You are on page 1of 4

In the Laboratory

Flat-Band Potential of a Semiconductor: W


Using the Mott–Schottky 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 A


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 en-
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ergy. With the current movement away from fossil fuel-based


energy towards more environmentally friendly, renewable en-
Downloaded via UNIV DU QUEBEC A MONTREAL on April 13, 2019 at 19:25:08 (UTC).

ergy sources, research in this area is again gaining momen-


tum. 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 in-
herent 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 B
one of the fundamental properties of any semiconductor–elec-
trolyte 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 semiconductor–electrolyte interface, together with the
Mott–Schottky equation for determining the flat-band po-
tential.

The Semiconductor–Electrolyte Interface


Of primary importance in the development of electro-
chemical photovoltaic cells is understanding the relationship
between semiconductor and electrolyte energy levels (2–5).
An energy-level diagram for both an n-type semiconductor
and a redox couple in an electrolyte solution is shown in Fig- C
ure 1A. For the semiconductor we have identified the valence-
and 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 semicon-
ductor potential, φ, changing as ᎑eφ 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 solu-
tion; that is,
Ox + e− Red (1)
The energies of the solution states depend on whether the
state is occupied (Red) or vacant (Ox), owing to the differ-
ent solvent-sheath energies, λ, around the Red and Ox spe-
cies. Since solvent molecule exchange between the
Figure 1. (A) Schematic of an n-type semiconductor showing the va-
coordination sphere of the redox-active species and the bulk
lence and conduction bands (VB and CB, respectively), Fermi level
electrolyte is a dynamic process leading to a range of solvent- (EF), band-gap energy (EG), and the redox states in solution (Ox and
sheath energies, the density of redox states is best described Red), with their corresponding Fermi level (EF(redox)) and solvent-reor-
in terms of separate Gaussian distributions (Figure 1A). The ganization energy (λ). (B) Electronic equilibrium between the n-type
redox Fermi level, EF(redox), is again the energy at which the semiconductor and redox couple in solution. (C) Situation when the
probability of a state being occupied by an electron is 0.5. semiconductor is at its flat-band potential Vfb.

www.JCE.DivCHED.org • Vol. 84 No. 4 April 2007 • Journal of Chemical Education 685


In the Laboratory

Here C and A are the interfacial capacitance and area, re-


spectively, ND the number of donors, V the applied voltage,
kB is Boltzmann’s constant, T the absolute temperature, and
e is the electronic charge. Therefore, a plot of 1兾C 2 against
V should yield a straight line from which Vfb can be deter-
mined from the intercept on the V axis. The value of ND can
also be conveniently found from the slope knowing ε and A.

Experimental
Electrode Preparation
A schematic of the semiconductor electrode used in this
work is shown in Figure 2. Essentially a compacted, sintered
Figure 2. Schematic of the assembled electrode.
disk of polycrystalline ZnO was mounted using chemically
resistant epoxy into a polypropylene tube. Before being en-
When an n-type semiconductor and a redox couple come capsulated, a contact wire was attached to the back of the
into contact, where EF is higher in energy compared to ZnO disk using Ag-loaded epoxy. The surface of the ZnO
EF(redox), equilibrium can be achieved through the transfer of was polished, thoroughly washed with ultra-pure water, and
electrons from the semiconductor to Ox so that the Fermi then patted dry prior to use.
levels for both phases are equal, as in Figure 1B. This has the
effect of charging the semiconductor positively, and since Electrochemical Protocol
semiconductor carrier densities are much lower than those The ZnO electrode was immersed in an aqueous solu-
in solution, the diffuse charge in the semiconductor (space tion of 7 × 10᎑4 M K3[Fe(CN)6] in 1 M KCl, together with
charge region) is counterbalanced essentially by a sheet of a saturated calomel reference electrode (SCE) and a Pt counter
charge in the electrolyte. Changing the voltage of the semi- electrode. Previously the electrolyte solution had been de-
conductor artificially through the use of a potentiostat causes gassed of oxygen by purging with nitrogen.
the semiconductor and redox couple Fermi levels to separate, The basis of an electrochemical impedance spectroscopy
and hence the level of band bending owing to electron deple- (EIS) experiment is to apply a small amplitude sinusoidal ac
tion in the semiconductor will change depending on the ap- voltage, V(t), and then measure the amplitude and phase angle
plied voltage. When the applied voltage is such that there is (relative to the applied voltage) of the resulting current, I(t).
no band bending, or charge depletion (Figure 1C), then the From this the impedance, Z(ω), can be determined using
semiconductor is at its flat-band potential, Vfb . Ohm’s law (5):

The Mott–Schottky Equation V (t ) = V0 + Vm sin (ω t ) (4a)

Under the circumstances shown in Figure 1A, that is,


where EF > EF(redox), the Mott–Schottky equation can be used I (t ) = I 0 + I m sin (ω t + θ) (4b)
to determine the flat-band potential of the semiconductor.
Understanding its derivation is essential for this experiment V (t )
because it reinforces many key concepts associated with the Z (ω ) = (4c)
semiconductor–electrolyte interface. For the complete deri- I (t )
vation the reader is referred to the Supplemental MaterialW Here V0 and I0 are the dc bias potential and the steady-state
for this experiment. However, in short, the starting point for current flowing through the electrode, respectively, when the
the derivation is Poisson’s equation in one dimension that impedance experiment is conducted, Vm and Im are the maxi-
describes the relationship between charge density and poten- mum voltage and current of the supplied sinusoidal signal,
tial difference, φ, in a phase, respectively, and θ is the phase angle of the resultant cur-
d2 φ ρ rent. An alternative, more convenient description is to ex-
2
= − (2) press Z(ω) in terms of orthogonal axes rather than polar
dx ε ε0 coordinates:
where ρ corresponds to the charge density at a position x
Z (ω ) = Z ′ + j Z ′′ (5a)
away from the semiconductor surface, ε is the dielectric con-
stant of the semiconductor, and ε0 is the permittivity of free
space. Using the Boltzmann distribution to describe the dis- Z ′ = Z (ω ) cos (θ) (5b)
tribution of electrons in the space charge region and Gauss’
law relating the electric field through the interface to the
charge contained within that region, Poisson’s equation can Z ′′ = Z (ω ) sin (θ) (5c)
be solved to give the Mott–Schottky equation:
1 2 k T Vm
2
= 2
V − Vf b − B Z (ω ) = (5d)
C ε ε 0 A e ND e (3) Im

686 Journal of Chemical Education • Vol. 84 No. 4 April 2007 • www.JCE.DivCHED.org


In the Laboratory

Here j is the imaginary number ( j = √᎑1). A range of fre-


quencies, ω, can be examined to generate an impedance spec-
trum.
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 with-
out the need for manual input.
Figure 3. Typical EIS response for ZnO immersed in 7 × 10᎑4 M
K3[Fe(CN)6] (+0.8 V versus SCE).
Specialized Equipment

• 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 Re-
search EG&G 273A), controlling software (ZPlot by
Schlumberger)
Figure 4. Modified Randles circuit used to model the ZnO elec-
Hazards trode interface. Terms are defined in the text.

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 experi-
ment. Both the Ag-epoxy and chemically resistant epoxy can circle seen in Figure 3. The impedance of a CPE in an ac
be hazardous if in contact with the skin. In terms of tech- circuit, ZCPE, is
niques, using a high-temperature furnace can be a consider-
able hazard. Any user should wear appropriate personal mπ mπ
protective equipment such as a face mask, lab coat, and ther- Z CPE = σ ω−m cos − j sin (6)
2 2
mally insulating gloves, as well as use long tongs when plac-
ing in or extracting samples from the furnace. When using
electrochemical apparatus, the user should always ensure cor- where σ is the CPE prefactor, ω is the angular frequency (ω
rect electrical contacts between the equipment and cell. Fur- = 2πf ), m is the CPE exponent (0 ≤ m ≤ 1), and j is the
thermore, equipment compliance should be evaluated using imaginary number ( j = √᎑1). Note that if m = 1 then ZCPE
a dummy cell. represents an ideal capacitor, ZC. RCT and ZCPE are in paral-
lel with each other because they represent alternate charge
Results and Discussion paths at the electrode surface. Also included in series with
RCT is a Warburg impedance, ZW, which takes into account
EIS Data and Analysis diffusion of electroactive species towards the electrode, which
A typical EIS result is shown in Figure 3. The first no- is most significant at low frequencies. ZW is essentially the
table feature is the depressed semicircular response to changes same as ZCPE, but with m = 0.5 in eq 6. The final compo-
in frequency. Interpretation of the EIS data was carried out nent is another resistance (RS) representing the voltage drop
by considering the possible faradaic and non-faradaic pro- in the electrolyte owing to the passage of current between
cesses that can occur at the ZnO surface and then relating the surface of the ZnO electrode and the reference electrode.
those back to a modified Randles circuit (6). EIS data collected in this work were then modeled by
The only possible faradaic process involves charge complex nonlinear least-squares regression (7, 8) using the
transfer in the [Fe(CN)6]3−兾[Fe(CN)6]4− redox couple. This total impedance of the modified Randles circuit (Figure 4).
is probably one of the more well-defined and reversible From the extracted parameters, the interfacial capacitance was
redox couples and so is ideal for this study. Charge trans- determined.
fer in this redox couple is represented by a resistance (RCT)
in the Randles circuit (Figure 4). In parallel with RCT is Mott–Schottky Plot
the non-faradaic electrode capacitance caused by the build
up of charge at the ZnO electrode surface. In the Randles To establish a Mott–Schottky plot as described above,
circuit we have represented this by a constant-phase ele- the interfacial capacitance, C, can be determined directly from
ment (CPE) to take into account any non-homogeneity σ (eq 6) if m = 1. However, m was substantially less than
of the ZnO surface; for example, surface roughness. Sur- unity over the entire voltage range considered owing most
face non-homogeneity is indicated by the depressed semi- probably to surface non-homogeneity. Nevertheless, m was

www.JCE.DivCHED.org • Vol. 84 No. 4 April 2007 • Journal of Chemical Education 687


In the Laboratory

slope of the Mott–Schottky plot, ND = 2 × 1024 m᎑3, which


is comparable to previously reported values (6 × 1024 m᎑3;
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


In this experiment, suitable for fourth-year under-
graduate and graduate students, we have explored the na-
ture of semiconductor materials through determination of
the flat-band potential using the Mott–Schottky 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 semicon-
ductor electrodes. Electrochemical impedance spectroscopy
was then employed to examine the semiconductor–elec-
trolyte, 7 × 10᎑4 M K3[Fe(CN)6] (1 M KCl), interface as
a function of applied voltage. To achieve this a modified
Randle’s circuit was developed to interpret the impedance
data, from which a value for the interfacial capacitance
was determined. A Mott–Schottky plot was then con-
Figure 5. Mott–Schottky plot for ZnO in 7 × 10᎑4 M K3[Fe(CN)6]
(1 M KCl).
structed that allowed a flat-band potential of ᎑0.316 ±
0.033 V versus SCE to be determined. The number of car-
riers, ND, was also determined: ND = 2 × 1024 m᎑3. Both
results were comparable to literature data emphasizing the
soundness of the technique.

W
Supplemental Material
constant over the entire voltage range, having an average value
of 0.57 ± 0.02 with no apparent trend in the data. There- Instructions for the students, including pre- and post-
fore, assuming that the electrode capacitance can be repre- lab questions and the complete derivation of the Mott–
sented directly by 1兾σ, a Mott–Schottky plot was constructed Schottky equation, and notes for the instructor are available
(Figure 5). According to eq 3, the flat-band potential of ZnO in this issue of JCE Online.
was ᎑0.316 ± 0.033 V versus SCE in 7 × 10᎑4 M K3[Fe(CN)6]
(1 M KCl). The steps apparent in Figure 5 most likely origi- Literature Cited
nate from the equivalent circuit fitting procedure applied to
data with some low frequency noise (Figure 3). The result- 1. Shockley, W. B. Proc. Electrochem. Soc. 1998, 98–1, 26.
ant σ values when squared then tend to vary only slightly, as 2. Morrison, S. R. Electrochemistry at Semiconductor and Oxidized
seen in the Mott–Schottky plot. Metal Electrodes; Plenum Press: New York, 1980.
In comparison with previous works, Freund and 3. West, A. R. Solid State Chemistry and Its Applications; John
Morrison (9) reported ᎑0.41 V versus SCE for a similar sys- Wiley and Sons: Chichester, United Kingdom, 1984.
tem. However, in this example a single crystal of ZnO was 4. Bockris, J. O’M.; Reddy, A. K. N. Modern Electrochemistry:
used, allowing for a well-defined crystal plane, [001], to be Plenum Press: New York, 1970; Vols. 1 and 2.
exposed to the electrolyte. The polycrystalline ZnO electrode 5. Impedance Spectroscopy: Emphasizing Solid Materials and Sys-
used here means that many crystal planes would be exposed, tems; Macdonald, J. R., Ed.; John Wiley and Sons: New York,
suggesting that the different Vfb values arise as a result of con- 1987.
ductivity differences along different crystallographic planes. 6. Randles, J. E. B. Discuss. Faraday Soc. 1947, 1, 11.
To determine ND from the Mott–Schottky equation 7. Boukamp, B. A. Solid State Ionics 1986, 18, 136.
(slope in Figure 5), Morrison (2) has quoted the use of a di- 8. Boukamp, B. A. Solid State Ionics 1986, 20, 30.
electric constant of 8.5 for this system. Therefore, from the 9. Freund, T.; Morrison, S. R. Surface Science 1968, 9, 119.

688 Journal of Chemical Education • Vol. 84 No. 4 April 2007 • www.JCE.DivCHED.org

You might also like