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

Efficient method for the calculation of dissipative quantum transport in quantum cascade lasers

Open Access Open Access

Abstract

We present a novel and very efficient method for calculating quantum transport in quantum cascade lasers (QCLs). It follows the nonequilibrium Green’s function (NEGF) framework but sidesteps the calculation of lesser self-energies by replacing them by a quasi-equilibrium expression. This method generalizes the phenomenological Büttiker probe model by taking into account individual scattering mechanisms. It is orders of magnitude more efficient than a fully self-consistent NEGF calculation for realistic devices. We apply this method to a new THz QCL design which works up to 250 K – according to our calculations.

© 2015 Optical Society of America

1. Introduction

A detailed understanding of carrier dynamics is crucial for the design and improvement of modern semiconductor nanodevices such as quantum cascade lasers (QCLs). However, neither a classical or semiclassical nor a strictly coherent, ballistic quantum mechanical theory can capture the tight interplay between incoherent relaxation processes and quantum interference effects. A general and rigorous framework to capture all of these effects is the nonequilibrium Green’s function (NEGF) theory. Today, it is well established that this theory is among the most general schemes for the prediction of quantum transport properties [13]. Unfortunately, the basic NEGF equations are complex, mathematically demanding, and a quantitative implementation is still a highly challenging task, even with the recent advances of modern computer hardware. Furthermore, it is very difficult to develop approximations within the NEGF formalism which maintain charge and current conservation and obey the Pauli principle [4].

In this work, we present a novel method to calculate stationary quantum transport properties that we termed multi-scattering Büttiker probe (MSB) model [5]. It generalizes the so-called Büttiker probe model [1, 6, 7] and accounts for individual scattering mechanisms. If the semiconductor device equations are discretized on a spatial grid, then the standard Büttiker probe scattering model associates a single momentum and energy sink with each device grid point. Unfortunately, it does not capture effects of different scattering mechanisms. In contrast, our new MSB model accounts for individual scattering mechanisms by considering multiple scattering potentials at each grid position [Fig. 1]. In this article we first outline the concepts of semiconductor heterostructures most readers are familiar with. Then we briefly summarize the NEGF scheme and the original Büttiker probe model before we explain the details of our novel method. Finally, we apply our MSB model to a new quantum cascade laser structure and present results which is the main motivation for this work.

2. Method

2.1. The one-particle Hamiltonian

A quantum cascade laser is a heterostructure consisting of different layers that are laterally homogeneous. Therefore we can separate the growth coordinate (z axis) and the homogeneous layers (xy plane). We describe the electronic structure by a single-band effective mass Hamiltonian

H(z,k)=HL(z)+HT(z,k),
with the longitudinal and transverse parts given as
HL(z)=22ddz(1m(z)ddz)+Ec,0(z)eϕ(z),
HT(z,k)=2k22m*(z),
where ħ is the reduced Planck constant, Ec,0(z) is the conduction band edge energy profile of the heterostructure and k = |k| is the absolute value of the two-dimensional wave vector in the xy plane. Throughout this work we assume the electron effective mass to be isotropic, m*=m*=m*. Additionally, we assume that the effective mass is independent of energy thereby neglecting any effects due to nonparabolicity. This assumption is reasonable for THz quantum cascade lasers where all relevant energy levels are close to the conduction band edge.

2.2. The Poisson equation

The electrostatic potential ϕ is obtained by solving the nonlinear Poisson equation

·(ε0εs(z)ϕ(z))=e(ND+(z)n(z,ϕ)).

The right hand side of this equation is the charge density which consists of the electron concentration n and the ionized donor concentration ND+. Here, e is the positive elementary charge, ε0 is the electric vacuum permittivity and εs is the static relative permittivity. The Poisson equation includes electron–electron scattering within the Hartree approximation. We solve this equation by a Newton–Raphson method. In order to solve it very efficiently, we apply a predictor–corrector approach to ‘predict’ the quantum density n(z) upon small changes dϕ(z) of the electrostatic potential [8]. We thus avoid the computationally expensive calculation of the Green’s functions [Eq. (5)(8)] within each Poisson iteration cycle.

2.3. The NEGF scheme

The quantum transport problem of a QCL can be regarded as a carrier scattering problem between reservoirs. Hence, the system of interest must be treated as an open quantum system. A common approach is to divide the quantum system into a device region and two lead regions (source and drain). The leads are assumed to be in equilibrium and consequently, all relevant scattering processes occur inside the device. The effect of the coupling of the device to the source and drain leads is described via a so-called contact self-energy ΣCR, which incorporates transitions between the device and the leads. A full implementation of the NEGF method requires the self-consistent solution of the retarded and lesser Green’s functions, GR and G<. The first one characterizes the width and energy of the scattering states and is related to the local density of states. The second quantity characterizes the state occupancy and determines the charge and current densities, and the optical gain. Under stationary conditions, the retarded and lesser Green’s functions are sufficient to completely describe a nonequilibrium system.

The self-consistent cycle involves the solution of four coupled integro-differential equations. In operator form, they read for a given energy E,

GR=(E1HΣR)1,
ΣR=ΣCR+GRDR+G<D<,
Σ<=ΣC<+G<D<,
G<=GRΣ<GR.

The term ‘self-consistent’ means that the Green’s functions determine the self-energies and at the same time the self-energies determine the Green’s functions. The self-energies [Eq. (6), Eq. (7)] are calculated in the self-consistent Born approximation. H is the electronic device Hamiltonian [Eq. (1)] that includes the electrostatic potential ϕ. Consequently, the Poisson equation [Eq. (4)] has to be solved self-consistently with the Green’s functions [Eq. (5)(8)], which introduces another self-consistent cycle. ΣR and Σ< are the retarded and lesser total scattering self-energies. These self-energies are built-up from the contact self-energy and the sum of all environmental Green’s functions DR and D< that contain the scattering vertices.

For a multi-quantum well structure such as a QCL grown along the z axis, the lateral momentum conservation simplifies the arguments to the following functional form, G = G (z, z′, k, E). All quantities are therefore functions of space, momentum and energy. The key simplifications in our new MSB approach consist in (i) sidestepping the self-consistent solution of Σ< and G< and (ii) performing the k integration analytically so that the Green’s functions become functions of space and energy only, G = G(z, z′, Ez). The total energy E is the sum of the energy Ez along the direction of propagation and energy E within a layer

E=Ez+E=Ez+2k22m*,
i.e. the Green’s functions for layered heterostructures within the MSB method do not depend on E and k separately but only on their difference
G(k,E)G(EE)G(Ez).

As we will later show, we reduce the mathematical complexity within the MSB model, so that the self-consistent cycle only involves two coupled equations, namely GR [Eq. (5)] and ΣR=ΣCR+GRDR [Eq. (6)]. The iterative solution of only two coupled equations is far more robust and faster than the iteration over four coupled equations within the full NEGF formalism. Moreover, current conservation and the Pauli principle are inherently guaranteed during every step of the iteration within the MSB model, since we explicitly calculate the virtual chemical potentials μ of the Büttiker probes as we will show further below. Therefore, even in a nonself-consistent Born approximation, current conservation is not violated.

The retarded self-energy

ΣR=ΣCR+ΣBR,
includes the contact self-energy
ΣCR=ΣSR+ΣDR,
which describes the coupling of the source (S) and drain (D) contacts to the device, and the Büttiker probe self-energy ΣBR [Eq. (25)] which describes scattering due to longitudinal acoustic and optical phonons. The contact self-energies due to the left (source) and right (drain) lead are nonzero within the contact–device interfaces only. They are written in the usual form describing the coupling to semi-infinite leads [1]. The only nonzero matrix elements are
ΣSR(1,1,Ez)=texp(ikSa),
ΣDR(N,N,Ez)=texp(ikDa),
where 1 and N refer to the indices of the leftmost and rightmost grid point. All matrices are represented using a finite differences discretization on N grid points in real space using a uniform grid spacing a. The kinetic coupling matrix element t is given by
t=22m*a2.

The wave vector k(Ez) is calculated for each energy Ez of the discretized energy grid from the lead dispersion which is given for a discrete lattice by

Ez(k)=Ec+2t(1cos(ka)),
where Ec = Ec,0 − eϕ is the conduction band edge energy at the source or drain, respectively.

 figure: Fig. 1

Fig. 1 Left: The standard Büttiker probe model associates a single momentum and energy sink with each device grid point i in position space using a phenomenological scattering parameter η, which is fixed during the calculation. Right: The proposed multi-scattering Büttiker probe model accounts for individual scattering mechanisms such as longitudinal acoustic (LA) or longitudinal optical (LO) phonon scattering by including multiple scattering potentials for each node that are calculated self-consistently.

Download Full Size | PDF

2.4. The multi-scattering Büttiker probe (MSB) model

One of the difficulties of the NEGF formalism is the fact that current conservation and the Pauli principle are not automatically guaranteed. Particularly, the calculation of the self-energies in any nonself-consistent approximation violates current conservation. To remedy such a result, the iteration over the four coupled Green’s functions and self-energies has to be carried out to infinite order [4]. In contrast, the main advantage of the Büttiker probe model is that it bypasses any self-consistent calculation between the Green’s functions and their corresponding self-energies. Moreover, it inherently guarantees current conservation, and the Pauli principle is always obeyed since a quasi-equilibrium expression for G< is assumed and thus, the fluctuation–dissipation theorem holds. However, the Büttiker probe model introduces a phenomenological scattering parameter η,

B(z,Ez)=iη(z),
which models scattering, and particularly the scattering strength, where η is defined as e.g. η = ħ/τ with τ being a material dependent scattering time that varies with spatial position z [Fig. 1]. In fact, within this model, scattering processes can be viewed as additional virtual floating contacts (so-called probes) within the device that remove, thermalize and/or reinject carriers. Typically, this scattering potential does not distinguish between individual scattering mechanisms. Furthermore, this parameter is fixed during the calculation. Thus neither the importance, nor the quantitative effects of individual scattering mechanisms can be estimated with such a model. Note that this is only one out of many possibilities for the scattering potential B(z,Ez) within the Büttiker probe model. In the following we show how to calculate the Büttiker probe self-energy within the MSB model where the treatment of the individual scattering mechanisms is retained. The total scattering strength of our multi-scattering Büttiker probes is the sum of the individual scattering strengths
B(z,Ez)=BLA(z,Ez)+BLO(z,Ez),
where we take into account scattering due to longitudinal acoustic (LA) and longitudinal optical phonons (LO). Electron–electron scattering is included within the Hartree approximation via the Poisson equation. The crucial concept of our approach is to replace the phenomenological scattering parameter η with a parameter derived from physical arguments and material constants that is calculated self-consistently during the iteration. This avoids the usage of a ‘fitting’ parameter.

In general, the derivation of the scattering kernels BLA and BLO for the MSB model follows the derivation of the scattering self-energies within the full NEGF scheme. In this scheme BLA depends on the retarded Green’s function GR, and BLO on both the retarded and the lesser Green’s functions, GR and G<. The MSB model, however, makes the following two approximations. First, we neglect the contribution of G< on BLO which essentially decouples the retarded and the lesser Green’s functions making the overall calculation faster and more robust. Secondly, we take only the diagonal part of GR into account by assuming that the scattering kernels only depend on the local density of states. The local density of states ρ(z,Ez) is the diagonal part (z = z′) of the spectral function A divided by 2π

ρ(z,Ez)=12πA(z,z,Ez).

The spectral function

A=i(GRGR)=2(GR),
depends on the imaginary part of the retarded Green’s function only. It can be viewed as a generalized density of states since it gives information about the location, energy and type of states regardless of their occupation. In our implementation the scattering strength of a single Büttiker probe that models carrier–acoustic phonon scattering reads
BLA(z,Ez)=iVD2kBT8πρMvs2ELAEzELAEz+ELAdEzρ(z,Ez),
where kB is the Boltzmann constant, VD is the scalar deformation potential, ρM is the mass density of the material, vs is the velocity of sound in the material and ELA is the maximum energy of longitudinal acoustic phonons. The probe scattering potential for the carrier–optical phonon scattering reads
BLO(z,Ez)=ie2ξELO32πε0(ε1εs1) [(1+NLO)ρ(z,EzELO)+NLOρ(z,Ez+ELO)],
where ξ is the Debye screening length, ε is the relative electric permittivity at the high frequency limit, and ELO is the energy of the optical phonons. Here we assume that the dispersion relation of the phonons is constant (ELO = const), and also that the phonon distribution at a given temperature T is constant,
NLO=(exp(ELOkBT)1)1.

The scattering potential B(z,Ez) is a functional of the local density of states ρ(z,Ez), i.e. it has to be calculated self-consistently. The local density of states can be directly calculated from the retarded Green’s function GR [Eq. (19)(20)]. Consequently, Eq. (21) and Eq. (22) constitute a self-consistent relation between ΣR and GR that, in principle, has to be solved in an iterative scheme. Fortunately, the density of states ρ(z,Ez) and thus also the self-energy ΣR depend on the diagonal part of GR only. Therefore, only the diagonal part of GR is required during the iteration which can be calculated very efficiently, e.g. with the FIND (Fast Inverse Using Nested Dissection) algorithm [9]. In order to decouple the lesser and the retarded Green’s functions, we neglected in Eq. (22) the contribution of the charge carrier density to enhance the convergence while the Pauli principle is still inherently guaranteed since the probes are assumed to be in a local equilibrium. This approximation essentially corresponds to dropping the term G<D< in Eq. (6). In principle, the scattering self-energy for LO phonon coupling depends on the momentum k. However, since we assume a Fröhlich coupling, which favors small momentum transfers, we can neglect the k dependence of the self-energies. Thus, they depend solely on the energy Ez along the growth axis [Eq. (9)]. We can therefore analytically perform integrations with respect to k [5] that appear in the calculation of observables like density, current and optical gain.

For the numbering of the spatial grid points we assume that i = 1 and i = N refer to the source and drain grid points, respectively. At all grid points i = 1,…,N we locate exactly one Büttiker probe p which has the scattering strength B(z,Ez) at this position zp. We can map the Büttiker scattering potential onto the retarded Büttiker probe self-energy of each probe as follows

ΣpR(z,z,Ez)=B(z,Ez)δ(zzp).

The δ function indicates that the self-energy ΣpR for probe p is nonzero only at the position coordinate zp where the probe is located. All other matrix elements apart from this one are zero. The retarded Büttiker probe self-energy ΣBR is the sum of the self-energies of each probe p,

ΣBR(z,z,Ez)=p=1NΣpR(z,z,Ez).

It has nonzero matrix elements only on the diagonal (z = z′). This self-energy corresponds to the term GRDR in Eq. (6).

For a system in equilibrium with chemical potential μ, the lesser Green’s function G< is calculated using

G<(z,z,Ez)=iF0(Ez,μ)A(z,z,Ez).

This equation is known as the fluctuation–dissipation theorem. Hereby, the spectral function is filled up according to the Fermi distribution function. Note that we use F0 which is the appropriate zero-order Fermi–Dirac integral for homogeneous semiconductor heterostructures defined in Eq. (33). From the diagonal part (z = z′) of G<, we can directly extract the energy resolved electron density

n(z,Ez)=i2πG<(z,z,Ez).

A device is driven out of equilibrium if the source and drain contacts have different chemical potentials, μSμD. In the special case of a ballistic situation where scattering is absent, G< can be calculated also in a nonequilibrium situation. However, we have to separate the total spectral function A into the spectral functions due to the source and drain contacts, A = AS + AD. We then fill up each spectral function by the corresponding Fermi distribution function of the source or drain contact, respectively,

G<(z,z,Ez)=i(F0,S(Ez,μS)AS(z,z,Ez)+F0,D(Ez,μD)AD(z,z,Ez)).

Within the MSB model, the total spectral function A is altered by adding a further term AB. This is the spectral function due to the Büttiker probe contacts. In fact, AB is the sum of the spectral functions of each individual Büttiker probe,

AB(z,z,Ez)=p=1NAp(z,z,Ez).

The spectral function of each probe Ap (z,z,Ez) is given by

Ap=GRΓpGR.

It is determined by its self-energy ΣpR through the broadening function Γp (z,z,Ez),

Γp=i(ΣpRΣpR).

In our implementation the broadening function Γp of each probe is only nonzero at the position coordinate of the probe (zp = z = z′). By replacing the index p with S or D in the last two equations, the corresponding quantities for the source and drain are obtained. The spectral function Ap of each probe is in general nonzero everywhere, i.e. each probe contributes to all other grid points and their correlations (z,z′) a specific spectral function, similar as the source and drain contacts. Therefore, we need a (local) Fermi distribution for each probe, in addition to the Fermi distributions of the source and drain contacts.

Now we introduce the approximation to calculate G< in nonequilibrium using the expression for an equilibrium system. By doing so, we use a local Fermi distribution F0(z,Ez) ≡ F0,p(Ez, μp) for the occupation of the spectral function Ap of each probe p, which therefore depends on the spatial coordinate z, i.e. on a local virtual chemical potential μ(z). Now G< is given by

G<(z,z,Ez)=i(i=S,DF0,i(Ez,μi)Ai(z,z,Ez)+p=1NF0,p(Ez,μp)Ap(z,z,Ez)),
taking into account source, drain and probe contacts. The standard Büttiker probe model uses the logarithmic function of the zero-order Fermi–Dirac integral
F0(z,Ez)=gm*(z)kBT2π2ln(1+exp(Ezμ(z)kBT)),
which takes care of the transverse modes k and includes the spin degeneracy factor g = 2. This is the Fermi distribution function of a semiconductor structure that is homogeneous in the xy plane. Although the virtual contacts are treated in a similar way as the real contacts, the crucial and important difference is that they must not carry any current, i.e. the current between the device and the probes must vanish. In order to guarantee current conservation, div j = 0, the net current jp at every probe must vanish,
jp= dEzjp(Ez)=0.

The physical motivation for this is straightforward. Scattering of carriers leads to a redistribution of their energy and momentum while charge and current conservation laws must remain valid. This leads to a constraint for determining the local Fermi level μ(z) which means that the virtual chemical potentials μ are not determined a priori. They can therefore be calculated in such a way that current conservation is fulfilled. Hence, scattering events can be viewed as absorption of charge carriers into the probes and a reinjection of equilibrated charge carriers back into the device because the virtual probe contacts are assumed to be in a thermodynamic equilibrium, analogously to the real contacts. This is physically motivated by the fact that scattering turns a distorted system back to equilibrium. However, in our implementation of two-contact devices we do not determine μ(z). Instead, current conservation can also be fulfilled by solving a linear system of equations for the coefficients c(z) ∈ [0,1] that are defined as follows

F(z,Ez)=c(z)FS(Ez)+(1c(z))FD(Ez).

The local distribution function F(z,Ez) of the probe at position z is assumed to be a linear combination of the Fermi distribution functions FS and FD of the source and drain contacts. If c = 0, the probe is filled up completely by the drain contact, and if c = 1, the probe is filled up completely by the source contact. For an equilibrium simulation, it obviously must hold c = 0.5. The assumption of a linear combination of the source and drain distributions for a probe at position z is reasonable because the carriers inside a two-contact device originate either from the source or drain and also leave through the source or drain. Since both leads are assumed to be in equilibrium, scattering mechanisms inside the device result in a mixture of the equilibrium distributions. Consequently, we do not have to calculate the virtual chemical potentials μ(z) that are typically used in the standard Büttiker probe model. Instead, it is sufficient to calculate the coefficients c(z) which define a local distribution function. This has numerical advantages and makes the overall iteration cycle more stable. In quantum cascade lasers electron–electron scattering is responsible for efficient energy redistribution and effective thermalization. This has been predicted theoretically [10] and observed experimentally in mid-infrared quantum cascade structures [11], resonant phonon THz QCLs [12] or bound-to-continuum THz QCLs [1315]. Thus we believe that using a local Fermi–Dirac distribution is reasonable. Unfortunately, our simplified model does not give insight on individual subband temperatures. Therefore such an influence on the thermal performance of quantum cascade lasers cannot be answered.

Although the novel MSB method is a rather rigorous approximation that is orders of magnitude faster in terms of computational work compared to the full NEGF model, it reproduces the results of the latter and also experimental data very nicely as demonstrated in [5] for a few QCL structures that we investigated. Concretely, we tested the MSB method against experiments [16] of an AlGaAs/GaAs THz QCL (three quantum well resonant phonon design), and against experiments [17, 18] and NEGF simulations [19] of another AlGaAs/GaAs THz QCL (four quantum well resonant phonon design). Furthermore, we compared our MSB model with NEGF simulations [20] of an InGaAs/GaAsSb THz QCL proposal with four quantum wells and a diagonal optical transition. Here, the MSB method reproduces the major observable quantities very well, although rough interfaces and alloy disorder are neglected. Finally, we benchmarked our calculations with respect to experimental data of a strain-balanced short-injector AlInAs/InGaAs mid-infrared QCL (diagonal transition design) [21] and an injectorless AlInAs/InGaAs mid-infrared QCL [22]. The peculiarity of the latter design is that there are in fact two different lasing frequencies depending on the applied bias. This feature was excellently reproduced by our calculations. In all these cases, the agreement was very encouraging. However, we were not able to simulate experimental results of the AlInAs/InGaAs mid-infrared QCL design of [23]. It turned out that the reason for this disagreement was the neglect of the nonparabolicity of the effective mass within our MSB model. This significantly changed the alignment of the higher-lying electronic states and thus the operation principle could not be reconstructed at all. This was proved by a full NEGF calculation using the software presented in [3] which includes nonparabolic effects and thus reproduced the experimentally observed transition correctly. On the other hand, assuming a parabolic mass in this full NEGF model leads to the same results as our MSB model for this structure. Obviously, the simplifications within the MSB model retain the main physical features very well for both THz and mid-infrared QCLs although special care has to be taken in terms of designs where nonparabolicity effects play an important role. In these cases we recommend complementary simulations using other models that take this into account. The MSB algorithm has been implemented into the nextnano.MSB software (http://www.nextnano.com) and allows finding an optimal QCL layout very quickly, as variations in alloy concentration, barrier thickness, and further parameters can be calculated in parallel.

3. Results

The emission wavelength of quantum cascade lasers ranges from the mid-infrared to the farinfrared or THz domain. QCLs are designed to enable transitions from an upper energy state to a lower energy state within the conduction band. The emitted photons have exactly the energy difference of the aforementioned states. This enables the engineering of the exact energy of the emitted photons. This photon energy is determined by the alignment of the states within the QCL and it can be tailored over a wide range through the variation of the well and barrier thicknesses, barrier heights, doping concentrations, as well as the employed materials. To optimize and design quantum cascade laser devices an efficient yet accurate calculation of dissipative transport properties is essential. Supported by recent calculations with our new MSB method, we propose a novel two-well injectorless THz quantum cascade laser design with a photon emission frequency of 2.85 THz which corresponds to a photon energy of Eγ = 11.8 meV. The design is based on the following principles. First, the device is based on the GaAs/AlGaAs material system. Although, devices based on InAs or AlSb intrinsically achieve better gain performances due to their lower effective masses [24], GaAs-based technology offers several advantages, e.g. the GaAs material system is well understood and mature processing technologies are available [25,26]. Secondly, we use an injectorless design which has the advantage of high gain [27] since the device is shorter and the intermediate injector–collector states are neither needed to fill the upper laser states nor to empty the lower laser states. Thus, in an injectorless two-well QCL the active region extends over almost the whole period and consequently the ratio of gain to nongain regions is much higher [28]. Thirdly, the design is based on narrow wells and barriers. This pushes the bound states to higher energies and therefore, so-called parasitic leakage currents from one period to the next are suppressed. Additionally, we do not impose the restriction of uniform barrier heights, i.e. we use different aluminum contents for the barriers. This additional degree of freedom for the QCL design can be used to improve temperature performance [29] and to find a design that is robust against fluctuations of material compositions. Finally, the design is based on the ‘complete thermalization’ of carriers [20]. Typical THz QCLs are designed such that the potential drop per period ϕth at threshold bias equals the sum of the emitted photon energy Eγ and the energy ELO of one longitudinal optical (LO) phonon, th = Eγ + ELO. As most electrons pass the period without emitting a photon, electrons might get heated up unless thermalized by electron–electron scattering. Hot electrons can be avoided by designing the threshold bias to be an integer multiple of the LO phonon energy, for instance th = 2ELO [20, 30], thereby maximizing the carrier thermalization within each period. Our design uses th = ELO and hence, for the energy balance of one period, the photon energy is negligible since only a minority of carriers emits a photon. Consequently, the majority of carriers, i.e. those that are not emitting light, are efficiently thermalized and nonperiodic effects are remedied [19,31].

Our concrete suggestion of a novel THz quantum cascade laser consists of GaAs wells and alternating Al0.1Ga0.9As and Al0.27Ga0.73As barriers. The barrier heights used in the calculation are 107 meV and 237 meV, respectively. The layer sequence for one period is 2.1/7.35/2.1/7.35 nm (Al0.1Ga0.9As/GaAs/Al0.27Ga0.73As/GaAs). The AlGaAs barriers are shown in bold and the GaAs wells in normal font. 7.35 nm correspond to 26 molecular layers of GaAs, and 2.1 nm to 7 molecular layers of AlxGa1−xAs, respectively. Only the underlined GaAs layer of each period is doped with a concentration of 3 × 1016 cm−3 where we assume the donors to be fully ionized. The widths of the layers as well as the heights of the barriers, which are defined by the alloy compositions, were optimized with a genetic algorithm based on the MSB method. Concretely, in terms of barrier widths wb and well widths ww, we were searching for structures that fit on a grid space resolution of a = 0.15 nm using the constraint wb = na and ww = ma with n and m being integers. Since the MSB method is very efficient in terms of computational memory and time, various alloy concentrations, barrier thicknesses, and further parameters can be calculated in parallel and an optimal QCL layout can be found quickly.

The operation scheme of this QCL is shown in Fig. 2 and Fig. 3(a). In principle, this QCL design consists of alternating upper and lower laser states only. The upper laser states are labeled ‘1a’ and ‘2a’ in Fig. 2 and are located in the wells labeled ‘A’ in Fig. 3(a). From this upper laser level a diagonal photon transition into the states ‘1b’ and ‘2b’, respectively, is possible. The laser transition between these states is diagonal and thus suppresses nonradiative transitions. Rather than emitting a photon, however, the majority of carriers scatters efficiently from state ‘1a’ into the upper laser state ‘2a’ of the next period because this energy difference is designed to be the energy of a longitudinal optical phonon. For the GaAs material system, the energy of a LO phonon is assumed to be ELO = 34 meV and thus, the threshold voltage of our novel QCL design is expected to be slightly below 34 mV where the states are aligned correctly for the LO phonon cascade. The threshold voltage is the applied bias voltage for which the optical gain compensates all optical losses, and thus the QCL starts to emit photons.

 figure: Fig. 2

Fig. 2 (a) Calculated conduction band profile (white line) and contour plot of the energy and position resolved spectral function A(z = z,Ez) at vanishing in-plane momentum k = 0 for the suggested QCL at the threshold bias voltage of 36 mV per period and a lattice temperature of 100 K. (b) Corresponding contour plot of the energy and position resolved current density.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 (a) Operation scheme of the novel QCL design. The red arrows indicate LO phonon emissions and the blue arrows indicate photon emissions. The two wells are labeled with ‘A’ and ‘B’ and only well ‘A’ is doped. (b) Calculated contour plot of the position resolved optical intensity gain as a function of the photon energy. The black solid contour line encloses the area of positive optical gain. The white solid line indicates the conduction band profile and is only meant to guide the eye. It is not related to the photon energy axis. The applied bias voltage is 36 mV and the temperature is 100 K.

Download Full Size | PDF

The effects of the narrow wells and high barriers are also shown in Fig. 2. The upper states labeled ‘1u’ and ‘2u’ are pushed more than 100 meV away from the conduction band edge. Thus they do not interact and interfere with other states of previous periods. Consequently, they do not carry any significant current, as shown in Fig. 2(b). The simulated structure extends from 0 nm to 105.6 nm. The left lead is positioned at 0 nm, followed by a 4.5 nm GaAs layer. Then we include 5 periods to minimize boundary effects, followed by a 2.1 nm Al0.1Ga0.9As barrier and an additional GaAs layer of 4.5 nm in front of the right lead. The labels ‘1a’ and ‘2a’ of Fig. 2 are located in the third period, the labels ‘2a’ and ‘2b’ in the fourth period.

We calculated the optical intensity gain as a function of the photon energy using the dipole approximation. We follow the approach of [32] where the gain is calculated for current-driven semiconductor heterostructures within the NEGF framework. The calculated position resolved optical intensity gain as a function of the photon energy is shown in Fig. 3(b). This plot gives insight into laser operation, i.e. one can easily see which parts of the structure are responsible for absorption or emission. Here, the black solid line encloses the area where the optical gain is positive. We find that almost the complete period contributes to the optical gain. Our calculated gain does not include optical losses which are typically at least of the order of 15 cm1 in other QCLs [16,21,33].

The calculated optical gain as a function of the lattice temperature is shown in Fig. 4(a). The corresponding calculated optical gain as a function of the photon energy is plotted for various temperatures in Fig. 4(b). We find significantly higher optical gain than for the four-well QCL published in [20]. Furthermore, we calculate gain even up to temperatures of 250 K which would be a new temperature record for THz quantum cascade lasers. This high temperature operation is confirmed by a complementary, independent simulation based on a different NEGF code [19]. Our previous simulations of a few other common THz QCL designs do not show such a high temperature operation, in agreement to experiment. In experiments, efficient cooling down to 230 K can be achieved easily with simple Peltier coolers in contrast to more sophisticated and more expensive cooling mechanisms which are required for lower temperatures [33,34]. The present temperature performance record of 199.5 K is achieved by a resonant phonon based three-well design [35].

 figure: Fig. 4

Fig. 4 (a) Calculated optical intensity gain as a function of the lattice temperature. The solid black line was calculated for the QCL proposed in [20] whereas the solid blue line is calculated for the novel QCL design. (b) Calculated optical intensity gain as a function of the photon energy for the novel QCL design for various temperatures.

Download Full Size | PDF

Finally, we also calculated the DC current as a function of the applied bias voltage for our proposed QCL design. The current–voltage characteristics are plotted in Fig. 5. We calculated I–V curves for various temperatures from 100 K to 250 K. Our calculations do not show any negative differential resistance below the threshold voltage of 36 mV. The simulation indicates an electrically stable operation of our novel QCL design. Note that usually a negative differential resistance can be observed above the threshold voltage in NEGF as well as in our MSB calculations (not shown). This is caused by state misalignment and originates mainly from the fact that the coupling of the photon field is not included within the calculations. Furthermore, we found that the gain is not very sensitive with respect to layer thickness fluctuations.

 figure: Fig. 5

Fig. 5 Calculated DC current as a function of the applied bias voltage for the novel QCL design for various temperatures. The current–voltage characteristics do not show any negative differential resistance and thus indicate an electrically stable design.

Download Full Size | PDF

4. Conclusion

We developed a novel method for the calculation of inelastic electron transport in quantum cascade lasers that takes into account individual scattering mechanisms. It is based on the nonequilibrium Green’s function formalism and treats the quantum cascade laser as an open quantum device. It offers the numerical efficiency of the simple Büttiker probe model but more importantly, it extends it by using realistic self-energies derived from the NEGF method for the probes. The assumption of local equilibrium simplifies the solution of the coupled equations. We presented a novel concrete THz QCL scheme based on a two-well design with alternating barrier heights where the threshold voltage is designed to match the energy of a LO phonon. The widths of the layers as well as the heights of the barriers were optimized with a genetic algorithm based on this newly developed multi-scattering Büttiker probe (MSB) model. Our calculations predict THz laser operation for temperatures up to 250 K.

Acknowledgments

S. Birner thanks C. Jirauschek, T. Grange, T. Kotani and T. Kubis for fruitful discussions. This work was supported by the German Research Foundation ( DFG) and the Technische Universität München within the funding programme Open Access Publishing.

References and links

1. S. Datta, “Nanoscale device modeling: the Green’s function method,” Superlattices Microstruct. 28 (4), 253–278 (2000). [CrossRef]  

2. A. Wacker, “Semiconductor superlattices: a model system for nonlinear transport,” Phys. Rep. 357 (1), 1–111 (2002). [CrossRef]  

3. T. Grange, “Nanowire terahertz quantum cascade lasers,” Appl. Phys. Lett. 105 (14), 141105 (2014). [CrossRef]  

4. T. Kubis and P. Vogl, “Assessment of approximations in nonequilibrium Green’s function theory,” Phys. Rev. B 83, 195304 (2011). [CrossRef]  

5. P. Greck, Efficient calculation of dissipative quantum transport properties in semiconductor nanostructures, Selected Topics of Semiconductor Physics and Technology (G. Abstreiter, M.-C. Amann, M. Stutzmann, and P. Vogl, eds.), vol. 105 (Verein zur Förderung des Walter Schottky Instituts der Technischen Universität München e.V., München, 2012).

6. M. Büttiker, “Role of quantum coherence in series resistors,” Phys. Rev. B 33 (5), 3020–3026 (1986). [CrossRef]  

7. M. Büttiker, “Symmetry of electrical conduction,” IBM J. Res. Develop. 32 (3), 317–334 (1988). [CrossRef]  

8. A. Trellakis, A. T. Galick, A. Pacelli, and U. Ravaioli, “Iteration scheme for the solution of the two-dimensional Schrödinger-Poisson equations in quantum structures,” J. Appl. Phys. 81 (12), 7880–7884 (1997). [CrossRef]  

9. S. Li and E. Darve, “Extension and optimization of the FIND algorithm: Computing Green’s and less-than Green’s functions,” J. Comput. Phys. 231 (4), 1121–1139 (2012). [CrossRef]  

10. P. Harrison, “The nature of the electron distribution functions in quantum cascade lasers,” Appl. Phys. Lett. 75 (18), 2800–2802 (1999). [CrossRef]  

11. M. Troccoli, G. Scamarcio, V. Spagnolo, A. Tredicucci, C. Gmachl, F. Capasso, D. L. Sivco, A. Y. Cho, and M. Striccoli, “Electronic distribution in superlattice quantum cascade lasers,” Appl. Phys. Lett. 77 (8), 1088–1090 (2000). [CrossRef]  

12. M. S. Vitiello, G. Scamarcio, V. Spagnolo, B. S. Williams, S. Kumar, Q. Hu, and J. L. Reno, “Measurement of subband electronic temperatures and population inversion in THz quantum-cascade lasers,” Appl. Phys. Lett. 86 (11), 111115 (2005). [CrossRef]  

13. M. S. Vitiello, G. Scamarcio, V. Spagnolo, T. Losco, R. P. Green, A. Tredicucci, H. E. Beere, and D. A. Ritchie, “Electron-lattice coupling in bound-to-continuum THz quantum-cascade lasers,” Appl. Phys. Lett. 88 (24), 241109 (2006). [CrossRef]  

14. M. S. Vitiello, G. Scamarcio, V. Spagnolo, C. Worrall, H. E. Beere, D. A. Ritchie, C. Sirtori, J. Alton, and S. Barbieri, “Subband electronic temperatures and electron-lattice energy relaxation in terahertz quantum cascade lasers with different conduction band offsets,” Appl. Phys. Lett. 89 (13), 131114 (2006). [CrossRef]  

15. M. S. Vitiello, T. Gresch, A. Lops, V. Spagnolo, G. Scamarcio, N. Hoyler, M. Giovannini, and J. Faist, “Influence of InAs, AlAs δ layers on the optical, electronic, and thermal characteristics of strain-compensated GaInAs/AlInAs quantum-cascade lasers,” Appl. Phys. Lett. 91 (16), 161111 (2007). [CrossRef]  

16. N. Jukam, S. S. Dhillon, D. Oustinov, J. Madéo, J. Tignon, R. Colombelli, P. Dean, M. Salih, S. P. Khanna, E. H. Linfield, and A. G. Davies, “Terahertz time domain spectroscopy of phonon-depopulation based quantum cascade lasers,” Appl. Phys. Lett. 94 (25), 251108 (2009). [CrossRef]  

17. A. Benz, G. Fasching, A. M. Andrews, M. Martl, K. Unterrainer, T. Roch, W. Schrenk, S. Golka, and G. Strasser, “Influence of doping on the performance of terahertz quantum-cascade lasers,” Appl. Phys. Lett. 90 (10), 101107 (2007). [CrossRef]  

18. A. M. Andrews, A. Benz, C. Deutsch, G. Fasching, K. Unterrainer, P. Klang, W. Schrenk, and G. Strasser, “Doping dependence of LO-phonon depletion scheme THz quantum-cascade lasers,” Mater. Sci. Eng. B 147 (2–3), 152–155 (2008). [CrossRef]  

19. T. Kubis, C. Yeh, P. Vogl, A. Benz, G. Fasching, and C. Deutsch, “Theory of nonequilibrium quantum transport and energy dissipation in terahertz quantum cascade lasers,” Phys. Rev. B 79 (19), 195323 (2009). [CrossRef]  

20. T. Kubis, S. R. Mehrotra, and G. Klimeck, “Design concepts of terahertz quantum cascade lasers: Proposal for terahertz laser efficiency improvements,” Appl. Phys. Lett. 97 (26), 261106 (2010). [CrossRef]  

21. A. Vizbaras, S. Katz, G. Boehm, and M.-C. Amann, “Short-injector quantum cascade laser emitting at 8-μ m wavelength with high slope efficiency,” IEEE Photon. Technol. Lett. 21 (19), 1384–1386 (2009). [CrossRef]  

22. A. Friedrich, G. Boehm, M. C. Amann, and G. Scarpa, “Quantum-cascade lasers without injector regions operating above room temperature,” Appl. Phys. Lett. 86 (16), 161114 (2005). [CrossRef]  

23. J. S. Yu, S. Slivken, and M. Razeghi, “Injector doping level-dependent continuous-wave operation of InP-based QCLs at λ ∼ 7. 3 μ m above room temperature,” Semicond. Sci. Technol. 25 (12), 125015 (2010). [CrossRef]  

24. E. Benveniste, A. Vasanelli, A. Delteil, J. Devenson, R. Teissier, A. Baranov, A. M. Andrews, G. Strasser, I. Sagnes, and C. Sirtori, “Influence of the material parameters on quantum cascade devices,” Appl. Phys. Lett. 93 (13), 131108 (2008). [CrossRef]  

25. C. Sirtori, H. Page, and C. Becker, “GaAs-based quantum cascade lasers,” Phil. Trans. R. Soc. Lond. A 359 (1780), 505–522 (2001). [CrossRef]  

26. M. Garcia, E. Normand, C. R. Stanley, C. N. Ironside, C. D. Farmer, G. Duxbury, and N. Langford, “An AlGaAs–GaAs quantum cascade laser operating with a thermoelectric cooler for spectroscopy of NH3,” Opt. Commun. 226 (1–6), 39–43 (2003). [CrossRef]  

27. S. Katz, A. Vizbaras, G. Boehm, and M.-C. Amann, “Nonlinear gain behavior in injectorless quantum cascade lasers,” Opt. Eng. 49 (11), 111107 (2010). [CrossRef]  

28. G. Scalari, M. I. Amanti, C. Walther, R. Terazzi, M. Beck, and J. Faist, “Broadband THz lasing from a photon-phonon quantum cascade structure,” Opt. Express 18 (8), 8043–8052 (2010). [CrossRef]   [PubMed]  

29. A. Matyas, R. Chashmahcharagh, I. Kovacs, P. Lugli, K. Vijayraghavan, M. A. Belkin, and C. Jirauschek, “Improved terahertz quantum cascade laser with variable height barriers,” J. Appl. Phys. 111 (10), 103106 (2012). [CrossRef]  

30. E. Dupont, S. Fathololoumi, Z. R. Wasilewski, G. Aers, S. R. Laframboise, M. Lindskog, S. G. Razavipour, A. Wacker, D. Ban, and H. C. Liu, “A phonon scattering assisted injection and extraction based terahertz quantum cascade laser,” J. Appl. Phys. 111 (7), 073111 (2012). [CrossRef]  

31. T. Kubis and P. Vogl, “Predictive quantum theory of current and optical emission in quantum cascade lasers,” Proc. SPIE 7230, 723019 (2009). [CrossRef]  

32. A. Wacker, “Gain in quantum cascade lasers and superlattices: A quantum transport theory,” Phys. Rev. B 66 (8), 085326 (2002). [CrossRef]  

33. M. A. Belkin, Q. J. Wang, C. Pflügl, A. Belyanin, S. P. Khanna, A. G. Davies, E. H. Linfield, and F. Capasso, “High-temperature operation of terahertz quantum cascade laser sources,” IEEE J. Sel. Topics Quantum Electron. 15 (3), 952–967 (2009). [CrossRef]  

34. T. Aellen, S. Blaser, M. Beck, D. Hofstetter, J. Faist, and E. Gini, “Continuous-wave distributed-feedback quantum-cascade lasers on a Peltier cooler,” Appl. Phys. Lett. 83 (10), 1929–1931 (2003). [CrossRef]  

35. S. Fathololoumi, E. Dupont, C. W. I. Chan, Z. R. Wasilewski, S. R. Laframboise, D. Ban, A. Mátyás, C. Ji-rauschek, Q. Hu, and H. C. Liu, “Terahertz quantum cascade lasers operating up to ~200 K with optimized oscillator strength and improved injection tunneling,” Opt. Express 20 (4), 3866–3876 (2012). [CrossRef]   [PubMed]  

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 (5)

Fig. 1
Fig. 1 Left: The standard Büttiker probe model associates a single momentum and energy sink with each device grid point i in position space using a phenomenological scattering parameter η, which is fixed during the calculation. Right: The proposed multi-scattering Büttiker probe model accounts for individual scattering mechanisms such as longitudinal acoustic (LA) or longitudinal optical (LO) phonon scattering by including multiple scattering potentials for each node that are calculated self-consistently.
Fig. 2
Fig. 2 (a) Calculated conduction band profile (white line) and contour plot of the energy and position resolved spectral function A(z = z,Ez) at vanishing in-plane momentum k = 0 for the suggested QCL at the threshold bias voltage of 36 mV per period and a lattice temperature of 100 K. (b) Corresponding contour plot of the energy and position resolved current density.
Fig. 3
Fig. 3 (a) Operation scheme of the novel QCL design. The red arrows indicate LO phonon emissions and the blue arrows indicate photon emissions. The two wells are labeled with ‘A’ and ‘B’ and only well ‘A’ is doped. (b) Calculated contour plot of the position resolved optical intensity gain as a function of the photon energy. The black solid contour line encloses the area of positive optical gain. The white solid line indicates the conduction band profile and is only meant to guide the eye. It is not related to the photon energy axis. The applied bias voltage is 36 mV and the temperature is 100 K.
Fig. 4
Fig. 4 (a) Calculated optical intensity gain as a function of the lattice temperature. The solid black line was calculated for the QCL proposed in [20] whereas the solid blue line is calculated for the novel QCL design. (b) Calculated optical intensity gain as a function of the photon energy for the novel QCL design for various temperatures.
Fig. 5
Fig. 5 Calculated DC current as a function of the applied bias voltage for the novel QCL design for various temperatures. The current–voltage characteristics do not show any negative differential resistance and thus indicate an electrically stable design.

Equations (35)

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

H ( z , k ) = H L ( z ) + H T ( z , k ) ,
H L ( z ) = 2 2 d d z ( 1 m ( z ) d d z ) + E c , 0 ( z ) e ϕ ( z ) ,
H T ( z , k ) = 2 k 2 2 m * ( z ) ,
· ( ε 0 ε s ( z ) ϕ ( z ) ) = e ( N D + ( z ) n ( z , ϕ ) ) .
G R = ( E 1 H Σ R ) 1 ,
Σ R = Σ C R + G R D R + G < D < ,
Σ < = Σ C < + G < D < ,
G < = G R Σ < G R .
E = E z + E = E z + 2 k 2 2 m * ,
G ( k , E ) G ( E E ) G ( E z ) .
Σ R = Σ C R + Σ B R ,
Σ C R = Σ S R + Σ D R ,
Σ S R ( 1 , 1 , E z ) = t exp ( i k S a ) ,
Σ D R ( N , N , E z ) = t exp ( i k D a ) ,
t = 2 2 m * a 2 .
E z ( k ) = E c + 2 t ( 1 cos ( k a ) ) ,
B ( z , E z ) = i η ( z ) ,
B ( z , E z ) = B LA ( z , E z ) + B LO ( z , E z ) ,
ρ ( z , E z ) = 1 2 π A ( z , z , E z ) .
A = i ( G R G R ) = 2 ( G R ) ,
B LA ( z , E z ) = i V D 2 k B T 8 π ρ M v s 2 E LA E z E LA E z + E LA d E z ρ ( z , E z ) ,
B LO ( z , E z ) = i e 2 ξ E LO 32 π ε 0 ( ε 1 ε s 1 )   [ ( 1 + N LO ) ρ ( z , E z E LO ) + N LO ρ ( z , E z + E LO ) ] ,
N LO = ( exp ( E LO k B T ) 1 ) 1 .
Σ p R ( z , z , E z ) = B ( z , E z ) δ ( z z p ) .
Σ B R ( z , z , E z ) = p = 1 N Σ p R ( z , z , E z ) .
G < ( z , z , E z ) = i F 0 ( E z , μ ) A ( z , z , E z ) .
n ( z , E z ) = i 2 π G < ( z , z , E z ) .
G < ( z , z , E z ) = i ( F 0 , S ( E z , μ S ) A S ( z , z , E z ) + F 0 , D ( E z , μ D ) A D ( z , z , E z ) ) .
A B ( z , z , E z ) = p = 1 N A p ( z , z , E z ) .
A p = G R Γ p G R .
Γ p = i ( Σ p R Σ p R ) .
G < ( z , z , E z ) = i ( i = S , D F 0 , i ( E z , μ i ) A i ( z , z , E z ) + p = 1 N F 0 , p ( E z , μ p ) A p ( z , z , E z ) ) ,
F 0 ( z , E z ) = g m * ( z ) k B T 2 π 2 ln ( 1 + exp ( E z μ ( z ) k B T ) ) ,
j p =   d E z j p ( E z ) = 0.
F ( z , E z ) = c ( z ) F S ( E z ) + ( 1 c ( z ) ) F D ( E z ) .
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.