— SOON — T-shaped quantum wire grown by cleaved edge overgrowth (CEO): wave functions and strain

Note

The tutorial is related to the PhD Thesis of R. Schuster [SchusterPhD2005]

Calculation of the strain tensor

First, we have to calculate the strain tensor by minimizing the elastic energy within continuum elasticity theory. Along the translationally invariant \(z\) direction the lattice commensurability constraint forced the \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\) layer to adopt the lattice constant of \(\mathrm{Al}_{0.3}\mathrm{Ga}_{0.7}\mathrm{As}\). The model for strain calculations can be specified inside the strain{ } group, where we choose the model: minimized_strain{ }.

In Figure 2.4.172 the calculated hydrostatic strain \(\epsilon_\mathrm{hyd} = \epsilon_{xx} + \epsilon_{yy} + \epsilon_{zz}\) (trace of the strain tensor) inside the structure is shown. The hydrostatic strain has its maximum at the intersection, where it leads to a reduced band gap, which is the requirement for confining the charge carriers. Thus, the quantum wire is formed in the \(\mathrm{GaAs}\) quantum well due to the tensile strain field induced by the \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\) layer.

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-InAlAs_strain_hydro_2D_sim_axis.jpg

Figure 2.4.172 In (a) the hydrostatic strain \(\epsilon_\mathrm{hyd}\) inside the T-shaped quantum wire structure is shown. In (b) a cross-section of \(\epsilon_\mathrm{hyd}\) along \(x\) at \(y = 0\) is shown.

Note that in a one-dimensional example, which is provided in the input file T-QWR_zb_III-V_Schuster_PhD_2005_1D_nnp_strained-QW.in, the strain tensor components of a \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\) layer that is strained pseudomorphically with respect to an \(\mathrm{Al}_{0.30}\mathrm{Ga}_{0.7}\mathrm{As}\) substrate are the following:

\[ \begin{align}\begin{aligned}\begin{aligned}\\\epsilon_{xx} &= 10.9 \cdot 10^{-3}\\\epsilon_{yy} &= \epsilon_{zz} = -12.4 \cdot 10^{-3}\\\epsilon_{xy} &= \epsilon_{xz} = \epsilon_{yz} = 0\\\epsilon_\mathrm{hyd} &= \mathrm{Tr}(\epsilon_{ij}) = -13.9 \cdot 10^{-3}\\\end{aligned}\end{aligned}\end{align} \]

Here, the growth direction is along the \(x\) direction, i.e. along [100]. The temperature is assumed to be \(40\,\mathrm{K}\) and the lattice constants are assumed to be temperature dependent (i.e. we use the \(40\,\mathrm{K}\) lattice constants).

In Figure 2.4.173 the individual strain tensor components (\(\epsilon_{xx}\), \(\epsilon_{yy}\), \(\epsilon_{xy}\)) with respect to the simulation coordinate system are presented. In our 2D simulation, the sample layout is homogeneous along the \(z\) direction, i.e. the lattice constant of \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\) is forced to have the same lattice constant as \(\mathrm{Al}_{0.3}\mathrm{Ga}_{0.7}\mathrm{As}\) along the \(z\) direction. Then the strain tensor component must be \(\epsilon_{zz} = -12.4 \cdot 10^{-3}\), in agreement with our 1D example, i.e. \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\), which has a larger lattice constant than \(\mathrm{Al}_{0.3}\mathrm{Ga}_{0.7}\mathrm{As}\) is strained compressively along the \(z\) direction. Similar to the 1D case, it is also expected that the \(\epsilon_{yy}\) component inside the \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\) barrier has a similar value to \(\epsilon_{zz}\), which is clearly the case. The dark blue area in Figure 2.4.173 (c) thus has a value around \(-12 \cdot 10^{-3}\). However, this value deviates from the ideal 1D value at the T-shaped intersection as expected (see also Figure 2.4.174). The same applies to the value of \(\epsilon_{xx}\), which is similar to the 1D value inside the \(\mathrm{In}_{0.16}\mathrm{Al}_{0.84}\mathrm{As}\) barrier: \(\epsilon_{xx} = 11 \cdot 10^{-3}\). The strain tensor components \(\epsilon_{xz}\) and \(\epsilon_{yz}\) with respect to the simulation coordinate system are equal to zero as in our 1D example.

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-InAlAs_strain_xxyyxy_2d_sim.jpg

Figure 2.4.173 In (a), (c), (e) the strain components \(\epsilon_{xx}\), \(\epsilon_{yy}\), \(\epsilon_{xy}\) are shown. In (b), (d), (f) a cut through the structure along \(x\) at \(y=0\) is shown.

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-InAlAs_strain_xx_quer_2d_sim.jpg

Figure 2.4.174 Strain tensor component \(\epsilon_{xx}\) along \(y\) direction at position \(x=0\).

The important difference with respect to the 1D case is the existence of a non-vanishing strain tensor component \(\epsilon_{xy}\) which brakes the symmetry of the sample layout. Usually, the \(\epsilon_{xy}\) component is attributed to be responsible for piezoelectricity. However, note that in the discussion before all strain tensor components refer to the simulation coordinate system (and not to the crystal coordinate system). So we have to plot the off-diagonal strain tensor components that are expressed with respect the crystal coordinate system orientation and then check if the off-diagonal components are non-zero, which is clearly the case as we can see from Figure 2.4.175.

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-InAlAs_strain_2d_cr.jpg

Figure 2.4.175 Strain tensor components \(\epsilon_{\tilde{x}\tilde{x}}\), \(\epsilon_{\tilde{y}\tilde{y}}\), \(\epsilon_{\tilde{x}\tilde{y}}=\epsilon_{\tilde{x}\tilde{z}}\) and \(\epsilon_{\tilde{y}\tilde{z}}\) with respect to the crystal coordinate system. The rotation with respect to the simulation system is a rotation of 45 degrees around the \(x\) axis, i.e. the [100] axis.

By comparing Figure 2.4.173 (a) and Figure 2.4.175 (a) we observe that \(\epsilon_{\tilde{x}\tilde{x}} = \epsilon_{xx}\), because the \(x\) coordinate axes coincide. Symmetry arguments show that the following holds:

\[ \begin{align}\begin{aligned}\begin{aligned}\\ \epsilon_{\tilde{y}\tilde{y}} &= \frac{1}{2} \left( \epsilon_{yy} + \epsilon_{zz} \right)\\ \epsilon_{\tilde{x}\tilde{y}} &= \epsilon_{\tilde{x}\tilde{z}} = \frac{1}{\sqrt{2}} \epsilon_{xy}\\\end{aligned}\end{aligned}\end{align} \]

Calculation of the piezoelectric charge density

The off-diagonal strain tensor components \(\epsilon_{\tilde{x}\tilde{y}}\), \(\epsilon_{\tilde{x}\tilde{z}}\) and \(\epsilon_{\tilde{y}\tilde{z}}\) are responsible for the piezoelectric polarization \(\mathbf{P}_\mathrm{piezo}\), given by

\[\begin{split}\mathbf{P}_\mathrm{piezo} = e_{14}\begin{pmatrix} 2\epsilon_{\tilde{y}\tilde{z}}\\ 2\epsilon_{\tilde{x}\tilde{z}}\\ 2\epsilon_{\tilde{x}\tilde{y}} \end{pmatrix},\end{split}\]

where \(e_{14}\) is the piezoelectric constant in units of \([\mathbf{C/m^2}]\). Once having determined the piezoelectric polarization, one is able to compute the piezoelectric charge density:

\[\rho_\mathrm{piezo}(x,y) = -\mathrm{div}\,\mathbf{P}_\mathrm{piezo}(x,y).\]

In Figure 2.4.176 the piezo electric charge density inside the quantum wire structure is shown. The strain-induced piezoelectric fields are then obtained from \(\rho_\mathrm{piezo}\) by solving Poisson’s equation.

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-piezo_charge_Tshaped2D.jpg

Figure 2.4.176 Piezoelectric charge density \(\rho_\mathrm{piezo}(x,y)\).

Calculation of the conduction and valence band edges

In Figure 2.4.177 the conduction and valence band edges of the structure are shown. The conduction and valence band edges were determined by taking into account the shifts and splittings due to the relevant deformation potentials as well as the changes due to the piezoelectric fields. We observe that the electron feels a conduction band minimum which is located left with respect to the T-shaped intersection. For the valance bands, we see that the valence band maximum for the heavy hole is not at the same position as the valence band maximum for the light hole.

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-InAlAs_bands.jpg

Figure 2.4.177 In (a), (c), (e) a 2D plot of the conduction, heavy hole and light hole band edge energies are shown. In (b), (d), (f) a cut through the conduction, heavy hole and light hole band edge energies at \(y=0\).

Electron and heavy hole wave functions

Figure 2.4.178 shows the square of the electron (e) and heavy hole (hh) wave functions (i.e. \(\psi^2\)). They were calculated within the effective-mass approximation (single-band).

../../../_images/nnpp_tu_T-shaped-QWR-strained_2D-InAlAs_with_without_piezo.jpg

Figure 2.4.178 In (a) the contour diagram of the square of the electron (e) and heavy hole (hh) wave functions (i.e. \(\psi^2\)) for the case where strain is included in the simulations, but piezoelectricity is not. Subplot (b) shows the same results as in (a), but this time including the piezoelectric effect. Note that in the plot the wave functions are normalized so that the maximum equals one, respectively.

In Figure 2.4.178 (a) the piezoelectric effect was not included. As one can clearly see in Figure 2.4.178 (b), the piezoelectric effect destroys the symmetry of the sample layout. The piezoelectric field results from the \(\epsilon_{xy}\) strain tensor component which is also not symmetric with respect to the T-shaped geometry.

Acknowledgement:

We would like to thank Robert Schuster from the University of Regensburg for providing experimental data and some figures for this tutorial.

Last update: 13/09/2024