We report atomic layer deposition (ALD) of ZnO thin films on O-polar surface crystalline ZnO substrates at the relatively low temperatures of 120, 150, and 200 °C. The as-grown ZnO films are studied with aberration-corrected transmission electron microscopy and diffraction contrast, photoluminescence (PL), and surface photovoltage (SPV) spectroscopy. We find that the homoepitaxial films have a monocrystalline structure with the density of basal stacking faults comparable to that of the substrate (1011 cm−2) and that the stacking faults can induce high lattice strain due to their interaction with the inversion domain boundaries. The narrow excitonic PL linewidth (2 meV at 8 K) and the sharp SPV bandgap transition confirm the high quality of the ZnO films. Despite similarities in the film properties, the growth temperature has an effect on the density and spatial distribution of intrinsic defects. Our results demonstrate a considerable potential of ALD ZnO homoepitaxy for fabricating high-quality ZnO nanostructures and attaining viable p-type ZnO.

ZnO is a widely studied semiconductor with a unique set of physical properties and tremendous potential for various applications.1,2 Its wide bandgap (3.37 eV) and large exciton binding energy (60 meV), high thermal and chemical stability, inexpensive growth technology, and non-toxitity make ZnO a prospective material for UV light emitters,3 gas4 and biosensors,5,6 transparent electronics,7 spintronics,8 and solar cells.9 Nevertheless, many technological problems remain to be solved to realize device-grade ZnO. In particular, many of the device applications require both donor and acceptor doping of ZnO. Whereas ZnO can be easily doped n-type,10 stable and reproducible p-type doping has been difficult to attain.11 Since ZnO is known to dope itself n-type, it is not clear what dopant could compensate the background shallow donors and ensure adequate abundance of shallow acceptors for p-type ZnO.

In the last two decades, there has been a significant effort to grow ZnO thin films with low concentrations of point and extended defects to reduce the unintentional n-type doping (for a recent review, see Ref. 12). So far, ZnO films of the highest crystalline quality have been obtained by homo- and heteroepitaxy, using methods that require growth temperatures in excess of 550 °C (see, e.g., Refs. 13–16). Here, we use atomic layer deposition (ALD) of ZnO thin films on a single-crystalline ZnO substrate at relatively low temperatures (200°C) to further lower the defect density of homoepitaxial ZnO. For this purpose, ALD is a promising technique in many respects.17 It provides high uniformity of the grown layers, precision, repeatability, simple doping protocols, and low thermal budget. In this Letter, we focus on extended defects and their impact on structural and optoelectronic properties of the ALD-grown homoepitaxial films, which are studied by transmission electron microscopy (TEM), photoluminescence (PL), and surface photovoltage (SPV) spectroscopy.

The pressurized melt-grown, O-terminated (0001) ZnO substrates used for the deposition are supplied by Cermet. The manufacturer reports the x-ray θ2θ rocking curve FWHM of 49 arc sec.18 Before being loaded into the ALD chamber, the substrates were degreased in ethanol and rinsed with de-ionized (DI) water. A Cambridge NanoTech Savannah 200 system was used for the deposition. The ZnO films are grown using electronic-grade diethylzinc and DI water as precursors and semiconductor-grade nitrogen with a constant flow of 20 sccm as a carrier and a purge gas. The temperature of the liquid precursors was 35 °C. Each ALD cycle consisted of a water dose of 15 ms, a nitrogen purge of 5 s, a diethylzinc dose of 15 ms, and a nitrogen purge of 5 s. All the films are deposited with 500 ALD cycles. The substrate temperatures were 120, 150, and 200 °C. When choosing the deposition temperature, we took into consideration reports of ZnO ALD on sapphire,19 glass, and silicon,20 using the same precursors and suggesting an optimum growth window of 130–180 °C, though ZnO homoepitaxy by ALD at 300 °C was reported.21 

Structural properties of the ZnO homoepitaxial films are studied by TEM. The electron transparent samples were prepared using a focused ion beam (FIB) Zeiss Crossbeam 340 with 30–2 kV Ga+ to reduce the amorphized ion-damaged layer. TEM images and electron diffraction patterns were obtained using a JEOL 2010F microscope, while high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) was performed with an aberration-corrected JEOL ARM200F. Both microscopes were operated at 200 keV.

The film thicknesses are determined from the TEM cross sections with the results listed in Table I. We observe a nearly linear increase in the growth rate with temperature, rather than a constant rate expected in the case of the self-limiting reaction. The increase in the film growth rate above the ALD temperature window was previously attributed to formation of atomic layers via Zn–Zn bonding.19 The above assertion, however, is not supported by our TEM studies.

TABLE I.

Thickness and growth rate of the ZnO films.

120 °C150 °C200 °C
Thickness (nm) 75 90 115 
Growth rate (Å/cycle) 1.5 1.8 2.3 
120 °C150 °C200 °C
Thickness (nm) 75 90 115 
Growth rate (Å/cycle) 1.5 1.8 2.3 

The films are imaged in the dark-field mode using the combined (01¯10) and (0002¯) reflections, along the 211¯0 zone axis. The TEM cross sections of all three samples are shown in Fig. 1, where the shorter red arrows indicate the [0001] growth direction. The selected area electron diffraction (SAED) patterns are obtained from the electron beam focalization on the film/interface regions. The SAED patterns, which are shown in the insets, provide interplanar distances of 0.5125 nm and 0.1382 nm for the (0002) and (01¯10) planes, respectively. These values are in accordance with the c and a3/4 parameters for ZnO. In addition, a magnified image of the interface region of the 150 °C sample and the fast Fourier transforms for the film, the substrate, and the interface are shown in the supplementary material, Fig. S1. From the TEM images and SAED patterns, we, thus, observe a monocrystalline hexagonal wurtzite crystal structure in all three samples with no noticeable change in the crystal orientation across the film/substrate interface, attesting high-quality homoepitaxial growth over the [0001] direction. Moreover, in the TEM images, there is no noticeable diffraction contrast between the substrate and the films, indicating their identical chemical composition.

FIG. 1.

(a)–(c) TEM dark-field images of the ZnO homoepilaxial films grown at (a) 120, (b) 150, and (c) 200° C, obtained by using the combined (01¯10) and (0002¯) reflections, along the 211¯0 zone axis. The dashed green line in (a) encircles a crystallite inclusion. The longer yellow and shorter red arrows highlight basal and prismatic stacking faults (BSFs and PSFs), respectively. The insets show the SAED patterns of the films and substrates combined. The (01¯10) and (0002¯) reflections are circled.

FIG. 1.

(a)–(c) TEM dark-field images of the ZnO homoepilaxial films grown at (a) 120, (b) 150, and (c) 200° C, obtained by using the combined (01¯10) and (0002¯) reflections, along the 211¯0 zone axis. The dashed green line in (a) encircles a crystallite inclusion. The longer yellow and shorter red arrows highlight basal and prismatic stacking faults (BSFs and PSFs), respectively. The insets show the SAED patterns of the films and substrates combined. The (01¯10) and (0002¯) reflections are circled.

Close modal

The TEM imaging is particularly useful in identifying structural defects, which is important for optimizing the growth and understanding their impact on optoelectronic properties. Figure 1(a) shows the presence of ZnO crystallite (circled with a broken green line), misoriented with respect to the surrounding crystal. The crystallite inclusions have been observed only in the 120 °C ZnO film. The extended defects, such as basal and prismatic stacking faults (BSFs and PSFs), are also noticeable in Fig. 1. The BSFs, which are parallel to the film/substrate interface and highlighted with longer yellow arrows, can be seen both in the films and in the substrate. The PSFs (shown with shorter red arrows) originate at the film/substrate interface and extend to the film surface (see also the supplementary material, Fig. S1). Because the film samples differ in thickness, the amount of the defects in a single image can be misleading. In order to obtain the defect densities, we counted and averaged the stacking faults over several TEM images of the same cross-sectional area for each sample. The area densities of BSFs and PSFs in the substrate and the films are listed in Table II. These defect densities (standard uncertainty, 5%) are less than in post-annealed ZnO films on sapphire grown by ALD22 and are comparable to that in ZnO films on sapphire grown by molecular beam epitaxy.23 Moreover, the BSF densities in all ZnO films are close to each other and to that of the substrate.

TABLE II.

Defect densities of the homoepitaxial ZnO/ZnO samples. The standard uncertainty is 5%.

Substrate120 °C150 °C200 °C
BSFs (cm–22.0 ×1011 1.6×1011 2.7×1011 1.8×1011 
PSFs (cm–2⋯ 1.7×1010 6.1×1010 3.3×1010 
Inclusions ⋯  ⋯ ⋯ 
Substrate120 °C150 °C200 °C
BSFs (cm–22.0 ×1011 1.6×1011 2.7×1011 1.8×1011 
PSFs (cm–2⋯ 1.7×1010 6.1×1010 3.3×1010 
Inclusions ⋯  ⋯ ⋯ 

In polar crystals such as the group III nitrides (Al, In, Ga)N, the PSFs are often identified as inversion domain boundaries (IDBs) by multiple dark-field TEM imaging along a non-centrosymmetric axis.24,25 This method is based on the violation of Friedel's law in wurtzite crystals,26 resulting in a strong contrast and contrast reversal in two adjacent regions of the image acquired under multiple beam conditions with g=±(0002). In Fig. 2, the inversion domains in the 200 °C ZnO film are shown in red rectangles with the IDBs indicated by white arrows. Conversely, no contrast reversal is observed under the multiple beam conditions in the absence of inversion domains. The IDBs are known to have an effect on bulk piezoelectricity in ZnO.27 By altering the strain fields, they create a local piezoelectric polarization.28 As a result, an electric field parallel to the stress in one of the inversion domains is canceled by an anti-parallel field in the other, thereby reducing the bulk response.29 The formation of inversion domains in the homoepitaxial ZnO films may be nucleated by accidental growth on the substrate or by nucleation and growth within a columnar grain due to the interaction with a BSF.

FIG. 2.

TEM images of the 200 °C ZnO film acquired under multiple dark-field conditions with (a) g=(0002¯) and (b) g=(0002). The inversion domains are shown inside red rectangles. The white arrows indicate the inversion domain boundaries (IDBs).

FIG. 2.

TEM images of the 200 °C ZnO film acquired under multiple dark-field conditions with (a) g=(0002¯) and (b) g=(0002). The inversion domains are shown inside red rectangles. The white arrows indicate the inversion domain boundaries (IDBs).

Close modal

The BSFs are responsible for the striation along the film/substrate interface in Fig. 1 [see also the supplementary material, Fig. S2(a)]. In hexagonal wurtzite structures, the BSFs can be looked at as a local deviation to the cubic zinc blende crystal structure.30 Since they introduce electronic levels just above the valence-band maximum, they can contribute to optical transitions and become noticeable in PL spectra.31 In contrast to GaN epilayers,32 long extended BSFs are not observed in the ZnO films, which can be related to the homoepitaxial growth. Otherwise, the observed stacking faults are closely similar to those in GaN epitaxial films.30,32Figure 3(a) shows a HAADF-STEM image of the 200 °C ZnO film, along the 211¯0 zone axis, featuring a BSF of I2 type. Following Ref. 32, we highlight the two {101¯0} planes of the perfect wurtzite structure in yellow, separated by 15 lattice spacings, so the wrong stacking sequences (underlined by a longer yellow line) can be seen. The I2 BSF is terminated by Shockley partial dislocations shown by shorter yellow lines. Since the two partial dislocations have opposite Burgers vectors, b=±13101¯0, the Burgers circuit drawn around the fault does not exhibit any closure failure. As suggested in Ref. 32, the I2 fault can be formed by a shear rather than dissociation of a perfect dislocation into two Shockley partials. A different case where an I1 BSF arrives at an IDB is shown in Fig. 3(b). While the IDB continues toward the film surface, the I1 BSF is terminated by a Shockley partial dislocation. Figure 3(c) shows a strain map of the interaction between the BSF and IDB, obtained from geometrical phase analysis.33 Normal strain component εyy is obtained in the direction perpendicular to the scanning direction in STEM to avoid a “fly-back” error.34 In contrast to the case of nearly strain-free I2 BSF of Fig. 3(a), the interaction between the I1 BSF and IDB results in a high strain of ±45% at the Shockley partial dislocation as shown in the inset in Fig. 3(c). The BSFs causing the lattice strain are also observed in the 120 and 150 °C films (supplementary material, Figs. S2 and S3). According to Ref. 35, the in-plane lattice strain has an effect on excitonic luminescence of ZnO homoepitaxual films.

FIG. 3.

HAADF-STEM images of stacking faults in the 200 °C ZnO film, along the 211¯0 zone axis. (a) I2-type BSF; the longer yellow line underlines wrong stacking sequences in the perfect wurtzite structure, while the shorter yellow lines show Shockley partial dislocations with Burgers vector b=±13101¯0. The Burgers circuit (highlighted in yellow) drawn around the fault does not exhibit any closure failure. (b) I1-type BSF interacting with IDB. (c) εyy strain map of the image (b). The inset in (c) shows a high-strain region of Shockley partial dislocation.

FIG. 3.

HAADF-STEM images of stacking faults in the 200 °C ZnO film, along the 211¯0 zone axis. (a) I2-type BSF; the longer yellow line underlines wrong stacking sequences in the perfect wurtzite structure, while the shorter yellow lines show Shockley partial dislocations with Burgers vector b=±13101¯0. The Burgers circuit (highlighted in yellow) drawn around the fault does not exhibit any closure failure. (b) I1-type BSF interacting with IDB. (c) εyy strain map of the image (b). The inset in (c) shows a high-strain region of Shockley partial dislocation.

Close modal

Optoelectronic properties of the ZnO films are studied with PL and SPV spectroscopies. The PL measurements employed a CW Kimmon IK5452R-E HeCd laser at a wavelength of 325 nm. The samples were mounted inside an evacuated Janis CCS-150 cryostat operating within a controllable 8–325 K temperature range. The PL spectra were probed by a Spex 1401 monochromator with a spectral resolution of 0.18 cm−1 and an RCA C31034 photomultiplier tube detector. As to SPV spectroscopy, it was performed under ultra-high vacuum (109 Torr). The SPV response was excited by monochromatic light of variable frequency (the energy range, 1.13–3.65 eV), sourced from a 250 W QTH lamp, filtered with an Oriel Cornerstone monochromator, and probed with a Besocke Kelvin Probe S and Kelvin Control 07.

Figure 4 shows PL spectra of the ZnO films at room temperature and 8 K. It should be noted that the film thicknesses are of the same order as the absorption depth (70 nm) of the laser radiation used for excitation. Thus, the PL spectra can be attributed to both the grown films and the underlying substrate (see the supplementary material, Fig. S4). The deep (2.5 eV) defect signatures in the spectra have relatively weak PL intensities, indicating low concentrations of visible luminescent centers. By the same token, the near-band edge luminescence is several orders of magnitude stronger than the deep level emission [see both Figs. 4(a) and 4(b)]. Furthermore, the SPV spectrum obtained for the 150 °C sample indicates a low content of surface mid-gap traps, while the bandgap transition is sharp and well-defined (supplementary material, Fig. S5).

FIG. 4.

(a) and (b) Semi-logarithmic plots of the PL intensity in the 120 (black), 150 (red), and 200 °C (green) ZnO samples at (a) 8 K and (b) room temperature. The PL spectra are displaced by two orders of magnitudes in turn for clarity of presentation. (c) and (d) Linear plots of the PL spectra of (c) structural defect bound excitons (Y), two electron satellites (TES) and phonon replicas, and (d) neutral donor bound excitons I4, I6, and I8.

FIG. 4.

(a) and (b) Semi-logarithmic plots of the PL intensity in the 120 (black), 150 (red), and 200 °C (green) ZnO samples at (a) 8 K and (b) room temperature. The PL spectra are displaced by two orders of magnitudes in turn for clarity of presentation. (c) and (d) Linear plots of the PL spectra of (c) structural defect bound excitons (Y), two electron satellites (TES) and phonon replicas, and (d) neutral donor bound excitons I4, I6, and I8.

Close modal
FIG. 5.

(a) Temperature dependence of the peak positions of the BEx lines I4, I6, and I8 for three ZnO samples. The same red shifts of the excitonic lines are observed in all three samples, regardless of the film growth temperature. (b) Temperature dependence of intensity of the I8 BEx peak (magenta circles) and the fit to Eq. (1) (black curves). The experimental data and the corresponding fitting curves for the films grown at the indicated temperatures are displaced by an equal amount in turn for clarity of presentation. The activation energies Ea obtained from the fit are shown on the graph.

FIG. 5.

(a) Temperature dependence of the peak positions of the BEx lines I4, I6, and I8 for three ZnO samples. The same red shifts of the excitonic lines are observed in all three samples, regardless of the film growth temperature. (b) Temperature dependence of intensity of the I8 BEx peak (magenta circles) and the fit to Eq. (1) (black curves). The experimental data and the corresponding fitting curves for the films grown at the indicated temperatures are displaced by an equal amount in turn for clarity of presentation. The activation energies Ea obtained from the fit are shown on the graph.

Close modal

The low-temperature PL spectra of Fig. 4(a) exhibit several emission features near the band edge, which originate from the radiative recombination of bound excitons (BEx). These exciton recombination lines are plotted in Figs. 4(c) and 4(d) with complementary spectral ranges. Arguably, all three ZnO samples produce rather similar PL spectra. The dominant BEx peaks have the same peak positions and similar intensities, except for the 3.330 eV peak (Y, indicated by a vertical blue arrow), which has a significantly higher relative intensity in the 120 °C sample. The same peak is observed in the substrate as well [supplementary material, Fig. S4(a)]. We, thus, attribute the 3.330 eV peak to excitons bound to the extended structural defects observed with TEM. For comparison, in Ref. 22, a peak at 3.321 eV was attributed specifically to the BSFs in ZnO epitaxial films grown on c-plane sapphire by ALD; in Ref. 36, a 3.333 eV line was ascribed to excitons bound to structural defects in ZnO single crystals. Furthermore, Ref. 37 reported sharp emission lines around 3.333 eV, accompanied by very weak green luminescence, in Cermet substrates such as used in the present study. Since the green luminescence originates from mobile point defects such as Zn or O vacancies, it was argued that these intrinsic defects are collected by dislocations, which result in the reduction of the 2.45 eV band and the intensity increase in the luminescence lines of excitons bound to extended structural defects.37 In Fig. 4(a), the 2.5 eV band is indeed relatively weak at 8 K in all ZnO samples. Interestingly, at room temperature, this luminescent band is stronger and red shifted in the 120 °C sample compared to the other samples (supplementary material, Fig. S6).

As to other emission lines in Fig. 4(c), we attribute a 3.307 eV peak to two electron satellite (TES) transitions of the I8 BEx with two phonon replicas at 3.235 and 3.163 eV, and a 3.288 eV peak to a phonon replica of the I8 BEx. In Fig. 4(d), the dominant peaks are related to impurity BEx lines in the ground state. Following Ref. 36, we assign the I4, I6, and I8 lines to excitons bound to hydrogen, aluminum, and gallium neutral donors, respectively. Note that the structure of the bound to impurities' exciton lines in the grown films is different from that of the substrate [see the supplementary material, Fig. S4(a)], except the I8 peak, which seems to originate from the substrate. This is hardly surprising because hydrogen participates in the film growth reaction, and aluminum is an omnipresent contaminant in ZnO growth.

The low-temperature PL spectra in the range of neutral donor BEx are analyzed with a Lorentzian peak fitting procedure to obtain temperature dependence of the peak intensities, spectral positions, and full widths at half maximum (FWHM). The temperature dependences of the peak positions for all ZnO samples are shown in Fig. 5(a). The observed red shifts of the BEx emission lines I4, I6, and I8 with increasing temperature are the same in all three samples, regardless of the growth temperature. The temperature dependence of the peak intensity of the I8 BEx emission line in all ZnO samples is shown in Fig. 5(b). It can be described by38,39

I(T)=I01+Aexp(Ea/kT),
(1)

where I0 is the peak intensity at 0 K, A is a constant, Ea is the activation energy of the thermal process, k is the Boltzmann constant, and T is the temperature. In Fig. 5(b), the black solid lines are fits to the experimental data (magenta circles). The obtained activation energies Ea (shown in the graph) are similar to all ZnO samples and are in agreement with previously reported data.37 The impurity BEx lines are remarkably narrow in all three samples,with a FWHM of 2 meV at 8 K [Fig. 4(d)]. For comparison, FWHM as small as of 1 to 8 meV were reported in ZnO films of the highest crystalline quality.16,21,40

The PL and SPV measurements, thus, confirm the high quality of the grown ZnO films, which seems to be independent of the growth temperature. The presence of crystallite inclusions in the 120 °C film does not affect the impurity BEx linewidth. It is larger in the films than in the substrate though, due to the presence of PSFs and higher impurity concentrations. On the other hand, the crystallite inclusions can be linked to the intrinsic defects and their spatial distributions, as discussed above. The increased intensities of the 2.5 and 3.330 eV peaks in the 120 °C ZnO sample are the signatures on that link.

In conclusion, we have used ALD to grow ZnO thin films on single-crystalline ZnO substrates at the relatively low temperatures of 120, 150, and 200 °C. The structural and optoelectronic properties of the homoepitaxial films are studied with high-resolution TEM, PL, and SPV spectroscopy. The TEM has revealed very good crystallinity, perfect alignment of the films with the substrate, and similar defect densities in all the films and substrates. The BSFs and PSFs are identified as dominant defect types in the ZnO films. The lack of prismatic defects on the ZnO substrate highlights the differences between our ALD method and the high-pressure melt growth process. Although the 120 °C film has crystallite inclusions, the growth temperature was found to have slight effect on the overall quality of the films. Narrow BEx emission lines and sharp SPV bandgap transition confirm the high quality of the ZnO films. In addition, the PL spectra identify the presence of point defects and unintentional contaminants (Al, H, and Ga) in the grown films.

The high quality of the homoepitaxial ZnO films grown under less than optimum conditions can be explained by the self-limiting nature of the ALD growth process. In the ideal case, the film is grown one atomic monolayer at a time. This controlled slow growth gives the atoms in the new layer enough time to settle into a proper lattice position. With less controlled and faster growth processes, such as chemical vapor deposition and sputtering, the atoms have less time to reach their lowest energy positions, and thus, more thermal energy is required to grow a good crystalline film. We expect that the ALD method can be a good foundation for fabrication of high-quality ZnO nanostructures and production of ZnO pn homojunctions.

See the supplementary material for Figs. S1–S6 referred in this Letter.

The TEM work has been performed at the Kleberg Advanced Microscopy Center. This research has been supported by the Air Force Office of Sponsored Research under Award No. FA9550-19-1-0359) (A.A.C.) and the U.S. Department of Defense under Award No. W911NF-18-1-0439 (A.P.).

The data that support the findings of this study are available within the article and its supplementary material.

1.
Ü.
Özgür
,
Y. I.
Alivov
,
C.
Liu
,
A.
Teke
,
M. A.
Reshchikov
,
S.
Doğan
,
V.
Avrutin
,
S.-J.
Cho
, and
H.
Morkoç
, “
A comprehensive review of ZnO materials and devices
,”
J. Appl. Phys.
98
,
041301
(
2005
).
2.
C.
Klingshirn
, “
ZnO: From basics towards applications
,”
Phys. Status Solidi B
244
,
3027
3073
(
2007
).
3.
D. C.
Look
,
B.
Claflin
,
Y. I.
Alivov
, and
S. J.
Park
, “
The future of ZnO light emitters
,”
Phys. Status Solidi A
201
,
2203
2212
(
2004
).
4.
L.
Zhu
and
W.
Zeng
, “
Room-temperature gas sensing of ZnO-based gas sensor: A review
,”
Sens. Actuator, A
267
,
242
261
(
2017
).
5.
A.
Tereshchenko
,
M.
Bechelany
,
R.
Viter
,
V.
Khranovskyy
,
V.
Smyntyna
,
N.
Starodub
, and
R.
Yakimova
, “
Optical biosensors based on ZnO nanostructures: Advantages and perspectives. A review
,”
Sens. Actuators, B
229
,
664
677
(
2016
).
6.
Y. Q.
Fua
,
J. K.
Luob
,
X. Y.
Du
,
A. J.
Flewitt
,
Y.
Li
,
G. H.
Markx
,
A. J.
Walton
, and
W. I.
Milne
, “
Recent developments on ZnO films for acoustic wave based bio-sensing and microfluidic applications: A review
,”
Sens. Actuators, B
143
,
606
619
(
2010
).
7.
Y.
Liu
,
Y.
Li
, and
H.
Zeng
, “
ZnO-based transparent conductive thin films: Doping, performance, and processing
,”
J. Nanomater.
2013
,
196521
.
8.
S. J.
Pearton
,
D. P.
Norton
,
Y. W.
Heo
,
L. C.
Tien
,
M. P.
Ivill
,
Y.
Li
,
B. S.
Kang
,
F.
Ren
,
J.
Kelly
, and
A. F.
Hebard
, “
ZnO spintronics and nanowire devices
,”
J. Electron. Mater.
35
,
862
868
(
2006
).
9.
R.
Vittal
and
K.-C.
Ho
, “
Zinc oxide based dye-sensitized solar cells: A review
,”
Renewable Sustainable Energy Rev.
70
,
920
935
(
2017
).
10.
T.
Minami
,
H.
Sato
,
H.
Nanto
, and
S.
Takata
, “
Group III impurity doped zinc oxide thin films prepared by RF magnetron sputtering
,”
Jpn. J. Appl. Phys., Part 2
24
,
L781
(
1985
).
11.
D. C.
Look
and
B.
Claflin
, “
P-type doping and devices based on ZnO
,”
Phys. Status Solidi B
241
,
624
(
2004
).
12.
M. A.
Borysiewicz
, “
ZnO as a functional material, A review
,”
Crystals
9
,
505
(
2019
).
13.
K.
Ogata
,
T.
Kawanishi
,
K.
Maejima
,
K.
Sakurai
,
S.
Fujita
, and
S.
Fujita
, “
Improvements of ZnO qualities grown by metal-organic vapor phase epitaxy using a molecular beam epitaxy grown ZnO layer as a substrate
,”
Jpn. J. Appl. Phys., Part 1
40
,
L657
L659
(
2001
).
14.
M. W.
Cho
,
C.
Harada
,
H.
Suzuki
,
T.
Minegishi
,
T.
Yao
,
H.
Ko
,
K.
Maeda
, and
I.
Nikura
, “
Issues in ZnO homoepitaxy
,”
Superlattices Microstruct.
38
,
349
363
(
2005
).
15.
I.
Bisotto
,
C.
Granier
,
S.
Brochen
,
A.
Ribeaud
,
P.
Ferret
,
G.
Chicot
,
J.
Rothman
,
J.
Pernot
, and
G.
Feuillet
, “
Residual doping in homoepitaxial zinc oxide layers grown by metal organic vapor phase epitaxy
,”
Appl. Phys. Express
3
,
095802
(
2010
).
16.
I. C.
Robin
,
P.
Marotel
,
A. H.
El-Shaer
,
V.
Petukhov
,
A.
Bakin
,
A.
Waag
,
M.
Lafossas
,
J.
Garcia
,
M.
Rosina
,
A.
Ribeaud
,
S.
Brochen
,
P.
Ferret
, and
G.
Feuillet
, “
Compared optical properties of ZnO heteroepitaxial, homoepitaxial 2D layers and nanowires
,”
J. Cryst. Growth
311
,
2172
2175
(
2009
).
17.
Z.
Gao
and
P.
Banerjee
, “
Review Article: Atomic layer deposition of doped ZnO films
,”
J. Vac. Sci. Technol. A
37
,
050802
(
2019
).
18.
J.
Nause
and
B.
Nemeth
, “
Pressurized melt growth of ZnO boules
,”
Semicond. Sci. Technol.
20
,
S45
S48
(
2005
).
19.
J.
Lim
and
C.
Lee
, “
Effects of substrate temperature on the microstructure and photoluminescence properties of ZnO thin films prepared by atomic layer deposition
,”
Thin Solid Films
515
,
3335
3338
(
2007
).
20.
E.
Guziewicz
,
I. A.
Kowalik
,
M.
Godlewski
,
K.
Kopalko
,
V.
Osinniy
,
A.
Wójcik
,
S.
Yatsunenko
,
E.
Łusakowska
,
W.
Paszkowicz
, and
M.
Guziewicz
, “
Extremely low temperature growth of ZnO by atomic layer deposition
,”
J. Appl. Phys.
103
,
033515
(
2008
).
21.
Ł.
Wachnicki
,
A.
Dużyńska
,
J. Z.
Domagala
,
B. S.
Witkowski
,
T. A.
Krajewski
,
E.
Przeździecka
,
M.
Guziewicz
,
A.
Wierzbicka
,
K.
Kopalko
,
S.
Figge
,
D.
Hommel
,
M.
Godlewski
, and
E.
Guziewicz
, “
Epitaxial ZnO films grown at low temperature for novel electronic application
,”
Acta Phys. Pol. B
120
,
A7
A10
(
2011
).
22.
S.
Yang
,
C. C.
Kuo
,
W.-R.
Liu
,
B. H.
Lin
,
H.-C.
Hsu
,
C.-H.
Hsu
, and
W. F.
Hsieh
, “
Photoluminescence associated with basal stacking faults in c-plane ZnO epitaxial film grown by atomic layer deposition
,”
Appl. Phys. Lett.
100
,
101907
(
2012
).
23.
J. W.
Lee
,
S. K.
Han
,
S.-K.
Hong
,
J. Y.
Lee
, and
T.
Yao
, “
Characterization of microstructure and defects in epitaxial ZnO (112¯0) films on Al2O3(11¯02) substrates by transmission electron microscopy
,”
J. Cryst. Growth
310
,
4102
4109
(
2008
).
24.
A.
Berger
, “
Inversion domains in aluminum nitride
,”
J. Am. Ceram. Soc.
74
,
1148
(
1991
).
25.
L. T.
Romano
and
J. E.
Northrup
, “
Inversion domains in GaN grown on sapphire
,”
Appl. Phys. Lett.
69
,
2394
(
1996
).
26.
R.
Serneels
,
M.
Snykers
,
P.
Delavignette
,
R.
Gevers
, and
S.
Amelinckx
, “
Friedel's law in electron diffraction as applied to the study of domain structures in non-centrosymmetrical crystals
,”
Phys. Status Solidi B
58
,
277
(
1973
).
27.
J.
Zúñiga-Pérez
,
V.
Muñoz-Sanjosé
,
E.
Palacios-Lidón
, and
J.
Colchero
, “
Polarity effects on ZnO films grown along the nonpolar [112¯0] direction
,”
Phys. Rev. Lett.
95
,
226105
(
2005
).
28.
C.
Shi
,
P. M.
Asbeck
, and
E. T.
Yu
, “
Piezoelectric polarization associated with dislocations in wurtzite GaN
,”
Appl. Phys. Lett.
74
,
573
(
1999
).
29.
H. D.
Espinosa
,
R. A.
Bernal
, and
M.
Minary-Jolandan
, “
A review of mechanical and electromechanical properties of piezoelectric nanowires
,”
Adv. Mater.
24
,
4656
4675
(
2012
).
30.
P.
Ruterana
,
A. M.
Sanchez
, and
G.
Nouet
, “
Extended defects in wurtzite GaN layers: Atomic structure, formation, and interaction mechanisms
,” in
Nitride Semiconductors: Handbook on Materials and Devices
(
Wiley
,
2003
), Chap. 8, pp.
379
438
.
31.
J.
Lähnemann
,
U.
Jahn
,
O.
Brandt
,
T.
Flissikowski
,
P.
Dogan
, and
H. T.
Grahn
, “
Luminescence associated with stacking faults in GaN
,”
J. Phys. D: Appl. Phys.
47
,
423001
(
2014
).
32.
V.
Potin
,
P.
Ruterana
, and
G.
Nouet
, “
HREM study of stacking faults in GaN layers grown over sapphire substrate
,”
J. Phys.: Condens. Matter
12
,
10301
10306
(
2000
).
33.
M. J.
Hýtch
,
E.
Snoeck
, and
R.
Kilaas
, “
Quantitative measurement of displacement and strain fields from HREM micrographs
,”
Ultramicroscopy
74
,
131
(
1998
).
34.
A. M.
Sanchez
,
P.
Galindo
,
S.
Kret
,
M.
Falke
,
R.
Beanland
, and
P. J.
Goodhew
, “
An approach to the systematic distortion correction in aberration-corrected HAADF images
,”
J. Microsc.
221
(
1
),
1
(
2006
).
35.
H.
Matsui
and
H.
Tabata
, “
Lattice strains and polarized luminescence in homoepitaxial growth of a-plane ZnO
,”
Appl. Phys. Lett.
101
,
231901
(
2012
).
36.
B. K.
Meyer
,
H.
Alves
,
D. M.
Hofmann
,
W.
Kriegseis
,
D.
Forster
,
F.
Bertram
,
J.
Christen
,
A.
Hoffmann
,
M.
Straßburg
,
M.
Dworzak
,
U.
Haboeck
, and
A. V.
Rodina
, “
Bound exciton and donor-acceptor pair recombinations in ZnO
,”
Phys. Status Solidi B
241
,
231
260
(
2004
).
37.
M. R.
Wagner
, “
Fundamental properties of excitons and phonons in ZnO: A spectroscopic study of the dynamics, polarity and effects of external fields
,” Ph.D. thesis (
Technische Universität
,
Berlin
,
2010
).
38.
P. O.
Holtz
,
B.
Monemar
, and
H. J.
Łożykowski
, “
Optical properties of Ag-related centers in bulk ZnSe
,”
Phys. Rev. B
32
,
986
996
(
1985
).
39.
N.
Kumar
,
R.
Kaur
, and
R. M.
Mehra
, “
Photoluminescence studies in sol–gel derived ZnO films
,”
J. Lumin.
126
,
784
788
(
2007
).
40.
Y.
Kashiwaba
,
K.
Haga
,
H.
Watanabe
,
B. P.
Zhang
,
Y.
Segawa
, and
K.
Wakatsuki
, “
Structures and photoluminescence properties of ZnO films epitaxially grown by atmospheric pressure MOCVD
,”
Phys. Status Solidi B
229
,
921
924
(
2002
).

Supplementary Material