Abstract
Materials with high electrical conductivity and optical transparency are needed for future flat panel display, solar energy, and other opto-electronic technologies. InxCd1-xO films having a simple cubic microstructure have been grown on amorphous glass substrates by a straightforward chemical vapor deposition process. The x = 0.05 film conductivity of 17,000 S/cm, carrier mobility of 70 cm2/Vs, and visible region optical transparency window considerably exceed the corresponding parameters for commercial indium-tin oxide. Ab initio electronic structure calculations reveal small conduction electron effective masses, a dramatic shift of the CdO band gap with doping, and a conduction band hybridization gap caused by extensive Cd 5s + In 5s mixing.
Substances in thin film form having high electrical conductivity and high optical transparency will be critical for next-generation flat-panel displays, photovoltaic cells, organic light-emitting diodes (LEDs), energy-efficient windows, and other opto-electronic technologies (1, 2). Although transparent conducting oxides (TCOs) such as Sn-doped In2O3 (ITO) and F-doped SnO2 are used extensively (3), their electrical and optical characteristics are inadequate for future applications (1), and fundamental scientific understanding of their properties has been hindered by their exceedingly complex microstructures. Recent advances in the TCO field include new ZnO conductors (refs. 4 and 5; ref. 6 and references therein), high carrier mobility Cd2SnO4 (7), new ternary and layered compounds (8, 9), p-type conductors (10), detailed analyses of phase relationships (ref. 11 and references therein), and detailed criteria for materials selection (12). Nevertheless, fundamental understanding of the constraints on maximum mobility and transparency imposed by crystal and electronic structure, film microstructure, and doping level is lacking. We report here a combined film growth, microstructure/charge transport/optical characterization, and first-principles electronic structure analysis of an InxCd1-xO (13, 14)‡‡ thin-film TCO family having simple crystal structures, exceptionally high electrical conductivity (≈2–5 × commercial ITO), and excellent optical transparency. The results reported here provide insight into criteria for optimizing TCO film properties.
For growth of metal oxide films, metal-organic chemical vapor deposition (MOCVD) offers the attraction of simple apparatus, conformal coverage, growth under high gaseous partial pressures, and adaptability to large-scale/large-area depositions (15). In the present case, it was necessary to develop the first easily handled Cd MOCVD precursor exhibiting high volatility and thermal stability, while introducing negligible carbon deposits. The synthesis of Cd(hfa)2(TMEDA) (hfa, hexafluoroacetylacetonate; TMEDA, tetramethylethylenediamine; Fig. 1A) is described in ref. 16 and references therein. In combination with In(dpm)3 (dpm, dipivaloylmethanate; Fig. 1B; ref. 5) in a low-pressure reactor (ref. 17 and references therein; 2 Torr, 360°C substrate temperature, Ar carrier gas, H2O-saturated O2 reactant gas), a series of InxCd1-xO films (18††), where x was systematically varied, was grown on smooth 1.25 cm × 0.50 cm glass substrates (float glass; growth rate ≈2.5 nm/min; film composition established by inductively coupled plasma spectrometry). Typical film thickness was 0.15 μm.
Figure 1.
Molecular structures of the metal-organic film growth precursors (A) Cd(hfa)2(TMEDA) is bis(hexafluoroacetylacetonato)(tetramethylethylenediamine)cadmium(II). (B) In(dpm)3 is tris(dipivaloylmethanato)indium(III). (C) θ-2θ x-ray diffraction scans of InxCd1-xO films grown on glass substrates by metal-organic chemical vapor deposition at 360°C. The unit cell dimensions contract linearly with In doping level until x ≈ 0.11.
X-ray diffraction reveals that for x ≤ 0.11, the as-grown InxCd1-xO films are phase-pure, crystalline, and highly [100] textured, with features assignable to a cubic CdO-type (NaCl) crystal structure (Fig. 1C). The lattice parameters contract monotonically with increasing x (indium content), with the reflections becoming broad for x ≥ 0.11, and with In2O3 features now evident. The nanoscale structure of the films in cross-sectional and plan-view was investigated by using cold field emission gun transmission electron microscopy. For the most conductive (vide infra) x = 0.05 film, the images clearly reveal pronounced [100] texture of the columnar, submicron (≈300 nm) grains, whereas electron diffraction confirms the cubic structure (Fig. 2). X-ray microchemical sampling evidences uniform presence of In throughout the film, with no evidence for significant grain boundary segregation, indicating a full solid solution of In in CdO. No other elements are detectable, with F− below the EDS (energy dispersive x-ray spectroscopy) and EELS (electron energy loss spectroscopy) detection limits of ≈1.2 wt %. Cross-sectional imaging indicates atomically abrupt substrate–film interfaces with no evidence for chemical reaction or solute segregation. Atomic force microscopy also indicates that the films are continuous and smooth with rms roughness of ≈10 nm over a 4 μm2 area.
Figure 2.
Nanoscopic images of the InxCd1-xO film microstructure for x = 0.05. (A) Plan-view transmission electron micrograph showing sub-micron columnar grains. (B) Electron diffraction pattern showing cubic crystal structure.
Charge transport measurements used instrumentation and techniques described elsewhere (19); reproducibility in conductivity measurements was ±1% on the same sample, ±5% from sample-to-sample. Data are compiled in Table 1. For x = 0.0, the four-probe electrical conductivity (3,000 S/cm; n-type) is approximately comparable to that of commercial ITO, whereas the charge carrier mobility of ≈150 cm2/Vs rivals or exceeds that of known TCO films, including that of CdO films grown by other deposition methods (ref. 20 and references therein). We speculate that the present rather “soft” growth technique provides higher O2 partial pressures as well as conditions for greater film microstructural orientation and crystallinity. As the In content is increased, the conductivity increases dramatically (Fig. 3A) as does n-type carrier concentration (Fig. 3B), whereas the mobility declines moderately (Fig. 3C). The highest conductivity is achieved at x = 0.05, where σ = 16,800 S/cm with an n-type carrier concentration (n) of 1.5 × 1021 cm−3 and a Hall mobility (μ) of 70 cm2/Vs. These parameters can be compared with those for films of the recently reported, high-mobility spinel TCO, Cd2SnO4, where σ = 8,300 S/cm, n = 9.0 × 1020 cm−3, and μ = 60 cm2/Vs, and to those of the high In-content spinel TCO, CdIn2O4, where σ = 4,300 S/cm, n = 6.1 × 1021 cm−3, and μ = 44 cm2/Vs (7). In the present case, increasing x beyond 0.05 results in declining conductivity and mobility, whereas the carrier concentration plateaus around x = 0.07 and then gradually declines (Fig. 3B). Variable-temperature data (Fig. 3D) reveal “metal-like” (∂σ/∂T < 0) behavior at low temperatures. Although carrier scattering mechanisms are not completely understood in polycrystalline TCO films (1), decreasing mobility with increasing carrier concentration combined with the modest conductivity increase with falling temperature argues for a predominant contribution from ionized impurity scattering, with a subordinate electron-phonon scattering role (21). Optical transmission spectra (Fig. 4A) reveal high transmittance and a widening visible transparency window with increasing In doping. Estimating the band gap from the absorption edge yields data (Fig. 4B) consistent with a classical Burstein-Moss (B-M) shift on conduction band filling (3), likely offset partially by gap shrinkage effects (22). From this analysis, the band gap of the x = 0.05 phase is ≈3.1 eV—rivaling or exceeding reported values for commercial ITO (3.0–3.7 eV).
Table 1.
Charge transport characteristics of InxCd1−xO films
Comp, x | Conductivity, S/cm | Carrier concentration, 1021 cm−3 | Mobility, cm2/V⋅s |
---|---|---|---|
0.00 | 3,560 | 0.152 | 146 |
0.02 | 9,370 | 0.640 | 91.5 |
0.03 | 15,600 | 1.26 | 76.9 |
0.04 | 12,300 | 1.27 | 60.5 |
0.05 | 16,800 | 1.51 | 69.2 |
0.09 | 9,010 | 1.65 | 34.0 |
0.11 | 6,740 | 1.06 | 39.6 |
All data at 25°C.
Figure 3.
Charge transport properties as a function of In doping level (x) for a series of InxCd1-xO films having 0.15 μm thickness. (A) Electrical conductivity at 300 K. (B) Charge carrier concentration at 300 K. (C) Charge carrier Hall mobility at 300 K. (D) Electrical conductivity as a function of temperature.
Figure 4.
Optical properties as a function of In doping content for a series of InxCd1-xO films. (A) Transmittance as a function of incident wavelength. (B) Derivation of the apparent optical band gaps.
Electronic band structure analysis used the precise FLAPW (full-potential linearized augmented plane wave) formalism (ref. 23 and references therein) with exchange-correlation energies treated via the local density approximation (LDA). Furthermore, the screened exchange LDA (sX-LDA) method was used for the first time in a TCO calculation to more accurately evaluate (24) band structure and optical properties. In addition to CdO, computations were performed for x = 0.031, 0.063, and 0.125 using 64-, 32-, and 16-atom supercells, respectively, and CdO atomic coordinates. The sX-LDA-derived CdO band structure (Fig. 5A) features a broad Cd 5s conduction band and O 2p-based valence band. The calculated direct gap at Γ of 2.36 eV (Table 2) compares well with experiment (2.28 eV; ref. 25). The carrier effective masses obtained from the conduction band curvature around Γ are 0.240 me, 0.247 me, and 0.259 me in the [100] [111], and [110] directions, respectively, also in excellent agreement with experiment (0.27 me; ref. 26). These rather small values contrast with 0.30–0.42 me calculated for In2O3 by LDA techniques (27) and suggest that CdO can function as a high mobility TCO host. Indium doping dramatically alters the CdO band structure (Fig. 5B) with a large Fermi energy displacement and B-M shift due to added In dopant electrons, which broadens the optical transparency window (similar effects are observed in LDA calculations on ITO; ref. 28). The present results also reveal extensive mixing of In 5s and Cd 5s states, yielding a hybridization gap (E) in the conduction bands, which further lowers optical absorption by weakening intraband transitions, and may reduce electron mobility at EF in heavily doped samples. Fig. 5C shows the wave vector k dependence of the conduction band energies as a function of doping. To obtain the optical band gap Eg (kF), which determines transparency in the visible, we calculated the optical transition matrices at kF and evaluated the edges of the interband transition energies. Next, Eg (k = 0) for x = 0.031 and 0.063 was evaluated by linearly interpolating the correction to the LDA band gap with respect to x, using sX-LDA data (1.33 eV for x = 0; 1.11 eV for x = 0.125). Table 2 summarizes results for Eg (k = 0) and Eg(kF) as a function of x. The minimum computed band gaps (in the [110] direction) are in excellent agreement with experiment (Fig. 5C), as is the band gap shrinkage of Eg at k = 0, which arises from many-body effects (22, 29) and counterbalances B-M shifting at high doping levels. Fig. 5C also shows a significant discrepancy from the rigid-band model neglecting any band gap shrinkage but fixing Eg (k = 0) and the effective mass at the experimental values.
Figure 5.
First-principles electronic structure calculations at the screened exchange-local density approximation level (sX-LDA) for InxCd1-xO. (A) Band structure of CdO with the direct band gap indicated. The origin of the energy is taken at the bottom of the conduction band. (B) Band structure of heavily doped InxCd1-xO, where x = 0.125. The origin of the energy is taken at the Fermi level. Dotted lines indicate states where mixing of Cd 5s with In 5s states is extensive; the direct gaps at k = 0 and k = kF and the hybridization gap are indicted. (C) Calculated band gap energies of InxCd1-xO as a function of kF (open symbols). The experimental band gaps using kF = (3π2n)1/3 (filled circles) and the rigid-band model, Eg(kF) = Eg(0) + (h2/2m*)kF2, where Eg(0) = 2.28 eV and m* = 0.27 me from experiment, are also shown for comparison.
Table 2.
Computed Fermi wave vectors, kF in atomic units, and band gaps, Eg in eV, for InxCd1−xO in the [100](Δ), [111](Λ), and [110](Σ) directions
x = 0.0 | x = 0.031 | x = 0.063 | x = 0.125 | |
---|---|---|---|---|
k![]() |
0.00 | 0.158 | 0.189 | 0.251 |
k![]() |
0.00 | 0.173 | 0.216 | 0.284 |
k![]() |
0.00 | 0.168 | 0.200 | 0.278 |
kF(n) | 0.00 | 0.174 | 0.220 | 0.277 |
Eg (k = 0) | 2.36 | 2.27 | 2.13 | 1.93 |
E![]() |
2.36 | 3.73 | 3.97 | 4.46 |
E![]() |
2.36 | 3.31 | 3.36 | 3.50 |
E![]() |
2.36 | 3.12 | 3.17 | 3.19 |
Fermi wave vectors in the free electron model kF(n) = (3π2n)1/3, where n = electron carrier density, are also listed.
For achieving high oxide conductivity and transparency, the present results are instructive, and in building on earlier qualitative models (30), point the way for design and discovery of new TCOs, both with and without Cd. The key to high InxCd1-xO carrier mobility is the broadly dispersed, free electron-like s-like conduction band, consistent with a large hopping integral between neighboring sites (28). This broad conduction band and the low carrier effective masses doubtless reflect the close metal–metal contacts possible in the cubic crystal structure where Cd-Cd = 3.320 Å in CdO (six-coordinate Cd+2 ionic radius = 1.09 Å), vs. closest In-In = 3.345 and 3.360 Å in In2O3 (six-coordinate In+3 ionic radius = 0.94 Å). Not only should an s-like conduction band promote uniform electronic charge density distribution, hence relatively low scattering, but the simple, isotropic cubic crystal structure should, a priori, afford fewer ionized and neutral scattering centers than more complex ITO-type structures (4, 21, 26, 31) having nonequivalent metal and O sites plus several types of vacancies and disorder. For enhancing optical transparency in the visible, the large doping-induced B-M shift more than compensates for gap shrinkage effects. In addition, doping splits the highly dispersed conduction band, and the resulting hybridization gap contributes to lower intraband optical absorption. Finally, to the extent that grain boundaries make some finite contribution to scattering (21), an isotropic cubic microstructure should help to minimize high-angle effects.
Acknowledgments
We thank Drs. T. Coutts and O. N. Mryasov for helpful comments. We thank the National Science Foundation MRSEC (Materials Research Science and Engineering Center) Program through the Northwestern Materials Research Center (Grant DMR–0076097) and the Department of Energy through the National Renewable Energy Laboratory (Contract AAD-9-18668-05) for support of this research.
Abbreviations
- LDA
local density approximation
- TCO
transparent conducting oxide
- ITO
Sn-doped In2O3
Footnotes
Whether Cd toxicity will limit large-scale implementation of all Cd-containing opto-electronic components is presently unresolved, although Cd-containing photovoltaic cells are moving to manufacture. See ref. 13 for discussions of relevant photovoltaic technology and ref. 14 for discussions of In and Sn toxicity.
Evidence for high conductivity in bulk CdO-In2O3 pellets sintered at 850°C.
References
- 1.Ginley, D. S. & Bright, C., eds. (2000) Mater. Res. Bull.25.
- 2.Kammler D R, Edwards D D, Ingram B J, Mason T O, Palmer G B, Ambrosini A, Poeppelmeier K R. Proc Electrochem Soc. 1999;99–11:68–77. [Google Scholar]
- 3.Hamberg I, Granqvist C G. J Appl Phys. 1986;60:R123–R159. [Google Scholar]
- 4.Wang R, King L L H, Sleight A W. J Mater Res. 1996;11:1659–1664. [Google Scholar]
- 5.Wang A, Dai J, Cheng J, Chudzik M P, Marks T J, Chang R P H, Kannewurf C R. Appl Phys Lett. 1998;73:327–329. [Google Scholar]
- 6.Minami T. Mater Res Bull. 2000;25:38–44. [Google Scholar]
- 7.Wu X, Coutts T J, Mulligan W P. J Vac Sci Technol A. 1997;15:1057–1062. [Google Scholar]
- 8.Edwards D D, Mason T O, Sinkler W, Marks L D, Goutenoire F, Poeppelmeier K R. J Solid State Chem. 1998;140:242–250. [Google Scholar]
- 9.Phillips J M, Cava R J, Thomas G A, Carter S A, Kwo J, Siegrist T, Krajewski J J, Marshall J H, Peck W F, Jr, Rapkine D H. Appl Phys Lett. 1995;67:2246–2248. [Google Scholar]
- 10.Kawazoe H, Yasukawa M, Hyodo H, Kurita M, Yanagi H, Hosono H. Nature (London) 1997;389:939–942. [Google Scholar]
- 11.Freeman A J, Poeppelmeier K R, Mason T O, Chang R P H, Marks T J. Mater Res Bull. 2000;25:45–51. [Google Scholar]
- 12.Gordon R G. Mater Res Bull. 2000;25:52–57. [Google Scholar]
- 13.Coutts T J, Young D L, Li X, Mulligan W P, Wu X. J Vac Sci Technol A. 2000;18:2646–2660. [Google Scholar]
- 14.Dunnick J K, Fowler B A. In: Handbook on Toxicity of Inorganic Compounds. Seiler H G, Sigel H, Siegel A, editors. New York: Dekker; 1988. pp. 155–174. [Google Scholar]
- 15.Schulz D L, Marks T J. In: CVD of Non-Metals. Rees W S, editor. New York: VCH; 1996. pp. 39–150. [Google Scholar]
- 16.Babcock J R, Wang A, Edleman N L, Benson D D, Metz A W, Metz M V, Marks T J. Mater Res Soc Symp Proc. 2000;623:317–328. [Google Scholar]
- 17.Belot J A, Wang A, McNeely R J, Liable-Sands L, Rheingold A L, Marks T J. Adv Mater Chem Vapor Dep. 1999;5:65–69. [Google Scholar]
- 18.Morozova L V, Komarov A V. Russ J Appl Chem. 1995;68:1240–1242. [Google Scholar]
- 19.Chung D-Y, Hogan T, Brazis P, Rocci-Lane M A, Kannewurf C R, Bastea M, Uher C, Kanatzidis M G. Science. 2000;287:1024–1027. doi: 10.1126/science.287.5455.1024. [DOI] [PubMed] [Google Scholar]
- 20.Murthy L C S, Rao K S R K. Bull Mater Sci. 1999;22:953–957. [Google Scholar]
- 21.Tahar R B H, Ban T, Ohya Y, Takahashi Y. J Appl Phys. 1998;83:2631–2645. [Google Scholar]
- 22.Dou Y, Fishlock T, Egdell R G, Law D S L, Beamson G. Phys Rev B Condens Matter. 1997;55:13381–13384. [Google Scholar]
- 23.Jansen H J F, Freeman A J. Phys Rev B Condens Matter. 1984;30:6177–6179. [Google Scholar]
- 24.Asahi R, Mannstadt W, Freeman A J. Phys Rev B Condens Matter. 1999;59:7486–7492. [Google Scholar]
- 25.Koffyberg F P. Phys Rev B Condens Matter. 1976;13:4470–4476. [Google Scholar]
- 26.Ueda N, Maeda H, Hosono H, Kawazoe H. J Appl Phys. 1998;84:6174–6177. [Google Scholar]
- 27.Odaka H, Iwata S, Taga N, Ohnishi S, Kaneta Y, Shigesato Y. Jpn J Appl Phys. 1997;36:5551–5554. [Google Scholar]
- 28.Mryasov, O. N. & Freeman, A. J. (2001) Phys. Rev. B Condens. Matter, in press. [DOI] [PubMed]
- 29.Charache G W, DePoy D M, Raynolds J E, Baldasaro P F, Miyano K E, Holden T, Pollak F H, Sharps P R, Timmons M L, Geller, et al. J Appl Phys. 1999;86:452–458. [Google Scholar]
- 30.Fan J C, Goodenough J B. J Appl Phys. 1977;48:3524–3531. [Google Scholar]
- 31.Nadaud N, Lequeux N, Nanot M, Jove J, Roisnel T. J Solid State Chem. 1998;135:140–148. [Google Scholar]