Control of Impurity Phase Segregation in a PdCrO2/CuCrO2 Heterostructure
Abstract
PdCrO2 films are synthesized on CuCrO2 buffer layers on Al2O3 substrates. This synthesis is accompanied by impurity phase segregation, which hampers the synthesis of high quality PdCrO2 films. The potential causes of impurity phase segregation were studied by using a combination of experiments and ab initio calculations. X-ray diffraction and scanning transmission electron microscopy experiments revealed impurity phases of CuxPd1-x alloy and chromium oxides, Cr2O3 and Cr3O4, in PdCrO2. Calculations determined that oxygen deficiency can cause the impurity phase segregation. Therefore, preventing oxygen release from delafossites could suppress the impurity phase segregation. The amounts of Cr2O3 and Cr3O4 depend differently on temperature and oxygen partial pressure. A reasonable theory-based explanation for this experimental observation is provided.
I Introduction
Delafossites are intriguing materials that can combine 2D electronic conductivity in the cation layers and magnetism in slightly distorted octahedra in the O6 layers, which stack alternately [1, 2, 3]. The abundant possible choices of monovalent and trivalent cations lead to a number of delafossite materials with diverse physical properties [4, 5]. The O2 delafossites were first reported in 1971 by a group of the DuPont Experimental Station [1, 2, 3, 4]. A quarter of century after delafossites were first reported, they received renewed attention when the transparent -type semiconductor CuAlO2 was discovered [6, 7]. Simultaneously, Tanaka et al.[8] reported the strong anisotropy of electronic conduction for the metallic PdCoO2 single crystals [3]. One decade later, Takatsu and Maeno et al., working on PdCoO2 and PdCrO2, reported the growth of single crystals of PdCrO2 [9]. These single crystals exhibit intriguing phenomena[4] such as the unconventional anomalous Hall effect in PdCrO2[10] and anomalous temperature dependence of specific heat and electrical resistivity that are driven by high-frequency phonons in PdCoO2[10]. Their seminal work originated the continuous study of delafossite metals to this day.
Delafossite metals have electronic conductivity comparable with the most conductive pure metals [3, 11, 8, 4] owing to their remarkably long electronic mean free paths of up to 20 m [4, 12, 13]. Among delafossite metals, PdCrO2 is especially interesting because it coexists with a layerwise non-collinear spin state [14, 15, 16, 17, 18] and exhibits high electronic conductivity [16]. Its topological properties, primarily caused by spin–orbit coupling in Pd, allow for the observation of an unconventional anomalous Hall effect [10, 19] in bulk PdCrO2. Additionally, PdCrO2 films and surfaces have been studied. Angle-resolved photoemission spectroscopy experiments showed that Pd-terminated PdCrO2 has surface ferromagnetism, which may originate from the Stoner-like instability[20, 21]. Experimental studies of PdCrO2 films established that the antiferromagnetic spin state remains stable down to a thickness of 3.6 nm [22].
Hybrid layered heterostructures, composed of PdCrO2 and other delafossite materials, could exhibit interesting and different phenomena than their parent compounds [23]. However, despite the interest in the material, the epitaxial growth of PdCrO2 films has not been widely studied [24, 25, 26, 27, 22]. The growth of PdCrO2 films on Al2O3 is sometimes accompanied by impurity phases (i.e., CuxPd1-x alloy and chromium oxides) [22]. Recent research discovered that a one-monolayer buffer layer of CuCrO2 on an Al2O3 substrate suppresses this instability [22]. However, a nonnegligible amount of impurity phase is still formed. Understanding the mechanism of the impurity phase segregation and how to suppress it is highly desired for the growth of heterostructures containing PdCrO2 or other Pd-based delafossites.
In this work, the mechanism of impurity phase segregation of a heterostructure of a PdCrO2 layer with a CuCrO2 buffer layer on an Al2O3 substrate was studied using a combination of experiments and ab initio calculations. X-ray diffraction (XRD) and scanning transmission electron microscopy (STEM) experiments were performed, and the segregation of CuxPd1-x alloy and chromium oxide (Cr2O3 and Cr3O4) impurity phases was observed. These experiments revealed that the formation of Cr2O3 negatively correlates with oxygen partial pressure, whereas the formation of Cr3O4 does not correlate with oxygen partial pressure. Moreover, the Cr2O3 (Cr3O4) formation weakly (strongly) positively correlates with temperature. The segregation of CuxPd1-x alloy and chromium oxide impurity phases must be accompanied by the appearance or disappearance of point defects because the segregation processes are not stoichiometric. In this scenario, calculations revealed that oxygen vacancies can cause the impurity phase segregation. Calculations also revealed that the segregation of Cr2O3 or Cr3O4 is energetically the most favorable among the chromium oxides, agreeing with the experiments described in Section III.1. Finally, the calculations also revealed that the formation of Cr2O3 and Cr3O4 depends on temperature and oxygen partial pressure.
II Experimental and calculation details
II.1 Experimental details
A PdCrO2 layer with thickness of approximately 10 nm was grown on a one-monolayer (0.38 nm) CuCrO2 buffer layer on an Al2O3 substrate via pulsed laser deposition using polycrystalline targets. Before the film growth, commercially available Al2O3 (0001) substrates (CrysTec, Germany) were annealed at 1100 ∘C for 1 h to achieve atomically flat surfaces with step-terrace structure. For PdCrO2 films, the growth conditions were widely varied: temperature () was 500–800 ∘C, and oxygen partial pressure () was 10–500 mTorr. The repetition rate and fluence of KrF excimer laser ( = 248 nm) were fixed at 5 Hz and 1.5 J/cm2, respectively. The cross-sectional STEM specimens were prepared using low-energy ion milling at LN2 temperature after mechanical polishing. High-angle annular dark field (HAADF) STEM measurements were performed on a Nion UltraSTEM200 operated at 200 kV. The microscope is equipped with a cold-field emission gun and a third- and fifth-order aberration corrector for sub-angstrom resolution. The convergence half-angle of 30 mrad was used, and the inner angle of the HAADF STEM was approximately 65 mrad.
II.2 Calculation details
Density functional theory (DFT) implemented in the VASP package [28] was used to understand the energetics of competing phases during the experimental growth process. The Perdew–Burke–Ernzerhof (PBE)+ method [29, 30] was used. The Hubbard correction was applied to the 3 shell of the Cr atoms. The value was 3.3 eV, which was optimized compared with the results of the HSE06 functional [31], as described in the Supporting Information. The core electrons were replaced with pseudopotentials made by the projector-augmented wave method accompanied by the VASP code [32, 33, 34]. The cutoff energy was 520 eV, and -spacing was 0.30 Å-1, which converged the Cr vacancy formation energy in CuCrO2 within 2 meV. Experimental lattice vectors for CuCrO2 [35], PdCrO2 [36], and Al2O3 were used 111The room temperature structure of Ref. [47] was used. The lattice vectors reported in the Materials project [38] were used for chromium oxides and chromium metal 222The database ID for each material: Cr = mp-90, CrO = mp-19091, Cr3O4 = mp-756253, and Cr2O3 = mp-19399.. The atomic coordinates were relaxed for the functional. The convergence criteria for the self-consistent field and ionic cycles were eV and eV, respectively.
III Results and Discussions
III.1 Segregation of impurity phases
Impurity phases including CuxPd1-x, Cr2O3, and Cr3O4 have been observed experimentally [40]. In Figure 1 we show in addition to 2– XRD spectrum, the intensity of the XRD data as a function of the growth conditions. As reported in Ref. [22], the high-quality PdCrO2 films can be achieved only within a relatively narrow growth window. Outside the growth window, the metallic properties are severely deteriorated by the impurity formation. The resistance could not be measured because of the high resistivity. The rectangular boxes in Figure 1 highlight the main impurities observed in XRD: Cr3O4 and Cr2O3. The bottom two panels of Figure 1 map the XRD intensities of Cr3O4 and Cr2O3 for temperature and oxygen partial pressure. The relative abundances between Cr3O4 and Cr2O3 are difficult to assess quantitatively using the XRD intensities because the XRD reflectivity varies with substances and angles. However, we use the intensities to assess qualitatively how the formation of each substance is affected by growth conditions. The XRD intensity of Cr3O4 strongly positively correlates with temperature (correlation coefficient 333 The correlation coefficient indicates the strength of correlation between two data-series, and . can be from -1 to +1. = +1 (-1) indicates the complete positive (negative) correlation between and . The correlation coefficient is defined by (1) Here, and are the standard deviations of and . is the covariance of the two data-series: (2) ), whereas the correlation between the XRD intensity of Cr2O3 and temperature is weak (). Moreover, Cr3O4’s peak strength does not depend on oxygen partial pressure (), but Cr2O3’s peak strength negatively depends on oxygen partial pressure (). These results are compared with our calculations in the last paragraph of Section III.6.
III.2 Point-defect formation energy
Because the system segregates oxygen-deficient oxides Cr2O3 or Cr3O4, the segregation process should be accompanied by the appearance or disappearance of point defects. The ratio of O to Cr (O/Cr) in Cr2O3 is . In Cr3O4, is about . In the delafossite materials, is 2.0. Therefore, the formation energies of multiple types of point self-defects in bulk Al2O3, CuCrO2, and PdCrO2 were calculated. To simplify the problem, defects in the bulk were calculated, even though samples with different thickness of PdCrO2 and CuCrO2 films have been grown in this and other works [40, 22].
The following describes the notation used for the point defects and explains how to evaluate the formation energies. indicate vacancies in the Cr, Cu, Pd, and O sites. The Cu (Pd) replacement defects in the Cr sites, or antisite defects, are indicated by CuCr (PdCr). Larger defect complexes such as CuCr& (PdCr&) can form when Cu (Pd) atoms move to a preformed , leaving the . The formation energies of these defects are given as follows: for CuCrO2,
(3) | |||||
(4) | |||||
and for PdCrO2,
(5) | |||||
(6) | |||||
Here, (CuCrO2)bulk and (PdCrO2)bulk are the total energies of pristine delafossite structures. (CuCrO2)X and (PdCrO2)X are the total energies of structures with type- defects. The chemical potential of atomic species is . The oxygen vacancy formation energy in the Al2O3 substrate was also calculated by
(7) |
Our experiments observed the Cu-Pd alloy and Cr oxide impurity phases on the composite sample of Al2O3, CuCrO2, and PdCrO2 (see § III.1). The defect formation energies should be evaluated for the experimental conditions: the chemical equilibrium states consisting of Al2O3, CuCrO2, PdCrO2, CuxPd1-x, and a chromium oxide. The exact value of in CuxPd1-x is not known experimentally. The ratio potentially depends on the volume comparison of CuCrO2 and PdCrO2. However, the change in the results is negligible when changes from 0.5 to 0.25 or 0.75 (variations of only 0.33 eV were observed), as described in the Supporting Information. Therefore, the results reported below assumed . Solving the following equations yields the chemical potentials. For example, if Al2O3, CuCrO2, PdCrO2, CuPd, and Cr3O4 coexist, then
(8) | |||||
(9) | |||||
(10) | |||||
(11) | |||||
(12) |
There exist as many independent linear equations as unknown chemical potentials, so the chemical potentials are trivially determined.
III.3 Formation energies of defects as a function of the chemical potentials
The formation energies of point defects in Al2O3, CuCrO2, and PdCrO2 were calculated for different chromium oxides, as described in Section III.2. We also considered the Cr metal as the Cr source of the Cr-rich limit. The results are summarized in Figure 2. The point-defect formation energies are all positive, so CuCrO2 and PdCrO2 are thermodynamically stable and stoichiometric under the considered chemical conditions. For low values of the oxygen chemical potential ( eV), the in CuCrO2 and PdCrO2 have the lowest formation energies; the in Al2O3 is much higher. As soon as the oxygen chemical potential increases, the and become the lowest formation energy defects. The Cr vacancies, by contrast, have much higher formation energies. The experimental chemical potentials are not well defined because the system is out of equilibrium, as described in Section III.2. However, each element’s stability corresponds to anywhere between the vertical lines that correspond to the oxygen chemical potentials with and . The , &, and & are all Cr-deficient point defects. For CuCrO2, the formation energy of & is lower than that of . Therefore, the Cr site does not have a vacancy because a neighboring Cu occupies the Cr site by forming next to .
III.4 Instability of CuCrO2 and PdCrO2 for oxygen-deficient samples
Experiments found the segregation of impurity phases of CuxPd1-x, Cr2O3, and Cr3O4 on a 10 nm PdCrO2 layer with a one-monolayer CuCrO2 buffer layer on an Al2O3 substrate. The samples were grown under low oxygen partial pressures. The simultaneous presence of seven compunds (CuCrO2, PdCrO2, CuxPd1-x, Cr2O3, Cr3O4, O2, and Al2O3) but only five chemical elements complicates the theoretical analysis. Finding a solution for the chemical potential equations is impossible when the equations outnumber the independent variables. In this case, the system is out of equilibrium. The chemical potentials may not be uniform throughuot the sample. For instance, near the surface, the oxygen chemical potential may be a function of temperature and oxygen partial pressure. By contrast, near the regions where Cr2O3 and Cr3O4 coexist, the chemical potentials of Cr and O are uniquely determined by the formation energies of the two solids. Alternatively, near the Al2O3 substrate, the oxygen chemical potential may be determined by temperature and the concentration of oxygen vacancies in Al2O3.
The impurity phases are oxygen deficient (i.e., chromium rich) relative to CuCrO2 and PdCrO2: the O/Cr ratios of Cr2O3 (1.5) and Cr3O4 (1.3) are smaller than that of CuCrO2 and PdCrO2 (2.0). To elucidate the possible cause of impurity phase segregation, several different possible reactions originating from out-of-equilibrium states were considered. Then their potential to destabilize CuCrO2 and PdCrO2 was examined. The analysis revealed that low oxygen partial pressures and high temperatures could explain the segregation of Cr2O3 and CuxPd1-x. Preexisting defects as energetic as oxygen vacancies in Al2O3 could enhance the segregation of Cr3O4 and CuxPd1-x.
III.5 Thermochemical reactions
To simplify the analysis, the CuxPd1-x alloy is assumed to be CuPd, as described in the last paragraph of Section III.2. The theoretical approach shows that the combination of CuCrO2 and PdCrO2 is stable against the CrO2 and CuPd impurity phase segregation, which is a stoichiometric process. The thermochemical equation of this segregation is
(13) | |||||
where is the energy gained, or lost if negative, to form CrO2. The value of was calculated to be eV per two formula units of CrO2, so this reaction is endothermic.
By contrast, the Cr/O ratios of CuCrO2 and PdCrO2 vs. Cr2O3 or Cr3O4 are different. Therefore, the impurity phase segregation may be caused by an impurity-absorbing defect. For Cr2O3+CuPd, the impurity phase segregation may be caused and promoted by an oxygen-adsorbent mechanism because the Cr/O ratios of CuCrO2 and PdCrO2 (1/2) and Cr2O3 (2/3) are different. This oxygen deficiency may be the result of low environmental oxygen concentration relative to chromium from either (i) defective CuCrO2, PdCrO2, or Al2O3 or (ii) low oxygen content in the vacuum growth chamber 444The samples were grown under mTorr oxygen partial pressure.. For mechanism (i), preexisting VO in CuCrO2, PdCrO2, or Al2O3 and formation of Cr-deficient defects such as VCr in CuCrO2 or PdCrO2 were considered to keep the Cr/O ratios constant before and after the process 555 Preexisting point defects with extra Cr, such as and , were not considered because their existence explains the appearance of CrxO1-x but not CuyPd1-y without involving multiple types defects to balance the chemical reaction. .
Therefore, the energy gain obtained by the (dis)appearance of point defects in CuCrO2, PdCrO2, and Al2O3 was compared with the release of oxygen molecules into the oxygen gas in the growth chamber. All these possibilities were considered as particle exchanges with a particle bath.
The energy cost of taking an atom ( O or Cr) from one of these particle baths is defined as
(14) |
Here, and are the energies of the particle bath without defects and with an O or Cr vacancy, respectively 666 is conceptually similar to the chemical potential of , but chemical potentials are not defined for a nonequilibrium process..
The thermochemical equations for the segregation of Cr2O3 when introducing O to or removing Cr from the particle bath are given as follows:
(15) | |||||
(16) | |||||
In equation (15), the term takes into account the effect of removing an oxygen vacancy in the particle bath, and considers the effect of creating a fraction of Cr vacancies in the particle bath.
Similarly, the segregation of Cr3O4 could be explained by the following reactions:
(17) | |||||
(18) | |||||
Their derivations are described in detail in Appendix A.
The exothermic energies, , are shown in eqs (15)–(18), for different values of and , depending on the particle baths in Table 1. The table shows that only and can be positive (i.e., exothermic reaction), whereas the reactions involving the formation of Cr-deficient defects are always endothermic. Therefore, the preexisting oxygen vacancies could explain the spontaneous segregation of Cr2O3, Cr3O4, and CuPd impurity phases.
Figure 3 shows and in Table 1 for different (i.e., different particle bath). The stability of oxygen atoms in each particle bath negatively correlates with : oxygen atoms are the most (least) stable in (O2 gas). changes depending on , as given in eq (15). Similarly, changes depending on , as given in eq (17). The impurity phase segregation is endothermic when O2 gas is the particle bath and exothermic for the other particle baths. The energetically favored chromium oxide changes from to with decreasing from to .
Bath location | ||||
---|---|---|---|---|
PdCrO2 | +2.703 | 0.990 | +2.466 | 3.192 |
CuCrO2 | +3.308 | 0.989 | +3.273 | 3.189 |
Al2O3 | +5.666 | – | +6.417 | – |
O2 in vacuum () | 1.619 | – | 3.297 | – |
III.6 Entropy contributions to the formation of
Cr2O3 and Cr3O4
In this section, entropy contributions to the positive and are considered. For convenience, and .
Entropy contributions depend on the temperature, point-defect densities, and oxygen partial pressure. The energies were replaced by Helmholtz free energies in eqs (15) and (17). For bulk structures, was evaluated by
(19) |
Here, is the vibrational free energy. For in a defective solid, the vacancy configurational entropy contribution was considered in addition to . If the vacancy density is , then the free energy change when removing one vacancy is
(20) |
(details in Appendix B). Here, is the Boltzmann constant. Therefore, and depend on vacancy density and temperature when the bath location is a defective solid.
When considering the case of O2 released into the growth chamber, because the experimental oxygen partial pressure is very low and the temperature is high, the translational entropy contribution could significantly stabilize the oxygen gas. This stabilization may change and from negative to positive. Without entropy contributions, they are negative, as shown in Table 1. The Helmholtz free energy of the oxygen gas per molecule is defined as
(21) |
Here, is the energy of an isolated oxygen molecule and , , and are free energies by vibrational, rotational, and translational entropies, respectively. Then and [45] are given by
(22) | |||||
(23) | |||||
(24) |
Here, is the Planck constant, and is the moment of inertia of an oxygen molecule. Therefore, and depend on oxygen partial pressure and temperature when the bath location is the dilute oxygen gas.
The entropy contributions in eqs (15) and (17) yield and for different bath locations and conditions. In these equations, the bulk free energies depend on only the temperature. When the bath is a defected crystal, depends on temperature and vacancy density. When the bath is the oxygen gas, depends on temperature and oxygen partial pressure. When ’s entropy contributions are ignored, and barely depend on the temperature: and do not change more than 60 meV from 600 to 1000 K, and does not change more than 11 meV from 600 to 1000 K. Therefore, the dependence of and on the conditions is almost equivalent to that of .
To understand the conditions under which different oxides might be generated experimentally, different baths for exchanging oxygen were systematically considered. When the bath is a defected crystal, was calculated for vacancy densities in the range of – per site and temperatures in the range of 600–1000 K. When the bath is the oxygen gas, was calculated for oxygen partial pressures in the range of – atm and temperatures in the range of 600–1000 K. Figure 4 shows a map of the calculated for different bath locations. The vertical width of each area indicates the variation width corresponding to vacancy densities in the range of – per site or oxygen partial pressures in the range of – atm. Figure 4 is divided into the three regions I–III according to the corresponding and values. The region I is : the impurity phase segregation does not proceed spontaneously. The region II is and : is spontaneously predominantly formed. The region III is : is spontaneously predominantly formed. Therefore, when the particle bath is oxygen gas, CuCrO2, or PdCrO2, the majority of chromium oxide is . When the particle bath is , the majority of chromium oxide is .
Furthermore, other bath locations than the above listed are realistically possible. For example, an oxygen-terminated PdCrO2 surface could lead to very high . By contrast, a Pd-terminated surface could lead to very low . Investigation of such further complicated mechanisms is a possible future work for theory and experiments.
These calculations revealed that in O2 gas decreases with decreasing oxygen partial pressure, and in defected crystals decreases with increasing (details in supporting information). This result is not surprising because it indicates that gaseous oxygen molecules are more stable under oxygen-poor conditions. In reality, would negatively correlates with oxygen partial pressure, so positively correlates with oxygen partial pressure in every bath location: lower oxygen partial pressure facilitates the segregation of impurity phases. This analysis agrees with the experimental finding described in Section III.1: formation negatively correlates with oxygen partial pressure.
However, this analysis does not explain the independence of formation on oxygen partial pressure. Rather, formation strongly depends on temperature, unlike . Some hypotheses are considered to explain the experimental results. (i) Most of the bath locations belong to region II in Figure 4, and the temperature determines how much the metastable Cr3O4 is segregated. (ii) Some of the bath locations belong to region III, but high barrier energy is required to transfer oxygen atoms, moving oxygen vacancies, so the temperature determines the oxygen exchange rate.
For example, the barrier energy required to transfer oxygen atoms from Al2O3 to the surface would be higher than from PdCrO2 to the surface. To verify the hypotheses, saddle state analyses by methods such as the nudged elastic band method, molecular dynamics, and/or modeling the sample’s surface should be applied in future works.
IV Conclusion
The mechanism of impurity phase segregation with the epitaxial growth of a PdCrO2 layer on a CuCrO2 buffer layer on an Al2O3 substrate was investigated via a combination of experiments and ab initio calculations. XRD experiments revealed the formation of CuxPd1-x alloy and chromium oxide (Cr2O3 and Cr3O4) impurity phases. Consequently, the impurity phase segregation should be involved with appearance or disappearance of point defects or oxygen migration because the possible segregation processes are not stoichiometric. In this scenario, several possible mechanisms of impurity phase segregation were considered with oxygen vacancy disappearance or chromium vacancy appearance into different particle baths: Al2O3, CuCrO2, PdCrO2, and the dilute oxygen gas. Calculations established that the oxygen vacancy consumption processes are energetically favorable and supported experimental evidence that Cr2O3 or Cr3O4 are the predominant chromium oxide impurity phases. Specifically, preventing the release of oxygen atoms from delafossite materials could suppress the impurity phase segregation.
Appendix
IV.1 Derivation of eqs (15) and (16)
For ease of explanation, let the particle bath be PdCrO2. Consider the following thermochemical equations for the segregation of Cr2O3 by removing preexisting O vacancies or creating Cr vacancies.
(25) | |||||
and
(26) | |||||
Here, for example, is the energy of f.u. bulk CuCrO2, is the energy of f.u. PdCrO2 with an oxygen vacancy, is the energy of f.u. PdCrO2 with a fraction of chromium vacancies, and is the energy of 1 f.u. bulk CuPd. For the bulk, the following relationships hold according to the definitions.
(27) | |||||
(28) |
Define the energy gain from removing oxygen or chromium vacancies in as follows:
(29) | |||||
(30) |
In the thermodynamic limit (), the following relationships should hold:
(31) | |||||
(32) |
Moreover, define
(33) |
These are the defined in eq (14). Applying eqs (27)–(33) to eqs (25) and (26), yields eqs (15) and (16).
IV.2 Configurational entropy of removing a vacancy.
When vacancies exist in sites, the configurational entropy is
(34) |
The entropy change achieved by adding one vacancy is given by
(35) | |||||
(36) |
For the limit of with fixed ,
(37) |
Therefore, the free energy change achieved by removing one vacancy is given by
(38) | |||||
(39) |
V Acknowledgments
We acknowledge E. Heinrich for valuable help with manuscript preparation. This work was supported by the US Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division (theory and synthesis) and as part of the Computational Materials Sciences Program and Center for Predictive Simulation of Functional Materials (structural characterization). VESTA [46] was used to draw the crystal structures.
References
- Shannon et al. [1971a] R. D. Shannon, D. B. Rogers, and C. T. Prewitt, Chemistry of noble metal oxides. I. Syntheses and properties of ABO2 delafossite compounds, Inorganic Chemistry 10, 713 (1971a), https://doi.org/10.1021/ic50098a011 .
- Shannon et al. [1971b] R. D. Shannon, C. T. Prewitt, and D. B. Rogers, Chemistry of noble metal oxides. II. Crystal structures of platinum cobalt dioxide, palladium cobalt dioxide, copper iron dioxide, and silver iron dioxide, Inorganic Chemistry 10, 719 (1971b), https://doi.org/10.1021/ic50098a012 .
- Shannon et al. [1971c] R. D. Shannon, D. B. Rogers, C. T. Prewitt, and J. L. Gillson, Chemistry of noble metal oxides. III. Electrical transport properties and crystal chemistry of ABO2 compounds with the delafossite structure, Inorganic Chemistry 10, 723 (1971c), https://doi.org/10.1021/ic50098a013 .
- Mackenzie [2017] A. P. Mackenzie, The properties of ultrapure delafossite metals, Reports on Progress in Physics 80, 032501 (2017).
- Abdelhamid [2015] H. N. Abdelhamid, Delafossite nanoparticle as new functional materials: Advances in energy, nanomedicine and environmental applications, Materials Science Forum 832, 28 (2015).
- Kawazoe et al. [1997] H. Kawazoe, M. Yasukawa, H. Hyodo, M. Kurita, H. Yanagi, and H. Hosono, -type electrical conduction in transparent thin films of CuAlO2, Nature 389, 939 (1997).
- Yanagi et al. [2000] H. Yanagi, S. ichiro Inoue, K. Ueda, H. Kawazoe, H. Hosono, and N. Hamada, Electronic structure and optoelectronic properties of transparent -type conducting CuAlO2, Journal of Applied Physics 88, 4159 (2000).
- Tanaka et al. [1996] M. Tanaka, M. Hasegawa, and H. Takei, Growth and Anisotropic Physical Properties of PdCoO2 Single Crystals, Journal of the Physical Society of Japan 65, 3973 (1996), https://doi.org/10.1143/JPSJ.65.3973 .
- Takatsu and Maeno [2010] H. Takatsu and Y. Maeno, Single crystal growth of the metallic triangular-lattice antiferromagnet PdCrO2, Journal of Crystal Growth 312, 3461 (2010).
- Takatsu et al. [2010] H. Takatsu, S. Yonezawa, S. Fujimoto, and Y. Maeno, Unconventional Anomalous Hall Effect in the Metallic Triangular-Lattice Magnet PdCrO2, Phys. Rev. Lett. 105, 137201 (2010).
- Tanaka et al. [1997] M. Tanaka, M. Hasegawa, and H. Takei, Crystal growth of PdCoO2, PtCoO2 and their solid-solution with delafossite structure, Journal of Crystal Growth 173, 440 (1997).
- Hicks et al. [2012] 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).
- Kushwaha et al. [2015] P. Kushwaha, V. Sunko, P. J. W. 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 5 delafossite oxide metal, Science Advances 1, 10.1126/sciadv.1500692 (2015).
- Doumerc et al. [1986] J.-P. Doumerc, A. Wichainchai, A. Ammar, M. Pouchard, and P. Hagenmuller, On magnetic properties of some oxides with delafossite-type structure, Materials Research Bulletin 21, 745 (1986).
- Mekata et al. [1995] M. Mekata, T. Sugino, A. Oohara, Y. Oohara, and H. Yoshizawa, Magnetic structure of antiferromagnetic PdCrO2 possible degenerate helices on a rhombohedral lattice, Physica B: Condensed Matter 213-214, 221 (1995).
- Takatsu et al. [2009] H. Takatsu, H. Yoshizawa, S. Yonezawa, and Y. Maeno, Critical behavior of the metallic triangular-lattice heisenberg antiferromagnet PdCrO2, Phys. Rev. B 79, 104424 (2009).
- Takatsu et al. [2014] H. Takatsu, G. Nénert, H. Kadowaki, H. Yoshizawa, M. Enderle, S. Yonezawa, Y. Maeno, J. Kim, N. Tsuji, M. Takata, Y. Zhao, M. Green, and C. Broholm, Magnetic structure of the conductive triangular-lattice antiferromagnet PdCrO2, Phys. Rev. B 89, 104408 (2014).
- Billington et al. [2015] D. Billington, D. Ernsting, T. E. Millichamp, C. Lester, S. B. Dugdale, D. Kersh, J. A. Duffy, S. R. Giblin, J. W. Taylor, P. Manuel, D. D. Khalyavin, and H. Takatsu, Magnetic frustration, short-range correlations and the role of the paramagnetic fermi surface of PdCrO2, Scientific Reports 5, 10.1038/srep12428 (2015).
- Ok et al. [2013] J. M. Ok, Y. J. Jo, K. Kim, T. Shishidou, E. S. Choi, H.-J. Noh, T. Oguchi, B. I. Min, and J. S. Kim, Quantum oscillations of the metallic triangular-lattice antiferromagnet PdCrO2, Phys. Rev. Lett. 111, 176405 (2013).
- Mazzola et al. [2018] 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 (2018).
- Harada [2021] T. Harada, Thin-film growth and application prospects of metallic delafossites, Materials Today Advances 11, 100146 (2021).
- Ok et al. [2020] 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).
- Lechermann and Richter [2020] F. Lechermann and R. Richter, Theoretical design of highly correlated electron states in delafossite heterostructures, Phys. Rev. Research 2, 013352 (2020).
- Harada et al. [2018] T. Harada, K. Fujiwara, and A. Tsukazaki, Highly conductive PdCoO2 ultrathin films for transparent electrodes, APL Materials 6, 046107 (2018).
- Sun et al. [2019] 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 Materials 7, 121112 (2019).
- Brahlek et al. [2019] 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. Materials 3, 093401 (2019).
- Yordanov et al. [2019] 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. Materials 3, 085403 (2019).
- Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54, 11169 (1996).
- Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett. 77, 3865 (1996).
- Dudarev et al. [1998] S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton, Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study, Phys. Rev. B 57, 1505 (1998).
- Heyd et al. [2003] J. Heyd, G. E. Scuseria, and M. Ernzerhof, Hybrid functionals based on a screened coulomb potential, The Journal of Chemical Physics 118, 8207 (2003), https://doi.org/10.1063/1.1564060 .
- Blöchl [1994] P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50, 17953 (1994).
- Kresse and Joubert [1999] G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B 59, 1758 (1999).
- Hobbs et al. [2000] D. Hobbs, G. Kresse, and J. Hafner, Fully unconstrained noncollinear magnetism within the projector augmented-wave method, Phys. Rev. B 62, 11556 (2000).
- Frontzek et al. [2011] M. Frontzek, J. T. Haraldsen, A. Podlesnyak, M. Matsuda, A. D. Christianson, R. S. Fishman, A. S. Sefat, Y. Qiu, J. R. D. Copley, S. Barilo, S. V. Shiryaev, and G. Ehlers, Magnetic excitations in the geometric frustrated multiferroic cucro2, Phys. Rev. B 84, 094448 (2011).
- Le et al. [2018] 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 and their effects on its magnetic structure, Phys. Rev. B 98, 024429 (2018).
- Note [1] The room temperature structure of Ref. [47] was used.
- Jain et al. [2013] A. Jain, S. P. Ong, G. Hautier, W. Chen, W. D. Richards, S. Dacek, S. Cholia, D. Gunter, D. Skinner, G. Ceder, and K. a. Persson, The Materials Project: A materials genome approach to accelerating materials innovation, APL Materials 1, 011002 (2013).
- Note [2] The database ID for each material: Cr = mp-90, CrO = mp-19091, Cr3O4 = mp-756253, and Cr2O3 = mp-19399.
- [40] S. Yoon, J. M. Ok, T. Ichiba, S.-H. Kang, A. Huon, M. Yoon, P. Ganesh, F. A. Reboredo, H. Miao, A. R. Lupini, and H. N. Lee, Templated epitaxy of PdCrO2 delafossites using atomic sacrificial layers (Unpublished).
-
Note [3]
The correlation coefficient
indicates the strength of correlation between two data-series, and . can be
from -1 to +1. = +1 (-1) indicates the complete positive (negative)
correlation between and . The correlation coefficient is defined by
(40) (41) - Note [4] The samples were grown under mTorr oxygen partial pressure.
- Note [5] Preexisting point defects with extra Cr, such as and , were not considered because their existence explains the appearance of CrxO1-x but not CuyPd1-y without involving multiple types defects to balance the chemical reaction.
- Note [6] is conceptually similar to the chemical potential of , but chemical potentials are not defined for a nonequilibrium process.
- Goodstein [1985] D. L. Goodstein, States of matter (Dover Publications, Inc., 1985) pp. 98–105.
- Momma and Izumi [2011] K. Momma and F. Izumi, VESTA3 for three-dimensional visualization of crystal, volumetric and morphology data, Journal of Applied Crystallography 44, 1272 (2011).
- Kondo et al. [2008] S. Kondo, K. Tateishi, and N. Ishizawa, Structural evolution of corundum at high temperatures, Japanese Journal of Applied Physics 47, 616 (2008).