Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Quantum theory of electroabsorption in semiconductor nanocrystals

Open Access Open Access

Abstract

We develop a simple quantum-mechanical theory of interband absorption by semiconductor nanocrystals exposed to a dc electric field. The theory is based on the model of noninteracting electrons and holes in an infinitely deep quantum well and describes all the major features of electroabsorption, including the Stark effect, the Franz-Keldysh effect, and the field-induced spectral broadening. It is applicable to nanocrystals of different shapes and dimensions (quantum dots, nanorods, and nanoplatelets), and will prove useful in modeling and design of electrooptical devices based on ensembles of semiconductor nanocrystals.

© 2015 Optical Society of America

1. Introduction

The effect of quantum confinement in semiconductor nanocrystals results in the discreteness of the nanocrystals’ energy spectra and leads to the modification of the nanocrystal’s interaction with both the confined and external elementary excitations [1–3]. This significantly differs the absorption and emission spectra of nanocrystals from the spectra of bulk semiconductors [4–14]. In particular, semiconductor nanocrystals exposed to an electrostatic field exhibit the specific quantum-confined Stark and Franz-Keldysh effects [15–21]. These effects show great promise for useful applications in many fields, including energy harvesting [22,23], near-field sensing [24], and the advancement of nanocrystal-based lasers [25–28].

The effect of electroabsorption has recently been experimentally studied for CdSe quantum dots, nanorods, and nanoplatelets in the form of colloidal solutions [29]. The experiment revealed some intriguing electroabsorption properties of semiconductor nanoplatelets, which are flat nanocrystals with almost zero thickness dispersion (the thickness of a nanoplatelet can be changed discretely, with a step of one monolayer). One of such properties is the pronounced electro-optical response of nanoplatelets compared to the response of quantum dots and nanorods. It was specifically found that the field-induced change in the absorption efficiency of the nanoplatelets is ten times that of the quantum dots. This feature makes nanoplatelets very promising for applications and is due to the relatively weak spatial quantum confinement in the planes of the nanoplatelets. Here we develop a fully quantum-mechanical theory of electroabsorption by semiconductor nanocrystals, which is needed for quantification of these and other experimental findings. We significantly simplify the mathematics by ignoring the Coulomb interaction between the confined charge carriers and assuming the confining potential to be infinite. The developed theory not only describes all the major features of electroabsorption, such as the Stark effect, the Franz-Keldysh effect, and the field-induced broadening of electroabsorption spectra, but also provides guidance toward novel applications.

2. Theoretical formulation

The quantum mechanical treatment of the electroabsorption phenomenon in semiconductor nanocrystals requires knowing the wave functions and energy spectrum of electrons and holes confined by the nanocrystals. We begin this treatment by formulating an analytically solvable model of the electronic subsystem of a semiconductor nanocrystal.

2.1. Electronic subsystem of a semiconductor nanocrystal

In order to analytically calculate the energy spectrum and wave functions of the confined charge carriers, we employ the two-band envelope k · p model [30] and assume that the nanocrystal is in the form of a rectangular parallelepiped lx × ly × lz. A state of the charge carriers is described by a set of three quantum numbers |n〉 = |nx, ny, nz〉, each representing the spatially confined motion along one of the Cartesian axes. The energies and wave functions obey the Shrödinger equation

En|n=(22m*Δ+V(r)+W(r))|n,
where m* is the effective mass of an electron or hole, W(r) is the confining potential, which is assumed to be zero inside the nanocrystal and infinity outside of the nanocrystal, and V(r) is the electric field potential.

The form of potential V(r) critically depends on the nanocrystal dimensions and the orientation of the nanocrystal with respect to the external electric field F, which is assumed to be homogenous far away from the nanocrystal. Due to the nonsphericity of the nanocrystal, the electric field inside it, Fint, is inhomogeneous. We first ignore this inhomogeneity by assuming that the external field is simply changed inside the nanocrystal by a constant factor, which is the ratio of permittivities of the nanocrystal and the environment, ε = εNC / εe, and take the inhomogeneity into account later using the perturbation theory. With this assumption, the electrostatic potential inside the nanocrystal is given by V(r) = qFr= ε, where q = −e for electrons and q = e for holes.

The linearity of potential V (r) allows one to separate variables in the Shrödinger equation. By taking |n〉= |nx〉|ny〉|nz〉 and considering the energy spectrum to be the sum of energies due to the confined motions in three spatial dimensions, En=Enx+Eny+Enz, we obtain the following one-dimensional Shrödinger equation inside the nanocrystal:

Env|nv=(22m*d2dv2+qεFvv)|nv(v=x,y,z),
where Fv is the vth component of the electric field. By introducing a new dimensionless variable
σ(v)=(2m*qFv2ε)1/3(vEnvεqFv),
we reduce Eq. (2) to the Airy equation d2|nv〉/dσ2 = σ|nv〉, whose general solution is a linear combination of the Airy functions of the first and second kinds, |nv=AnvAi(σ)+BnvBi(σ).

Since the nanocrystal surface is assumed to be impenetrable for electrons and holes, the wave functions at this surface must vanish. The application of this boundary condition at facet v = lv/2 relates the two integration constants in the general solution as Bnv=AnvAi[σ(lv/2)]/Bi[σ(lv/2)]. Using the normalization condition, 〈nv|nv〉 = 1, then gives [31]

Anv=π(2m*qFv2ε)1/3Bi[σ(lv/2)]Bi[σ(lv/2)]{Bi2[σ(lv/2)]Bi2[σ(lv/2)]}1/2.

Finally, the energy spectrum Env can be calculated from the dispersion equation, which follows from the boundary condition at facet v = −lv/2,

Ai[σ(lv/2)]Bi[σ(lv/2)]=Ai[σ(lv/2)]Bi[σ(lv/2)].

Some algebra shows that in weak fields Envlv4.

The exact electric field inside the nanocrystal can be decomposed into homogeneous and inhomogeneous contributions as Fint(r) = F/ε + δF(r). If the permittivities of the nanocrystal and its environment do not differ too much, then |δF|≪|F| and electrostatic potential δV(r) of the inhomogeneous contribution can be treated as a small perturbation of potential V(r) = qFr. By solving the equation ∇δV = δF and using the stationary perturbation theory, we can write the wave functions and energy spectrum of the nanocrystal in the form

|n=|n+mnn|δV|mEnEm|m,n=En+n|δV|n.

Note that even though δV is determined up to a constant, this constant does not affect the perturbed wave functions and interband transition energies, because it results in the shift of the nanocrystals energy spectrum as a whole.

2.2. Electroabsorption by a semiconductor nanocrystal

Consider electroabsorption of a semiconductor nanocrystal upon an interband transition between a hole state m and an electron state n, which is induced by light of intensity I and frequency ω. The transition frequency is given by

ωnm=(n+m+Eg)/,
where Eg is the semiconductor band gap. The strength of electroabsorption is characterized by the electron–hole pair generation rate [32]
Wnm=(4πeP)2I3c3ω2nNCΓnm|n|m|2,
where P is the Kane parameter, nNC is the refractive index of the nanocrystal material at frequency ω, and the spectral line shape of the transition characterized by a full dephasing rate γnm can be approximated by the Lorenzian
Γnm=1πγnm(ωωnm)2+γnm2.

2.3. Electroabsorption by an ensemble of semiconductor nanocrystals

The absorption coefficient of a monodisperse ensemble of N equally oriented nanocrystals is given by Knm = N(ħω/I)Wnm. The absorption coefficient of a monodisperse ensemble with randomly oriented nanocrystals can be found by the averaging

KnmΩ=14πKnm(ϑ,φ)sinϑdϑdφ,
where ϑ and φ are the polar angle and azimuth specifying the nanocrystal orientation with respect to the external field.

If the nanocrystal dimensions in the ensemble are distributed about their mean values 〈lv〉 (v = x,y,z), and the distribution is characterized by the probability density G(〈lx〉, 〈ly〉, 〈lz〉, lx, ly, lz), then the inhomogeneous size broadening of the electroabsorption spectrum of the ensemble is described by the absorption coefficient

Knmg=Knm(lx,ly,lz)G(lx,ly,lz,lx,ly,lz)dlxdlydlz.

The three dimensions of nanocuboids are usually independent of one another and characterized by the same distributions, which results in the following factorization of the probability density: G(〈lx〉, 〈ly〉, 〈lz〉, lx, ly, lz) = g (〈lx〉, lx) g (〈ly〉, ly) g (〈lz〉, lz). The size distribution of the nanocrystal edges is often described by the lognormal density function

g(lv,lv)=12πlvδexp(ln2(lv/lv)2δ2),
where δ is the standard deviation of ln Lv and 〈ly Med[Lv] is the geometric mean of Lv.

3. Results and discussion

The developed theory describes all the significant features of electroabsorption of semiconductor nanocrystals, including the Stark effect, the Franz-Keldysh effect, and the field-induced broadening of electroabsorption spectra of nanocrystal ensembles.

The Stark effect is the field-induced shifting and splitting of absorption peaks in the nanocrystal spectrum. When a nanocrystal is exposed to a dc electric field, the size-quantized energies of its electrons and holes shift, as illustrated in Fig. 1(a), changing the frequencies of interband transitions. The change of the transition energies is described by Eqs. (5), (6), and (7). If the field is sufficiently strong, then the energies of some transitions are smaller than the band gap of bulk semiconductor, which makes the nanocrystal to absorb light at otherwise forbidden frequencies. The developed theory also predicts that the strength of the Stark effect critically depends on the dimensions of the nanocrystal and its orientation with respect to the electric field. In particular, the red shifts of the absorption peaks are maximal when the largest nanocrystal dimension is aligned with the field. The theory also describes the lifting of the degeneracy of the nanocrystal states and splitting of the respective peaks in the absorption spectrum due to the symmetry breaking of the confining potential by the electric field.

The Franz-Keldysh effect is the suppression of interband transitions that were dipole allowed without the field and the emergence of the forbidden transitions in the absorption spectrum of the nanocrystal. Both features are described by the overlap integral 〈〈n|m〉〉 in Eq. (8). In the absence of the electric field, the confined electrons and holes are described by the same set of wave functions, which means that only electric dipole transitions between the states with the same quantum numbers are allowed, i.e. 〈n|m〉 = δnm. The electric field shifts electrons and holes to the opposite sides of the nanocrystal, making the overlap integral 〈〈n|m〉〉 decrease with the field strength and reducing the rates of the dipole-allowed transitions. At the same time, the overlap integral of the originally orthogonal wave functions of electrons and holes becomes nonzero, which leads to the appearance of additional absorption peaks in the nanocrystal spectrum. The variation of the overlap integral is illustrated by Fig. 1(b). Like the Stark effect, the Franz-Keldysh effect is the most pronounced when the electric field is directed along the largest dimension of the nanocrystal, because the spatial separation of electrons and holes in a given field grows with the reduction of the quantum confinement.

The field-induced broadening of electroabsorption spectrum is a result of random orientation of the nanocrystals in space. This type of broadening comes from the averaging of the absorption spectra of individual nanocrystals, which differ due to the strengths of the Stark and Franz-Keldysh effects being dependent on the nanocrystal orientation relative to the external field. Indeed, nanocrystals that have their largest dimensions aligned to the electric field exhibit the strongest Stark and Franz-Keldysh effects, featuring spectra that are the most modified by the field. On the other hand, the spectra of nanocrystals with the smallest dimensions parallel to the field are almost unchanged by the field. As a result of spectra averaging, described by Eq. (10), each line in the absorption spectrum of the nanocrystal acquires a more or less pronounced low-frequency wing, like those shown in Fig. 1(c). The nonmonotonic dependence of the absorption peak on the field strength, evidenced by the figure, is another clear manifestation of the Franz-Keldysh effect. Also noteworthy is that the field-induced broadening is the strongest in ensembles of nanorods and nanoplatelets, whose three dimensions are significantly different, and is absent in the electroabsorption spectra of spherical nanocrystals.

 figure: Fig. 1

Fig. 1 (a) Stark shift of electronic energies Enx, (b) overlap integral 〈nx|mx〉, and (c) in-homogeneously broadened absorption peak (111) → (112) of randomly oriented 2 × 20 × 20 nm3 CdSe nanoplatelets. In (a) and (b) lx = 20 nm. In all panels δV = 0, the field strength is given inside the nanoplatelets, and the effective masses of electrons and holes are 0.11m0 and 0.44m0, where m0 is the free-electron mass.

Download Full Size | PDF

Acknowledgments

This work was funded by Grant 14.B25.31.0002 and Government Assignment No. 3.17.2014/K of the Ministry of Education and Science of the Russian Federation. M.Y.L. also gratefully acknowledges the financial support from the Ministry of Education and Science of the Russian Federation through its scholarship of the President of the Russian Federation for young scientists and graduate students (2013–2015).

References and links

1. A. V. Fedorov, A. V. Baranov, I. D. Rukhlenko, T. S. Perova, and K. Berwick, “Quantum dot energy relaxation mediated by plasmon emission in doped covalent semiconductor heterostructures,” Phys. Rev. B 76, 045332 (2007). [CrossRef]  

2. I. D. Rukhlenko and A. V. Fedorov, “Propagation of electric fields induced by optical phonons in semiconductor heterostructures,” Opt. Spectrosc. 100, 238–244 (2006). [CrossRef]  

3. I. D. Rukhlenko and A. V. Fedorov, “Penetration of electric fields induced by surface phonon modes into the layers of a semiconductor heterostructure,” Opt. Spectrosc. 101, 253–264 (2006). [CrossRef]  

4. A. S. Baimuratov, I. D. Rukhlenko, V. K. Turkov, I. O. Ponomareva, M. Y. Leonov, T. S. Perova, K. Berwick, A. V. Baranov, and A. V. Fedorov, “Level anticrossing of impurity states in semiconductor nanocrystals,” Sci. Rep. 4, 6917 (2014). [CrossRef]   [PubMed]  

5. A. S. Baimuratov, I. D. Rukhlenko, V. K. Turkov, A. V. Baranov, and A. V. Fedorov, “Quantum-dot supercrystals for future nanophotonics,” Sci. Rep. 3, 1727 (2013). [CrossRef]  

6. A. S. Baimuratov, I. D. Rukhlenko, and A. V. Fedorov, “Engineering band structure in nanoscale quantum-dot supercrystals,” Opt. Lett. 38, 2259–2261 (2013). [CrossRef]   [PubMed]  

7. I. Sagnes, A. Halimaoui, G. Vincent, and P. Badoz, “Optical absorption evidence of a quantum size effect in porous silicon,” Appl. Phys. Lett. 62, 1155–1157 (1993). [CrossRef]  

8. Y. Wang and N. Herron, “Nanometer-sized semiconductor clusters: Materials synthesis, quantum size effects and photophysical properties,” J. Phys. Chem. 95, 525–532 (1991). [CrossRef]  

9. A. S. Baimuratov, I. D. Rukhlenko, V. K. Turkov, M. Y. Leonov, A. V. Baranov, Y. K. Gun’ko, and A. V. Fedorov, “Harnessing the shape-induced optical anisotropy of a semiconductor nanocrystal: A new type of intraband absorption spectroscopy,” J. Phys. Chem. C 118, 2867–2876 (2014). [CrossRef]  

10. I. D. Rukhlenko, M. Y. Leonov, V. K. Turkov, A. P. Litvin, A. S. Baymuratov, A. V. Baranov, and A. V. Fedorov, “Kinetics of pulse-induced photoluminescence from a semiconductor quantum dot,” Opt. Express 20, 27612–27635 (2012). [CrossRef]   [PubMed]  

11. S. Furukawa and T. Miyasato, “Quantum size effects on the optical band gap of microcrystalline Si:H,” Phys. Rev. B 38, 5726–5729 (1988). [CrossRef]  

12. E. V. Ushakova, A. P. Litvin, P. S. Parfenov, A. V. Fedorov, M. V. Artemyev, A. V. Prudnikau, I. D. Rukhlenko, and A. V. Baranov, “Anomalous size-dependent decay of low-energy luminescence from PbS quantum dots in colloidal solution,” ACS Nano 6, 8913–8921 (2012). [CrossRef]   [PubMed]  

13. Y. Kanemitsu, H. Uto, Y. Masumoto, T. Matsumoto, T. Futagi, and H. Mimura, “Microstructure and optical properties of free-standing porous silicon films: Size dependence of absorption spectra in Si nanometer-sized crystallites,” Phys. Rev. B 48, 2827–2831 (1993). [CrossRef]  

14. I. D. Rukhlenko, A. D. Fedorov, A. S. Baymuratov, and M. Premaratne, “Theory of quasi-elastic secondary emission from a quantum dot in the regime of vibrational resonance,” Opt. Express 19, 15459–15482 (2011). [CrossRef]   [PubMed]  

15. X. Liu, T. Iimori, R. Ohshima, T. Nakabayashi, and N. Ohta, “Electroabsorption spectra of PbSe nanocrystal quantum dots,” Appl. Phys. Lett. 98, 161911 (2011). [CrossRef]  

16. H. Spector and J. Lee, “Stark effect in the optical absorption in cubical quantum boxes,” Physica B 393, 94–99 (2007). [CrossRef]  

17. P. Jin, C. Li, Z. Zhang, F. Liu, Y. Chen, X. Ye, B. Xu, and Z. Wang, “Quantum-confined Stark effect and built-in dipole moment in self-assembled InAs/GaAs quantum dots,” Appl. Phys. Lett. 85, 2791–2793 (2004). [CrossRef]  

18. S. Ramanathan, G. Petersen, K. Wijesundara, R. Thota, E. Stinaff, M. Kerfoot, M. Scheibner, A. Bracker, and D. Gammon, “Quantum-confined Stark effects in coupled InAs/GaAs quantum dots,” Appl. Phys. Lett. 102, 213101 (2004). [CrossRef]  

19. I. Itskevich, S. Rybchenko, I. Tartakovskii, S. Stoddart, A. Levin, P. Main, L. Eaves, M. Henini, and S. Parnell, “Stark shift in electroluminescence of individual InAs quantum dots,” Appl. Phys. Lett. 76, 3932–3934 (2000). [CrossRef]  

20. D. Miller, D. Chemla, and S. Schmitt-Rink, “Relation between electroabsorption in bulk semiconductors and in quantum wells: The quantum-confined Franz-Keldysh effect,” Phys. Rev. B 33, 6976–6982 (1986). [CrossRef]  

21. D. Miller, D. Chemla, T. Damen, A. Gossard, W. Wiegmann, T. Wood, and C. Burrus, “Band-edge electroabsorption in quantum well structures: The quantum confined Stark effect,” Phys. Rev. Lett. 53, 2173–2176 (1984). [CrossRef]  

22. A. S. Baimuratov, I. D. Rukhlenko, V. K. Turkov, A. V. Baranov, and A. V. Fedorov, “Quantum-dot supercrystals for future nanophotonics,” Sci. Rep. 3, 1727 (2013). [CrossRef]  

23. H. Hillhouse and M. Beard, “Solar cells from colloidal nanocrystals: Fundamentals, materials, devices, and economics,” Curr. Opin. Colloid Interface Sci. 14, 245–259 (2009). [CrossRef]  

24. K. Park, Z. Deutsch, J. J. Li, D. Oron, and S. Weiss, “Single molecule quantum-confined Stark effect measurements of semiconductor nanoparticles at room temperature,” ACS Nano 6, 10013–10023 (2012). [CrossRef]   [PubMed]  

25. Y. Wang, K. S. Leck, V. D. Ta, R. Chen, V. Nalla, Y. Gao, T. He, H. V. Demir, and H. D. Sun, “Blue liquid lasers from solution of CdZnS/ZnS ternary alloy quantum dots with quasi-continuous pumping,” Adv. Mater. 27, 169–175 (2015). [CrossRef]  

26. Y. Wang, S. Yang, H. Yang, and H. D. Sun, “Quaternary alloy quantum dots: Toward low-threshold stimulated emission and all-solution-processed lasers in the green region,” Adv. Mater. 3, 652–657 (2014).

27. D. Bimberg, “Quantum dots for lasers, amplifiers and computing,” J. Phys. D: Appl. Phys. 38, 2055 (2005). [CrossRef]  

28. H. Saito, K. Nishi, and S. Sugou, “Ground-state lasing at room temperature in long-wavelength InAs quantum-dot lasers on InP(311)B substrates,” Appl. Phys. Lett. 78, 267 (2001). [CrossRef]  

29. A. W. Achtstein, A. V. Prudnikau, M. V. Ermolenko, L. I. Gurinovich, S. V. Gaponenko, U. Woggon, A. V. Baranov, M. Y. Leonov, I. D. Rukhlenko, A. V. Fedorov, and M. V. Artemyev, “Electroabsorption by 0D, 1D, and 2D nanocrystals: A comparative study of CdSe colloidal quantum dots, nanorods, and nanoplatelets,” ACS Nano 8, 7678–7686 (2014). [CrossRef]   [PubMed]  

30. M. Ehrhardt and T. Koprucki, Multi-Band Effective Mass Approximations: Advanced Mathematical Models and Numerical Techniques, 1st ed., Lecture Notes in Computational Science and Engineering94 (Springer International Publishing, 2014).

31. O. Olendski, “Comparative analysis of electric field influence on the quantum wells with different boundary conditions,” Annalen Phys. 527, 278–295 (2015). [CrossRef]  

32. S. Y. Kruchinin and A. V. Fedorov, “Spectroscopy of persistent hole burning in the quantum dot-matrix system: Quantum-confined Stark effect and electroabsorption,” Phys. Solid State 49, 968–975 (2007). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (1)

Fig. 1
Fig. 1 (a) Stark shift of electronic energies E n x, (b) overlap integral 〈nx|mx〉, and (c) in-homogeneously broadened absorption peak (111) → (112) of randomly oriented 2 × 20 × 20 nm3 CdSe nanoplatelets. In (a) and (b) lx = 20 nm. In all panels δV = 0, the field strength is given inside the nanoplatelets, and the effective masses of electrons and holes are 0.11m0 and 0.44m0, where m0 is the free-electron mass.

Equations (12)

Equations on this page are rendered with MathJax. Learn more.

E n | n = ( 2 2 m * Δ + V ( r ) + W ( r ) ) | n ,
E n v | n v = ( 2 2 m * d 2 d v 2 + q ε F v v ) | n v ( v = x , y , z ) ,
σ ( v ) = ( 2 m * q F v 2 ε ) 1 / 3 ( v E n v ε q F v ) ,
A n v = π ( 2 m * q F v 2 ε ) 1 / 3 Bi [ σ ( l v / 2 ) ] Bi [ σ ( l v / 2 ) ] { Bi 2 [ σ ( l v / 2 ) ] Bi 2 [ σ ( l v / 2 ) ] } 1 / 2 .
Ai [ σ ( l v / 2 ) ] Bi [ σ ( l v / 2 ) ] = Ai [ σ ( l v / 2 ) ] Bi [ σ ( l v / 2 ) ] .
| n = | n + m n n | δ V | m E n E m | m , n = E n + n | δ V | n .
ω nm = ( n + m + E g ) / ,
W nm = ( 4 π e P ) 2 I 3 c 3 ω 2 n NC Γ nm | n | m | 2 ,
Γ nm = 1 π γ nm ( ω ω nm ) 2 + γ nm 2 .
K nm Ω = 1 4 π K nm ( ϑ , φ ) sin ϑ d ϑ d φ ,
K nm g = K nm ( l x , l y , l z ) G ( l x , l y , l z , l x , l y , l z ) d l x d l y d l z .
g ( l v , l v ) = 1 2 π l v δ exp ( ln 2 ( l v / l v ) 2 δ 2 ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.