Introduction

The InAs/GaSb and InAs/InAs1−xSbx superlattices (SLs) have been considered as an alternative to HgCdTe. These materials were first presented by Smith et al. in 1987 initiating the process of the HgCdTe use for infrared (IR) applications [1,2,3,4]. Recently, researchers have focused on the fact that InAs/InAs1−xSbx SLs are more flexible in the optimization of device performance. The type-2 superlattices (T2SLs) InAs/InAs1−xSbx are used for the fabrication of barrier detectors to include nBn and pBn design [5, 6], low-noise interband cascade infrared photodetectors (ICIP) [7, 8], and avalanche photodiodes (APD) [9], for wide range of applications in the field of science, medicine, technology, safety, industry, medical diagnostic imaging, night vision devices, and spectroscopy.

Figure 1 shows the bandgap (Eg) dependence on the lattice constant (a) for bulk A3B5 materials (InAs, InSb, GaSb, and GaAs) being used for SLs fabrication. In the case of the ternary materials, Vegard’s law was used to determine lattice constant [10,11,12].

Figure 1
figure 1

Lattice constant and bandgap of the bulk A3B5 materials at T = 0 K

The GaSb has a similar lattice constant to InAs and InSb, while the GaAs exhibits a large lattice mismatch in comparison with the mentioned materials leading to the high strains. When growing a SL, it was important to reduce the tension between the SL constituting layers and the substrate on which the SL was deposited. That is why it is important to choose the buffer layer and minimal thickness (with reference to the thickness of the SL) in which the tension between the SL and the buffer layer was averaged. As a result, the proper buffer layer selection reduced the dislocations density having a great influence on the characteristics of the processed detectors. We decided to apply the interfacial misfit array (IMF) technique in order to get a defect-free buffer layer.

In this work, a numerical modeling by APSYS platform of medium wavelength infrared (MWIR) and long wavelength infrared (LWIR) T2SLs InAs/InAs1−xSbx IR structure is presented. In the simulation, the strains created by GaSb buffer in InAs and InAs1−xSbx layers were considered. It was shown that the use of T2SLs InAs/InAs1−xSbx with the “strain-balanced” condition makes it possible to grow high-quality SLs for IR applications.

T2SL InAs/InAs1−xSbx structure

The InAs/InAsSb T2SLs were grown on semi-insulating GaAs (001) substrates with 2° offcut toward <110> in RIBER Compact 21-DZ solid-source molecular beam epitaxy (MBE) system. The substrates were thermally deoxidized, and a 250-nm-thick GaAs layer was deposited at 665 °C in order to smooth the surface after deoxidization. Then, a 1.2-µm-thick GaSb layer was grown using IMF technique to reduce the large lattice mismatch between GaAs substrate and the SLs. Finally, a 300–450 periods, non-intentionally doped InAs/InAs1−xSbx T2SL absorber layer, were deposited at 425 °C. The growth processes have been monitored in situ by RHEED system. The T2SL InAs/InAs1−xSbx structure is presented in Fig. 2. The n describes the number of SL periods, and L is the thickness of SL components. Detailed description of the T2SLs growth procedure is given in [13].

Figure 2
figure 2

Schematic representation of the T2SL InAs/InAs1−xSbx structure

High-resolution X-ray diffraction (PANalytical X’Pert) was utilized to assess the structural and crystallographic properties of the grown samples. The thickness of the several layers in SLs could be determined by the distance between the satellite peaks appearing on the radial scan. In order to know the antimony composition in InAsSb layer, the secondary ion mass spectrometry (SIMS) was used. The PL emission was analyzed using a Bruker Vertex v70 Fourier transform infrared (FTIR) spectrometer to determine the bandgap energy of the InAs/InAsSb T2SLs absorber. Table 1 presents the basic parameters of the grown samples.

Table 1 The T2SLs parameters: thickness L, Sb molar composition xSb, number of periods n

All samples were checked for compliance with the “strain-balanced” condition [14, 15]. This means that the strain occurring between the components of the SL (InAs and InAs1−xSbx) and the buffer layers (GaSb) was small, not causing a large number of dislocations. Hence, the conclusion was made that these strains will not have a significant impact on the characteristics of the device.

The several expressions for designation of the compressive (LInAsSb) and tensile (LInAs) layer thickness have been used. An average lattice method arose from the notion that a “strain-balanced” structure could be reached by averaging of the compressive and tensile lattice parameters. This could be derived assuming identical elastic material properties (material elastic constants), and the balance equation is a form:

$$ a_{\text{GaSb}} \cdot (L_{\text{InAs}} + L_{\text{InAsSb}} ) = a_{\text{InAs}} \cdot L_{\text{InAs}} + a_{\text{InAsSb}} \cdot L_{\text{InAsSb}} , $$
(1)

where ai is the lattice constant of “i” materials.

The thickness of a “strain-balanced” structure for the samples presented in Table 1 was determined by the equation [14, 16]:

$$ L_{\text{InAsSb}} = \left( {\frac{L}{{x_{\text{Sb}} }}} \right) \cdot \left( {\frac{{a_{\text{GaSb}} - a_{\text{InAs}} }}{{a_{\text{InSb}} - a_{\text{InAs}} }}} \right), $$
(2)

where L is the thickness of one SL period.

The next approach suggested that the “strain-balanced” structure could be reached by assuming equal thickness for the tensile and compressive layers. It could be assumed that the structure is balanced if the following condition is met [15,16,17,18,19]:

$$ \epsilon_{xx}^{\text{InAs}} \cdot L_{\text{InAs}} + \epsilon_{xx}^{\text{InAsSb}} \cdot L_{\text{InAsSb}} = 0, $$
(3)

\( \epsilon \) is the stresses related to the buffer layer.

In order to express the strain, the following equation was used:

$$ L_{\text{InAsSb}} = L \cdot \left( {\frac{{a_{\text{InAsSb}} \cdot (a_{\text{GaSb}} \cdot a_{\text{InAs}} )}}{{a_{\text{GaSb}} \cdot (a_{\text{InAsSb}} \cdot a_{\text{InAs}} )}}} \right). $$
(4)

Comparison of these two equations suggested that the discrepancy in the thickness of the InAsSb layer was less than 5%, not causing significant changes in the devices performance. The sample satisfaction of the condition “strain-balanced” is marked by in the Table 1.

Theoretical simulation of the SLs optical characteristics

The commercial APSYS platform allowed to study the SLs optical characteristics. The 4 bands method (kp 8×8) with periodic boundary was used, which is well described in the papers [20,21,22]. In addition, to obtain symmetric wave functions Ψe(h), the Kane parameter, F = 0 was assumed [22]. The standard value of k = 0.06 was used in simulations with the exception for the determination of the effective masses, for which the smaller interval of the vector k = 0.03 was applied. Up to date, many different values of the parameters required for the band structure theoretical modeling have been used. The parameters implemented in our calculations are presented in “Appendix” in Table 3 [20, 23,24,25,26,27,28,29,30,31,32].

The strain could also influence the band structure of the material. In our calculations, the solid-theory model was used [23, 24]. Figure 3 shows the SLs InAs/InAs1−xSbx band structure for unstrained and strained conditions. Electron affinity for unstrained materials InAs and InSb is presented in Table 3. Under the influence of strain, an electron affinity changed and the valence band splitted on the heavy and light holes subbands.

Figure 3
figure 3

The energy band profiles of unstrained and strained SLs

The set of dispersion curves E = f (k, q), allowing to estimate electron and hole effective masses (heavy and light holes) for the SL, was simulated. The effective masses were consistent with those presented in [21]. Figure 4 shows the absorption coefficient (α) for three samples: Sample-A; Sample-B, and Sample-D with selected absorber thicknesses and xSb.

Figure 4
figure 4

The simulated absorption coefficient versus photon energy at T = 230 K

The presented “strain-balanced” structures, Sample-A; Sample-B; and Sample-D, were optimized for 12 μm, 18 μm, and 7 μm cutoff wavelengths, respectively. In comparison with the MWIR devices where α ≈ 6000 cm−1, the LWIR detector absorption coefficient was less than 1500 cm−1. The sharp change in the absorption characteristics (e.g., Sample-D) was caused by two transitions: c1hh1 and c1lh1. This was presented in more detail in [21].

The correct determination of the SL energy gap could be reached by a derivative of the absorption coefficient which was shown for Sample-C at T = 300 K in Fig. 5. In this case, a derivative of the absorption coefficient on energy (this graph is shown upper left) had to be calculated in order to determine the maximum value.

Figure 5
figure 5

The simulated absorption coefficient at T = 300 K for Sample-C (method for the determining of the bandgap)

Figure 6 represents a theoretical simulation of the absorption coefficients for Sample-C at selected temperatures 120 K, 180 K, 230 K, and 300 K. It can be seen that when temperature increased, the width of the bandgap decreased and the cutoff wavelength increased, correspondingly. Also, the maximum value of the absorption coefficient decreased with temperature.

Figure 6
figure 6

The simulated absorption coefficient versus temperature for Sample-C

In order to determine the T2SLs electron affinity χ (position of the conductivity band Ec), the band structure of one of the SL samples with period L = 18.3 nm and selected thicknesses of the InAs and InAs1−xSbx layers were simulated. Since APSYS platform gives only the relevant values of Ec, the simulated value was adjusted to that extracted from the linear relationship for the thickness LInAsSb of the layer InAs1−xSbx corresponding to the “strain-balanced” structure.

Figure 7 shows the linear dependence of the SL electron affinity χ (red line) versus the thickness of the InAsSb layer and the theoretical simulation (black curve) intersecting at the point LInAsSb = 3.95 nm corresponding to the balanced SL (see eq. 2). The different slope of these curves could be explained by additional strains between the layers when their thickness deviates from the thickness in the “strain-balanced” structure. For thicker InAsSb layers, the electron affinity was reduced, while for thinner ones it was higher than the linear averaging. Even though these changes were low, a 20% deviation from the “strain-balanced” thickness resulted in a reduction in electron affinity by 0.002 eV.

Figure 7
figure 7

The theoretical simulation of the SL InAs/InAs0.59Sb0.41 electron affinity for Sample-B (red line—linear averaging; black line—theoretical simulations)

Results

In order to compare the simulated absorption curves (Sample-C at temperature T = 220 K) with the experimental data, two characteristics, photoluminescence (PL) and absorption coefficient, were compared (Fig. 8). This figure illustrates the difference between the energy corresponding to 100% α cutoff and the energy corresponding to 50% α cutoff, being mostly determined. It is shown that energy corresponding to the 50% PL maximum correlated with the width of the bandgap (the definition was described above), but the maximum of the PL curve agreed with the half of the absorption coefficient value allowing to draw the conclusion that the bandgap energy extracted from PL should be determined very carefully.

Figure 8
figure 8

The simulated α and PL experimental data for Sample-C at T = 220 K

We determined bandgap based on the calculated absorption curve and PL measurements using the method specified above. Figure 9 shows the temperature dependence of the energy gap for Sample-C. The green and blue lines correspond to the theoretical calculations, while the purple squares correspond to the experimental results (reading PL data from the maximum and from the 50% cutoff). It is worth mentioning that a proper coincidence between theory and experiment was reached by accepting the temperature dependence of almost all the necessary parameters for theoretical simulations.

Figure 9
figure 9

The simulated and measured bandgap versus temperature for Sample-C

Table 2 presents a comparison of simulated and measured bandgap energy for the samples described in Table 1.

Table 2 The theoretical and experimental data for Eg and CBO at T = 230 K

The detector’s performance was fully confirmed by our calculations. Particularly interesting were the results for photoconductors made of Sample-B optimized for VLWIR range. After the growth, standard photolithography technique with wet-etching was used to define the active (75 × 85 µm) and contact areas of the IR photoconductors. The vacuum evaporation of Au/Ti electroplating was applied to fabricate ohmic contacts. At the end of the processing, the monolithic GaAs immersion lenses were formed to increase optical area of the detectors [13].

Figure 10 shows the relative responsivity characteristics (Ri) normalized to unity depending on the wavelength at the temperatures 210 K, 230 K, and 300 K. The Ri temperature dependence in arbitrary units was presented to show the cutoff wavelength changes versus temperature. The 50% cutoff could be then estimated from this figure. For example, at the temperature T = 230 K cutoff wavelength was λCO ≈ 16 μm (corresponding to energy of 0.0775 eV). Comparing these data with the previous results (estimated convergence of the energy gap by photoluminescence measurement and theoretical simulations), we saw a fairly high agreement of the theory and experimental data. At 300 K, c–1h1 transitions were more decisive than chh1, whereas at low temperatures the situation was reverse. Comparison of the simulated absorption coefficient (see Fig. 4, Sample-B at T = 230 K) and the experimentally measured Ri (Fig. 10) gave a good bandgap match.

Figure 10
figure 10

The normalized responsivity characteristics (experimental data) at temperatures: 210 K, 230 K and 300 K for Sample-B

Figure 11 presents the detectivity (D*) of the two samples (Sample-A and Sample-B). It should be noted that the D*, as well as the absorption coefficient, depend on the spectral range of the detectors. At T = 230 K, D* reached ~ 8 × 109 cm Hz1/2W−1 (Sample-A) and ~ 9 × 108 cm Hz1/2W−1 (Sample-B) for V = −0.5 V. Those experimental data were similar to the one presented in [6].

Figure 11
figure 11

The experimental data D* characteristics at T = 230 K for the Sample-A and Sample-B

Conclusions

This work presents the characterization of the T2SL absorber made of the InAs and InAs1−xSbx materials on GaAs substrate with GaSb buffer layer. In our simulations, two options were considered: when the SL layer thickness corresponded to the “strain-balanced” condition and when this requirement was not met. By the “strain-balance,” we meant to achieve zero average in-plane stress in the tensile/compressively strained layer of the SL and of the buffer layer, such that no shear force was generated at the interfaces between layers of the SL. By producing the SL structures that were “strain-balanced” to the buffer layer, we eliminated the possibility of the strain and dislocations occurrence suppressing the device performance. The experimental data’s confirmed our theoretical simulations. The proposed method of determining the energy gap from the composition of the PL and absorption characteristic gave results consistent with the data from the responsivity detectors characteristic.