
T. K.
Kim
,
M.
Hoesch
,
A. P.
Mackenzie
, and
P. D. C.
King
, “
Itinerant ferromagnetism of the Pd-terminated polar surface of PdCoO2
,”
Proc. Natl. Acad. Sci. U. S. A.
115
,
12956
12960
(
2018
).
15.
V.
Sunko
,
H.
Rosner
,
P.
Kushwaha
,
S.
Khim
,
F.
Mazzola
,
L.
Bawden
,
O. J.
Clark
,
J. M.
Riley
,
D.
Kasinathan
,
M. W.
Haverkort
,
T. K.
Kim
,
M.
Hoesch
,
J.
Fujii
,
I.
Vobornik
,
A. P.
Mackenzie
, and
P. D. C.
King
, “
Maximal Rashba-like spin splitting via kinetic-energy-coupled inversion-symmetry breaking
,”
Nature
549
,
492
496
(
2017
).
16.
F.
Mazzola
,
C. M.
Yim
,
V.
Sunko
,
S.
Khim
,
P.
Kushwaha
,
O. J.
Clark
,
L.
Bawden
,
I.
Marković
,
D.
Chakraborti
,
T. K.
Kim
,
M.
Hoesch
,
A. P.
Mackenzie
,
P.
Wahl
, and
P. D.
King
, “
Tuneable electron–magnon coupling of ferromagnetic surface states in PdCoO2
,”
npj Quantum Mater.
7
,
20
(
2022
).
17.
P. F.
Carcia
,
R. D.
Shannon
,
P. E.
Bierstedt
, and
R. B.
Flippen
, “
O2 electrocatalysis on thin film metallic oxide electrodes with the delafossite structure
,”
J. Electrochem. Soc.
127
,
1974
1978
(
1980
).
18.
T.
Harada
,
K.
Fujiwara
, and
A.
Tsukazaki
, “
Highly conductive PdCoO2 ultrathin films for transparent electrodes
,”
APL Mater.
6
,
046107
(
2018
).
19.
P.
Yordanov
,
W.
Sigle
,
P.
Kaya
,
M. E.
Gruner
,
R.
Pentcheva
,
B.
Keimer
, and
H.-U.
Habermeier
, “
Large thermopower anisotropy in PdCoO2 thin films
,”
Phys. Rev. Mater.
3
,
085403
(
2019
).
20.
M.
Brahlek
,
G.
Rimal
,
J. M.
Ok
,
D.
Mukherjee
,
A. R.
Mazza
,
Q.
Lu
,
H. N.
Lee
,
T. Z.
Ward
,
R. R.
Unocic
,
G.
Eres
, and
S.
Oh
, “
Growth of metallic delafossite PdCoO2 by molecular beam epitaxy
,”
Phys. Rev. Mater.
3
,
093401
(
2019
).
21.
J. M.
Ok
,
M.
Brahlek
,
W. S.
Choi
,
K. M.
Roccapriore
,
M. F.
Chisholm
,
S.
Kim
,
C.
Sohn
,
E.
Skoropata
,
S.
Yoon
,
J. S.
Kim
, and
H. N.
Lee
, “
Pulsed-laser epitaxy of metallic delafossite PdCrO2 films
,”
APL Mater.
8
,
051104
(
2020
).
22.
J.
Sun
,
M. R.
Barone
,
C. S.
Chang
,
M. E.
Holtz
,
H.
Paik
,
J.
Schubert
,
D. A.
Muller
, and
D. G.
Schlom
, “
Growth of PdCoO2 by ozone-assisted molecular-beam epitaxy
,”
APL Mater.
7
,
121112
(
2019
).
23.
T.
Harada
,
K.
Sugawara
,
K.
Fujiwara
,
M.
Kitamura
,
S.
Ito
,
T.
Nojima
,
K.
Horiba
,
H.
Kumigashira
,
T.
Takahashi
,
T.
Sato
, and
A.
Tsukazaki
, “
Anomalous Hall effect at the spontaneously electron-doped polar surface of PdCoO2 ultrathin films
,”
Phys. Rev. Res.
2
,
013282
(
2020
).
24.
Q.
Song
,
J.
Sun
,
C. T.
Parzyck
,
L.
Miao
,
Q.
Xu
,
F. V. E.
Hensling
,
M. R.
Barone
,
C.
Hu
,
J.
Kim
,
B. D.
Faeth
,
H.
Paik
,
P. D. C.
King
,
K. M.
Shen
, and
D. G.
Schlom
, “
Growth of PdCoO2 films with controlled termination by molecular-beam epitaxy and determination of their electronic structure by angle-resolved photoemission spectroscopy
,”
APL Mater.
10
,
091113
(
2022
).
25.
R.
Wei
,
P.
Gong
,
M.
Zhao
,
H.
Tong
,
X.
Tang
,
L.
Hu
,
J.
Yang
,
W.
Song
,
X.
Zhu
, and
Y.
Sun
, “
Solution-processable epitaxial metallic delafossite oxide films
,”
Adv. Funct. Mater.
30
,
2002375
(
2020
).
26.
Y.
Tarn
,
M.
Sinha
,
C.
Pasco
,
D. G.
Schlom
,
T. M.
McQueen
,
K. M.
Shen
, and
B. D.
Faeth
, “
An algorithm for subtraction of doublet emission lines in angle-resolved photoemission spectroscopy
,”
J. Electron Spectrosc. Relat. Phenom.
265
,
147323
(
2023
).
27.
H.-J.
Noh
,
J.
Jeong
,
J.
Jeong
,
E.-J.
Cho
,
S. B.
Kim
,
K.
Kim
,
B. I.
Min
, and
H.-D.
Kim
, “
Anisotropic electric conductivity of delafossite PdCoO2 studied by angle-resolved photoemission spectroscopy
,”
Phys. Rev. Lett.
102
,
256404
(
2009
).
28.
V.
Sunko
,
Angle Resolved Photoemission Spectroscopy of Delafossite Metals
, 1st ed. (
Springer
,
Cham
,
2019
).
29.
T.
Harada
, “
Thin-film growth and application prospects of metallic delafossites
,”
Mater. Today Adv.
11
,
100146
(
2021
).

Supplementary Material