You are on page 1of 9

Superlattices and Microstructures 100 (2016) 228e236

Contents lists available at ScienceDirect

Superlattices and Microstructures


journal homepage: www.elsevier.com/locate/superlattices

Electronic transport mechanism in intrinsic and doped


nanocrystalline silicon films deposited by RF-magnetron
sputtering at low temperature
D. Benlakehal a, *, A. Belfedal a, b, Y. Bouizem a, J.D. Sib a, L. Chahed a,
K. Zellama c
a
LPCMME, D epartement de Physique, Universite d’Oran 1, Algeria
b
Laboratoire de Chimie Physique des Macromol ecules et Interfaces Biologiques, Universit
e de Mascara, Algeria
c
LPMC, UFR des Sciences, Universit
e de Picardie Jules Verne, 33 rue Saint-Leu, 80039, Amiens, France

a r t i c l e i n f o a b s t r a c t

Article history: The dependence on the temperature range, T, of the electronic transport mechanism in
Received 17 August 2016 intrinsic and doped hydrogenated nanocrystalline silicon films, deposited by
Received in revised form 22 September 2016 radiofrequency-magnetron sputtering at low substrate temperature, has been studied.
Accepted 26 September 2016
Electrical conductivity measurements s(T) have been conducted on these films, as a
Available online 28 September 2016
function of temperature, in the 93e450 K range. The analysis of these results clearly shows
a thermally activated conduction process in the 273e450 K range which allows us to es-
Keywords:
timate the associated activation energy as well as the preexponential conductivity factor.
Activation energy
Range hopping
While, in the lower temperature range (T < 273 K), a non-ohmic behavior is observed  for
the conductivity changes. The conductivity s(T) presents a linear dependence on T 4 ,
1
Hopping energy
Nanocrystalline silicon and a hopping mechanism is suggested to explain these results. By using the Percolation
theory, further information can be gained about the density of states near the Fermi level
as well as the range and the hopping energy.
© 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
2. Experimental details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
3. Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
3.1. Structural results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
3.2. Electronic results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
3.2.1. Thermally activated process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
3.2.2. Hopping transport mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

* Corresponding author. Tel.: þ213 557052503; fax: þ213 41581941.


E-mail address: djbenlakehal@gmail.com (D. Benlakehal).

http://dx.doi.org/10.1016/j.spmi.2016.09.035
0749-6036/© 2016 Elsevier Ltd. All rights reserved.
D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236 229

1. Introduction

Hydrogenated nanocrystalline silicon (nc-Si:H) material grown at low temperature has attracted considerable attention in
the last decades due to its remarkable electronic properties [1e3], and for its promising applications such as thin film
transistors, flat panel displays, and also for its increased stability against light-induced degradation [3e5]. This material
exhibits a variety of different mixed complexes structures consisting in nano-sized crystallites embedded in an amorphous
matrix with high fraction of voids and hydrogen. This heterogeneous structure is directly connected to the growth mechanism
and, therefore, to the plasma conditions parameters and the deposition technique. However, growth of nc-Si:H with good
structural and optoelectronic properties at relatively low temperature (<150  C), compatible with the use of flexible poly-
meric substrates, is the challenge for many research groups [1,3e8]. The relationship between the deposition parameters and
growth mechanism of nc-Si:H is still a subject of debate [7e10].
In ours previous studies we have investigated the effects of different plasma conditions on the structural and the opto-
electronic properties of intrinsic and doped nc-Si:H films prepared by radiofrequency-magnetron sputtering (RFMS) at low
temperature [11e15]. The results obtained in these studies show that the optimal plasma conditions for the growth of nc-Si:H
by RFMS technique are a total pressure of 3 Pa, a threshold of RF-power (200e250 W) and a gas mixture of 30% of Ar and 70%
of H2.
The heterogeneous structure of this kind of material and its complex configurations make very difficult all attempts to
correlate the structural properties with the electrical transport mechanism. The electrical properties of amorphous semi-
conductors films generally depend on the density of states (DOS), the position of the Fermi level and the temperature [16,17].
The tunneling transition of electronic carriers from occupied to unoccupied states involves one or more phonons so-
called «hopping stands». When both the number of phonons and the energy decrease, the hopping between states that are
closer in energy (even if they are widely spaced) becomes more predominant than that between the nearest neighbors whose
energies differ substantially. This mechanism is known as variable range hopping (VRH) conductivity, as previously reported
[16e18].
In the present work we focus on the study of the electronic transport mechanism and its dependence on the temperature,
in intrinsic and doped nc-Si:H films deposited by RFMS technique.

2. Experimental details

The intrinsic and n/p-doped hydrogenated nanocrystalline silicon films (nc-Si:H), were deposited by RFMS, under a total
pressure of 3 Pa, at low substrate temperatures TS ¼ 100  C and room temperature. The p and n type doping were achieved by
using boron-doped (100 ppm) and phosphorus-doped (50 ppm) crystalline silicon targets, respectively. For the intrinsic films,
we used undoped single crystalline silicon target of 3 inch in diameter and 99.9% purity. Two deposition times, respectively,
equal to 3 and 30 min are used. The RF-power of 220 W, the plasma gas mixture of 30% Ar þ 70% H2 and the target-sample
holder distance, fixed at 70 mm, were maintained constant for all the films, and are considered as optimal plasma conditions
that favor the growth of crystallized films at low substrate temperature [11e15]. These plasma conditions lead, as previously
reported [14,15], to film thicknesses and bonded hydrogen content, respectively, in the range of 380e470 Å and in the range of
12.5e16 at.%, for the films deposited for 3 min and, respectively, in the range of 3890e4960 Å and in the range of 7.5e11 at.%,
for those grown for 30 min. The electrical conductivity measurements were carried out in the planar configuration on quartz
substrates, as a function of temperature, in the range 273e450 K for the thermally activated process, and in the 93e273 K
range, for the hopping mechanism.

3. Results and discussion

3.1. Structural results

Fig. 1 shows the typical Raman spectra obtained in the transverse optic (TO)-like mode corresponding to the intrinsic and
doped films deposited at 100  C, for 3 and 30 min. This figure clearly shows that the samples deposited during 3 min present
two peaks: the first one is centred towards 480 cm1, and is attributed to the amorphous structure, the second one is centred
towards 520 cm1, and suggests the presence of the crystalline Si in the bulk of the films [11]. The structure of these samples is
composite by a mixture of amorphous phase and by crystallized regions. Although, the relative proportion of the crystallized
regions is more important in the p doped sample. For the films deposited during 30 min and at 100  C. All the spectra present
very intense asymmetric peaks centred between 516 and 520 cm1, indicating that these samples contain an important
crystalline fraction [19,20]. In our previous work [14] we shown that the substrate temperature does not seems to affect in a
significant way the Raman spectra for the films of nanocrystalline silicon intrinsic and doped deposited by RFMS technique.
These results are well confirmed by spectroscopic ellipsometric analysis for more details can consult reference [14].
More information about the structural changes and the average grain size of the crystallites in the bulk of the films can be
also gained from the TEM and HRTEM experiment. Fig. 2 (a, b) are typical bright and dark field TEM images obtained for films
deposited for 3 min at room temperature, which show the (100) silicon substrate and the film. From the silicon substrate we
can distinguish clearly two main regions: an amorphous layer and a nanocrystalline one, confirming the nanocrystalline
230 D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236

1600

Raman Intensity (u.a)

Raman Intensity (u.a)


n3min
i3min
n30min
1400 p3min i30min
p30min
1200

1000

800
Wavenumber (cm )

600

400

200

0
400 420 440 460 480 500 520 540 560
-1
Wavenumber (cm )
Fig. 1. Typical Raman spectra obtained in the TO-like mode for films deposited at 100  C during 30 min and 3 min of growth (inset of figure).

Fig. 2. TEM bright (a) and dark (b) field images for the intrinsic film deposited during 3 min.

character and the process of crystallization can be occur from 3 min of deposition. This important result is confirmed by the
HRTEM image presented in Fig. 3, when the average grain size obtained for these films is around 15 nm.

3.2. Electronic results

3.2.1. Thermally activated process


From the point of view of the electronic transport properties, the DOS distribution in nc-Si:H is similar to that of a-Si:H. It
consists in a pseudo-gap, usually defined by the energy separation between two « mobility edges» [16e18], in band tails due
essentially to the localized states in grain boundaries, and in a reduced deep-gap density within the pseudo gap. The results
obtained for the electrical conductivity s(T) for intrinsic and doped films deposited at 100  C, for 3 and 30 min are sum-
marized in Fig. 4, where ln(s(T)) is plotted as a function of the inverse of temperature (Ahrenius plot), in the range of
280e450 K. This figure clearly shows a linear behavior of the conductivity in the whole considered range, for all the films. We
also observed the same trend for the film deposited at room temperature, not shown in the figure. The conduction mechanism
in this range is well described by a thermally activated process following the expression:
 
Ea
sðTÞ ¼ s0 exp (1)
kB T

Where, Ea, KB and s0 denote, respectively, the associated activation energy, the Boltzmann constant and the preexponential
conductivity factor. The analysis of this linear regression, leads to the estimation of the activation energy, Ea, the room
D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236 231

Fig. 3. Typical HRTEM image obtained for the intrinsic film deposited during 3 min, which clearly shows nanocrystalline regions.

temperature conductivity sRT and the preexpontial factor s0. These results are summarized in Table 1. Moreover, Fig. 4 also
shows that, for each doping type, when the temperature is varied from 280 to 450 K, only a relatively small variation, less than
one order of magnitude, is observed for the conductivity. This variation of s(T) is due to the nanocrystalline structure of the
films, which correctly agrees with the relatively high values of the conductivity, obtained for all the films, mainly at room
temperature (sRT varies from 6.18  105 to 2.72  102 (Ucm)1), as shown in Table 1, in comparison with the values usually
observed for the amorphous silicon films. Furthermore, it also agrees with the low values obtained for the activation energy Ea
(0.11e0.2 eV) (Table 1).
On the other hand, the values of the activation energy do not seem to be affected by the doping type. However, the doping
seems to induce an increase in sRT as well as in s0 for all the doped films, compared to the intrinsic ones. It is also worth to
mention, that the values of Ea are systematically lower and around 0.12 eV for the films deposited for 3 min, than those
obtained for those deposited for 30 min, which are around 0.14e0.2 eV. In addition, the results presented in Table 1 indicate
that the increase in the crystalline fraction XC, deduced previously from the Raman experiments [14,15], is not accompanied
by a decrease in the activation energy. However, the increase in XC induces an increase in the values of sRT and s0, which are
systematically higher for the thick and less hydrogenated films, deposited for 30 min, compared to the thin and more hy-
drogenated ones, deposited for 3 min. Furthermore, Fig. 4 clearly shows that, for each doping type, the thick films exhibit
higher values of s(T), around two orders of magnitude, than the thin ones. These results agree with the fact that a highly
hydrogenated surface layer can facilitate crystallization to take place which, therefore, allows the films to have good electronic
properties.

0,1

n30m

0,01
-1

p30m
σ (Ω cm)

1E-3 i30m

n3m
i3m
1E-4 p3m

1E-5
2,5 2,6 2,7 2,8 2,9 3,0 3,1 3,2 3,3 3,4 3,5
-1
1000/T(K )
Fig. 4. Variation of lns(T) as a function of the inverse of T, in the temperature range 280e450 K, obtained for intrinsic and doped nc-Si:H films grown at 100  C.
232 D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236

Table 1
Summary of the deposition conditions and of the values of the different parameters deduced from the electrical conductivity measurement as the activation
energy, Ea, the room temperature conductivity, sRT, preexpontial factor s0 and of the crystalline volume fraction, XC, deduced from the analysis of the Raman
spectroscopy results [15].

Samples Type Deposition times (min) s0 (U cm)1 sRT (U cm)1 Ea (eV) XC (at.%)
3
n3min n-doped 3 7.23$10 6.18$105 0.12 35
i3min intrinsic 3 5.38$103 6.03$105 0.12 e
p3min p-doped 3 6.69$103 1.91$104 0.11 38
n30min n-doped 30 2.94 2.72$102 0.14 72
i30min intrinsic 30 1.22 1.45$103 0.18 74
p30min p-doped 30 1.50 8.34$104 0.20 73.5
p30mRT p-doped 30 1.50 8.34$104 0.20 71

RT: room temperature.

3.2.2. Hopping transport mechanism


Generally, at average temperatures, the conduction process is assumed to be due to the states localized in band tails, and
for low temperature the electronic transport is governed by states localized around the Fermi level. When one dopes a
semiconductor by donor or acceptor atoms, one induces a shift in the Fermi level position which is, consequently, accom-
panied by a change in the activation energy values. The results obtained for the electrical conductivity in the temperature
range 100  T  290 K are presented in Fig. 5, in terms of variation of ln(s(T)) as a function of the inverse of temperature. Fig. 5
does not show a linear behavior. This suggests that the conduction process in this temperature range is rather governed by
another mechanism than the thermally activated one discussed above for the range 280e450 K, and it is well described by the
relationship so-called «Mott's relation» [16,18]:
  g 
T0
sðTÞ ¼ s00 exp  (2)
T

where g (1/4  g  1/2) is a factor determining the dimension of electrical transport, and the parameters s00 et T0 of this
classical model are given by:

uph 16 a3
s00 ¼ e2 NðEF Þ ; T0 ¼ (3)
a2 hB NðEF Þ

where e is the electron charge, uph is the characteristic frequency of the photon, a is the parameter of wave, defined by a¼1/a0
(a0 is the Bohr radius), and N(EF) is the density of states localized around the Fermi level [21,22].
It is worth noticing that it has been suggested in previous works conducted on amorphous and nanostructured films
prepared in different conditions [23e25], that a localization parameter LP can be defined as LP ¼ N(EF)a3, in the range 105
LP  1, when the localized states have an exponential distribution with T1/4, and a linear dependence between lns00 and T0
1=4

was observed. Moreover, these results are confirmed by other previous ones obtained for intrinsic and p-doped microcrys-
talline silicon [26]. Also, using a correlation with classical percolation theory [27] one can deduce the parameter T*0 given by
the relation:

a3
T0* ¼ C0 (4)
kB NðEF Þ

whereC0 is a constant depending on the model. Two values are reported in the literature for C0, equal to 16 [27] and 310 [21].
The parameter 1/a typically varies from 0.3 to 3 nm.
Other parameters can be deduced from the hopping conduction model such as the range distance Rh and the energy range
hopping Wh characterizing the transport mechanism. These parameters are given by the following expressions:
 1=4  
9 3 T0 1=4
Rh ¼ ¼
8p a kB T NðEF Þ 8a T
 1=4 (5)
3
Wh ¼ ¼ k B T0 T 3
4pR3h NðEF Þ
Two other equivalent parameters are also obtained by using the percolation theory and are given by the following
relations:
 
A T0 1=4
R*h ¼
Ca T
(6)
B  1=4
Wh* ¼ 4 kB T0 T 3
C
D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236 233

i30m
p30m
1E-3 p30mRT

-1
σ (Ω cm)
1E-4

1E-5

1E-6
3 4 5 6 7 8 9 10
-1
1000/T(K )
Fig. 5. Variation of lns(T) as a function of the inverse of T, in the temperature range 100e280 K, obtained for some selected films deposited during 30 min at
100  C and room temperature.

Where C ¼ 2.06, A¼(9/8p) andB¼(128/9p) [28].


Let us now analyze our results obtained for the intrinsic and doped nc-Si:H films in the range 100e290 K, in the context of
the hopping conduction model. We present in Fig. 6 the variation of ln(s(T)) as a function of (T1/4), which clearly shows a
linear behavior in the whole range of T, in good agreement with the results previously obtained in the same range of tem-
perature for the electrical conductivity of different materials grown in different conditions [23e26]. The values of the
different parameters deduced from the hopping model and the percolation theory are summarized in Table 2.
From the results of Table 2 we can make the following remarks:

 For all the studied films, when N(EF) increases, the distance Rh and the hopping energy Wh decrease. This behavior can be
explained by the increase of the crystallites sizes, and by the reduction of the defect density in grain boundaries, inducing a
decrease of the distance and hopping energy between the crystallites.
 The values of the average distance Rh and hopping energy Wh determined for our films are in good agreement with those
previously obtained for films elaborated by other techniques [22,27].
 The values determined for the parameter LP(LP¼N(EF)a3) for our films, well agree with the range of validity of the
Variable range Hopping (VRH) model ((105 e 1) eV1), with the parameter a1 equal to 1 nm.

0,01
n30m

1E-3
-1
σ (Ω cm)

1E-4

1E-5 i30m

p30mRT
i3m
1E-6 p3m

0,23 0,24 0,25 0,26 0,27 0,28 0,29 0,30 0,31


-1/4 -1/4
T (K )
Fig. 6. Variation of lns(T) as a function of (T1/4), in the temperature range 100e290 K, obtained for films deposited at 100  C and room temperature.
234 D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236

Table 2
Summary of the different parameters determined for our films from the use of the VRH model as well as the percolation theory.

Samples Hopping Percolation theory Diffusion model

s00 (U cm1) T0 ( K) N(EF) LP (eV1) Rh (cm) at Wh (eV) at room Rh* (cm) at room Wh* (eV) at room
(eV1 cm3) room temperature temperature temperature temperature
n3min 8.29$102 2.24$107 8.29$1018 8.29$103 6.20$107 0.427 2.87$107 0,107
i3min 7.30$102 2.12$107 8.76$1018 8.76$103 6.11$107 0.421 2.83$107 0,106
p3min 3.38$102 1.30$107 1.43$1019 1.43$102 5.41$107 0.373 2.51$107 0,094
n30min 2.26 1.46$105 1.27$1021 1.27 1.76$107 0.121 8.17$108 0,030
i30min 2.96$107 9.98$107 1.86$1018 1.86$103 9.01$107 0.620 4.17$107 0,155
p30min 3.33$106 6.92$107 2.68$1018 2.86$103 8.22$107 0.566 3.81$107 0,142
p30mRT 4.81$107 1.16$108 1.60$1018 1.60$103 9.35$107 0.644 4.33$107 0,162

 The values deduced for s00 and T0, in the considered temperature range are quite well explained in the framework of the
VRH model, as it is clearly shown in Fig. 7, in which we present the variation of ln(s00) as a function of T1/4
0 , and which
indicates a linear behavior in the whole temperature range.

On the other hand, we present in Fig. 8 the variation of N(EF) as function of T1/4
0 by using the Mott constant C0 equal to 16,
according to relation (4). This figure shows that N(EF) increases when T0 decreases, as expected, to the limit of localized states
defined by N(EF)a31. It is also to mention that we obtain values for N(EF) of one order of magnitude higher when a value of
C0 equal to 310 is used.
The values of the parameters Rh and Wh, determined using either the percolation theory or the diffusion model are
presented as a function of temperature, respectively, in Fig. 9 (a,b). This figure shows that when the temperature increases, Rh
decreases progressively while, the hopping energy Wh it increases following a linear behavior, and indicates that Wh ~ kBT. It
also shows that the values of both Rh and Wh determined for the thin films deposited for 3 min are systematically lower than
those obtained for the thick ones deposited for 30 min. The thin films exhibit, indeed, values for the activation energy lower
than those for the thick ones, as presented above in Table 1. This behavior might be explained by the presence of oxygen atoms
in the films, as it was observed in the infrared spectra, reported in previous work [15], and which seem to play an important
role in the electrical conduction mechanism.
As a final remark, if we consider the results obtained for the electrical conductivity, summarized in Table 1, one can say that
the behavior, observed for the activation energy and electrical conductivity in the temperature range 273e450 K, cannot be
explained only by the change in the crystalline fraction, determined from the Raman spectroscopy measurements [15]. Other
phenomena can also play an important role in the conduction process, such as the presence of defects at grain boundaries and
oxygen. Indeed, several authors have previously suggested [27e31] that the oxygen incorporated in the material plays the role
of donor and can induce a decrease in the activation energy associated with the electrical conductivity. This might explain the
results obtained for the films deposited during 3 min which exhibit a high fraction of oxygen and a reduced crystalline
fraction [14,15], compared to the films deposited during 30 min. However, we cannot exclude that the crystalline fraction and

8
10
7
10
6
10
-1
σ00(Ω cm)

5
10
4
10
3
10
2
10
1
10
0
10
20 40 60 80 100 120
1/4 1/4
T0 (K )

Fig. 7. Variation of s00 as function of T1/4


0 . The dashed line is a guide for the eye.
D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236 235

10

N(EF) (eV cm )
-3
10

-1 10

10

20 40 60 80 100 120
1/4 1/4
T0 (K )

Fig. 8. Variation, as a function of (T1/4), of the estimated N(EF), obtained using the Mott's constant C0 ¼ 16. The dashed line is a guide for the eye.

1,3x10
0,65 n3m
a n3m
i3m i3m b
1,2x10 p3m 0,60 p3m
i30m i30m
p30m 0,55 p30m
1,1x10 n30m
n30m
0,50
1,0x10
0,45
Wh (eV)
Rh(cm)

9,0x10 0,40
0,35
8,0x10
0,30
7,0x10
0,25

6,0x10 0,20
0,15
5,0x10
100 150 200 250 300 100 150 200 250 300
T (K) T (K)

Fig. 9. Variation, as a function of T, of the average hopping distance Rh (a) and hopping energy (b) deduced using the VRH model, for selected films deposited at
100  C, during 3 min (open symbols) and 30 min (closed symbols). The dashed lines are guides for the eye.

the size of crystallites may play a role in the sRT changes. Indeed, an increase in sRT is observed with the increase in the
crystalline fraction, independently of the doping type. Finally, it is to be mentioned that the deposition temperature seems to
influence predominantly the hydrogen content incorporated in the films [14,15].

4. Conclusion

The activation energy and the room temperature conductivity values, deduced from the electrical conductivity mea-
surements, indicate that our intrinsic and doped thin films deposited by RFMS technique at substrate temperatures, as low as
100  C or room temperature, present quite good electronic properties. The lower activation energy values obtained for the
thin films grown during 3 min may be principally due to the presence of defects in grain boundaries as well as to oxygen
atoms which might be incorporated in the these films. In the low temperature range (100  T  290 K), the electrical con-
ductivity results are better explained in the framework of the variable range hopping (VRH) model. The different parameters
such as the average hopping distance and hopping energy, determined by using this model, show a strong variation with
temperature and are in good agreement with those reported in the literature for films grown by other deposition techniques.

References

[1] A. Matsuda, Thin Solid Films 337 (1999) 1.


[2] S. Klein, F. Finger, R. Carius, M. Stutzmann, J. Appl. Phys. 98 (2005) 24905.
[3] A.V. Shah, J. Meier, E. Vallat-Sauvan, N. Wyrsch, U. Kroll, C. Droz, U. Graf, Sol. Energy Mater. Sol. Cells 78 (2003) 469.
[4] B.T. Li Hongbo, R.H. Franken, J.K. Rath, R.E.I. Schropp, Sol. Energy Mater. Sol. Cells 93 (2009) 338.
[5] J. Sancho-Parramon, D. Gracin, M. Modreanu, A. Gajovic, Sol. Energy Mater. Sol. Cells 93 (2009) 1768.
[6] J. Meier, R. Fluckiger, H. Keppner, A.V. Shah, Appl. Phys. Lett. 65 (1994) 860.
236 D. Benlakehal et al. / Superlattices and Microstructures 100 (2016) 228e236

[7] S. Hamma, P.Roca i Cabarrocas, J. Non-Cryst. Solids 227e230 (1998) 852.


[8] T. Sugano, T. Kitagawa, Y. Sobajima, T. Toyama, H. Okamoto, J. Appl. Phys. 97 (2005) 94910.
[9] J.E. Gerbi, J.R. Abelson, J. Appl. Phys. 89 (2001) 1463.
[10] Y. Leconte, C. Dufour, B. Garrido, R. Rizk, J. Non-Cryst. Solids 299 (2002) 87.
[11] D. Senouci, R. Baghdad, A. Belfedal, L. Chahed, X. Portier, S. Charvet, K.H. Kim, P. Roca i Cabarrocas, K. Zellama, Thin Solid Films 522 (2012) 186.
[12] Y. Bouizem, K. Kefif, J.D. Sib, D. Benlakehal, A. Kebab, A. Belfedal, L. Chahed, J. Non-Cryst. Solids 358 (2012) 854.
[13] Y. Bouizem, C. Abbes, K. Kefif, J.D. Sib, D. Benlakehal, A. Kebab, L. Chahed, Thin Solid Films 545 (2013) 245.
[14] A. Belfedal, Y. Bouizem, D. Benlakehal, R. Baghdad, D.J. Sib, A. Kebab, L. Chahed, K. Zellama, Phys. Status Solidi C 7 (3e4) (2010) 565.
[15] A. Belfedal, D. Benlakehal, Y. Bouizem, R. Baghdad, M. Clin, A. Zeinert, O. Durand-Drouhin, J.D. Sib, L. Chahed, K. Zellama, Mater. Sci. Semicond. Process.
26 (2014) 231.
[16] N.F. Mott, R.A. Davis, Electronic Processes in Non-Crystalline Materials, second ed., Oxford University, Oxford, 1979.
[17] P.W. Anderson, Phys. Rev. 109 (1958) 1492.
[18] N.F. Mott, Philos. Mag. 19 (1969) 835.
[19] H.S. Mavi, A.K. Shukla, S.C. Abbi, K.P. Jain, J. Appl. Phys. 66 (1989) 5322.
[20] C. Goncalves, S. Chervet, A. Zeinert, M. Clin, K. Zellama, Thin Solid Films 91 (2002) 403.
[21] S.Y. Myong, K.S. Lim, M. Konagai, Appl. Phys. Lett. 88 (2006) 103120.
[22] A. Dussan, R.H. Buitrago, J. Appl. Phys. 97 (2005) 043711.
[23] C. Godet, J. Non-Cryst. Solids 299e302 (2002) 333.
[24] G. Lazar, M. Clin, S. Charvet, M. Therasse, C. Godet, K. Zellama, Diam. Relat. Mat. 12 (2003) 201.
[25] C. Godet, Philos. Mag. B 81 (2001) 205.
[26] S.B. Concari, R.H. Buitrago, J. Non-Cryst. Solids 338e340 (2004) 331.
[27] J. Meier, Mat. Res. Soc. Symp.Proc. 420 (1996) 3.
[28] M. Thamilselvan, K. Premnazeer, D. Mangalarj, S.K. Narayandass, Phys. B 337 (2003) 404.
[29] E. Simanek, Sol. State Commun. 40 (1981) 1021.
[30] P. Sheng, F. Abeles, Phys. Rev. Lett. 31 (1973) 44.
[31] M.J. Williams, C. Wang, G. Luckovsky, J. Non-Cryst. Solids 137e138 (1991) 737.

You might also like