Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

This article was downloaded by: [Dicle University]

On: 07 November 2014, At: 18:37


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Philosophical Magazine
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tphm20

Effect of Al doping on structural,


optical and electrical properties of
SnO2 thin films synthesized by pulsed
laser deposition
a b ac
S.K. Sinha , S.K. Ray & I. Manna
a
Department of Metallurgical and Materials Engineering, Indian
Institute of Technology, Kharagpur 721302, India
b
Department of Physics and Meteorology, Indian Institute of
Technology, Kharagpur 721302, India
c
Department of Materials Science and Engineering, Indian
Click for updates Institute of Technology, Kanpur, U.P. 208016, India
Published online: 14 Oct 2014.

To cite this article: S.K. Sinha, S.K. Ray & I. Manna (2014) Effect of Al doping on structural, optical
and electrical properties of SnO2 thin films synthesized by pulsed laser deposition, Philosophical
Magazine, 94:31, 3507-3521, DOI: 10.1080/14786435.2014.962641

To link to this article: http://dx.doi.org/10.1080/14786435.2014.962641

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [Dicle University] at 18:37 07 November 2014
Philosophical Magazine, 2014
Vol. 94, No. 31, 3507–3521, http://dx.doi.org/10.1080/14786435.2014.962641

Effect of Al doping on structural, optical and electrical properties of


SnO2 thin films synthesized by pulsed laser deposition
S.K. Sinhaa, S.K. Rayb and I. Mannaa,c*
a
Department of Metallurgical and Materials Engineering, Indian Institute of Technology,
Kharagpur 721302, India; bDepartment of Physics and Meteorology, Indian Institute of
Technology, Kharagpur 721302, India; cDepartment of Materials Science and Engineering,
Indian Institute of Technology, Kanpur, U.P. 208016, India
(Received 1 April 2014; accepted 2 September 2014)
Downloaded by [Dicle University] at 18:37 07 November 2014

Aluminium-doped (Al = 0–5 wt.%) SnO2 thin films with low-electrical resis-
tivity and high optical transparency have been successfully synthesized by
pulsed laser deposition technique at 500 °C. Structural, optical and electrical
properties of the as-deposited and post-annealed thin films were investigated.
X-ray diffraction patterns suggest that the films transform from crystalline to
amorphous state with increasing aluminium content. The root mean square
(Rq ) surface roughness parameter, determined by atomic force microscopy
decreases upon annealing of the as-deposited film. While resistivity of the film
is the lowest (9.49 × 10−4 Ω-cm) at a critical doping level of 1 wt.% Al, opti-
cal transparency is the highest (nearly 90%) in the as-deposited condition.
Temperature dependence of the electrical resistivity suggests that the Mott’s
variable range hopping process is the dominant carrier transport mechanism in
the lower temperature range (40–135 K) for all the films whereas, thermally
activated band conduction mechanism seems to account for conduction in the
higher temperature region (200–300 K).
Keywords: thin films; X-ray diffraction; atomic force microscopy; transparent
conducting oxides; amorphous

1. Introduction
Demands for materials exhibiting both high optical transparency as well as high
electrical conductivity are rapidly growing for a variety of technological applications in
optoelectronic devices including architectural windows, thin film photovoltaic cells [1],
ion-storage devices [2], flat panel displays [3], defrosting windows in refrigerators and
airplanes [4], gas sensors [5], etc. Transparent conducting oxides (TCO) are used as
filters in heat-efficient window applications that reflect in the infrared and remain trans-
parent in the visible region [6,7]. Most of the well known and widely used TCO thin
films are based on ZnO, CdO, SnO2 and In2O3 and have been investigated for more
than 40 years. These are wide band gap n-type semiconducting material, but studies on
their corresponding p-type substitute for transparent conducting applications have been
seldom reported. Among the many TCO materials investigated so far, the best n-type
TCOs are In2O3:Sn, SnO2:F and ZnO:Al thin films [8]. These materials show optical

*Corresponding author. Email: imanna@metal.iitkgp.ernet.in

© 2014 Taylor & Francis


3508 S.K. Sinha et al.

transparency in excess of 90% and resistivity of the order of 10−4 Ωcm [9,10]. In addi-
tion to developing n-type TCO materials, novel p-type transparent conducting oxides
such as CuAlO2 [11,12], CuGaO2 [13,14], SrCu2O2 [15,16], (La1−xSrxO) CuS [17] and
ZnO:(Ga,N) [18,19] have been actively studied in the recent past. Apart from exploring
new p-type transparent conductors, there is a parallel effort to convert the usual n-type
oxides such as ZnO and SnO2 to p-type compounds. The so-called ‘transparent elec-
tronics’ [20,21] uses the concept of n-type and p-type TCOs in the field of optoelec-
tronic device technology where visible portion of the solar radiation get transmitted, yet
generates electricity by absorbing the ultraviolet rays to form p–n junctions. Tin oxide
(SnO2) is a transparent conducting semiconductor with a direct optical band gap of
about 3.87–4.3 eV [22]. SnO2 thin films can either be polycrystalline or amorphous,
depending upon the deposition technique, post-heat treatment and type and/or level of
doping. It is now recognized that amorphous TCOs possess several benefits as com-
pared to their crystalline counterparts. Since amorphous materials are deposited at lower
Downloaded by [Dicle University] at 18:37 07 November 2014

temperatures or even room temperatures [23], the deposition process proves more con-
venient and simpler. Moreover, it increases the variety of substrate materials that can be
used (such as plastics for flexible electronics) for deposition. Amorphous materials offer
additional advantages in terms of more uniform etchability (as they are isotropic and
devoid of grain boundaries [24]) and hence, lower surface roughness [25]. It may be
pointed out that amorphous TCO materials offer the above advantages without signifi-
cantly compromising with two of their most important properties; namely, electrical
conductivity and optical transparency [26].
Although a limited number of investigations on synthesis of Al-doped SnO2 has
been conducted [27,28], transport properties of such Al-doped SnO2 films at low tem-
perature have not been determined to date. It is, therefore, imperative to understand the
transport mechanism of charge carriers of doped SnO2 thin films and assess the scope
of exploiting these films in device applications. Thus, in the present paper, we report
the systematic investigation of the structural, optical and electrical properties of pulsed
laser deposited Al-doped SnO2 thin films both in as-deposited and annealed condition.

2. Experimental
Pure tin oxide (99.99% pure, Aldrich) and aluminium oxide (99.99% pure, Aldrich)
containing 1, 3 and 5 wt.% Al2O3 were properly weighed, mixed and thoroughly
ground in a mortar by mechanical means to produce homogeneous distribution of Al in
SnO2. These powder blends were subsequently used to produce 25.4 mm diameter disks
through cold pressing by applying 6 ton load followed by sintering in air at 1000 °C
for 4 h. The sintering routine included heating and cooling at a rate of 6 °C per minute
in oxygen ambience prior to and after isothermal holding at 1000 °C for 4 h. Films
were deposited on borosilicate glass substrates (to study the optical transmittance spec-
tra only) and p-Si (100) wafers (resistivity 7–14 Ω-cm) of 10 mm × 10 mm dimensions
located 45 mm away from the target. The substrates were mounted on a special sub-
strate holder inside the deposition chamber and placed parallel to the target. The sub-
strate holder was heated to maintain an isothermal temperature (500 °C) and ensure a
uniform film thickness. KrF excimer laser (Lambda Physik COMPEX, wave-
length = 248 nm and pulsed duration = 25 ns) was used to deposit the films. During
deposition, the target was continuously rotated to ensure uniform ablation and avoid
Philosophical Magazine 3509

denting or drilling. The optimum conditions for deposition were: laser energy = 300 mJ,
base pressure = 5 × 10−5 mbar, oxygen pressure = 1.1 × 10−1 mbar, substrate tempera-
ture = 500 °C, repetition rate = 10 Hz and deposition length = up to 15,000 pulses.
Although the as-deposited films tend to vary in thickness, only the films with uniform
thicknesses (750 ± 25 nm) were selected to study the optical and electrical properties.
The as-deposited Al-doped SnO2 films were annealed at 600 °C for 1 h in oxygen
ambience (2 × 10−1 mbar). Recently, the present authors have reported deposition of
high-quality ZnO-SnO2 composite thin films by the same pulse laser deposition tech-
nique, albeit with a very different composition than the present system (Al-doped SnO2
film) and for a different scope of application [29]. The structural properties of the
as-deposited and annealed films have been studied by grazing incidence X-ray diffrac-
tometer (PANalytical-X’pert PRO, Cu Kα radiation at λ = 0.15418 nm) in the 2θ range
20° – 65°. The microstructure and surface morphology of the films have been analyzed
using field emission scanning electron microscope (FESEM, Carl Zeiss, Supra 40) and
Downloaded by [Dicle University] at 18:37 07 November 2014

Veeco nanoscope-IV atomic force microscopy (AFM) in tapping mode, respectively.


Optical transmittance spectra were recorded by Perkin-Elmer Lambda 45 spectropho-
tometer in the wavelength range 300–1100 nm. The sheet resistance (Rs) of the
Al-doped SnO2 thin films was measured by standard four-probe technique. Resistance
of each film represents an average of at least five different measurements at selected
and equivalent locations on the film surface. The resistivity of the Al-doped SnO2 films
was measured in the temperature range 10–300 K using a liquid nitrogen cryostat sys-
tem (Model HC-4E, Sumitomo Cryogenics) using the four-point probe technique. A
331-Lakeshore temperature controller was used for maintaining isothermal temperature.
The measurements were taken during both cooling and heating processes in the
above-mentioned temperature range, considering a predetermined film thickness.

3. Results and discussion


The GIXRD patterns shown in Figure 1(a–f) present the overall state of crystallinity of
the as-deposited and annealed Al-doped SnO2 thin films with varying aluminium con-
centration. It is evident from the GIXRD plots that with increasing Al concentration,
the films transformed from a perfectly crystalline (1 wt.% Al: SnO2) to completely
amorphous (5 wt.% Al: SnO2) state, both in the annealed and in the as-deposited condi-
tions. The presence of only a broad and diffused halo in the diffraction profile suggests
that the deposited film is mostly amorphous in nature, particularly at higher Al concen-
tration. The (110), (200), (111), (211), (220) and (310) planes of tetragonal rutile struc-
ture of SnO2 (JCPDS card no. 72-1147) can be readily indexed in the diffraction
pattern of the annealed films. The peak found at ~25.6° in the as-deposited state of
Al-doped SnO2 films shows the formation of α-Al2O3 (JCPDS card no. 42-1468) phase
which primarily comes from the PLD target. Moreover, the GIXRD patterns does not
detect any peak from the impurities, such as unreacted Sn or any other tin oxides
(except SnO), confirming the high purity of the deposited product. Furthermore, no
diffraction peak corresponding to aluminium or its oxides has been detected. A slight
shift of the major SnO2 peaks suggests that a measurable amount of Al has been
dissolved in the SnO2 films. The decrease in crystallinity and subsequent structural
transformation of the film to completely amorphous state at higher Al concentration
(5 wt.%) might be attributed to the stress generated due to the difference between Sn4+
3510 S.K. Sinha et al.
Downloaded by [Dicle University] at 18:37 07 November 2014

Figure 1. (colour online) GIXRD patterns of the as-deposited (plot a-c) and thermally annealed
(plot d–f) Al-doped SnO2 thin films with various Al concentrations (1, 3 and 5 wt.% Al).

(0.071 nm) and Al3+ (0.051 nm) ionic radii. The substitution of Sn4+ ions by about
30% smaller Al3+ ions and possible creation of interstitialcy beyond the permissible
limit may promote transformation of the crystalline film into disordered amorphous
state. The segregation of Al3+ dopant in the grain boundary with increasing dopant con-
centration is also believed to promote lack of long-range periodicity in the deposited
film. Indeed, an earlier investigation confirms that crystallinity of Al-doped SnO2 films
decreases with increasing Al concentration [30].
Figure 2(a–f) shows the FESEM images of Al-doped SnO2 thin films with 1, 3 and
5 wt.% Al concentration. The films with 1 wt.% Al, both in annealed and in as-deposited
state reveals a homogeneous onion-like grain structure resembling crystalline SnO2, as
reported in our previous paper [31]. The nanocrystalline grain size (and subsequently the
grain boundary specific area) has important consequences on controlling the electrical,
optical and magnetic properties in doped or undoped metal oxides. Straumal et al. [32,33]
have shown that doped or undoped ZnO becomes ferromagnetic only if its grain bound-
ary-specific area reaches a certain threshold value. In our films too, it can be inferred that
the films with crystalline grain structure (1 wt.% Al: SnO2) exhibits distinctly better opti-
cal and electrical properties as compared to that of its amorphous counterpart, as discussed
in the subsequent sections. As the Al concentration increases, the films reveals
amorphous-like morphology containing no distinct grain or grain boundary-like feature.
Moreover, particle coalescence occurs in a few places both in 3 and 5 wt.% Al-doped
SnO2 thin films. Figure 3(a)–(f) show the AFM deflection images (5 μm × 5 μm) of
as-deposited and annealed Al-doped SnO2 thin films on Si (100) substrates. As all the
Philosophical Magazine 3511
Downloaded by [Dicle University] at 18:37 07 November 2014

Figure 2. FESEM images of the Al-doped SnO2 thin films deposited on Si (100) substrate at
500 °C and annealed at 600 °C for 1 h in oxygen. (a) 1 wt.% Al: SnO2 (as-deposited) (b) 1 wt.%
Al: SnO2 (annealed) (c) 3 wt.% Al: SnO2 (as-deposited) (d) 3 wt.% Al: SnO2 (annealed) (e) 5
wt.% Al: SnO2 (as-deposited) (f) 5 wt.% Al: SnO2 (annealed).

films have been deposited at the same substrate temperature (500 °C), the surface
roughness and constitution of the films are influenced mainly by Al concentration and
post-deposition annealing conditions. Between these two parameters, as the dopant con-
centration differs by 4 wt.%, the surface morphology is more significantly influenced by
the post-deposition annealing treatment. The AFM images reveal that the annealed films
exhibit different surface morphological features and relatively smoother (ignoring lump
formation in few places) topology than that of the as-deposited ZnO films. The root mean
square (Rq ) surface roughness values of the annealed films decrease considerably from the
Downloaded by [Dicle University] at 18:37 07 November 2014

3512
S.K. Sinha et al.

Figure 3. (colour online) 2D-AFM images (5 μm × 5 μm area) of Al-doped SnO2 thin films prepared by PLD (a) 1 wt.% Al: SnO2 (as-deposited) (b) 3
wt.% Al: SnO2 (as-deposited) (c) 5 wt.% Al: SnO2 (as-deposited) (d) 1 wt.% Al: SnO2 (annealed) (e) 3 wt.% Al: SnO2 (annealed) (f) 5 wt.% Al: SnO2
(annealed).
Philosophical Magazine 3513

corresponding value in as-deposited condition, as shown in the inset of AFM images.


However, the as-deposited and annealed films with 1 wt.% Al show crystalline morphol-
ogy with distinct grain structure, as already confirmed by XRD (Figure 1) and FESEM
image (Figure 2). With increase in temperature during post-deposition annealing, the sur-
face of the films turns increasingly smoother due to enhanced mobility of surface atoms.
Furthermore, annealing aids deeper penetration of Al into the film. Figure 4 shows the
measured optical transmittance of the as-deposited and annealed Al-doped SnO2 thin
films in the range 300 to 1100 nm. The plots indicate that the average optical
transmittance of the samples in the visible region (420–780 nm) remains between 66 and
90%. This makes the Al doped-SnO2 films suitable for use as window material. The
decreasing trend of the optical transmittance with increasing Al concentration as observed
from Figure 4 might be attributed to an increase in scattering centers (increase in
interstitialcy or decrease in voids due to inadequate solid solubility of Al in SnO2). It can
be seen that the average transmittance in the visible wavelength region is about 80% for
Downloaded by [Dicle University] at 18:37 07 November 2014

all the films (except the film with 1 wt.% Al annealed) with a sharp fundamental absorp-
tion edge. It is also noticed that the absorption edges shift towards lower energies i.e.
longer wavelengths (except 5 wt.% Al:SnO2 annealed film) and hence, lower band gap
with increasing Al concentration in the films. The absorption edge for direct interband
transition is given by [34]:

1=2
ahm ¼ Cðhm  Eg Þ (1)
where C is a constant for direct transition, α is the optical absorption coefficient, Eg is
the optical band gap and hν denotes the photon energy.

Figure 4. (colour online) Optical transmission spectra of Al-doped SnO2 thin films in as-deposited
and annealed conditions deposited by PLD using SnO2 targets containing 1, 3 and 5 wt.% Al.
3514 S.K. Sinha et al.

Eg can be obtained by extrapolating the linear region near the onset in a plot of
(αhν)2 versus hν, as shown in Figure 5(a). Accordingly, the Eg values are 3.03, 3.404,
3.20, 3.26, 2.92 and 3.49 eV for the films with 1 wt.% Al (as-deposited), 1 wt.% Al
(annealed), 3 wt.% Al (as-deposited), 3 wt.% Al (annealed), 5 wt.% Al (as-deposited)
and 5 wt.% Al (annealed)-doped SnO2 films, respectively. The direct optical band gap
of pure polycrystalline SnO2 film is ~3.6 eV. The decrease in optical band gap of the
Al-doped SnO2 thin films might be attributed to the Burstein–Moss effect [35,36]. The
crystallite or particle size effect [37] on band gap can be ignored here as the films are
essentially amorphous in nature particularly at higher (>3 wt%) Al concentration.
It is known that higher optical transmittance and lower sheet resistance of thin film
are primary requirement for enhanced opto-electronic performance of transparent
Downloaded by [Dicle University] at 18:37 07 November 2014

Figure 5. (colour online) (a) Band gap (Eg) estimation for Al-doped SnO2 thin films using Tauc’s plot
and (b) Figure of merit plot of Al-doped SnO2 thin films in as-deposited and annealed conditions.
Philosophical Magazine 3515

conductive oxide materials. In practice, there is a trade-off between these two


parameters. In order to estimate the quality of Al-doped SnO2 thin films, the figure of
merit (ΦTC) defined by Haacke [38], is calculated as a criterion for the quality or
performance of the films, and is expressed as follows:
10
Top
UTC ¼ (2)
Rs
where Top is the optical transmittance at λ = 550 nm (%) and Rs is the sheet resistance
(Ω/unit area). The result shows that the figure of merit value is higher in the annealed
state compared to that in the as-deposited films. The maximum figure of merit is
obtained for the film with 3 wt.% Al. Figure 5(b) shows a comparative plot for figure
of merit of annealed and as-deposited Al-doped SnO2 thin films. It is evident that
annealing of the as-deposited films increases ΦTC values due to the decrease in sheet
resistance as well as enhancement of optical transparency.
Downloaded by [Dicle University] at 18:37 07 November 2014

To understand electronic conduction and electrical transport mechanism of the


Al-doped tin oxide thin films (annealed only), low temperature electrical conductivity
was studied, in the range 10 K to room temperature (300 K). Figure 6(a–b) shows the
variation of resistivity with temperature for the annealed Al-doped SnO2 thin films. It is
apparent that resistivity decreases with increase in temperature (in the range 75–300 K)
showing the semiconducting behavior of the deposited films. At low temperature
regions, the transport shows a metallic behavior with the metal-to-insulator transition
temperature (Tc) tending to decrease with increase in Al%. A slight deviation of increas-
ing resistivity in the lower temperature regime (<30 K) of the films with 1 and 5 wt.%
Al could be attributed to the associated structural disorder which dominates at very low
temperature. For the temperature range 30–150 K, the decrease in resistivity is steeper
for the 3 and 5 wt% Al-doped films compared to that in the 1 wt.% Al-doped film. The
resistivity value increases with Al concentration indicating the incorporation of dopant
in the interstitial sites. The temperature dependence of electrical conductivity (σ) as a
function of temperature (T) for the disordered materials can be expressed as [39]:
r ¼ r0 expðEa =kT Þ (3)

where σ0 is a pre-exponential constant, Ea is the activation energy and k is the


Boltzmann constant. Ea were calculated for each film from the linear Arrhenius plots
(Figure 7) fitted to the data of ln σ plotted against 1/T of the annealed Al-doped tin
oxide thin films in the desired temperature range. The straight line fits well with the
experimental data in the upper range of temperature (200–300 K). In the lower tempera-
ture region (<180 K), regression analysis does not conform to Arrhenius relationship
(i.e. absence of straight line fit), which substantiates that thermally activated conduction
mechanism does not prevail in the low-temperature region though the same may
dominate in the higher temperature range. The spread in the activation energy values
with varying Al concentration is confirmed by the different slopes of the Arrhenius-type
distribution of conductivity in the high-temperature region of the ln σ versus 1/T plots.
The activation energy of the post annealed Al-doped SnO2 films with 1, 3 and
5 wt.% Al were calculated as 68.04, 82.03 and 102.6 meV, respectively. The activa-
tion energy for the thermally activated band conduction varies with the donor or
acceptor carrier concentration and the impurity energy levels of the semiconductor.
3516 S.K. Sinha et al.
Downloaded by [Dicle University] at 18:37 07 November 2014

Figure 6. (colour online) Temperature dependence resistivity plots for the annealed Al-doped
SnO2 thin films deposited with (a) 1 wt.% Al and (b) 3 and 5 wt.% Al concentration.

It is found that the solid solubility of Al in SnO2 is limited to less than 1 at.% only
[40]. Therefore, it can be assumed that Al does not act as dopant at higher Al con-
centration in Al-doped SnO2 films. Instead it goes to interstitial sites of the SnO2
crystal. This in turn increases the non-stoichiometric oxygen in the SnO2 lattice,
subsequently reducing the oxygen vacancy concentration, which in turn leads to
increase in resistivity (or decrease in conductivity) of the film. This explanation
agrees well with the Kröger–Vink defect model reported by Freeman et al. [41].
The resistivity of the Al-doped SnO2 films thus increase with increase in Al content.
Earlier, in case of pure SnO2, it has been found that increasing the oxygen partial
pressure (i.e. decreasing the oxygen vacancy) leads to a decrease in electrical
Philosophical Magazine 3517
Downloaded by [Dicle University] at 18:37 07 November 2014

Figure 7. (colour online) Temperature dependence of electrical conductivity plotted as ln σ versus


T−1 for annealed Al-doped SnO2 thin films with varying Al (1, 3 and 5 wt.%) concentration. The
regression analysis plots are presented as broken lines.

conductivity of the films after annealing [42,43]. Furthermore, the increase in resis-
tivity could also be attributed to an increase in the amorphous phase and degree of
disorder with increasing Al content, as evident from GIXRD patterns (Figure 1). Xu
et al. [40] have shown that there is a sharp decrease in carrier concentration (n) of
Al-doped SnO2 powders with an increase in Al3+ amount as dopant from 0.001 to
1%. The Al3+ ions, with three valence electrons, substitute Sn4+ ion in the SnO2 lat-
tice resulting in the creation of a hole in the cationic sub-lattice for every three sub-
stitution. Ahmed et al. [30] have shown that up to a critical amount (12.5%) the
conductivity of Al-doped SnO2 thin films decreases. In our films prepared by PLD,
the room temperature (300 K) conductivity value decreases from 12.3 × 103 S/cm to
3.52 × 103 S/cm, with increasing Al concentration from 1 to 5%. Table 1 shows the
optical and electrical properties of Al-doped SnO2 thin films obtained by various
deposition techniques, as reported in the literature.

Table 1. Comparison of optical and electrical properties of aluminium-doped tin oxide thin films.

Optical Electrical
Deposition transmittance Optical band conductivity
Technique Composition (% T) gap (eV) (S cm−1) Refs.
Spray 3.5–12.1 at.% >75 3.98–4.10 14.92–1.11 [28]
pyrolysis Al–SnO2
Sol gel 2.31–18.56 at.% >80 3.87–4.21 0.21–1.36 [31]
Al–SnO2
Spray 3 at% Al–SnO2 >85 3.7 104 [46]
pyrolysis
PLD 1–5 wt.% Al: >75 3.03–3.49 12.3 × 103 Present
SnO2 −3.52 × 103 work
3518 S.K. Sinha et al.

At sufficiently low temperature, the carrier transport mechanism can be explained as


the phonon-assisted hopping conduction mechanism or the variable range hopping
(VRH) model, originally proposed by Mott [44,45]. As per this model, the electrons
hop between the localized states, close to the Fermi level, as localized states exist inside
the energy distribution region of the impurity level within the energy band gap. In case
of polycrystalline materials, the VRH conduction mechanism occurs in the grain bound-
ary regions at sufficiently higher temperatures where the carriers do not have enough
energy to cross the potential barrier and transfer themselves into the crystallite state by
thermionic emission. According to the Mott’s VRH model, the relationship between the
conductivity and the temperature of disordered semiconductor material is expressed by
the following relation,
"   #
1=2 T0 1=4
r ¼ r0 T exp  (4)
T
Downloaded by [Dicle University] at 18:37 07 November 2014

where σ0 is pre-exponential factor and T0 is a parameter related to the disorder of the


system. These parameters are described by the following expressions:
 
N ðEF Þ 1=2 ka3
r0 ¼ 3e mph
2
; T0 ¼ (5)
8paK kN ðEF Þ
Here, N(EF) is the density of localized states near the Fermi level, λ is a dimensionless
constant (~18), K is the Boltzmann constant, α is the inverse localization length of wave
function associated with the localized state near the Fermi level and νph is the Debye
frequency for SnO2 (~1013).
Mott also expressed average hopping energy (W) and average hopping distance (R)
as [44]
 1=4
9 4
R¼ ; W ¼ (6)
8pakTN ðEF Þ 3pR3 N ðEF Þ
The values of T0 are calculated from the slopes of the fitted straight lines obtained from
Equation (4) and those of σ0 are obtained by its intercepts. Using T0, σ0, Equations (5)
and (6), the values of N(EF), α, R and W are calculated and are summarized in Table 1.
These data are comparable to the previously reported F-doped tin oxide films prepared
by spray-pyrolysis technique [45]. It may be noted that the Mott parameters satisfy the
condition αR >> 1 and W >> kT for all composition of Al-doped SnO2 thin films in the
present study, thereby justifying the applicability of Mott’s variable range hopping con-
duction in low temperature (<135 K) region.
Figure 8 shows the temperature dependence of the conductivity for the Al-doped
SnO2 films deposited by pulsed laser deposition at 500 °C in oxygen ambience, plotted
as ln (σT1/2) versus T−1/4 within 30–135 K temperature regimes. As shown in Figure 8,
all the plots exhibit the linear relationship within the region 50–125 K with the fit being
better for the higher doped films with 3 or 5 wt.% Al. The VRH conduction of elec-
trons, in general, occurs between localized band-tail states of amorphous or disordered
materials. The films with 3 and 5 wt.% Al shows an improved fitting regime since they
are amorphous in nature as compared to the predominantly crystalline film with 1 wt.%
Al doping. This difference in fitting may arise primarily due to the relatively higher
Philosophical Magazine 3519
Downloaded by [Dicle University] at 18:37 07 November 2014

Figure 8. (colour online) Temperature dependence conductivity of annealed Al-doped SnO2 thin
films plotted as ln (σT1/2) versus T−1/4. The broken lines represent regression analysis plots as per
Equation (4).

extent of crystallinity in the concerned film with 1 wt.% Al. However, the extrapolation
and related results may still be acceptable due to the nearly horizontal (very shallow
gradient) nature of the linear portion of the curve fitted to the VRH model proposed by
Mott. The rising slope at the end of the curve for the film with 1 wt.% Al may also be
related to the composite nature of the film with both crystalline and amorphous regions.
Moreover, with increasing doping concentration, a considerable number of impurity
atoms were introduced in the band-tail region of host SnO2 structure that could account
for improved linear fitting of the other two curves for film with higher dopant
concentration.
From Table 2, it can be concluded that the films with higher Al concentration have
a lower density of the localized electron state and have a much lower localization
length of the wave function compared with the films with lower Al doping concentra-
tion. Finally, the carrier transport mechanism in the Al-doped SnO2 films can be attrib-
uted to Mott’s variable range hopping in the temperature range below 135 K and
thermally activated band conduction above 200 K.

Table 2. Summary of the values of Mott’s variable range hopping parameters.

N(EF)
Film composition T0 (K) α (cm−1) (cm− 3 eV− 1) R (cm) αR W (eV)
5 10 21 −8 2
1 wt.% Al 14.6 × 10 1.85 × 10 8.15 × 10 4.56 × 10 8.44 × 10 0.548
3 wt.% Al 11.03 × 103 2.63 × 108 3.37 × 1018 10.10 × 10−7 2.66 × 102 0.122
5 wt.% Al 3.90 × 103 1.84 × 107 3.27 × 1015 18.12 × 10−6 3.33 × 102 0.021
3520 S.K. Sinha et al.

4. Conclusions
In an attempt to produce low-cost and novel transparent conducting oxides, p-type
Al-doped SnO2 thin films (containing 1, 3 and 5 wt.% Al) were deposited by pulsed
laser deposition technique on p-Si (100) and glass substrates at 500 °C to study the
effect of aluminium concentration and post-deposition annealing on structural, optical
and low-temperature resistivity properties. GIXRD and SEM results indicate that an
increase in dopant concentration (Al3+) promotes amorphous instead of crystalline state
and deteriorates optical transparency and electrical conductivity. Only the films with
1 wt.% Al are essentially polycrystalline single phase SnO2 (rutile) both in as-deposited
and annealed condition. The optical transmittance of the Al-doped SnO2 films was
higher than 80% (except the film with 1 wt.% Al annealed) in the visible wavelength
spectrum with a sharp fundamental absorption edge. The combined transparent conduct-
ing oxide property (expressed by the figure of merit), shows that the opto-electronic
property of the films can be improved by annealing compared to that in as-deposited
Downloaded by [Dicle University] at 18:37 07 November 2014

condition and the highest ΦTC has been achieved in the film with 3 wt.% Al in annealed
condition. Temperature-dependent resistivity from 10 to 300 K reveals that the carrier
transport processes are governed by thermally activated band conduction between
200 and 300 K, and Mott’s VRH conduction mechanism below 135 K. The decrease in
Al–SnO2 film conductivity for higher Al fraction may be attributed to the decrease in
oxygen-vacancy arising from increasing Al2O3 content, owing to its limited solid solu-
bility in SnO2. The results obtained thus indicate that high-quality Al-doped tin oxide
thin films can be produced by pulsed laser deposition technique at elevated tempera-
tures and offer a potential and cost effective material for transparent conducting and
other opto-electronic applications.

Acknowledgements
Partial financial supports from the Department of Science and Technology (DST) to I. Manna
(J C Bose fellowship) and Council of Scientific and Industrial Research (CSIR), Government of
India (Project: GAP 0280 and SRF support to S.K. Sinha) are gratefully acknowledged.

References

[1] K. Omura, P. Veluchamy, M. Tsuji, T. Nishio and D. Murojono, J. Electrochem. Soc. 146
(1999) p.2113.
[2] B. Orel, U. Lavrencic-Stangar and K. Kalcher, J. Electrochem. Soc. 141 (1994) p.L127.
[3] S.J. Laverty, H. Feng, P. Maguire. J. Electrochem. Soc. 144 (1997) p.2165.
[4] Z. Hamberg, J.S.E.M. Svesson, T.S. Eriksson, C.G. Granqvist, P. Arrenius and F. Narin,
Appl. Opt. 26 (1987) p.2131.
[5] A.V. Tadeev, G. Delabouglise and M. Labeau, Mater. Sci. Engg. B 57 (1998) p.76.
[6] K.L. Chopra, S. Major and D.K. Pandya, Thin Solid Films 102 (1983) p.1.
[7] A.L. Dawar and J.C. Joshi, J. Mater. Sci. 19 (1984) p.1.
[8] A.J. Freeman, K.R. Poeppelmeier, T.O. Mason, R.P.H. Chang and T.J. Marks, MRS Bull. 25
(2000) p.45.
[9] D.S. Ginley and C. Bright, MRS Bull. 25 (2000) p.15.
[10] B.G. Lewis and D.C. Paine, MRS Bull. 25 (2000) p.22.
[11] Y. Wang, H. Gong, F. Zhu, L. Liu, L. Huang and A.C.H. Huan, Mater. Sci. Eng. B 85
(2001) p.131.
Philosophical Magazine 3521

[12] K. Tonooka, K. Shimokawa and O. Nishimura, Thin Solid Films 411 (2002) p.129.
[13] H. Kawazoe, M. Yasukawa, H. Hyodo, M. Kurita, H. Yanagi and H. Hosono, Nature 389
(1997) p.939.
[14] H. Yanagi, H. Kawazoe, A. Kudo, M. Yasukawa and H. Honoso, J. Electroceram. 4 (2000)
p.427.
[15] A. Kudo, H. Yanagi, H. Hosono and H. Kawazoe, Appl. Phys. Lett. 73 (1998) p.220.
[16] H. Ohta, K.I. Kawamura, M. Orita, M. Hirano and H. Hosono, Electron. Lett. 36 (2000)
p.984.
[17] H. Hiramatsu, K. Ueda, H. Ohta, M. Orita, M. Hirano and H. Hosono, Thin Solid Films 411
(2002) p.125.
[18] T. Yamamoto H. Katayama-Yoshida, Japan. J. Appl. Phys. Part 2: Lett. 38L (1999). p. 166
[19] M. Joseph, H. Tabata, H. Saeki, K. Ueda and T. Kawai, Physica B 302 (2001) p.140.
[20] H. Ohta, K.I. Kawamura, M. Orita and M. Hirano, Appl. Phys. Lett. 77 (2000) p.475.
[21] M.K. Jayaraj, A.D. Draeseke, J. Tate, R.L. Hoffman and J.F. Wager, Mater. Res. Soc. Symp.
Proc. 666 (2001) p. F311.
Downloaded by [Dicle University] at 18:37 07 November 2014

[22] K.L. Narayanan, R. Rajaraman, M.C. Valsakumar, K.G.M. Nair and K.P. Vijayakumar,
Mater. Res. Bull. 34 (1999) p.1729.
[23] D.B. Buchholz, D.E. Proffit, M.D. Wisser, T.O. Mason and R.P.H. Chang, Prog. Nat Sci:
Mat. Int. 22 (2012) p.1.
[24] K. Nomura, H. Ohta, A. Takagi, T. Kamiya, M. Hirano and H. Hosono, Nature 432 (2004)
p.488.
[25] W.M. Kim, D.Y. Ku, I.K. Lee, Y.W. Seo, B.K. Cheong, T.S. Lee, I.H. Kim and K.S. Lee,
Thin Solid Films 473 (2005) p.315.
[26] M. Warasawa, A. Kaijo and M. Sugiyama, Thin Solid Films 520 (2012) p.2119.
[27] C. Xu, J. Tamaki, N. Miura and N. Yamazoe, Talanta 38 (1991) p.1169.
[28] M. Mohagheghi and M. Saremi, J. Phys. D: Appl. Phys. 37 (2004) p.1248.
[29] S.K. Sinha, T. Rakshit and S.K. Ray, I. Manna, Appl. Surf. Sci. 257 (2011) p.10551.
[30] Sk.F. Ahmed, S. Khan, P.K. Ghosh, M.K. Mitra, K.K. Chattopadhyay, J. Sol-Gel. Sci.
Techn. 39(2006) p. 241.
[31] S.K. Sinha, R. Bhattacharya, S.K. Ray and I. Manna, Mater. Lett. 65 (2011) p.146.
[32] B. B. Straumal, S.G. Protasova, A.A. Mazilkin, T. Tietze E. Goering, G. Schütz, P.B. Straumal,
B. Baretzky, J. Beilstein Nanotechnology 4 (2013) p. 361.
[33] B.B. Straumal, S.G. Protasova, A.A. Mazilkin, G. Schütz, E. Goering, B. Baretzky and
P.B. Straumal, J. Exp. Theor. Phys. Lett. 97 (2013) p.367.
[34] J.I. Pankove, Optical Processes in Semiconductors, Dover, New York, 1971.
[35] E. Burstein, Phys. Rev. 93 (1954) p.632.
[36] T.S. Moss, Proc. Phys. Soc. London. 67 (1954) p.775.
[37] J. Pal and P. Chauhan, Mater. Charact. 60 (2009) p.1512.
[38] G. Haacke, J. Appl. Phys. 47 (1976) p.4086.
[39] K.L. Chopra, Thin Film Phenomena, McGraw-HillNew York, 1969.
[40] C. Xu, J. Tamaki, N. Miura and N. Yamazoe, Talanta 38 (1991) p.1169.
[41] A.J. Freeman, K.R. Poeppelmeier, T.O. Mason, R.P.H. Chang and T.J. Marks, MRS Bull 25
(2000) p.45.
[42] Z.M. Jarzebski and J.P. Morton, J. Electrochem. Soc. 123 (1976) p.299.
[43] S. Samson and C.G. Fonstad, J. Appl. Phys. 44 (1973) p.4618.
[44] N.F. Mott, J. Non-Cryst. Solids 1 (1968) p.1.
[45] K.S. Ramaiah and V.S. Raja, Appl. Surf. Sci. 253 (2006) p.1451.
[46] M. Benhaliliba, C.E. Benouis, Y.S. Ocak and F. Yakuphanoglu, J. Nano – Electron. Phys 4
(2012) p.01011.

You might also like