10
PHYSICAL REVIEW 8 VOLUME 44, NUMBER 3 15 JULY 1991-, I Infrared transitions between shallow acceptor states in GaAs-Ga1 Al As quantum wells Samuele Fraizzoli Institut Rornand de Recherche Nurnerique en Physique des Materiaux (IRRMA), PIIB Ecu-biens, CII-2025 Lausanne, Switzerland and Scuola Norrnale Superiore, I-$6200 Pisa, Italy Alfredo Pasquarello Institut de Physique Theor~que, Ecole Polytechnique Federale de Lausanne, PIIB-Ecublens, CII-2025 Iausanne, Switzerland (Received 14 September 1990; revised manuscript received 11 February 1991) We calculate energies and oscillator strengths of infrared transitions between ground and excited shallow acceptor states in quantum wells (QW's) for varying well width. The impurity states are calculated within a four-band effective-mass theory, which accounts for the valence-band mixing as well as for the mismatch of the band parameters and the dielectric constants between well and barrier materials. The envelope function is expanded into a basis set consisting of products of two-dimensional hydrogeniclike functions and impurity-free QW eigenfunctions at k~~ —— 0. The present method is suited for s-type ground states as well as for p-type excited states. We obtain the absorption spectra for well widths ranging from 50 to 200 A. We find an overall increase of the transition energies for decreasing well widths. For polarizatioxi in the layer planes, the oscillator strengths of the lines corresponding to the bulk C and D lines maintain their oscillator strengths for decreasing well widths, whereas the lines corresponding to the bulk C line are very weak for small well widths. On the other hand, in the case of polarization along the QW axis, the oscillator strengths of the main lines decrease considerably as the well width decreases. We compare our results with absorption spectra of a recent experiment, and find a fairly good agreement. I. INTRODUCTION In the last decade growing interest has been devoted to the study of shallow impurities in GaAs-Gai Al As quantum wells (QW's). A good understanding of the eA'ects due to this kind of impurity is important not only from a fundamental point of view but also be- cause of technological applications. Shallow impurities can be present either because of intentional doping or because of uncontrolled growth conditions. In fact, a background concentration of impurities can be found in nominally undoped samples. Therefore a precise char- acterization of the impurities is necessary not only in order to fully exploit their properties but also to identify their presence. Owing to continuous improvements in the growth techniques as well as to a successful transfer of experimental techniques from the study of bulk materi- als to the study of heterostructures, it has been possible to gain detailed information on the properties of shal- low impurities in QW's. Extensive studies with a large variety of techniques such as photoluminescence, rnagnetospectroscopy, 5 far-infrared absorption, reso- nant Raman scattering, » and two-hole transitions, have been performed. These measurements have pro- vided accurate determinations of transition energies be- tween ground and excited states as a function of the well width. Furthermore selective doping has allowed one to study these properties for various impurity positions in the QW. Experiments have been performed on shallow donor states as well as on shallow acceptor states. Some previous calculations of impurity states in QW's were based on one-band eA'ective-mass models. ~ In these models, hydrogenic Hamiltonians, which have been applied successfully in bulk calculations, have been ex- tended to the case of QW's by including a square-well potential. These models are particularly suited for donor states, which are associated with the nondegenerate con- duction band. After the pioneering work by Bastard, who considered infinite barriers, these models have been implemented taking successively into account the mis- match of the band parameters and the dielectric con- stants between well and barrier materials. The problem of calculating energy levels and wave functions of a shallow acceptor in a quantum well is much more complicated than the similar problem for a donor. This point has been nicely discussed by Greene and Bajaj. In bulk GaAs the uppermost valence band, to which the acceptor states are associated, is fourfold degenerate at the center of the Brillouin zone. More- over, in the acceptor problem, band warping must be accounted for. The QW potential lifts in part the degen- eracy of the band producing a splitting between heavy and light holes, but the acceptor binding energy is so large that an accurate calculation has to include the full QW dispersion. Moreover, because of the more localized wave functions of the acceptor ground states the eA'ect of 1118 1991 The American Physical Society

As quantum wells

  • Upload
    alfredo

  • View
    226

  • Download
    11

Embed Size (px)

Citation preview

Page 1: As quantum wells

PHYSICAL REVIEW 8 VOLUME 44, NUMBER 3 15 JULY 1991-,I

Infrared transitions between shallow acceptor statesin GaAs-Ga1 Al As quantum wells

Samuele FraizzoliInstitut Rornand de Recherche Nurnerique en Physique des Materiaux (IRRMA), PIIB Ecu-biens,

CII-2025 Lausanne, Switzerlandand Scuola Norrnale Superiore, I-$6200 Pisa, Italy

Alfredo PasquarelloInstitut de Physique Theor~que, Ecole Polytechnique Federale de Lausanne,

PIIB-Ecublens, CII-2025 Iausanne, Switzerland(Received 14 September 1990; revised manuscript received 11 February 1991)

We calculate energies and oscillator strengths of infrared transitions between ground and excitedshallow acceptor states in quantum wells (QW's) for varying well width. The impurity states arecalculated within a four-band effective-mass theory, which accounts for the valence-band mixingas well as for the mismatch of the band parameters and the dielectric constants between well andbarrier materials. The envelope function is expanded into a basis set consisting of products oftwo-dimensional hydrogeniclike functions and impurity-free QW eigenfunctions at k~~

——0. Thepresent method is suited for s-type ground states as well as for p-type excited states. We obtainthe absorption spectra for well widths ranging from 50 to 200 A. We find an overall increase of thetransition energies for decreasing well widths. For polarizatioxi in the layer planes, the oscillatorstrengths of the lines corresponding to the bulk C and D lines maintain their oscillator strengthsfor decreasing well widths, whereas the lines corresponding to the bulk C line are very weak forsmall well widths. On the other hand, in the case of polarization along the QW axis, the oscillatorstrengths of the main lines decrease considerably as the well width decreases. We compare our resultswith absorption spectra of a recent experiment, and find a fairly good agreement.

I. INTRODUCTION

In the last decade growing interest has been devotedto the study of shallow impurities in GaAs-Gai Al Asquantum wells (QW's). A good understanding of theeA'ects due to this kind of impurity is important notonly from a fundamental point of view but also be-cause of technological applications. Shallow impuritiescan be present either because of intentional doping orbecause of uncontrolled growth conditions. In fact, abackground concentration of impurities can be found innominally undoped samples. Therefore a precise char-acterization of the impurities is necessary not only inorder to fully exploit their properties but also to identifytheir presence. Owing to continuous improvements in thegrowth techniques as well as to a successful transfer ofexperimental techniques from the study of bulk materi-als to the study of heterostructures, it has been possibleto gain detailed information on the properties of shal-low impurities in QW's. Extensive studies with a largevariety of techniques such as photoluminescence,rnagnetospectroscopy, 5 far-infrared absorption, reso-nant Raman scattering, » and two-hole transitions,have been performed. These measurements have pro-vided accurate determinations of transition energies be-tween ground and excited states as a function of the wellwidth. Furthermore selective doping has allowed one tostudy these properties for various impurity positions in

the QW. Experiments have been performed on shallowdonor states as well as on shallow acceptor states.

Some previous calculations of impurity states in QW'swere based on one-band eA'ective-mass models. ~ Inthese models, hydrogenic Hamiltonians, which have beenapplied successfully in bulk calculations, have been ex-tended to the case of QW's by including a square-wellpotential. These models are particularly suited for donorstates, which are associated with the nondegenerate con-duction band. After the pioneering work by Bastard,who considered infinite barriers, these models have beenimplemented taking successively into account the mis-match of the band parameters and the dielectric con-stants between well and barrier materials.

The problem of calculating energy levels and wavefunctions of a shallow acceptor in a quantum well ismuch more complicated than the similar problem for adonor. This point has been nicely discussed by Greeneand Bajaj. In bulk GaAs the uppermost valence band,to which the acceptor states are associated, is fourfolddegenerate at the center of the Brillouin zone. More-over, in the acceptor problem, band warping must beaccounted for. The QW potential lifts in part the degen-eracy of the band producing a splitting between heavyand light holes, but the acceptor binding energy is solarge that an accurate calculation has to include the fullQW dispersion. Moreover, because of the more localizedwave functions of the acceptor ground states the eA'ect of

1118 1991 The American Physical Society

Page 2: As quantum wells

INFRARED TRANSITIONS BETWEEN SHALLOW ACCEPTOR. . . 1119

central-cell corrections has to be considered.An attempt to address the acceptor problem includ-

ing the complicated valence-band structure was made byMasselink, Chang, and Morkoq who used a four-bandeffective-mass Hamiltonian. These authors were ableto calculate binding energies of the two s-type groundstates, using a symmetry-adapted basis set in r space.Subsequently, Pasquarello, Andreani, and Buczko useda similar eR'ective-mass model, but within a k-space ap-proach, in order to calculate excited acceptor states.However, the formalism turned out to be unsatisfactoryfor the strongly bound ground states, because the wavefunctions had been expanded on a basis which includedonly the discrete subband spectrum. Very recently wehave been able to include the eA'ect due to the subbandcontinuum yielding results for the s-type ground and ex-cited states. ~s The comparison of calculated 1s-2s transi-tion energies with Raman scattering and two-hole transi-tions measurements ' showed very good agreement. Ourresults also agree with recent calculations by Einevoll andChang, who obtained binding energies of s-type acceptorstates using an effective tight-binding method.

In this work we are interested in infrared transitionsbetween ground and excited acceptor states in GaAs-Gar Al As QW's. Infrared transitions between impu-rity states in QW's are more diKcult to observe with re-spect to the bulk casez because the absorption spectrapresent broader peaks on a stronger background. Themain reasons for this diA'erence are the following ones.In the QW the transition energy varies as a functionof the position of the impurity center. Therefore thedoped regions have to be small and as a consequencehigher doping concentrations than in the bulk are neededin order to observe the absorption lines. The relativesmaller distance between the impurity centers producesline broadening because of the increased interaction be-tween the impurities. Broadening also occurs because ofthe finite spatial width of the doped layers and becauseof the interface roughness or other defects of the sam-ple. Infrared transitions between QW donor states havebeen studied extensively from an experimental as wellas from a theoretical ' point of view and a satisfactoryagreement has been reached. On the other hand, muchless work has been done on infrared transitions betweenacceptor states in QW's: to our knowledge only one ex-perimental paper has been published on this subject untilnow. From a theoretical point of view, the calculation ofoscillator strengths of transitions between s-type groundand excited acceptor states and p-type excited acceptorstates requires accurate wave functions and even in thebulk case calculations have not been performed until veryrecently. 2r ~~ In the QW case, until now, there has beenno theory which gives wave functions of both ground andexcited acceptor states with a reasonable degree of ac-curacy. In this work we use the theory which we havepreviously developed for the calculation of the bindingenergies of s-type states and show that it is suited foIacceptor states of any symmetry. We are thus able to cal-culate oscillator st, rengths of infrared transitions betweenground and excited acceptor states.

In Sec. II we expose the theory which gives the ener-

gies and wave functions of acceptor states in QW's. Theresults are presented for well widths ranging from 50 to200 A. . The symmetry properties of the acceptor statesare brie8y discussed. Section III is devoted to the cal-culation of the oscillator strengths of the acceptor tran-sitions, for each of the two nonequivalent polarizations.The results are compared with experimental far-infraredabsorption spectra by Reeder et al Co.ncluding remarkscan be found in Sec. IV.

II. SHALLOW ACCEPTOR STATES

A. Theory

In this subsection we expose a comprehensive theoryfor the calculation of energy levels and wave functions ofacceptors in GaAs-Gar Al As QW's. A major virtueof the present theory is that, in contrast with othertheories, "~9 it gives accurate results for ground andexcited acceptor states as well as excited acceptor statesof any symmetry. We consider a quantum well grownin the [001) direction (z axis) with a single on-center (atz = 0) acceptor impurity. We solve the acceptor problemwithin a four-band efI'ective mass theory.

The acceptor Hamiltonian is a 4 x 4 matrix operator

IIkin + IIQw + Igt + II@

where II"'" represents the kinetic energy of the holes,II& the confinement potential due to the valence-banddiscontinuity, II+ the potential of the impurity and of theimage-charges due to the dielectric mismatch, and II' theself-energy term, also due to the dielectric mismatch. InHamiltonian (1) the kinetic energy of the hole is chosento be positive.

The kinetic energy term H"'" describes the dispersionof the bulk I'8 valence band. H"'" is quadratic in k =—iV' and is given by the I uttinger-Kohn Hamiltonian. ~

Inside the quantum well (~z~ ( I/2) the l,uttingerparameters correspond to those of the well material, inthe barrier to those of the barrier material. We use theI uttinger-Kohn Hamiltonian in the axial approximation,which has been shown to be accurate in the bulk casewithin 0.5 meV. The confinement potential H~ is asquare-well potential of barrier height V and well widthL. The Coulomb potential of a point charge in a systemof three dielectrics separated by two infinite planes is con-tained in H . In order to satisfy the Maxwell boundaryconditions H+ must contain an infinite series of imagecharges. In our method of solution we will take thiscomplete series explicitly into account. The self-energyterm H' represents the interaction between the hole andthe polarization charge induced at the interfaces of theQW by the hole itself, because of the dielectric mismatchbetween well and barrier materials. ~4 Vfe have neglectedthis interaction because it shifts the energy levels of thehole in both the acceptor and the subband states, andtherefore an efkctive cancellation of this eA'ect occursin the Coulomb binding energies. We have checked thispoint in first-order perturbation theory, finding negleg-ible contributions (less than 0.1 meV) for wells ranging

Page 3: As quantum wells

1120 SAMUELE FRAIZZOLI AND ALFREDO PASQUARELLO

F "(p, e, z) = e'& 'l'f "(p, z). (2)

Because of the angular momentum of the Bloch functionsa phase factor e'& '& appears in the envelope functionand therefore its spin components have different trans-formation properties under rotations around the z axis.We expand the functions f ' into a basis set of functionswhich are separable in the coordinates p and z:

f '(C»z) = ).&. '(~)~.'(z).

from 50 to 200 A. .We discuss now corrections which are not accounted

for by Harniltonian (1), namely the efFects due to thesplit-off band, to the spatially-dependent screening, andto the particular chemical properties of the consideredimpurity. All these corrections are effective very nearto the acceptor site. They affect mainly the ground statebecause its wave function is more localized, with the high-est density on the acceptor site. Acceptor states of p-typesymmetry, which have a vanishing wave function on theimpurity site, are barely affected. The inclusion of thesecorrections produces a twofold effect on the spectra calcu-lated here. First, the transition energies are all shifted byan amount corresponding to the shift of the ground state.For the case of Be acceptors, as discussed in Ref. 18, thedifferent corrective effects almost cancel and the ground-state shift can be neglected to a first approximation in thewell width range considered here. Second, the modifica-tion of the ground-state wave function affects directly theoscillator strengths. However, a study of these correctiveeffects in the bulk has shown that although the absolutevalues of the oscillator strengths are considerably modi-fied, their relative values remain almost unchanged.We expect that the same conclusions hold for the QW.

The effective-mass Hamiltonian we have used is invari-ant for rotations around the z axis and for inversion withrespect to the center of the QW. The acceptor eigenstatescan be classified by two quantum numbers: the z com-ponent of the total angular momentum m and the parity(+) with respect to inversion. The acceptor levels aretwofold degenerate because of time-reversal symmetry.The two states of the Kramers doublet have opposite mbut the same parity, and the doublet can therefore unam-biguously be labeled by (~m~, +). In order to distinguishbetween the two states of a doublet, it is useful to con-sider the operator u, which represents the reflection in thez = 0 plane. 26 The two states of a doublet can be chosento have opposite parity with respect to o. The relationbetween the symmetry group D2d, which accounts for thezinc-blende structure, and the symmetry group D p ofthe efFective mass Hamiltonian (1) is given in Ref. 17.

The Hamiltonian (1) acts on a four-component enve-lope function I". Each component I"' is labeled by thespin index s running from 2 to —2. The s componentof an acceptor envelope function of definite angular mo-mentum I reads

1

(p2 + z2)1~2 dq e—I'l~ Jo(pq)

where Jo is the Bessel function of order zero. With thistransformation all the integrals over the space coordi-nates are calculated analytically: only the remaining in-tegral over the auxiliary variable q is calculated numeri-cally. In this way we are able to include as many as 40 g„functions and up to 9 exponentials in Eq. (4). This largenumber of basis functions turns out to be necessary forthe calculation of the ground states, which are stronglybound. The number of basis functions can be reduced inthe calculations of states of different symmetry. The useof a set of basis functions which are separable in the zand in-plane (p, 0) coordinates was originally suggestedby the quasi-two-dimensional character of the QW sys-tem with narrow well width. However, this approachturned out to be very effective also in the case of verylarge well widths. By a comparison with the results forthe bulk case, we are able to estimate that the conver-gence of our calculations for the binding energies of theacceptor states is within 0.2 meV. As we will see in thenext section, a further advantage of our basis set is thatalso the transition matrix elements, which are needed forthe calculation of the oscillator strengths, can be carriedout analytically.

B. Results

QW subband state at k~~= 0. The envelope function

g„ is a solution of the Hamiltonian II '" + II~ whichdescribes the impurity-free QW system. The basis setof functions g„ is complete if discret, e as well as contin-uum states of the impurity-free QW are included. Thelatter states are accounted for by discretizing the con-tinuum introducing sufIiciently distant infinite barriers.It is necessary to include continuum states in order toachieve numerical convergence for the binding energiesof the acceptor ground states. The g„satisfy currentconserving boundary conditions at the interfaces. 2

The radial functions R„"(p) in Eq. (3) are developedin an expansion of functions with the correct asymptoticbehavior for large and small p:

~m, s( ) (rn s[ )— gm, & —n(p (4)

where the exponents o, ~ are chosen in a geometrical seriesand cover the relevant physical region, and where the co-eKcients A„&" are taken as variational parameters. Theeigenvalue problem can be turned into a linear problemfor the parameters A„&". We note that, for a given sym-metry, the lowest state as well as the excited ones arecalculated at the same time.

A great virtue of the present basis set is that mostof the integrals which appear in the matrix elements ofHamiltonian (1) can be carried out analytically. In thecalculation of the matrix elements of the Coulomb poten-tial an auxiliary integral which decouples p and z coordi-nates is introduced, using the well-known transformation

The function g„' is chosen to be the s component of thefour-component envelope function g„, which describes a

In this subsection we present numerical results for thebinding energies of on-center acceptors in QW's. We have

Page 4: As quantum wells

44

= 6.85, pg ——2. ,= 2-12 53 for GaAs,—1

t k thethe ie e a

interpolation.

on'

and valence band oonduction anbetween con'

an

resed.

he energy e11

and 2 we h

Th tor ener-t accepto

f the erst

imp urity-freef the ground tsta esgies of

exci

) ('+) d( )f the reduced sy

1 te bulk grounecause o e

d statel

ld degeneratestates: t e moo old degenerate s: e mo

d tate of (in o

g

ited statesround s athe light-hole gr d

re orted andtr have been repor' g

esults are in goodR f. 18. These resu s a dd reviouslydiscusse pt with exper'agreemen

10

5

I

100I

150050 200

L(A)

d arity force tor states of od pgy o ptoIG. 2. Binding en

a functionas

pthe o orespect to

ALLOW ACC EPTQRBETWEEN SHTRANSITIONSINFRARED T

I

30

20

050

~as% ~

~ W

fI

100L(A)

I

150 200

we resent also

and (2, +) syan ~ s mmetry as we a

of the lowesto —' state.

energies olk

n'

. we plot the b dngc tates of odd syrnm

I's) and 2Ps(2 s ib lk 2P (I'dou

degenerate and dotted-das he cur

' el.s mmetry, p

t

energies of t

0QW subband eig ns. "

variationa methe present vard even to imprresults an e

t s of even par-or states o

)with respect to t e

ABSORP TIONIII. INFRARED AB

ths ofoscillator strengthculate the osci a on'

tion we calcu oscIn this sec io

t thtransi iotions from g

f t}1ates inexcite ace a

infrared absorption

Page 5: As quantum wells

1122 SAMUELE FRAIZZOLI AND ALFREDO PASQUARELLO

A. Theory B. Results

The absorption cross section, defined as the absorbedenergy per unit time divided by the electromagnetic flux,is given by~~

2~2h2p ( e2 l0(~, e) = '

~ ~ ) f o(&') ~(E' —Eo —&~),nmo /he)

where n is the refractive index. The sum is over all thedoublets of Kramers degenerate excited acceptor statesand the label 0 indicates the ground-state doublet of theacceptor. In Eq. (6) f;o is the oscillator strength of tran-sitions from the ground doublet to the excited doubletlabeled by i,

(7)

where E0, F0& and E;, I","k, are the energies and envelopefunctions of ground and excited states, respectively, g0 ——

2 is the degeneracy of the ground state, and e is thepolarization vector of the electromagnetic radiation. Thesum indices k, k' indicate the two components of theground and excited doublet, respectively.

We now give the selection rules for the electromagnetictransitions. ~ Only transitions between acceptor statesof opposite parity with respect to inversion are allowed.Therefore only odd states can be reached starting fromthe even ground states. For z polarization, transitionsare allowed only between acceptor states with Am = +1and of the same parity with respect to the o reflection.On the other hand, z-polarized radiation induces onlytransitions between states with Lrn = 0 and of oppositeparity with respect to o. . It should be noted that theselection rules for a D2p symmetry group allow additionaltransitions. However, these transitions are expected tohave smaller oscillator strengths.

In the bulk, for any given polarization e, the oscillatorstrengths as defined in Eq. (7) satisfy a sum rule. si Inthe QW, because of the nonequivalence of the in-planeand z directions such a sum rule does not hold. It ishowever possible to derive the following sum rule (see theAppendix), which relates oscillator strengths for differentpolarizations to each other:

40X POLARIZATION

30

20

In order to interpret the absorption spectrum it is use-ful to start from the bulk limit. In the bulk, theabsorption spectrum is characterized by a strong dou-blet, whose components have been called D and C. Atlower energies another line is found which is called G andis considerably weaker. The transition energies and theintensities 2 of these lines have been explained. In thiswork we study QW's with well widths ranging from 50to 200 A. In the bulk the wave function of the acceptorground state decreases at large distances as e "~"' with

25 A. .s2 Therefore, we expect that the acceptor spec-trum at 200 A can be interpreted starting from the bulklimit. As we have seen in the previous section, the bulkstates split in the QW because of the lower symmetry. Asa consequence, the G, D, and C lines split into differentcomponents in the QW.

In Fig. 3 we plot the energies of the allowed transi-tions in the case of z polarization in GaAs-Gai Al AsQW's as a function of the well width. Analogously, theenergies of the allowed transitions in the case of z polar-ization are plotted in Fig. 4. The QW lines have beenlabeled with G, D, or C according to the bulk line towhich they correspond. We note that the transition en-ergies generally increase for decreasing well widths. This

(8)

50 100 150 200This sum rule can be used to check the precision of ourcalculation. In fact, we cannot perform the sum over allthe exact eigenstates which appears in Eq. (8) as we diag-onalize the Hamiltonian on a finite basis. In a variationalcalculation only the solutions of lowest energy accuratelyrepresent the true eigenstates. Variational solutions ofhigher energy have in general no specific meaning. How-ever, it is interesting to observe that if we take the sumin (8) over the whole set of variational solutions, we findthat the sum rule is verified within 4%%uo in all cases.

L (A)

FIG. 3. Transition energies of the allowed transitions in xpolarization between ground and excited acceptor states foran on-center impurity in GaAs-Gao 7Alo 3As quantum wells,as a function of the well width. Solid and dashed curvesrepresent transitions from the (2, +) (heavy-hole) and (2, +)(light-hole) ground state, respectively. Dotted curves repre-sent the onset of the transitions between the ground statesand the continuum of the first heavy-hole subband.

Page 6: As quantum wells

INFRARED TRANSITIONS BETWEEN SHALLOW ACCEPTOR. . . 1123

40

230

0

20

50 100L(A)

I

150 200

FIG. 4. Transition energies of the allowed transitions in zpolarization between ground and excited acceptor states foran on-center impurity in GaAs-Gao 7Alo 3As quantum wells,as a function of the well width. Solid and dashed curvesrepresent transitions from the (-, +) (heavy-hole) and (-,+)(light-hole) ground state, respectively. Dotted curves repre-sent the onset of the transitions between the ground statesand the continuum of the first heavy-hole subband.

is mainly due to the increase of the binding energies of theground states (see Fig. I), which is more significant thanthe variation of the binding energies of the excited states(see Fig. 2). The relative positions of the lines changesignificantly as a function of the well width. We notethat in z polarization QW lines which correspond to thebulk C line cannot be excited starting from the heavy-hole ground state but only can from the light-hole groundstate. In z polarization the increase of the transition en-ergies from the heavy-hole ground state is considerablylarger than in the case of z polarization. In particularfor small well widths the lines which correspond to thebulk C line occur at energies very near to the onset ofthe transitions from the heavy-hole ground state to thecontinuum of the first heavy-hole subband. This can beunderstood as follows. The lines which correspond to theC line originate from transitions from the ground stateto excited states of (2, —) symmetry. These states aremainly associated with the light-hole subband which ispushed towards higher energies when the well width de-creases. Hence the corresponding transition energies in-crease. Moreover, the contribution of the light-hole sub-band to the lower (2, —) acceptor states decreases andthese states are mainly associated with the heavy-halesubb and.

bQ

CCO

61w3/2x

2

z

050 100

L(A)150 200

FIG. 5. Oscillator strengths of the acceptor transitionscorresponding to the bulk G line in GaAs-Ga0, 7Alo, 3As quan-tum wells as a function of the well width.

We present in Figs. 5, 6, and 7 results for the oscilla-tor strengths of the lines which correspond to the G, D,and C lines of the bulk spectrum, respectively. Solid anddashed lines are used for transitions from the heavy- andlight-hole ground state, respectively. In Fig. 5 we plot asa function of the well width the oscillator strengths of allthe lines corresponding to the bulk G line. The oscilla-tor strengths of the five lines are rather small and of thesame order of magnitude as in the bulk. The line Q,l~1/2

has a negligible oscillator strength in the scale of Fig. 5and has not been reported. In Fig. 6 we plot the oscil-lator strengths of the lines corresponding to the bulk Dline. These lines are the strongest of the QW absorptionspectrum. The oscillator strengths of transitions whichare allowed in z polarization are almost independent ofthe well width. Only the line D, which is allowed inz polarization loses its oscillator strength for decreasingwell widths. In Fig. 7 we give the oscillator strengthsof the lines which correspond to the bulk C line. Theline labeled with C', / represents the transition to thethird level of (z, —) symmetry, corresponding to the bulk3Ps/g [I's] state. In the bulk this transition occurs at al-most the same energy as the transition to the 2Ps/2[I'7]state and contributes to the intensity of the C line. Inthis figure the oscillator strengths of the main lines de-crease for decreasing well width. This is understood asthe acceptor states of (2, —) symmetry corresponding tothe bulk C line are mainly associated with the first lighthole subband. For narrow well widths, as these states arepulled towards the continuum of the first heavy-hole sub-band, their oscillator strengths are transferred to higherstates because of mixing eKects.

When the binding energies of the ground and excitedacceptor states and the first-excited acceptor states arereferred to the edge of the bulk valence band, it is foundthat at 200 A. they are close to the bulk limit. s The com-parison between QW oscillator strengths and their bulkvalues is not as straightforward. The oscillator strengthsof all QW transitions which correspond to the same bulktransition must be considered together. 3~ We find that

Page 7: As quantum wells

1124 SAMUELE FRAIZZOLI AND ALFREDO PASQUARELLO

20 Dh-+5/2X

bQ

0

00

15

10

Dial/2X

5

Dl~ 1/2Zr

//

//

//

//

Dh~ 1/2X

050 100

L(A)150 200

FIG. 6. Oscillator strengths of the acceptor transitionscorresponding to the bulk D line in GaAs-Ga0. 7A10.3As quan-tum wells as a function of the well width.

20 D

bQC

0iD

10 D

==D

ground state, respectively. In the figures the onset of thetransitions between the ground states and the continuumof the first heavy-hole subband is indicated by thin verti-cal lines. The main difference between the two polariza-

at 200 A the correspondence is not as close as for thebinding energies. In order to check the reliability of ourcalculations in the limit of large well widths, we haveperformed a calculation in the spherical approximation(yi ——6.85, p = 0.7) for which bulk oscillator strengthsare provided in literature. ~i We have considered a QWwith a well width of 265 A. For the G line, we obtainoscillator strengths of 9.6 x 10 for z polarization and8.0 x 10 for z polarization which have to be comparedwith the bulk value of 8.25 x 10 . We are not able todetermine to what extent the residual discrepancy is dueto the anisotropy of the QW system or to the finite vari-ational basis set of our calculations. We conclude thatthe oscillator strengths for well widths ranging from 50to 200 A. , presented here, are reliable at least within thisaccuracy.

In order to reproduce the absorption spectrum in thecase of z polarization we show the oscillator strengths asa function of the transition energy in Fig. 8, for variouswell widths. The same has been done for z polarizationin Fig. 9. Solid and dashed lines indicate the oscillatorstrengths of transitions from the heavy- and light-hole

C

0CD

bQ

0iD

20

10

p ~

20

10

20

D

D

== D

I==:—~ ~ i

D

D

~ I'I

~ I~ I~ I~ I~ I

I~ II II I~ II II III ~~ II I~ ~I II ~I II ~~ ~I ~~ I~ ~~ ~~ ~I ~I ~I ~I II ~~ I~ I

I I~ II II1I I~ I~ II II ~~ II ~I II II II II II II II II II II II II II II II III II II I

D

i III

IIIIIII

II

\IIIII

I

III

IIIIIII

IIII

II

\

IIIIIIIIIIIIIIII

III

I

(b)

(d)

bQ

C

20

15

10

C1~3/2X

0

CD

10

20

D

DcI

30Transition energy ( meV )

40

50 100L(A)

150 200

FIG. 7. Oscillator strengths of the acceptor transitionscorresponding to the bulk C line in GaAs-Ga0. 7Alo 3As quan-tum wells as a function of the well width.

FIG. 8. Oscilla. tor strengths of the lowest lines in the ac-ceptor spectrum for x polarization as a function of the transi-tion energy in GaAs-Ga0, 7A10.3As quantum wells. Solid anddashed lines indicate the oscillator strengths of transitionsfrom the (~, +) and (2, +) ground state, respectively. Thindashed lines indicate the onset of the transitions between theground states and the continuum of the first heavy-hole sub-band. The well width is (a) 200 A, (b) 150 A, (c) 100 A, and

(d) 50 A.

Page 8: As quantum wells

INFRARED TRANSITIONS BETWEEN SHALLOW ACCEPTOR. . . 1125

tions is that the spectrum in z polarization is in generalmaintained for decreasing well widths whereas in the caseof z polarization the intensity of the lines is suppressedfor small well widths. We remark that this observationrefers to the oscillator strengths of the lowest transitions,which are accurately given by our variational method andare presented in the figures. In the case of z polarization

(Fig. 8), the D is always the strongest line of thespectrum. As mentioned above, the line which corre-sponds to the bulk C line can only be reached startingfrom the light-hole ground state. Moreover, the oscillatorstrength of this line decreases for decreasing well widths.In the case of z polarization we expect that transitionsfrom the heavy-hole ground state give rise to a strong

20D (a)

bQ

OCD

10

C

0 ~

20 (b)

bQC

O&D

10

G0 ~

D

20

bQ

~ m

O

~ I

Dac

20 30Transition energy ( meV )

40

FIG. 9. Oscillator strengths of the lowest lines in the ac-ceptor spectrum for z polarization as a function of the transi-tion energy in GaAs-Ga0 7Alo. 3As quantum wells. Solid anddashed lines indicate the oscillator strengths of transitionsfrom the (2, +) and (2, +) ground state, respectively. Thindashed lines indicate the onset of the transitions between theground states and the continuum of the first heavy-hole sub-band. The well width is (a) 200 A, (b) 150 A, and (c) 100 A..

doublet as long as the well is not too narrow. In thebulk limit these two peaks correspond to the C line. Thecase of 50-A well width has not been reported in Fig. 9because in this case all the transitions have a negligibleoscillator strength on the scale of the figure.

In order to compare our results with experimentalspectra we have to consider the relative occupation of theheavy- and light-hole ground states. In fact the relativeintensity of the transitions from the two ground statesdepends not only on their oscillator strengths but alsoon the relative population of the ground states, which isdetermined by the Boltzmann factor. The energy sepa-ration between the ground states as a function of the wellwidth, can be obtained from the curves plotted in Fig. 1.

The only far-infrared absorption experiment on QW'shas been performed by Reeder et al. .7 In this experimentabsorption spectra of z-polarized radiation have been ob-tained at a temperature T = 4.2 K in Be-doped QW'swith well widths of 100, 150, and 200 A. . It is rather dif-ficult to interpret the experimental re ults because thebackground is strong and quite diA'erent for each sample.We limit ourselves to the spectra of the 150- and 200-A samples which show more evident peaks, and com-pare them with Figs. 8(a) and 8(b), respectively. As wehave already mentioned in the previous section, we ex-pect that, for Be impurities in this well width range, cor-rections not included in our pure Coulombian calculationdo not modify the energies of the transitions and leavetheir relative intensities unchanged. Therefore we com-pare theoretical and experimental results without consid-ering a global shift of the computed transition energies.The spectra at 200 and 150 A. present a major featureat 22 and 24 meV, respectively, which is consider-ably broadened on the low-energy side (particularly inthe second spectrum). We identify this feature with theD line, which from Fig. 8 is expected to have the highestoscillator strength. Because of the lower QW symmetry,the bulk D line splits in three components. At 200 A.

they are close together. We find the strongest compo-nent at 21.4 meV and the other two at 22.0 and 22.2meV. At 150 A the three components of the D line aremore separated, the strongest being at 22.5 meV and theothers at 24.2 and 24.7 meV. At higher energies than theD line a smaller feature is recognizable in both the 200-and 150-A spectra, at 24 and 26 meV, respectively.We identify this line with the C line, which we predictat 24.2 and 26.3 meV, respectively. In the spectrum ofthe 200-A sample, at lower energies than the D line, asmall feature at 17 meV can be identified as the G line,which we predict at about 16 meV (there are three com-ponents rather close to each other). In the other sample,because of the noise of the background, we are not able torecognize an analogous feature in the same energy region.

In conclusion, the comparison with the experiment byReeder e7 a/. is encouraging. However, because of itslimited resolution it does not constitute a definitive testof the present theory and other experiments are calledfor. An analysis of the intensity of the observed linesversus temperature as well as of the separation of thelines in a magnetic field may be useful for identifying thepeaks in the spectrum.

Page 9: As quantum wells

1126 SAMUELE FRAIZZOLI AND ALFREDO PASQUARELLO

IV. CONCLUSIONS ACKNOWLEDC MENTS

We have calculated energies and oscillator strengths oftransitions between ground and excited acceptor statesand excited shallow acceptor states in GaAs-Gai Al Asquantum wells. The acceptor states have been calcu-lated variationally within a four-band eA'ective mass the-ory which accounts for the mixing in the valence bandand for the mismatch of the material parameters at theinterfaces of the QW. i This theory is able to yield ac-curate wave functions for ground and excited acceptorstates as well as excited acceptor states, which is an es-sential requisite for the calculation of oscillator strengths.

We have obtained oscillator strengths for z as well asfor z polarization. The acceptor spectra for the two po-larizations are identical in the bulk by symmetry. In thequantum well, the two polarizations are no longer equiva-lent and a considerable polarization dependence is foundin the spectra. For both polarizations a general shift to-wards higher energies of all the transitions is found for de-creasing well widths. This shift can mainly be attributedto the increase of the binding energies of the two groundstates.

In the case of z polarization, the absorption spectrumis essentially similar to the bulk case. For narrow wellwidths some lines corresponding to transitions to excitedstates associated with the light-hole subbands decreasestrongly. Another diA'erence with respect to the bulk caseis that the transitions from the light-hole ground statecan only be observed if this state is suKciently populated.The splitting between the two ground states can be usedin order to estimate their relative population at a giventemperature. For example, the intensity of the QW linewhich corresponds to the bulk C line is strongly reducedat low temperature.

The situation is diA'erent in the case of z polarization.In this case the intensities of the lines decrease for de-creasing well widths. At, 50 A. the lines are expected tobe too weak to be observable. On the other hand, inthe limit of large well widths (150—200 L), there are twostrong QW lines which correspond to the bulk C line.This doublet originates from transitions from the heavy-hole ground state and is therefore expected to be easilyobservable.

At present, only one experiment on infrared absorptionin acceptor-doped quantum wells has been performed.The absorption spectra have been obtained only in thecase of x polarization. Comparison between theory andthis experiment is encouraging but not completely satis-factory because of the low experimental resolution. Asthe present theory is able to reproduce with great ac-curacy the measured transition energies between 18 and28 acceptor states, ' we believe more experimental dataare required in order to test our predictions also for tran-sitions between ground and excited acceptor states andp-type excited acceptor states in quantum wells.

We are grateful to R. Buczko for an important contri-bution in the early stages of the work. We also acknowl-edge interesting and fruitful discussions with L. C. An-dreani, A. Baldereschi, F. Bassani, N. Binggeli, and S.Rodriguez. In particular, we wish to thank L. C. An-dreani for a critical reading of the manuscript. One ofthe authors (S.F.) acknowledges partial support from theSwiss National Science Foundation under Grant No. 20-5446.87 and from the European Economic Communityunder Contract No. ST23-0254-1-1.

AP PEND IX: SUM RULE

For bulk systems described by Hamiltonians whose ki-netic part is the Luttinger-Kohn Hamiltonian, a sum rulefor the oscillator strengths holds. The purpose of thisappendix is to generalize this sum rule in order to copewith QW systems.

Using the double commutator of a spatial coordinatex& with the Hamiltonian as in the derivation of the usualThomas-Reiche-Kuhn sum rule, we derive

(Al)

where the sum over It,' is over the degenerate states of

a same multiplet. The identity (Al) holds for the bulksystem as well as for the QW.

A more useful sum rule may be obtained applyingthe following result of group theory which can easily beworked out:

where, for example, I" indicates that the envelope func-tion (or the operator in the case of I'y) transforms asthe ith row of the irreducible representation I' of thesymmetry group of the Hamiltonian, and g is the de-generacy of I' . In the bulk, the symmetry group of theHamiltonian is Oi„and application of (A2) gives

(A3)

where gp —4 is the degeneracy of the ground state. Inthe QW, because of the lower symmetry, the z directionis no longer equivalent to the in-plane directions z andy. Application of Eq. (A2) leads to the result stated inthe text in Eq. (8).

R. C. Miller, A. C. Gossard, W. T. Tsang, and O.Munteanu, Phys. Rev. B 25, 3871 (1982).X. Liu, A. Petrou, B. D. McCombe, 3. Ralston, and G.Wicks, Phys. Rev. B 38, 8522 (1988).

P. O. Holtz, M. Sundaram, R. Simes, j. I. Merz, A. C.Gossard, and J. H. English, Phys. Rev. B 39, 13 293 (1989).P. O. Holtz, M. Sundaram, K. Doughty, j. L. Merz, and A.C. Gossard, Phys. Rev. B 40, 12338 (1989); G. C. Rune,

Page 10: As quantum wells

INFRARED TRANSITIONS BETWEEN SHALLOW ACCEPTOR. . . 1127

P. O. Holtz, M. Sundaram, J. L. Merz, A. C. Gossard. , andB. Monemar (unpublished).X. Liu, A. Petrou, A. L. Moretti, F. A. Chambers, and G.P. Devane, Superlatt. Microstruct. 4, 141 (1988).R. J. Wagner et al. , in G'aAs and Related Compounds, Biar-ritz, 198$, Proceedings of the 11th International Sympo-sium, IOP Conf. Proc. No. 74 (Institute of Physics andPhysical Society, London, 1985), p. 315; N. C. Jarosik etal. , in Proceedings of the 17th International Conference onPhysics of Semiconductors, San Francisco, 198$, edited byD. J. Chadi and W. Harrison (Springer, New York, 1985),p. 507; N. C. Jarosik, B. D. McCombe, B. V. Shanabrook,j. Comas, J. Ralston, and G. Wicks, Phys. Rev. Lett. 54,1283 (1985); B. V. Shanabrook, Surf. Sci. 170, 449 (1986);N. C. Jarosik et al. , ibid. 170, 459 (1986).A. A. Reeder, B. D. McCombe, F. A. Chambers, and G. P.Devane, Phys. Rev. 38, 4318 (1988).B. V. Shanabrook, J. Comas, T. A. Perry, and R. Merlin,Phys. Rev. B 29, 7096 (1984); T. A. Perry, R. Merlin, B.V. Shanabrook, and J. Comas, Phys. Rev. Lett. 54, 2623(1985).D. Gammon, R. Merlin, W. T. Masselink, and H. Morkoq,Phys. Rev. B 33, 2919 (1986).G. Bastard, Phys. Rev. B 24, 4714 (1981).C. Mailhiot, Yia-Chung Chang, and T. C. McGill, Phys.Rev. B 26, 4449 (1982).R. L. Greene and K. K. Bajaj, Solid State Commun. 45,825 (1983); Phys. Rev. B 34, 951 (1986); R. L. Greene andP. Lane, ibid 34, 86. 39 (1986).S. Fraizzoli, F. Bassani, and R. Buczko, Phys. Rev. B 41,5096 (1990).R. L. Greene and K. K. Bajaj, Solid State Commun. 53,1103 (1985).W. T. Masselink, Y.-C. Chang, and H. Morkop, Phys. Rev.B 28, 7373 (1983); 32, 5190 (1985).J. M. Luttinger and W. I(ohn, Phys. Rev. 97, 869 (1955).A. Pasquarello, L. C. Andreani, and R. Buczko, Phys. Rev.B 40, 5602 (1989).' S. Fraizzoli and A. Pasquarello, Phys. Rev. B 42, 5349

(199O).G. T. Einevoll and Y.-C. Chang, Phys. Rev. B 41, 1447(199O).R. A. Cooke, R. A. Hoult, R. F. Kirkman, and R. A.Stradling, J. Phys. C ll, 345 (1978); R. F. Kirkman, R.A. Stradling, and P. J. Lin-Chung, ibid. 11, 419 (1978).N. Binggeli and A. Baldereschi, Solid State Commun. 66,323 (1988); N. Binggeli, A. Baldereschi, and A. Qu at-tropani, in Shallow Impuri ti es in Semiconductors 1988,edited by B. Monemar (IOP, Bristol, 1989), pp. 521—526.R. Buczko and F. Bassani, in Shallow Impurities in Semi-conductors 1988 (Ref. 21), pp. 107—112.J. M. Luttinger, Phys. Rev. 102, 1030 (1956).M. Kumagai and T. Takagahara Phys. Rev. B 40, 12359(1989).N. Binggeli, Ph. D. thesis, Ecole Polytechnique Federale,Lausanne, 1990.L. C. Andreani, A. Pasquarello, and F. Bassani, Phys. Rev.B 36, 5887 (1987).

"This point is reviewed in L. C. Andreani, S. Fraizzoli, andA. Pasquarello, Phys. Rev. B 42, 7641 (1990).M. Altarelli, in IIeterojunctions and Semiconductor Super-lattices, edited by G. Allan, G. Bastard, N. Boccora, andM. Uoos (Springer, Berlin, 1986).Numerical Data and Functiona/ Relationships in Scienceand Technology, edited by O. Madelung, Landolt-Bornstein,New Series, Group III, Vol. 17 (Springer, Berlin, 1982).H. C. Casey and M. B.Panish, IIeterostructure Lasers (Aca-demic, New York, 1978).S. M. Kogan and A. F. Polupanov, Solid State Commun.27, 1281 (1978).A. Baldereschi and N. O. Lipari Phys. Rev. B 8, 2697(1973).In order to obtain these binding energies, the quantizationenergy of the first heavy-hole subband has to be subtractedfrom the binding energies given in the present work.In fact they have to be summed and divided by two. Thefactor 2 accounts for the average of the oscillator strengthsof transitions from heavy- and light-hale ground states.