Sie sind auf Seite 1von 4

In the Laboratory

Flat-Band Potential of a Semiconductor:


Using the MottSchottky Equation
K. Gelderman, L. Lee, and S. W. Donne*
Discipline of Chemistry, University of Newcastle, Callaghan, NSW 2308, Australia; *scott.donne@newcastle.edu.au

It has long been known that metals are good conductors


of electricity. However, the discovery of semiconductors and
the transistor effect by Bardeen, Shockley, and Brattain in 1948
(summarized in ref 1) generated considerable interest in the
electronic properties of all materials and paved the way for
the development of the myriad of electronic devices we have
today. Of particular relevance to this article are photovoltaic
cells, which can be used to convert light into electrical energy. With the current movement away from fossil fuel-based
energy towards more environmentally friendly, renewable energy sources, research in this area is again gaining momentum. Previously, it was at a peak during the energy crisis of
the early 1980s, when there was a global deficiency of fossil
fuels. At the present time, however, the motivation is the inherent realization that fossils fuels are a finite resource, as well
as being detrimental to the environment when combusted.
In this article we describe an experiment to determine
one of the fundamental properties of any semiconductorelectrolyte system; namely, its flat-band potential. To gain a more
thorough understanding of this semiconductor property,
which will be of significance to both senior undergraduate
and graduate students, we begin by describing the nature of
the semiconductorelectrolyte interface, together with the
MottSchottky equation for determining the flat-band potential.

The SemiconductorElectrolyte Interface


Of primary importance in the development of electrochemical photovoltaic cells is understanding the relationship
between semiconductor and electrolyte energy levels (25).
An energy-level diagram for both an n-type semiconductor
and a redox couple in an electrolyte solution is shown in Figure 1A. For the semiconductor we have identified the valenceand conduction-band edges (VB and CB, respectively), the
band-gap energy (EG), and the Fermi level (EF), which is the
energy at which the probability of an electronic state being
occupied is 0.5. These bands are dependent on the semiconductor potential, , changing as e where e is the charge on
an electron.
The energy levels for redox-active species in solution arise
by virtue of the donors (Red) and acceptors (Ox) in solution; that is,
(1)
Ox + e
Red
The energies of the solution states depend on whether the
state is occupied (Red) or vacant (Ox), owing to the different solvent-sheath energies, , around the Red and Ox species. Since solvent molecule exchange between the
coordination sphere of the redox-active species and the bulk
electrolyte is a dynamic process leading to a range of solventsheath energies, the density of redox states is best described
in terms of separate Gaussian distributions (Figure 1A). The
redox Fermi level, EF(redox), is again the energy at which the
probability of a state being occupied by an electron is 0.5.
www.JCE.DivCHED.org

Figure 1. (A) Schematic of an n-type semiconductor showing the valence and conduction bands (VB and CB, respectively), Fermi level
(EF), band-gap energy (EG), and the redox states in solution (Ox and
Red), with their corresponding Fermi level (EF(redox)) and solvent-reorganization energy (). (B) Electronic equilibrium between the n-type
semiconductor and redox couple in solution. (C) Situation when the
semiconductor is at its flat-band potential Vfb.

Vol. 84 No. 4 April 2007

Journal of Chemical Education

685

In the Laboratory

Here C and A are the interfacial capacitance and area, respectively, ND the number of donors, V the applied voltage,
kB is Boltzmanns constant, T the absolute temperature, and
e is the electronic charge. Therefore, a plot of 1C 2 against
V should yield a straight line from which Vfb can be determined from the intercept on the V axis. The value of ND can
also be conveniently found from the slope knowing and A.
Experimental

Figure 2. Schematic of the assembled electrode.

When an n-type semiconductor and a redox couple come


into contact, where EF is higher in energy compared to
EF(redox), equilibrium can be achieved through the transfer of
electrons from the semiconductor to Ox so that the Fermi
levels for both phases are equal, as in Figure 1B. This has the
effect of charging the semiconductor positively, and since
semiconductor carrier densities are much lower than those
in solution, the diffuse charge in the semiconductor (space
charge region) is counterbalanced essentially by a sheet of
charge in the electrolyte. Changing the voltage of the semiconductor artificially through the use of a potentiostat causes
the semiconductor and redox couple Fermi levels to separate,
and hence the level of band bending owing to electron depletion in the semiconductor will change depending on the applied voltage. When the applied voltage is such that there is
no band bending, or charge depletion (Figure 1C), then the
semiconductor is at its flat-band potential, Vfb .

Electrode Preparation
A schematic of the semiconductor electrode used in this
work is shown in Figure 2. Essentially a compacted, sintered
disk of polycrystalline ZnO was mounted using chemically
resistant epoxy into a polypropylene tube. Before being encapsulated, a contact wire was attached to the back of the
ZnO disk using Ag-loaded epoxy. The surface of the ZnO
was polished, thoroughly washed with ultra-pure water, and
then patted dry prior to use.
Electrochemical Protocol
The ZnO electrode was immersed in an aqueous solution of 7 104 M K3[Fe(CN)6] in 1 M KCl, together with
a saturated calomel reference electrode (SCE) and a Pt counter
electrode. Previously the electrolyte solution had been degassed of oxygen by purging with nitrogen.
The basis of an electrochemical impedance spectroscopy
(EIS) experiment is to apply a small amplitude sinusoidal ac
voltage, V(t), and then measure the amplitude and phase angle
(relative to the applied voltage) of the resulting current, I(t).
From this the impedance, Z(), can be determined using
Ohms law (5):

The MottSchottky Equation


Under the circumstances shown in Figure 1A, that is,
where EF > EF(redox), the MottSchottky equation can be used
to determine the flat-band potential of the semiconductor.
Understanding its derivation is essential for this experiment
because it reinforces many key concepts associated with the
semiconductorelectrolyte interface. For the complete derivation the reader is referred to the Supplemental MaterialW
for this experiment. However, in short, the starting point for
the derivation is Poissons equation in one dimension that
describes the relationship between charge density and potential difference, , in a phase,
d2
dx

(2)

Journal of Chemical Education

(4a)

I (t ) = I 0 + I m sin ( t + )

(4b)

Z ( ) =

V (t )
I (t )

(4c)

Here V0 and I0 are the dc bias potential and the steady-state


current flowing through the electrode, respectively, when the
impedance experiment is conducted, Vm and Im are the maximum voltage and current of the supplied sinusoidal signal,
respectively, and is the phase angle of the resultant current. An alternative, more convenient description is to express Z() in terms of orthogonal axes rather than polar
coordinates:

where corresponds to the charge density at a position x


away from the semiconductor surface, is the dielectric constant of the semiconductor, and 0 is the permittivity of free
space. Using the Boltzmann distribution to describe the distribution of electrons in the space charge region and Gauss
law relating the electric field through the interface to the
charge contained within that region, Poissons equation can
be solved to give the MottSchottky equation:
k T
1
2
V Vf b B
=
2
2
e
C
0 A e ND
(3)

686

V (t ) = V0 + Vm sin ( t )

Vol. 84 No. 4 April 2007

Z ( ) = Z + j Z

(5a)

Z = Z ( ) cos ()

(5b)

Z = Z ( ) sin ()

(5c)

Z ( ) =

Vm
Im

www.JCE.DivCHED.org

(5d)

In the Laboratory

Here j is the imaginary number ( j = 1). A range of frequencies, , can be examined to generate an impedance spectrum.
In this experiment, the impedance of the ZnO electrode
was measured at bias potentials ranging from +0.8 to 0.5 V
(versus SCE) in 50-mV increments, with 15 minutes allowed
for equilibration at each new potential. The frequency range
was from 20 kHz to 0.1 Hz, with Vm set at 5 mV. Clearly
this is a long experiment and so it is highly preferable to use
an automated system that can control the experiment without the need for manual input.
Specialized Equipment

Figure 3. Typical EIS response for ZnO immersed in 7 104 M


K3[Fe(CN)6] (+0.8 V versus SCE).

Pellet press and hydraulic ram


High-temperature oven capable of at least 1200 C
Equipment to carry out EIS; e.g., gain-phase analyzer
(Solartron 1253), potentiostat (Princeton Applied Research EG&G 273A), controlling software (ZPlot by
Schlumberger)
Figure 4. Modified Randles circuit used to model the ZnO electrode interface. Terms are defined in the text.

Hazards
K3[Fe(CN)6] is toxic if swallowed or by skin contact;
however, the quantities used in this experiment are small.
ZnO and KCl do not pose a serious hazard in this experiment. Both the Ag-epoxy and chemically resistant epoxy can
be hazardous if in contact with the skin. In terms of techniques, using a high-temperature furnace can be a considerable hazard. Any user should wear appropriate personal
protective equipment such as a face mask, lab coat, and thermally insulating gloves, as well as use long tongs when placing in or extracting samples from the furnace. When using
electrochemical apparatus, the user should always ensure correct electrical contacts between the equipment and cell. Furthermore, equipment compliance should be evaluated using
a dummy cell.
Results and Discussion

EIS Data and Analysis


A typical EIS result is shown in Figure 3. The first notable feature is the depressed semicircular response to changes
in frequency. Interpretation of the EIS data was carried out
by considering the possible faradaic and non-faradaic processes that can occur at the ZnO surface and then relating
those back to a modified Randles circuit (6).
The only possible faradaic process involves charge
transfer in the [Fe(CN)6]3[Fe(CN)6]4 redox couple. This
is probably one of the more well-defined and reversible
redox couples and so is ideal for this study. Charge transfer in this redox couple is represented by a resistance (RCT)
in the Randles circuit (Figure 4). In parallel with RCT is
the non-faradaic electrode capacitance caused by the build
up of charge at the ZnO electrode surface. In the Randles
circuit we have represented this by a constant-phase element (CPE) to take into account any non-homogeneity
of the ZnO surface; for example, surface roughness. Surface non-homogeneity is indicated by the depressed semi-

www.JCE.DivCHED.org

circle seen in Figure 3. The impedance of a CPE in an ac


circuit, ZCPE, is
Z CPE = m cos

m
2

j sin

m
2

(6)

where is the CPE prefactor, is the angular frequency (


= 2f ), m is the CPE exponent (0 m 1), and j is the
imaginary number ( j = 1). Note that if m = 1 then ZCPE
represents an ideal capacitor, ZC. RCT and ZCPE are in parallel with each other because they represent alternate charge
paths at the electrode surface. Also included in series with
RCT is a Warburg impedance, ZW, which takes into account
diffusion of electroactive species towards the electrode, which
is most significant at low frequencies. ZW is essentially the
same as ZCPE, but with m = 0.5 in eq 6. The final component is another resistance (RS) representing the voltage drop
in the electrolyte owing to the passage of current between
the surface of the ZnO electrode and the reference electrode.
EIS data collected in this work were then modeled by
complex nonlinear least-squares regression (7, 8) using the
total impedance of the modified Randles circuit (Figure 4).
From the extracted parameters, the interfacial capacitance was
determined.
MottSchottky Plot
To establish a MottSchottky plot as described above,
the interfacial capacitance, C, can be determined directly from
(eq 6) if m = 1. However, m was substantially less than
unity over the entire voltage range considered owing most
probably to surface non-homogeneity. Nevertheless, m was

Vol. 84 No. 4 April 2007

Journal of Chemical Education

687

In the Laboratory

slope of the MottSchottky plot, ND = 2 1024 m3, which


is comparable to previously reported values (6 1024 m3;
ref 2 ). The deviation is most probably due to the action of
surface states in the polycrystalline electrode capturing and
immobilizing the carriers.
Summary and Conclusions

Figure 5. MottSchottky plot for ZnO in 7 104 M K3[Fe(CN)6]


(1 M KCl).

In this experiment, suitable for fourth-year undergraduate and graduate students, we have explored the nature of semiconductor materials through determination of
the flat-band potential using the MottSchottky equation.
Experimentally, a technique was developed for preparing
a suitable polycrystalline ZnO electrode for study. Note
that a similar approach could be used for other semiconductor electrodes. Electrochemical impedance spectroscopy
was then employed to examine the semiconductorelectrolyte, 7 104 M K3[Fe(CN)6] (1 M KCl), interface as
a function of applied voltage. To achieve this a modified
Randles circuit was developed to interpret the impedance
data, from which a value for the interfacial capacitance
was determined. A MottSchottky plot was then constructed that allowed a flat-band potential of 0.316
0.033 V versus SCE to be determined. The number of carriers, ND, was also determined: ND = 2 1024 m3. Both
results were comparable to literature data emphasizing the
soundness of the technique.
W

constant over the entire voltage range, having an average value


of 0.57 0.02 with no apparent trend in the data. Therefore, assuming that the electrode capacitance can be represented directly by 1, a MottSchottky plot was constructed
(Figure 5). According to eq 3, the flat-band potential of ZnO
was 0.316 0.033 V versus SCE in 7 104 M K3[Fe(CN)6]
(1 M KCl). The steps apparent in Figure 5 most likely originate from the equivalent circuit fitting procedure applied to
data with some low frequency noise (Figure 3). The resultant values when squared then tend to vary only slightly, as
seen in the MottSchottky plot.
In comparison with previous works, Freund and
Morrison (9) reported 0.41 V versus SCE for a similar system. However, in this example a single crystal of ZnO was
used, allowing for a well-defined crystal plane, [001], to be
exposed to the electrolyte. The polycrystalline ZnO electrode
used here means that many crystal planes would be exposed,
suggesting that the different Vfb values arise as a result of conductivity differences along different crystallographic planes.
To determine ND from the MottSchottky equation
(slope in Figure 5), Morrison (2) has quoted the use of a dielectric constant of 8.5 for this system. Therefore, from the

688

Journal of Chemical Education

Supplemental Material

Instructions for the students, including pre- and postlab questions and the complete derivation of the Mott
Schottky equation, and notes for the instructor are available
in this issue of JCE Online.
Literature Cited
1. Shockley, W. B. Proc. Electrochem. Soc. 1998, 981, 26.
2. Morrison, S. R. Electrochemistry at Semiconductor and Oxidized
Metal Electrodes; Plenum Press: New York, 1980.
3. West, A. R. Solid State Chemistry and Its Applications; John
Wiley and Sons: Chichester, United Kingdom, 1984.
4. Bockris, J. OM.; Reddy, A. K. N. Modern Electrochemistry:
Plenum Press: New York, 1970; Vols. 1 and 2.
5. Impedance Spectroscopy: Emphasizing Solid Materials and Systems; Macdonald, J. R., Ed.; John Wiley and Sons: New York,
1987.
6. Randles, J. E. B. Discuss. Faraday Soc. 1947, 1, 11.
7. Boukamp, B. A. Solid State Ionics 1986, 18, 136.
8. Boukamp, B. A. Solid State Ionics 1986, 20, 30.
9. Freund, T.; Morrison, S. R. Surface Science 1968, 9, 119.

Vol. 84 No. 4 April 2007

www.JCE.DivCHED.org

Das könnte Ihnen auch gefallen