Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

Chemical Physics Letters 423 (2006) 225–228

www.elsevier.com/locate/cplett

Electronic transport in Si–SiO2 nanocomposite films


a,*
M.L. Ciurea , V.S. Teodorescu a, V. Iancu b, I. Balberg c

a
National Institute of Materials Physics, P.O. Box MG-7, Bucharest-Magurele 77125, Romania
b
Department of Physics, University ‘Politehnica’ of Bucharest, Bucharest 060042, Romania
c
The Racah Institute of Physics, Hebrew University, Jerusalem 91904, Israel

Received 13 March 2006


Available online 5 April 2006

Abstract

We report experimental investigations and modeling of the electronic transport in Si–SiO2 nanocomposite films. The current–voltage
characteristics measured at room temperature are interpreted as due to high field-assisted tunneling. The activation energies from the
current–temperature curves are given by the energy separations between quantum confinement electronic states, determined from a quan-
tum well model. Consequently, the calculated mean diameter of a nanodot (5.2 nm) is in good agreement with the microstructure data
(5 nm). Also, the potential barrier between nanocrystalline Si and amorphous SiO2, previously obtained for nanocrystalline oxidized por-
ous Si (2.2 eV), is confirmed.
 2006 Elsevier B.V. All rights reserved.

Semiconducting nanomaterials and nanostructures have films, taking into account the potential barrier between nc-
an increasing number of applications in biomedical devices, Si and amorphous silicon dioxide (a-SiO2).
sensors, optoelectronics, and ultra-large scale integrated The Si–SiO2 films were prepared by co-sputtering of Si
electronics. The understanding of the specific processes that and SiO2 on an elongated quartz substrate, followed by
take place at this scale represents a major challenge. In the annealing in N2 flow at 1100 C to form Si crystallites
addition, the behavior of the indirect gap semiconductors [12,13]. This method offers the advantage of an almost lin-
raises several problems [1–5]. Thus, several conduction ear variation of the volume content of the nc-Si phase (x)
mechanisms were proposed to describe the various aspects from one end of the substrate (x  0%) to the other one
of the electronic transport in nanocrystalline silicon (nc-Si) (x  100%). The films thickness was in the 5–9 lm range.
and silicon-based nanostructures. The Ohmic conduction For the transport measurements, 50 coplanar Al electrodes
and surface states contributions were discussed for silicon were deposited at 1 mm distance from each other. In this
nanowires [6]. Fowler–Nordheim tunneling was used to Letter, we present investigations performed on samples in
describe carrier injection from the substrate of a nanocrys- the x = 65–75% range.
talline floating-gate MOS transistor [7], and the Schottky The films microstructure was investigated with a JEOL
tunneling was investigated for ultra-small diodes [8]. Other 200CX microscope for transmission electron microscopy
proposed mechanisms are the Poole–Frenkel tunneling in (TEM) and a TOPCON 002B instrument for the high res-
nanocrystalline porous silicon (nc-PS) [9] and nc-Si/CaF2 olution measurements (HRTEM). The current–voltage (I–
multilayer structures [10], and fractal percolation for nc- V) and current–temperature (I–T) characteristics were
PS [11]. taken using a Keithley 642 electrometer and an Agilent
The purpose of the present work is to investigate the E3631A d.c. power supply.
electronic transport mechanism in Si–SiO2 nanocomposite The microstructure measurements show the segregation
of the nc-Si and a-SiO2 phases at nanometric scale. This
segregation decreases with the increase of the nc-Si concen-
*
Corresponding author. Fax: +40 21 493 0267. tration. Fig. 1 presents a typical non-uniform distribution
E-mail address: ciurea@infim.ro (M.L. Ciurea). of nc-Si in a-SiO2 matrix at x  66%. Differently oriented

0009-2614/$ - see front matter  2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.cplett.2006.03.070
226 M.L. Ciurea et al. / Chemical Physics Letters 423 (2006) 225–228

Fig. 1. HRTEM image from x  66% region.

Fig. 3. I–T characteristics for x  66% region, measured at 4, 5 and 25 V.


silicon crystallites of practically the same size (around
5 nm) and quasi-spherical shape form a network with chain
mental point is farther from the fitting line than its exper-
morphology, while the width of the amorphous regions
imental error.
does not exceed 5 nm. The apparent nc-Si phase concentra-
The electronic transport presents two aspects, the trans-
tion is smaller than the real one, as not all the crystallites
port inside and between the nanodots. Inside the nanodots,
are properly oriented to exhibit (1 1 1) lattice fringes con-
the Coulomb blockade (CB) effect appears. For the mean
trast in the HRTEM image.
nanodot diameter of 5 nm, the CB energy is ECB = e2/
A typical I–V characteristic, taken at room temperature
(2pe0erd) = 46 meV. At RT this means 1.8 kBT, so that no
(RT) at x  66% is shown in Fig. 2. The curve is superlin-
shape modification of the I–V curves could be observed
ear and practically symmetric. Three I–T curves, measured
between liquid nitrogen temperature (LNT) and RT. Inside
on the same sample, are presented in Fig. 3. At low bias (4
the nanodots the carriers will perform ballistic movements,
and 5 V) one finds three activation energies (E1 = 0.22 ±
due to the absence of impurities and/or point defects [14].
0.02 eV, E2 = 0.32 ± 0.02 eV, and E3 = 0.44 ± 0.02 eV),
The transport between the nanodots is mainly performed
while at high bias (25 V) only the values (0.33 ± 0.02 eV
by tunneling under high field (eU  kBT) through a perco-
and 0.44 ± 0.02 eV, practically equal with E2 and E3,
lation network (see [15]). From Fig. 1 it results that the car-
respectively) can be observed. The activation energies were
riers move preferentially along nc-Si chains and tunnel
determined under two restrictions: (i) the linear correlation
through a number of barriers. Then, the high field-assisted
coefficient must be greater than 0.9995 and (ii) no experi-
tunneling (HFAT) current is well described by the relation
[16]
 pffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
I ¼ a½u exp dv u  ðu þ qU b Þ expðdv u þ qU b Þ;
ð1Þ
where u and d are the mean height and width of the poten-
tial barrier between nanodots, jqj = e (the elementary
charge) and v = (8m*/h2)1/2 (m* being the effective carrier
mass), while a is a constant proportional to the number
of equivalent paths and the ionic transport through
a-SiO2 was neglected. If one considers the tunneled barriers
as roughly equal, then Ub = U/N, where U is the applied
bias and N the mean number of barriers. For superlinear
curves, the relation (1) can be rewritten in the form
h pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi i
I ¼ I 0 signðU Þ ð1  jU j=U 0 Þ expða 1  jU j=U 0 Þ  expðaÞ ;
ð2Þ
1/2
with I0 = jaju, U0 = Nu/e, a = dvu and q = e. From
the experimental results one cannot determine all the four
Fig. 2. I–V characteristic for x  66% region. The solid line represents the theoretical parameters (a, N, d and u), but only the fit
theoretical fit. parameters I0, U0 and a.
M.L. Ciurea et al. / Chemical Physics Letters 423 (2006) 225–228 227

The solid line in Fig. 2 represents the fit of the experi- mated from Eq. (3). On the other hand, Eq. (3) can be
mental characteristic to Eq. (2). The fit parameters are applied to any kind of nanomaterials.
I0 = (3.383 ± 0.003) A, U0 = (191.2 ± 0.2) V and a = The energy differences between consecutive levels for
14.74 ± 0.01, with a correlation coefficient q = 0.9996. d  5 nm are an order of magnitude greater than both
For a potential barrier height u = (2.2 ± 0.2) eV (previ- ECB and kBT, even at RT. Consequently, at a given T,
ously obtained for oxidized nc-PS) [17], one finds the carrier concentration will be given by the Boltzmann
N = 87 ± 8 and d = (1.0 ± 0.1) nm. Then, the CB effect law, n  exp(Ea/kBT), where the activation energy Ea is
should appear for jUj6UCB = N(ECB-kBT)/e. The CB the absolute value of the energy difference between the last
threshold, UCB, is equal to 1.75 V at RT and 3.45 V at occupied level and the following one. The drift velocity is
LNT, respectively. At the same time, eUmax/Nu  1/6, jus- constant (ballistic transport inside the nanodots), so that
tifying the choice of the HFAT mechanism (corresponding the I–T curves present an Arrhenius-like behavior. When
to a trapezoidal-like barrier) instead of the other tunneling a level is practically filled, a following one starts to be
mechanisms (corresponding to triangular-like ones). As the excited and, at this ‘jump’ temperature, the activation
carrier path is along the chains and the width of a barrier energy is modified. Therefore, the ratio of the two consec-
between neighboring crystallites in a chain is an inter- utive activation energies is
atomic distance (twinned crystallites), the mean width of 2 2
a barrier is much smaller than the distance between the E00a el00 ;p00  el0 ;p0 xl00 ;p00 þ1  xl0 ;p0 þ1
RE  ¼ ¼ . ð4Þ
chains. The number of tunneled barriers does not give E0a el0 ;p0  el;p x2l0 ;p0 þ1  x2l;pþ1
the total electron path through the sample (see Fig. 1). This
means that the rest of the path is covered through less resis- This ratio depends on the excitation conditions. For ther-
tive mechanisms (e.g. the resistance of a full crystallite mal excitation (eU  kBT), the first three levels correspond
chain is much smaller than the equivalent resistance of a to l = 0, 1, 2 and p = 0, so that RE = 1.26. If the excitation
single mean width barrier and the ionic transport short cir- is made under a high field, the first three levels correspond
cuits most of the barriers), which do not influence the I–V to l = 0 (angular momentum conservation) and p = 0, 1, 2,
characteristic in a significant manner. Similar investiga- with RE = 1.67. If the field is applied at a temperature
tions made on different samples in the x = 65–75% region where the (1, 0) level is thermally excited, RE = 1.50, and
showed that d varies with less than 10% and N with more so on.
than 50%, in a fluctuating manner (due to the randomness As mentioned before, three activation energies are
of the nc-Si space distribution). observed in the I–T characteristics measured at 4 and
The nanodot surface acts as the wall of a quantum well 5 V. Their ratios are RE ” E2/E1 = 1.45 and R0E 
that generates the quantum confinement (QC) energy lev- E3 =E2 ¼ 1:37. For E1 = e1,1  e1,0, E2 = e1,2  e1,1 and
els. These levels are located in the band gap because the E3 = e1,3  e1,2, the differences between the theoretical
fundamental QC level coincides with the highest occupied and the experimental ratios are of the order of 3%. From
level at absolute zero (top of the valence band). Different Eq. (3) we derive d = 5.2 ± 0.4 nm, in agreement with the
types of potentials (rectangular, parabolic, Woods-Saxon microstructure results. At 25 V, only the last two activation
etc.) were used to describe the QC [17–19]. If the nanodot energies (E2 and E3) appear. The disappearance of E1
is small enough, a spherical shape and an infinite rectangu- means that this bias is high enough to excite the (1, 1) level
lar quantum well adequately describe the QC (less than 5% from the (1, 0) one, which should be expected at this tem-
differences for the first 3–4 levels, in comparison with a 2– perature range. The energy shift induced by the voltage
5 eV deep rectangular quantum well, for sizes between 1 increase from 5 to 25 V can be evaluated on the basis of
and 10 nm) [17]. Then, the QC energy levels are given by the I–V fit. For N = 87, a bias variation of 20 V gives an
the expression average bias per barrier of (0.23 ± 0.02) V, almost equal
to E1/e. This does not only suggest that the nc-Si/a-SiO2
2h2  
el;p ¼  2 x2l;pþ1  x20;1 . ð3Þ barrier height is practically the same for oxidized nc-PS
md and Si–SiO2 nanocomposite films, but also proves the
Here d is the nanodot diameter and xl,p the p-th zero of the self-consistency of the different results presented above.
spherical Bessel function jl(x) for the angular quantum On different samples, where the average bias per barrier
number l (by convention, e0,0 ” 0). is significantly smaller than E1/e, the first activation energy
The expression (3) is correct only in the effective mass value appears at all applied biases.
approximation (EMA) limits. The LCAO method [20] In conclusion, the analysis of the electronic transport
proved that the diameter dependence of the band gap has proves that: (i) the I–T characteristics depend on the
the form Eg  da, with a = 1.39 for spherical nanodots, transitions between the electronic states and (ii) between
instead of a = 2 as obtained from EMA. Exciton measure- nanodots the dominant mechanism observed in the I–V
ments [5] suggest an even lower exponent, a = 0.6–0.8. characteristics is the HFAT. The electron energies and
However, recent investigations [21] proved that EMA can the mean nanodot diameter of 5.2 nm, evaluated by
be applied if one introduces a size dependence of the effec- means of a simple quantum well model, are in good
tive mass. Therefore the nanodot diameter can be esti- agreement with the activation energies found in the I–T
228 M.L. Ciurea et al. / Chemical Physics Letters 423 (2006) 225–228

characteristics and the microstructure data. The potential [6] I. Ionica, L. Montes, S. Ferraton, J. Zimmermann, V. Bouchiat, L.
barrier between nc-Si and a-SiO2, previously obtained for Saminadayar, in: Techn. Proceedings of 2004 NSTI Nanotech. Conf.
& Trade Show, Boston, vol. 3, 2004, p. 165 (Chapter 4).
oxidized nc-PS, is also confirmed. [7] R.J. Walters, G.I. Bourianoff, H.A. Atwater, Nat. Mater. 4 (2005)
143.
Acknowledgements [8] G.D.J. Smit, S. Rogge, T.M. Klapwijk, Appl. Phys. Lett. 81 (2002)
3852.
This work was supported in part by CERES Contract [9] M. Ben-Chorin, F. Möller, F. Koch, Phys. Rev. B 49 (1994) 2981.
[10] V. Ioannou-Sougleridis, T. Ouisse, A.G. Nassiopoulou, F. Bassani, F.
No. 9/2001 (Romanian National RD Plan), in part by Is- Arnaud d’Avitaya, J. Appl. Phys. 89 (2001) 610.
rael Science Foundation (ISF) and in part by Intel [11] M. Ben-Chorin, F. Möller, F. Koch, Phys. Rev. B 51 (1995) 2199.
Corporation. [12] M. Dovrat, Y. Oppenheim, J. Jedrzejewski, I. Balberg, A. Sa’ar, Phys.
Rev. B 69 (2004) 155311.
[13] V. Iancu, M. Draghici, L. Jdira, M.L. Ciurea, J. Optoelectron. Adv.
References Mater. 6 (2004) 53.
[14] T.V. Torchynska, J. Appl. Phys. 92 (2002) 4019.
[1] A.G. Cullis, L.T. Canham, P.D.J. Calcott, J. Appl. Phys. 82 (1997) [15] D. Toker, D. Azulay, N. Shimoni, I. Balberg, O. Millo, Phys. Rev. B
909. 68 (2003) 041403(R).
[2] L. Tsybetskov, K.D. Hirschman, S.P. Duttagupta, M. Zacharias, [16] M.R. Reshotko, A. Sa’ar, I. Balberg, Phys. Status Solidi A 197 (2003)
P.M. Fauchet, G.P. McCaffrey, D.J. Lockwood, Appl. Phys. Lett. 72 113.
(1998) 43. [17] M.L. Ciurea, I. Baltog, M. Lazar, V. Iancu, S. Lazanu, E. Pentia,
[3] D. Kovalev, H. Heckler, G. Polisski, F. Koch, Phys. Status Solidi B Thin Solid Films 325 (1998) 271.
215 (1999) 871. [18] W.P. Yuen, Phys. Rev. B 48 (1993) 17316.
[4] L. Pavesi, L. Dal Negro, C. Mazzoleni, G. Franzò, F. Priolo, Nature [19] K. Clemenger, Phys. Rev. B 44 (1991) 12991.
408 (2000) 440. [20] C. Delerue, G. Allan, M. Lannoo, Phys. Rev. B 48 (1993) 11024.
[5] J. Heitmann, F. Müller, L.X. Yi, M. Zacharias, D. Kovalev, F. [21] J. Wang, A. Rahman, A. Ghosh, G. Klimeck, M. Lundstrom, IEEE
Eichhorn, Phys. Rev. B 69 (2004) 195309. Trans. Electron. Dev. 52 (2005) 1589.

You might also like