<<

RAPID COMMUNICATIONS

PHYSICAL REVIEW B VOLUME 58, NUMBER 20 15 NOVEMBER 1998-II

Short-range versus long-range -hole exchange interactions in semiconductor quantum dots

A. Franceschetti, L. W. Wang, H. Fu, and A. Zunger National Renewable Energy Laboratory, Golden, Colorado 80401 ͑Received 25 August 1998͒ Using a many-body approach based on atomistic pseudopotential wave functions we show that the electron- hole exchange interaction in semiconductor quantum dots is characterized by a large, previously neglected long-range component, originating from monopolar interactions of the transition density between different unit cells. The calculated electron-hole exchange splitting of CdSe and InP nanocrystals is in good agreement with recent experimental measurements. ͓S0163-1829͑98͒51144-7͔

One of the most intriguing features of the spectroscopy of clear, however, to what extent the exchange interaction in semiconductor quantum dots is the energy shift between the quantum dots is affected by dielectric screening. Recent di- zero-phonon absorption and emission peaks observed in size- rect calculations8,15 for InP and CdSe nanocrystals revealed selective spectroscopies of monodispersed samples.1–7 The that the unscreened splitting is significantly larger redshift of the emission peak, of the order of 10 meV, has than the measured redshift, suggesting that the screening of been measured in Si,1 CdSe,2–5 InP,6 and InAs ͑Ref. 7͒ the LR interaction may be important. nanocrystals grown with different techniques. Although sev- ͑iv͒ How to calculate the exchange splitting. In standard eral models have been proposed to explain the redshifted approaches based on the effective-mass approximation emission, recent measurements and calculations for CdSe ͑EMA͒ the e-h exchange interaction is described by a short- ͑Refs. 2–5͒ and InP ͑Refs. 6 and 8͒ nanocrystals indicate range phenomenological Hamiltonian acting on the exciton that, as first proposed by Calcott et al.,1 the redshift origi- envelope function:2–4 nates from the exciton splitting induced by the electron-hole (e-h) exchange interaction. According to this model, absorp- HEMAϭϪ␰ a3 ⌬E ␦ r Ϫr ␴ J , ͑1͒ tion takes place in a -allowed state, while emission oc- exch x x ͑ h e͒ • curs from a lower energy spin-forbidden state. There are four open issues regarding the e-h exchange where ␰ is a structure-dependent parameter (␰ϭ␲/4 for cu- interaction in semiconductor quantum dots. bic lattice structure and ␰ϭ␲/3 for hexagonal lattice struc- ͑i͒ The magnitude of the long-range exchange interaction. ture͒, ␴ and J are the electron and hole spin operators, ax is In general, the effective exchange interaction contains a the bulk exciton radius, and ⌬Ex is the bulk exchange split- short-range ͑SR͒ component, which decays exponentially ting. This model relies heavily on the knowledge of ax and 2 with an effective length comparable with the bulk lattice ⌬Ex ͓ϳ0.1 meV, 0.001 meV ͑Ref. 7͔͒, which are often constant, and a long-range ͑LR͒ component, which decays subject to large experimental uncertainties. The solutions of instead as a power law and extends over several lattice the model Hamiltonian ͑1͒ fit well the observed redshift in constants.9 In bulk periodic semiconductors the LR interac- CdSe nanocrystals,2–4 but in the case of spherical zinc- tion is responsible for the nonanalytic behavior of the exciton blende quantum dots Eq. ͑1͒ predicts a 1/R3 scaling of the dispersion as the exciton wave vector approaches zero,10 redshift with size,3 which is not observed in either InP ͑Ref. which is observed spectroscopically as a longitudinal- 6͒ or InAs ͑Ref. 7͒ nanocrystals where a ϳRϪ2 dependence transverse splitting. However, the magnitude of LR exchange is seen. While the prefactor of Eq. ͑1͒ has been recently interactions in zero-dimensional quantum dots is still an questioned,16 the inability of the conventional EMA to repro- open question. duce the correct scaling law is particularly troublesome. ͑ii͒ The physical origin of the LR exchange interaction. In this work we investigate the nature of the e-h exchange Conventional wisdom2–4,11 suggests that the LR exchange interaction in semiconductor quantum dots using a many- interaction in quantum dots originates, as in bulk semicon- body approach based on atomistic pseudopotential wave ductors, from dipole-dipole coupling of the transition density functions having a degree of accuracy comparable to ab ini- between unit cells. Under this assumption, the LR contribu- tio wave functions. We find the following. tion to the exchange splitting of s-like in spherical ͑i͒ For direct excitons the e-h exchange interaction has a quantum dots vanishes.11 It is not obvious, however, that sizable LR contribution, comparable in magnitude ͑even af- only dipolar LR interactions exist in quantum dots. ter screening͒ with the SR contribution. The LR and SR ͑iii͒ The screening of the LR exchange interaction. The components have distinct dependencies on the quantum dot effects of dielectric screening on the e-h exchange interac- size. tion are quite controversial.12–14 It is generally understood ͑ii͒ The LR component does not originate from dipole- that in bulk semiconductors the SR exchange interaction dipole interactions between unit cells, as in the case of bulk should be unscreened, while the LR exchange interaction excitons, but from monopole-monopole interactions that are should be screened by the bulk dielectric tensor.14 It is not peculiar to quantum-confined systems.

0163-1829/98/58͑20͒/13367͑4͒/$15.00PRB 58 R13 367 ©1998 The American Physical Society RAPID COMMUNICATIONS

R13 368 A. FRANCESCHETTI, L. W. WANG, H. FU, AND A. ZUNGER PRB 58

͑iii͒ Using a phenomenological, distance-dependent di- electric constant, we show that the LR component of the exchange interaction in quantum dots is significantly reduced by screening effects. ͑iv͒ We are able to calculate the exciton splitting of semi- conductor quantum dots without any a priori empirical in- formation on the bulk exchange splitting or exciton radius. Our results are in good agreement with the experimentally measured redshift in a number of systems. The low-lying excited states of a quantum dot are calcu- lated by a configuration-interaction expansion of the many- particle in terms of single-substitution Slater determinants ͕⌽vc͖, obtained from the ground state ⌽0 by exciting an electron from the occupied single-particle state ␺v(x) of energy ⑀v to the unoccupied single-particle state ␺c(x) of energy ⑀c ͓here xϵ(r,␴), where ␴ϭ , is the spin variable͔. By taking the ground-state energy↑ as↓ energy zero, the matrix elements of the many-particle Hamiltonian in the representation ͕⌽ ͖ are H vc

͗⌽vc͉ ͉⌽vЈcЈ͘ϭ͑⑀cϪ⑀v͒ ␦v,vЈ␦c,cЈϪJvc,vЈcЈϩKvc,vЈcЈ , H ͑2͒ where J and K are the Coulomb and exchange integrals:

␺* x ͒ ␺* x ͒ ␺ x ͒ ␺ x ͒ vЈ͑ 1 c ͑ 2 v͑ 1 cЈ͑ 2 Jvc,vЈcЈϭ FIG. 1. ͑a͒ The unscreened exchange integral K0 (S) ͑nor- ͵͵ ¯⑀͑r ,r ͒ ͉r Ϫr ͉ vc,vc 1 2 1 2 malized by its value at S ϱ) as a function of the cutoff radius S → ϫdx dx , ͑3͒ for GaAs (Rϭ22.5 Å), InP (Rϭ17.4 Å), and CdSe (Rϭ19.2 Å) 1 2 0 spherical quantum dots. The asymptotic values of Kvc,vc(ϱ) are * * 18.1, 21.3, and 18.2 meV, respectively. The arrows denote the ␺v ͑x1͒ ␺c ͑x2͒ ␺cЈ͑x1͒ ␺v͑x2͒ K ϭ Ј Wigner-Seitz radii of the three materials. ͑b͒ shows, for the same vc,vЈcЈ ͵͵ ¯ ¯ ⑀͑r1 ,r2͒ ͉r1Ϫr2͉ quantum dots, the distance-dependent dielectric constant ⑀(S) used in Eqs. ͑3͒ and ͑4͒ to screen the e-h interaction. ϫdx1 dx2 . ͑4͒ states. Thus, we can selectively solve Eq. ͑5͒ for the band- Note that J and K have both diagonal (vϭvЈ,cϭcЈ) and edge states using the folded spectrum method.19 off-diagonal components. The screening of the e-h interac- We consider spherical nanocrystals made of zinc-blende tion, caused by the polarization of the medium, is described GaAs and InP and wurtzite CdSe, with effective radii rang- phenomenologically by the microscopic, position-dependent ing from ϳ6Åtoϳ23 Å. The effective radius is defined in dielectric constant ¯⑀, and will be discussed later. A diago- terms of the number of in the dot (Ndot)asR 1/3 nalization of the Hamiltonian matrix ͑2͒ yields the low- ϭa0(␥Ndot) , where a0 is the bulk lattice constant, ␥ energy excited states of the quantum dot. ϭ3/32␲ for zinc-blende dots, and ␥ϭ3ͱ3c0 /32␲a0 for wurtzite dots. The surface dangling bonds are passivated us- The single-particle wave functions ␺i(r,␴) and energies ing hydrogenlike potentials. Crystal-field effects, which were ⑀i entering Eqs. ͑2͒–͑4͒ are obtained here using an atomistic 2–4 pseudopotential approach that avoids effective-mass approxi- previously treated perturbatively, are naturally described here by specifying the atomic positions. mations. The total, microscopic pseudopotential V ps of the system consisting of the quantum dot and the surrounding The exchange integrals Kvc,vЈcЈ of Eq. ͑4͒ include both barrier is written as superposition of atomic potentials, which the SR part and the LR part of the e-h exchange interaction. 8,17 The effective range of the SR interaction in bulk periodic are derived from measured bulk transition energies, effec- 9 tive masses, and deformation potentials, and from ab initio solids is comparable with the Wigner-Seitz radius Rws . To bulk wave functions calculated in the framework of density- establish the effective range of the exchange interaction in Ϫ1 functional theory. The single-particle states are the solutions quantum dots, we replace the Coulomb potential ͉r1Ϫr2͉ of the Schro¨dinger equation in Eq. ͑4͒ with a cutoff potential ␪(SϪ͉r1Ϫr2͉)•͉r1 Ϫ1 Ϫr2͉ ; the e-h interaction is thus set to zero if ͉r1Ϫr2͉ 2 ˆ ¯ 0 -Ϫ͑1/2͒ ٌ ϩV ps͑r͒ϩVso͔␺i͑r,␴͒ϭ⑀i ␺i͑r,␴͒ , ͑5͒ ϾS. The unscreened (⑀ϭ1) exchange integral Kvc,vc be͓ tween the valence-band maximum and the conduction-band ˆ 17,18 where Vso is the spin-orbit operator. Since the single- minimum in the spin-up configuration is shown in Fig. 1͑a͒ particle levels of a quantum dot are widely spaced in energy as a function of the e-h separation S. We see that the integral 0 compared to the characteristic exchange energies, only a few Kvc,vc(S) keeps increasing far beyond SϳRws , unequivo- single-particle states near the band edges are expected to cally indicating the existence of LR exchange interactions in contribute to the exchange splitting of the lowest exciton quantum dots. RAPID COMMUNICATIONS

PRB 58 SHORT-RANGE VERSUS LONG-RANGE ELECTRON-HOLE... R13 369

TABLE I. Multipole components ͑Ref. 20͒͑Mϭmonopole, Dϭdipole͒ of the LR contribution to the unscreened exchange in- 0 tegral Kvc,vcϭESRϩELR . The numbers in parentheses denote the components of the indirect (X-derived͒ GaAs exciton exchange en- ergy. All energies are in meV.

M-M M-D D-D R (Å) ELR ELR ELR ELR ESR CdSe 19.2 12.7 1.3 0.1 15.4 2.8 InP 17.4 11.7 2.2 0.3 16.9 4.4 GaAs 22.5 14.6 0.5 0.0 15.6 2.6 ͑Ϫ0.1͒͑0.0͒͑0.0͒͑0.1͒͑1.3͒

To understand the physical origin of the LR exchange interaction, we analyze the unscreened exchange integral 0 9 Kvc,vc in terms of the standard multipole-expansion method. By partitioning the system into N eight- unit cells 0 ͕⍀n ,nϭ1,...N͖, the LR contribution to Kvc,vc can be written as

N ␹m*͑r1͒ ␹n͑r2͒ E ϭ dr dr , ͑6͒ LR ͚ ͵͵ Ϫ 1 2 mÞn ͉r1 r2͉ where ␹n(r)ϭ͚␴ ␺v(r,␴) ␺c*(r,␴)ifr෈⍀n and 0 20 otherwise. ELR is further decomposed into multipole con- tributions; in particular, the monopole-monopole (M-M) term is given by FIG. 2. Exciton splitting ⌬ of InP and CdSe spherical quantum dots, as a function of the dot radius R. The experimental data for N CdSe were taken from Ref. 3 ͑crosses͒ and Ref. 5 ͑circles͒. The qm* qn M-M dotted line in the InP plot is a fit to the experimental results of Ref. ELR ϭ ͚ , ͑7͒ mÞn ͉RmϪRn͉ 6. Calculated results are shown both with and without spin-orbit coupling. where qnϭ͐␹n(r) dr is the monopole moment of the transi- tion density ␹n(r) and Rn denotes the position of the unit Ϫ1 Ϫ1 cell ⍀n . The leading terms in the multipole expansion of ⑀ (k) is the Fourier transform of ⑀ (r1Ϫr). The inverse Ϫ1 ELR are shown in Table I for GaAs, InP, and CdSe nano- dielectric constant ⑀ consists of an electronic ͑high- Ϫ1 crystals. In the case of direct (⌫-derived͒ excitons, ELR is frequency͒ contribution ⑀el and an ionic ͑low-frequency͒ Ϫ1 dominated by the monopole-monopole contribution, while contribution ⌬⑀ion , which are approximated here by the the dipole-dipole contribution is small due to the spherical Thomas-Fermi model of Resta23 and by the polaronic model symmetry of the nanocrystals.21 The situation is thus quali- of Haken:24 tatively different from the case of bulk excitons, where the monopole-monopole interaction vanishes (qnϭ0) because k2ϩq2sin kR / dot kR Ϫ1 ͑ ϱ͒ ͑⑀ϱ ϱ͒ of the local orthogonality ͑within each bulk unit cell͒ of ␺v ⑀el ͑k͒ϭ ͑8a͒ k2ϩq2 and ␺c . We also find that the monopole-monopole term scales approximately as 1/R with the dot size, whereas the SR part ͑as well as the dipole-dipole contribution to the LR 1 1 1/2 1/2 3 ⌬⑀Ϫ1͑k͒ϭ Ϫ ϩ . part͒ scales as ϳ1/R . Our results contradict the predictions ion dot dot 2 2 2 2 22 ͩ⑀ ⑀ ͪͩ1ϩ␳ k 1ϩ␳ k ͪ of k–p models that the monopole term vanishes in direct- 0 ϱ h e gap II-VI quantum dots. ͑8b͒ Having established the importance of LR exchange inter- Ϫ1/2 2 1/3 Here qϭ2 ␲ (3␲ n0) ͑where n0ϭelectron density͒ actions in quantum dots, we consider next the effects of di- is the Thomas-Fermi wave vector, and R is the electric screening. The degeneracy between the longitudinal ϱ solution of the equation sinh(qR )/(qR )ϭ⑀dot . Also, ␳ exciton mode and the upper polariton branch at kϭ0 in bulk ϱ ϱ ϱ h,e ϭ(ប/2m ␻ )1/2, where m (m ) is the hole ͑electron͒ ef- semiconductors13 suggests that the LR exchange interaction h,e LO h e fective mass and ␻ is the frequency of the bulk LO pho- should be screened by the dielectric constant as is the polar- LO non mode. dot and dot are the quantum-dot macroscopic iton splitting. We will thus use a phenomenological distance- ⑀ϱ ⑀0 dependent dielectric constant to screen the exchange interac- high-frequency and low-frequency dielectric constants, tion in quantum dots according to the e-h separation. The which are obtained here from a modified Penn model where screened Coulomb potential of Eqs. ͑3͒ and ͑4͒ can be the effective-mass gap is replaced by the pseudopotential- 23 Ϫ1 Ϫ1 calculated gap. The dielectric constant ¯⑀ obtained with the rewritten as ͐⑀ (r1 ,r) ͉rϪr2͉ dr. Assuming that Ϫ1 Ϫ1 present model is plotted in Fig. 1͑b͒ as a function of the e-h ⑀ (r1 ,r)Ϸ⑀ (r1Ϫr), the screened Coulomb potential can be calculated in reciprocal space as 4␲⑀Ϫ1(k)/k2, where separation S. Note that ¯⑀ 1 when S 0. Thus, the SR ex- → → RAPID COMMUNICATIONS

R13 370 A. FRANCESCHETTI, L. W. WANG, H. FU, AND A. ZUNGER PRB 58

change interaction is unscreened, whereas the LR exchange our calculations for CdSe nanocrystals, but not for InP interaction is significantly screened. nanocrystals. Recent ab initio calculations for bulk CdSe The exciton splitting ⌬, defined as the energy difference ͑Ref. 25͒ have shown that the exchange parameter ⌬Ex used between the lowest spin-allowed and the lowest spin- in Ref. 3 is roughly twice as large as the calculated short- forbidden exciton states, is calculated by direct diagonaliza- range parameter. tion of the Hamiltonian matrix ͑2͒. We find that the basis set ͑iii͒ For zinc-blende InP dots, both the calculated exciton generated from the four single-particle states at the top of the splitting8 and the experimentally measured redshift6 scale as valence band and the two single-particle states at the bottom ϳ1/R2 with the dot radius R. This is in contrast with the of the conduction band is sufficient to obtain the exciton 1/R3 scaling law predicted by Eq. ͑1͒. splitting with an accuracy of about 10% relative to a larger In conclusion, we have shown that the e-h exchange in- configuration-interaction expansion. The exciton splitting of teraction in quantum dots is characterized by a previously InP and CdSe nanocrystals is shown in Fig. 2 as a function of unexpected LR contribution originating from the local non- size. For InP dots we compare the singlet-triplet splitting orthogonality between valence and conduction states of the obtained by solving Eq. ͑5͒ with vanishing spin-orbit cou- quantum dot within each unit cell. This LR interaction is pling to the exciton splitting obtained in the presence of spin- specific to quantum-confined systems, and cannot be derived orbit coupling; we see that the inclusion of spin-orbit inter- from the bulk exchange splitting by simple scaling argu- action lowers ⌬ by ϳ30%. We also find the following. ments. Our method permits reliable predictions of the exci- ͑i͒ Our calculated exciton splitting is in reasonably good ton splitting in quantum dots even in the absence of experi- agreement with available experimental results,3,5,6 even mental information on the magnitude of the bulk exchange though no empirical adjustments are used. splitting and exciton radius. ͑ii͒ The screened LR component of the exchange interac- tion is comparable in magnitude with the unscreened SR component, and in a parameter-free calculation cannot be This work was supported by the U.S. DOE, OER-BES, ignored. Interestingly, the phenomenological model of Eq. Division of Materials Science, under Grant No. DE-AC36- ͑1͒, which neglects LR exchange interactions, agrees with 83CH10093.

1 P.D.J. Calcott et al., J. Phys.: Condens. Matter 5, L91 ͑1993͒. 15 A. Franceschetti and A. Zunger, Phys. Rev. Lett. 78, 915 2 M. Nirmal et al., Phys. Rev. Lett. 75, 3728 ͑1995͒. ͑1997͒. 3 Al. L. Efros et al., Phys. Rev. B 54, 4843 ͑1996͒. 16 S.V. Goupalov and E.L. Ivchenko, J. Cryst. Growth 184/185, 393 4 D.J. Norris et al., Phys. Rev. B 53, 16 347 ͑1996͒. ͑1998͒. 5 M. Chamarro et al., Phys. Rev. B 53, 1336 ͑1996͒. 17 L.W. Wang and A. Zunger, Phys. Rev. B 51, 17 398 ͑1995͒. 6 O.I. Micic et al., J. Phys. Chem. 101, 4904 ͑1997͒. 18 The present calculations for GaAs quantum dots do not include 7 U. Banin et al., Superlattices Microstruct. 22, 559 ͑1997͒. spin-orbit coupling. 8 H. Fu and A. Zunger, Phys. Rev. B 56, 1496 ͑1997͒. The dielec- 19 L.W. Wang and A. Zunger, J. Chem. Phys. 100, 2394 ͑1994͒;J. tric constant used in this paper differs from the one used here in Phys. Chem. 98, 2158 ͑1994͒. the way ionic screening is included. 20 The exact value of ELR, and its decomposition shown in Table I, 9 R.S. Knox, Solid State Phys., Suppl. 5,1͑1963͒. depend slightly on the definition of the domains ͕⍀ ͖, but the 10 W.R. Heller and A. Marcus, Phys. Rev. 84, 809 ͑1951͒. n resulting qualitative conclusions are unchanged. 11 T. Takagahara, Phys. Rev. B 47, 4569 ͑1993͒. 21 In the case of indirect (X-derived͒ excitons the LR part is 12 Y. Abe, Y. Osaka, and A. Morita, J. Phys. Soc. Jpn. 17, 1576 negligible, and only the SR part contributes to the exchange ͑1962͒; L.J. Sham and T.M. Rice, Phys. Rev. 144, 708 ͑1966͒. 13 integral. L.C. Andreani, F. Bassani, and A. Quattropani, Lett. Nuovo Ci- 22 mento 10, 1473 ͑1988͒; K. Cho, Solid State Commun. 33, 911 I. Kang and F. Wise, J. Opt. Soc. Am. B 14, 1632 ͑1997͒. 23 ͑1980͒. R. Resta, Phys. Rev. B 16, 2717 ͑1977͒. 24 14 V.A. Kiselev and A.G. Zhilich, Sov. Phys. Solid State 14, 1233 H. Haken, Nuovo Cimento 10, 1230 ͑1956͒. 25 ͑1972;͒; 15, 1351 ͑1972͒. H. Fu, L.W. Wang, and A. Zunger ͑unpublished͒.