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

Analysis of threshold current of uniaxially tensile stressed bulk Ge and Ge/SiGe quantum well lasers

Open Access Open Access

Abstract

We propose and design uniaxially tensile stressed bulk Ge and Ge/SiGe quantum well lasers with the stress along <100> direction. The micro-bridge structure is adapted for introducing uniaxial stress in Ge/SiGe quantum well. To enhance the fabrication tolerance, full-etched circular gratings with high reflectivity bandwidths of ~500 nm are deployed in laser cavities. We compare and analyze the density of state, the number of states between Γ- and L-points, the carrier injection efficiency, and the threshold current density for the uniaxially tensile stressed bulk Ge and Ge/SiGe quantum well lasers. Simulation results show that the threshold current density of the Ge/SiGe quantum well laser is much higher than that of the bulk Ge laser, even combined with high uniaxial tensile stress owing to the larger number of states between Γ- and L- points and extremely low carrier injection efficiency. Electrical transport simulation reveals that the reduced effective mass of the hole and the small conduction band offset cause the low carrier injection efficiency of the Ge/SiGe quantum well laser. Our theoretical results imply that unlike III-V material, uniaxially tensile stressed bulk Ge outperforms a Ge/SiGe quantum well with the same strain level and is a promising approach for Si-compatible light sources.

© 2017 Optical Society of America

Corrections

27 October 2017: A typographical correction was made to the author affiliations.

1. Introduction

The Si-based light source has long been considered to be the “holy grail” of Silicon Photonics. Many efforts have been made to achieve light sources on Si platform, such as Si nanocrystals [1], erbium-doped Si [2], band-engineered Ge [3], GeSn alloy [4], Ge Zener-Emitter [5] and III-V material on Si [6]. Among those techniques, the band-engineered Ge has attracted great interest due to its Si-compatible fabrication process and ability to be electrically driven. The idea of band-engineered Ge was originally proposed by a MIT group [3], conceiving that the energy difference between the Γ and L points can be reduced by introducing tensile strain and the remaining electronic states can be further filled by extrinsic electrons from n-type doping so that more injected electrons can occupy the Γ valley where efficient radiative recombination occurs. Subsequently, the optically pumped [7] and electrically driven [8] Ge lasers have been demonstrated but with unsatisfying threshold current density as high as 280 kA/cm2. In order to lower the threshold current density of Ge laser, the tensile strain needs to be enhanced and therefore a variety of strain techniques have been explored including SiNx stressor [9] and suspended microstructures [10, 11]. The highest reported tensile strain values, to the best of our knowledge, are 4.9% [12] and 1.9% [11] in the case of uniaxially and biaxially tensile stressed, respectively.

On the other hand, it is widely accepted that the lasers using low dimensional quantum structures as gain media have lower threshold current density due to the modified density of state [13] and scaling law [14]. Indeed, the III-V lasers with quantum structures have achieved desirable performance, motivating a series of theoretical [15, 16] and experimental investigations [17–20] of the Ge/Si1-xGex quantum structure. Compared with quantum wire and quantum dot, quantum well features easy epitaxial process and can be combined with tensile stress. It has been discussed in [21] that the unstrained Ge/Si1-xGex quantum well underperforms bulk Ge since the quantum confinement of Γ valley is stronger than that of L valley, making the energy difference between Γ- and L-points even larger. Recently, uniaxial tensile stress and biaxial tensile stress have been realized in Ge/Si0.19Ge0.81 quantum well [22], signifying that it is possible to introduce high tensile strain in Ge/Si1-xGex quantum well. Therefore, a comparison of threshold currents of lasers based on tensile strained bulk Ge and Ge/Si1-xGex quantum well is in demand. We calculate the optical gain of uniaxially tensile stressed Ge/Si1-xGex quantum well in [23]. Nevertheless, a laser structure based on uniaxially tensile stressed Ge/Si1-xGex quantum well has rarely been proposed. In [24], we propose a Distributed Bragg Reflector (DBR) bulk Ge laser using the micro-bridge structure to introduce high uniaxial stress. The DBR laser has a low threshold current since it combines high tensile stressed Ge, low loss optical cavity and quasi-heterojunction. Similar structure can be utilized to obtain a uniaxially tensile stressed Ge/Si1-xGex quantum well laser.

In this paper, we analyze and compare the threshold current densities of DBR lasers based on <100> uniaxially tensile stressed bulk Ge and Ge/Si1-xGex quantum well growing along the <001> direction. The effects of strain and quantum confinement on the densities of state of L valley and carrier injection efficiency are analyzed. Differing from the laser structure in [24], the new design adopts a full etched ultra-broadband circular Bragg grating to enable larger fabrication tolerance of the gratings. Besides, a lateral p+-n-n+ junction is designed for the bulk Ge laser to obtain longer optical cavity so that the threshold current can be reduced. Through band structure calculation, optical and electrical transport simulations, we show that the threshold current density of the DBR laser based on the uniaxially tensile stressed Ge/Si1-xGex quantum well is about two orders of magnitude higher than that of its bulk Ge counterpart. Therefore, we suggest that bulk Ge rather than Ge/SiGe quantum well should be used as the gain medium for Ge laser.

2. Design of the DBR lasers

2.1 Micro-bridge for quantum well and quantum well material stack

Tensile strained Ge as the gain medium is indispensable for a low threshold Ge laser. SiNx stressor [9] and suspended micro-structures [10, 11] are the two main techniques to introduce tensile strain in Ge, which can also be applied to Ge/Si1-xGex quantum well. Compared with SiNx stressor technique, the micro-bridge structure can incorporate low loss optical cavity and p-n junction for carrier injection, which is demonstrated in our previous work [24]. Therefore, we adopt the micro-bridge structure to introduce tensile strain. The high tensile strain is obtained by redistributing the thermal mismatch tensile strain in Ge-on-Si or Germanium on Insulator (GeOI) wafer using the suspended micro-bridge structure. The bulk Ge micro-bridge has been widely investigated while the Ge/Si1-xGex quantum well micro-bridge is less straightforward but can be conceived analogously. In the annealing process of SiGe buffer, thermal expansion coefficients mismatch between SiGe alloy and Si also gives rise to residual tensile strain which depends on the annealing temperature and Ge composition of the buffer [25]. Therefore, the buffer layer can be used as a stressor layer. Figure 1(a) shows the micro-bridge structure for introducing tensile strain in Ge/SiGe quantum well. Note that the upper quantum well layer needs to be etched to form a ridge waveguide in order to enhance the tensile strain in the quantum well and confine the light. The micro-bridge is along the x-direction, i.e. the <100> crystal orientation. The thicknesses of buffer layer Hbuffer and multiple quantum wells (MQWs) layer Hr are 300 nm and 450 nm, respectively. The width and length of the narrow bridge are 2 μm and 10 μm, respectively. The total length of the bridge is 740 μm and the taper angle θ is 30°. The width of the ridge is 600 nm. 3-D Finite Element Method (FEM) simulation is performed using COMSOL Multiphysics software to obtain the strain distribution of the Ge/Si1-xGex quantum well micro-bridge. In the simulation, linear elastic deformation is assumed. The initial biaxial stress conditions in the buffer layer are τxx=τyy=(C11+C122C122/C11)εxx and τzz=0 with εxx=0.18%. Fixed constraint is used for the bottom of the structure and symmetry condition is used in the mirror planes of the structure to reduce computational complexity. Other boundaries are considered to be free. Generally, strain-balanced quantum well stack is utilized to achieve high crystalline quality when there exists lattice mismatch between the quantum well and barrier. In the case of Ge/Si1-xGex strain-balanced quantum well, the SiGe quantum barrier and Ge quantum well are tensile and compressive strained, respectively. To simplify the simulation of strain redistribution process, we assume that the effects of well and barrier counteract due to the strain balance mechanism. Namely, only the lower SiGe buffer layer is slightly biaxial tensile strained while the material above the buffer layer is considered to be unstrained. It should be noted that this assumption is only used in the FEM simulation. For the subsequent optical gain and electrical simulation, the strain of micro-bridge needs to be modified to obtain the exact strain of the well and barrier. Before the micro-bridge is fabricated, the in-plane lattice constants of the well and barrier are equal to that of the buffer layer which is abuf(1+εbi) if the thermal mismatch strain εbi is taken into account. After the micro-bridge forms, tensile stress is introduced in the narrow bridge and the lattice constant along the x-direction becomes abuf(1+εxxbri), where εxxbri is the strain component of the micro-bridge, directly obtained from the FEM simulation. Therefore, the εxx strain component of the well and barrier can be expressed as

εxx=abufa(1+εxxbri)1
where abuf is the unstrained lattice constant of the buffer layer. a is the lattice constant of the well or barrier. Since the total strain contains the lattice mismatch strain and the strain induced by uniaxial stress, εyy and εzz can be divided into two parts and are described by
εyy=abufa(1+εbi)1C12C11+C12(εxxbriεbi)
εzz=2C12C11(abufa(1+εbi)1)C12C11+C12(εxxbriεbi)
where C11 and C12 are the elastic constants.

 figure: Fig. 1

Fig. 1 (a) Schematic diagram of the mirco-bridge structure for introducing uniaxial stress in Ge/SiGe quantum well (not to scale). (b) Distribution of the strain component εxxbri calculated by 3-D FEM, when Hbuffer=300nm, Hr=450nm, Wr=600nm, θ=30°, L1=10μm, L2=740μm and the width of narrow bridge is 2μm.

Download Full Size | PDF

Figure 1(b) shows the strain component εxxbri distribution of the micro-bridge simulated by 3-D FEM. As can be seen, both the buffer layer and MQWs layer are uniaxially tensile stressed with εxxbri=4.4%. Moreover, εxxbri reaches its maximum value 7% at the corner of the taper, which is still below the breakdown threshold [26]. It is worth notice that the achievable strain value depends on the crystalline quality of the material. Therefore, it is beneficial to use the SiGe on insulator (SiGeOI) wafer which is similar to the GOI wafer [27] so that high uniaxially tensile stressed quantum well can be attained.

The strain-balanced quantum well is designed using an elastic energy minimization method [28]. First, the average lattice constant of the MQWs region is calculated by

a||MQW=(abGbhb+awGwhw)/(Gbhb+Gwhw)
G=2(C11+2C12)(1C12/C11)
where hb and hw are the thicknesses of barrier and well, respectively. For a pseudomorphic interface, the lattice constants in the plane a|| remain the same over the structure. In other words, the in-plane lattice of the buffer layer should match that of the MQWs region. In this way, Ge composition of the buffer layer can be determined. Note that the residual thermal strain should be included in the calculation of the lattice constant of SiGe buffer. The unstrained lattice constant of Si1-xGex alloy is fitted from the experimental data [29].
aSiGe=0.0282x2+0.1981x+5.4315
where aSiGe is in the unit of A°. For a laser diode, type-I quantum well and large band offset are required for strong quantum confinement and efficient carrier injection. The conduction band minima of unstrained Ge and SiGe alloy with high Ge composition are located at L-points. Nevertheless, tensile strain shifts the band edges of the quantum well and barrier and hence the optimal Ge composition of the barrier is changed. According to the model solid theory and deformation potential theory [28], we calculate the band edges of Γ-, L- and Δ-valleys and then estimate the optimal Ge composition of barrier for maximum conduction band offsets under various strain values. The energy shifts of valley i are described by [28]
ΔEci=[Ξd1+Ξu{a^ia^i}]:ε
where Ξd and Ξu are the dilatation and uniaxial deformation potential, respectively. 1 is the unit tensor and a^i is the unit vector parallel to the k vector of valley i. ε is the strain tensor. {} denotes dyadic product. The energy shift of Γ-valley is

ΔEcΓ=ac(εxx+εyy+εzz)

There are eight equivalent L-valleys and six equivalent Δ-valleys for unstrained Ge and SiGe. Uniaxial tensile stress along [100] direction breaks the degeneracy of Δ-valleys, while the L valleys remain degenerate. For L valleys, the energy shifts are

ΔEcL=aL(εxx+εyy+εzz)

For Δ-valleys along the [100] and [1¯00] directions, the energy shifts are expressed as

ΔEcΔ2=ΞdΔ(εxx+εyy+εzz)+ΞuΔεxx

For Δ-valleys along the [010], [001], [01¯0] and [001¯] directions, the energy shifts are

ΔEcΔ4=ΞdΔ(εxx+εyy+εzz)+ΞuΔεyy
The band offsets of Ge and SiGe alloy interface are determined using the model solid theory [28]. First, the unstrained band edges of conduction and valence band are determined by
Ev=Ev.av+Δso/3
Eci=Ev+Egi
where Ev.av and Δso are the average energy of valence band and spin-orbit split off energy, respectively. Egi is the bandgap of the i conduction band. The bandgap of Γ-valley for Si1-xGex alloy at 300 K is [16]
EgΓ=0.7985x+4.185(1x)0.14x(1x)
The bandgaps of L- and Δ-valleys for Si1-xGex alloy at 300 K are fitted from the data of Braunstein [30]

EgL=1.861.2x
EgΔ=1.0870.487x+0.264x2

Then the conduction band edges and band offsets under strain are evaluated using Eqs. (7)-(11). The valence band edges are obtain by the 8-Band k∙p method. The material and band structure parameters of Ge and Si at 300 K are presented in Table 1. The parameters for SiGe alloy are calculated by linear interpolation. Figure 2(a) shows the optimal Ge compositions of barrier for maximum conduction band offsets at different strain values. It can be seen that the optimal Ge composition of barrier rises from 0.847 to 0.926 with the increase of εxxwell from 0 to 5%. Figure 2(b) depicts the band alignment of the Ge/Si0.09Ge0.91 which is the optimal configuration for εxxwell=4%. The conduction band minima of uniaxially tensile stressed Si0.09Ge0.91 are located at L points. Meanwhile, the Δ-points along the <100> and <1¯00> approximately overlap the L points due to the strain splitting effects. Therefore, choosing Ge composition lower than 0.91 will downshift the Δ-points, leading to the decrease of the conduction band offset.

Tables Icon

Table 1. Material and band structure parameters at 300 K for Ge and Sia

 figure: Fig. 2

Fig. 2 (a) Optimal Ge composition of the barrier as a function of the strain component εxxwell. (b) Band alignment of the Ge/Si0.09Ge0.91 quantum well with εxxwell=4%.

Download Full Size | PDF

After obtaining the Ge composition of barrier, the SiGe buffer layer should be designed to make the quantum well stack under a strain balance status. Since we focus on the uniaxially tensile stressed quantum well with εxxwell=4%, Si0.09Ge0.91 is chosen to be the barrier. The widths of well and barrier are 17 nm and 21 nm, respectively. 0.18% biaxial tensile strain induced by thermal mismatch is considered in the buffer layer. Using Eqs. (4)-(6), the corresponding Ge composition of the buffer layer is calculated to be 0.91.

2.2 Optical cavity and optical gain spectra

In our previous work, we have designed a bulk Ge DBR laser utilizing the micro-bridge structure to introduce high uniaxial stress. Simulation results show that the laser has a low threshold current thanks to the high uniaxial stress, low loss optical cavity and quasi-heterojunction. However, 3-D Finite Different Time Domain (FDTD) simulation indicates that a 5 nm deviation of the grating period can result in a 20 nm wavelength shift of the reflection peak, leading to the misalignment of the two reflection peaks. Therefore, the original design requires high fabrication accuracy. To relax the fabrication requirement, we propose a new optical cavity design which involves a full-etched circular grating with high reflectivity over ultra-broad band and a normal waveguide grating as the output mirror, as depicted in Figs. 3(a) and 3(b). Figure 3(c) illustrates the ridge waveguide used in the lasers. For the bulk Ge laser, the waveguide only supports fundamental TE-mode. For the quantum well laser, owing to the restriction of the waveguide dimensions, fundamental TE- and TM-mode are supported in the waveguide. The single-mode operation and two-mode operation of the waveguides indicate that the photons are well confined in the active region for both bulk Ge and Ge/SiGe quantum well lasers. The ultra-broad band high reflectivity of the circular grating results from the considerable refractive index contrast between the air gap and grating blade. However, the air gap can evoke intense diffraction loss if deployed in a normal straight waveguide. In order to reduce the diffraction loss, Wang [32] has proposed and demonstrated a circular grating where the wave front of the diffraction wave matches the grating blade. We adopt a similar design with a different two-section taper region, as shown in Fig. 3(d). The first taper with small taper angle acts as a transition region that can reduce the refractive index difference between the straight waveguide and the second taper. The second taper with large angle serves as a diffraction region that allows the wave to form a circular wave front. 3-D FDTD simulation signifies that there will exist ripples in the reflectance spectrum without the first taper. It should be noted that the first trench of the circular grating is designed to be narrower to avoid intense vertical diffraction loss. For the bulk Ge laser, the sidewall grating can be used as the output mirror on the other side of the optical cavity in consideration of its easy fabrication process. However, for the Ge/SiGe quantum well laser, sidewall grating fails to realize reflectivity higher than 90%. Therefore, surface grating is employed as the output mirror of the quantum well laser.

 figure: Fig. 3

Fig. 3 Schematic diagrams of the uniaxially tensile stressed (a) bulk Ge laser and (b) Ge/SiGe quantum well laser. (c) Cross section view of the straight waveguide. For bulk Ge laser, Wr=600nm, H1=100nm and H2=200nm. For Ge/SiGe quanum well laser, Wr=600nm, H1=450nm and H2=300nm. (d) Top view of the circular grating. Reflection coefficient spectra of the fundamental TE-mode for the gratings of (e) bulk Ge laser and (f) Ge/SiGe quantum well laser.

Download Full Size | PDF

3-D FDTD simulation is performed to simulate the reflection spectra and optimize the structure parameters. For the bulk Ge laser, the ridge width Wr, the ridge height H1 and the total height of the ridge waveguide are 600 nm, 100 nm and 300 nm, respectively. The taper angles θ1 and θ2 are chosen to be 20° and 70°, respectively. The lengths of the first taper and the second taper are 3 μm and 4 μm, respectively. The width of the first trench is 100 nm. The widths of the other trenches and blades are 200 nm and 318 nm. The number of the trenches is 10. The etched width of the sidewall grating is 100 nm. The grating period is 420 nm with a duty cycle of 50% and a total number of 120. For the quantum well laser, the dimensions of the waveguide are the same as those used in the strain simulation. At the same time, the second taper angle changes to 60°. The widths of the other trenches and blades are both 250 nm. The etched depth of the surface grating is 200 nm with a grating period of 327 nm. The duty cycle is 50% and the number of the periods is 350.

Figures 3(e) and 3(f) show the reflection coefficient spectra of the fundamental TE-mode for the gratings of bulk Ge laser and quantum well laser, respectively. As can be seen, the bandwidths of the circular gratings of bulk Ge laser and quantum well laser with reflection coefficients above 95% are ~500 nm, which allows for relatively large fabrication tolerance. The peak reflection coefficients of the sidewall grating and the surface grating are 97% and 95%, located at 2524 nm and 2294 nm, respectively. The reflection peaks of the output gratings are determined by estimating the net gain peak positions of the gain media when the lasers are near threshold. Models based on 8-Band k∙p method are employed to calculate the net gain spectra of the uniaxially tensile stressed bulk Ge and Ge/SiGe quantum well, which can be found in our previous works [23, 24]. Figures 4(a) and 4(b) depict the net optical gain coefficient spectra of the gain media of the bulk Ge laser and the Ge/Si0.09Ge0.91 quantum well laser when the lasers perform at threshold. As can be seen, the peak gains of the TE-polarized light for the bulk Ge and Ge/SiGe quantum well are 180 cm−1 and 600 cm−1, located at 2524 nm and 2294 nm, respectively, which correspond to the reflection peaks of the output gratings. Furthermore, simulation results indicate that the optical gain of TE-polarized light is higher than that of the TM-polarized light. As a result, only TE-mode can lase in the quantum well laser even though the ridge waveguide supports two modes.

 figure: Fig. 4

Fig. 4 (a) Net gain coefficient spectrum of the TE-polarized light for bulk Ge with a strain of 4%, a doping concentration of 7×1018cm3 and an injected carrier density of 2.3×1018cm3. (b) Net gain coefficient spectra of the undoped Ge/Si0.09Ge0.91 quantum well with a strain of 4% and an injected carrier density of 2.1×1019cm3.

Download Full Size | PDF

2.3 Electrical structure

Electrical structure for carrier injection is essential for an electrically driven laser. For the bulk Ge micro-bridge structure, the quasi-heterojunctions along the bridge utilizing the band offsets induced by strain distribution has been proposed [24]. However, the longitudinal heterojunction design limits the length of the optical cavity since the carrier densities experience an attenuation along the optical cavity. Therefore, longer optical cavity for lower threshold current cannot be deployed. To address this concern, a lateral Ge homo-junction is designed, which is shown in Fig. 5(a). To enhance the confinement of the carriers, the doping concentrations of the p+-region, n-region and n+-region are 1.5×1019cm3, 7×1018cm3 and 1.5×1019cm3, respectively. Figure 5(b) shows the p+-i-n+ junction of the Ge/SiGe quantum well laser. Since we target at the uniaxially tensile stressed quantum well with εxxwell=4%, the structure has a 400nm-thick Si0.09Ge0.91 buffer layer, followed by 5 pairs of Ge/Si0.09Ge0.91 quantum wells and a 140nm-thick n-doped Si0.09Ge0.91 layer. The doping concentrations of the p+- and n+-region are 1.5×1019cm3. The MQWs region and part of the buffer layer are undoped in order to reduce the free carrier absorption (FCA). It should be pointed out that only the narrow bridge region is doped for both the bulk Ge and quantum well laser to avoid the FCA loss in the gratings and passive waveguide.

 figure: Fig. 5

Fig. 5 Cross section of (a) the p+-n-n+ junction of bulk Ge laser and (b) the p+-i-n+ junction of Ge/Si0.09Ge0.91 quantum well laser. Electron concentration profiles of the (c) p+-n-n+ junction and (d) p+-i-n+ junction at a bias voltage of 0.6V.

Download Full Size | PDF

Electrical transport simulation is performed using drift-diffusion model. The parameters of the strained material, such as bandgap, electron affinity, mobility and effective mass are determined using the method described in [24]. Figures 5(c) and 5(d) depict the calculated electron concentration profiles of the lateral p+-n-n+ junction and the vertical p+-i-n+ junction under a bias voltage of 0.6 V, respectively. For the bulk Ge laser, the active region has a high electron concentration varying from 1.1×1019cm3 to 1.2×1019cm3 with a homogeneous distribution along the z direction. For the quantum well laser, the electrons are confined in the quantum well with a concentration of 6×1018cm3. These simulation results indicate that the designed electrical structures for the bulk Ge and Ge/SiGe quantum well laser can effectively inject carriers into the active region where radiative recombination mainly occurs.

3. Analysis of densities of state, carrier injection efficiencies and threshold current densities

The success of III-V quantum well in low threshold laser is the main drive for the investigation of Ge/SiGe quantum well. Theoretically, there are two widely accepted reasons for the low threshold of quantum well laser. First, owing to the step-like density of state, the injected carrier density of the quantum well is lower than that of the conventional bulk material when the peak gain of them are the same, making contribution to reducing the threshold current [13]. On the other hand, the scaling laws can also explain why lasers with quantum structures have lower threshold current. The scaling equation for the quantum well laser is described below [14]

Ith(ηγτe)=N2Dtr(WL)+Wdg2DlnR1+WLdg2Dαscat
where η is the carrier injection efficiency. γ=1b/g2D, b is the coefficient for FCA withαfca=bN2D/Lz and g2D is the areal density of state. N2D is the areal density of carrier, Lz is the width of quantum well. τ is the carrier lifetime. N2Dtr is the transparency carrier density. W, L and d are the width, length and thickness of the active region, respectively. Assuming the facet reflectivity R approaches 1 and the scattering loss αscat can be neglected, the minimum threshold current density can be expressed as Jth=N2Dthe/ηγτ. As a consequence, the scaling law predicts a threshold current density as low as 60 A/cm2 for the III-V quantum well [14]. However, as far as the GeSi material system is concerned, those two mechanisms should be treated differently. Firstly, it has been demonstrated for untrained Ge/SiGe quantum well that the energy difference between the Γ- and L-points become larger than that of bulk Ge due to the stronger quantum confinement effect of Γ-valley [21]. Consequently, more electrons are needed to fill the states between the Γ- and L-points which raises the threshold current. To obtain better understanding and extend to the case of strained quantum well, we compare the density of state of L-valley and the number of states between the Γ- and L-points of the strained Ge/SiGe quantum well with their bulk Ge counterpart. Since the effective mass approximation is adopted to calculate the band structure of L-valleys, the strain is considered to have little effect on the density of state of L-valley. The volumetric density of state of L-valley for the quantum well is calculated numerically by
g3DQW=2π2LzΔEnLHeaviside(Ec+ΔEEnLL(kt))dkxdky
where Heaviside() is the unit step function. Ec and ΔE are the band edge of L-valley and the energy range, respectively. Note that the area density of state is divided by the length of quantum well Lz to obtain the volumetric density of state which can be compared with its bulk Ge counterpart. For the bulk Ge, the average density of state of L-valleys in the energy range of ΔE can be expressed as g3DBulk=8π(2mn*)3/2ΔE/3h3, where mn*=0.55 is the density of state effective mass. When the energy range ΔE is chosen to be kBT, g3DBulk is calculated to be 2.97×1020cm3eV1 while g3DQW is calculated to be 3.6×1020cm3eV1 which is larger than g3DBulk, indicating that the Ge/SiGe quantum well has no advantage over bulk Ge in terms of the density of state. Figure 6 shows the number of states between the Γ- and L-points for the bulk Ge and Ge/SiGe quantum well under various tensile strain. The Ge compositions of barrier are chosen to be the optimized values that maximize the conduction band offsets. As can be seen, bulk Ge has less states between the Γ- and L-points than Ge/SiGe quantum well when εxxwell>0.5%. Even though Ge/GeSi quantum well with εxxwell=0 has less states to fill than unstrained bulk Ge, positive net gain cannot be achieved due to the large number of states between the Γ- and L-points.

 figure: Fig. 6

Fig. 6 Number of states between the Γ- and L- points for bulk Ge and Ge/SiGe quantum well as a function of the strain componentεxxwell.

Download Full Size | PDF

Another factor that should be carefully treated is the carrier injection efficiency in the scaling equation. A higher carrier injection efficiency reflects a better confinement of the injected carrier. Most of the theoretical investigations assume a high carrier injection efficiency, indicating an ideal carrier confinement [14, 21, 33]. However, this is not true in the case of Ge/SiGe quantum well. Electrical transport simulation using drift-diffusion model is performed to obtain the relationship between the current density J and carrier density N. The carrier injection efficiency is derived by η=eLz(AN+BN2+CN3)/J, where A=1/τSRH is the Shockley–Reed–Hall (SRH) recombination coefficient. B is the radiative recombination coefficient and C is the Auger recombination coefficient. The defect limited SRH life time is set to 3 ns according to the experimental data of a GOI sample [34]. The radiative recombination coefficient for uniaxially tensile stressed Ge/SiGe quantum well with εxxwell=4% is calculated to be 3.3×1013cm3s1. The Auger recombination coefficient C of tensile strained Ge is chosen to be 1.6×10-30cm6s1 [35]. It should be noted that the strain effects on the parameters used in the electrical simulation should be taken into account. The models for the strain-dependent parameters are described in our previous work [24]. A uniaxial tensile stress with εxxwell=4% is considered in our simulation. Figure 7(a) shows the calculated injected carrier density as a function of the current density for the strained quantum well p+-i-n+ junction. The relationship between the carrier injection efficiency and injected carrier density is shown by the red curve in Fig. 7(b). It can be seen that the carrier injection efficiency of the strained quantum well p+-i-n+ junction is below 1%, indicating that the carrier confinement is weak. Moreover, the injection efficiency deteriorates with the increase of injected carrier density. This is because the quasi-Fermi level of electron raises with the increase of carrier density, making it easier for the electrons to escape from the quantum well. However, an injected carrier density of ~1×1019cm3 is required for the quantum well to produce net gain. At such high injection level, the carrier injection efficiency is less than 103. To get insight into the cause of the low carrier injection efficiency, we change the parameters of the electrical simulation and find that the effective mass and band offset are two crucial factors. According to the 8-Band k∙p band structure calculation, the effective mass of hole for Ge decreases from 0.33 to 0.16. If an unstrained effective mass is used in the simulation, the carrier injection efficiency turns out to be higher especially in the high injection level, as shown by the blue curve in Fig. 7(b). On the other hand, the small conduction band offset also results in a low carrier injection efficiency. In the simulation for comparison, the conduction band offset is set to 0.15 eV instead of the original value 0.0284 eV. The simulation results, shown by the purple curve in Fig. 7(b), signify that the carrier injection efficiency is improved by two orders of magnitude. To summarize, the actual carrier injection efficiency for the Ge/SiGe quantum well is several orders of magnitude lower than the ideal 100% which is assumed in the scaling law and a number of theoretical models. As a result, the threshold current of Ge/SiGe quantum well laser can be much higher than envisioned.

 figure: Fig. 7

Fig. 7 (a) Injected carrier density as a function of current density for the p+-i-n+ junction of Ge/Si0.09Ge0.91 quantum well laser. (b) Carrier injection efficiency as a function of injected carrier density for the p+-i-n+ junction of Ge/Si0.09Ge0.91 quantum well laser. The red curve and blue curve represent simulation results using strained effective mass and unstrained effective mass, respectively. The purple curve represents simulation results using unstrained effective mass as well as a larger conduction band offset of 0.15 eV.

Download Full Size | PDF

Finally, we compare the threshold current densities of uniaxially tensile stressed bulk Ge laser and Ge/SiGe quantum well laser. The threshold conditions are expressed below.

(ΓactgbulkΓpαpΓnαn)Lg=ln(1R1R2)
(ΓQWgQWΓbufferαbufferΓbarrierαbarrierΓnαn)Lg=ln(1R1R2)
where g and α represent the optical gain and FCA loss, respectively. The overlap integration between the optical field and the active region of the bulk Ge laser Γact is 75%, calculated by the Finite Difference Method (FDM). For the quantum well laser, the overlap integration ΓQW is 8.5%. Lg denotes the length of the gain region rather than the total length of the optical cavity and is 10 μm. The reflection coefficients of the circular mirrors and output mirrors are 97% and 95%, respectively. The calculation procedure of the FCA loss can be found in our previous work [23, 24]. The threshold gain for the bulk Ge laser is calculated to be 180 cm−1. Using the optical gain model described in [24], the threshold carrier density for uniaxially tensile stressed bulk Ge with εxx=4% and a n-doping concentration of 7×1018cm3 is 2.3×1018cm3. Utilizing the relationship of carrier density and current density obtained by the electrical transport simulation, the threshold current density is 37 kA/cm2. For the Ge/SiGe quantum well laser, assuming the FCA loss of barrier and buffer layer can be neglected, the threshold gain is calculated to be 600 cm−1, corresponding to a threshold carrier density of 2.1×1019cm3. Such a carrier density requires a current density as high as 1×104 kA/cm2. Moreover, the confinement of electron becomes weak due to the rise of the quasi Fermi level of electron, giving rise to a high carrier density of the barrier and intrinsic buffer region. As a result, the FCA loss of barrier and intrinsic buffer region should be taken into account. Therefore, the actual threshold current density for the quantum well laser is even higher, signifying that the laser using uniaxially tensile stressed Ge/SiGe quantum well as the gain medium is inferior to its bulk Ge counterpart in a realistic strain range. Nevertheless, we suggest that future works toward group IV quantum well laser can be devoted to GeSn/SiGeSn material system since direct bandgap gain medium as well as larger band offsets can be achieved [36, 37].

4. Conclusion

The electrically driven uniaxially tensile stressed bulk Ge laser and Ge/SiGe quantum well laser are designed. Full etched circular gratings with high reflectivity bandwidths of ~500 nm are utilized to relax the fabrication requirement. Lateral p+-n-n+ homo-junction is designed for the bulk Ge laser, allowing for longer optical cavity compared with the quasi-heterojunction along the optical cavity. As a consequence, the threshold current can be reduced as long as a longer micro-bridge can be realized experimentally. For the Ge/SiGe quantum well laser, the Ge compositions of barrier under different strain levels are optimized to attain maximum band offsets. The quantum effects on the density of state and number of states between the Γ- and L- points are analyzed, indicating that Ge/SiGe quantum well has no advantages over bulk Ge in terms of those two factors. Additionally, electrical transport simulations show that the carrier injection efficiency of the quantum well laser is lower than 103, resulting from the reduced effective mass, small conduction band offset and high injection level. The threshold current density of uniaxially tensile stressed bulk Ge laser with a realistic εxx value of 4% is calculated to be 37 kA/cm2. However, the Ge/SiGe quantum well laser at the same strain level has a threshold current density higher than 1×104 kA/cm2, revealing that the Ge/SiGe quantum well underperforms even though combined with tensile strain. Therefore, we suggest that future works of group IV quantum well laser aim at GeSn/SiGeSn material system.

Funding

National Natural Science Foundation of China (NSFC) (Grant No. 61435004).

References and links

1. L. Pavesi, L. Dal Negro, C. Mazzoleni, G. Franzò, and F. Priolo, “Optical gain in silicon nanocrystals,” Nature 408(6811), 440–444 (2000). [PubMed]  

2. G. Franzò, F. Priolo, S. Coffa, A. Polman, and A. Carnera, “Room-temperature electroluminescence from Er-doped crystalline Si,” Appl. Phys. Lett. 64(17), 2235–2237 (1994).

3. J. Liu, X. Sun, D. Pan, X. Wang, L. C. Kimerling, T. L. Koch, and J. Michel, “Tensile-strained, n-type Ge as a gain medium for monolithic laser integration on Si,” Opt. Express 15(18), 11272–11277 (2007). [PubMed]  

4. S. Wirths, R. Geiger, N. von den Driesch, G. Mussler, T. Stoica, S. Mantl, Z. Ikonic, M. Luysberg, S. Chiussi, J. M. Hartmann, H. Sigg, J. Faist, D. Buca, and D. Grützmacher, “Lasing in direct-bandgap GeSn alloy grown on Si,” Nat. Photonics 9(2), 88–92 (2015).

5. R. Koerner, D. Schwaiz, I. A. Fischer, L. Augel, S. Bechler, L. Haenel, M. Kern, M. Oehme, E. Rolseth, B. Schwartz, D. Weisshaupt, W. Zhang, and J. Schulze, “The Zener-Emitter: A novel superluminescent Ge optical waveguide-amplifier with 4.7 dB gain at 92 mA based on free-carrier modulation by direct Zener tunneling monolithically integrated on Si,” in Proceedings of the 2016 IEEE International Electron Devices Meeting (IEDM), 22.5.1–22.5.4 (2016).

6. A. W. Fang, E. Lively, Y.-H. Kuo, D. Liang, and J. E. Bowers, “A distributed feedback silicon evanescent laser,” Opt. Express 16(7), 4413–4419 (2008). [PubMed]  

7. J. Liu, X. Sun, L. C. Kimerling, and J. Michel, “Direct-gap optical gain of Ge on Si at room temperature,” Opt. Lett. 34(11), 1738–1740 (2009). [PubMed]  

8. R. E. Camacho-Aguilera, Y. Cai, N. Patel, J. T. Bessette, M. Romagnoli, L. C. Kimerling, and J. Michel, “An electrically pumped germanium laser,” Opt. Express 20(10), 11316–11320 (2012). [PubMed]  

9. M. El Kurdi, M. Prost, A. Ghrib, S. Sauvage, X. Checoury, G. Beaudoin, I. Sagnes, G. Picardi, R. Ossikovski, and P. Boucaud, “Direct Band Gap Germanium Microdisks Obtained with Silicon Nitride Stressor Layers,” ACS Photonics 3(3), 443–448 (2016).

10. M. J. Süess, R. Geiger, R. A. Minamisawa, G. Schiefler, J. Frigerio, D. Chrastina, G. Isella, R. Spolenak, J. Faist, and H. Sigg, “Analysis of enhanced light emission from highly strained germanium microbridges,” Nat. Photonics 7(6), 466–472 (2013).

11. A. Gassenq, K. Guilloy, G. Osvaldo Dias, N. Pauc, D. Rouchon, J.-M. Hartmann, J. Widiez, S. Tardif, F. Rieutord, J. Escalante, I. Duchemin, Y.-M. Niquet, R. Geiger, T. Zabel, H. Sigg, J. Faist, A. Chelnokov, V. Reboud, and V. Calvo, “1.9% bi-axial tensile strain in thick germanium suspended membranes fabricated in optical germanium-on-insulator substrates for laser applications,” Appl. Phys. Lett. 107(19), 191904 (2015).

12. A. Gassenq, S. Tardif, K. Guilloy, G. O. Dias, N. Pauc, I. Duchemin, D. Rouchon, J.-M. Hartmann, J. Widiez, J. Escalante, Y.-M. Niquet, R. Geiger, T. Zabel, H. Sigg, J. Faist, A. Chelnokov, F. Rieutord, V. Reboud, and V. Calvo, “Accurate strain measurements in highly strained Ge microbridges,” Appl. Phys. Lett. 108(24), 241902 (2016).

13. W. T. Tsang, “Quantum Confinement Heterostructure Semiconductor Lasers,” Semicond. Semimet. 24, 397–458 (1987).

14. A. Yariv, “Scaling laws and minimum threshold currents for quantum-confined semiconductor lasers,” Appl. Phys. Lett. 53(12), 1033–1035 (1988).

15. G. Pizzi, M. Virgilio, and G. Grosso, “Tight-binding calculation of optical gain in tensile strained [001]-Ge/SiGe quantum wells,” Nanotechnology 21(5), 055202 (2010). [PubMed]  

16. W. J. Fan, “Tensile-strain and doping enhanced direct bandgap optical transition of n+ doped Ge/GeSi quantum wells,” J. Appl. Phys. 114(18), 183106 (2013).

17. E. Gatti, E. Grilli, M. Guzzi, D. Chrastina, G. Isella, and H. von Känel, “Room temperature photoluminescence of Ge multiple quantum wells with Ge-rich barriers,” Appl. Phys. Lett. 98(3), 031106 (2011).

18. P. Chaisakul, D. Marris-Morini, G. Isella, D. Chrastina, N. Izard, X. Le Roux, S. Edmond, J.-R. Coudevylle, and L. Vivien, “Room temperature direct gap electroluminescence from Ge/Si0.15Ge0.85 multiple quantum well waveguide,” Appl. Phys. Lett. 99(14), 141106 (2011).

19. Z. Liu, W. Hu, C. Li, Y. Li, C. Xue, C. Li, Y. Zuo, B. Cheng, and Q. Wang, “Room temperature direct-bandgap electroluminescence from n-type strain-compensated Ge/SiGe multiple quantum wells,” Appl. Phys. Lett. 101(23), 231108 (2012).

20. G. Lin, N. Chen, L. Zhang, Z. Huang, W. Huang, J. Wang, J. Xu, S. Chen, and C. Li, “Room Temperature Electroluminescence from Tensile-Strained Si0.13Ge0.87/Ge Multiple Quantum Wells on a Ge Virtual Substrate,” Materials (Basel) 9(10), 803 (2016). [PubMed]  

21. Y. Cai, Z. Han, X. Wang, R. E. Camacho-Aguilera, L. C. Kimerling, J. Michel, and J. Liu, “Analysis of Threshold Current Behavior for Bulk and Quantum-Well Germanium Laser Structures,” IEEE J. Sel. Top. Quantum Electron. 19(4), 1901009 (2013).

22. M. Xue, X. Chen, J. Chen, M.-Y. Kao, C. Shang, K. Zang, Y. Huo, C.-Y. Lu, Y. Chen, H. Deng, T. Kamins, and J. Harris, “Ge/SiGe Quantum-well Micro-bridges with High Tensile Strain,” in Conference on Lasers and Electro-Optics, OSA Technical Digest (online) (Optical Society of America, 2017), JTu5A.125.

23. J. Jiang and J. Sun, “Theoretical analysis of optical gain in uniaxial tensile strained and n+-doped Ge/GeSi quantum well,” Opt. Express 24(13), 14525–14537 (2016). [PubMed]  

24. J. Jiang, J. Sun, Y. Zhou, J. Gao, H. Zhou, and R. Zhang, “Design and analysis of a CMOS-compatible distributed Bragg reflector laser based on highly uniaxial tensile stressed germanium,” Opt. Express 25(6), 6497–6510 (2017). [PubMed]  

25. E. T. Fei, X. Chen, K. Zang, Y. Huo, G. Shambat, G. Miller, X. Liu, R. Dutt, T. I. Kamins, J. Vuckovic, and J. S. Harris, “Investigation of germanium quantum-well light sources,” Opt. Express 23(17), 22424–22430 (2015). [PubMed]  

26. L. T. Ngo, D. Almécija, J. E. Sader, B. Daly, N. Petkov, J. D. Holmes, D. Erts, and J. J. Boland, “Ultimate-Strength Germanium Nanowires,” Nano Lett. 6(12), 2964–2968 (2006). [PubMed]  

27. V. Reboud, A. Gassenq, G. Osvaldo Dias, K. Guilloy, J. M. Escalante, S. Tardif, N. Pauc, J. M. Hartmann, J. Widiez, E. Gomez, E. Bellet Amalric, D. Fowler, D. Rouchon, I. Duchemin, Y. M. Niquet, F. Rieutord, J. Faist, R. Geiger, T. Zabel, E. Marin, H. Sigg, A. Chelnokov, and V. Calvo, “Ultra-high amplified strains in 200-mm optical germanium-on-insulator (GeOI) substrates: towards CMOS-compatible Ge lasers,” Proc. SPIE 9752, 97520F (2016).

28. C.G. Van de Walle and R. M. Martin, “Theoretical calculations of heterojunction discontinuities in the Si/Ge system,” Phys. Rev. B Condens. Matter 34(8), 5621–5634 (1986). [PubMed]  

29. J. P. Dismukes, L. Ekstrom, and R. J. Paff, “Lattice Parameter and Density in Germanium-Silicon Alloys,” J. Phys. Chem. 68(10), 3021–3027 (1964).

30. R. Braunstein, A. R. Moore, and F. Herman, “Intrinsic Optical Absorption in Germanium-Silicon Alloys,” Phys. Rev. 109(3), 695–710 (1958).

31. C. G. Van de Walle,“Band lineups and deformation potentials in the model-solid theory,” Phys. Rev. B Condens. Matter 39(3), 1871–1883 (1989). [PubMed]  

32. Y. Wang, S. Gao, K. Wang, H. Li, and E. Skafidas, “Ultra-broadband, compact, and high-reflectivity circular Bragg grating mirror based on 220 nm silicon-on-insulator platform,” Opt. Express 25(6), 6653–6663 (2017). [PubMed]  

33. G. E. Chang, S. W. Chang, and S. L. Chuang, “Strain-Balanced GezSn1−z-SixGeySn1−x−y Multiple-Quantum-Well Lasers,” IEEE J. Quantum Electron. 46(12), 1813–1820 (2010).

34. D. Nam, J.-H. Kang, M. L. Brongersma, and K. C. Saraswat, “Observation of improved minority carrier lifetimes in high-quality Ge-on-insulator using time-resolved photoluminescence,” Opt. Lett. 39(21), 6205–6208 (2014). [PubMed]  

35. S. Dominici, H. Wen, F. Bertazzi, M. Goano, and E. Bellotti, “Numerical evaluation of Auger recombination coefficients in relaxed and strained germanium,” Appl. Phys. Lett. 108(21), 211103 (2016).

36. D. Stange, N. von den Driesch, D. Rainko, C. Schulte-Braucks, S. Wirths, G. Mussler, A. T. Tiedemann, T. Stoica, J. M. Hartmann, Z. Ikonic, S. Mantl, D. Grützmacher, and D. Buca, “Study of GeSn based heterostructures: towards optimized group IV MQW LEDs,” Opt. Express 24(2), 1358–1367 (2016). [PubMed]  

37. C. S. Fenrich, X. Chen, R. Chen, Y.-C. Huang, H. Chung, M.-Y. Kao, Y. Huo, T. I. Kamins, and J. S. Harris, “Strained Pseudomorphic Ge1–xSnx Multiple Quantum Well Microdisk Using SiNy Stressor Layer,” ACS Photonics 3(12), 2231–2236 (2016).

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 (7)

Fig. 1
Fig. 1 (a) Schematic diagram of the mirco-bridge structure for introducing uniaxial stress in Ge/SiGe quantum well (not to scale). (b) Distribution of the strain component ε x x b r i calculated by 3-D FEM, when H b u f f e r = 300 n m , H r = 450 n m , W r = 600 n m , θ = 30 ° , L 1 = 10 μ m , L 2 = 740 μ m and the width of narrow bridge is 2μm.
Fig. 2
Fig. 2 (a) Optimal Ge composition of the barrier as a function of the strain component ε x x w e l l . (b) Band alignment of the Ge/Si0.09Ge0.91 quantum well with ε x x w e l l = 4 % .
Fig. 3
Fig. 3 Schematic diagrams of the uniaxially tensile stressed (a) bulk Ge laser and (b) Ge/SiGe quantum well laser. (c) Cross section view of the straight waveguide. For bulk Ge laser, W r = 600 n m , H 1 = 100 n m and H 2 = 200 n m . For Ge/SiGe quanum well laser, W r = 600 n m , H 1 = 450 n m and H 2 = 300 n m . (d) Top view of the circular grating. Reflection coefficient spectra of the fundamental TE-mode for the gratings of (e) bulk Ge laser and (f) Ge/SiGe quantum well laser.
Fig. 4
Fig. 4 (a) Net gain coefficient spectrum of the TE-polarized light for bulk Ge with a strain of 4%, a doping concentration of 7 × 10 18 c m 3 and an injected carrier density of 2.3 × 10 18 c m 3 . (b) Net gain coefficient spectra of the undoped Ge/Si0.09Ge0.91 quantum well with a strain of 4% and an injected carrier density of 2.1 × 10 19 c m 3 .
Fig. 5
Fig. 5 Cross section of (a) the p+-n-n+ junction of bulk Ge laser and (b) the p+-i-n+ junction of Ge/Si0.09Ge0.91 quantum well laser. Electron concentration profiles of the (c) p+-n-n+ junction and (d) p+-i-n+ junction at a bias voltage of 0.6V.
Fig. 6
Fig. 6 Number of states between the Γ- and L- points for bulk Ge and Ge/SiGe quantum well as a function of the strain component ε x x w e l l .
Fig. 7
Fig. 7 (a) Injected carrier density as a function of current density for the p+-i-n+ junction of Ge/Si0.09Ge0.91 quantum well laser. (b) Carrier injection efficiency as a function of injected carrier density for the p+-i-n+ junction of Ge/Si0.09Ge0.91 quantum well laser. The red curve and blue curve represent simulation results using strained effective mass and unstrained effective mass, respectively. The purple curve represents simulation results using unstrained effective mass as well as a larger conduction band offset of 0.15 eV.

Tables (1)

Tables Icon

Table 1 Material and band structure parameters at 300 K for Ge and Si a

Equations (20)

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

ε x x = a b u f a ( 1 + ε x x b r i ) 1
ε y y = a b u f a ( 1 + ε b i ) 1 C 12 C 11 + C 12 ( ε x x b r i ε b i )
ε z z = 2 C 12 C 11 ( a b u f a ( 1 + ε b i ) 1 ) C 12 C 11 + C 12 ( ε x x b r i ε b i )
a | | M Q W = ( a b G b h b + a w G w h w ) / ( G b h b + G w h w )
G = 2 ( C 11 + 2 C 12 ) ( 1 C 12 / C 11 )
a S i G e = 0.0282 x 2 + 0.1981 x + 5.4315
Δ E c i = [ Ξ d 1 + Ξ u { a ^ i a ^ i } ] : ε
Δ E c Γ = a c ( ε x x + ε y y + ε z z )
Δ E c L = a L ( ε x x + ε y y + ε z z )
Δ E c Δ 2 = Ξ d Δ ( ε x x + ε y y + ε zz ) + Ξ u Δ ε x x
Δ E c Δ 4 = Ξ d Δ ( ε x x + ε y y + ε zz ) + Ξ u Δ ε y y
E v = E v . a v + Δ s o / 3
E c i = E v + E g i
E g Γ = 0.7985 x + 4.185 ( 1 x ) 0.14 x ( 1 x )
E g L = 1.86 1.2 x
E g Δ = 1.087 0.487 x + 0.264 x 2
I t h ( η γ τ e ) = N 2 D t r ( W L ) + W d g 2 D ln R 1 + W L d g 2 D α s c a t
g 3 D Q W = 2 π 2 L z Δ E n L H e a v i s i d e ( E c + Δ E E n L L ( k t ) ) d k x d k y
( Γ a c t g b u l k Γ p α p Γ n α n ) L g = ln ( 1 R 1 R 2 )
( Γ Q W g Q W Γ b u f f e r α b u f f e r Γ b a r r i e r α b a r r i e r Γ n α n ) L g = ln ( 1 R 1 R 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.