+ All Categories
Home > Documents > Optimization of optical waveguide modulators based on Wannier-Stark localization: an experimental...

Optimization of optical waveguide modulators based on Wannier-Stark localization: an experimental...

Date post: 19-Nov-2023
Category:
Upload: laborange
View: 1 times
Download: 0 times
Share this document with a friend
10
214 IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 28, NO. 1, JANUARY 1992 Optimization of Optical Waveguide Modulators Based on Wannier-Stark Localization: An Experimental Study E. Bigan, M. Allovon, M. Carre, C. Braud, A.Carenco, and P. Voisin Abstract-We present optical waveguide modulation results obtained using Wannier-Stark localization in InGaAs-InAlAs superlattices grown by molecular beam epitaxy on InP sub- strates. We show that our results are in good agreement with a previously reported wavefunction model. We experimentally investigate the modulation behavior as a function of the elec- troabsorptive superlatticethickness. For this purpose we intro- duce the extinction ratio per unit waveguide length and per unit electric field as a relevant figure of merit that is broadly appli- cable to any electroabsorptive waveguide modulator. We show that inhomogeneousbroadening imposes an optimum thickne5s for the electroabsorptive superlattice. Using a 20 period 65 A InGaAs-20 A InAlAs superlattice located in the core of an op- tical waveguide, we obtain a 14 dB extinction ratio by applying a drive voltage as low as 0.6 V to a 320 pm long waveguide device operating at 1.57 pm incident light wavelength under TE-polarization mode. I. INTRODUCTION IRECTLY modulated semiconductor lasers exhibit a D dynamic wavelength shift called wavelength chirping [l] that is detrimental to long distance and high bit rate data transmission [2] because of the optical fiber chro- matic dispersion. This chirp can be reduced using external modulation [3], [4]. Light intensity modulators using electroabsorption in semiconductors under waveguide configuration are receiving much attention because such components are suitable for monolithic integration with a DFB laser source: first demonstrations operating in the 1.5 pm wavelength range were realized using the Franz- Keldysh effect in bulk semiconductors [5], [6] and they have led to the recent demonstration of low chirp and high bit rate transmission experiments using monolithically in- tegrated DFB lasers and electroabsorptive modulators [7], [8]. Electroabsorption effects in multiple quantum well structures (quantum-confined Stark effect) were subse- quently introduced [9], [ 101 and optimized [ 1 11-[ 151 be- Manuscript received January 4, 1991; revised May 7, 1991 This work was supported in part by “Ministkre de la Recherche et de la Technolo- gie.” CNET-Bagneux and LPMC-ENS are Unit& Associkes au Centre Na- tional de la Recherche Scientifique (CNRS). E. Bigan, M. Allovon, M. C a d , C. Braud, and A. Carenco are with the Centre National d’Etudes des TCltcommunications, Laboratoire de Bagneux, 92220 Bagneux, France. P. Voisin is with the Laboratoire de Physique de la Matihe CondensCe de I’Ecole Normale SupCrieure (LPMC-ENS), 75005 Pans, France. IEEE Log Number 9103676. cause larger absorption variations can be obtained with similar applied electric fields [16], [17]. However drive voltages are most often higher than a few volts which im- plies a large energy dissipation under high speed modu- lation. More recently [18], [19] we have shown that elec- troabsorption effects in strongly coupled superlattices named as Wannier-Stark localization [20]-[22] could be used to achieve very low drive voltage optical waveguide modulation in the 1.5 pm wavelength range. Here we pre- sent improved device performance resulting from an ex- perimental study of modulation characteristics depen- dence on the electroabsorptive superlattice thickness. In order to compare our different structures we introduce a figure-of-merit that is broadly applicable to any electroab- sorptive waveguide modulator. 11. PHYSICAL BACKGROUND When quantum wells are separated by very thin barriers in a periodic structure the discrete energy levels broaden into minibands because of the resonant tunneling effect. Carriers are delocalized and the superlattice (SL) struc- ture exhibits to some extent a three-dimensional (3-D) be- havior. The application of a low electric field to such a structure breaks the resonance because the energy levels of adjacent quantum wells are misaligned by eFd where: e is the elementary electron charge F is the applied electric field d is the SL period. Carriers tend to localize and the structure recovers a two- dimensional (2-D) behavior [20]-[22]. This is very unlike the quantum-confined Stark effect that is a quadratic phe- nomenon [23] and where no significant change is ex- pected under low applied electric fields. More precisely, Bleuse et al. [20] have shown that the SL absorption spec- trum could be viewed as the sum of absorption steps cor- responding to transitions connecting holes and electrons localized in wells separated by p periods and occurring at energies E,, + peFd( p = 0, st 1, +_2, * * ) where E,, is the fundamental transition energy associated to the iso- lated quantum well, The corresponding oscillator strengths are related to the overlap between hole and elec- tron wavefunctions and are given by J: (A/2eFd) where 0018-9197/92$03.00 0 1992 IEEE
Transcript

214 IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 28, NO. 1, JANUARY 1992

Optimization of Optical Waveguide Modulators Based on Wannier-Stark Localization:

An Experimental Study E. Bigan, M. Allovon, M. Carre, C. Braud, A.Carenco, and P. Voisin

Abstract-We present optical waveguide modulation results obtained using Wannier-Stark localization in InGaAs-InAlAs superlattices grown by molecular beam epitaxy on InP sub- strates. We show that our results are in good agreement with a previously reported wavefunction model. We experimentally investigate the modulation behavior as a function of the elec- troabsorptive superlattice thickness. For this purpose we intro- duce the extinction ratio per unit waveguide length and per unit electric field as a relevant figure of merit that is broadly appli- cable to any electroabsorptive waveguide modulator. We show that inhomogeneous broadening imposes an optimum thickne5s for the electroabsorptive superlattice. Using a 20 period 65 A InGaAs-20 A InAlAs superlattice located in the core of an op- tical waveguide, we obtain a 14 dB extinction ratio by applying a drive voltage as low as 0.6 V to a 320 pm long waveguide device operating at 1.57 pm incident light wavelength under TE-polarization mode.

I. INTRODUCTION IRECTLY modulated semiconductor lasers exhibit a D dynamic wavelength shift called wavelength chirping

[ l ] that is detrimental to long distance and high bit rate data transmission [2] because of the optical fiber chro- matic dispersion. This chirp can be reduced using external modulation [3], [4]. Light intensity modulators using electroabsorption in semiconductors under waveguide configuration are receiving much attention because such components are suitable for monolithic integration with a DFB laser source: first demonstrations operating in the 1.5 pm wavelength range were realized using the Franz- Keldysh effect in bulk semiconductors [5], [6] and they have led to the recent demonstration of low chirp and high bit rate transmission experiments using monolithically in- tegrated DFB lasers and electroabsorptive modulators [7], [8]. Electroabsorption effects in multiple quantum well structures (quantum-confined Stark effect) were subse- quently introduced [9], [ 101 and optimized [ 1 11-[ 151 be-

Manuscript received January 4, 1991; revised May 7, 1991 This work was supported in part by “Ministkre de la Recherche et de la Technolo- gie.” CNET-Bagneux and LPMC-ENS are Unit& Associkes au Centre Na- tional de la Recherche Scientifique (CNRS).

E. Bigan, M. Allovon, M. C a d , C. Braud, and A. Carenco are with the Centre National d’Etudes des TCltcommunications, Laboratoire de Bagneux, 92220 Bagneux, France.

P. Voisin is with the Laboratoire de Physique de la Matihe CondensCe de I’Ecole Normale SupCrieure (LPMC-ENS), 75005 Pans, France.

IEEE Log Number 9103676.

cause larger absorption variations can be obtained with similar applied electric fields [16], [17]. However drive voltages are most often higher than a few volts which im- plies a large energy dissipation under high speed modu- lation. More recently [18], [19] we have shown that elec- troabsorption effects in strongly coupled superlattices named as Wannier-Stark localization [20]-[22] could be used to achieve very low drive voltage optical waveguide modulation in the 1.5 pm wavelength range. Here we pre- sent improved device performance resulting from an ex- perimental study of modulation characteristics depen- dence on the electroabsorptive superlattice thickness. In order to compare our different structures we introduce a figure-of-merit that is broadly applicable to any electroab- sorptive waveguide modulator.

11. PHYSICAL BACKGROUND

When quantum wells are separated by very thin barriers in a periodic structure the discrete energy levels broaden into minibands because of the resonant tunneling effect. Carriers are delocalized and the superlattice (SL) struc- ture exhibits to some extent a three-dimensional (3-D) be- havior. The application of a low electric field to such a structure breaks the resonance because the energy levels of adjacent quantum wells are misaligned by eFd where:

e is the elementary electron charge F is the applied electric field d is the SL period.

Carriers tend to localize and the structure recovers a two- dimensional (2-D) behavior [20]-[22]. This is very unlike the quantum-confined Stark effect that is a quadratic phe- nomenon [23] and where no significant change is ex- pected under low applied electric fields. More precisely, Bleuse et al. [20] have shown that the SL absorption spec- trum could be viewed as the sum of absorption steps cor- responding to transitions connecting holes and electrons localized in wells separated by p periods and occurring at energies E,, + peFd( p = 0, st 1, +_2, * * ) where E,, is the fundamental transition energy associated to the iso- lated quantum well, The corresponding oscillator strengths are related to the overlap between hole and elec- tron wavefunctions and are given by J: (A/2eFd) where

0018-9197/92$03.00 0 1992 IEEE

BIGAN et al.: OPTIMIZATION OF OPTICAL WAVEGUIDE MODULATORS 215

Jp is the pth order Bessel function A is the SL total miniband width.

Then the SL electroabsorption spectrum per unit period can be expressed as

- (EQW + peFd)) (1)

where

hvo is the incident light photon energy a. is the absorption step height of the discrete quan-

H is the heaviside function defined by: H(x) = 0 if

It can be shown [20] that in the zero-field limit this expression asymptotically tends towards the superlattice arcosine absorption spectrum corresponding to the super- lattice miniband dispersion as first introduced by Esaki et al. [24]. On the other hand, when eFd approaches A, the Bessel function argument gets sufficiently small so that only the ( p = 0) and ( p = f 1) transitions have signifi- cant oscillator strengths [20]. The corresponding absorp- tion spectra of a superlattice at zero-field and in this so- called high-field limit eFd/A = l are given in Fig. l . It should be noted that neither excitonic contributions nor broadening effects have been taken into account. Both play a major role in determining the actual absorption spec- trum shape. Excitonic effects have been observed by many authors primarily in the localization regime because of the two-dimensional-like behavior [ 191, [25]-[30], but also under flatband conditions [27]-[30]. Excitonic contribu- tions as well as coulombic enhancement of the contin- uums associated with oblique and direct transitions have been considered recently [31]-[34] and remain a subject under investigation. How the noninclusion of these effects will limit the significancy of this model will be discussed when we will compare experimental results to calculated data. Two interesting electroabsorption configurations can be distinguished on this schematic:

i) incident photon energy is higher than the zero-field SL bandgap and lower than the single quantum well band- gap (i.e., : E,, - A/2 < hvo 5 EQw): there is a negative absorption variation named as ‘blue-shift” [25] which has been successfully used to realize a self electrooptical bi- stable device [26] electroabsorptive modulator arrays [27], as well as a normally-off asymmetric Fabry-Perot reflection modulator [35], [36], all in the GaAs-GaAlAs material system. We have already pointed out [18], [19] that this configuration is not suitable for any application requiring both high contrast and low attenuation because excitonic effects as well as broadening mechanisms con- tribute to increase significantly the on-state absorption originating from the ( p = - 1) oblique transition.

ii) incident photon energy is lower than the zero-field SL bandgap (i.e.,: hvo < E,, - A/2): there is a positive

tum well

x < 0 and H(x) = 1 otherwise.

WAVELENGTH (a.u.1

Fig. 1 . Absorption spectra (schematic) of an ideal superlattice without ap- plied electric field (solid line) and in the “high-field’’ limit (dashed line).

absorption variation originating from the ( p = -1) oblique transition which is similar to conventional “red- shift” electroabsorption effects except it can be obtained with much smaller electric fields. In our previous work [IS], [I91 we have proposed the use of this ( p = - 1) oblique transition to achieve efficient very low drive volt- age optical waveguide modulation and we have demon- strated it using a InGaAs-InAl As superlattice featuring both high extinction ratio and low on-state attenuation. All of the following will concentrate on this ii) electroab- sorptive configuration.

111. EXPERIMENTAL Our structures (sample I, 11, and 111) are InGaAlAs p-

i-n double heterostructures grown by MBE onto n + -1nP substrates (Sn doping level: 1.5 X 10l8 ~ m - ~ ) . The nom- inally undoped region consists of a 10 (!ample I), 20 (!ample 11), or 30 (sample 111) period 65 A InGaAs-20 A InAlAs electroabsorptive SL (absorption edge: 1.5 pm) surrounded on toth sides by 15. (sample I) or 8 (sample 11) period 25 A InGaAs-30 A InAlAs confining SL. Sample I11 has no confining SL. The nominally undoped region width is thus kept constant (0.26 pm). The nomi- nally undoped region was surrounded by a 0.1 pm thick n-InAIAs buffer layer (doping level: 5 x 10” cmP3) and a 1.8 pm thick p-InAlAs confining layer (doping level: 5 X loL7 A 0.1 pm thick pf-InGaAs contact layer (doping level: 2 x l O I 9 cmP3) was grown on top of the structures. The refractive index of the confining SL is in- termediate between the refractive index of the electroab- sorptive SL and the refractive index of the InAlAs confin- ing layers. The purpose of its inclusion is to increase the optical confinement factor over the electroabsorptive SL. Magnitudes for the optical confinement factor of the three samples will be specified when we will compare the per- formances of the different modulator structures. Fig. 2 gives a schematic of the structures. 50 pm wide TiAu electrodes were evaporated on top of 100 pm wide mesas etched down to the InP substrate. After thinning and Au electrode deposition on the backside of the samples broad- area modulators of various lengths ranging from 100 to 500 pm were cleaved. They were characterized using a tunable NaC1-OH F-center laser. We only investigated the

216 IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 28, NO. I , JANUARY 1992

f P + - I n G a A s

I P - I n A l A s I I - I n G a A s / I n A I A s confining SL

I - I n G a A s / I n A I A s e l e c t r o a b s o r o t i v e SL

N+ - InP s u b s t r a t e

Fig. 2. Cross-sectional view of the modulator structures (schematic).

TE-guided modes in this work. Relative transmission was measured as a function of reverse bias at different incident light wavelengths of the electroabsorptive SL transparent region. Photocurrent versus reverse bias was simulta- neously measured and the combination of both measure- ments allowed a determination of the on-state attenuation 1371. These characterizations were carried out for differ- ent waveguide lengths and for each length on different modulators. Averaged results allowed us to obtain a meaningful comparison of the three different structures. In order to analyse the electroabsorption effects involved in our structure we have carried out photocurrent (PC) spectroscopy under waveguide configuration. PC spectra are representative of electroabsorption spectra if only a small fraction of the incident light is absorbed. This con- dition is most often satisfied when light propagates per- pendicularly to the epitaxial layers but this is not the case under waveguide configuration with usual waveguide lengths. In order to obtain PC spectra that are represen- tative of electroabsorption spectra we fabricated devices having one cleaved facet and one etched facet by applying conventional photolithography and wet chemical etching techniques to individual devices. The waveguide length could thus be reduced to 5 pm.

IV. RESULTS AND DISCUSSION A . Isolated Devices

PC spectra are given in Fig. 3. They were measured on short waveguides for different reverse bias voltages. A very similar behavior is observed for the three different samples: at +0.35 V (forward bias) the electroabsorptive SL is only partially depleted and the corresponding PC spectra are given by the dashed lines. When the bias volt- age goes from +0.20 V (forward bias) to -0.40 V (re- verse bias) the electric field is increased and three features that are characteristic from Wannier-Stark localization can be clearly observed:

i) a strong increase of the excitonic oscillator strength which results from the going from a 3-D behavior to a 2- D behavior. This occurs around 1.48 pm incident light wavelength.

1.45 1.50 1.55 1. WAVELENGTH lpml

0

Sample !l

1.50 1.55 1.60 WAVELENGTH ipml

I I 1 Sample m L=5pm

Fig. 3 . Photocurrent spectra of the three samples for different bias. Wave- guide length is 5 pm. The 0.35 V forward bias spectra are given by the dashed lines.

ii) an expected blue-shift of the absorption edge around 1.52 pm incident light wavelength.

iii) the (- 1) oblique transition occurring in the elec- troabsorptive SL transparent region (around 1.58 pm in- cident light wavelength).

The electron-to-heavy hole excitonic absorption peak position gives us direct information on the actual well width. The 1.485 pm ayerage position is in reasonable agreement with the 65 A nominal well width. Further- more, the dispersion in this peak position is only 5 nm (2.8 meV) over the three samples which indicates excel- lent growth control.

Fig. 4 shows typical relative transmission curves versus reverse bias at different wavelengths. Apart from the mag- nitude of the extinction ratio that is sample dependent and that will be discussed in the next subsection, the three samples exhibit a similar general behavior in the same

BIGAN et al.: OPTIMIZATION OF OPTICAL WAVEGUIDE MODULATORS 211

100

901 - 80- E 70- - 60-

5 5 0 - m

a + 4 0 - > ‘4 30-

1

0- - 0 5 0 0.5 1 1 5 2

REVERSE BIAS IV)

Sample 1 L I 3 2 0 p m

1.57pm 70 1.58pm

1.59pm 1.60pm

VI ,b I 10

0 -0.5 0 0.5 1 1.5 2

REVERSE BIAS [ V I

100

9 0 1

1.59pm 1.60pm

Sample Q l L=400pm TE i

scribed in our previous work [ 191. Furthermore we notice that:

i) the optimum voltage for minimum relative transmis- sion shifts linearly with increasing incident light wave- length

ii) once this optimum has been reached and relative transmission is increasing again the relative transmission curves for the different wavelengths superimpose to each other. Points i) and ii) will be clarifed in the following discussion. From (1) we determine the absorption varia- tion A a as a function of the electric field F for a given photon energy hvo lower than the SL bandgap E,, - A / 2 (zero-field absorption is zero):

At low electric field F, only high order (-p) oblique tran- sitions contribute to the absorption. The corresponding oscillator strengths are low because the Bessel function orders are high. When F varies from zero to infinity, ab- sorption undergoes jumps for electric field values of E,, - hvo/med, m being an integer. These jumps correspond to the apparition of the (-m) oblique transitions in our operating wavelength region. Between these jumps the absorption decreases [38]. This F-’ oscillatory behavior has already been observed in a GaAs-GaAlAs superlattice at low temperature using electroreflectance [22]. Maxi- mum absorption variation is obtained under the (-1) oblique transition [38]. It can be expressed as [38]

- 0 5 0 0 5 1 1.5 2 REVERSE BIAS ( V I

Fig. 4. Relative transmission of the TE-guided mode versus reverse bias voltage for three individual devices taken out from samples I, 11, and 111. Measurements were carried out for four different wavelengths.

wavelength range. This last point, combined with the low dispersion gf the well width, indicates that the dispersion of the 20 A nominal barrier width is low over the three structures. This precise control of the MBE growth was confirmed by X-ray diffqction measurements. Measured periods are 83 and 53 A for the electroabsorptive SL (samples I , 11, and 111) and the confining SL (samples I and 11). Maximum relative transmission is obtained under forward bias because the p-i-n built-in electric field is sufficiently large to partially induce Wannier-Stark local- ization. We did not investigate behavior above 0.35 V forward bias in order to avoid any perturbation due to for- ward current flowing through the diode. When forward bias (reverse bias, respectively) is decreased (increased, respectively) relative transmission drops very rapidly, reaches a minimum value, then increases again. Such be- havior has already been observed and qualitatively de-

(3)

(4)

with the condition

eFd 2 E,, - hvo.

Because we operate below the SL bandgap the Ressel function argument is smaller than unity and Aahuo is a de- creasing function of F. Therefore maximum absorption variation is obtained when

eFd = E,, - hvo. (5) Now we are going to verify that our results are in good

agreement with (5). Neglecting the exciton binding en- ergy (a few meV), as well as any displacement of the ex- citon due to the quantum-confined Stark effect (using Bas- tard variational approach [23] the calculated Stark shift for an isolated quantum well at 0.4 V reverse bias is 0.75 meV) we estimate the quantum well band-gap energy E,, from the excitonic absorption peak position at 0.4 V re- verse bias: E,, = 835 * 3 meV (* stands for dispersion over the three samples). First let us consider device op- eration at X = 1.57 pm that is to say hvo = 790 meV. The mean electric field over the intrinsic region of opti- mum modulation Fopt is estimated by

218 IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 28, NO. 1, JANUARY 1992

- 1.57pm _-_ 1.58pm

where

Vopt is the reverse bias for optimum modulation Vbi is the built-in potential tl is the intrinsic layer thickness.

Using$tructure parameters tl = 0.26 pm, Vbi = 0.8 V, d = 85 A and measuring Vopt on the relative transmission curves: VOpt = 0.5 f 0.1 V (i- stands for dispersion over the three samples) we get: eFoptd = 43 &- 3 meV (i- stands for dispersion over the three samples). We see that ( 5 ) is verified within a few meV. Equation ( 5 ) also im- plies that optimum electric field dependence on incident light energy is linear [point i)] and is only related to the superlattice period. From ( 5 ) we derive this dependence

1 ed WO,, = - - 6(hvo) . (7)

Using the structure parameters and the expression of the electric field (6) we calculate the optimum voltage de- pendence on the incident light wavelength: 6Vop,/6X = 0.15 V/O.O1 pm which compares favorably to the exper- imental value determined from the relative transmission curves: 6Vopt/6X = 0.13 V/O.O1 pm. This clarifies point i). Equation (3) implies that, provided we operate under the ( - 1) oblique transition, the absorption variation is a decreasing function of the electric field F that only de- pends on the superlattice parameters A and d , but does not depend on the incident light wavelength. This [point ii)] can be verified on our relative transmission curves ex- cept at short wavelength and for large reverse bias where relative transmission starts decreasing again. We find a good agreement between the experimental results and the calculated data and this raises up a few comments: ( 5 ) means that the shift between the direct and (- 1) oblique transition is equal to the energy level misalignment be- tween two adjacent wells, which is the basis of a Stark ladder formation. In determining these transition posi- tions we have neglected the related excitonic binding energies. It should be noted that these energies (a few meV) are much smaller than the misalignment values we are dealing with (45 meV) and that such a further correc- tion would fall within the precision we are interested in (a few meV). Excitonic contributions as well as Coulom- bic enhancement of the continuums associated to direct and oblique transitions will affect the measured absorp- tion variations but not the general behavior. This is why we did not attempt to compare measured extinction ratios with calculated absorption variations. The same remarks can be made about broadening effects that will smooth out Wannier-Stark localization features, but that should not change considerably their positions.

B. Averaged Results

Fig. 5 shows the averaged extinction ratio per unit waveguide length simultaneously with the averaged on-

state attenuation per unit waveguide length versus the number of periods of the electroabsorptive superlattice. They are given for four different incident light wave- lengths. For a given sample (number of periods is fixed) the larger extinction ratios are obtained close to the su- perlattice bandgap which has been explained in the pre- vious analysis. For a given wavelength the on-state atten- uation increases with increasing the number of periods. This is expected because of the increase in the optical confinement factor. To estimate the optical confinement factor for the three samples, we carried out multilayer cal- culations [39]. We used bulk InGaAs and InAlAs refrac- tive index values determined from single effective oscil- lator modes [40] for the well and barrier refractive index values. Oscillator and dispersion energies were taken out from the data reported by Nojima et al . [41] for InGaAs and InAlAs and by Broberg et al . [42] for the InP sub- strate. Calculations were performed at the mean wave- length 1.585 pm and the corresponding values of the con- stitutive materials are: nInGaAs = 3.61, nInAIAs = 3.21, and nInP = 3.17. One must be aware that such an approxi- mation neglects the enhanced dispersion near the band- edge of quantum structures [43], [44]. A more precise de- termination of the modal properties of the SL waveguide could be achieved using grating coupling techniques [45], [46]. Estimated values of the optical confinement factor are 0.13, 0.27, and 0.40 for samples I , 11, and 111, re- spectively. At long wavelengths (1.59 pm, 1.60 pm) the extinction ratio increases with increasing the number of periods. At shorter wavelength (1.58 pm) this extinction ratio saturates and eventually slightly decreases when going from 20 to 30 periods at 1.57 pm incident light wavelength. This behavior seems contradictory with a mere increase of the optical confinement factor. The re- sults are replotted in Fig. 6 in a different way: the aver- aged extinction ratio per unit waveguide length is divided by the averaged on-state attenuation per unit waveguide length.

Using the notations developed in the appendix, this fig- ure is equal to Aa/ao. It has already been used for opti-

BIGAN et al.: OPTIMIZATION OF OPTICAL WAVEGUIDE MODULATORS 219

O 0-0 NUMBER OF PERIODS

Fig. 6. Average extinction ratio divided by average on-state attenuation (modulation depth parameter: A a / a o ) versus number of periods of the electroabsorptive SL. Measurements were carried out for four different wavelengths.

mizing quantum-confined Stark effect-based modulators [12] and it can be referred to as the modulation depth pa- rameter. We can see on this plot that

i) A a / a o decreases with increasing the number of pe- riods for each investigated wavelength

ii) A a / a o increases with increasing the operating wavelength. We believe that this behavior can be explained by broadening. The overall effect of broadening is to in- crease a. and to decrease A a which makes A a / a o a pa- rameter that is very sensitive to broadening. We attribute the A a / a o increase with operating wavelength to homo- geneous broadening: although the largest values of A a are obtained close to the absorption edge cyo values are also very large in this range because of homogeneous broaden- ing. These large values of a. then decrease faster than A a when increasing the operating wavelength. The A a /ao decrease with increasing the number of periods is a char- acteristic feature of inhomogeneous broadening. The ma- jor source of such broadening can be electric field inho- mogeneity arising from finite residual doping in the nominally undoped layer. Background carrier concentra- tion within the nominally undoped region was estimated to be 1.2 X loL6 cmT3 from C-V measurements as well as from photocurrent level at short wavelength. The ori- gin of such a high carrier concentration is not clear at present but it has already been observed by Wakita et al. [15] in MQW modulators realized in the same material system using the same growth technique.

It is clear from Fig. 5 and Fig. 6 that A a / a o is not a relevant figure of merit for modulator optimization: a value of A a / a o as large as 25 is obtained at 1.60 pm operating wavelength using the 10 period electroabsorp- tive superlattice (sample I). However this corresponds to a very poor modulator since the averaged extinction ratio per unit waveguide length is lower than 1 dB/ 100 pm.

On the other hand the averaged extinction ratio per unit waveguide length divided by the applied electric field is a relevant figure-of-merit as fully explained in the appen- dix. It takes into account the compromises arising be-

tween extinction ratio, diode capacitance, and drive volt- age through waveguide length and intrinsic layer thickness. This figure-of-merit is labelled as r * A a / A F where

r is the optical confinement factor A a is the absorption variation A F is the applied electric field difference between on-

It is shown in Fig. 7, versus the number of periods. For this figure-of-merit determination we took A F = AV/ t , . The on-state was taken as the 0.2 V forward bias state in order to satisfy the depletion condition that is necessary to ensure full comparison capability of this figure-of-merit (see Appendix).

This value of 0.2 V forward bias for intrinsic layer de- pletion was determined from photocurrent measurements at short wavelength (1.45 pm) on sample I11 in which the whole intrinsic layer contributes to photocurrent. It was confirmed by capacitance-voltage measurements that were carried out on the three samples. The off-state was taken as the bias value that gave the largest extinction ratio. We should notice that on-state and off-state could have been chosen using the bias values that maximize I' * A a / A F . Slightly higher values of this figure-of-merit can thus be obtained corresponding to lower drive voltage configura- tions. For simplicity purposes, this further refinement of modulation optimization was not investigated in this work. For long wavelengths (1.59-1.60 pm) r * A a / A F in- creases with increasing the number of periods. At shorter wavelength (1.58 pm) it decreases slightly between 20 and 30 periods and eventually exhibits a sharp maximum for 20 periods at 1.57 pm incident light wavelength. Far away from the absorption edge, optimum operation is achieved using a large number of periods. Closer to the absorption edge where broadening effects play a major role, optimum operation is achieved with a smaller num- ber of periods. Operation is optimum as close as possible to the absorption edge as expected from the analysis de- veloped in the individual device section: larger absorption variation is obtained with smaller electric field when op- erating wavelength is closer to the absorption edge. As mentioned in the appendix, one should however check that on-state attenuation is not prohibitive. For the 20 period sample (sample 11) we obtained a A a / a o parameter value of 3 and a r - Acy/AF parameter of 1.7 dB/100 pm/V/pm at 1.57 pm operating wavelength. Both pa- rameters were determined from the averaged measure- ments with the on-state defined as the 0.2 V forward bias state to guarantee intrinsic layer depletion. This is the rea- son why the A a / a o value is lower than shown in Fig. 5 where the on-state was defined as the 0.35 V forward bias state.

Our best individual device was a 320 pm long wave- guide device. On-state attenuation was 4.6 dB, extinction ratio was 14 dB, and the drive voltage was 0.6 V (from 0.2 V forward bias to 0.4 V reverse bias) at 1.57 pm op- erating wavelength. These data correspond to a A a / a o parameter value of 3.05 and a r A a / A F parameter value

state and off-state.

220 IEEE

0 L

-2 - 0.

\ 0.5 - U-

0 - -

JOURNAL OF QUANTUM ELECTRONICS, VOL. 28, NO. 1, JANUARY 1992

,’,,,/’, ’ ’,, / -

’ / / 1 LO

Fig. 7 . Average extinction ratio per unit waveguide length divided by ap- plied electric field (figure-of-merit: r . A a / A F ) versus number of periods of the electroabsorptive SL. Measurements were camed out for four dif- ferent wavelengths.

of 1.9 dB/ 100 pm/V/pm. Both values are very close to the averaged values. Such low dispersions of the modu- lator characteristics were obtained for the three samples. This demonstrates excellent growth uniformity.

V. CONCLUSION We have reported Wannier-Stark localization obser-

vation in InGaAs-InAlAs superlattice waveguides. We have achieved efficient waveguide intensity modulation using the (- 1) oblique transition connecting a hole local- ized in a well with an electron localized in the adjacent well. We show that our results are in fairly good agree- ment with the superlattice wavefunction model developed by Bleuse et al . [20]. We have carried out an experimen- tal study of the dependence of modulation characteristics on the electroabsorptive superlattice thickness. In order to compare our different structures we propose the use of the extinction ratio per unit waveguide length and per unit applied electric field as a relevant figure of merit for elec- troabsorptive waveguide modulator optimization. Using both this figure of merit and the modulation depth param- eter we show that inhomogeneous broadening effects im- pose an optimum electroabsorption superlattice thickness for optimum modulation. Using a 20 period electroab- sorptive superlattice we have realized a modulator having a 14 dB extinction ratio and a 4.6 dB on-state attenuation with a drive voltage as low as 0.6 V. The device is a 320 pm long waveguide operating at 1.57 pm under TE po- larization mode.

APPENDIX In this section we introduce a figure of merit that is

applicable to any electroabsorptive waveguide modulator. We consider a device consisting of a p-i-n waveguide diode. The geometrical parameters of the device are listed below

L is the waveguide length W is the p-i-n diode width tI is the intrinsic layer thickness.

The intrinsic layer contains the electroabsorptive ma- terial and the electric field is applied by reverse biasing the diode. Let us precise the electrical parameters of the p-i-n diode. First we assume that the residual doping in the intrinsic layer is sufficiently low to ensure that the mean electric field value F over the intrinsic layer thick- ness tI is meaningful. Electric field inhomogeneity result- ing from finite residual doping can be further introduced as an additional inhomogeneous broadening factor in the electroabsorption analysis. In order to avoid dramatic in- crease of the p-i-n diode capacitance the intrinsic layer should be kept depleted all the time. This is our second assumption: intrinsic layer remains depleted even in the lowest mean electric field state (which is the on-state for conventional positive absorption variation configurations, off-state for “blue-shift” configurations). Third we as- sume that the p and n doping levels are sufficiently high so that the extension of the depletion region in the p and n doped region is negligible. Under these assumptions we can write

where

V is the applied bias F is the mean electric field value

index on or 08 stands for on-state or off-state.

This can be rewritten as

A V ~ = AF

tl where

A V is the drive voltage A F is the electric field difference between on-state and

Now we introduce the electroabsorption properties. Let a. be the on-state absorption and Act the absorption vari- ation occurring under variation of the mean electric field F. If we define the optical confinement factor r as the fraction of guided energy contained in the electroabsorp- tive material then we can express the extinction ratio on /off and the on-state attenuation att. (notwithstanding coupling losses) as [ 171 :

off-state.

on/off = er’Aa.L (A31 art. = er.ao.L (A41

Another important parameter of the electroabsorptive modulator is the electric 3 dB bandwidth. Here we assume that it is RC limited. This has been discussed in great de- tail by Wood [17]. It means we neglect the photogener- ated carriers transit time over the intrinsic layer thickness. Experimental data on quantum confined Stark effect-based modulators show that this is a valid assumption up to speeds in excess of 20 GHz [15]. This can be qualitatively explained first by the fact that such modulators operate at

BIGAN et al.: OPTIMIZATION OF OPTICAL WAVEGUIDE MODULATORS 22 1

large electric fields and the transit time is known to de- crease with increasing electric fields [47], and second be- cause the intrinsic layer of such guided wave modulators is thin (around 0.5 vm) which shortens the transit time of photogenerated camers. Although up to now no experi- mental data is available about Wannier-Stark localiza- tion-based modulators, we can notice that i) although the operating electric field is low, transit is expected to be fast because of the very small bamer width ii) second the total intrinsic layer is even thinner in the structures that have been considered up to now [19], [this work]. Then the 3 dB bandwidth V3dB can be expressed as

1 (A51

1 tl =-. T * R CPi-,, T * R E/ * (W * L ) v3dB =

where

R is a load resistor connected in parallel with Cp-l-,, (the assembly being driven by a matched voltage generator with impedance R)

Cp+" is the p-i-n capacitance cl is the intrinsic layer dielectric constant.

Here we have neglected the extension of the depleted region in the p and n regions which is valid provided p and n doping levels are sufficiently high. The four char- acteristics of an electroabsorptive optical waveguide mod- ulator are given by (A2-A5). System applications require a large extinction ratio simultaneously with a low on-state attenuation. Therefore the A a / a o parameter must be suf- ficiently large. However this single parameter cannot be considered as a relevant figure of merit because it does not take into account neither the 3 dB bandwidth nor the drive voltage. From the four characteristics expressions given above, it can be seen that extinction ratio, drive voltage, and 3 dB bandwidth are related to each other. One wishes to get simultaneously a large extinction ratio, a low drive voltage, and a high 3 dB bandwidth. To de- termine a figure-of-merit taking into account the relation- ship between these characteristics we should consider two different electroabsorption configurations determined by the following parameters

Ll,2

(t l) 1,2

AF1.2. Indexes 1 and 2 refer to configurations 1 and 2. cl is

assumed to display no significant change from one config- uration to another which is valid for 111-V materials. We wish to compare the potentialities of different electroab- sorption configurations to achieve efficient optical wave- guide modulation but we do not plan to determine an op- timum device geometry or technology. This is why we do not discuss neither the influence of the W parameter (in- trinsic layer width) that depends on technological choices for achieving transverse single-mode operation, nor the

contribution of the bonding pad capacitance to the total device capacitance. These are independent of the elec- troabsorption configurations to a large extent.

If we let the extinction ratios as well as the drive volt- ages be equal, i.e.:

then

which means the higher the extinction ratio per unit wave- guide length and per unit applied electric field the higher the bandwidth.

Similarly, if we let the extinction ratios as well as the bandwidths be equal, i.e.:

then:

which means the higher the extinction ratio per unit length and per unit applied electric field the lower the drive volt- age.

Eventually, if we let the drive voltages as well as the bandwidths be equal, i.e.,

then

(r * Aa)l * LI (I' - Aa)l (t1)l AFI (r * & (r Aa)2 (t1)2 (r - Aa),

AF2

(A1 1) . - - - - -

which means the higher the extinction ratio per unit length and per unit applied electric field the higher the extinction ratio.

We have shown that if any two of the three character- istics: extinction ratio, drive voltage, and bandwidth are equal, the third one compares to the extinction ratio per unit length and per unit applied electric field expressed by

222 IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 28, NO. 1, JANUARY 1992

r A~ A F

This demonstrates why this parameter is a relevant fig- ure-of-merit that takes into account the compromises be- tween extinction ratio and bandwidth through the wave- guide length and between drive voltage and bandwidth through the intrinsic layer thickness. But it does not suf- fice by itself since it does not take into account the on- state attenuation. Both A a / a o and r - A a / A F parame- ters should be considered. However these two figures of merit do not play a symmetric role. We have demon- strated that a linear increase of the figure-of-merit r - A a / A F results in a linear improvement of either the ex- tinction ratio, the drive voltage or the 3 dB bandwidth. Let us have a closer look to the effect of A a / a o : this parameter is equal to the extinction ratio divided by the on-state attenuation (both being expressed in dB). If we impose a given extinction ratio ../of the output power drop through the modulator in the on-state (on-state atten- uation) is given (in dB) by: o n / o $ l ( A a / a o ) where on/o$is expressed in dB and A a / a o is the dimensionless modulation depth parameter. This means that once A a /ao is sufficiently large for the on-state attenuation to be small an important increase in A a / a o does not bring any sig- nificant improvement. Therefore A a / a o must be given a minimum value that is determined from system consid- erations while the actual optimization should be made on the A a / A F figure-of-merit.

ACKNOWLEDGMENT We wish to thank J . Landreau for processing assistance

and J . Primot for X-ray measurements.

REFERENCES

[l] T. L. Koch and J. E. Bowers, “Nature of wavelenth chirping in di- rectly modulated semiconductor lasers,” Electron. Lett., vol. 20, pp.

[2] J. C. Cartledge and G. S. Burley, “The effect of laser chirping on lightwave system performance,” J. Lightwave Technol., vol. 7, pp. 568-573, 1989.

[3] F. Koyama and K. Iga, “Frequency chirping of external modulation and its reduction,” Electron. Lett., vol. 21, pp. 1065-1066, 1985.

[4] -, “Frequency chirping in external modulators,” J . Lightwave Technol., vol. 6, pp. 87-93, 1988.

[5] Y. Noda, M. Suzuki, Y. Kushiro, and S. Akiba, “High-speed elec- troabsorption modulator with strip-loaded GaInAsP planar wave- guide,”J. Lightwave Technol., vol. 4, pp. 1445-1453, 1986.

[6] H. Soda, K. Nakai, H. Ishikawa, and H. Imai, “High-speed Ga- InAsP/InP buried-heterostructure optical intensity modulator with semi-insulating InP burying layers,” Electron. Lett., vol. 23, pp.

[7] H. Soda, M. Furutsu, K. Sato, N. Okazaki, S. Yamazaki, H. Nishi- moto, and H. Ishikawa, “High-power and high-speed semi-insulating BH structure monolithic electroabsorption modulator/DFB laser light source,” Electron. Lett., vol. 26, pp. 9-10, 1990.

[8] M. Suzuki, H.Taga, S. Yamamoto, H. Wakabayashi, and Y. Mat- sushima, “Sub-volt operation of E-A modulator integrated with X/4- shifted DFB laser for 2.4 Gb/s-135 km transmission,” presented at 16th Euro. Conf. Opt. Commun., Amsterdam, The Netherlands, 1990, post-deadline paper ThG2.5.

[9] U. Koren, B. 1. Miller, T. L. Koch, G. Eisenstein, R. S. Tucker, I. Bar-Joseph, and D. S. Chemla, “Low-loss InGaAs/InP multiple quantum well optical electroabsorption waveguide modulator,” Appl. Phys. Lett., vol. 5 1 , pp. 1132-1134, 1987.

1038-1039, 1984.

1232-1234, 1987.

K. Wakita, Y. Kawamura, M. Nakao, and H. Asahi, “Long wave- length waveguide multiple quantum well optical modulators,” IEEE J. Quantum Electron., vol. QE-23, pp. 2210-2215, 1987. H. Temkin, D. Gershoni, and M. B. Panish, “InGaAsP/InP quan- tum well modulators grown by gas source molecular beam epitaxy,” Appl. Phys. Lett., vol. 50, pp. 1776-1778, 1987. S. Nojima and K. Wakita, “Optimization of quantum well materials and structures for excitonic electroabsorption effects,” Appl. Phys. Lett., vol. 53, pp. 1958-1960, 1988. J. E. Zucker, I. Bar-Joseph, B. I. Miller, U. Koren, and D. S. Chemla, “Quatemary quantum wells for electro-optic intensity and phase modulation at 1.3 pm and 1.55 bm,” Appl. Phys. Lett., vol.

E. Bigan, M. Allovon, M. C a d , and A. Carenco, “Strained-layer InGaAs/InAlAs multiple quantum wells for efficient optical wave- guide modulation at 1.55 pm,” Electron. Lett., vol. 26, pp. 355- 356, 1990. K. Wakita, I. Kotaka, 0. Mitomi, H. Asahi, Y. Kawamura, and M. Naganuma, “High-speed InGaAlAs/InAlAs multiple quantum well optical modulators,” J . Lightwave Technol., vol. 8, pp. 1027-1031, 1990. D. A. B. Miller, D. S. Chemla, T. C. Damen, A. C. Gossard, W. Wiegmann, T. H. Wood, and C. A. B u m s , “Electric field depen- dence of optical absorption near the band gap of quantum-well struc- tures,” Phys. Rev. B , vol. 32, pp. 1043-1060, 1985. T. H. Wood, “Multiple quantum well (MQW) waveguide modula- tors,” J . Lightwave Technol., vol. 6, pp. 743-757, 1988. E. Bigan, M. Allovon, and P. Voisin, French patent No. 89 15909, 1989. E. Bigan, M. Allovon, M. Carrk, and P. Voisin, “Very low drive voltage optical waveguide modulation in an InGaAs /InAIAs super- lattice,” Appl. Phys. Lett., vol. 57, pp. 327-329, 1990. J. Bleuse, G. Bastard, and P. Voisin, “Electric-field-induced local- ization and oscillatory electro-optical properties of semiconductor su- perlattices,” Phys. Rev. Lett., vol. 60, pp. 220-223, 1988. E. E. Mendez, F. Agullo-Rueda, and J . M. Hong, “Stark localization in GaAs-GaAlAs superlattice under an electric field,” Phys. Rev. Lett., vol. 60, pp. 2426-2429, 1988. P. Voisin, J. Bleuse, C. Bouche, S. Gaillard, C.Alibert, and A. Re- greny, “Observation of the Wannier-Stark quantization in a semi- conductor superlattice,” Phys. Rev. Lett., vol. 61, pp. 1639-1642, 1988. G. Bastard, E. E. Mendez, L. L. Chang, and L. Esaki, “Variational calculations on a quantum well in an electric field,” Phys. Rev. B ,

L. Esaki and R. Tsu, “Superlattice and negative differential conduc- tivity in semiconductors,” IBM J. Res. Develop., vol. 14, pp. 61- 65, 1970. J. Bleuse, P. Voisin, M. Allovon, and M. Quillec, “Blue shift of the absorption edge in AIGaInAs-GaInAs superlattices: Proposal for an original electro-optical modulator,” Appl. Phys. Lett., vol. 53, pp.

I. Bar-Joseph, K. W. Goossen, J. M. Kuo, R. F. Kopf, D. A. B. Miller, and D. S. Chemla, “Room-temperature electroabsorption and switching in a GaAs/AlGaAs superlattice,” Appl. Phys. Lett., vol. 55, pp. 340-342, 1989. G. R. Olbright, T. E. Zipperian, J. Klem, and G. R. Hadley, “Op- tical switching in N x N arrays of individually addressable electroab- sorption modulators based on Wannier-Stark carier localization in GaAs/GaAlAs superlattices,” J. Opt. Soc. Amer. B, vol. 8, pp. 346- 354, 1991. R. H. Yan, R. J. Simes, H. Ribot, L. A. Coldren, and A. C. Gossard, “Room-temperature two-dimension exciton exchange and blue-shift of absorption edge in GaAs/AlGaAs superlattices under an electric field,” Appl. Phys. Lett., vol. 54, pp. 1549-1551, 1989. B. Pezeshki, D. Thomas, and J. S. Hams, Jr., “Wannier-Stark lo- calization in a strained InGaAs/GaAs superlattice,” Appl. Phys.

H. Schneider, H. T. Grahn, K. v. Klitzing, and K. Ploog, “Reso- nance-induced delocalization of electrons in GaAs-AIAs superlat- tices,” Phys. Rev. Lett., vol. 65, pp. 2720-2723, 1990. R. H. Yan, F. Lamelle, and L. A. Coldren, “Tight-binding analysis on exciton energy in field-induced Stark-localized superlattices, ” Appl. Phys. Lett., vol. 5 5 , pp. 2002-2004, 1989. D. M. Whittaker, “Theory of excitons in short-period superlattices,” Phys. Rev. B , vol. 41, pp. 3238-3241, 1990. M. M. Dignam and J. E. Sipe, “Exciton Stark ladder in GaAs/Ga-

54, pp. 10-12, 1989.

vol. 28, pp. 3241-3245, 1983.

2632-2634, 1988.

Lett., vol. 57, pp. 2116-2117, 1990.

BIGAN et al . : OPTIMIZATION OF OPTICAL WAVEGUIDE MODULATORS 223

I,AI,As superlattices,” Phys. Rev. Lett., vol. 64, pp. 1797-1800, 1990. K. J . Moore, G. Duggan, A. Raukema, and K. Woodbridge, “Ex- citons associated with miniband dispersion in (1nGa)As-GaAs strained layer superlattices,” Superlatt. and Microstruct., vol. 7, pp. 303- 308, 1990. K.-K. Law, R. H. Yan, J. L. Merz, and L. A. Coldren, “Normally- off high-contrast asymmetric Fabry-Perot reflection modulator using Wannier-Stark localization in a superlattice,” Appl. Phys. Lett., vol.

K.-K. Law, R. H. Yan, L. A. Coldren, and J. L. Merz, “Self-elec- tro-optic device based on a superlattice asymmetric Fabry-Perot mod- ulator with an on/off ratio > 100: I , ” Appl. Phys. Lett., vol. 57, pp.

T. H. Wood, “Direct measurement of the electric-field-dependent ab- sorption coefficient in GaAs/AlGaAs multiple quantum-wells,” Appl. Phys. Lett., vol. 48, pp. 1413-1415, 1986. For a review of Bessel functions properties see for example: M. A. Abramovitz and I. A. Stegun, Handbook of mathematical functions. I. M. Skinner, R. Shail, and B. L. Weiss, “Modal propagation within MQW waveguides,” IEEEJ. Quantum Electron., vol. 25 , pp. 6-11, 1989. S. H. Wemple and M. DiDomenico, Jr., “Behavior of the electronic dielectric constant in covalent and ionic materials,” Phys. Rev. B ,

S. Nojima and H. Asahi, “Refractive index of InGaAs/InAIAs mul- tiquantum-well layers grown by molecular-beam epitaxy,” J . Appl.

B. Broberg and S. Lindgren, “Refractive index of Inl-xGaxAsyP1-y layers and InP in the transparent wavelength region,” J. Appl. Phys.,

Y. Suzuki and H. Okamoto, “Refractive index of GaAs-A1As super- lattice grown by MBE,” J . Electron. Mater., vol. 12, pp. 397-411, 1983. f . P. Leburton, K. Hes, N. Holonyak, Jr., J. J . Coleman, and M. Camras, “Index of refraction of AlAs-GaAs superlattice,” Appl. Phys. Lett., vol. 54, pp. 4230-4231, 1983. G. J. Sonek, J. M. Ballantyne, Y. J. Chen, G. M. Carter, S. W. Brown, E. S. Koteles, and J. P. Salemo, “Dielectric properties of GaAs/AIGaAs multiple quantum well waveguides,” ZEEE J . Quan- tum Electron., vol. QE-22, pp. 1015-1018, 1986.

56, pp. 1886-1888, 1990.

1345-1347, 1990.

vol. 3, pp. 1338-1351, 1971.

Phys., vol. 63, pp. 479-483, 1988.

vol. 55, pp. 3376-3381, 1984.

[46] S. I. Hansen, J. H. Marsh, J. S. Roberts, and R. Gwilliam, “Refrac- tive index changes in a GaAs multiple quantum well structure pro- duced by impurity-induced disordering using boron and fluorine,”

[47] R. J. Manning, P. J . Bradley, A. Miller, J . S. Roberts, P. Mistry, and M. Pate, “Electric-field-dependent photoresponse of multiple quantum well modulator,” Electron. Lett., vol. 25, pp. 269-270, 1989.

Appl. Phys. Lett., vol. 58, pp. 1398-1400, 1991.

E. Bigan, photograph and biography not available at the time of publica- tion.

M. Allovon, photograph and biography not available at the time of publi- cation.

M. C a d , photograph and biography not available at the time of publi- cation.

C. Braud, photograph and biography not available at the time of publi- cation.

A. Carenco, photograph and biography not available at the time of pub- lication.

P. Voisin, photograph and biography not available at the time of publica- tion.


Recommended