Marc
Kamel
a,
Hatem M.
Titi
b,
Mohamad
Ataya
b,
Antranik
Jonderian
b,
Kirk H.
Bevan
*ac and
Eric
McCalla
*bc
aDivision of Materials Engineering, Faculty of Engineering, McGill University, Montreal, Canada. E-mail: kirk.bevan@mcgill.ca
bDepartment of Chemistry, McGill University, Montreal, Canada. E-mail: eric.mccalla@mcgill.ca
cCentre for the Physics of Materials, McGill University, Montreal, Canada
First published on 30th July 2024
Transparent semiconductors continue to be of high interest for a variety of electronic applications with the ability to tune transport properties being highly sought-after. Herein, we develop novel AgxCoO2−δ delafossite materials with silver content being varied for the first time. The silver content is controlled by the ion-exchange temperature used during synthesis with single-crystal NaCoO2 precursors. By contrast, materials made from polycrystalline precursors show a constant silver content (x = 0.8) regardless of ion-exchange temperature. Importantly, we observe for the first time that the single-crystal precursors in fact separate into many sheets during ion-exchange resulting in partial exfoliation, and this occurs even at low ion-exchange temperatures. The materials were characterized further with X-ray diffraction and X-ray photoemission spectroscopy, both of which were consistent with the Ag content varying both at the surface and the bulk of the material. The results also show a high oxygen deficiency at least at the surface. The impact of varying silver and oxygen content of the crystals was further explored with hall measurements with higher silver content yielding a higher charge carrier (hole) concentration. The samples made from single-crystals showed higher electronic conductivities, hole densities and mobilities than those made from powders. The hole concentration was found to be further tunable by varying the oxygen content of the material by heating in an oxygen atmosphere. The role of silver vacancies in creating holes was confirmed by DFT calculations. Surprisingly, the calculations also indicated that oxygen deficiency results in a band rearrangement that also provides accessible holes at the Fermi level. This study illustrates the benefit that optimized synthesis conditions can have in tuning overall composition to yield optimal transport properties in delafossites. This work also helps elucidate the consequences of competing cationic and anionic off-stoichiometries.
Another delafossite that received considerable attention in the past is AgCoO2.1,16–19 Delafossite AgCoO2 has been synthesized in both polycrystalline20 and crystal forms16 by performing ion-exchange on the precursor NaCoO2. Delmas et al. demonstrated that the stacking of the delafossite can be varied by using various NaCoO2 polymorphs.17 Of highest interest for the current work, two different AgCoO2 polymorphs (D2 and D3) can be obtained in this manner and their structures as well as the structure of their respective precursors are shown in Fig. 1. This shows that the ion-exchange process where Na is replaced by Ag has very little impact on the overall stacking of the CoO2 slabs. Previous work on this delafossite also clearly showed that the materials synthesized by ion-exchange are p-type semiconductors, which is attributed to Ag vacancies acting as p-dopants.18,19 Although this is qualitatively logical, metal vacancies do in fact act as p-dopants while oxygen vacancies act as n-dopants, it does not make quantitative sense as the stoichiometry identified is Ag0.75CoO2 and the charge carrier concentrations are only on the order of 1019 cm−3,18,19,21,22 which certainly do not require 25% metal vacancies to be created, in fact 25% vacancies would result in an extremely high carrier concentration. There are therefore important questions to answer related to how the materials stoichiometry leads to the observed transport properties and this will be addressed in detail herein. To date, there has been no study that explored variations in the stoichiometry of the material and, as far as we know, only the x ∼ 0.75 material has been reported. It is also important to note that, to our knowledge, DFT calculations have only been performed on the stoichiometric AgCoO223 which is shown to be an intrinsic semiconductor, again implying a gap with respect to experiments where clear p-type transport is observed. These earlier studies calculations were done using the GGA functional which is well known to under-estimate band gaps. Hence, we include DFT calculations using the more accurate hybrid functional herein on a variety of compositions to better understand how the values of x and δ affect the band structure and thus the transport properties in AgxCoO2−δ.
Another result of note from prior work on this delafossite is that single-crystals have been reported by ion-exchange.16 Ref. 16 shows SEM images of crystalline AgCoO2 with particle sizes of about 50 μm and showing only top view of the samples. As will be discussed thoroughly herein, when viewed from the side these crystals show significant fracturing during ion-exchange. Importantly, we obtain comparable images from the top using the same synthesis route, but the side views of the crystals that we report here for the first time show that the crystals are now semi-detached flakes and as such they are no longer single-crystals. Thus, in this work, we focus on updating our understanding of AgCoO2 delafossites made from single-crystal precursors. The powder samples are well-known to take the D3 structure shown in Fig. 1, while the single-crystal samples take the D2 structure20 using the Delmas notation. The structural differences between D2 and D3 types lie in the orientation of successive CoO2 slabs as illustrated in Fig. 1. In the D2 structure, all CoO2 cations align along the c-axis, and each CoO2 slab is rotated by π/3 relative to the preceding one. In contrast, the D3 structure exhibits uniform orientation for all CoO2 slabs.24 It is certainly of note that one of the potential applications for AgCoO2 is as a catalyst for oxidation of CO.25 In such applications, high-surface areas are required and thus the D3 powder samples are ideal. However, here we focus on potential electronic applications and it is the samples made from single-crystal precursors (we refer to these as crystal samples) that show tunability in terms of electronic transport properties. By contrast, the powder samples show no variation in the current study and serve as a reference to better understand the changes that are only seen in the crystal samples.
For the synthesis of the larger crystals, the same method was utilized but with NaCoO2 single crystals as the precursors. Here, to make NaCoO2 single crystals, we followed the method of Fujita et al.27 We mixed Na2CO3 with Co3O4 and used NaCl as a flux in a Co3O4:Na2Co3:NaCl mass ratio of 1:7.5:7. The mixture was placed in an Al2O3 crucible (ID = 1.1 cm, OD = 1.3 cm) and covered with a MgO plate and then placed in the furnace to be heated to 950 °C at a rate of 10 °C min−1 with a dwell time of 12 hours followed by a slow cooling at 0.5 °C min−1 down to 850 °C. NaCoO2 crystals were then obtained by washing away the NaCl flux with distilled water. The same ion-exchange process was done for the crystal synthesis, except the crystals were simply covered above and below with pre-mixed AgNO3 and KNO3 powder. Fig. 1 details the experimental procedures, illustrating the workflow for both powders and crystals highlighting their differences throughout synthesis and characterization. This figure shows the contrast in the two precursors, illustrates the setup for the ion-exchange steps, demonstrates the differences in the final morphology and also shows how the contacts were made for the Hall measurements. The powders were pressed into pellets and then indium contacts were added while for the crystals we sputtered contacts onto the crystal using a mask.
Powder X-ray diffraction (PXRD) was performed on the powder samples in a Panalytical Empyrean diffractometer equipped with a Mo anode (60 kV, 40 mA, 10 min scan, 4–30° scattering angles) and a GalaPIX area detector. The crystals were characterized using single-crystal X-ray diffraction (SCXRD) data on a Bruker D8 Venture diffractometer equipped with a Photon 200 CMOS area detector, and IμS microfocus X-ray source (Bruker AXS, CuKα source).
The microstructure and elemental mapping of selected samples were observed by scanning electron microscopy (SEM) using a Hitachi SU3500 microscope at 10 kV equipped with an Al Kα micro-focused monochromator and an energy dispersive spectroscopy (EDS) detector. The X-ray photoelectron spectroscopy (XPS) measurements were done at a pressure of mbar with an X-ray spot size of 200 μm. The survey and high-resolution spectra were collected with a pass energy of 200 and 50 eV, respectively.
We also performed thermogravimetric analysis (TGA) with the TA Discovery 5500 on polycrystalline and crystal samples under oxygen atmosphere. The samples were heated to either 190 °C or 590 °C at a rate of 1 °C min−1, dwelling there for an hour and finally cooling back down to room temperature at a rate of 10 °C min−1. Finally, to measure electronic properties of our samples we used the Hall measurement technique in the Van der Pauw geometry using an Ecopia HMS-3000 device that allowed us to do measurements at room temperature and at 77 K by immersing the sample in liquid nitrogen. For the crystal samples, the contacts were made by first sputtering 10 nm of Ti, then 50 nm of Au, and finally attaching wires using silver epoxy on the corners as shown in Fig. 1f.
Finally, we performed DFT calculations on 4 different structures. The D2 and D3 AgCoO2 structures were obtained from the materials project (mp-19178 for D2 and mp-753001for D3).28 The silver and oxygen deficient structures (D2-Ag0.5CoO2 and D2-AgCoO1.5, respectively) were generated using the program supercell29 using the parent D2-AgCoO2 structure. A 2 × 1 × 1 D2-Ag0.5CoO2 supercell was generated with 14 atoms to represent the silver deficient delafossite. Similarly, a 2 × 1 × 1 D2-AgCoO1.5 oxygen deficient supercell with 14 atoms was generated and shown in Fig. S7 (ESI†) for the oxygen deficient AgCoO1.5 delafossite. The use of the supercell was required in both cases to keep the vacancy concentration low enough to obtain a stable structure. Density functional theory (DFT) was carried out using the vienna ab initio simulation package (VASP) version 5.4.1. The generalized gradient approximation (GGA) in Perdew–Burke–Ernzerhof formulation was used for the structure relaxation. The density of states (DOS) and band diagram plots were generated using the hybrid functional, HSE06 with the standard 0.25 fraction of Hatree–Fock exchange. Based on the cutoff energy and K-point convergence test, the calculations were done with a cutoff energy of 550 eV in a 4 × 8 × 2 Γ-centered mesh. The DOS was plotted using the vasplab package in MATLAB.30 We also used VASPKIT31 code to perform the band structure calculations. The values of the valence band maximum, conduction band minimum and bandgap were calculated using the pymatgen python library.32
By contrast to the powder samples, SCXRD shows significant variation in the crystal samples as we change TIE. Fig. 3a shows that the single-crystal synthesis of the NaCoO2 precursor was successful: we obtained a structure that yields sharp peaks and they index to the P1c space group with lattice parameters a = b = 2.8349(2) Å and c = 10.8916(8) Å consistent with ref. 34, the structure being found in the Crystallography Open Database as COD153298.35 SEM images in Fig. S2 (ESI†) also show that the precursor is made up of large crystals that range in size with the largest being more than 2 mm across. Despite the single-crystal precursor, the SCXRD patterns after ion-exchange at both 180 and 300 °C show arcs in Fig. 3b instead of the individual points associated with scattering from single-crystals. However, the arcs are short in length, showing that the samples are by no means powders, but are no longer single-crystals either. By integrating over these arcs, we produce the simulated PXRD patterns shown in Fig. 2b for these crystals (we will refer to them as crystals throughout this article to distinguish them from the powder samples, but we remind the reader to not confuse them with single-crystals). The results index to the D2-AgCoO2 structure with varying lattice parameters shown in Fig. 3d–f with larger cell volumes as the ion-exchange temperature, TIE, increases. This shows that there is variation in the bulk structure as TIE changes. It is also important to recognize the massive increase in c lattice parameter occurring during ion-exchange (10.89 Å to 12.35 Å represents a 12% increase), while the a-axis does not increase by more than 1.5%. The consequences of this will be discussed extensively below.
Fig. 4 shows SEM images of the D2-AgCoO2 crystals synthesized with various TIE as well as one made without any silver present during the ion-exchange. The results as viewed from above the particles are very consistent with previous literature: the samples look like the single-crystal precursors when viewed from above. However, the profile views show varying extent of exfoliation with layers separating noticeably during the ion-exchange process. This phenomenon was particularly pronounced in crystals subjected to ion-exchange at 300 °C and 180 °C, as evidenced by the SEM images (a, b, d, and e). However, Fig. 4f shows that the exfoliated sheets are first formed during the departure of sodium ions from the crystal, and this process was then further intensified when silver ions were introduced into the structure in Fig. 4d and e. Although this phenomenon has not been reported previously in AgCoO2, in hindsight, it is not surprising considering the 12% increase in c-axis lattice parameter taking place during ion-exchange such that fracturing occurs in that direction only.
Fig. 4 SEM images of D2-AgxCoO2−δ crystals made at 300 °C (a, d), 180 °C (b, e) and 250 °C with no silver added during the ion-exchange step (c, f). |
The exfoliation does not, however, explain why the lattice parameters increase with the ion-exchange temperature. For this, we turn to EDX measurements on the crystals shown in Fig. 2a. Fig. S3 and Table S1 (ESI†) show EDX results for the crystal samples, which indicate that all samples contain very little sodium after the ion-exchange, and the data also shows high tunability of silver content in the crystals. For the polycrystalline samples, the Ag:Co ratio (x) is constant at about 0.8 and in fact shows silver deficiency consistent with ref. 18 and 19. By contrast, in the crystals, the silver content is stoichiometric (x = 1) for high TIE (300 °C) and decreases dramatically all the way down to x = 0.42 at the lowest TIE. Fig. S4 (ESI†) shows that with an ion-exchange temperature of 150 °C, a noisy XRD signal is obtained, and for no silver present we are unable to identify the phases present. Although EDX is a surface measurement, the fact that the bulk lattice parameters increase with TIE strongly supports the conclusion that this non-stoichiometry is systemic and not a surface effect. Whereas we expected lower TIE to lead to slower ion-exchange, it instead leads to a lower equilibrium Ag content. This provides us with a unique opportunity to study a solid solution by controlling the synthesis conditions rather than the stoichiometry of the precursors utilized. It is also, to our knowledge, the first ever study of AgxCoO2 with varied x values.
To further understand the chemistry of these new Ag-deficient structures, we perform XPS to determine the oxidation states of Ag and Co. Fig. 5 and Fig. S5 (ESI†) show XPS patterns of crystals and powders, respectively, synthesized at both low and high ion-exchange temperatures. The figures show the XPS fits and Table 1 and Table S3 (ESI†) show the extracted values for both silver (peaks identified using ref. 36 and 37) and cobalt (peaks identified according to ref. 38 and 39) at low and high TIE. For the powder samples, the average oxidation states for Ag and Co are 1.16 and 2.41, respectively, at 300 °C, giving an overall formula Ag0.8CoO1.67 at the surface of the particles (this is based on SEM/XPS only). Although we are cautious to not fully trust average oxidation states extracted from XPS, this result for the powder samples certainly shows that both metal and oxygen vacancies will need to be taken into account in order to understand the transport in these materials. When doing the synthesis at TIE = 150 °C, the crystals are even more silver deficient than the powder samples, therefore we expect such vacancies to be compensated by higher oxidation of cobalt and silver and/or lower oxygen content. At the lowest silver content of x = 0.42, Table 1 shows an average oxidation state of 2.54 for cobalt and this drops to 2.40 in the sample with x = 1 (TIE = 300 °C) showing oxidation of cobalt does help compensate for the silver deficiency. Similarly, we could see the same trend for silver where the average oxidation state goes from 1.69 to 1.10 at low and high temperature of ion-exchange, respectively, again showing that silver oxidation also helps compensate for silver vacancies and it actually compensates more than the cobalt does. The data reported here therefore demonstrates that the material at the surface (per XPS sensitivity being limited to the surface region only) is both silver and oxygen deficient and the silver content can be tuned in the crystals over a large range (0.4 to 1.0). Below, we discuss consequences of these surface off-stoichiometries in terms of transport properties measured using the Hall technique. Since such measurements are heavily impacted by the surface of the materials, we feel this focus on the surface composition is highly justified. We leave the depth of these non-stoichiometries (i.e. do they extend fully into the bulk?) as an open question worthy of further study, with neutron diffraction being a particularly viable probe of both bulk oxygen and silver.
Fig. 5 XPS patterns of the Ag(3d) and Co(2p) peaks for D2-AgxCoO2−δ crystals made by ion-exchange at 150 °C (a, b) and 300 °C (c, d). |
Sample | Co | Ag |
---|---|---|
Powder TIE = 150 °C | 2.42 | 1.15 |
Powder TIE = 300 °C | 2.41 | 1.16 |
Crystal TIE = 150 °C | 2.54 | 1.69 |
Crystal TIE = 300 °C | 2.4 | 1.1 |
From the structural study performed here we can therefore conclude that both the silver and oxygen content can be highly tuned in delafossite AgxCoO2−δ, and both silver/oxygen non-stoichiometries may therefore play a role in determining charge carrier transport properties. To explore the potential consequences of these compositional variations, DFT was performed on various structures (Fig. 6). Table S4 (ESI†) shows the lattice parameters from DFT and from XRD, while Table S5 (ESI†) shows the relatively small changes in lattice parameters taking place during relaxation. The lattice parameters are in reasonable agreement between experiment and calculation. Band structure and density of states plots were generated for key structures. First, D2 and D3 AgCoO2 show similar features in term of band gap which is around 2.7 eV, as well as the positions of the conduction and valence bands having similar values in both structures. Thus, the vacancy-free AgCoO2 structures show features of intrinsic semiconductors. DFT was also performed on silver deficient D2-Ag0.5CoO2, which has a similar Ag content to the low-Ag crystals we obtained by ion-exchange (0.42). In this calculation, the Fermi level is clearly within the valence band demonstrating p-doping. This is not unexpected, oxygen vacancies are well-known n-dopants, so it follows that metal cation vacancies serve as p-dopants. The bandgap of this system is fairly close to that of the original structure, so we view the Ag-deficient delafossite as a p-doped semiconductor where the Ag deficiencies do not dramatically distort the band structure. Finally, we performed DFT calculation on oxygen vacant samples as we concluded from EDX/XPS that the samples are oxygen deficient. The band structure is dramatically restructured compared to the stoichiometric AgCoO2 band diagrams, such that this level of oxygen deficiency should be viewed as an altered structure altogether. Of highest significance, the Fermi level again crosses the valence band as shown in Fig. 6d such that we expect an electrically metallic material based on this band diagram – with low energy indirect band offsets near the Fermi level that may allow for the promotion of hole carriers as well. This shows that the oxygen vacancies are not acting as simple dopants, there are so many of them that it results in a re-structuring of the bands entirely. The contribution of each element to the band structures can be seen in Fig. S8 (ESI†) showing the projected band diagrams. In the silver deficient material (Fig. S8c, ESI†), both Ag and Co show valence band maxima above the Fermi level. In the case of the low oxygen material, the band rearrangement results in cobalt being heavily present at the top of the valence band which again crosses the Fermi level giving a metallic like band structure.
Fig. 6 Band Structure and density of states plots for (a) D2-AgCoO2, (b) D3-AgCoO2, (c) D2-Ag0.5CoO2 and (d) D2-AgCoO1.5. |
To further understand/confirm the significance of the DFT results and the structural results above, we performed Hall measurements in the van der Pauw configuration in order to determine the transport properties in these delafossites. Fig. 7 and Table 2 show Hall measurement results on both crystals and polycrystalline samples at room temperature and at liquid nitrogen temperature. We extracted bulk charge carrier concentration, mobility, and conductivity for all samples. Firstly, all measurements show p-type conduction in the Hall measurements, largely in agreement with the DFT calculations. The results for the 300 °C crystal (silver content of 1, expected oxygen deficiency based on XPS) show p-type conduction therefore indicates that very low oxygen content can allow for the existence of hole carriers in such a phase as predicted by the DFT calculations. There are also clear trends in the charge carrier concentrations as a function of TIE in Fig. 7. The bulk concentrations of the polycrystalline samples show a near-stable trend (ranging from 5.68 × 1019 to 3.3 × 1019 cm−3) which is consistent with the EDX results that showed a constant silver content in the polycrystalline samples (the small variation seen is attributed to the small difference in the amount of silver metal, Ag0, that tends to decrease with a higher TIE as seen in Table S2, ESI†). We observed the same trend for polycrystalline samples when measured at liquid nitrogen temperatures but with an overall lower bulk concentration. At room temperature there are more charge carriers (holes) which indicates a certain amount of free carrier freeze-out occurring at 77 K. Similar trends are seen in conductivity: a slight decrease in conductivity in the polycrystalline samples can be attributed to the difference in silver metal content depending on TIE, though part of the effect (particularly at 77 K) is due to a small change in mobility. At room temperature, though, mobility does not change significantly (0.10 to 0.15 cm2 V−1 s−1). Overall, the polycrystalline samples show very little tunability in the transport properties, consistent with the low tunability of silver content seen in the structure results above.
x | T IE (°C) | 300 K | 77 K | ||||
---|---|---|---|---|---|---|---|
n (1018 cm −3) | μ (cm2 V−1 s−1) | σ (S cm−1) | n (1018 cm−3) | μ (cm2 V−1 s−1) | σ (S cm−1) | ||
a T indicates the sample was first heated in the TGA in oxygen up to the temperature indicated (either 190 or 590 °C). | |||||||
0.80 PC | 150 | 56.8(1.0) | 0.15 | 1.15 | 2.8(1.2) | 0.392 | 0.0847 |
0.85 PC | 180 | 57.2(1.4) | 0.11 | 0.91 | 3.1(0.5) | 0.227 | 0.103 |
0.82 PC | 215 | 66.9(0.5) | 0.1 | 1.05 | 1.3(0.5) | 0.573 | 0.086 |
0.84 PC | 250 | 56.7(1.5) | 0.12 | 0.8 | 1.5(0.4) | 0.387 | 0.073 |
0.78 PC | 275 | 50.6(0.9) | 0.1 | 0.75 | 2.4(0.9) | 0.344 | 0.045 |
0.84 PC | 300 | 33.0(0.7) | 0.11 | 0.27 | 0.47(0.15) | 0.173 | 0.001 |
PC | 275-T190a | 80.8(1.6) | 0.13 | 1.45 | |||
PC | 275-T590a | 6.4(1.4) | 0.37 | 0.29 | |||
0 X | 93(3) | 1.01 | 1.8 | 0.80(0.27) | 21.49 | 2.81 | |
0.42 X | 150 | 48(1) | 5.39 | 39 | 17.4(0.2) | 9.44 | 25.74 |
0.56 X | 180 | 122.0(0.2) | 1.39 | 25.55 | 24.4(0.7) | 15.96 | 51.69 |
0.58 X | 215 | 89(3) | 7.88 | 38.75 | 15.8(0.4) | 24.78 | 57.59 |
0.7 X | 250 | 321(1) | 0.75 | 36.22 | 16.8(0.3) | 26.49 | 65.508 |
0.81 X | 275 | 284.0(0.1) | 1.36 | 62.47 | 37.4 s | 23.8 | 259 |
1 X | 300 | 320.0(0.2) | 1.51 | 79.25 | 29.4(0.7) | 53.59 | 197.4 |
0.42 X | 180-T590a | 216(1) | 8.09 | 137.38 | 45.6(0.6) | 22.42 | 156.98 |
1 | 300-T590a | 159(0.26) | 3.99 | 88.85 | 16.7(0.6) | 88.71 | 209.53 |
By contrast, the crystals show a marked increase in tunability of transport properties. Fig. 7 shows that charge carrier concentrations can be varied by about an order of magnitude with the highest concentrations being obtained for the highest TIE which results in the highest silver content (x ≈ 1). At low ion-exchange temperatures, crystals showed a bulk concentration close to that of the polycrystalline samples (which are also both silver and oxygen deficient at the surface). However, when increasing the temperature of ion-exchange, the bulk concentration of carriers increased gradually up to 3.2 × 1020 cm−3. We observed the same trend in the conductivity which increased from 25 to 79 S cm−1 as TIE increased. There is no strong trend in the mobility (μ) of the crystals where it seems to be generally constant at room temperature, and increases slightly with TIE at 77 K. Importantly, this indicates that band transport and not hopping transport is responsible for the carrier motion.40 The trends in bulk concentration are similar at RT and 77 K, though the values are lower at 77 K (ranging between 1.74 × 1019 to 2.94 × 1019 cm−3). The conductivity of the crystals kept the same trend with an overall increase going from 25.74 S cm−1 to 197.4 S cm−1 at 77 K. In addition, to understand further the ion-exchange process, we measured the electronic properties of a sample that was synthesized without silver nitrate. Table S1 (ESI†) shows that during this process 92.2% of the sodium left the crystal such that this material is nearly Na free and of course has no silver. As one would expect, this sample showed the lowest conductivity with 1.80 S cm−1 at room temperature and 2.81 S cm−1 at 77 K, and a relatively low bulk concentration compared to the other silver-containing crystal samples with only 9.3 × 1019 cm−3 at room temperature and 8 × 1017 cm−3 at 77 K. This shows consistency with the rest of the samples (Fig. 7) where lower silver content in the crystal yields lower conductivities. This, however, is counterintuitive given that the DFT calculations show that Ag vacancies should yield holes based on the band structures (see Fig. 6c). The fact that the opposite happens, as is also seen for the carrier concentration as a function of the TIE in Fig. 7, suggests that the silver vacancies may interfere with the oxygen content and prevent/limit the band reorganization seen in the oxygen deficient DFT calculation (see Fig. 6d). In this sense, we speculate that the two non-stoichiometries compete to a certain extent thereby setting a bound for how far the transport can be improved by varying Ag & O content. However, the fact that Hall measurements indicate holes are the dominant carrier indicates that Ag deficiency likely dominates the properties observed. From Fig. 6d we can see that strong oxygen deficiency would result in a metallic electronic structure reconstruction that would not be as amenable to generating a p-type Hall voltage (as for example, by contrast, the Ag deficiency is in Fig. 6c). This suggests that the band rearrangement predicted from DFT for the oxygen deficient material did not fully occur in our samples.
In addition, we performed TGA experiments in an oxygen atmosphere for both polycrystalline and crystal samples, shown in Fig. S6 (ESI†), in order to determine to what extent the oxygen content can be tuned in the materials after synthesis and the resulting impact on transport properties. Fig. S6 (ESI†) shows the results for two TGA runs: the first up to 190 °C shows a mass gain (uptake of oxygen) while the second up to 590 °C shows a mass decrease (release of oxygen). For both samples, it was confirmed that the XRD patterns did not vary (this was not the case for samples heated above 590 °C). Fig. 7 shows that the polycrystalline sample heated up to 190 °C gave an increase in the bulk concentration (light blue circle in Fig. 7) up to 8.1 × 1019 cm−3 consistent with an increase in holes by increasing the oxygen content (oxygen vacancies are n-dopants so oxygen excess acts as p-dopants). The sample had a negligible change in mobility (from 0.10 to 0.13 cm2 V−1 s−1) and an increase in conductivity going from 0.75 to 1.45 S cm−1 consistent with the charge carrier concentration increasing. On the other hand, the polycrystalline sample heated up to 590 °C (orange circle in Fig. 7) showed a decrease in the number of charge carriers down to 6.5 × 1018 cm−3 which we attribute to a decrease in oxygen content as seen in the TGA mass drop. The two TGA heated polycrystalline samples are therefore consistent with the transport properties being tunable by varying the oxygen content such that the hole concentration can either be reduced with less oxygen or increased with more oxygen.
Having established the temperatures at which oxygen is taken up and released on the powders, we repeated the experiments on two crystals (TIE = 180 and 300 °C). It should be noted that the crystals do not weigh enough to obtain measurable mass changes in the TGA. Fig. 7 shows that the sample with TIE = 300 °C gave the expected decrease in charge carrier density after heating to 590 °C, the bulk concentration decreased down to 1.6 × 1020 cm−3 with a conductivity staying relatively stable with a value of 209.53 S cm−1. These results are fully consistent with the results for the powders: decreased oxygen content yields a decrease in charge carrier densities. For the TIE = 180 °C sample, however, the bulk concentration increased slightly, going up to 2.16 × 1020 cm−3. This is the second counterintuitive transport trend seen in the samples with both Ag and O deficiency. It should however be mentioned that the increase in concentration is small and the values are within uncertainty of each other at RT. Nonetheless, the non-intuitive trends seen in the Ag/O deficient materials warrant further study, as do the impact of the competing cationic and anionic off-stoichiometries in the same complex oxide. This sample also showed an increase in conductivity from 25.55 S cm−1 for the as synthesized sample to 137.38 S cm−1 for a sample synthesized at the same temperature but that was then heated in the TGA. These results therefore show that the crystals’ carrier transport properties can indeed be further optimized by varying the oxygen content following post synthesis treatments. It is also important to re-emphasize that the cation oxidation states are also playing a critical role in tuning the transport properties. In particular, the reduction of Ag at the higher ion-exchange temperatures (as determined by XPS) results in a greater number of holes available for transport (accounting for some of the increase in the hole carrier concentration seen in Fig. 7). There is thus a delicate balance between cationic oxidation state and oxygen content that occurs in the Ag-deficient delafossites.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4tc02403f |
This journal is © The Royal Society of Chemistry 2024 |