Sie sind auf Seite 1von 8

IL T

ELSEVIER
Journal of Electroanallytical Chemistry 396 (1995) 219-226

,IOtll~AI. OF

A generalized theory of intensity modulated photocurrent spectroscopy (IMPS) 1


E.A. Ponomarev 2 L.M. Peter
School of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, UK
Received 20 January 1995; in revised form 27 April 1993

Abstract

A generalized theoretical model of intensity modulated photocurrent spectroscopy (IMPS) is proposed, and analytical equations for the frequency dependence of the photocurrent and photovoltage responses are derived. The theory treats the influence of the space charge, the surface state and the Helmholtz capacitances on the IMPS response. In the model, charge transfer and recombination at the semiconductorlelectrolyte interface are described by kinetic equations, whereas surface state charging and the current through the external circuit are treated using linear circuit equations. The calculated IMPS plots exhibit a number of useful diagnostic features. The treatment also shows that it should be possible to measure the space charge capacitance directly from the photovoltage response, even if it is comparable in magnitude with the Helmholtz capacitance. Equivalent circuits representing the processes occurring during IMPS measurements are discussed, and it is shown that single passive elements do not provide a satisfactory description of the kinetic steps responsible for the photocurrent response.
Keywords: Mathematical model; Photochemistry; Doped semiconductors; Microwave reflectivity; Intensity modulated photocurrent spectroscopy

1. Introduction Non-stationary techniques have made major contributions to the understanding of the kinetics and mechanisms of elementary processes involved in photoelectrochemical reactions at the semiconductor ]electrolyte interface. The most common types of external perturbation involve changes of illumination intensity or electrode potential, and various equivalent electrical circuits have been proposed to explain the response [1-12]. One of the most powerful of these techniques is intensity modulated photocurrent spectroscopy (IMPS), which can be used to measure the rate constants of charge transfer and recombination. The theory of this method has been developed elsewhere [12-15], but there are a number of important features that have not been considered so far. The present paper therefore re-examines the basic theoretical framework for IMPS and explores the more general case where

Dedicated to Professor Kenichi Honda on the occasion of his 70th birthday. 2 Permanent address: Institute of Energy Problems of Chemical Physics (Branch), Russian Academy of Sciences, 142432 Chernogolovka, Moscow Region, Russia. 0022-0728/95/$09.50 1995 Elsevier Science S.A. All rights reserved SSDI 0 0 2 2 - 0 7 2 8 ( 9 5 ) 0 4 1 15-X

Csc and C a are of similar magnitude. The diagnostic features of the calculated IMPS responses should extend the range of application of the technique. Earlier derivations of the IMPS response have been based on the assumption that the space-charge capacitance Csc is much smaller than the Helmholtz capacitance C H. While this is a good approximation for many semiconductor electrodes, it is not appropriate for heavily doped semiconductors such as titanium dioxide. A second complication which has not been considered in detail in the theory of IMPS is the fact that many semiconductor lelectrolyte interfaces exhibit 'Fermi level pinning' due to the electronic charge accumulated in surface states. In these systems, the potential dependence of the stored surface charge defines a surface state capacitance Css which changes the potential distribution. The influence of the dynamics of surface state charging on the modulated photocurrent response has not been considered previously, although de Wit et al. [16] have discussed the effect of charged photodecomposition intermediates on the IMPS response of CdS. The generalized theory developed in the present paper is important not only for IMPS but also for a powerful new technique which involves measuring the microwave reflectivity of the illuminated semicon-

220

E.A. Ponomarev,LM. Peter~Journalof Electroanalytical Chemistry396 (1995) 219-226 Css Rss

ductor lelectrolyte interface under modulated illumination [17].

2. Analytical model
2.1. Preliminary considerations

Csc

-41
R Fig. 1. The electrical circuit used in the theoretical IMPS analysis to represent the transfer of charge through the external circuit and the charging of surface states. Note that the circuit does not represent the internal processesof chargerecombinationand interfacialcharge transfer. These processesare describedby kineticequations.

Illumination of a semiconductor electrode generates electron-hole pairs that are separated rapidly within the depletion layer. Subsequently, the majority carriers move to the back contact, whereas the minority carriers move to the semiconductor[electrolyte interface where they are removed by either charge transfer or surface recombination (here we neglect recombination in the space-charge region). Unless the illumination intensity is low, the redistribution of photogenerated charge associated with these processes may change the band bending, altering the space-charge capacitance and the concentration of majority carders within the depletion layer. The photo-induced changes can give rise to additional contributions to the current flowing through the external circuit, making the IMPS response more difficult to interpret. Therefore expressions are required that allow estimation of the change of band bending brought about by illumination. In practice it is desirable to reduce it to an acceptably small value in order to simplify the IMPS response. In this paper we derive expressions for the photocurrent response of the semiconductor [electrolyte junction to periodic illumination of the electrode under reverse bias conditions. The process of charge separation in the space charge region is not considered since it is very fast (in the picosecond region) [3,4]. The approach is also simplified by the assumption that light is mainly absorbed within the depletion layer, so that the diffusion of minority carders towards the depletion layer from the bulk of semiconductor can be neglected. Electron-hole pair photogeneration followed by charge separation within the double layer leads to an additional potential drop across the depletion layer which tends to change the electrode potential. Under potentiostatic conditions, the potentiostat reacts to maintain constant potential difference between the bulk of the electrode and electrolyte by driving electric current through the external circuit. In practice, the speed of this control response is limited by the bandwidth of the potentiostat and the RC time constant of the electrochemical cell, which is determined by the electrode capacitance and the total series resistance associated with the electrolyte and ohmic contact. Since the accuracy of potentiostatic control is affected by the IR drop, it is essential to minimize series resistance and to restrict measurements to low currents. Even in the absence of IR effects, the photoinduced charging of the depletion layer causes a shift of the Fermi level which leads to charging of the surface state capacitance (note that this charging process is not the same as the

accumulation of charged intermediates in multistep photodecomposition reactions as considered in [16]). The rate of this charging process is proportional to the photo-induced potential drop across the depletion layer, and therefore it can be represented as a charging of the surface state capacitance Css through a series resistor Rss. It follows that the transfer of charge through the external circuit and charging the surface state capacitance can be represented by the equivalent circuit shown in Fig. 1. However, as we show in Section 3, the internal processes of charge recombination and charge transfer across the interface cannot be represented satisfactorily by single passive linear electric elements. Instead, they are best described by rate constants and concentration terms. In fact, detailed considerations of energy conservation show that the first-order rate constants for recombination and charge transfer can only be represented in principle by equivalent 'RC time constants' if the external resistance is negligible. In any case, the R and C values have no physical significance on their own (the reader is referred to the work of Vanmaekelbergh and coworkers [18-20] for a discussion of the impedance elements describing surface and space-charge recombination). If the light intensity use for IMPS measurements is sufficiently low, the band bending, space-charge capacitance and density of majority carriers are not changed significantly from their values in the dark. Provided that charge transfer, recombination and capture by surface states are linear processes, they can be described by first-order kinetic equations, greatly simplifying the theoretical treatment. In general, this implies that we are considering the low intensity limit of the IMPS response. Similarly, the magnitude of the ac photovoltage response is considered to be sufficiently small that the density of majority carriers is almost constant. This condition may be more difficult to guarantee experimentally than in the IMPS case. In the discussion that follows, consideration is limited to n-type semiconductors for the purposes of illustration.

E.A. Ponomarev, L.M. Peter/ Journal of Electroanalytical Chemistry 396 (1995) 219-226

221

2.2. Direct charge transfer and recombination via the valence band
Consider the situation where photogenerated holes can transfer into solution across the interface directly via the valence band with a rate constant k~ or else recombine with electrons with a rate constant k 2 (Fig. 2). We assume that, at low levels of illumination, k 1 and k 2 are independent of light intensity. It is also assumed that the surface state capacitance is negligible. Under these conditions, the appropriate differential equations for the photogenerated charge Qsc across Csc and for the photogenerated charge QH across C H are dQsc dt = I(1 + ei'') - j - k 2 ( Q s c dQr~ -= k l ( Q s c - Qn) - J dt Qn) (1) (2)

for the photocurrent response to pulsed illumination, and their solution gives the expression for the current through the external circuit as

k, +ito(C/Csc) (
J=l kl + k z + i t o

1
w------~ 1+ i

ei~t-q-/0--

kl (4)

kl + k 2

where
C=

Csc Cn Csc + C.

(5)

Here, ~"= RC is the time constant of the electrochemical cell (R is the total series resistance arising from the electrolyte, the semiconductor bulk and the ohmic contact). The ac component of the photocurrent is

kl + ito(C/Csc) ( j(w)=lo
k,+k2+ito

1
1 +itor

)
(6)

where Io is the amplitude of the photogenerated hole current towards the surface, to is the light modulation frequency and j is the current through the external circuit caused by illumination. Eq. (1) shows that the photogenerated charge across the depletion layer increases owing to the minority carrier flux to the surface and decreases owing to recombination and discharge through the external circuit. Similarly, Eq. (2) shows that the photogenerated charge across the Helmholtz capacitance increases as a result of charge transfer across the interface and decreases owing to discharge through the external circuit. The rates of charge transfer and recombination are proportional to the minority carrier concentration at the surface, which equals (Qsc-QH). The fact that the potentiostat maintains a constant potential across the system gives rise to the third equation asc Csc

Q, CH

=jR

(3)

Eqs. (1)-(3) resemble those considered previously [ 1,21 ]

. o

EF-I

k2 i ,.~ kl

In the limiting case where Csc << C H, Eq. (6) is identical with the expression obtained in [13]. If the time constant r is small compared with the time constant associated with charge transfer and recombination (i.e. r << l/(k I + k2)) , Eq. (6) shows two semicircles in the complex plot of Im[j(to)] vs. Re[j(to)] as a function of frequency o~. In the low frequency limit, the plot tends towards the real axis at the point j(to) [ ,o-~ 0 = lokt/(kl + kz). As the frequency increases, the imaginary component of the photocurrent passes through a maximum at a frequency tOma = k 1 + k 2 before decreasing to intersect the x real axis at loCrt/(Csc + Cn) (note that in the limit C H >> Csc the intercept is located at I 0 as expected). At this point in the IMPS plot the frequency is large in comparison with (k I + k 2) but still low in comparison with r - J. At higher frequencies the IMPS plot exhibits a second semicircle in the lower quadrant arising from the RC time constant of the cell, and the response tends towards the origin on the complex plane plot as the photocurrent becomes completely attenuated. The general shape of the IMPS response discussed above is very similar to the form derived previously, i.e. the semicircles arising from charge transfer/recombination and from RC attenuation appear in the upper and lower quadrants of the complex plane respectively. However, Fig. 3 shows that if the depletion layer capacitance is large and charge transfer is fast, the IMPS plots develop two semicircles in the lower complex plane. This kind of behaviour should be an interesting diagnostic feature of highly doped semiconductor photoelectrodes.

2.3. The intensity modulated photovoltage response


Fig. 2. The simplest model of the elementary processes of charge transfer and recombination taking place at the illuminated semiconductor 1electrolyte interface under reverse bias.

IMPS experiments are normally designed so as to minimize the cell time constant ~" as far as possible in order to increase the range of rate constants which is accessible to measurement. In practice, the upper limit to the rate con-

222

E.A. Ponomarev, L.M. Peter/Journal of Electroanalytical Chemistry 396 (1995) 219-226


I I ' I '

0.4 . . . . .

~.....
~:~--"~ /

+.....
I "~'~'--(3 "

i. . . . .

Io 11 4-~--~
IV

<-/',A,_I
0.2

"%

I0 51
-{
__L

41

- < q 4 f - - - - -'-" ~-U~- ~- . . . . .

l----'q.---4 ....
_

"~

0.0 . . . . .

~.....

a__

. . . .

--r. . . . .

-- . . . . . . . . . .

-0.2

-~

7 - - f -~k-/~z---- - o j ~ v<,.,,. ~ \ ~ jo <,vl,,,,. \~ _jo I


I
I

- --i - ~5-kz I I
] I I
I I

-0.4

..... ~__Z_

7 I I

-I I I

I I

It is interesting to note that the high frequency intercept is inversely proportional to the depletion layer capacitance even if Csc << C H. This means that it should be possible to measure the space-charge capacitance from photopotential measurements under periodic illumination. In contrast, the conventional method of measuring Csc by potential modulation cannot give the information about Csc alone because the potential perturbation is applied to Csc and C n connected in series, i.e. the measured capacitance is given by Eq. (5). The advantage of the intensity modulated photovoltage technique is that the ac perturbation (light in this case) influences only the depletion layer, allowing deconvolution of Csc.

0.0

0.2

0.4

0.6

0.8

.0

2.4. Photo-induced changes in band bending


In the derivation of the IMPS expressions, it has been assumed that the band bending remains constant during modulated illumination. In practice, this is always an approximation to the real behaviour of the photoelectrode. It is important to establish that the photoinduced changes in band bending are sufficiently small to allow use of the linear equations developed in the present treatment. The excess charge due to minority carriers at the surface is Qsc- QH, and it can be easily found from Eqs. (1)-(3):

Re(j ( co)II o)
Fig. 3. The IMPS responses calculated f r o m Eq. (6) with k, = 10 s - I , Csc = 4 x 10 - 6 F, C n = 10 - 5 F, R = 10 ~'~, a n d (1) k I = 0, (2) k I = 10 s - 1 , (3) k I = 5 0 s - 1 , (4) k 1 = 100 s -1 and (5) k I = 2 0 0 s - l . Note that the IMPS plots develop t w o semicircles in the l o w e r q u a d r a n t w h e n charge transfer is fast a n d Csc a n d C n are o f similar magnitude.

stants that can be determined is set either by the resistance associated with the solution, the semiconductor electrode and its ohmic contact, or by the frequency response of the potentiostat. However, there is an interesting possibility of measuring very fast processes by avoiding potentiostatic control. Instead, r can be increased deliberately by adding a large series resistor R in the external circuit. In this case, at frequencies to >> r - I , the measurement occurs under coulostatic rather than potentiostatic conditions. The ac component of photopotential across the large resistor R is U(to) =j(to)R. Provided that the dc and ac photopotentials are of the order of a few millivolts or less, k ! and k 2 can be considered as approximately constant, and an expression for the product toU(to) in this low intensity limit is easily derived from Eq. (6):

Qsc - Qn =

el'' + ~ k I + k 2 + i to k I + k2

I0

10

(9)

I I
0.4
. . . . . . . . .

I I
. . . . .

I I
,
. . . . .

I o I

~/

i ~..--c~A_

t,,

1o kl + ito( C/Csc)
toU(to) = ~k i -#- k 2 + i to (7)
~,

i o

=o

oo, ....
-0.2 ,x "D,. -"~----7 .....

The complex plane plot of toU(to) exhibits a semicircle similar to that seen at low frequencies under potentiostatic conditions, although it is of course scaled by lo/C. The maximum of the semicircle is located at to,~x = (k~ + k2). The low frequency intercept occurs at coU(to) I <o-., o =
lo
C

I I I

i I i ~- . . . . .
\1
I

i " , ~ / , ~ [ " ~ V I J O ly O - / g T - ~ . . . . . .

,
",~ l

:oo

,
i
I

XJ" I
I

',

/q
kI + k2

(8a) 0.0 0,2 0.4 0.6

7
0.8 1.0

The high frequency limit of toU(to) is given by


tou(to) I ~,.-, = =

ReO((o)li o)
(8b)
Fig. 4. The IMPS response calculated f r o m Eq. (19) with k 2 = 5 0 s - I , C s c = 1 0 - 7 F, C H = 1 0 - 5 F, Css = 5 1 0 - 6 F, "rss = 3 1 0 - 4 s, R = l 0 O , a n d ( l ) k I = 0, (2) k I = 2 0 s - i, (3) k t = 5 0 s - l, (4) k t = 2 0 0 s - 1 and (5) k 1 = 500 s -1 .

Io Csc

so that k l, k 2 and Csc can be obtained.

E.A. Ponomarev, L.M. Peter~Journal of Electroanalytical Chemistry 396 (1995) 219-226 I I


4- . . . . .
I

223

I
I

Im
I

0.4

.....

d . . . . .

I I

I [

I. .
I i
.

. . .

IA i

.....

1
2

2.5. The influence of high surface state capacitance in the case of direct charge transfer and recombination via the valence band
The separation of photogenerated charge across the depletion layer changes A~osc. In the presence of surface states, this change causes charging of the surface state capacitance. If this process is linear and proportional to the photoinduced potential drop across the depletion layer, it can be described as a charging of surface state capacitance Css through a resistor Rss. The corresponding differential equations are dQsc =10( 1 + ei't) - J - J s s - k2(Qsc - Qa) dt dQa dr = k ' ( Q s c - Qa) - J Qsc Qa + =jR Csc Ca dQss dt =Jss
--

!
. . . . .

I
I
. . .

1 kHzl

02

.....

0.0 . . . . .

~......

I-'' / ~ - - 2-~!.-m- 4

-o.2

-%----,

~.~

~,'..~

' \ "~ -",L,


1-

.....

200 IkHz
\ I

~-. . . . .

~. . . . .
I
I I

\ ~ \l

~
I I

~/~ i ~ I

r( I

~. . . . .

-0.4
0.0

I
I
I ~

~
I I ~

....
0.4 0.6

2.....
[ I ,

(14) (15) (16)

0.2

0.8

1.0

ReG(o)/I o)
Fig. 5. The IMPS response calculated from Eq. (19) with k] = 3 0 0 s -~ , k 2 = 50 s - 1, Cs C = 10-7 F, C H = 10 -~ F, Css = 5 X 10 -6 F, ~'ss = 3 X 10 -4 s, and (1) R = 0 ,O and (2) R = 10 ,O.

(17)
(18)

The charge Qsc at low frequency (tot<< 1) is given by Csc ( Io el,O,+ 1o ) Qsc = Csc + CH k I + k 2 + i to k I + k----"~ or
Csc

Qss Qsc -Cs s +Jss Rss Cs c

(10)

where Jss is the current due to charging the surface state capacitance. The solution of Eqs. (14)-(18) is j(to) = 10{kl -I- i to( C/Csc)}{( k I + k 2 + i to)( 1 + i to'r)

Qsc =

Csc + C H

(Qsc - QH)

(11)

+ [Css/(Csc + Ca) ] [ito/(1 + ito~-ss)]


X[1 +RCH(k I + i t o ) ] } -l vhere rss = RssCss.
o ] I -~----+ ..... 2~ zx 3 I v , I

If the space-charge capacitance is small (Csc .~c CH) the charge Qsc across the depletion layer (which is the charge of the excess majority carriers at the edge of the depletion layer) is much smaller than the charge due to minority carriers at the surface: Csc Qsc = -~H (Osc - Oa) << (Qsc - OH) (12)

(19)

0.4

I *13Hlr I \~ 4-. . . . . ~ . . . .

b ~L_4x,~ ~ t I ,,OJ"t~'---h ICr" I

I ' I 4.....

The photogenerated excess of majority carriers at the edge of the depletion layer in the bulk semiconductor changes the potential difference across the space-charge region. If the perturbation of k I and k 2 is small, the value of the ac component of the potential change A(Psc is given to a good approximation by &Psc -- Csc Ca k~ + k 2 + i to

0.2 . . . . .

L _

"-"

-0.2

--~ .....

Eq. (13) can be used to estimate the change of band bending under illumination and to adjust the light intensity so as to minimize the effect. For example, if C a = l0 -5 F ' ~ m - 2 , *to = 1 0 - t A c m - 2 and (ki + k 2 ) = 102 s - l , A t p = 1 mV in the low frequency limit. Clearly, this estimation method is only valid if A~osc is a few millivolts, since the dependence of k 2 and k 1 on band bending is not considered.

-0.4
I

0.0

0.2

0.4

0.6

08

.0

Re(j(o )/Io)
Fig. 6. The IMPS response calculated from Eq. (19) with kj = 2 0 s - I , k 2 = 5 0 s -1, Csc = 1 0 -7 F, C H = 1 0 -5 F, Css = 5 X 1 0 -6 F, R = 1 0 1'~ and (1) l"ss = 0 . 1 s, (2) 7"ss=10 -3 s, (3) ~ ' s s = 2 1 0 -4 s and (4) ~'ss = 5 X 10 - s s.

224

E.A. Ponomarev,L.M. Peter~Journalof ElectroanalyticalChemistry396 (1995) 219-226

The low frequency photocurrent limit is given by kl j(co) I o,- 0 = 10 k, + k 2 (20)


EF

I
-

and, as before, is determined only by the rate constants k 1 and k 2. The shape of the complex plane plots described by Eq. (19) for the general case is rather complicated to analyse. Some of the main features of the IMPS signatures are illustrated in Figs. 4-7. The analysis is relatively straightforward if charge transfer and recombination occur on a longer time-scale than the charging of the surface states capacitance. In this case, the plot shows a semicircle at low frequencies with intercepts on the real axis at lokl/(k 1 + k 2) and loCH/(C . + Css + Csc), and with tOm~ = (k I + k2). This semicircle arises essentially from x the 'freezing out' of the charge transfer and recombination processes as the frequency of modulation is increased until only the displacement component due to charging remains. At even higher frequencies, the charging of surface states is eventually 'frozen out' as well, and the IMPS plot develops a second semicircle where the photocurrent changes from IoCH/(Crl + Css + Csc) to IoC./(C n + Csc). Finally, at very high frequencies, the plot shows the usual semicircle in the lower quadrant arising from RC attenuation. These considerations show that the surface state capacitance is expected to influence the IMPS response only when its magnitude is similar to that of the Helmholtz capacitance. This contrasts with the situation encountered in impedance measurements, where the condition for surface states to influence the response is that Css should be of comparable magnitude to Csc instead of C . . It follows that undistorted IMPS responses may be obtained even for

f I
] P2 a kl
P

I:

Fig. 8. A more general model of the elementary processes of charge transfer via valenceband and surface states and recombinationvia surface states taking place at the illuminated semiconductorlelectrolyte interface under reverse bias. semiconductor electrodes where conventional Mott-Schottky plots fail to produce correct values of the depletion layer capacitance due to 'Fermi level pinning' by surface states.

2.6. Charge transfer via the valence band and via surface states
The discussion can be extended to consider the situation where the minority carriers can transfer to the solution directly from the valence band and indirectly via surface states. At the same time, minority carriers captured by the surface states can recombine with majority carriers (Fig. 8). To simplify the treatment, it is assumed that the surface state capacitance is negligible. The appropriate differential equations describing these processes are

__ dt

dQsc

= -j-

k4p 2 + I 0 ( 1

+e iwt)

(21) (22)
(23)

I
0.4 . . . . . (3 [] -', 1 2
. .

.....

4-.

I
. . . . ~ . . . . . I

I
+. . . . . I I
I

dQ. d---T-= - j + kl Pl + k3 P2
- -

I t I

Qsc

Qn

=jR

i
. . . .

Csc C. dp~ dt = I ( I + ei'') -Pl(k~ + k2) dp2


--

(24)

0.0 . . . . . .

....

-o.2

-0.4

= k2 P , - P2(ks + k4) (25) dt where Pl is the concentration of minority carders in the valence band and P2 is the concentration of minority carriers captured by surface states. The solution of Eqs. (21)-(25) gives the expression for the ac component of the photocurrent:

k, + ico( C/Csc )
0.0 0.2 0.4 0.6 0.8 1.0 J(co) = I 0 k, + k2 +ico

Re(j( o.))/I o)
Fig. 7. The IMPS response calculated from F-Xl.(19) with kI = 200 s -L, k 2 = 5 0 s - I , C s c = 10 - 7 F, C H = 10-s F, Css = 5X 10-6 F, R = 10 .O, and (1) ~'ss= 0.1 s, (2) Zss = 3 X 10-4 s and (3) ~'ss= 5 X 10-s s.

+(

k2
~kl + k 2 + i c o

)k3+ico(C/Csc)
k3 + k 4 + i c o

1 +icor (26)

E.A. Ponomarev, L.M. Peter/ Journal of Electroanalytical Chemistry 396 (1995) 219-226

225

An interesting feature of Eq. (26) is that it predicts that, for certain values of the parameters, the absolute value of the ac photocurrent can appear to be greater than the photogenerated hole current I 0 (Fig. 9). The reason for this apparently anomalous effect is that the total light induced current through the external circuit contains a contribution from electrons that move towards the surface to recombine with captured holes with rate constant k 4. The time-dependent concentration of captured holes exhibits a phase shift relative to the light modulation owing to the finite value of k 2, and in consequence the total phase shift between the illumination and the electron flux may exceed 90 . Since electrons and holes differ in sign, a phase shift of more than 90 results in a positive contribution of electrons to the photocurrent, i.e. the total current (holes + electrons) becomes greater than the photogenerated hole current I0.

+TIlt "ec,
j( w) =I0
1

"

Fig. 10. An electrical equivalent electric circuit for the analysis of the transient photoresponse of the semiconductor[electrolyte junction. The limitations of the circuit are discussed in the text.

capacitance (Fig. 10). The external resistance is assumed to be negligibly small. Analysis of this circuit shows that the ac photocurrent through the external circuit is given by

3. Equivalent circuits

l l [ Ru( Csc + CH)] + i~[ CHl( C H + Csc)] (l/Rt~ + llRr~)[1/(Csc + CH) ] + iw


(27)

In order to compare the equations for the IMPS response obtained here with the behaviour of an equivalent circuit, consider the electric circuit where the charge transfer and recombination processes are represented by resistors Rt~ and R,ec respectively and the surface state capacitance is negligibly small. It is reasonable to assume that R,~c is connected in parallel with the depletion layer capacitance, but the placement of Rt, encounters conceptual difficulties that have been reviewed recently by Kenyon et al. [11]. It is assumed here that Rtr is in parallel with C n . The process of charge photogeneration is represented by a current generator in parallel with the space-charge

It can be seen that this equation coincides with Eq. (6) (w~" << 1) if it assumed that

k~ = R~(Csc + C.)
1

(28)

k,~< = R,oc(Csc + C . )

(29)

'

~. I~_Z

r ~

0.6 0.4 0.2 0.0 I -0.2 -0.4 -0.6


__~__A t

J-!2 7L
~ --i ~ ~
-r - -

--

I I

t f ~

Is6k~z , i ,
[ i

l I ~

b _~ I.O
J ]

i i J_.~

L_

I I L

io
I

L__ 112 2 IA I

3-

Eqs. (6) and (27)-(29) show that the IMPS response under potentiostatic conditions ( w r << 1) can be described in the framework of the electric equivalent circuit. However, it seems that this approach has little merit because it replaces the rate constants k 1 and k 2 by "resistors of charge transfer and recombination" [7,11,22] that can only have meaning in combination with the real capacitances Csc and C a . Moreover, this approach is not consistent with the conclusions of Vanmaekelbergh and coworkers [18-20] who have shown that the recombination processes must be represented by an impedance, leading to an increased number of elements in the equivalent circuit. The utility of the equivalent circuit approach is questionable in this case. In contrast, k I and k z are clearly related to the elementary steps involved in the photoelectrochemical behaviour of the interface. However, the present treatment of the charging of surface states using an RC equivalent circuit seems justified since the process is assumed to be linear and proportional to the photo-induced potential drop across the depletion layer.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Re(j(o)/Io)
Fig. 9. The IMPS response calculated from Eq. (26) with k ) = 0 s - t , k2=100 s -1, k 4 = 1 0 0 s - t , C s c = 1 0 -7 F, C n = I 0 - s F, R = 1 0 and (1) k 3 = 0, (2) k 3 = 200 s - l and (3) k 3 = 500 s -1.

4. Conclusions

The generalized theoretical model of IMPS outlined here predicts a characteristic response for heavy doped semiconductor electrodes where the magnitudes of the

226

E.A. Ponomarev, L.M. Peter~Journal of Electroanalytical Chemistry396 (1995) 219-226

space-charge capacitance and Helmholtz capacitances are comparable. The expressions obtained for the intensity modulated photovoltage show that it should be possible to investigate very fast processes by measuring the photopotential across a large external resistor under coulostatic conditions. At the same time, the high frequency limit o f photopotential response can be used to measure the spacecharge capacitance o f a semiconductor electrode even if it is comparable in magnitude to the Helmholtz capacitance. The expressions describing the influence o f the surface state capacitance on the IMPS response should prove useful for the interpretation o f IMPS results for electrodes that exhibit Fermi level pinning. The generalized approach can also be extended to other non-stationary techniques, in particular to impedance spectroscopy o f the illuminated semiconductor l electrolyte junction [23,24].

Acknowledgements
EP acknowledges the Royal Society for the award o f a postdoctoral fellowship. This work was also supported by the University o f Bath.

References
[1] R.H. Wilson, T. Sakata, T. Kawai and K. Hashimoto, J. Electrochem. Soe., 132 (1985) 1082. [2] F. Willig, Bet. Btmsenges. Phys. Chem., 92 (1988) 312.

[3] K. Bitterling, F. Wiilig and F. Decker, J. Electroanal. Chem., 228 (1987) 29. [4] F. Willig, K. Bitterling, K.-P. Charte and F. Decker, Ber. Bunsenges. Phys. Chem., 88 (1984) 374. [5] R.L. Cook, R.C. MacDuff and A.F. Sanunels, J. Electrochem. Soc., 136 (1989) 1468. [6] A.P. Norton, S.L. Bernasek and A.B. Bocarsly, J. Phys. Chem., 92 (1988) 6009. [7] J.J. Rarnsden and R. T6th-Bocon~di, J. Chem. Soc. Faraday Trans., 86 (1990) 1527. [8] S. Gottesfeld, Ber. Bunsenges. Phys. Chem., 91 (1987) 362. [9] A. Frippiat and A. Kitsch-De Mesmaeker, J. Phys. Chem., 89 (1985) 1285. [10] S.B. Deutscher, J.H. Richardson, S.P. Perone, J. Rosenthal and J. Ziemer, Faraday Discuss. Chem. Soc., 70 (1980) 33. [I 1] C.N. Kenyon, G.N. Ryba and N.S. Lewis, J. Phys. Chem., 97 (1993) 12928. [12] L.M. Peter, Chem. Rev., 90 (1990) 753. [13] R. Peat and L.M. Peter, J. Electroanal. Chem., 228 (1987) 351. [14] J. Li and L.M. Peter, J. Electroanal. Chem., 193 (1985) 27. [15] L.M. Peter in M. Schiavello (Ed.), Photocatalysis and the Environment: NATO ASI Series C. Vol. 237. Kluwer, Dordrecht, 1988, p. 243. [16] A.R. de Wit, D. Vamaekelberghand J.J. Kelly, J. Electrochem. Sot., 139 (1992) 2508. [17] G. Schlichthtid, E.A. Ponomarev and L.M. Peter, J. Electrochem. Soc., in press. [18] D. Vanmaekelbergh and F. Cardon, Electrochim. Acta, 37 (1992) 837. [19] D. Vanmaekelberghand F. Cardon, J. Phys. D, 19 (1986) 643. [20] D. Vanmaekelbergh, A.R. de Wit and F. Cardon, J. Appl. Phys., 73 (1993) 5049. [21] E.A. Ponomarev and S.D. Babenko, J. Electroanal. Chem., 371 (1994) 27. [22] J. Schefold, J. Electroanal. Chem., 341 (1992) l I 1. [23] E.A. Ponomarev and L.M. Peter, submitted. [24] D.J. Fermin, E.A. Ponomarev and L.M. Peter, in preparation.

Das könnte Ihnen auch gefallen