You are on page 1of 5

Current-transport studies and trap

extraction of hydrothermally grown ZnO


nanotubes using gold Schottky diode

solidi

status

pss

physica

Phys. Status Solidi A 207, No. 3, 748752 (2010) / DOI 10.1002/pssa.200925547

www.pss-a.com

applications and materials science

G. Amin*, I. Hussain, S. Zaman, N. Bano, O. Nur, and M. Willander


Department of Science and Technology, Campus Norrkoping, Linkoping University, 60174 Norrkoping, Sweden
Received 26 October 2009, revised 10 December 2009, accepted 15 December 2009
Published online 26 January 2010
PACS 61.46.Fg, 61.46.Km, 68.37.Hk, 84.37.q, 85.30.Hi
* Corresponding

author: e-mail gulam@itn.liu.se, Phone: 46 11363016, Fax: 4611363270

High-quality zinc oxide (ZnO) nanotubes (NTs) were grown by


the hydrothermal technique on n-Si substrate. The room
temperature (RT) current-transport mechanisms of Au Schottky
diodes fabricated from ZnO NTs and nanorods (NRs) reference
samples have been studied and compared. The tunneling
mechanisms via deep-level states was found to be the main
conduction process at low applied voltage but at the trap-filled
limit voltage (VTFL) all traps were filled and the space-charge-

limited current conduction was the dominating currenttransport mechanism. The deep-level trap energy and the trap
concentration for the NTs were obtained as 0.27 eV and
2.1  1016 cm3, respectively. The same parameters were also
extracted for the ZnO NRs. The deep-level states observed
crossponds to zinc interstitials (Zni), which are responsible for
the violet emission.

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1 Introduction Zinc oxide (ZnO) has a large direct


bandgap (3.34 eV) and a high exciton binding energy
(60 meV) that can provide more efficient excitonic emission
even above room temperature (RT). ZnO, with a number
of advantages compared to other wide-bandgap semiconductors including high thermal/chemical stabilities and
the possibility of wet chemical etching, have led to the
demonstration of ZnO as an alternative material to nitride
semiconductors [1, 2]. ZnO has a wide range of applications
in opto-electronics, transparent electronics, surface acoustic
waves, piezoelectronic transducers, spintronics, and chemical gas sensors [25].
Zinc oxide has a rich family of nanostructures such as
nanorods (NRs), nanorings, nanobelts, nanoparticles, nanotips, and nanotubes (NTs) [2, 57]. On the other hand ZnO
NTs have attracted considerable attention due to the link
between the tubular structure and high porosity and surface
area to volume ratio that can lead to attractive performance
characteristics [8]. Although nanotubular structures of ZnO
have become a popular topic, ZnO NTs have not yet been
exploited for electrical/electronic devices [911]. Besides
the above-mentioned properties of ZnO, it also possesses a

large number of radiative intrinsic and extrinsic defects that


emit light covering the whole range [12]. The dominant
intrinsic defects in ZnO, e.g., oxygen vacancies (Vo), zinc
vacancies (VZn), zinc interstitials (Zni), oxygen interstitials
(Oi), zinc antisites (Zno), and extrinsic impurities like Cu, Li,
Fe, and their existence in different samples depend on the
growth method used [12]. The free carrier concentration,
doping compensation, minority carrier lifetime, and luminescence efficiency are directly or indirectly related to the
defects [12]. The causes of the formation of these defects
and spatial distribution are still controversial. ZnO-based
electrical devices depend upon the defect chemistry and the
understanding of the electrical properties, both of which are
the subject of current theoretical and experimental studies.
The electrical transport mechanism in ZnO NTs has not yet
been studied and to the best of our knowledge no results have
been reported for the electrical transport of ZnO/n-Si NTs. In
this study, we report the investigation of current-transport
mechanisms in ZnO/n-Si NTs. Moreover, deep-level traps
were extracted from the currentvoltage (IV) measurements of ZnO NRs and NTs and these traps were correlated
to Zni.
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Original
Paper
Phys. Status Solidi A 207, No. 3 (2010)

2 Experiment The process to obtain ZnO NTs is a


two step process. First we grow the ZnO NRs on n-type
Si substrate using a low-temperature chemical growth
technique. In this method 0.075 M zinc nitrate hexahydrate
(Zn (NO3)6H2O) were mixed with 0.075 M hexamethyl
tetramine (HMT) (C6H12N4). The sample was placed
in the solution and was heated at 96 8C for 5 h. After that,
it was washed with deionized water and dried in air.
In the second step we converted the ZnO NRs into
NTs by immersing the NRs in KCl solution of a
concentration in the range from 3.0 to 3.4 M for time periods
ranging from 6 to 13 h. The temperature of the solution
was varied from 80 to 90 8C [7]. The immersion time was
optimized by taking into account the NRs dimensions
in 3.4 M KCl solution at 85 8C. After growth the top
and bottom metal contacts were performed to get the final
device. For the ohmic contact on n-type Si we first etch a
small portion of the n-type Si, which we cover before the
growth of ZnO NTs. After that a thin layer of Ag was used
as ohmic contact to the substrate. Prior to the evaporation
of the Schottky contact on the ZnO NTs an insulating PMMA
layer was deposited between the NTs. In our case, we
spin coated the sample with PMMA at a spin speed of
2000 rpm for 30 s. After spinning, the samples were baked
for 2 min at 90 8C to increase the adhesion. To remove the
additional layer of PMMA oxygen plasma reactive ion
etching is used with an oxygen flow rate of 100 sccm, 250 Pa
of pressure for 30 s. To ensure that no PMMA was on the top
of NTs a scanning electron microscope (SEM) was used to
investigate the top of the ZnO NTs prior to the contact metal
deposition. If residues were still observed, the reactive ion
etching step is repeated. It is important to mention that this
oxygen plasma step will serve two purposes: the first is to
remove all PMMA from the top of the ZnO NTs and NRs, and
the second is that it acts as a surface pretreatment step. Then,
Au circular contacts of 1 mm diameter and a thickness
of 120 nm were evaporated onto a group of NTs. After
completion of the device processing, the diodes were tested
for rectification. Good rectifying properties were oberved
from all diodes. Previously, it was reproted that Au contacts
on ZnO thin films were ohmic, and Schottky rectifying
contacts can only be acheived by hydrogen peroxide
pretreatment [13], or by a 1 h oxygen plasma treatment of
the fabrciated Au/n-ZnO thin film [14]. Nevertheless, we
here show that a short oxygen plasma pretreatment for ZnO
NTs and NRs will also lead to a Schottky rectifying contact.
The configuration of the device is illustrated schematically in
Fig. 1.
3 Results and discussion The device structure was
characterized by SEM and IV. The ZnO NTs were grown
vertically aligned as shown in the SEM image in Fig. 2. As
shown in Fig. 2 the yield was almost 100% and all NRs were
converted to NTs.
Figure 3 shows the resulting Au/ZnO NTs heterojunction
exhibiting good rectifying IV characteristics. Such rectification behavior is best described by the thermionic emission
www.pss-a.com

749

Figure 1 Schematic illustration of Au/ZnO NTs Schottky diode.

theory. According to this theory, the current in such a device


could be expressed as:
 


qV  IRs
1 ;
(1)
I Is exp
nkT
where Is is the saturation current, Rs is the series resistance, k
is the Boltzmann constant, T is the absolute temperature, q is
the elementary electric charge, V is the applied voltage, and
n is the ideality factor. The saturation current Is is given as:


qFb
 2
Is AA T exp 
;
(2)
kT
where A is the active device area, A is the effective
Richardson constant, and Fb is the barrier height.
The ideality factor from Eq. (1) was found to be in the
range 23 for the Schottky diodes investigated. The barrier
height of the ZnO NTs obtained from Eq. (2) is determined to
be 0.91  0.02 eV. The higher value of the ideality factor
indicates that the transport mechanism is no longer
dominated by thermionic emission. Non-ideal behavior is
often attributed to defect states in the bandgap of the
semiconductor providing other current-transport mechanisms such as structural defects, surface contamination,
barrier tunneling, or generation recombination in the space

Figure 2 SEM image of ZnO NTs grown on n-Si substrate.


2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

750

solidi

status

physica

pss

G. Amin et al.: Current transport studies and trap extraction of hydrothermally grown ZnO nanotubes

Figure 3 Typical currentvoltage characteristics of Au/ZnO NTs


Schottky diode.

charge region and to variations in interface composition [15,


16]. To understand which mechanism dominates the junction
behavior, the IV characteristics of the device are studied on
a loglog scale.
The loglog plot of the RT IV data is shown in Fig. 4 and
it illustrates that the current-transport mechanism exhibits
three different regions. The current in region I follows a
linear dependence, i.e., I  V. This indicates the current
transport is dominated by tunneling at low voltages. The
boundary for this region was determined to be below
0.034 V. In region II (0.043.4 V) the current increases
exponentially, I  exp(cV). The ideality factor (23) is
determined in this region and the dominating transport
mechanism is recombination tunneling. Finally, above 3.4 V
the current follows a power law (I  V2), indicating a spacecharge limited current (SCLC) transport mechanism. The
SCLC and the other regions observed in the present study
have been also reported in different n-ZnO NRs/p-Si

heterojunctions [9, 10, 17] and in Schottky contacts to ZnO


NRs [11].
Lampert and Mark [18] have developed a single-carrier
SCLC model with the presence of a trap above the Fermi
level. As the applied voltage V is lower than the turn on
voltage (Von) (V < Von) the thermally generated carrier
density n0 dominates over the injected carrier density and
the carrier transit time tx d2/mVon is larger than the
dielectric or ohmic relaxation time tV ee0/qn0m. The
injected carrier would thus undergo dielectric relaxation to
maintain the charge neutrality rather than transport across the
sample. In this region, the trap levels are not completely
filled. For V > Von the carrier transit time is smaller than the
dielectric relaxation time (tx < tV) and the injected carriers
dominate over the thermally generated carriers. The increase
of the applied voltage also shifts the quasi-Fermi level
towards the conduction band and the effect would be the
filling up of the traps at an energy level Ec  Et [9, 18]. At an
applied voltage of V > trap-filled limit voltage (VTFL) (TFL
stands for trap-filled limit) all the traps are filled and the
conduction would become space-charge limited and the
current follow according to the Mott and Gurney SCLC
expression [19] given by:


9ee0 mAV 2
;
(3)
J
8d3
where J is the current density, m is the field-independent
carrier mobility, d is the thickness of active layer, A is the
area, e is the dielectric constant, and e0 is the permittivity of
free space.
The VTFL is given as [20]:
VTFL

Nt qd2
;
2ee0

(4)

where Nt is the concentration of the unoccupied states (trap


concentration), located approximately at the estimated
effective Fermi level. The effective carrier concentration n0
in the active region is given by the expression:
J2VTFL Nt
 ;
JVTFL
n0

(5)

where J(VTFL) is the current density at VTFL and J(2VTFL)


the current density at a voltage twice the VTFL. From the
calculated value of n0, the position of the effective Fermi
level (quasi-Fermi level) can be estimated from the
expression:
n0 Nc eEF Ec =kT :

Figure 4 Loglog plot for the IV data of Au/ZnO NTs Schottky


diode.
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

(6)

The position of the deep-level state is taken as being


within an energy band of a few kT from the effective Fermi
level.
The slow rise in current at low voltages is an indication of
a distribution of states in the energy gap. When all these
states are filled with carriers the current rises rapidly to
the space charge limit value. Deep-level parameters were
www.pss-a.com

Original
Paper
Phys. Status Solidi A 207, No. 3 (2010)

calculated from the experimental VTFL at RT. In these


calculations the depletion region thickness at zero bias
capacitance was used as the active layer thickness. The
values of n0, Nt, and the location of the deep-level states
(traps) below the conduction band for ZnO NTs were
obtained to be 3.3  1015 cm3, 2.1  1016 cm3, and
0.27  0.02 eV, respectively. Similarly for ZnO NRs the
corresponding values of n0, Nt, and the location of the deeplevel states (traps) were obtained to be 2.9  1017 cm3,
8.1  1018 cm3, and 0.20  0.02 eV, respectively. The
deep-level states observed in ZnO NTs and NRs are in
approximate agreement with the values obtained from the
full potential linear muffin-tin orbital method, and it
indicates the transition energy from Zni level to the valance
band in ZnO corresponding to 3.1 eV [21]. This agrees well
with our experimental results, the transition energy from the
observed trap (Zni) to the valence band is 3.13  0.02 eV. In
our case for the ZnO NTs and NRs we correlate the traps to
Zni that was recently shown to be a radiative deep level
emitting a violet line [21]. During the growth we have nonequilibrium conditions. Due to this, Zn interstitial are
present. However Janotti and Van de Walle [22] used their
calculation assuming that equilibrium prevails. Hence their
conclusion cannot be used in our case. Electroluminescence
spectra for Schottky contacts in the present ZnO NRs (not
shown here) showed a clear violet emission line in agreement
with other reports [21]. Furthermore, it was shown that for
hydrothermally grown ZnO bulk samples, Zni can be
transformed to Zno in the presence of Vo leading to a change
in the Schottky diode properties [23]. Several groups
reported that the deep-level emission (DLE) for the violet,
green, and orange red are due to Zni, Vo, and Oi, respectively
[12, 21, 24]. As mentioned above the trap concentration
observed in ZnO NRs (2.9  1017 cm3) is two orders of
magnitude higher than that observed for ZnO NTs
(3.3  1015 cm3). According to two separate recent investigations using high-resolution electron microscopy (HRTEM) [25] and depth-resolved cathodoluminescnce (CL)
[26], it was shown that the density of defects is higher near
the substrate than further up in ZnO NRs, i.e., along the
growth direction. Moreover, as the growth of the ZnO NRs
was achieved using a seed layer, which upon heating to
250 8C dissociate to form ZnO nanoparticles of a size 90 nm
[27] initiating the growth, a radial dependence of the traps is
also expected. The combination of radial and vertically
upward reduction of traps is believed to be the reason for the
higher density of traps in ZnO NRs compared to etched ZnO
NTs.
The IV characteristics of the ZnO NTs and the ZnO NRs
are shown in Fig. 5. ZnO NTs show a relatively low turn-on
voltage and a much higher current growth rate. Furthermore,
due to the large surface area and with the same forward bias,
ZnO NTs offer higher electric current compared to the NRs
as shown in Fig. 5. It is reported that the shape of the IV
characteristics depends on the Rs of the device. If the Rs is
high, the IV curve will show wide curvature, and if the effect
of Rs is less, then the nonlinear region of forward bias IV
www.pss-a.com

751

Figure 5 Currentvoltage characteristics of Au/ZnO NTs and NRs


Schottky diode.

curve will be small [28]. As shown in the figure, the IV curve


for NRs shows wider curvature than NTs, so the Rs for the
NTs and NRs is calculated to be 1.5 and 90 kV, respectively,
using the method developed by Cheung and Cheung [29].
The Au/ZnO NRs Schottky diode is much more resistive than
that of the Au/ZnO NTs Schottky diode [30]. There are two
possibilities why NRs have higher resistance than NTs.
These are the higher trap concentration and/ or the lower
mobility in the case of NRs.
In summary, the current-transport mechanism of Au
Schottky diodes fabricated on ZnO NTs has been studied.
The results illustrate a current-transport mechanism exhibiting three different regimes. The tunneling mechanism via
deep-level states is the main conduction process at low
applied voltage, while space-charge-limited current conduction dominated the current transport at higher applied
voltage. We have found lower diode turn-on threshold
voltage for the NTs compared to NRs. Some deep-level
parameters, namely, the effective carrier concentration, the
trap concentration, and the energy of the traps were all
calculated. The trap concentration in NRs was found to be
two orders of magnitude higher than for NTs. This is
probably due to the inhomogeneity in trap concentration
along the radial as well as along the vertical growth direction.
The deep-level states observed are attributed to Zni, which
are responsible for violet emission.
References
[1] C. Y. Chang, F. C. Tsao, C. J. Pan, G. C. Chi, H. T. Wang, J. J.
Chen, F. Ren, D. P. Norton, S. J. Pearton, K. H. Chen, and L.
C. Chen, Appl. Phys. Lett. 88, 173503 (2006).
[2] D. C. Look, Mater. Sci. Eng. B 80, 383 (2001).
[3] C. Klingshirn, Phys. Status Solidi B 244, 3027 (2007).
zgur, Ya. I. Alivov, C. Liu, A. Teke, M. Reshchikov, S.
[4] U. O
Dogan, V. Avrutin, S. J. Cho, and H. Morkoc, J. Appl. Phys.
98, 041301 (2005).
[5] Z. L. Wang, Mater. Sci. Eng. R 64, 33 (2009).
[6] A. Wadeasa, O. Nur, and M. Willander, Nanotechnology 20,
065710 (2009).
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

752

solidi

status

physica

pss

G. Amin et al.: Current transport studies and trap extraction of hydrothermally grown ZnO nanotubes

[7] J. Elias, R. Tena-Zaera, G.-Y. Wang, and C. Levy-Clement,


Chem. Mater. 20, 6633 (2008).
[8] L. Yu, G. Zhang, S. Li, Z. Xi, and D. Guo, J. Cryst. Growth
299, 184 (2007).
[9] L. Xu. Liao, Q. Zhang, J. X. Ai, and D. Xu, J. Phys. Chem. C
111, 4549 (2007).
[10] L. Vayssieres, K. Keis, A. Hagfeldt, and S. Lindquist, Chem.
Mater. 13, 4395 (2001).
[11] Y. Tang, L. Luo, Z. Chen, Y. Jiang, B. Li, Z. Jia, and L. Xu, J.
Electrochem. Commun. 9, 289 (2007).
[12] C. H. Ahn, Y. Yi. Kim, D. C. Kim, S. K. Mohanta, and H. K.
Cho, J. Appl. Phys. 105, 013502 (2009).
[13] Q. L. Gu, C. K. Cheung, C. C. Ling, A. M. C. Ng, A. B.
Djurisic, L. W. Lu, X. D. Chen, S. Fung, C. D. Beling, and H.
C. Ong, J. Appl. Phys. 103, 093706 (2008).
[14] H. L. Mosbacker, Y. M. Strzhemechny, B. D. White, P. E.
Smith, D. C. Look, D. C. Reynolds, and C. W. Litton, Appl.
Phys. Lett. 87, 012102 (2005).
[15] M. W. Allen and S. M. Durbin, Appl. Phys. Lett. 92, 12110 (2008).
[16] D. T. Quan and H. Hbib, Solid-State Electron. 36, 339
(1993).
[17] N. Koteeswara Reddy, Q. Ahsanulhaq, J. H. Kim, and Y. B.
Hahn, Appl. Phys. Lett. 92, 043127 (2008).
[18] M. A. Lampert and P. Mark, Current Injection in Solids
(Academic Press, New York, 1970).
[19] N. F. Mott and R. W. Gurney, Electronic Processes in Ionic
Crystals (Oxford University Press, London, 1940).

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

[20] E. G. Bylandeer, J. Appl. Phys. 49, 1188 (1978).


[21] H. Zeng, Z. Li, W. Cai, and P. Liu, J. Appl. Phys. 102, 104307
(2007).
[22] A. Janotti and C. G. Van de Walle, Phys. Rev. B 76, 165202
(2007).
[23] H. Noor, P. Klason, O. Nur, Q. Wahab, M. Asghar, and M.
Willander, J. Appl. Phys. 105, 123510 (2009).
[24] A. B. Djurisic, Y. H. Leung, K. H. Tam, Y. F. Hsu, L. Ding,
W. K. Ge, Y. C. Zhong, K. S. Wong, W. K. Chan, H. L. Tam,
K. W. Cheah, W. M. Kwok, and D. L. Phillips, Nanotechnology 18, 095702 (2007).
[25] M. Willander, O. Nur, Q. X. Zhao, L. L. Yang, M. Lorenz,
B. Q. Cao, J. Zuniga Perez, C. Czekalla, G. Zimmermann,
M. Grundmann, A. Bakin, A. Behrends, M. Al-Suleiman,
A. El-Shaer, A. Che Mofor, B. Postels, A. Waag, N. Boukos,
A. Travlos, H. S. Kwack, J. Guinard, and D. Le Si Dang,
Nanotechnology 20, 332001 (2009).
[26] N. Bano, I. Hussain, O. Nur, M. Willander, Q. Wahab, A.
Henry, H. S. Kwack, and d. Le Si Dang, in press (2009).
[27] S. Aydogan, M. Saglam, and A. Turut, Microelectron. Eng.
85, 278 (2008).
[28] C. Li, G. Fang, J. Li, L. Ai, B. Dong, and X. Zhao, J. Phys.
Chem. C 112, 990 (2008).
[29] S. K. Cheung and N. W. Cheung, Appl. Phys. Lett. 49, 85
(1986).
[30] H. Guo, Z. Lin, Z. Feng, L. Lin, and J. Zhou, J. Phys. Chem. C
113, 12546 (2009).

www.pss-a.com

You might also like