Electrochemical Properties of Silicon Carbide

1 downloads 0 Views 210KB Size Report
b present address: Institute for Applied Photovoltaics, Haydnstr.19,. D–45884 Gelsenkirchen, Germany c. Department of Physics, Carl-von Ossietzky-University, ...
J. Electrochem. Soc., revised version submitted Feb. 13, 1996

Electrochemical

Iver

Properties

Lauermanna,b , Rüdiger

of

Silicon

Carbide

Memming* , a,c and Dieter Meissner * ,a,d

a Institute for Solar Energy Research (ISFH), Sokelantstr.5, 30165 Hanno-

ver, Germany address: Institute for Applied Photovoltaics, Haydnstr.19, D–45884 Gelsenkirchen, Germany c Department of Physics, Carl-von Ossietzky-University, P.O.Box 2503, 26111 Oldenburg, Germany d present address: Forschungszentrum Juelich GmbH (KFA), Institute of Energy Process Engineering, D–52425 Juelich, Germany b present

Abstract The electrochemical properties of single crystalline n- and p-type electrodes of two different modifications of SiC in aqueous solutions were investigated in the dark and under illumination. The main emphasis was laid on investigations of the problem of water splitting in photoelectrolysis cells. No oxygen was detected at SiC electrodes. Instead anodic oxidation of SiC was found leading to the formation of CO2 , CO and silicon oxides. Since the oxide film is very porous the anodic corrosion current remained on a high level. In the presence of concentrated hydrofluoric acid a different surface layer was formed possibly due to disproportionation processes and the formation of amorphous SiC. Flatband potentials at different

pH-values

were

measured

and

the

charge

transfer

processes

with various redox systems were examined. A quantitative analysis was very difficult or even impossible, probably due to the complicated surface chemistry and layer formation.

* Electrochemical Society Active Member

- 2 -

Introduction Since 1974 the photoelectrochemical properties of SiC have been investigated by various authors.1-9 Because of the large bandgap and the excellent chemical stability, it was assumed that SiC should not be subject to anodic corrosion under conditions at which photoelectrolysis of H2 O was expected although there are contradictory reports in the literature. Values for the flatband potential of n-SiC given in the literature vary in a range of -0.7 to -1.8 V vs. SCE even for the same pH and there is yet no agreement about a possible corrosion of SiC photoanodes nor about the corrosion products. Some of these results can be attributed to the inferior quality of SiC crystals that were available at the time those experiments were performed. More

recently

new

interest

in

SiC

as

a

semiconductor

material

for

devices like blue light emitting diodes has led to the investigation of (photo)electrochemical ested

in

the

etching

methods.10,11

photoelectrochemistry

of

We

large

were

mainly

bandgap

inter-

semiconduc-

tors for the use in photoelectrochemical solar energy devices and examined two SiC modifications with bandgaps of 2.4 and 3.1 eV, respectively. The use of large bandgap semiconductors ( Eg > 2 eV) allows band

a

much

processes

bandgap those

better in

distinction

fundamental

semiconductors

with

smaller

tend

bandgaps.

to The

between

valence

and

conduction

investigations.

Furthermore,

be

more

chemically

cubic

modification

stable of

SiC

large than (3C-

SiC) with a bandgap of 2.4 eV would in theory be especially suitable for

water

cathodic

splitting

since

reactions,

and

overpotentials thermodynamic

for

both,

the

restrictions

anodic

and

demand

that

the energy separation between the excited carriers must be consid-

- 3 -

erably higher than the theoretical value of 1.23 eV. In addition, the absorption of the cubic SiC is still in the visible range of the solar spectrum. At present single crystalline SiC is only available in the hexagonal

modification

Eg = 3.1 eV) or as cubic material

(6H-SiC,

(3C-SiC), the latter being deposited as a thin layer on a Si substrate in a CVD process. Another advantage of SiC over many other large bandgap materials is the availability of p-doped crystals. This fact makes it interesting to be used as a photocathode in a photoelectrolysis cell. In such a case corrosion problems may be avoided because in this configuration H 2 would be formed at the p-type SiC photocathode and O 2 at the metal counter electrode. Other semiconductors which are available as n- and p-type material such as GaAs and InP have a too small bandgap for a water splitting reaction.

Experimental Electrochemical

experiments

were

Procedures performed

in

an

electrically

shielded

dark box using the common three electrode configuration. The reference electrode chloride

was

a

saturated

electrode

in

a

calomel

saturated

electrode potassium

(SCE) chloride

or

a

silver/silver

solution.

In

this

paper all potentials are given with respect to SCE. The potentiostat was a Solartron 1255, operated through a Macintosh II via a GPIB interface using

a

home

made

LabVIEW

program

(National

Instruments,

Austin,

Texas, USA). A Tacussel bipotentiostat BI-PAD was used for the rotating ring disc experiments. In this case the data acquisition was performed through a HP 7090 digital plotter. Illumination sources used

- 4 -

for the photoelectrochemical experiments were 1000 or 450 W high pressure Osram XBO xenon lamps with a high percentage of UV light (Müller lamp or Oriel 140). The n- and p-type crystals used in our experiments, had a doping level ranging from 1016 to 101 9 cm- 3 . Crystal sources and a few

physical

properties are given in Table 1. Exact doping levels were provided only for samples obtained from source 4; for all other samples they were not known to the manufacturer. Although the doping density can be calculated in theory from the slope of a straight Mott-Schottky plot of impedance data, in many cases no reliable values were obtained because the curves were not straight over a sufficiently large potential

range.

Ohmic contacts to n- doped SiC crystals were achieved by rubbing liquid indium on the back of the samples followed by annealing under hydrogen atmosphere at 450° C. To obtain an ohmic contact on p-doped samples aluminum was evaporated onto the sample with subsequent annealing in hydrogen at 900° C. Although annealing of evaporated aluminum on p-SiC has been reported by some authors to produce a Schottky contact, we found that annealing led to the formation of a reliable

ohmic

contact,

in

accordance

with.12 The contacted samples

were glued onto a brass conductor mounted in a Teflon sample holder using silver epoxy and subsequently the sample was covered with epoxy so that only the polished [0001- ]- or [0001]-surface was exposed. Prior to each electrochemical experiment the electrodes were polished using diamond powder (0.25 µm) to give a mirror like surface, and then etched in 48% aqueous HF for 15 to 60 s to remove any native oxide surface. In the case of the 3C-SiC samples different approaches had to be used. These

- 5 -

samples consisted of thin SiC films (6 to 12 µm) grown on a highly ndoped and

silicon the

wafer.

Si/SiC

To

contact

eliminate on

the

possible

effects

electrochemical

of

the

Si-substrate

experiments,

samples

were prepared with ohmic contacts on both sides, one on the SiC front surface and the other on the back of the Si substrate. Another set of samples was prepared by forming at first an ohmic contact to the SiC front surface and protecting it by embedding the entire surface in an epoxy resin. The Si substrate was then etched away by concentrated HF/HNO3

solutions. The SiC surface appearing after etching was then

used for the electrochemical measurements. Because of the mechanical instability of these thin layers polishing was not possible and etching in HF was used as the only surface pretreatment. Fig. 1 illustrates the different

electrode

discs

of

6H-SiC

preparation (5 mm

methods.

diameter)

For

were

ring cut

disc from

electrodes crystalline

circular platelets

using an ultrasonic saw (KLN) and then pressed into thin machined Teflon tubes.

These

were

then

squeezed

into

exactly

fitting

platinum

tubes

which formed the ring electrodes of the RRDE assembly so that SiC, Teflon and platinum were sealed against each other. Gases generated in the electrochemical cell were analyzed using a gas chromatograph (Shimadzu GC 8A) placed inside the same glove box in which the experiments were conducted, thus avoiding contamination of gas samples with outside air. In order to separate O2 , N2 , Ar, H2 and CO, a 50 m

long porous-layer-open-tube

(PLOT)

quartz

capillary

column,

the

inner surface of which was covered by a molecular sieve was used at 0 ° C. To separate CO2 and the remaining gases, a 2 m stainless steel column filled with alumina powder was employed. All chemicals were of pro analysi grade and were not further purified.

- 6 -

Cobalt sepulchrate was prepared according to the procedure given by Creaser 13. The deposition of metals and metal oxides on SiC was performed by using different methods. Copper (up to 10 nm) were deposited in vacuum by evaporation at an argon pressure of 5.10- 2 mbar. Platinum was deposited on SiC by photoelectrochemical and photochemical reduction from a solution of hexachloroplatinic acid in H2 O by UV-radiation.

Results Capacity

measurements: For any photoelectrochemical system the knowl-

edge of the energetic position of the band edges at the semiconductor is crucial because a charge transfer takes place only if there is a sufficient overlap between the electronic states in the semiconductor and in the electrolyte.

Corresponding

informations

can

be

obtained

by

measure-

ments of the space charge capacity. In Fig. 2 Mott-Schottky curves are presented as obtained with the hexagonal n- and p-type (6H-SiC) electrodes and in Fig. 3 for the cubic n-type crystals (n-3C-SiC). We found that in many cases the plot of 1/ C2 vs. UE does not yield straight lines as

shown

in

extrapolating yield

very

Fig. 2. the

plots

determination

Mott-Schottky

reproducible

Mott-Schottky

The

results.

vary

curves

of

considerably

flatband

potential

by

1/C2 vs. UE does also not

to

Especially

the

with

with

(6H-SiC)

time.

More

electrodes

the

reliable

data

were obtained with the cubic crystals (as shown in Fig. 3). In this case only a week drift of the Mott-Schottky have been found. The time effects observed with the (6H-SiC)-electrodes have been found in all electrolytes including 48% HF - solutions and for all electrodes. No surface pretreatment such as etching, polishing or soaking in the solution for a certain period of time, led to more reliable capacity measurements. The

- 7 -

Mott-Schottky curves given in Fig.2 have been obtained with (6H-SiC)electrodes during the first potential sweep after having kept the electrodes in the solution for at least one hour. They are given here in order to get a reasonable basis for comparing the flatband potentials and consequently the position of energy bands of the 6H- and 3C-SiC electrodes (see discussion section). Electrochemical

behavior: In order to examine the feasibility of splitting

water at SiC- electrodes, current-voltage measurements were conducted in aqueous electrolytes without any additional redox system. Some of these experiments are similar to those we have reported earlier.1,2 Since now also cubic SiC (3C-SiC) is available which has a smaller bandgap, it is interesting to compare the electrochemical behavior of both modifications.

In

Fig. 4 a

the

j/UE curve of a n-type 6H-SiC electrode in 0.1 M

solution of Na2 SO4 under

chopped

illumination

(white

light,

0.4 Wcm- 2 )

is shown. The photocurrent curve measured under illumination by chopped light shows a very slow rise between - 1 and +1.5 Volts with pronounced current transients which occur over the whole potential range when the light is turned on. Although the j/UE curve presented in Fig. 4a is typical for n-6H-SiC, the onset of photocurrent and the quantum efficiency varied considerably from one sample to the other. This is probably due to the heterogeneous nature of most rather crude SiC crystals from which we cut our individual samples. In Fig. 4 b a current-voltage dependence is shown as obtained with the cubic material (n-3C-SiC). This electrode consists of a thin n-3C-SiC layer on a highly doped silicon substrate and the

ohmic

contact

is

on

the

back

side

of

the

silicon

substrate

as

described in the experimental section. The j/UE -curve is very similar to that of the n-(6H-SiC) electrode (compare Figs. 4 a and b). The main difference lies in the more positive onset of the photocurrent and in the

- 8 -

current transients which are less pronounced and occur in a smaller potential range in the case of 3C-SiC. In the case of these two modifications we also measured the excitation spectra with n- type electrodes., as given in Fig. 5. These spectra were measured keeping the electrode at a potential where the photocurrent is about saturated, i.e. at 1 V (SCE). Note that two different current scales are used because the quantum efficiency is much smaller for 3C-SiC. The cathodic dark currents increase more or less exponentially with the electrode potential. According to our chemical analysis the corresponding

cathodic

current

is

entirely

due

to

hydrogen

formation,

i.e.

no

cathodic corrosion was found. In the case of the n-type 3C-SiC electrode the current - potential curves were quite well reproducible (Fig 6a). A cathodic shift of the curve with increasing pH was observed as quantitatively shown in Fig. 6b, similarly as in the Mott-Schottky measurements (Fig. 3).

The

shift

of

the

current-potential

curve

which

represents

the

hydrogen formation, amounts to about 60 mV (Fig. 6b). This value agrees well with the shift of the capacity curves measured with 3C-SiC electrodes. In both cases the shift agrees with what have been expected theoretically. The electrochemical behavior of p-type electrodes was examined in order to evaluate their possible use as photocathodes in a water splitting system. Since no p-doped cubic material was available only the hexagonal material was investigated in more detail. A current potential curve as obtained with a p-6H-SiC electrode is given in Fig. 7. In most cases the cathodic dark current is relatively high. Also the photocurrent- potential curve is far from an ideal shape. Although a photocurrent was detected positive of +1 V, which is close to the reported flatband potential, the

- 9 -

full photocurrent is only reached at rather negative potentials, whereas in the range between +1.2 and 0 V a smaller photocurrent plateau is observed . The curves given in Fig. 7 were obtained if an electrolyte of 40% HF in water was used in order to remove oxide layers from the electrode. In a further set of experiments we investigated the oxidation products formed during anodic polarization of p-type electrodes in the dark and of n-type samples upon illumination. We especially checked the head space of the electrolysis cell by GC-analysis for gaseous oxidation products such as oxygen. According to our experimental results performed with both types of electrodes in a sealed cell, however, no oxygen was found even after passing more than 1017 Coulomb across the interface. We also found that no oxygen was formed with electrodes, on which small Ptislands have been deposited. Since no H2 O2 could be detected either, the anodic current must be entirely due to anodic corrosion. With n-type electrodes we observed in addition that the photocurrent decreased with time in acidic solutions without HF as well as in concentrated hydrofluoric acid. In both cases the photocurrent-time curves showed a fast initial and later a slow degradation of the photocurrent over a period of several hours (Fig. 8). This behavior is in contrast to that of silicon electrodes

insofar

as

with

silicon

in

solutions

without

HF

the

anodic

current drops always to a very low value within a few seconds due to the formation of a thin impermeable insulating oxide layer. After

prolonged

current

flow

in

sulfuric

acid

solutions

the

electrodes

are covered with a thick white layer of silicon oxides which can be readily dissolved in HF. The head space analysis of gaseous products reveals CO and CO2 in a ratio close to 1:1 as products of the oxidation. On the

- 10 -

other hand, if the electrolysis was carried out in 48% hydrofluoric acid, then

the

surface

of

the

electrode

was

covered

with

a

yellow-brown

layer, which was insoluble in HF. These layers show surface features similar to those on the original surface and they were identified as silicon carbide by XPS and by chemical analysis. Current flow over periods of days have led to a complete destruction of the electrode. Redox

reactions: In a previous paper we have presented some current-

potential curves in the presence of a redox system at hexagonal n-type SiC

electrodes

(n-6H-SiC).2 In the present paper we also want to report

on redox processes at cubic electrodes. In Fig. 9 a-c current- potential curves for the electrochemical reduction of various redox systems at n3C-SiC

electrodes

are

given.

The

cathodic

saturation

currents

are

entirely due to mass transport limitation in the electrolyte. The further current increase beyond the saturation current corresponds to hydrogen evolution. In the case of cobalt (II/III) sepulchrate we have also measured the current-potential dependence at various pH-values (not shown). There not

it

shift

was

interesting

although

to

according

note to

that

the

capacity

current-potential measurements

curves

(Fig.

3)

did the

energy bands are shifted upwards upon increasing the pH of the solution. Another peculiar result is that no reoxidation of cobalt (II) sepulchrate could be found although its standard potential (-0.5 V vs.SCE) is rather negative. In addition we studied the reduction of CO2 at SiC electrodes which has been reported in the literature.5,14 When bare SiC- electrodes were used, no difference can be detected between j/UE -curves obtained in CO2 containing solutions and those in CO2

-free

electrolytes.

For

electrodes

that have been plated electrochemically with thin layers or islands of

- 11 -

copper, however, a clear difference is visible (Fig. 10). Copper was chosen as a catalyst because other authors had used copper electrodes to reduce CO2 .15 The reduction wave caused by CO2 ,however, is often small and ill defined. In order to obtain more evidence for the reduction of CO2 the following experiment was conducted: The electrode was held at a fixed potential of -1.35 V and the current was measured over a period of 600 s. In one case the solution was purged with nitrogen only, in the other case CO2 was bubbled into the solution for the first 300 s and then it was switched back to nitrogen (see insert in Fig. 10). It is clearly seen that CO2 is reduced at the copper plated SiC-electrode. This reduction was only observed at more or less neutral solutions, it did not occur in strongly acidic nor in strongly alkaline electrolytes. There is also no reduction of sodium carbonate solutions without purging it with CO2 . Since these solutions were buffered, a simple pH-effect due to the presence of CO2 in water according to the reaction CO2 + H2 O -----> CO3 2- + 2 H+

( 1 )

can be ruled out. In that case the reduction of protons might have been the observed electrochemical process. The reduction products were analyzed by gas chromatography. The only volatile product to be found was methane. In a further set of experiments we investigated the oxidation of some redox systems. These investigations were mainly focussed on the competition between the oxidation of a redox system and the anodic decomposition reaction. Since corresponding measurements had to be performed with RRDE-assemblies they could only be done with the hexagonal electrodes. It was not possible to make a rotating ring-disc electrode from the very thin cubic samples because the silicon substrate could not be

- 12 -

sufficiently isolated from the electrolyte. In solutions without any redox system we did not observe any ring current when the disc potential was scanned to positive values. This also proves that water is not oxidized at 6H-SiC. Upon addition of certain redox systems, however, the oxidation of the SiC electrode is clearly reduced in favor of the oxidation of the redox system itself. The best stabilization was obtained with redox systems, the oxidation of which occurs at lower anodic potentials than the anodic decomposition of SiC. This is illustrated in Fig. 11 a and b for the redox couple Cu+ / Cu2 + in chloride solutions. In the same range also the ring current rises but decreases to a very low value at a potential around this

+2

V

(Fig. 11 b).

hysteresis

(Fig. 11 a).

The

corresponding

This

result

disc

indicates

current that

also

the

reflects

oxidation

of

the redox system is blocked once a corrosion current is reached exceeding that caused by the redox reaction. The original situation can not be restored

by

cathodic

polarization

but

only

by

polishing

the

electrode

with a diamond paste. Similar observations have been made for the oxidation of bromide, i.e. the latter process only took place if a large corrosion process was avoided (not shown). Only in the case of the I- / I3 - couple the oxidation continued even in the range of higher corrosion currents probably due to the adsorption of the iodide.

Discussion Potential obtain

distribution:

straight

As

Mott-Schottky

already plots

mentioned and

it

is

above

it

is

difficult

to

even

more

difficult

to

derive a reliable flatband potential because of changes with time. Our measurements have shown that, even with identical electrodes and under identical conditions, values of the flatband potential can vary by more

- 13 -

than one volt. This result indicates that the surface chemistry and especially the formation of a surface layer plays a dominant role. A comparison of papers published by different authors also shows a wide range of values for the flatband potential of both, n- and p-type electrodes (see Table 2). This scattering of data has probably the same origin as in our case. On the other hand, the quality of crystals plays an important role. This becomes quite obvious by comparing the data obtained with epitaxially grown cubic crystals and the hexagonal SiC produced in high temperature

process.

Since the Mott-Schottky curves were measured under the same conditions, the flatband potentials obtained with n-6H (Uf b = -1.25 V at pH0) and n-3C-SiC electrodes (Uf b =-0.35 V at pH0) can be compared. According to these results the difference of the flatband potentials amount to ∆ U =0.9 V, i.e. a value which is slightly above the difference of the fb bandgaps of the two modifications (∆ Eg=0.7 eV). Taking into account the uncertainty in the determination of the flatband potential it can be con-

cluded from this result that the conduction band of 3C-SiC is by 0.7-0.9 eV lower than that of 6H-SiC whereas the position of the valence band is nearly identical for both types of electrodes. There also seems to be no agreement on the pH- dependence. Whereas some authors report a pH- dependence of about 59 mV per unit pH, others have seen no pH dependence at all. We found with the n-3C-SiC electrodes clearly a pH dependence of 52 mV per unit pH, which is slightly smaller than the expected 59 mV (Fig. 3). The same shift by pH was found in the onset of photocurrent. The pH dependence does certainly also exist for 6H-SiC but it was less pronounced and could not be quantified because of poor reproducibility. It can be concluded at least from these

- 14 -

results that the surface should be OH-terminated. In general, the pH dependence of the flatband potential is attributed to the dissociation of surface groups which depends on the pH as it has been found with various semiconductors

(see

e.g.19). The surface of SiC is probably terminated

with Si-OH groups or perhaps also with C-OH groups. It should also be mentioned that the Mott-Schottky curves are shifted upon illumination as observed with many semiconductors (see e.g.20). This result

indicates

that

minority

carriers

are

not

transferred

across

the

interface at a very high rate but that they are trapped in some surface states. This has not yet been further investigated. Hydrogen evolution and anodic decomposition: As proved analytically, SiC is stable under cathodic polarization and the corresponding current is entirely due H2

-formation.

The

shift

of

the

current-potential

curves

with pH (Fig. 6b) amounts to about 60 mV per pH unit which is in accordance with the shift of the Mott-Schottky plots (Fig. 3). The cathodic photocurrent

observed

with

p-type

electrodes

exhibits

a

considerable

overvoltage (Fig. 7) because the onset only occurs at around +1.2 V whereas the flatband potential can be estimated to be at around +2 V (Fig.

2 b).

The

nature

of

the

shoulder

in

the

photocurrent-potential

dependence at p-6H-SiC is not yet understood. The current-potential curves for the n-type 6H-SiC and the 3C-SiC materials are rather similar although with the latter electrode higher photocurrent densities and a steeper photocurrent onset were found (compare Figs. 4 a and b). This may be partly due to the better quality of the cubic crystal. On the other hand the cubic crystal absorbs already at higher wavelengths as shown by the excitation spectrum (Fig. 5). However, the quantum efficiency found with the cubic electrodes in the long wave-

- 15 -

length region is considerably smaller than that found for hexagonal electrodes. Since the energy gap of the cubic crystal corresponds to an indirect transition, the absorption coefficient is expected to be small, which would lead to a relatively large penetration depth of light. Obviously the diffusion length of holes in the epitaxially grown films on highly doped silicon substrates is smaller than the penetration depth of light which causes the low quantum yield. Concerning the onset of the anodic photocurrents for both electrodes, n-(6H-SiC) and n-(3C-SiC) a considerable overvoltage was found (compare Fig. 4 with Figs.2 and 3). We will return to this effect below. The analysis of the oxidation products during anodic current flow in the dark or upon illumination has clearly shown that no oxygen was formed but

only

corrosion

takes

place.

In

solutions

without

HF

white

silicon

oxide layers were formed. It has been suggested by other authors 2 1 an eight hole process for the dissolution as given by: SiC + 8 h+ + 4 H2 O ----> SiO2 + CO2 + 8 H+ .

(2)

This reaction describes the formation of CO2 . The production of CO can only be explained by a 4-hole process such as: SiC + 4 h+ + 2 H2 O ----> SiO + CO + 4 H+ .

(3)

If both reactions proceed at the same rate the resulting net reaction is given by: 2 SiC + 12 h+ + 6 H2 O ----> SiO2 + SiO + CO2 + CO + 12 H+ (4) i.e. a 6-hole reaction per SiC molecule is assumed. Recently, J. Shor found experimental evidence for a 6-hole corrosion reaction by comparing the consumed charge with the amount of material etched away from

- 16 -

the surface 10. These experiments were conducted with cubic 3C-SiC electrodes

whereas

we

used

the

hexagonal

6H-modification.

Thus

far,

however, the chemistry of both modification must be assumed to be very similar

18

.

Another interesting finding is the formation of an amorphous layer of SiC on the surface of the SiC electrodes when anodically etched in concentrated HF solutions. This behavior is very similar to the formation of porous Si on the surface of silicon electrodes etched under similar conditions

22

by

similar

a

. A paper by Brander and Boughey suggests the formation of SiC mechanism.23 Such a reaction, however,

disproportionation

would be very unlikely because it involves the disproportionation of the free species such as CF2 and SiF2 to form SiC, SiF4 and CF4 . More likely is the gradual braking and forming of bonds as illustrated in Fig. 12. Recent SEM investigations by J. Shor have shown that these layers have a porous texture 24 and that their formation can indeed be compared with the formation of porous silicon. All these investigations have proven that no oxygen is formed although SiC fulfills all energetic conditions, such as width of bandgap and positions of the conduction and valence bands with respect to the reduction potential of the H2 O/H2 couple and the oxidation potential of the H2 O/O2 system,

respectively,

splitting.

These

towards

oxidation

as

findings

required reflect

for

the

the

photoelectrochemical

thermodynamic

instability

water of

SiC

(∆ G= -1180 kJ/mole for the reaction SiC + 2O2 -->

CO2 + SiO2 ) and show that the well known chemical stability has only kinetic reasons. On the other hand, this instability toward anodic corrosion

offers

conditions

the rather

possibility than

of

using

etching

SiC

electrochemically

molten

salts

at

very

high

under

mild

temperatures

- 17 -

which for device fabrication will be of great importance. This has been demonstrated by Shor and Kurtz who produced nice etch patterns by photoelectrochemistry

11

.

The question arises whether a photoelectrolysis cell, consisting of a ptype SiC and a Pt counter electrode could be made, in which H2 is formed at the p-type electrode and oxygen at the Pt- electrode during illumination. Although the thermodynamic requirements are fulfilled (Fig. 7), not a sufficient photovoltage can be developed because of the rather large overvoltage for the onset of the photocurrent at the p-type electrode. Accordingly, such a cell could only be operated by applying an external bias. Charge transfer processes with redox systems: According to our experimental results there are a number of redox systems which can be reduced by an electron transfer from the conduction band to the oxidized species of the redox couple. This can already be concluded from the currentpotential (Fig. 9 a-c)

curves since

which the

were

obtained

corresponding

with

currents

n-type occur

SiCat

less

electrodes cathodic

potentials than the H2 -evolution. From this result it can be concluded that the rate constants for these redox reactions are considerably higher than that for the reduction of water. Since this finding is in agreement with common charge transfer models it may not be worth mentioning. However,

with

several

other

semiconductor

electrodes,

such

as

for

instance n-GaAs which is much better defined than SiC, this result is not obtained 25,29. From this point of view we expected SiC to be a very suitable semiconductor for studying basic charge transfer processes in more detail. For instance, according to the usual charge transfer model it is required that the cathodic current should be proportional to the concen-

- 18 -

tration of the electron acceptor in the solution and to the electron density at the surface, i.e. j c-

= - kc- cox ns = - kc- cox nso exp

( eUE -

eUredox

)

/ kT .

(5)

in which j-c is the cathodic exchange current density via the conduction band, k-c the corresponding charge transfer rate constant, cox the concentration of the oxidized species at the electrode surface, nso the electron density

at

the

semiconductor

surface

at

the

redox

potential,

UE and

Uredox the electrode potential and the electrochemical potential of the redox couple, respectively. The current is essentially determined by ns and the rate constant kc- , the latter depending on the reorientation energy. Since a relatively large overvoltage with respect to the redox potential was found for the reduction of Fe3+ we assume at present that the rate constant is rather small for this process. Accordingly, the current should increase by one order of magnitude when the electrode potential is varied by 60 mV. Unfortunately, an evaluation of the current-voltage curves in Fig. 9 yielded for all examples a much weaker current increase. It is not sure whether this is due to an oxide layer across which some of the externally applied potential occurs, or whether the flatband potential shifts in the range of the cathodic current flow as it has been found for GaAs.27 In principle, this can be analyzed by using impedance spectroscopy. Since, however, the capacity drifts considerably, it was impossible to obtain reliable impedance data. Similar

observations

were

made

for

the

oxidation

of

redox

systems.

However, a p-type SiC electrode changes irreversibly if polarized so far anodically that the corrosion current considerably exceeds the diffusion current of the redox system. Here, the oxide formed in the corrosion

- 19 -

reaction,

inhibits

the

redox

reaction

(Fig. 11).

Although

the

spongy

nature of the oxide layer still permits a current flow, an electron donor such as Cu+ or Br- obviously cannot diffuse through this oxide layer. An exception is the oxidation of I– with which no hysteresis occurred even after strong anodic polarization. This is probably caused by the strong adsorption of the I2 / I -

on the SiC surface which seems to prevent effi-

ciently the oxidation of these adsorption sites.

Conclusion We have shown that in spite of the relatively large bandgap and the favorable position of the band edges at the surface neither with n- nor with

p-type

should

be

investigations

SiC

electrodes

mentioned with

that

SiC

H2 O we

-photoelectrolysis

have

particles

also

loaded

performed with

was

detected.

some

metals

It

preliminary

and/or

metal

oxides as catalysts. Also in this case no photoelectrolysis was observed. Interestingly, however, fuels such as methane can be produced in ethanol containing solutions. The mechanism is not yet understood. Further studies with SiC monograin membranes in which the particles are geometrically fixed and on which a catalyst can be deposited in a defined way, may help to solve the reaction mechanism.

- 20 -

Acknowledgments: The authors are indebted to Prof. Dr. B. Kastening, Hamburg, for many valuable discussions and support. We also thank Marion Bahnemann and Ursula Neff for performing some of the electrochemical measurements. The supply of SiC-samples by various companies and institutions listed in Table 1 is acknowledged. This work was financially supported by the German Federal Minister for Research and Technology (BMFT) under contracts No. 329081A and 329580 as well as by the Volkswagen Foundation under contract number I/72 365.

- 21 -

References 1.

R. Memming, Photochem.

Photobiol., 1 6 , 325 (1972).

2.

M. Gleria and R. Memming, J. Electroanal. Chem., 6 5 , 163 (1975).

3.

Y.V. Pleskov and Y.Y. Gurevich, in Modern Aspects of Electrochem-

i s t r y , Vol.16, B.E. Conway, R. E. White and J. O. Bockris, eds., p. 189, Plenum Press, New York 1985. 4.

M. D. Krotova and Y. V. Pleskov, Soviet

Electrochemistry, 1126

(1981). 5.

R. L. Cook, R. C. MacDuff and A. F. Sammells, J. Electroanal. Chem., 1 3 5 , 3069 (1988).

6.

S. Yamamura, H. Kojima, J. Ioda and W. Kawai, J.

Electroanal.

Chem., 2 2 5 , 287 (1987); J. Electroanal. Chem., 2 4 7 , 333 (1988). 7.

S. Mirsagatow and M. Duisenaev, Geliotechnika , 3 , 8 (1977).

8.

T. Inoue and T. Yamase, in 4th Int. Conf. of Photochemical Conver-

sion and Storage of Solar Energy, Book of Abstracts, J. Rabani, ed., p. 37, Jerusalem (1982). 9.

T. Inoue and T. Yamase, Chem. Lett., 7 , 869 (1985).

10.

J. S. Shor, I. Grimberg, B.-Z. Weiss and A. D. Kurtz, Appl.

Phys.Lett.,

6 2 , 2836. (1993). 11.

J. S. Shor and A. D. Kurtz, J. Electrochem. Soc., 1 4 1 , 778 (1994).

12.

H. Daimon, M. Yamanaka, E. Sakuma, S. Misawa and S. Yoshida, Jap.

J. Appl. Phys., 2 5 , L 592 (1986). 13.

I. I. Creaser, R. Geue, J. M. Harrowfield, A. Herrit, A. M. Sargeson, M. R. Snow and J. Springborg, J. Am. Chem. Soc., 1 0 4 , 6016 (1982).

14.

J. O. Bockris and K. Uosaki, J. Electrochem. Soc., 1 2 4 , 1348 (1977).

15.

Y. Hori, K. Kikuchi, A. Murata and S. Susuki, Chem. Lett., 897 (1986)

16.

H. Morisaki, H. Ono and K. Yazawa, J. Electrochem. Soc.,1 3 1 , 2081 (1984)

- 22 -

17.

A. Mannivannan, A. Fujishima and G. V. S. Rao, Ber. Bunsenges. Phys.

Chem., 9 2 , 1522 (1988). 18.

I. Lauermann, D. Meissner, R. Memming, R. Reineke and B. Kastening in Dechema-Monographie, chaft,

19.

Frankfurt

Vol.

124,

p.617,

VCH

Verlagsgesells-

(1991).

R. Memming in Comprehensive

Treatise

of

Electrochemistry, Vol.

7, B. E.Conway and J. O. Bockris eds., p. 529, Plenum Press, New York (1983). 20.

R. Memming in Topics

of

Current

Chemistry, Vol.169, M. Mattey,

ed., p.105, Springer-Verlag, Berlin (1994). 21.

H. Hirayama, T. Kawabuko, A. Goto and T. Kaneko, J. Am. Ceram.

Soc., 7 2 , 2049 (1989). 22.

R. Memming and G. Schwandt, Surf. Sci., 5 , 109 (1966).

23.

R. W. Brander and A. L. Boughey, Brit. J. Appl. Phys., 1 8 ,

905

(1967). 24.

J. S. Shor, J. Grimberg, B. Z. Weiss and A. D. Kurtz in Meeting of the

Electrochem. Soc., Abstract No.634, Honolulu (1993). 25.

D. Meissner, Ch. Sinn, R. Memming, P. H. L. Notten and J. J. Kelly in

Homogeneous and Heterogeneous Photocatalysis, E. Pelizzetti and N. Serpone, eds., p.343, D. Reidel Publ., Dordrecht (1986). 26.

R. Memming, Ber. Bunsenges. Phys. Chem., 9 1 , 353 (1987).

27.

I. Uhlendorf, R. Reineke-Koch and R. Memming, J.Phys.Chem.,in press

28.

R. Memming in Topics in Current Chemistry, Vol.143, E. Steckhan, ed., p. 79, Springer Verlag, Berlin (1988).

- 23 -

Table 1: Crystal sources

source

crystal

doping

doping

type

density

type

color

_________________________________________________________ 1

3

4

5

6H

high

n

green

6H

high

p

blue

low

p

blue

6H

high

n

green

6H

high

p

blue

3C

10 1 7 - 1 0 1 8 cm- 3

n

yellow

3C

10 1 6 - 1 0 17 cm - 3

n

yellow

3C

high

n

dark

6H

high

n

green

_______________________________________________________ 1

Elektroschmelze Kempten GmbH, München.

2

Siemens AG, Zentrale Forschung und Entwicklung, München.

3

Showa Denko, Ceramics Div., Tokyo.

4

Dr. R. F. Davis, North Carolina State University, Dpt. of Material Science and Engineering, Raleigh, NC.

5

Sharp Corporation, Tokyo.

- 24 -

Table 2:

Values of flatband potentials in Volt (SCE) for n- and p-(6HSiC) as reported by different author

n-SiC

pH

p-SiC

pH

Reference

________________________________________________________ -1,6

1.3

2

+1.44

0

14

+1.14

14

14

13.0

16

-1.45 to -1.65 -0.7

1

+1.6

0.3

9

-0.96

0.3

+1.52

0.3

17

-1.66

13.0

+1.3

13.0

17

- 25 -

Figure Fig. 1:

Captions

Contacting methods for 3C-SiC on Si substrates, a) via SiC; b) via Si; c) via SiC with subsequent removal of the substrate by etching

Fig. 2:

Mott-Schottky plots for 6H-SiC electrodes in acid and alkaline solutions. a): n-type b) p-type Measuring frequency; 20 kHz; scan rate 0.05 V/s

Fig. 3:

Mott-Schottky plot for an n-type 3C-SiC electrode at various pH-values

Fig. 4:

Current-voltage dependence for a n-6H (a) and a n-3C-SiC (b) electrode

under

intermittent

light

excitation

at

pH 7;

N2 -

purged; scan rate: 50 mVs- 1 Fig. 5:

Photocurrent excitation spectra of 6H- and 3C-SiC electrodes in 1 M NaOH at UE= + 1 V

Fig. 6:

a)Current-voltage curves at an n-type 3C-SiC electrode with respect to the reduction of H2 O at different pH-values. b) Potentials at which a current of 50 µ A was measured vs. pH

Fig. 7:

Current-voltage dependence for a p-type 6H-SiC electrode in a 40% HF solution

- 26 -

Fig. 8:

Anodic photocurrent vs. time at an n-type 6H-SiC electrode at pH 4,

electrode

potential:

+1

V,

illumination:

white

light

(0.2 Wcm - 2 ) Fig. 9:

Reduction of various redox systems at an n-type 3C-SiC electrode in 1 M HCl; scan rate: 10 mVs- 1

Fig. 10:

Reduction of CO2 at an n-type 6H-SiC electrode loaded with Cu in the dark in 0.1 M Na2 SO4 solution. Scan rate: 0.1 Vs- 1 Insert: reduction current vs. time at UE= -1.35 V

Fig. 11:

Current-voltage curves for a p-type 6H-SiC electrode at pH 1 in the presence of 25 µM Cu1 + , 1000 rpm; scan rate: 0.1 Vs- 1

Fig. 12:

Model for the rearrangement of the surface layer during anodic polarization

in

HF-solutions

- 27 -

a

a

In In

epoxy

epoxy

ccontact ontact

SiC SiC

SSii

b

In

contact

SiC

Si

c

In

epoxy contact

SiC

Figure 1

- 28 -

[10 - 1 2cm 4 / F 2 ]

8

- 2

a

1 0

2

pH 14 6

pH 0

C

4

0

- 2 -1.5 - 1 -0.5 potential

0 [V

0.5 vs

1

1.5

2

SCE]

b

Figure 2

[10 - 1 2cm 4 / F 2 ]

- 29 -

pH 14

C

- 2

pH 10

pH 4,7 pH 0

- 2

- 1

0 1 2 potential [V vs SCE]

3

Figure 3

b [ µ A / cm 2]

current

[ µ A / cm 2]

a

current

- 30 -

250

500 1 2

1

2

Figure 4

0.24

1

0.2

0.8

0.16 n-6H-SiC

n-3C-SiC

0.6

0.12

0.4

0.08

0.2

0.04

0

0 -0.04

-0.2 300

[ µ A / cm 2]

current

1.2

current

[ µ A / cm 2]

- 31 -

380

460 wavelength [nm]

540

620

Figure 5

- 32 -

a

potential [V vs SCE]

current

pH 14

[ µ A / cm 2]

pH 4.7 pH 6.8

pH 2

pH 0

pH 10 pH 11.5

b

Figure 6

- 33 -

current

[ mA / cm 2 ]

0.25 0.2 0.15 0.1 0.05 dark

0 -0.05

illuminated

-0.1 -0.15 - 1

-0.5

0

0.5

1

1.5

2

Figure 7

- 34 -

3.8

current

[ µ A / cm 2]

3.4 3 2.6 2.2 1.8 1.4 1 0

2000

4000 time [s]

6000

Figure 8

- 35 -

a

0.5

-0.5 [V vs SCE]

3 -

16.7 mM [Fe(CN) 6

]

current [ µ A/cm 2]

potential

-500

b -0.5

potential

[V vs SCE]

Fe 3 +

-1000

[ µA / cm 2]

4.7 mM

current

- 1

c potential

current

- 1

[V vs SCE] -0.5

0.4 mM

1 mM

Co(sep.) 3 +

[ µA / cm 2]

-100

Figure 9

- 36 -

N2 CO2

-0.4 -0.6 -0.8 -1.6

[ µ A / cm 2 ]

-0.2 -0.05

-0.1

current

current

[ µ A / cm 2 ]

0

-0.15

stop purging CO 2

CO -0.2

0

100

200

300

time

400

500

600

[s]

-1.2 -0.8 -0.4 potential [V vs SCE]

0

Figure 10

- 37 -

a) disc current

0 0 0

without Cu

current

[ µA ]

-0.8 -1.6

with Cu -2.4 -3.2

b) ring current - 4

-0.8

0 0.8 1.6 potential [V vs SCE]

2.4

Figure 11

- 38 -

F C C

F

C

Si

C

C

Si

C

C

Si

C

C

Si

C

F F

F

Si

C

C

Si

C

C

Si

C

F OH

C

Si

C

C

OH

C

F

C

F F Si

F

F

F F

F C

F

F

F F F F F C

C Si

F F C F

C

F

Si F C

F OH

Si

C

C

OH

F

F

F

F

F

F C

F

F

F Si

C

F OH

F

Si

C

C

OH

F

Figure 12