Shuqun
Chen
a,
Giorgio
Carraro
b,
Davide
Barreca
c,
Andrei
Sapelkin
d,
Wenzhi
Chen
e,
Xuan
Huang
e,
Qijin
Cheng
e,
Fengyan
Zhang
e and
Russell
Binions
*a
aSchool of Engineering and Materials Science, Queen Mary University of London, London E1 4NS, UK. E-mail: r.binions@qmul.ac.uk
bDepartment of Chemistry and INSTM, Padova University, 35131 Padova, Italy
cCNR-IENI and INSTM, Department of Chemistry, Padova University, 35131 Padova, Italy
dDepartment of Physics and Astronomy, Queen Mary University of London, London E1 4NS, UK
eSchool of Energy Research, Xiamen University, Xiamen 361005, China
First published on 19th May 2015
High quality Ga-doped ZnO thin films for use as energy efficient glazing coatings were deposited onto glass substrates by low cost single source aerosol assisted chemical vapour deposition (AACVD) of zinc and gallium acetylacetonates (in methanol) at a temperature of 350 °C. The effect of Ga content ranging from 0.4 at% to 6.1 at% on the structural and functional properties of ZnO films was investigated. Highly c-axis oriented films are easily formed in the case of pure ZnO with hexagonal (002) surfaces observed. This texture is gradually weakened in 0.4 at% to 3.0 at% Ga doped samples, and the deposit morphology is transformed to granular particles, irregular platelets, agglomerated particles and wedge-like structures, respectively, which may result from retarded grain boundary growth and increasing exposed non-(002) surfaces. Further gallium addition to 4.3 at% suppresses the grain growth and deteriorates the system crystallinity, with a concomitant change to a (102) preferential orientation in the heavily 6.1 at% Ga doped sample. The ZnO:Ga coatings exhibit high carrier concentration (up to 4.22 × 1020 cm−3) and limited carrier mobility (<5 cm2 V−1 s−1), and the minimum resistivity value obtained is 1.16 × 10−2 Ω cm. Due to their large band gaps (3.14–3.42 eV) and favourable carrier numbers, high visible transmittance (83.4–85.3%) and infrared reflection (up to 48.9% at 2500 nm) are observed in these films, which is one of the best AACVD ZnO reported for low emissivity application and close to the optical requirements for commercial energy saving glazing.
Transparent conducting oxides (TCO) with a large enough band gap to transmit the visible spectrum of light and with a high charge carrier concentration to reflect infrared radiation can be used as low-E coatings.8,12 One typical example is Pilkington K-glass, where a thin fluorine-doped tin oxide (FTO) coating is deposited on a glass surface by an atmospheric pressure chemical vapour deposition (CVD) process.13 In recent years, wide-band-gap zinc oxide has been investigated as new energy efficient coating because it is cheap, biocompatible, chemically stable and easy to fabricate.14–16 The intrinsic ZnO, however, has a low electron concentration of 1018 to 1019 cm−3 (ref. 17) and doping becomes an indispensable approach to meet the low-E optical requirements. Among group-III elements (such as Al, Ga and In), common n-type dopants in ZnO, gallium is regarded as a better choice because its ionic and covalent radii (0.62, 1.26 Å) is closer to those of zinc (0.74, 1.31 Å) than to those of aluminium (0.5, 1.26 Å) or indium (0.81, 1.44 Å), so the lattice distortion under a high doping input can be minimized.18–20 In addition, Ga is relatively oxidation resistant, so the formation of non-conductive gallium oxide in ZnO can be suppressed.21,22
High quality ZnO:Ga films (carrier density superior to 1020 cm−3) can be produced by magnetron sputtering,18 pulsed laser deposition24 and spray pyrolysis25 on glass substrates, but they are hardly fabricated by conventional CVD processes probably due to the absence of appropriate Zn and/or Ga precursors. In recent years, aerosol assisted CVD (AACVD) has been increasingly utilized to fabricate TCO materials because it could provide a wider availability of precursors for high quality CVD products.26–28 The architecture of AACVD thin films can be easily tailored by tuning deposition conditions, i.e. the physical properties of the carrier solvent (boiling point, heat of combustion and viscosity), the gas flow rate and the substrate temperature.29–31 This is important because the morphology of a film could directly influence its optical performance and application.32 For instance, to improve the absorption efficiency of silicon thin film solar cells, a ZnO layer with rough pyramid-like surfaces can be introduced to scatter and trap light into the absorber material,33,34 whereas low-sized surface features make the coatings suitable for architectural glazing.14 Moreover, for TCO material deposition, the introduction of doping atoms could also alter film growth processes, resulting in different film structure and spatial organization. As a result, the morphology and bulk properties of TCO films can be modified by adding various type and amount of dopant material. A variety of dopant elements, including Al, Ga, In, F, Cu and Ag, in ZnO film deposition has been reported in earlier AACVD works.14,35–37 However, to our knowledge, there has been hardly any systematic study on the structural and functional properties of AACVD zinc oxide films as a function of doping concentration.
Based on the above observations, in this work, a detailed investigation of the influence of Ga content on the growth behaviour and functional properties of ZnO films prepared by AACVD process has been undertaken. The main goal of this work was to characterize the film composition, structure, morphology, electrical and optical properties as a function of doping content, discussing the interplay between the system structural parameters and opto-electronic performances.
Sample I.D. | Elemental composition [at%] | Film thickness [μm] | TC (hkl) | Lattice constant a [Å] | Lattice constant c [Å] | Surface roughness [nm] | ||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
Zn | O | Ga | (002) | (101) | (102) | (103) | (112) | |||||
ZnO | 46.9 | 53.1 | 0 | 0.76 | 3.51 | 0.03 | 0.30 | 1.05 | 0.12 | 3.2504 | 5.2082 | 13.5 |
ZnO:Ga(0.4) | 42.2 | 57.8 | Not detected | 0.41 | 3.01 | 0.09 | 0.58 | 1.03 | 0.28 | 3.2501 | 5.2101 | 8.2 |
ZnO:Ga(0.8) | 46.1 | 53.1 | 0.8 | 0.49 | 2.78 | 0.13 | 0.77 | 0.98 | 0.34 | 3.2518 | 5.2107 | 28.2 |
ZnO:Ga(2.3) | 46.2 | 51.5 | 2.3 | 0.72 | 2.66 | 0.14 | 0.91 | 0.90 | 0.39 | 3.2511 | 5.2102 | 36.8 |
ZnO:Ga(3.0) | 45.4 | 51.5 | 3.0 | 0.48 | 2.01 | 0.29 | 1.12 | 0.94 | 0.63 | 3.2504 | 5.2067 | 14.1 |
ZnO:Ga(4.3) | 43.9 | 51.8 | 4.3 | 0.51 | 1.54 | 0.70 | 1.36 | 0.79 | 0.62 | 3.2527 | 5.2063 | 7.1 |
ZnO:Ga(6.1) | 41.8 | 52.1 | 6.1 | 0.49 | 0.68 | 1.27 | 1.61 | 0.68 | 0.76 | 3.2530 | 5.1921 | 4.6 |
![]() | ||
Fig. 1 Wide-scan XP spectra of ZnO and ZnO:Ga films with various Ga contents. The insets show the Ga 2p3/2 and O 1s XP bands. |
Irrespective of the specific processing conditions, Zn 2p3/2 peak positions (average BE = 1021.3 eV) pointed out the presence of Zn(II) in ZnO environment. This indication could be confirmed by the analysis of the Zn LMM Auger signal and the calculation of the corresponding Auger parameters [αZn = BE(Zn 2p3/2) + KE(Zn LMM) = 2010.4 eV] further verify this indication.39,40 The insets of Fig. 1 display high-resolution Ga 2p3/2 and O 1s XPS photoelectron signals. The Ga 2p3/2 peak position and shape indicate the presence of Ga(III) in an oxide environment.41,42 In addition, the intensity of the Ga 2p3/2 signal is monotonically enhanced with an increase of gallium dopant content, indicating a progressive enhancement of Ga content in the obtained systems (compare Table 1). Regarding oxygen, the main peak is centred at 530.1 eV, in agreement with the position expected for O in ZnO lattice.43 The asymmetry of the O 1s signals on the high BE side at ≈531.8 eV suggests the co-presence of hydroxyl groups.39,40 Moreover, it is seen that an increase in the Ga content in the specimens produced a concomitant chemical shift to higher BEs of the main O 1s component. This phenomenon can be attributed to the fact that Ga atoms are bonded more strongly to oxygen, as the Ga–O covalent bond length is smaller than that of Zn–O.18,23 Due to the low gallium loading, the Ga 2p3/2 peak signal in the first doped sample was undetectable by XPS measurement and the Ga atomic content in other five ZnO:Ga films can be calculated at 0.8 at%, 2.3 at%, 3.0 at%, 4.3 at% and 6.1 at%, respectively. In spite of this, we can still estimate its content to be approximately 0.4 at% based on a linear relationship between the nominal Ga/Zn molar ratio in precursor solutions and the final gallium content in films, as illustrated in Fig. S2.† The details of the film elemental composition and the corresponding sample I.D. are listed in Table 1.
The lattice parameters, a and c, of hexagonal ZnO and ZnO:Ga films are listed in Table 1. Among five strong c-axis oriented samples, the c value in pure ZnO is lower than that of others except for the ZnO:Ga(3.0). This result might be apparently surprising because Ga3+ holds a smaller ionic radius with respect to Zn2+, whereby the substitution of Zn2+ with Ga3+ at lattice sites could decrease the lattice constant.23 The first possible reason could due to the presence of high oxygen vacancy density in ZnO causing lattice distortion and decreasing the interplanar spacing.47,48 The existence of oxygen vacancies is indeed demonstrated by the analysis of film electrical properties, where the ZnO obtains a carrier density of 0.2 × 1020 cm−3. Fig. 3 shows the photoluminescence spectra of pure ZnO and two doped samples ZnO:Ga(0.4) and ZnO:Ga(3.0). The spectra display two intense peaks near 540 nm and 610 nm. It is generally believed the green emission at 540 nm is due to transition in defects, in particular the oxygen vacancies.49–51 The orange emission at 610 nm is less commonly reported, and could be related to the presence of interstitial oxygen ions.52,53 The reduced peak intensity at 540 nm indicates the generation of oxygen vacancies in ZnO is suppressed with Ga addition since the oxygen atoms are boned more strongly.54,55 The other more likely reason is the existence of interstitial gallium atoms in ZnO lattice, which expands the lattice parameters considerably.56 Also this kind of defect is hard to avoid under non-vacuum deposition conditions. Therefore, larger c-axis lattice constants are observed in ZnO:Ga(0.4) to ZnO:Ga(2.3) compared with pure ZnO. A further increase of Ga content to 3.0 at% starts suppressing the c-axis oriented crystal growth, associated with a reduction of (002) peak intensity, and the lattice parameter decreases to a value of 5.2067 Å. For ZnO:Ga(4.3) and ZnO:Ga(6.1), the crystal lattice is heavily distorted and the (002) signal intensity is significantly lowered, leading to a minimum c value of 5.1921 Å. Also the general larger lattice parameter a in doped samples could due to the reduction of oxygen vacancies, existence of interstitial gallium atoms and the promotion of near a-axis oriented crystal growth. For example, the normal direction of (102) plane is 58.03° deviated from the c-axis,25 so the promoted (102) crystal growth could help increase the a parameter.
The interplay between Ga doping and morphology is shown by the SEM images in Fig. 4. It is seen that the pure ZnO is composed of uniform regular grains, with hexagonal faces parallel to the substrate, although not so obvious as those in ZnO nanorods.57 After adding 0.4 at% Ga, the hexagonal grains disappear and turn into granular particles, and then transform into irregularly shaped platelets in ZnO:Ga(0.8). Agglomerated particles exhibit in the sample with 2.3 at% Ga and the ZnO:Ga(3.0) film morphology is largely wedge-like in shape. Further Ga addition appears to suppress the grain growth and poorly connected particles are observed both in ZnO:Ga(4.3) and ZnO:Ga(6.1), a phenomenon which correlates with the reduced peak intensities in their XRD patterns. It is also worth mentioning that both pure ZnO and samples with a low doping level exhibit typical columnar grain structure, as seen from their cross-section SEM images in Fig. S3,† while the ZnO:Ga(4.3) and ZnO:Ga(6.1) are more likely thickened by overlapped particles without any evidence of macro-texture, indicating that coalescence processes are largely suppressed.
The film crystallization occurs through sequential nucleation, initial growth and coalescence processes.58 Texture can be formed during the first nucleation stage, driven by surface energy minimization, or developed in the subsequent growth phase because only grains with lower surface energy can survive during the coalescence process.59 In pure ZnO, the polar (002) planes have higher surface energy, so the fastest crystal growth rate is usually along the c-axis to reduce the (002) facet areas as well as the system energy.60 Moreover, only those [002]-oriented crystallites with their c-axis orientation normal or near normal to the underlying substrate could grow all the way upwards, all differently oriented crystals stop their growth at earlier stages, resulting in columnar grain features and strong (002) texturing.61 The surface morphology of crystalline films is also affected by the preferred growth direction but in many cases they are more related to the exposed crystal planes. For instance, c-axis oriented ZnO films could preserve a hexagonal surface feature with their (002) planes exposed or exhibit a pyramidal structure by exposing the (101) planes, whose normal direction is 62° deviated from that of the basal (002) planes.62 The introduction of extrinsic doping atoms could greatly influence the ZnO film growth as well as the resulting morphology. On the one hand, the dopant atoms could alter the surface energy of crystallographic planes. For example, Liu et al. reported the growth of the a-axis-oriented (100) plane is more active than the growth of the c-axis-oriented (002) planes in ZnO:F films due to F− anions filling O vacancies or substituting O sites.63 This phenomenon would promote the growth of wedge-like grains parallel to the substrate rather than columnar ones.17 On the other hand, the dopant impurities are prone to segregating at the non-crystalline grain boundary areas, especially when the doping content suppresses its saturation point in ZnO, and drag the grain boundary movement, which reduce the grain size as well as the film crystallinity.64,65
For our pure ZnO sample, its pronounced columnar grains and strong (002) texture represent the microstructure evolution process have been fully developed. Also the observed hexagonal surface feature suggests the growth rate of ingrain and grain boundary areas are identical along the film thickening direction. For the coatings with a low doping level (0.4 at% to 3.0 at%), their highly c-axis oriented textures indicate (i) the (002) planes still hold much higher surface energy and growth rate than the others under the given gallium content, and (ii) the film coalescence processes are also greatly developed in these samples. In spite of this, the enhanced (102) texture coefficient suggests the incorporation of Ga atoms in ZnO lattice may increase the surface energy of (102) facets and populate their growth. A similar phenomenon has also been reported in Ga-doped ZnO nanowires, where the wire growth direction was changed from [001] direction in pure ZnO to [102] in ZnO:Ga samples.66 Thus, the c-axis textures are less significant in doped coatings and also weaken with increasing gallium content. Moreover, the grain boundary growth in ZnO:Ga samples would be retarded compared with the ingrain areas, resulting in the disappearance of hexagonal grains and the exposure of other low-index facets. These non-[002] oriented surfaces, which are formed at the final deposition stage, could successfully avoid to be incorporated into the columnar grain structure. It is likely that Ga will preferentially move to the polar (002) surfaces as a way of charge compensation in the crystal. This will also contribute to the retardation of growth in this direction. But in order to identify exactly which facets are exposed in ZnO:Ga(0.4) to ZnO:Ga(3.0), cross-section TEM investigation is required in our future work. When the Ga content exceeds 4.3 at%, the segregation of gallium atoms at grain boundaries become pronounced and suppress the grain growth greatly. So the [002]-oriented crystallites can no longer overgrow other crystallites with different orientations during the film thickening, leading to an obvious reduction in (002) peak intensity as well as the disappearance of a columnar texture structure. Also the obtained [102] preferential orientation in ZnO::Ga(6.1) should originate from a preferred nucleation in the early growth stage as the heavy doping inhibits all the crystallites growth significantly and refined grains are observed in this sample. The coarse grain boundaries, with amorphous-like contrast to the grain interiors, in its HRTEM image (Fig. 5) could verify the segregation of dopant atoms in ZnO:Ga(6.1). The doping dependent ZnO film growth behavior is schematically illustrated in Fig. S4.†
![]() | ||
Fig. 5 Bright field HRTEM images of the nanograins in ZnO:Ga(6.1). The inset shows the selected area electron diffraction pattern and the red arrows refer to grain boundary areas. |
Sample I.D. | Carrier concentration | Carrier mobility | Resistivity | Resistance | Doping efficiency | T λ550 | T λ380−780 | Plasma wavelength | R λ2500 | Band gap |
---|---|---|---|---|---|---|---|---|---|---|
[×1020 cm−3] | [cm2 (V s)−1] | [×10−2 Ω cm] | [Ω sq−1] | [%] | [%] | [%] | [nm] | [%] | [eV] | |
ZnO | 0.20 | 25.0 | 1.28 | 168.9 | — | 82.5 | 84.6 | — | — | 3.14 |
ZnO:Ga(0.4) | 0.80 | 4.8 | 1.62 | 395.6 | 63.9 | 83.9 | 85.3 | — | — | 3.19 |
ZnO:Ga(0.8) | 1.70 | 2.6 | 1.42 | 289.4 | 68.2 | 81.6 | 83.4 | 2440 | 17.5 | 3.27 |
ZnO:Ga(2.3) | 3.55 | 0.9 | 1.92 | 266.8 | 45.0 | 83.4 | 83.6 | 1825 | 41.1 | 3.39 |
ZnO:Ga(3.0) | 4.22 | 1.3 | 1.16 | 241.9 | 42.4 | 86.3 | 84.7 | 1670 | 48.9 | 3.40 |
ZnO:Ga(4.3) | 3.60 | 0.6 | 2.97 | 582.4 | 25.6 | 86.5 | 84.5 | 1920 | 35.0 | 3.42 |
ZnO:Ga(6.1) | 1.14 | 0.1 | 57.57 | 11![]() |
5.8 | 83.6 | 84.7 | 2380 | 18.6 | 3.37 |
The pure ZnO film has a carrier concentration of 0.2 × 1020 cm−3 and these charge carriers can be identified as oxygen vacancies since the films are deposited under an oxygen-deficient atmosphere. The addition of Ga atoms enhances ZnO carrier density to a range of 1020 cm−3. As the Ga content increased, the concentration of carrier shows a rapid enhancement from 0.80 × 1020 cm−3 in ZnO:Ga(0.4) to 4.22 × 1020 cm−3 in ZnO:Ga(3.0), and then gradually decreases to 1.14 × 1020 cm−3 at higher Ga loadings. These concentration values are quite high and comparable to the previously reported values in sputtered ZnO:Ga films (upto ∼5 × 1020 cm−3).23,67 The electron carriers in gallium-doped ZnO films are generated by substituting Zn2+ ions with Ga3+ ions, and this substitution efficiency is essential for the film electrical performance because those inactive doping atoms, such as interstitial gallium atoms, cannot generate free electrons but act as electron scattering centres.17 The doping efficiency (ηDE) can be defined as the ratio of the electron concentration to the Ga atomic concentration in ZnO:Ga films under an assumption that every incorporated Ga cation provides one free electron with substitution of a Zn ion:17,68
![]() | (1) |
A high carrier mobility of 25.0 cm2 (V s)−1 can be observed in pure ZnO as such it is easy to appreciate that scattering by the carriers is limited. After the Ga atoms are introduced, the mobility values decrease dramatically. Upon enhancing Ga content, the mobility in ZnO:Ga samples declines gradually from a maximum value of 4.8 cm2 (V s)−1 in ZnO:Ga(0.4) to a lowest value of 0.1 cm2 (V s)−1 in ZnO:Ga(6.1). It is known that the mobility of free carrier is determined by the electron scattering arising mainly from grain boundaries, ionized impurities and neutral impurities in doped ZnO films.25 The dominance of the scattering effects varies with the carrier density and the potential barrier at the grain boundary has been considered to be negligible when the carrier concentration is superior to 1020 cm−3.69 Moreover, the ionized impurity scattering cannot be the only dominant mechanism in our case because even more charge carriers are generated by ionized impurities in sputtered ZnO:Ga films, their mobility (10–30 cm2 V−1 s−1) could still one order higher than our results.23,67 Thus, the inferior mobility performance in present samples could mainly result from an insufficient doping efficiency, where many inactive dopant atoms locate in the ZnO lattice as interstitial defects or segregate at grain boundary areas as neutral impurities. By comparison, the higher mobility values in sputtered ZnO:Ga coatings should due to an improved incorporation efficiency. We attribute this to the high vacuum condition during their depositions, which typically leads to films of higher purity and a reduced number of defects.70 In Fig. 6, the charge carrier mobility is plotted versus the carrier density and doping efficiency for the studied coatings. It is seen that the mobility values depend linearly on the doping efficiency, and a similar trend is also observed with respect to carrier density, except for the most heavily doped sample ZnO:Ga(6.1). This indicates the dominant electron scattering mechanism is transformed from a combined ionized and neutral impurity scattering in ZnO:Ga(0.4) to ZnO:Ga(4.3) to the neutral impurity scattering only in ZnO:Ga(6.1).
![]() | ||
Fig. 6 The carrier mobility as a function of (a) carrier concentration and (b) doping efficiency for ZnO:Ga films. The dash circle marks the sample deviating from the linear trend. |
Through a combination of carrier concentration and mobility, a lowest resistivity value of 1.16 × 10−2 Ω cm was obtained in ZnO:Ga(3.0) film with an estimated carrier density 4.22 × 1020 cm−3 and mobility 1.27 cm2 (V s)−1. Due to the inferior mobility performance, our film resistivity is one order of magnitude higher than previously reported highly conductive ZnO:Ga film.23,67 In spite of this, these coatings could still have a potential application in energy efficient glazing. According to our observations, carrier density is the most important electric parameter rather than resistivity, though clearly they are related.
![]() | ||
Fig. 7 (a) Optical transmission spectra of ZnO and various ZnO:Ga films. (b) Optical reflection spectra of selected ZnO:Ga films. |
The calculated average transmittance in the visible light region (380 nm to 760 nm) of pure ZnO and ZnO:Ga films are 84.6%, 85.3%, 83.4%, 83.6%, 84.7%, 84.5% and 84.7%, respectively, a little lower than the values in sputtered ZnO:Ga samples (90–95%),23,67 thus a majority of visible light could transmit through the coatings. The minor light loss in TCO films is expected to mainly depend on the light scattering as a function of grain size, surface roughness and the level of defects.71,72 Among the studied coatings, a high visible transmission value of 84.7% can be observed in ZnO:Ga(6.1) even though a greater number of boundary areas is existed in this sample because its grain size is much smaller than that of others. This indicates the grain boundary light scattering is unlikely to be the main reason for visible light loss. Furthermore, the transmittance of pure ZnO and ZnO:Ga(3.0) are basically same despite their large difference in carrier density, representing the carrier scattering effect is also limited for present samples. Based on this, the obtained lower transmission values in ZnO:Ga(0.8) and ZnO:Ga(2.3) could be due to their rougher surface. This idea can be directly verified by two representative AFM images in Fig. 8, where the most transparent sample ZnO:Ga(0.4) exhibits a low surface roughness of 8.2 nm and this value in ZnO:Ga(2.3) could reach as high as 36.8 nm.
![]() | ||
Fig. 8 AFM images for samples (a) ZnO:Ga(0.4) and (b) ZnO:Ga(2.3). Roughness values of other samples are summarised in Table 1. |
The reduction of transmittance and increase of reflectivity in near infrared region in ZnO:Ga films are caused by a coherent oscillation of conduction electrons (plasmons) with incident electromagnetic radiation.73 The reflection onset occurs at the plasma wavelength (λp) which can be defined as follows:74
λp2 = c2m*ε/Nee2 | (2) |
The optical band gaps of pure ZnO and various ZnO:Ga films were determined by constructing Tauc plots using the (ahv)2 relation.77 The resulting plots are shown in Fig. 9(a) and the corresponding values are listed in Table 2. It is seen the band gap of pure ZnO film at 3.14 eV is lower than the reported value of bulk ZnO (3.24 eV) and the introduction of Ga atoms could enhance this band gap obviously with a maximum value of 3.42 eV obtained in sample ZnO:Ga(4.3). Such a band-gap widening phenomenon has been reported in many literatures for doped ZnO and can be explained through the Burstein–Moss effect. Accordingly, the excess free electrons with the addition of donor Ga3+ ions would fill the bottom levels of conduction band, thereby leading to an increase in the Fermi level.58,78,79 This band gap broadening (ΔEg) is related to the electron concentration Ne through the following equation:80
![]() | (3) |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ta02163d |
This journal is © The Royal Society of Chemistry 2015 |