You are on page 1of 7

A lot of attention has been directed towards ZnO@TiO2 core-shell nanostructures for use in dye-

sensitized solar cells (DSSCs) and water-splitting photo-electrodes. Both materials have a wide band gap
above 3 eV (Kwiatkowski et al). However, ZnO has a much higher electron mobility and longer carrier
lifetimes, which make it a better conductor, and a large exciton binding energy to prevent
recombination of electrons and holes (Fan et al). In spite of those qualities, ZnO-based DSSCs give poor
results due to the instability of the ZnO surface in the presence of the dyes (Quintana et al). TiO2, in
contrast, shows a high surface stability, indicating that it would make a good boundary between the ZnO
and the dye molecules (Goh et al). In order to encapsulate the ZnO in TiO2 and still take advantage of its
conductivity, many people have been growing nanowire assemblies of ZnO on fluorine-doped or indium-
doped tablets of tin oxide.

Many techniques for synthesizing ZnO@TiO2 nanowire (NW) arrays have been developed, including
ultra-high vacuum techniques such as vapor phase transport and chemical vapor deposition, and wet-
chemical techniques such as sol-gel synthesis or electrochemical deposition. The latter are significantly
less expensive than the former due to the former’s need for UHV conditions and lack of scalability for
large area solar cells (Hernandez et al). The various wet-chemical methods differ by use of colloidal
particles (sol-gel), heat (hydrothermal), or electric current (electrochemical) in a liquid-phase chemical
reaction. Most of the methods used, regardless of expense, tend to produce tightly packed ZnO
nanowires with diameters on the order of 10-100 nm and lengths that range from 100 nm to many tens
of microns (Goh et al, Irannejad et al). Many of the same techniques used for creating the ZnO NW
arrays can be adapted to deposit TiO2 onto the ZnO NWs. These methods can be used to generate
rugged or smooth TiO2 shells with thicknesses ranging from 3 nm to 22 nm (Goh et al). The prevalence
of wet-chemical synthesis of ZnO@TiO2 NWs in research is such that it is worth reviewing how these
processes work in more detail.

In most wet-chemical techniques for synthesizing ZnO nanowire arrays, a seed layer of ZnO is deposited
onto a tin oxide plate doped with something like fluorine or indium by soaking the plate in a solution of
ethanol and Zn(CH3COO)2 and annealing it at a high temperature (Goh et al, Kwiatkowski et al). The
seeded plate is then immersed in a solution containing precursor chemicals at a raised temperature for
an extended period of time. There are many different combinations of precursor chemicals that work for
the various wet-chemical methods. As an example, in one hydrothermal treatment, a seeded plate is
immersed in the combination of zinc chloride, hexamethylenetetramine, and ammonium hydroxide at
95 C for 12 hours, then rinsed and annealed at 300 C for 1 hour to synthesize a ZnO NW array with rod
diameters in the range 50-100 nm and lengths on the order of 10 µm (Feng et al). Depositing the TiO2
onto the ZnO NWs is similarly done, although the precursor solution will be particular to the TiO2, and
the times and temperatures of immersion and annealing will generally be optimized for the TiO2
deposition as well. Regardless of the particular wet-chemical method used, the concentration of the
precursors, the length of immersion time, and the temperature of the heat treatment can be tweaked to
change the resulting geometries and chemical compositions of the ZnO@TiO2 NW assemblies.
Consequent influences of morphology, doping, and chemical composition on the photocatalytic and
photoelectrochemical properties of ZnO@TiO2 solar cells are well researched and show a great deal of
promise for obtaining an inexpensive means of producing DSSCs and water-splitting photoelectrode
cells.

Morphological characteristics, such as the length of the ZnO NWs, the thickness of the TiO2 shells, and
the roughness of the TiO2 shell, and their influences on the solar cell characteristics are well-
documented. In one study, it was found that by changing the concentration of ammonium hydroxide in
the aforementioned hydrothermal treatment, the ZnO NWs could be synthesized with predetermined
wire lengths (Feng et al). By doubling the concentration of the ammonium hydroxide, researchers found
that the ZnO NW lengths changed from 5 µm to 14 µm. The current vs voltage for the various lengths
can be seen in [insert Feng figure number here]. The current density was found to increase by more than
a factor of two as the length of the wires increased, and the cell efficiencies increased from 1.20% to
3.80% (Feng et al).

Figure 1: Current density vs voltage for ZnO and ZnO@TiO2 NW arrays of various lengths. Obtained from Feng et al.

The effect of varying the thickness of the TiO2 shell has been investigated in several studies. In UHV
depositions, the shell thickness varies with number of deposition cycles (Irannejad et al). In wet-
chemical techniques, the thickness can be controlled directly by the length of time immersed in the
precursor solution (Goh et al), or by repeating the deposition process some number of cycles (Feng et
al). In one study, the thickness of the TiO2 shell was varied from about 10 nm after a 5 minute
immersion to about 35 nm after a 15 minute immersion. The resulting currents vs voltages can be seen
in [insert Goh figure number here]. These results demonstrate that the thickness of the shell doesn’t
have much influence up to a certain threshold, at which point the shell becomes too thick to allow
efficient transport of the electrons from the TiO2 into the ZnO before recombination takes place.
However, in another UHV-technique study, the shell thickness was varied from 7 nm to 21 nm, with the
current vs voltage shown in [insert Irranejad figure number here]. These results would appear to
indicate that the thickness has more of an influence. It is worth noting that the current densities for
[insert Goh figure number here] are over twice as high as those of [insert Irranejad figure number here],
indicating that there is possibly another factor that accounts for the differences between the two
results.
Figure 2: Current density vs voltage for ZnO/TiO2 NW arrays with various TiO2 shell thicknesses. Obtained from Goh et al.

Figure 3: Current density vs voltage for ZnO/TiO2 NW arrays with various TiO2 shell thicknesses. Obtained from Irannejad et al.

In another study, it was shown that by adjusting the sol-gel treatment for TiO2 deposition from one 60-
minute deposition cycle to three 30-minute deposition cycles, the TiO2 shell’s roughness can be changed
from a rugged coat to a smooth one, as can be observed in the schematic shown in [insert schematic
from Kwiatkowski]. The roughness of the TiO2 shell influences the band gap of the core-shell structures,
lowering it from 3.05 eV for pure ZnO NWs to 2.93 eV for the rough shell and 2.85 eV for the smooth
shell. The smooth shell’s lower band gap is attributed to a more continuous interface between the ZnO
and the TiO2. The rugged shell showed a higher rate of photocatalytic activity, which was attributed to
the increased accessibility of the solution species to the interface between the ZnO and the TiO2
(Kwiatkowski et al). These results alone are enough to demonstrate the importance of the ZnO@TiO2
NW morphologies when designing DSSCs with higher efficiency.

Figure 4: Schematic illustrating the difference in surface roughness of the TiO2 shell before and after the heat treatment.
Obtained from Kwiatkowski et al.
The modification of the chemical composition of the ZnO@TiO2 NW arrays, by processes such as
calcination, chemical doping, etc, can also influence the characteristics of the solar cells. In one study, it
was found that subjecting the ZnO/TiO2 array to a calcination at 450 C for 3 hours resulted in a
decreased carrier recombination rate, leading to a higher open circuit potential than the un-calcined
cells (Kwiatkowski et al). SEM images of the calcined NW arrays showed evidence of voids appearing at
the ZnO/TiO2 interface and zinc ions diffusing into the TiO2 shell, as can be seen in [insert diagram
number from Kwiatkowski here]. The voids are due to the diffusion of the zinc ions, which cause the
interface between the ZnO and the TiO2 to move as the zinc ions move from one side to the other. This
movement of the interface due to diffusion is called the Kirkendall effect. The Kirkendall effect typically
only takes place at much higher temperatures, but it is hypothesized that the defective surface structure
of the ZnO nanorods facilitated the diffusion of the zinc ions (Kwiatkowski et al).

Figure 5: Diagram illustrating the formation of Kirkendall voids at the ZnO/TiO2 interface. Obtained from Kwiatkowski et al.

In another study, doping the TiO2 shell with niobium led to an increase in short circuit current and
power conversion efficiency. The currents vs voltages for various dopant concentrations are shown in
[insert figure number for Saurdi]. The current density peaks at 5% Nb concentration, after which it
begins to decrease, possibly due to the defects associated with the high Nb levels. In addition, the Nb
doping led to an increase in the PCE from 2.5% for the pure ZnO/TiO2 nanowires to over 5.3% for the 5%
Nb cell (Saurdi et al).

Figure 6: Current density vs voltage for ZnO/TiO2:Nb NW arrays with various concentrations of Nb doping. Obtained from Saurdi
et al.

The studies examined in this paper reflect only a small portion of the literature existing on these arrays.
Although they show that ZnO@TiO2 DSSCs still need some work, these DSSCs are nonetheless showing
themselves to be a serious contender for inexpensively made DSSCs, so long as researchers continue to
investigate the various influences on their characteristics and improve existing synthesis methods.
Works Cited

Ako, Rajour Tanyi, et al. "DSSCs with ZnO@ TiO2 core–shell photoanodes showing improved
Voc: Modification of energy gradients and potential barriers with Cd and Mg ion dopants." Solar
Energy Materials and Solar Cells 157 (2016): 18-27.
Ding, Yong, et al. "Continuous electron transport pathways constructed in TiO 2 sub-
microsphere films for high-performance dye-sensitized solar cells." RSC Advances 5.23 (2015):
17493-17500.
Fan, Jiandong, et al. "Solution-growth and optoelectronic performance of ZnO: Cl/TiO2 and
ZnO: Cl/ZnxTiOy/TiO2 core–shell nanowires with tunable shell thickness." Journal of Physics
D: Applied Physics 45.41 (2012): 415301.
Feng, Yamin, et al. "Synthesis of ZnO@ TiO2 core–shell long nanowire arrays and their
application on dye-sensitized solar cells." Journal of Solid State Chemistry 190 (2012): 303-308.
Goh, Gregory Kia Liang, et al. "Low temperature grown ZnO@ TiO2 core shell nanorod arrays
for dye sensitized solar cell application." Journal of Solid State Chemistry 214 (2014): 17-23.
Govender, Kuveshni, et al. "Understanding the factors that govern the deposition and
morphology of thin films of ZnO from aqueous solution." Journal of Materials Chemistry 14.16
(2004): 2575-2591.
Greene, Lori E., et al. "ZnO− TiO2 core− shell nanorod/P3HT solar cells." The Journal of
Physical Chemistry C 111.50 (2007): 18451-18456.
Hernández, Simelys, et al. "Optimization of 1D ZnO@ TiO2 core–shell nanostructures for
enhanced photoelectrochemical water splitting under solar light illumination." ACS applied
materials & interfaces 6.15 (2014): 12153-12167.
Irannejad, A., et al. "Effect of the TiO2 shell thickness on the dye-sensitized solar cells with
ZnO–TiO2 core–shell nanorod electrodes." Electrochimica Acta 58 (2011): 19-24.
Ji, In, et al. "One-dimensional core/shell structured TiO 2/ZnO heterojunction for improved
photoelectrochemical performance." Bulletin of the Korean Chemical Society 33.7 (2012): 2200-
2206.
Kwiatkowski, Maciej, et al. "Improvement of photocatalytic and photoelectrochemical activity of
ZnO/TiO2 core/shell system through additional calcination: Insight into the mechanism."
Applied Catalysis B: Environmental 204 (2017): 200-208.
Kwiatkowski, Maciej, Igor Bezverkhyy, and Magdalena Skompska. "ZnO nanorods covered
with a TiO 2 layer: simple sol–gel preparation, and optical, photocatalytic and
photoelectrochemical properties." Journal of Materials Chemistry A 3.24 (2015): 12748-12760.
Quintana, Maria, et al. "Comparison of dye-sensitized ZnO and TiO2 solar cells: studies of
charge transport and carrier lifetime." The Journal of Physical Chemistry C 111.2 (2007): 1035-
1041.
Saurdi, I., et al. "Effect of Nb-doped TiO2 on nanocomposited aligned ZnO nanorod/TiO2: Nb
for dye-sensitized solar cells." AIP Conference Proceedings. Vol. 1733. No. 1. AIP Publishing
LLC, 2016.
Tiwana, Priti, et al. "Electron mobility and injection dynamics in mesoporous ZnO, SnO2, and
TiO2 films used in dye-sensitized solar cells." ACS nano 5.6 (2011): 5158-5166.
Zarębska, Kamila, et al. "Electrodeposition of Zn (OH) 2, ZnO thin films and nanosheet-like Zn seed
layers and influence of their morphology on the growth of ZnO nanorods." Electrochimica Acta 98 (2013):
255-262.

You might also like