You are on page 1of 15

Materials Research Express

PAPER

AZO nanocolumns grown by GLAD: adjustment of optical and structural


properties
To cite this article: L G Daza et al 2019 Mater. Res. Express 6 1050b9

View the article online for updates and enhancements.

This content was downloaded from IP address 130.238.7.40 on 25/10/2019 at 00:15


Mater. Res. Express 6 (2019) 1050b9 https://doi.org/10.1088/2053-1591/ab4055

PAPER

AZO nanocolumns grown by GLAD: adjustment of optical and


RECEIVED
22 February 2019
structural properties
REVISED
18 August 2019
ACCEPTED FOR PUBLICATION
L G Daza1 , R Castro-Rodríguez1 and A Iribarren2
2 September 2019 1
Department of Applied Physics, CINVESTAV-IPN, Unidad Mérida. 97310 Mérida, Yucatán, México
2
PUBLISHED Instituto de Ciencia y Tecnología de Materiales, Universidad de La Habana, Zapata y G, Vedado, La Habana 10400, Cuba
13 September 2019
E-mail: germandaza13@hotmail.com and luis.daza@cinvestav.mx

Keywords: zinc oxide, aluminum, nanorods, thin films, sputtering, optical properties

Abstract
ZnO doped with aluminum (AZO) was grown on a commercial glass substrate using the technique of
Glancing Angle Deposition (GLAD). The thin films of the aligned nanocolumns were obtained using
the rf-sputtering as a vapor source. To adapt the physical properties, the films are prepared by
controlling the parameters of the incident flux angle (α) and the substrate rotation (ω) as follows: (1)
different angles of incident deposition flux of α=0°, 40°, 60° and 80° without rotation of the
substrate, and (2) the angle of flux of 80° was fixed and the films were grown using a substrate rotation
of 0.6, 1.2 and 10 rpm respectively. The characterization of the physical properties was carried out,
including structural (scanning electron microscopy and x-ray diffractometry) and optics (UV-visible
spectrometry). The scanning electron microscopy images showed that by controlling both parameters,
the nanocolumn arrays can grow in the form of vertical and inclined structures aligned with uniform
thicknesses. Combining these parameters of growth, the refractive index can be adapted between 1.37
and 1.73 at a wavelength of 700 nm and the energy band gap in the range of 3.45 and 3.64 eV
respectively. The orientation of the nanocolumns grows normal to the c-crystallographic plane (002)
with a crystallite size between 26 nm and 36. Shadowing at 80° results in a remarkable increase of the
film’s porosity.

1. Introduction

Sculpted thin films with well-nano-designed morphology are actually a new class of materials with physical
properties that can be controllably adapted to be applicable in nanoscale technology [1–3]. This has
revolutionized the field of materials science in recent years by allowing the manipulation of the interactions of
light with matter. These advantages are very attractive for practical applications, such as transparent electrodes
[4, 5] solar cells [6] and other devices. The interest in the ZnO semiconductor for optoelectronics has increased
its application due to its high exciton binding energy of 60 meV and broadband energy of 3.37 eV at room
temperature [7]. ZnO nanostructures such as nanorods, nanotextured surfaces and nanocolumnar crystallinity
have aroused great interest in advanced photon management design devices and detection applications due to
their high surface/volume ratio [8–10]. Transparent electrodes are also an important component in organic and
hybrid solar cells because their properties such as electrical conductivity and optical transparency can be
combined with other functional requirements for these devices, such as large surface area, effective light
management, directions of preferential charge transport, mechanical flexibility and compatibility with
polymeric substrates [11, 12]. In particular, the nanostructures morphology in thin films of transparent
conductive oxide (TCO) combined with thickness, grain sizes, texture, agglomeration, surface geometry, and
porosity have a significant influence on the main characteristics of the new generation of solar cells [13, 14]. For
most photovoltaic devices, management light on devices through the TCO substrate and a well-aligned array of
nanostructures grown on glass substrates is of particular interest because they can improve the performance of
the device [15].

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 1. Experimental setup with GLAD method combined with sputtering rf. Side left show the schematic system utilized and the
side right the configuration of GLAD with angle deposition (α) and speed rotation (ω).

To develop sculpted nanostructure material with low cost, Glancing Angle Deposition (GLAD) [16, 17] is
one of the useful physical deposition methods because, unlike conventional evaporation methods, a substrate
without (OAD) or with rotation is tilted against the flux of vapor. The mechanism of growth in OAD is called the
shadow effect. This effect begins in the early stages of deposition when nuclei are randomly formed on the
surface of the substrate. As the deposit process continues, the growth of the cores brings the arrival of any other
evaporated material to areas that have no cores, and that produces shaded areas. Therefore, the higher cores will
receive more material from the steam flow compared to the smaller ones; this prevents the smaller islands from
growing larger and leading to the formation of different growth. finally, only the upper part of the longer
columns will continue the growth, so the nuclei and the smaller columns will stop growing causing a columnar
growth inclined towards the vapor flux with a columnar angle β, this process is called columnar extinction [16],
If a substrate speed rotation is added (GLAD), the shadow effect is reduced to having no nucleation zones with
shading and the resulting columns would not have columnar inclination angle β. By controlling the angle of
incidence of the vapor flux and the rotation of the substrate, it is possible to manufacture structures such as
nanorods, nanocolumns and zig-zags [17]. In photovoltaic devices [18–21], for example, to improve the
performance of solar cells, different three-dimensional architectures have been created from the fabrication of a
three-dimensional union n/p of vertically aligned arrays of semiconductors with well-defined nanostructured
forms, greater surface area and excellent charge transport property.
The adjustment of well-aligned nanocolumns arrays of AZO thin films combined with its optical and
structural properties may be of interest as transparent electrodes to improve the performance of solar cell
devices. In this work, we studied selective samples of AZO thin films with aligned nanocolumns structure
developed under different GLAD parameters in commercial glass substrates. We focus mainly on the effect of
nanocolumns morphology on its optical and structural characteristics, which shows that using variations of
these morphologies, it is possible to obtain and control significant changes in the physical properties of AZO
thin films.
Growths of thin films of AZO have been developed with similar techniques and characteristics made in this
work, with some variations and different results, for example with changes in the type of substrate have been
obtained nanostructured films of AZO that when used in cells improve the efficiency [22], also growth on
flexible substrate in electronic devices or flexible solar cells [23] and improvement in properties opto-electronics
changing deposition temperature [24].

2. Experimental details

AZO nanocolumnar thin films with an area of 25 mm×25 mm were deposited by GLAD using the rf-
sputtering technique (power of 80 W), figure 1(a), as a source of vapor flux reaching different oblique angles
from the normal line on the substrate with adequate conditions limited enough to create thin films with vertical
or inclined nanocolumns. The deposition chamber was first evacuated to a pressure of ~2×10−3 Pa. The AZO
films were prepared at the temperature of the substrate TS ∼120 °C using a target to substrate-distance of 60 mm.
We use an AZO commercial target with a mixture of ZnO (99.99% purity) and 2% weight of Al2O3 (99.99%

2
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

purity), argon gas with a total pressure of 1.333 Pa and a deposition time of 40 min. Before the deposition, the
substrates were ultrasonically cleaned in acetone, ethylethanol, isopropyl and deionized water for 10 min in
sequence, then dried by N2 flow and fixed in the support of the deposition system. As for a selective deviation
glancing angle and substrate rotation, the AZO nanostructures can be deposited on glass substrates, we prepare
AZO nanocolumns inclined samples of three different thin films without rotation of the substrate at oblique
angles of α=40°, 60° and 80° respectively (denoted S2, S3 and S4 respectively), and after another three samples
with vertical nanocolumnar morphologies grown by changing the rotation speed (ω) of the substrate with
ω=0.6, 1.2 and 10 rpm at an oblique angle selected from 80° (denoted S5, S6 and S7 respectively) as can be
shown in the figure 1(b). Sample S1 was a polycrystalline thin film of AZO deposited using conventional normal
incidence evaporation (α=0°, without substrate rotation). The morphological, optical and crystallographic
characterizations of the samples prepared were carried out using scanning electron microscopy (SEM JEOL-
7600F), UV–vis spectroscopy (the wavelengths of the light varied from 300 to 1200 nm with the angle of
incidence of the beam of light 90°) and x-ray diffraction (XRD) respectively. The XRD measurements were made
in symmetric geometry (θ-2θ), exploring the dispersed intensity in the range 2θ of 20°–80°, with a step size of
0.02°, and acquisition times of 10 s per step. The data were obtained by means of a D5000 Siemens x-ray
diffractometer with a filtered monochromatic radiation beam CuKα (λ=0.14518 nm) for 40 kV with 35 mA
and an aperture diaphragm of 0.2 mm.

3. Results and discussion

Top and cross views SEM images of synthesized films are shown in figures 2(a)–(g), respectively. Successfully
formed porous and very uniform packed arrays of AZO with nanocolumnar alignments are evident with
uniformity of length. It was found that the length and angle of inclination (β) of the inclined nanocolumnar
could be controlled well by changing the angles of incident deposition flow (α). Figure 2(a) shows images of
sample S1, which shows a thin film of polycrystalline AZO grain deposited at α=0° without rotation of the
substrate, the deposition rate was ∼11.2 nm min−1. In the cross-section film, a thickness of 450±10 nm with a
continuous and slightly rough surface morphology is observed. Inset in the lower part of figure 2(a) shows an
amplification of the image area of the cross-section of the film, which is shown in the red square. It is clear that
there is a homogenous surface porosity formation in the film of transverse morphology. Figures 2(b)–(d)
corresponds to samples S2, S3 and S4 prepared under different incident deposition flux angles of α=40°, 60°
and 80° without substrate rotation, respectively. The inclined nanocolumns shape formation occurs as the
incident deposition flux angles increase, and the vertical nanocolumnar shape occurs when the oblique angle of
the substrate reaches 80° and the thin films grow using the rotation of the substrate. Figures 2(e)–(g) corresponds
to samples S5, S6 and S7 with substrate rotation of ω=0.6, 1.2 and 10 rpm respectively. As can be seen, the
uniform nanocolumns arrays are vertically aligned on the substrate and the clearly defined dimensions of
diameter and height are observed, the type of finished conical surfaces of the vertical nanocolumns are also
observed in samples S5, S6 and S7. The formation of the cracks observed in all the transverse images of figure 2 is
due to mechanical breakage during the preparation of the sample.
In general, the incident flow angle (α) is greater than the columnar inclination angle (β) and there have been
many mathematical correlation attempts [25–28]; some deposit parameters change the value of β, such as, the
temperature of the substrate, the deposition rate, the angular distribution of the vapor flow and the pressure
within the vacuum chamber [29]; However, the most widely used heuristic expressions to correlate α and β are
the tangent rule proposed by Nieuwenhuizen and Haanstra [30], tan b = 12 tan a, which is valid for small
values of α (α„ 50 °) [26] Or the cosine rule proposed by Tait [31] b = a - arcsin 1 - 2cos a Using a purely
( )
ballistic model, which applies to high values of α. These expressions are related to the flow of steam that is
directed towards the substrate with an angle of incidence. The another way, a flow that arrives perpendicular to
the surface of the substrate leads to an almost columnar growth, resulting in a compact structure [25, 26].
The model does not consider any modification of the distribution function of the incident angle of the
deposition particles due to the rate deposition and the gas pressure into de chamber growth, for example,
therefore, the dependency of these parameters is not considered. Figure 3 shows the behavior of the angle of
inclination of the nanocolumns β40°, β60° and β80° of the samples S2, S3 and S4 as a function of α. However, by
substituting our values of α in the Tait model, the β values are not closer to our experimental value. Therefore,
the above-mentioned model is not suitable for reproducing the experimental evolution illustrated in figure 3
because it only considers shadowing effects. A model that allows the calculation β as a function of α during the
growth process and that supports these experimental findings and that explains the main processes responsible
for the formation of the nanostructure of these films according to the mechanisms of self-shadowing in the
surface, deposition rate, and gas pressure is necessary. In this work we only show an adjustment between β and
α. The nanocolumn tilt angle α increases following an exponential fitting that describes the relation β as a

3
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 2. Surface and cross-sectional FESEM images of AZO thin films growth to ω=0 rpm and α= 0° (a), 40° (b), 60° (c), and 80°
(d). Also with α= 80° and ω=0.6 rpm (e), 1.2 rpm (f) and 10 rpm (g).

Figure 3. Nanocolumn tilt angle of the AZO films as a function of the substrate inclination angle. The insets show the experimental
setup with GLAD method.

4
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Table 1. AZO thin films nanocolumns parameters of samples with different deposition angle and rotation.

Sample α (°) ω (rpm) β (°) Nanocolumn high (nm) Nanocolumndiameter (nm) q

S1 0 0 0 — — —
S2 40 0 12.6 — — —
S3 60 0 14.2 — — —
S4 80 0 18.6 — — —
S5 80 0.6 0 ∼346 ∼30.8 0.0035
S6 80 1.2 0 ∼294 ∼46.2 0.0074
S7 80 10 0 ∼242 ∼38.5 0.0091

Figure 4. Film thickness of the AZO films as a function of the substrate inclination angle.

function of α by the expression:


b = - 24.04e -(a 53.75) + 24.04 (1)
This result show a relation between β and α in all range of α measured, for our material and grown variables. For
the samples S5-S7 there is no present a nanocolumn tilt angle, the results are vertical arrangements, with changes
in nanocolumns diameters (table 1). The image inset of figure 3 shows a schematic representation of the GLAD
technique, showing the incident deposition flux angle. Figure 4 shows that the thickness (t) of the film of the
samples S1, S2, S3 and S4 decreases as the angles of incident deposition flux increase. The SEM images of the
upper surface show spherical shaped grains. The nanocolumns thickness decreases following an exponential fit
that relation t as a function of α:
t = - 20.27e (a 6.65) + 468.27 (2)
As shown in figure 4, a change of the substrate inclination angle from 0° to 80° has a significant impact on the
thickness of the films. The temperature of the substrate can also have an impact on the thickness, however, the
temperature of the substrate that is reached in each deposition is approximately the same, so it cannot be
responsible for the changes in thickness observed. The AZO films deposited at a glancing-angle suffer a severe
decrease in the growth rate, for example, the deposition rate of the sample S1 is approximately 11.2 nm min−1,
but the glancing-angle increase of 40° to 80° leads to a severe drop in the growth rate of the film as indicated by
equation (2) with the growth time that remains constant at 40 min, while all other processes parameters remain
the same. Likewise, at a variation of the distance between the target and the substrate, it can also induce a
significant change in the growth rate of the film, but this parameter was not modified. Therefore, the change in
the thickness of the sample S1 to S4 can be attributed to the reduction of the deposition rate due to the geometry
of the deposition as the angle of incidence of the species and the effect of self-shadowing. Then, in the inclined
substrate, the angle of the arrival flux is reduced and the total number of incident particles per unit area of the
substrate decreases, resulting in low deposition rates as shown in the equation (2). In other words, due to the
small number of incoming species per substrate site available for nucleation, the probability of adhesion will be
lower compared to that of the deposition of the normal incidence angle. Consequently, the species that adhere to
the substrate will cause early shading effects and, as a result, the nucleation islands will be more dispersed. On the
other hand, the rotations of the substrate also induce a decrease in the thickness of the films (table 1). Samples S5,

5
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 5. Transmittance spectrum of AZO thin films of samples S1-S4, the inset show the measured for the samples S5-S7 (a). Also
Refection spectrum of samples S1-S4, the inset show the measured for the samples S5-S7 (b).

S6 and S7 with a substrate rotation of ω=0.6, 1.2 and 10 rpm show a thickness of 346 nm, 294 nm and 242 nm
respectively which decreases the growth rate. This suggests that the spatial arrangement of the sputtering
particles in a substrate can also be controlled by the incident directions using the rotations of the substrate. In
this situation, the particles reaching the substrate are energetically favored by the substrate sites due to their
rotation, i.e., crystalline sites. The crystalline sites increase as an increase in the rotation speed of the substrate
occurs. Then, the deposition is close to a normal deposition of 0°, and the sputtered particles are distributed
homogeneously, favoring the vertical nanocolumns, because the sputtering particles and nucleation islands are
likely to develop at a speed comparable along the surface x-y or the z-axis.
Figure 5(a), shows the optical transmittance (T) in the wavelength range of 300–1100 nm (UV-visible region)
to samples S1, S2, S3 and S4, respectively. In a large part of the wavelength spectrum measured, the absorption is
very low, with almost null values so we take the expression [1-T(λ)] as the reflectance
R (l )(A  0: T + R » 1) and it shows in the figure 5(b). The figures inserted in both images correspond to
the spectrum of T(λ) and [1-T(λ)] of the samples S5, S6 and S7 respectively. These samples show significant
differences in their optical response in comparison with sample S1. The edge absorption of the sample S1 was
∼345 nm and the edge absorption of the others samples shows a small change towards the shorter wavelength.
The films show a transmittance that varies from ∼82% in the visible region, which show sufficient transparency
for use in photovoltaic applications. The transmittance for the sample S1 at the wavelength of 400 nm observed
in figure 5(a) is less than 60%. This can be attributed in part to the decrease in the volume of material because the
thickness of the film decreases as α increases (as shown in figure 4). A similar observation can be observed in the
inset of figure 5(a), where the edge absorption of samples S5, S6 and S7 shows a small change towards the shorter
wavelength as ω increases. Between 500 and 1100 nm, the 1-T of the thin films was ∼18% (figure 5(b)) and
fluctuated around 4%, due to the effects of surface interference. Transmittance in ultraviolet region does not
change significantly in figure 5(a). This because the columns diameter in the films deposited are less than 100 nm
and the Mie and Rayleigh scattering can be neglected, and as consequence the optical absorption in the films is
very low [32]. In a thin film the reflectance can be controlled by changing the layer thickness for obtaining
constructive or destructive interference in a specific wavelength. In our samples, between 500 and 1100 nm the
reflectance spectra [1-T(λ)] of the thin films was ~16% (figure 5(b)) and fluctuated around 4%, because of
surface interference. The sum of transmittance and reflectance is nearly 100% at long wavelengths, indicating
that the films possess nearly smooth surfaces. Minimum and maximum reflectance values in this spectrum are
found. The minimum reflectance value obtained at any wavelength depends on the optical thickness
(multiplication of the refractive index and thickness) of the film and is an integral multiple of that wavelength
following the expression for normal light incidence as: 2nt = (m - l ) l , where n is the refractive index of the
film, t is the film thickness, λ is the wavelength, m is an integer, and l is ½ or 0 for constructive or destructive
interference of reflected light. The wavelength value for the minimum reflectance of the samples S1 and S2 is
close to 610 nm, while the minimum reflectance of the samples S3 and S4 is close to 475 nm showing a blue shift
when the angle of inclination of the substrate increase. This blue displacement is due to the decrease of the
optical thickness of the AZO film with the increase of the inclination angle of the substrate. A similar behavior is
observed in the wavelength value for the minimum reflectance with the effect of rotations of the substrate of
samples S5, S6 and S7, at an inclination angle of the fixed substrate of 80°. The wavelength value for the
minimum reflectance of the sample S5 is close to 675 nm, while the minimum reflectance of the samples S6 and

6
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 6. SWT of the AZO thin films as a function of the substrate inclination angle. The inset show the SWR as a function of the
substrate inclination angle.

S7 are close to 520 nm and 600 nm respectively, also showing a blue shift change when increases the rotation of
the substrate.
In order to compare the optical properties between the samples, we calculate the solar weight of the
transmittance (SWT) and the 1-T (SW1-T), respectively, defined as:

SWT (l) =
ò S (l) T (l) dl (3a)
ò S (l) dl
SW 1 - T (l) =
ò S (l)(1 - T )(l) dl (3b)
ò S (l) dl
where S (λ) is the solar radiation spectrum AM 1.5, T(λ) and 1-T(λ) are the optical transmittances and the
reflectance, respectively. SWT and SWR can be estimated by the normalization of the transmittance and
reflectance spectra, and this was done by integrating the flux of solar spectral photons in the wavelength range of
300–900 nm. Figure 6 shows the SWT and the figure inset the SW1-T of samples as a function of α.
From figure 6 we can see that SWT increases monotonically from ∼79.6% reaching a maximum value
~82.4% for S4, after SWT takes values that fluctuate around ∼82.4% to samples S5, S6 and S7 respectively.
Similar characteristics were observed in the SW1-T, in the inset of figure 6 we can see that SW1-T decreases
monotonically from ∼20.4% reaching a minimum value ~17.4%, after SWR takes values that fluctuate around
∼17.4%. The band gap energy (EG) of the films was calculated by extrapolation of a linear region of the square of
the absorption coefficient as a function of the photon energy (as shown in figure 7), and compared to the
standard ZnO value of 3.37 eV we can see a difference between these EG that indicate the effects of the possible
contribution of quantum confinement due to the formations of some structures of small size dimensions in the
nanocolumns structures. In figure 7 we can see that the EG tends a high energy as indicated by the arrow line.
Figure 8 shows the EG as a function of α, we can see that EG monotonically increases the value of 3.53 eV to S4,
after that the EG takes values around 3.53 eV to samples S5, S6 and S7 respectively, the maximum value of EG
takes place at ω∼1.2 rpm with EG=3.64 eV. An increase in the value of EG is appreciated with respect to the
standard value 3.37 eV [33] possibly by effect Moss- Burstein, in which there is a negative correlation between
carrier density and edge position of IR absorption, and a positive correlation between the density of the carriers
and the edge of UV absorption, where the band gap energy increase with a higher density of carriers [34], another
possible factor is the change of residual stress and crystallinity [24].
The refractive indexes of the films deposited were calculated using equation [35]:
⎛ 1 + R (l ) ⎞ 4R (l)
n (l ) = ⎜ ⎟+ - k 2 (l ) (4)
⎝ 1 - R (l ) ⎠ (1 - R (l))2
where k (l ) is the extinction coefficient k (l ) = la (l ) /4p, a (l ) is the absorption coefficient, R (λ) is the
reflectance R (l ) = 1 - T (l ) - A (l ), where the absorbance A(λ) tends to zero in the visible region under the
energy value of band gap (λ > 400 nm), so that the reflectance calculation of light is reduced to

7
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 7. Tauc plot of AZO films growth at different incident deposition flux angles. The inset shows the Tauc plot for the samples
with different substrate rotation to α=80°.

Figure 8. The band gap behavior of AZO film as a function of the substrate inclination angle. The line only indicates the tendency.

R (l ) = 1 - T (l ). However, numerous methods have been used for the determination of the refractive index
and the extinction coefficient of thin films. One of their, is the combination of a normal incidence transmission
measurement and a method of measuring near-normal reflectance [ R (l ), T (l )]. This method requires at least
these direct measurements of R (l ) and T (l ). In our case, we use a single transmittance spectrum, since these
require the least experimental effort and allow a ‘virtual’ measure of R (l ) as a second variable from
R (l ) = 1 - T (l ). On the other hand, generally in thin films it is not possible appreciated a defined value of
refractive index due at the oscillations of the optical interference. To reduce this effect, we apply the adjustment
reported in the literature [36]. Figure 9 plots the refractive index of the films as a function of the wavelength and
show that the refractive indexes tend a low values as increasing the flux angle α, but increase as increasing the
substrate rotation. These behaviors can be observed in figure 10. At values of the refractive indexes at λ=700
nm as indicate by the arrow line in figure 9, the refractive index increases monotonically with increasing α,
reaching a minimum value for S4 simple. After that, when the rotation of the substrate is applied, the refractive
index increases as it increases ω as showed by samples S5, S6 and S7 respectively. This is caused by the kind and
formation of the nanocolumns array.
This controllable refractive index allows of graded refractive index in the thin films, indicates the possible
application of these thin films also as antireflective coatings.
The porosity (P) of the nanocolumnar samples is defined as the ratio between the volume of the gaps
between the nanocolumns and the total volume of gaps plus nanocolumns. It is well known that the refractive
indices of nanoporous materials depend on the porosity, as reported by Yoldas [37]:

8
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 9. Behavior of the refractive index as a function of the wavelength grown at different incident deposition flux angles. The inset
shows the refractive index for the samples with different substrate rotation to α=80°.

Figure 10. Behavior of the refractive index at λ=700 nm as a function of the substrate inclination angle for the AZO films. The
dashed line indicates the standard pattern.

n 2 (l ) - 1
P=1- (5)
nd2 (l) - 1

where n(λ) and nd(λ) represent the refractive index for a porous and dense thin film, respectively. We assume n
as the refractive index of our samples calculated from equation (4) and nd∼1.753 as the value of the refractive
index of AZO crystalline [38]. The porosity values are showed in figure 11, we can see that the porosity increases
significantly as α increases, and there are a real relationship between the porosity and the rotation of the
substrate, the porosity increases as it increases ω as showed by samples S5, S6 and S7 respectively. Clearly, the
optical properties are affected by changing GLAD parameters such as tilde and substrate rotation.
To investigate the possible diffuser phenomenon in samples S5, S6 and S7, we apply the Shirley-George
model [39, 40] used for strong light diffusers in films with vertical nanocolumnar geometry arrays. In this model,
it is considered a surface where the nanocolumn ends its growth with a conical or parabolic tendency. For a
conical profile on the surface of the films, the model assumes an autocorrelation parameter (q) defined by:

q = ld (2ph)2 (6)

where the angle of incidence of the light is normal to the arrangement of the thin films, h is the height and d the
mean value of the diameter of the nanocolumns in the thin film respectively, λ is the wavelength of the light in
the visible spectrum and assumed as λ=550 nm. In this model, values less than q<0.1 in the visible light
spectrum indicate nanocolumnar arrays in the strong diffuser category. The height and diameters were

9
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 11. Behavior of the porosity as a function of the substrate inclination angle for the AZO films.

Figure 12. XRD normalized pattern for AZO thin film growth at different incident deposition flux angles (a) and grown with different
substrate rotation to α=80°(b).

calculated from SEM images and the value of q to the vertical nanoscale of samples S5, S6 and S7 are reported in
table 1. The values q are well below 0.1 indicate that the arrays of columnar nanostructures are in the category of
highly light diffusers.
Figures 12(a) and (b) show the normalized XRD pattern respect to the peak (002), of the samples grown
without and with substrate rotations, figure 12(a) corresponds to samples S1, S2, S3 and S4, and figure 12(b)
correspond to samples S5, S6 and S7, respectively. The samples growth without substrate rotation revealed a
strongly preferential (002) orientation with c-axis perpendicular to the substrate surface. The observed
diffraction peaks (002) and (103) for the samples S1-S4 are consistent with the Wurtzite ZnO hexagonal phase
(JPCDS 36-1451). Figure 12(a) shows only the formations of the peaks (002) and (103), however in figure 12(b),
due to the grown process under oblique flux incidence and the rotation of the substrate, the peak (103) tends to
decrease while the peak (101) appears and increases as the rotational velocity of the substrate increases. Change
in the crystalline texture of the films is promoted, as reported in other articles [41], this orientation has been
attributed to lateral growth and changed of the textura on the Surface layer and they are congruent with recent
articles [24]. XRD shows that the peaks (002) are a displacement on the left side relative the 2θ002 positions of the
standard ZnO powder pattern, as indicated by the discontinuous vertical lines shown in figures 13(a) and (b).
A Voigt adjustment was applied to the peak (002) to obtain the position 2θ002 and the full-width at half-
maximum (FWHM). Figure 13(a) shows that there is no correlation between the position 2θ002 with α, however,
a correlation with the rotation of the substrate is clear as shown in figure 13(b) indicated by the arrow line, where
the position 2θ002 tends at the value of the standard ZnO pattern as the rotation of the substrate increases. Table 2
shows the variation of the position of the peak 2θ002 and the FWHM. The deviations of the diffraction peaks
(002) towards lower angles with respect to ZnO pattern can be attributed to the imperfect stoichiometry and/or
to the presence of a tensile stress in the unit cells of the ZnO. We calculate the value of the crystalline size DXRD of

10
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 13. Normalized diffractograms of the (002) peaks for AZO thin film grown at different incident deposition flux angles (a) and
grown with different substrate rotation to α=80°(b).

Table 2. Results of structural analysis for AZO samples with different


deposition angle and rotation.

Sample 2θ (°) FWHM (°) DXRD (nm) c002-axis (nm)

S1 34.39 0.274 30.31 0.5211


S2 34.28 0.423 19.63 0.5228
S3 34.38 0.319 26.07 0.5212
S4 34.44 0.316 26.31 0.5204
S5 34.31 0.249 33.31 0.5223
S6 34.33 0.230 36.12 0.5219
S7 34.09 0.258 32.13 0.5208

the sample and the residual stress (σ) of the thin films AZO in the direction of the c002-lattice parameter (vertical
direction in the substrate). DXRD and σ depend on the position of the 2θ002 peak that allows us to determine the
effect of the peak broadening using the Scherrer and biaxial strain methods, respectively. From the formula
Scherrer [42]:

0.9l
DXRD = (7)
FWHM cos q002

where FWHM is the full-width at half-maximum of the (002) peak yields to DXRD values as reported in table 2,
and the σ based on the biaxial strain model was calculated by the formula [43]:

s = - 233(c 002 - c 0 ) c 0 GPa (8)

where c002=λ/sinθ002, λ=1.5418 Å wavelength of x-ray source, θ002 Bragg diffraction angle in (002) plane
and c0 is the unstrained lattice parameter of 5.2075 Å from a ZnO powder sample.
The x rays show (table 2) that the lattice is in tensile strain, we observe higher values along the c-axis, in
comparison with the ZnO pattern, which decreases towards the bulk value for the samples S5, S6 and S7.
However, there is no correlation between c002 with α, the degree of crystallinity of samples S5, S6 and S7 of the
FWHM values is much better than samples S1, S2, S3 and S4. The crystalline sizes varied between 19.63 and
36.12 nm, shown nanostructured thin films. The calculated residual stress indicates a very strong compressive
stress along the vertical surfaces of the films, indicating a possible interaction of force between the nanocolumns
and their environment. Figure 14 shows the residual stress as a function of α. Sample S1 shows a relatively low
compressive stress compared to the horizontal line indicating that there is low stress.
When the inclined substrate angle is applied, a strong compressive stress is generated, as shown in S2, and
then α increasing, the stress tends to a state of tensile as indicated by the arrow line inclined between S1, S2 and
S3. A similar behavior is observed with the rotation of the substrate, at low rotation of the substrate a strong
compressive stress is produced as shown at S5, and then ω increasing, the stress tends to a state of tensile as
indicated by the vertical arrow line between S5, S6 and S7.

11
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

Figure 14. Residual stress as a function of the substrate inclination angle. The diagonal and vertical dashed line represents a possible
tendency for the change of substrate inclination and rotation, respectively.

4. Conclusion

AZO nanocolumnar arrays thin films of vertical and with various nanocolumn tilt angle were fabricated on
unheated glass substrates by sputtering rf-reactive magnetron with GLAD deposition geometry at different
glancing-angles and substrate rotations. It was found that the role of deposition geometry is crucial for the
selective formation of nanostructured thin films combined with significant optical and microstructural changes,
such as the refractive index, band-gap energy and residual stress of GLAD AZO films. The increases in the angle
of inclination of the nanocolumns follow an exponential adjustment with the glancing-angle, this is due to the
contribution at the shadowing effect of the decrease in the rate of deposition and the pressure of the gas used
during growth. However, the rotation of the substrate induces the vertical alignment of the nanocolumns. The
hexagonal phase of wurtzite ZnO with preferential diffraction peaks of (002) and (103) was obtained in all
prepared samples. Adjustable band gap energy values (from ~3.45 to ~3.64 eV) and refractive index at λ=700
nm (~1.4 to~1.72) were obtained, which would allow for adjustments in the values the glancing-angle and the
rotation of the substrate to perform engineering of these properties in multiple optoelectronics applications. The
optical and structural properties of the nanostructure’s arrays show differences with the nanocolumnar
morphology and are influenced by changing the tilt or vertical position of the nanocolumns. The vertical
position on nanostructures can provide an essential input for the optimization of a solar cell. Analysis based on
the Shirley-George model for strong nanocolumnar array diffusers, indicate that enhanced scattering light in the
films with low light absorption, showing that the arrays of the vertical columnar of AZO nanostructures are in
the category of highly light diffusers. The deposition of GLAD reactive rf-magnetron sputtering at both,
glancing-angles and the rotation of the substrate is an effective method for the low temperature preparation of
thin films of hexagonal phase of AZO thin films with high porosity with optical and structural quality properties

Acknowledgments

Acknowledgments from the authors to Dr Patricia Quintana for the help of their experimental characterization
laboratories of Conacyt Fomix-Yucatán with contract 2008-108160, and Conacyt-Lab [2009-01-123913, 29-
(2692, 4643), 188345, 204822]. Also thanks Oswaldo Gómez, Mario Herrera, Dora Huerta and Daniel Aguilar
for technical support and Ms Lourdes Pinelo for secretarial assistance.

ORCID iDs

L G Daza https://orcid.org/0000-0002-2334-8187
A Iribarren https://orcid.org/0000-0003-0465-1516

12
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

References
[1] Akhlesh Lakhtakia R M 2005 Sculptured Thin Films: Nanoengineered Morphology and Optics
[2] Hawkeye M M and Brett M J 2007 Glancing angle deposition: fabrication, properties, and applications of micro-and nanostructured
thin films J. Vac. Sci. Technol. A Vacuum, Surfaces, Film. (https://doi.org/10.1116/1.2764082)
[3] Chaudhary A, Klebanov M and Abdulhalim I 2015 PbS nanosculptured thin film for phase retarder, anti-reflective, excellent absorber,
polarizer and sensor applications Nanotechnology (https://doi.org/10.1088/0957-4484/26/46/465703)
[4] Hecht D S, Hu L and Irvin G 2011 Emerging transparent electrodes based on thin films of carbon nanotubes, graphene, and metallic
nanostructures Advanced. Materials. (https://doi.org/10.1002/adma.201003188)
[5] Linnet J, Walther A R, Wolff C, Albrektsen O, Mortensen N A and Kjelstrup-Hansen J 2018 Transparent and conductive electrodes by
large-scale nano-structuring of noble metal thin-films Opt. Mater. Express (https://doi.org/10.1364/OME.8.001733)
[6] Ali N, Hussain A, Ahmed R, Wang M K, Zhao C, Haq B U and Fu Y Q 2016 Advances in nanostructured thin film materials for solar cell
applications Renew. Sustain. Energy Rev. (https://doi.org/10.1016/j.rser.2015.12.268)
[7] Litton C W, Reynolds D C and Collins T C 2011 Zinc Oxide Materials for Electronic and Optoelectronic Device Applications
[8] Gonzalez-Valls I and Lira-Cantu M 2009 Vertically-aligned nanostructures of ZnO for excitonic solar cells: a review Energy Environ.
Sci. (https://doi.org/10.1039/B811536B)
[9] Wang Z 2004 Nanostructures of zinc oxide Mater. Today (https://doi.org/10.1016/S1369-7021(04)00286-X)
[10] Mudusu D, Nandanapalli K R, Dugasani S R, Park S H and Tu C W 2016 Zinc oxide nanorods shielded with an ultrathin nickel layer:
tailoring of physical properties Sci. Rep. (https://doi.org/10.1038/srep28561)
[11] Rowell M W, Topinka M A, McGehee M D, Prall H J, Dennler G, Sariciftci N S, Hu L and Gruner G 2006 Organic solar cells with carbon
nanotube network electrodes Appl. Phys. Lett. (https://doi.org/10.1063/1.2209887)
[12] Wu Y, Zhang X, Jie J, Xie C, Zhang X, Sun B, Wang Y and Gao P 2013 Graphene transparent conductive electrodes for highly efficient
silicon nanostructures-based hybrid heterojunction solar cells J. Phys. Chem. C (https://doi.org/10.1021/jp402529c)
[13] Huang M, Hameiri Z, Gong H, Wong W C, Aberle A G and Mueller T 2014 Novel hybrid electrode using transparent conductive oxide
and silver nanoparticle mesh for silicon solar cell applications Energy Procedia (https://doi.org/10.1016/j.egypro.2014.08.043)
[14] Vanecek M, Babchenko O, Purkrt A, Holovsky J, Neykova N, Poruba A, Remes Z, Meier J and Kroll U 2011 Nanostructured three-
dimensional thin film silicon solar cells with very high efficiency potential Appl. Phys. Lett. (https://doi.org/10.1063/1.3583377)
[15] Zhu K, Neale N R, Miedaner A and Frank A J 2007 Enhanced charge-collection efficiencies and light scattering in dye-sensitized solar
cells using oriented TiO2 nanotubes arrays Nano Lett. (https://doi.org/10.1021/nl062000o)
[16] Hawkeye M M, Taschuk M T and Brett M J 2014 Glancing Angle Deposition of Thin Films: Engineering the Nanoscale
[17] Barranco A, Borras A, Gonzalez-Elipe A R and Palmero A 2016 Perspectives on oblique angle deposition of thin films: from
fundamentals to devices Prog. Mater. Sci. (https://doi.org/10.1016/j.pmatsci.2015.06.003)
[18] Yin G, Sun M, Liu Y, Sun Y, Zhou T and Liu B 2017 Performance improvement in three–dimensional heterojunction solar cells by
embedding CdS nanorod arrays in CdTe absorbing layers Sol. Energy Mater. Sol. Cells (https://doi.org/10.1016/j.solmat.2016.09.040)
[19] Fan Z, Ruebusch D J, Rathore A A, Kapadia R, Ergen O, Leu P W and Javey A 2009 Challenges and prospects of nanopillar-based solar
cells Nano Res. (https://doi.org/10.1007/s12274-009-9091-y)
[20] Tian B, Kempa T J and Lieber C M 2009 Single nanowire photovoltaics Chem. Soc. Rev. (https://doi.org/10.1039/B718703N)
[21] Fan Z et al 2009 Three-dimensional nanopillar-array photovoltaics on low-cost and flexible substrates Nat. Mater. (https://doi.org/
10.1038/nmat2493)
[22] Mbule P, Wang D, Grieseler R, Schaaf P, Muhsin B, Hoppe H, Mothudi B and Dhlamini M 2018 Aluminum-doped ZnO thin films
deposited on flat and nanostructured glass substrates: quality and performance for applications in organic solar cells Solar Energy
(https://doi.org/10.1016/j.solener.2018.03.007)
[23] Khatami S, Fekri L and Pour G B 2018 Investigation of nanostructure and optical properties of flexible AZO thin films at different
powers of RF magnetron sputtering Nano Brief Reports and Reviews (https://doi.org/10.1142/S1793292018500625)
[24] Cosme I, Vázquez-Y-Parraguirre S, Malik O, Mansurova S, Carlos N, Tavira-Fuentes A, Ramirez G and Kudriavtsev Y 2019 Differences
between (103) and (002) x-ray diffraction characteristics of nanostructured AZO films deposited by RF magnetron sputtering Surface
and Coatings Technology (https://doi.org/10.1016/j.surfcoat.2019.05.033)
[25] Robbie K, Sit C and Brett M J 1998 Advanced techniques for glancing angle deposition Journal of Vacuum and Science Technology B
(https://doi.org/10.1116/1.590019)
[26] Akkari F C, Kanzari M and Rezig B 2008 Growth and properties of the CuInS2 thin films produced by glancing angle deposition
Materials Science and Engineering: C (https://doi.org/10.1016/j.msec.2007.10.012)
[27] Vicsek T 1992 Fractal growth phenomena World Scientific 2nd edn (Budapest: Eötvös University)
[28] Ye D X, Zhao Y P, Yang G R, Zhao Y G, Wang G C and Lu T M 2002 Manipulating the columna tilt angles of nanocolumnar films by
glancing-angle deposition Nanotechnology (https://doi.org/10.1088/0957-4484/13/5/314)
[29] SiyanakiF H, Dizaji H R, Ehsani M H and Khorramabadi S 2015 The effect of substrate rotation rate on physical properties of cadmium
telluride films prepared by a glancing angle deposition method Thin Solid Films.
[30] Nieuwenhuizen J M and Haanstra H B 1966 Microfractography of thin films Philips Technical Review
[31] Tait R N, Smy T and Brett M J 1993 Modelling and characterization of columnar growth in evaporated films Thin Solid Films (https://
doi.org/10.1016/0040-6090(93)90378-3)
[32] Xi J Q, Schubert M F, Kim J K, Schubert E F, Chen M, Lin S Y, Liu W and Smart J A 2007 Optical thin-film materials with low refractive
index for broadband elimination of Fresnel reflection Nat. Photonics (https://doi.org/10.1038/nphoton.2007.26)
[33] Berginski M, Hüpkes J, Schulte M, Schöpe G, Stiebig H, Rech B and Wuttig M 2007 The effect of front ZnO:Al surface texture and
optical transparency on efficient light trapping in siliconthin-film solar cells Journal of Applied Physics (https://doi.org/10.1063/
1.2715554)
[34] Grilli M L, Sytchkova A, Boycheva S and Piegari A 2013 Transparent and conductive Al-doped ZnO films for solar cells applications
Physica Status Solidi (a) (https://doi.org/10.1002/pssa.201200547)
[35] Takci D K, Senadim Tuzemen E, Kara K, Yilmaz S, Esen R and Baglayan O 2014 Influence of Al concentration on structural and optical
properties of Al-doped ZnO thin films J. Mater. Sci. Mater. Electron. (https://doi.org/10.1007/s10854-014-1843-0)
[36] Martín-Tovar E A, Denis-Alcocer E, Diaz E C Y, Castro-Rodríguez R and Iribarren A 2016 Tuning of refractive index in Al-doped ZnO
films by rf-sputtering using oblique angle deposition J. Phys. D. Appl. Phys. (https://doi.org/10.1088/0022-3727/49/29/295302)
[37] Tamar Y, Tzabari M, Haspel C and Sasson Y 2014 Estimation of the porosity and refractive index of sol-gel silica films using high
resolution electron microscopy Sol. Energy Mater. Sol. Cells (https://doi.org/10.1016/j.solmat.2014.07.020)

13
Mater. Res. Express 6 (2019) 1050b9 L G Daza et al

[38] Treharne R E, Seymour-Pierce A, Durose K, Hutchings K, Roncallo S and Lane D 2011 Optical design and fabrication of fully sputtered
CdTe/CdS solar cells Journal of Physics: Conference Series (https://doi.org/10.1088/1742-6596/286/1/012038)
[39] Shirley L G and George N 1988 Diffuser radiation patterns over a large dynamic range: I. Strong diffusers Appl. Opt. (https://doi.org/
10.1364/AO.27.001850)
[40] Cansizoglu M F, Engelken R, Seo H W and Karabacak T 2010 High optical absorption of indium sulfide nanorod arrays formed by
glancing angle deposition ACS Nano (https://doi.org/10.1021/nn901180x)
[41] Toledano D, Escobar-Galindo R, Yuste M, Albella J M and Sánchez O 2013 Compositional and structural properties of nanostructured
ZnO thin films grown by oblique angle reactive sputtering deposition: Effect on the refractive index Journal of Physics D: Applied
(https://doi.org/10.1088/0022-3727/46/4/045306)
[42] Fang G, Li D and Yao B L 2002 Fabrication and vacuum annealing of transparent conductive AZO thin films prepared by DC
magnetron sputtering Vacuum (https://doi.org/10.1016/S0042-207X(02)00544-4)
[43] Shi Q, Zhou K, Dai M, Hou H, Lin S, Wei C and Hu F 2013 Room temperature preparation of high performance AZO films by MF
sputtering Ceram. Int. 39 (2) 1135–1141

14

You might also like