Electronic devices would benefit from a low-cost amorphous, dopable, bipolar oxide semiconductor. However, p-type oxides are quite rare, largely due to self-compensation by native defects. Our simulations find that the amorphous phase of TeO2 is chemically ordered, forms shallow, uncompensated acceptor substitutional AsTe and NO centers, and uses materials that are processable at low temperatures.

There is presently great interest in industrially relevant p-type amorphous oxide semiconductors. So far, various p-type oxides have been found, but each has its limitations. The first p-type oxide, Cu2O, used a noble metal (Cu) with a s-shell lying near the O 2p level to raise the valence band maximum (VBM) energy above the O 2p level, making it a moderately shallow state.1 However, the bandgap of Cu2O was too small to be fully transparent. Cu2O was superseded by CuAlO2 where strong Al-O interactions widened the optical bandgap.2 However, the layered structure of CuAlO2 and similar compounds2,3 did not favor disorder. SnO is a p-type oxide with a low effective hole mass and 3 eV direct bandgap.4,5 However, its 0.7 eV indirect gap gives extremely high leakage current densities to be useful. The ZnRh2O4 spinel has a wide gap and is p-type and amorphous (a-).6 However, its cost and high processing temperature inhibit its use in back-end-of-line (BEOL) CMOS devices with, for example, n-type a-InGaZnO4 (IGZO),7 which use low (<400 C) processing temperatures. There were various data-based searches identifying, say, Bi2BaTa2O6 or SnTa2O6 as having low hole masses.8,9 However, experiments showed that these oxides had a relatively low p-type conductivity, due to compensation by intrinsic defects.10,11 Thus, there is still no fully satisfactory p-type oxide.

Bulk TeO2 is a glassy oxide already known for its large nonlinear optical coefficient.12–14 Guo et al.15,16 identified TeO2 as a possible p-type oxide from its calculated low effective hole mass and the high calculated mobility of its upper valence states. This is due to the isotropic nature of the now filled Te 5s states. Zavabeti et al.17 measured various experimental properties of crystalline β-TeO2, such as a hole mobility of 146 cm2/V.s, an effective hole mass of 0.51, and the weakly p-type behavior of the undoped material. Shi et al.18 produced β-TeO2 by pulsed laser and by sol-gel deposition, but again undoped. Extrinsically doped TeO2 by, say, As has not yet been produced or a high conductivity measured.

We show here that a-TeO2 would have a short range order similar to that of β-TeO2 with minimal like-atom bonds. We then show that its VBM state would lie close to the approximate p-type doping limit19–22 and that As-doped a-TeO2 doped would be a good uncompensated p-type semiconductor with its Fermi energy (EF) lying in the valence band. This occurs because TeO2 is mostly covalently bonded and maintains its bonding when doped to a p-type doping limit energy, beyond that typical of ionic oxides, so the conduction is not compensated by intrinsic defects.

Here, the electronic structures of the rutile, α-, β-, γ-, and a-TeO2 phases are calculated by density functional theory (DFT) using Heyd–Scuseria–Ernzerhof (HSE) hybrid functionals23 to give good bandgaps. The HSE method is also used to correct the band edge energies [electron affinity (EA) and ionization potentials (IP)] against the vacuum level in Table I with supercells having a 15 Å vacuum spacing between TeO2 slabs.

TABLE I.

Comparison of structural and electronic properties of idealized rutile, α-, β-, and amorphous TeO2, calculated by HSE.

Rutile TeO2 α-TeO2 β-TeO2 γ-TeO2 a-TeO2
Symmetry  P42/mnm  P4121 Pbca  P212121  ⋯ 
Crystal system  Rutile  Distorted rutile  Orthorhombic  Orthorhombic  Amorphous 
Eform eV/f.u.  −1.32  −1.54  −1.55  −1.49  −1.31 
Gap (eV)  0.72  2.65  3.29  ⋯  3.2 
IP (eV)  6.09  7.31  6.45  ⋯  6.2 
EA (eV)  5.37  4.66  3.16  ⋯  3.2 
Bandgap  Indirect  Indirect  Direct  Indirect  ⋯ 
Rutile TeO2 α-TeO2 β-TeO2 γ-TeO2 a-TeO2
Symmetry  P42/mnm  P4121 Pbca  P212121  ⋯ 
Crystal system  Rutile  Distorted rutile  Orthorhombic  Orthorhombic  Amorphous 
Eform eV/f.u.  −1.32  −1.54  −1.55  −1.49  −1.31 
Gap (eV)  0.72  2.65  3.29  ⋯  3.2 
IP (eV)  6.09  7.31  6.45  ⋯  6.2 
EA (eV)  5.37  4.66  3.16  ⋯  3.2 
Bandgap  Indirect  Indirect  Direct  Indirect  ⋯ 

TeO2 has a parent rutile phase with sixfold Te sites, Fig. 1(a), and three polytypes. The rutile structure has a relatively narrow gap.24 The α and β polytypes of TeO2 are distorted rutile orthorhombic phases with much wider gaps of ∼3.5 eV, with partly layered structures consisting of covalently bonded chains of fourfold coordinated Te sites and twofold coordinated O sites.14,15 The γ phase is less stable by 0.06 eV per formula unit (fu). The calculated properties of these polytypes are summarized in Table I and Figs. 1 and 2. We also give the total energies per TeO2 fu. for each polytype in Table I. These values are consistent with those of Deringer et al.25 

FIG. 1.

(a)–(d) Lattice structures of idealized rutile, α-, β-, and γ-TeO2, respectively (orange = Te, red = O).

FIG. 1.

(a)–(d) Lattice structures of idealized rutile, α-, β-, and γ-TeO2, respectively (orange = Te, red = O).

Close modal
FIG. 2.

(a) and (b) HSE band structures of idealized rutile and β polytypes of TeO2 and (c) partial density of states of β-TeO2.

FIG. 2.

(a) and (b) HSE band structures of idealized rutile and β polytypes of TeO2 and (c) partial density of states of β-TeO2.

Close modal
To study the doping properties and defect self-compensation in β- or a-TeO2, we calculate the defect formation energy ΔHq,20 
as a function of the ideal host total energy EH, the defect charge (q), the relevant chemical potential μ, and the Fermi energy (Ef) with respect to the valence band edge. Here, nα is the number of atoms of element α and μα is the chemical potential of element α.

The HSE band structures of two crystalline phases of TeO2 are shown in Fig. 2. They are similar to those of earlier work,26 except for the HSE correction. Figure 2(a) shows that the parent rutile phase is a narrow gap semiconductor with a bulk indirect bandgap of 0.72 eV. Due to this narrow gap, the calculated electron affinity is quite high, 6.09 eV.

β-TeO2 forms the stable layer-like structure shown in Fig. 1(c). Its band structure in Fig. 2(b) is calculated in a similar fashion. Its direct bandgap at Γ is 3.29 eV.15,17 It has a Te s-like valence band maximum, which gives a high theoretical hole mobility. The ionization potential or VBM energy is 6.45 eV below the vacuum level. The β phase is stabilized from the rutile structure by 0.2 eV/fu.

The valence band maximum seen in the calculated partial density of states (PDOSs) has Te s-like character in Fig. 2(c), as found earlier.15,16,18,26–28 The highest valence band is Te s-like. Below this lies the predominantly O 2p-like states centered on −2.5 eV below the VBM. This band then extends down to −7.5 eV, followed by the O 2s band. The conduction band is largely Te p-like. It has a direct bandgap.

It is important to confirm that the TeO2 structure can withstand disorder to be useful as an amorphous oxide. Experimentally, TeO2 is known to be glass-forming like SiO2 based on β-TeO2 units.12,13 On the other hand, many of the recently deposited metal oxides like, for example, ZnO have more ionic bonding to oppose homopolar bonding. A-TeO2 is more covalent, so some homopolar bonds are a strong possibility. Thus, we subject a 96-atom supercell of crystalline β-TeO2 to an ab initio molecular dynamics (AIMD) calculation, in which the supercell is heated to 2000 K for 4ps, then cooled to 300 K in 20 ps, and then relaxed at 300 K for 5 ns. The structure is shown in Fig. 3(a), and the calculated radial distribution function (RDF) and partial RDFs are shown in Fig. 3(b). The Te-O nearest neighbor bond distribution is centered on 1.96 Å, while the O-O second neighbor peak lies at around 2.8 Å. These results are quite similar to the earlier AIMD results of Pietrucci et al.26 and also reasonably similar to experimental results.29,30 We see there are very few direct O–O bonds at 1.8 Å or direct Te-Te bonds in the total RDF at 3.8 Å in this relatively small supercell. Thus, a-TeO2 is very strongly chemically ordered, despite an absence of ionic bonding; the main disorder arises within the weaker Te-O secondary bonds.24 TeO2 is the only polytype of Te and O on the solid phase diagram and has a narrow composition range,31 consistent with this. The existence of glassy a-TeO2 and that it can be made as a plasma-deposited (PLD) film18 suggest that it can be made by atomic layer deposition (ALD).32 The calculated PDOS of a-TeO2 is seen in Fig. 3(c), the s-like upper valence bands being slightly mixed with O 2p states.

FIG. 3.

(a) Structure of amorphous TeO2 by ab initio molecular dynamics, (b) pair distribution function of a-TeO2, and (c) calculated (HSE) PDOS of a-TeO2.

FIG. 3.

(a) Structure of amorphous TeO2 by ab initio molecular dynamics, (b) pair distribution function of a-TeO2, and (c) calculated (HSE) PDOS of a-TeO2.

Close modal

Zavabeti et al.17 have shown that undoped crystalline β-TeO2 has a high p-type mobility but have not yet shown this in doped films, perhaps because of safety concerns with As doping.

It is still necessary to show that doped p-type films with high conductivity can be made without self-compensation of carriers by native defects. For this, the absolute band edge energies of the TeO2 phases referred to the vacuum level are shown in Fig. 4(a) and should lie within the approximate doping limit guidelines. It shows that the band edges of β-TeO2 are much higher than those of α-TeO2 and are then more compatible with p-type doping. This arises because of the layer structure of β-TeO2 where the dipoles of the lone pairs act to raise the electrostatic potential within the bonding units.

FIG. 4.

(a) Calculated HSE band edge energies referenced to the vacuum level of different phases of TeO2, compared to the approximate doping limits.21 (b) Calculated HSE defect formation energies for the oxygen vacancy and for O-rich and O-poor β-TeO2. The formation energies in a-TeO2 are very close to those in β-TeO2.

FIG. 4.

(a) Calculated HSE band edge energies referenced to the vacuum level of different phases of TeO2, compared to the approximate doping limits.21 (b) Calculated HSE defect formation energies for the oxygen vacancy and for O-rich and O-poor β-TeO2. The formation energies in a-TeO2 are very close to those in β-TeO2.

Close modal

Self-compensation itself can be checked by calculating the formation energy of the main native defect, in this case the oxygen vacancy, as a function of EF within the bandgap, and as a function of the O chemical potential μO, between the limits of O-rich TeO2 and O-poor TeO2, as shown in Fig. 4(b). If the formation energy drops to near 0 eV for EF lying within the bandgap, then an O vacancy can form spontaneously, effective doping will be compensated,19–21 and the overall mobility will be low. Figure 4(b) shows that in general β-TeO2 will be a good host. However, self-compensation could occur in O-poor films for EF at the VBM.

As the average calculation based on “doping limits” is based on defects in mainly polar oxides whereas TeO2 is a much more covalent oxide, we will directly calculate the effect of substitutional p-type doping in amorphous TeO2. We, therefore, test compensation by substituting two AsTe sites for Te sites into the a-TeO2 lattice and disorder the network by additional AIMD calculation. Two atoms are used to avoid spin-polarization effects. Figure 5(a) shows the resulting network for this converged calculation. Figure 5(b) shows that EF is found to lie inside the top of the valence band and that the As acceptor state is shallow in this HSE calculation. This acceptor level is not seen to be deep as are many p-type dopants in oxides like ZnO or Ga2O3.33 We attribute this behavior to the direct covalent bonding of the As atoms lying at fourfold Te sites, and also to the fact that substituting for Te atoms rather than for O atoms does not lead to strong polaronic effects at the acceptor site as would occur on O sites in say ZnO. Interestingly, however, separately substituting two acceptor N atoms on O sites as NO sites would remain quite shallow, but these sites disproportionate to having different coordination, to form N2 - N3+ valence alternation pairs34 as in Fig. 5(c). Here the subscript means coordination and the superscript means charge.

FIG. 5.

(a) AIMD structure of a-TeO2 with two AsTe substitutional sites, (b) calculated HSE DOS of a-TeO2 with two AsTe showing Fermi level lying in the top of the valence band, (c) two substitutional two NO sites, (d) calculated HSE PDOS for a-TeO2 containing two NO sites.

FIG. 5.

(a) AIMD structure of a-TeO2 with two AsTe substitutional sites, (b) calculated HSE DOS of a-TeO2 with two AsTe showing Fermi level lying in the top of the valence band, (c) two substitutional two NO sites, (d) calculated HSE PDOS for a-TeO2 containing two NO sites.

Close modal

We have shown by calculation that amorphous TeO2 is a dopable p-type oxide and that it is not self-compensated by intrinsic defects. Further, we have found that the valence band edge energy lies within the “doping limits” guidelines. We have also noted that substitutional As or Sb atoms lying at Te sites will produce degenerate conduction with the Fermi energy lying below the valence band maximum. This compound should be producible by low-cost atomic layer or sputtering deposition.

The authors have no conflicts to disclose.

John Robertson: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal). Xuewei Zhang: Formal analysis (equal); Investigation (equal). Qingzhong Gui: Investigation (equal). Yuzheng Guo: Formal analysis (equal); Investigation (equal); Methodology (equal); Supervision (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
2.
H.
Kawazoe
,
M.
Yasukawa
,
H.
Hyodo
,
M.
Kurita
,
H.
Yanagi
, and
H.
Hosono
,
Nature
389
,
939
(
1997
).
3.
H.
Ohta
,
M.
Orita
,
M.
Hirano
,
I.
Yagi
,
K.
Udea
, and
H.
Hosono
,
J. Appl. Phys.
91
,
3074
(
2002
).
4.
Y.
Ogo
,
H.
Hiramatsu
,
K.
Nomura
,
H.
Yanagi
,
T.
Kamiya
,
M.
Hirano
, and
H.
Hosono
,
Appl. Phys. Lett.
93
,
032113
(
2008
).
5.
J. B.
Varley
,
A.
Schleife
,
A.
Janotti
, and
C. G.
Van der Walle
,
Appl. Phys. Lett.
103
,
082118
(
2013
).
6.
T.
Kamiya
,
S.
Narushima
,
H.
Mizoguchi
,
K.
Shimizu
,
K.
Ueda
,
H.
Ohta
,
M.
Hirano
, and
H.
Hosono
,
Adv. Funct. Mater.
15
,
968
(
2005
).
7.
K.
Nomura
,
H.
Ohta
,
A.
Takagi
,
T.
Kamiya
,
M.
Hirano
, and
H.
Hosono
,
Nature
432
,
488
(
2004
).
8.
G.
Hautier
,
A.
Miglio
,
G.
Ceder
,
G. M.
Rignanese
, and
X.
Gonze
,
Nat. Commun.
4
,
2292
(
2013
).
9.
Y.
Hu
,
X.
Yao
,
D. G.
Schlom
,
S.
Datta
, and
K. J.
Cho
,
Chem. Mater.
33
,
221
(
2021
).
10.
M.
Barone
,
M.
Foody
,
Y. Q.
Hu
,
J. X.
Sun
,
B.
Frye
,
S. S.
Perera
,
B.
Subedi
,
H.
Paik
,
J.
Hollin
,
M.
Jeong
,
K.
Lee
,
C. H.
Winter
,
N. J.
Podraza
,
K. J.
Cho
,
A.
Hock
, and
D. G.
Schlom
,
J. Phys. Chem. C
126
,
3764
(
2022
).
11.
Z.
Zhang
,
Y.
Guo
, and
J.
Robertson
,
Chem. Mater.
34
,
643
(
2022
).
12.
R. A. H.
El-Mallawany
,
Tellerurite Glasses Handbook
, 2nd ed. (
CRC Press
,
2016
).
13.
M.
BenYahia
,
E.
Orhan
,
A.
Beltran
,
O.
Masson
,
T.
Merle-Mejean
,
A.
Mirgorodski
, and
P.
Thomas
,
J. Phys. Chem. B
112
,
10777
(
2008
).
14.
Y.
Li
,
W.
Fan
,
H. G.
Sun
,
X. F.
Cheng
,
P.
Li
, and
X.
Zhao
,
J. Appl. Phys.
107
,
093506
(
2010
).
15.
S.
Guo
,
Z.
Zhu
,
X.
Hu
,
W.
Zhou
,
X.
Song
,
S.
Zhang
,
K.
Zhang
, and
H.
Zeng
,
Nanoscale
10
,
8397
(
2018
).
16.
S.
Guo
,
H.
Qu
,
W.
Zhou
,
S. A.
Yang
,
Y. S.
Ang
,
J.
Lu
,
H.
Zeng
, and
S.
Zhang
,
Phys. Rev. Appl.
17
,
064010
(
2022
).
17.
A.
Zavabeti
,
P.
Aukarasereenont
,
H.
Tuohey
,
N.
Syed
,
A.
Jannat
,
A.
Elbournem
,
K. A.
Messalea
,
B. Y.
Zhang
,
B. J.
Murdoch
,
J. G.
Partridge
,
M.
Wurdack
,
D. L.
Creedon
,
J.
vanEmbden
,
K.
Kalatar Zadeh
,
S. P.
Russo
,
C. F.
McConville
, and
T.
Daeneke
,
Nat. Electron.
4
,
277
(
2021
).
18.
J.
Shi
,
Z.
Sheng
,
L.
Zhu
,
X.
Xu
,
Y.
Gao
,
D.
Tang
, and
K. H. L.
Zhang
,
Appl. Phys. Lett.
122
,
101901
(
2023
).
19.
W.
Walukiewicz
,
Physica B Condens. Matter
302
,
123
(
2001
).
20.
A.
Zunger
,
Appl. Phys. Lett.
83
,
57
(
2003
).
21.
J.
Robertson
and
S. J.
Clark
,
Phys. Rev. B
83
,
075205
(
2011
).
22.
H.
Hosono
,
Jpn. J. Appl. Phys.
52
,
090001
(
2013
).
23.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
,
J. Chem. Phys.
118
,
8207
(
2003
).
25.
V. L.
Deringer
,
R. P.
Stoffel
, and
R.
Dronskowski
,
Cryst. Growth Des.
14
,
871
(
2014
).
26.
M.
Ceriotti
,
F.
Pietrucci
, and
M.
Bernasconi
,
Phys. Rev. B
73
,
104304
(
2006
).
27.
F.
Piertrucci
,
S.
Caravati
, and
M.
Bernasconi
,
Phys. Rev. B
78
,
064203
(
2008
).
28.
R.
Raghvender
,
A.
Bouzid
,
S.
Cadars
,
D.
Hamani
,
P.
Thomas
, and
O.
Masson
,
Phys. Rev. B
106
,
174201
(
2022
).
29.
E. R.
Barney
,
A. C.
Hannon
,
D.
Holland
,
N.
Umesaki
,
M.
Tatsumiago
,
R. G.
Orman
, and
S.
Feller
,
J. Phys. Chem. Lett.
4
,
2312
(
2013
).
30.
O. L. G.
Alderman
,
C. J.
Benmore
,
S.
Feller
,
E. I.
Kamitsos
,
E. D.
Simandiras
,
D. K.
Liakos
,
M.
Jesuit
,
M.
Boyd
,
M.
Packard
, and
R.
Weber
,
J. Phys. Chem. Lett.
11
,
427
(
2020
).
31.
V. P.
Itkin
and
C. B.
Alcock
,
J. Phase Equilib.
17
,
533
(
1996
).
32.
V.
Pore
,
T.
Hatanpaa
,
M.
Ritala
, and
M.
Leskela
,
J Am. Chen Soc.
131
,
3478
(
2009
).
33.
H.
Peelaers
,
J. L.
Lyons
,
J. B.
Varley
, and
C. G.
van de Walle
,
APL Mater.
7
,
022519
(
2019
).
34.
M.
Kastner
,
D.
Adler
, and
H.
Fritzsche
,
Phys. Rev. Lett.
37
,
1504
(
1976
).
Published open access through an agreement with University of Cambridge