This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

First-principles theory of direct-gap optical emission in hexagonal Ge
and its enhancement via strain engineering

Christopher A. Broderick [email protected] Materials Department, University of California, Santa Barbara, California 93106-5050, U.S.A. School of Physics, University College Cork, Cork T12 YN60, Ireland Tyndall National Institute, University College Cork, Lee Maltings, Dyke Parade, Cork T12 R5CP, Ireland    Xie Zhang School of Materials Science and Engineering, Northwestern Polytechnical University, Xi’an 710072, China    Mark E. Turiansky Materials Department, University of California, Santa Barbara, California 93106-5050, U.S.A.    Chris G. Van de Walle Materials Department, University of California, Santa Barbara, California 93106-5050, U.S.A.
Abstract

The emergence of hexagonal Ge (2H-Ge) as a candidate direct-gap group-IV semiconductor for Si photonics mandates rigorous understanding of its optoelectronic properties. Theoretical predictions of a “pseudo-direct” band gap, characterized by weak oscillator strength, contrast with a claimed high radiative recombination coefficient BB comparable to conventional (cubic) InAs. We compute BB in 2H-Ge from first principles and quantify its dependence on temperature, carrier density and strain. For unstrained 2H-Ge, our calculated spontaneous emission spectra corroborate that measured photoluminescence corresponds to direct-gap emission, but with BB being approximately three orders of magnitude lower than in InAs. We confirm a pseudo-direct- to direct-gap transition under 2\sim 2% [0001] uniaxial tension, which can enhance BB by up to three orders of magnitude, making it comparable to that of InAs. Beyond quantifying strong enhancement of BB via strain engineering, our analysis suggests the dominance of additional, as-yet unquantified recombination mechanisms in this nascent material.

Introduction

Si photonics serves as an enabling platform for applications ranging from datacoms and optical computing to sensing and quantum computing. Zhou et al. (2023) However, despite significant progress, the indirect band gaps of Si and Ge render them inefficient light emitters, limiting the realization of active photonic components including light-emitting diodes (LEDs), laser and optical interconnects for monolithic integration on Si. It has therefore remained a persistent objective to develop novel direct-gap group-IV semiconductors compatible with complementary metal-oxide semiconductor (CMOS) fabrication. Geiger et al. (2015)

Advances in nanowire (NW) fabrication have enabled growth of Ge in the metastable lonsdaleite (2H, “hexagonal diamond”) phase. Hauge et al. (2017) Predictions of a “pseudo-direct” fundamental band gap in 2H-Ge De and Pryor (2014); Rödl et al. (2019) – originating from back-folding of the L6c conduction band (CB) minimum of conventional cubic (3C, diamond-structured) Ge and characterized by weak oscillator strength Rödl et al. (2019) – have been confirmed via experimental demonstrations of direct-gap photoluminescence Fadaly et al. (2020) (PL) and stimulated emission van Tilburg et al. (2024) from 2H-Ge NWs. Exploiting crystal structure as a novel degree of freedom for band-structure engineering to realize a CMOS-compatible direct-gap semiconductor is driving a surge of interest in 2H-Ge as a light-emitter for Si photonics. It is therefore critical to develop a detailed understanding of radiative recombination in 2H-Ge, to interpret experimental measurements and inform development of this nascent material.

In this Article, we analyze radiative recombination in 2H-Ge via first-principles calculations. We compute the spontaneous emission (SE) rate, which we compare to measured PL spectra. Fadaly et al. (2020) Via the integrated SE rate we explicitly compute the radiative recombination coefficient BB and lifetime τrad\tau_{\scalebox{0.7}{\text{rad}}}, and quantify their dependence on carrier density nn and temperature TT. We also investigate the impact of [0001] uniaxial tension on BB and τrad\tau_{\scalebox{0.7}{\text{rad}}}, confirming that a strain-induced pseudo-direct- to direct-gap transition Suckert et al. (2021) drives significant enhancement of BB. The SE calculations corroborate that measured PL from 2H-Ge NWs corresponds to radiative recombination across the fundamental (direct) band gap. However, our analysis suggests that the initial assumption of purely radiative recombination in 2H-Ge, which underpinned inference of a high BB coefficient, is not consistent with theory. Calculated τrad\tau_{\scalebox{0.7}{\text{rad}}} values suggest that the measured carrier lifetime is dominated by as-yet unquantified recombination mechanisms. Application of uniaxial tension along the [0001] NW axis can enhance BB by up to three orders of magnitude, making it comparable to that in a conventional (cubic) direct-gap III-V semiconductor.

Unstrained 2H-Ge

We employ the band structure and momentum matrix elements calculated using density functional theory (DFT) – via the Tran-Blaha modified Becke-Johnson (TB-mBJ) exchange potential – to compute SE spectra, with the BB coefficient then computed via the integrated SE rate (cf. Methods).

Refer to caption
Figure 1: (a) Band gap vs. temperature for cubic 3C-Ge (upper) and hexagonal 2H-Ge (lower), from parameterized TB-mBJ DFT calculations (lines) and experiment (circles). The dashed line is the calculated 2H-Ge band gap; the solid line has been redshifted to align with experiment at T=0T=0 K. Experimental data for 3C-Ge and 2H-Ge are from Refs. Thurmond, 1975 and Fadaly et al., 2020, respectively. (b) DFT-calculated band structure of unstrained 2H-Ge, with TB-mBJ parameterization corresponding to T=300T=300 K. (c) Measured PL (circles) and calculated SE (lines) spectra for unstrained 2H-Ge, for temperatures T=T= 4, 50, 100, 200 and 300 K. Vertical arrows denote the calculated band-gap energy at each temperature (as in (a)). Experimental PL data are from Ref. Fadaly et al., 2020.

In a direct-gap semiconductor, Bf~ΓEg2TD/2B\propto\widetilde{f}_{\Gamma}\,E_{g}^{2}\,T^{-D/2} at temperature TT in DD spatial dimensions, Landsberg (1991) where EgE_{g} is the band gap and f~Γ\widetilde{f}_{\Gamma} is the polarization-averaged zone-center oscillator strength between the CB and VB edge states. The 2H-Ge band gap has been measured to decrease from 0.352 eV at T=0T=0 K to 0.279 eV at T=300T=300 K. Fadaly et al. (2020) Since BEg2B\propto E_{g}^{2}, given this sizeable 21\approx 21% decrease of the 2H-Ge band gap up to room temperature, we include a TT-dependent band gap in our DFT calculations. We achieve this by treating the Becke-Roussel mixing parameter cBRc_{\scalebox{0.7}{\text{BR}}} in the TB-mBJ exchange potential as an empirical parameter, which we fit to the experimental TT-dependent fundamental (indirect) L6c-Γ8v\Gamma_{8v} band gap of conventional cubic 3C-Ge. Thurmond (1975) Keeping this fit cBR(T)c_{\scalebox{0.7}{\text{BR}}}(T) fixed, we then apply it to predict the TT-dependent band gap of 2H-Ge. The results of these calculations are shown in Fig. 1(a). Our calculated T=0T=0 K (cBR=1.215c_{\scalebox{0.7}{\text{BR}}}=1.215) 2H-Ge band gap of 0.360 eV is in excellent quantitative agreement with experiment, exceeding the measured band gap by only 8 meV. This simple treatment implicitly encapsulates the lattice thermal expansion and electron-phonon coupling that drive the reduction of EgE_{g} with increasing TT, accurately capturing the measured TT-dependent band gap of 2H-Ge. Figure 1(b) shows the calculated band structure of 2H-Ge, with cBRc_{\scalebox{0.7}{\text{BR}}} (=1.185=1.185) chosen to correspond to T=300T=300 K, where our calculated Eg=0.281E_{g}=0.281 eV exceeds the measured value by only 2 meV. We note the presence of a direct fundamental band gap, between the Γ8c\Gamma_{8c} CB and Γ9v\Gamma_{9v} VB extrema, with the former originating from back-folding of the L6c CB minimum of 3C-Ge (visible in the lowest CB along Δ\Delta in Fig. 1(b)). De and Pryor (2014); Rödl et al. (2019)

To validate our analysis, we compare calculated SE spectra rsp(ω)r_{\scalebox{0.7}{\text{sp}}}(\hbar\omega) for unstrained 2H-Ge to the PL measurements of Ref. Fadaly et al., 2020. The NWs of Ref. Fadaly et al., 2020 have diameters in excess of 150 nm, such that no quantum confinement effects are expected to impact the optoelectronic properties, thereby allowing to compare our bulk DFT calculations to experiment. The results of this qualitative comparison are summarized in Fig. 1(c), for T=4T=4 – 300 K. Beginning at T=T= 4 K we adjust the carrier density nn so that the calculated SE peak energy aligns with the measured PL peak energy, yielding n=3×1017n=3\times 10^{17} cm-3. Next, we adjust the spectral width δ\delta of the hyperbolic secant lineshape used to compute rsp(ω)r_{\scalebox{0.7}{\text{sp}}}(\hbar\omega) (cf. Methods), finding that δ=14\delta=14 meV provides a good description of the measured PL lineshape. The remaining SE spectra in Fig. 1(c) are then computed by increasing TT while keeping nn and δ\delta fixed, motivated by the fact that the PL spectra were measured at fixed excitation intensity (I=1.9I=1.9 kW cm-2). We note that the TT-dependent PL peak energy is distinct from the Eg(T)E_{g}(T) data of Fig. 1(a); the latter were extracted via Lasher-Stern-Würfel fits to the measured PL spectra. Fadaly et al. (2020) Similarly, the SE peak at temperature TT is blueshifted with respect to Eg(T)E_{g}(T) (vertical arrows, Fig. 1(c)) due to the Burstein-Moss effect. The calculated SE spectra track the measured PL spectra accurately with increasing TT, capturing the redshift of the PL peak and the emergence of a high-energy tail driven by the increasing carrier temperature.

Having validated our theoretical approach vs. experiment for unstrained 2H-Ge, we turn our attention to the BB coefficient. Figure 2(a) compares the calculated BB coefficient vs. TT of 3C-InAs and 2H-Ge in the non-degenerate regime (i.e. at low carrier density, n=1015n=10^{15} cm-3). 3C-InAs is chosen as a reference material due to its possessing a narrow, direct band gap (=0.354=0.354 eV at T=300T=300 K Vurgaftman and Meyer (2001)) comparable to the pseudo-direct fundamental band gap of 2H-Ge. We note that our calculated B=1.64×1011B=1.64\times 10^{-11} cm3 s-1 for 3C-InAs at T=300T=300 K is in excellent quantitative agreement with the value B=1.81×1011B=1.81\times 10^{-11} cm3 s-1 recently reported by Hader et al., Hader et al. (2022) who employed input from DFT calculations to solve the many-body semiconductor luminescence equations. At T=300T=300 K we calculate B=7.44×1015B=7.44\times 10^{-15} cm3 s-1 for 2H-Ge, predicting that the room temperature BB coefficient of unstrained 2H-Ge is approximately three orders of magnitude lower than that of 3C-InAs.

Refer to caption
Figure 2: (a) BB coefficient vs. temperature TT for unstrained 2H-Ge (blue circles) and 3C-InAs (red squares) at carrier density n=1015n=10^{15} cm-3. Inset: log-log plot of BB vs. TT in 2H-Ge for T=200T=200 – 300 K (circles); the line is a linear fit of slope xx, where BTxB\propto T^{x}. (b) BB coefficient vs. nn for 2H-Ge at T=4T=4, 50, 100, 200 and 300 K. Circles denote DFT-calculated values of BB; lines are fits following Eq. (1). (c) Radiative lifetime τrad\tau_{\scalebox{0.7}{\text{rad}}} vs. nn for 2H-Ge at T=4T=4, 50, 100, 200 and 300 K. Circles denote τrad\tau_{\scalebox{0.7}{\text{rad}}} computed using the DFT-calculated BB values of (b) via τrad1=Bn\tau_{\scalebox{0.7}{\text{rad}}}^{-1}=Bn; lines were obtained similarly via the Eq. (1) fits of (b).

This low BB coefficient is a consequence of the low Γ8c\Gamma_{8c}-Γ9v\Gamma_{9v} optical (momentum) matrix element, due to the fact that the Γ8c\Gamma_{8c} CB minimum is a back-folded L6c state from 3C-Ge. This Γ8c\Gamma_{8c} state thus contains predominantely pp-like orbital character, closely matching its symmetry to that of the pp-like Γ9v\Gamma_{9v} VB maximum. For 2H-Ge at T=4T=4 K we compute oscillator strength f~(Γ8c-Γ9v)=EP/Eg=6.81×103\widetilde{f}(\Gamma_{8c}\text{-}\Gamma_{9v})=E_{P}/E_{g}=6.81\times 10^{-3}, where EP=|p~(Γ8c-Γ9v)|2/2m0=2.45E_{P}=|\widetilde{p}(\Gamma_{8c}\text{-}\Gamma_{9v})|^{2}/2m_{0}=2.45 meV is the Kane parameter and p~(Γ8c-Γ9v)\widetilde{p}(\Gamma_{8c}\text{-}\Gamma_{9v}) is the polarization-averaged interband momentum matrix element. This oscillator strength is in close agreement with f~(Γ8c-Γ9v)=3.95×103\widetilde{f}(\Gamma_{8c}\text{-}\Gamma_{9v})=3.95\times 10^{-3} obtained by polarization-averaging the TB-mBJ results of Ref. Rödl et al., 2019. We note that our analysis confirms the optical selection rules identified in Ref. Rödl et al., 2019, consistent with recent polarization-resolved PL measurements. van Tilburg et al. (2023)

To quantify the TT dependence of BB we assume BTxB\propto T^{x} and compute xx as the slope of a linear fit to logB\log B vs. logT\log T. This is shown in the inset to Fig. 2(a) for T=200T=200 – 300 K, where we compute BT2.18B\propto T^{-2.18} for 2H-Ge (vs. BT1.80B\propto T^{-1.80} for 3C-InAs). This exceeds the BT3/2B\propto T^{-3/2} dependence expected for an idealized bulk semiconductor, and is partially a consequence of capturing the measured Eg(T)E_{g}(T) in our calculations. Increasing to a degenerate carrier density n=1018n=10^{18} cm-3 we calculate BT1.76(1.47)B\propto T^{-1.76(-1.47)} for 2H-Ge (3C-InAs), suggesting that the radiative recombination rate τrad1=Bn\tau_{\scalebox{0.7}{\text{rad}}}^{-1}=Bn in 2H-Ge decreases more rapidly with increasing TT than in 3C-InAs.

Figure 2(b) summarizes the BB coefficient vs. nn, where open circles denote the results of our DFT calculations. At fixed TT we note a strong reduction of BB as nn is increased beyond the non-degenerate regime, in line with the expected impact of phase-space filling. Hader et al. (2005); Hader and Moloney (2006) For example, at T=300T=300 K we compute that BB decreases weakly between n=1015n=10^{15} and 101810^{18} cm-3, from 7.44 to 6.14×10156.14\times 10^{-15} cm3 s-1, beyond which carrier density it decreases rapidly due to phase-space filling. At fixed TT this phase-space filling can be parameterized by fitting to the DFT-calculated BB vs. nn using the empirical relation Grein et al. (2002)

B(n)=B01+nn0,B(n)=\frac{B_{0}}{1+\frac{n}{n_{0}}}\,, (1)

where B0B_{0} is the best-fit value of BB in the non-degenerate regime. The results of this fitting are shown using solid lines in Fig. 2(b) where, e.g., we obtain B0=1.63×1012B_{0}=1.63\times 10^{-12} (7.01×10157.01\times 10^{-15}) cm3 s-1 and n0=3.78×1016n_{0}=3.78\times 10^{16} (5.80×10185.80\times 10^{18}) cm-3 at T=4T=4 (300) K.

Using the DFT-calculated (circles) and Eq. (1) fits (lines) of Fig. 2(b), we compute the radiative lifetime τrad\tau_{\scalebox{0.7}{\text{rad}}} vs. nn at fixed TT via τrad1=Bn\tau_{\scalebox{0.7}{\text{rad}}}^{-1}=Bn. The resulting lifetimes are summarized in Fig. 2(c). At n=1015n=10^{15} cm-3 we compute τrad=0.59\tau_{\scalebox{0.7}{\text{rad}}}=0.59 ms at T=4T=4 K, in good agreement with the values calculated by Rödl et al. Rödl et al. (2019) and Suckert et al. Suckert et al. (2021) As TT increases, our calculated non-degenerate τrad\tau_{\scalebox{0.7}{\text{rad}}} increases strongly, reaching 135 ms at T=300T=300 K. This diverges strongly from Refs. Rödl et al., 2019 and Suckert et al., 2021, which predicted a TT-independent τrad\tau_{\scalebox{0.7}{\text{rad}}} up to T400T\approx 400 K. We note that the TT independence predicted in Refs. Rödl et al., 2019 and Suckert et al., 2021 is a consequence of employing Maxwell-Boltzmann statistics, which artificially pins the quasi-Fermi levels by enforcing unit occupancy at the band extrema – i.e. generating an explicit change of nn as TT is varied. The use of Maxwell-Boltzmann statistics therefore precludes reliable TT-dependent prediction of B=(τradn)1B=(\tau_{\scalebox{0.7}{\text{rad}}}\,n)^{-1}. This contrasts with the use of Fermi-Dirac statistics in the present analysis, which allows to independently vary nn and TT. At T=300T=300 K we expect that τrad\tau_{\scalebox{0.7}{\text{rad}}} exceeds its T=4T=4 K value by a factor B(4K)/B(300K)B(4~{}\text{K})/B(300~{}\text{K}), with τradT+2.18\tau_{\scalebox{0.7}{\text{rad}}}\propto T^{+2.18} at fixed nn in 2H-Ge. The strong increase of τrad\tau_{\scalebox{0.7}{\text{rad}}} with increasing TT arises in our analysis due to the decrease (increase) in electron (hole) quasi-Fermi level, relative to the CB minimum (VB maximum), required to maintain n=1015n=10^{15} cm-3 with increasing TT. The calculated τrad\tau_{\scalebox{0.7}{\text{rad}}} at fixed TT decreases strongly with increasing nn, reaching a value of 17.4 μ\mus (162.8 μ\mus) at T=4T=4 (300) K for n=1018n=10^{18} cm-3.

We note that these calculated radiative lifetimes exceed the measured 2H-Ge carrier lifetime, τ=0.98\tau=0.98 (0.46) ns at T=4T=4 (300) K, Fadaly et al. (2020) by approximately four orders of magnitude. We recall that our calculated SE spectra qualitatively, but closely, reproduce the measured TT-dependent PL spectra of Ref. Fadaly et al., 2020. This supports the key conclusion of Ref. Fadaly et al., 2020, that the observed PL is consistent with direct-gap band-to-band optical emission from bulk-like 2H-Ge. However, our analysis – which is consistent with the low-temperature DFT calculations of Refs. Rödl et al., 2019 and Fadaly et al., 2020, but which allows to accurately predict the BB coefficient and its temperature dependence – indicates that the measured carrier lifetime τ\tau is not consistent with the expected radiative lifetime. Firstly, τrad\tau_{\scalebox{0.7}{\text{rad}}} for a direct-gap semiconductor is expected to increase with increasing TT; the data of Ref. Fadaly et al., 2020 demonstrate an 50\approx 50% decrease of τ\tau between T=4T=4 and 300 K. Secondly, a high 2H-Ge BB coefficient, comparable to that of direct-gap 3C-InAs or 3C-GaAs, was inferred in Ref. Fadaly et al., 2020 by assuming τ=τrad\tau=\tau_{\scalebox{0.7}{\text{rad}}} (i.e. by assuming an internal quantum efficiency of 100%). Our explicit calculation of BB for 2H-Ge – beyond the previously investigated non-degenerate regime Fadaly et al. (2020); Suckert et al. (2021) – suggests that BB is approximately three orders of magnitude lower than inferred in Ref. Fadaly et al., 2020. The fundamental origin of this behavior is the weak Γ8c\Gamma_{8c}-Γ9v\Gamma_{9v} optical matrix element. Our calculated EP=2.45E_{P}=2.45 meV is at odds with the recent value EP3.8E_{P}\geq 3.8 eV estimated by van Lange et al. van Lange et al. (2024) However, we note that the source of this discrepancy is as described above: the analysis of van Lange et al. is predicated on the same underlying assumption of Ref. Fadaly et al., 2020, that τ=τrad\tau=\tau_{\scalebox{0.7}{\text{rad}}}. The oscillator strength associated with the fundamental direct band gap of 2H-Ge thus remains a topic of active investigation. As an experimental test of these competing interpretations, we note that our calculated SE spectra predict that – at fixed excitation and temperature – the PL intensity emitted by a 2H-Ge NW should be approximately three orders of magnitude lower than that emitted by an equivalent 3C- or 2H-InAs NW.

Generally, we expect τ1=τrad1+τnon-rad1\tau^{-1}=\tau_{\scalebox{0.7}{\text{rad}}}^{-1}+\tau_{\scalebox{0.7}{\text{non-rad}}}^{-1} for the carrier recombination rate. τrad1\tau_{\scalebox{0.7}{\text{rad}}}^{-1} can include contributions from phonon-assisted radiative recombination in addition to the direct (Δ𝐤=0\Delta\mathbf{k}=0) radiative recombination considered herein. We note the close theory-experiment correspondence of Fig. 1(c), where the absence of visible phonon replicas in the measured PL spectra indicates that phonon-assisted radiative recombination does not contribute appreciably to τrad1\tau_{\scalebox{0.7}{\text{rad}}}^{-1}. τnon-rad1\tau_{\scalebox{0.7}{\text{non-rad}}}^{-1} is the total non-radiative recombination rate, and can include contributions from Auger-Meitner recombination, defect-related (Shockley-Read-Hall) recombination and/or surface recombination at NW facets. Willem-Jan Berghuis et al. (2024) Initial investigations of native defects in 2H-Ge NWs have not confirmed the presence of localized trap states lying within the 2H-Ge band gap. Fadaly et al. (2021) However, we note the presence of a weak “s-shape” in the experimental TT-dependent 2H-Ge band gap (cf. Fig. 1(a)). This is typically a signature of trapped carriers undergoing thermionic emission from near-band-edge localized states with increasing TT. Imhof et al. (2010) The measured τ1\tau\lesssim 1 ns carrier lifetime suggests the presence of additional recombination mechanisms that, while evidently not precluding direct-gap optical emission, nonetheless dominate recombination in first-generation 2H-Ge NWs. Further work is therefore required to quantify additional recombination mechanisms in 2H-Ge.

Strained 2H-Ge

We now turn our attention to the possibility of increasing BB (decreasing τrad\tau_{\scalebox{0.7}{\text{rad}}}) via strain engineering. Suckert et al. Suckert et al. (2021) predicted theoretically that application of [0001] uniaxial tension to 2H-Ge shifts the pseudo-direct Γ8c\Gamma_{8c}-Γ9v\Gamma_{9v} (direct Γ7c\Gamma_{7c}-Γ9v\Gamma_{9v}) band gap upwards (downwards) in energy. This produces a pseudo-direct- to direct-gap transition for tensile cc-axis strain ϵzz\epsilon_{zz} close to 2%, such that the fundamental band gap becomes optically bright, leading to a strong reduction of τrad\tau_{\scalebox{0.7}{\text{rad}}}. Suckert et al. (2021) We note that, due to the close structural similarity of [111] 3C-Ge and [0001] 2H-Ge, this transition corresponds to the known ability to drive an indirect- to direct-gap transition in cubic 3C-Ge via uniaxial [111] tension. Zhang et al. (2009)

Figure 3 shows the calculated band gaps of 2H-Ge under [0001] uniaxial tension for (a) an idealized case in which applying a cc-axis strain in isolation produces lattice parameter c=(1+ϵzz)c0c=(1+\epsilon_{zz})\,c_{0}, and (b) the physically realistic case in which the uniaxial stress applied to bring about the same extension of cc is accompanied by relaxation of the cc-plane lattice parameter aa and internal parameter uu. Case (b) is equivalent to pseudomorphic growth of compressively strained 2H-Ge on an [0001]-oriented substrate, where lattice relaxation in response to cc-plane compression produces tension along [0001]. In the unrelaxed case we identify a pseudo-direct- (solid blue) to direct-gap (dashed blue) transition for ϵzz=1.7\epsilon_{zz}=1.7%, confirming the prediction of Suckert et al. Suckert et al. (2021) Comparing Figs. 3(a) and 3(b) we observe that lattice relaxation drives a more rapid downward shift in energy of the U5c CB minima (dash-dotted red), leading to the emergence of an indirect-gap regime for a narrow range of tensile strains, ϵzz=2.1\epsilon_{zz}=2.1 – 2.3%, coinciding with the pseudo-direct Γ8c\Gamma_{8c} to direct Γ7c\Gamma_{7c} crossover at Γ\Gamma, beyond which the band gap is direct. This contrasts with Ref. Suckert et al., 2021, in which an indirect-gap regime was not observed in TB-mBJ DFT calculations employing self-consistent evaluation of cBRc_{\scalebox{0.7}{\text{BR}}}. Our test calculations utilizing this approach confirm its underestimation of the experimental band gap of unstrained 2H-Ge by 15\approx 15%, Rödl et al. (2019); Suckert et al. (2021) and suggests that Fig. 3(c) in Ref. Suckert et al., 2021 may correspond to the unrelaxed case. Our analysis suggests that Brillouin zone- (BZ-) edge U-valley CB states can lie close in energy to the CB minimum in direct-gap tensile strained 2H-Ge. From the perspective of LED/laser operation, engineering the Γ8c\Gamma_{8c}-U5c CB valley splitting then becomes an important consideration: the highly degenerate U valleys present a large density of states, which can drive rapid intervalley scattering of Γ\Gamma-point electrons to the BZ edge.

Refer to caption
Figure 3: Evolution with [0001] uniaxial tensile strain ϵzz\epsilon_{zz} of the T=300T=300 K 2H-Ge band gaps between the Γ9v\Gamma_{9v} VB maximum and the pseudo-direct Γ8c\Gamma_{8c} (solid blue), direct Γ7c\Gamma_{7c} (dashed blue), indirect U5c (dash-dotted red), and indirect M5c (dotted red) CB minima. (a) Band gaps vs. applied [0001] strain ϵzz\epsilon_{zz}, but without allowing relaxation of the cc-plane lattice parameter aa or internal parameter uu. (b) As in (a), but including relaxation of the lattice and internal parameters aa and uu.
Refer to caption
Figure 4: (a) BB coefficient vs. [0001] uniaxial tensile strain ϵzz\epsilon_{zz} for 2H-Ge at T=300T=300 K and n=1015n=10^{15} cm-3 (closed blue circles). The horizontal dashed red line shows the calculated BB coefficient of unstrained 3C-InAs at the same TT and nn (cf. Fig. 2(a)). (b) BB coefficient vs. nn for 2H-Ge at ϵzz=0\epsilon_{zz}=0, 1, 2, 3 and 4%. Circles denote DFT-calculated values of BB; lines are fits following Eq. (1). (c) Radiative lifetime τrad\tau_{\scalebox{0.7}{\text{rad}}} vs. nn for 2H-Ge at ϵzz=0\epsilon_{zz}=0, 1, 2, 3 and 4%. Circles denote τrad\tau_{\scalebox{0.7}{\text{rad}}} computed using the DFT-calculated BB values of (b) via τrad1=Bn\tau_{\scalebox{0.7}{\text{rad}}}^{-1}=Bn; lines were obtained similarly via the Eq. (1) fits of (b).

The results of our strain-dependent analysis are summarized in Fig. 4, where we restrict our attention to the relaxed case of Fig. 3(b). Figure 4(a) shows the calculated BB coefficient of 2H-Ge vs. ϵzz\epsilon_{zz} at T=300T=300 K in the non-degenerate limit (blue). Between ϵzz=0\epsilon_{zz}=0 and 1% we note a decrease in BB. This is a consequence of the rapid reduction in Γ8c\Gamma_{8c}-U5c splitting (cf. Fig. 3(b)), with these BZ-edge states then being partially occupied via the high-energy tail of the electron distribution function. Since we consider only direct radiative transitions, these BZ-edge electrons cannot recombine with zone-center holes, thereby reducing the number of electron-hole pairs available to contribute to the SE rate. As ϵzz\epsilon_{zz} is increased above 1% the Γ8c\Gamma_{8c}-Γ7c\Gamma_{7c} splitting reduces rapidly, leading to an increasing fraction of electrons occupying CB states in the Γ7c\Gamma_{7c} valley. The Γ7c\Gamma_{7c}-Γ9v\Gamma_{9v} transition, which derives from the direct Γ7c\Gamma_{7c}-Γ8v\Gamma_{8v} transition in 3C-Ge, is optically bright. In unstrained 2H-Ge it has Eg=0.873E_{g}=0.873 eV and EP=18.18E_{P}=18.18 eV, with oscillator strength f~(Γ7c-Γ9v)=20.82\widetilde{f}(\Gamma_{7c}\text{-}\Gamma_{9v})=20.82, which is 103\sim 10^{3} times larger than that associated with the fundamental (pseudo-direct) Γ8c\Gamma_{8c}-Γ9v\Gamma_{9v} band gap. For ϵzz>1\epsilon_{zz}>1% recombination via the Γ7c\Gamma_{7c}-Γ9v\Gamma_{9v} transition dominates the SE rate, leading to a strong increase in BB. For example, at ϵzz=3\epsilon_{zz}=3% we compute B=1.44×1012B=1.44\times 10^{-12} cm3 s-1 with direct band gap Eg=0.299E_{g}=0.299 eV – i.e. band gap close to that of unstrained 2H-Ge, with BB increased (τrad\tau_{\scalebox{0.7}{\text{rad}}} decreased) by two orders of magnitude vs. unstrained 2H-Ge.

Figures 4(b) and 4(c) respectively show the strain-dependent BB and τrad\tau_{\scalebox{0.7}{\text{rad}}} vs. nn at T=300T=300 K. As in Figs. 2(b) and 2(c) open circles denotes DFT-calculated values, while solid lines are fits following Eq. (1). As in unstrained 2H-Ge, phase-space filling decreases both BB and τrad\tau_{\scalebox{0.7}{\text{rad}}} with increasing nn. Again selecting ϵzz=3\epsilon_{zz}=3% as an example, as the band gap is close to that of unstrained 2H-Ge, we fit B0=1.65×1012B_{0}=1.65\times 10^{-12} cm3 s-1 and n0=1.28×1018n_{0}=1.28\times 10^{18} cm-3. Between n=1015n=10^{15} and 101810^{18} cm-3 BB decreases weakly, from 1.44×10121.44\times 10^{-12} to 9.15×10139.15\times 10^{-13} cm-3, beyond which carrier density it decreases rapidly. Similarly, at ϵzz=3\epsilon_{zz}=3% we compute τrad=0.69\tau_{\scalebox{0.7}{\text{rad}}}=0.69 ms, decreasing to 1.1 μ\mus by n=1018n=10^{18} cm-3 and 0.56 μ\mus at n=1019n=10^{19} cm-3. We note that these radiative lifetimes are approximately two orders of magnitude lower than those computed for unstrained 2H-Ge above. These results summarize that, in addition to allowing to tune the emission wavelength, strain engineering of 2H-Ge can be expected to strongly enhance τrad1\tau_{\scalebox{0.7}{\text{rad}}}^{-1} for mid-infrared emission, provided adequate energy separation between the Γ\Gamma- and U-point CB valley minima can be maintained.

Conclusions

In summary, we analyzed radiative recombination in lonsdaleite Ge from first principles, including the BB coefficient and its dependence on temperature, carrier density and strain. For unstrained 2H-Ge, calculated SE spectra corroborate that experimentally observed PL from 2H-Ge NWs is consistent with direct-gap optical emission. The calculated BB coefficient is approximately three orders of magnitude lower than in 3C-InAs, indicating a significant difference between the radiative lifetime τrad\tau_{\scalebox{0.7}{\text{rad}}} and measured carrier lifetime τ\tau, associated with as-yet unquantified recombination mechanisms. We confirmed the presence of a pseudo-direct- to direct-gap transition under [0001] unaxial tension, and elucidated the role played by internal relaxation. We observed an indirect band gap in a narrow strain range straddling this transition, with a direct band gap emerging for tensile strains 2.3\gtrsim 2.3%. This highlights CB valley splitting as an important consideration for band-structure engineering of 2H-Ge. We predict that tensile strain strongly enhances BB, which can approach that of 3C-InAs in the direct-gap regime. Analysis of additional recombination mechanisms is required to inform further development of this emerging semiconductor, with the aim of realizing a direct-gap emitter for Si integrated photonics.

Methods

Density functional theory:

DFT calculations were performed using the projector-augmented wave (PAW) method, as implemented in the Vienna Ab-initio Simulation Package (VASP). Blöchl (1994); Kresse and Joubert (1999) All calculations use a plane-wave cut-off energy of 500 eV, include spin-orbit coupling, and employ PAW potentials in which the semi-core (3d)10(3d)^{10} orbitals of Ge and (4d)10(4d)^{10} orbitals of In are treated as valence states. BZ integration for the 2H and 3C crystal phases was performed using Γ\Gamma-centered 11×11×711\times 11\times 7 and 11×11×1111\times 11\times 11 Monkhorst-Pack (MP) k-point grids, respectively. The lattice parameters a0=3.961a_{0}=3.961 Å and c0=6.534c_{0}=6.534 Å, and internal parameter u0=0.3744u_{0}=0.3744, of 2H-Ge were computed in the LDA and are in good quantitative agreement with previous calculations. Rödl et al. (2019) Electronic structure calculations were performed in the meta-generalized gradient approximation, employing the TB-mBJ exchange potential in conjunction with the LDA to the electronic correlation. Tran and Blaha (2009) As described above, the Becke-Roussel mixing parameter cBR(T)c_{\scalebox{0.7}{\text{BR}}}(T) was determined via an empirical fit to the experimental TT-dependent L6c-Γ8v\Gamma_{8v} band gap of 3C-Ge (cf. Fig. 1(a)). For 3C-InAs we use the LDA-calculated lattice parameter a0=6.030a_{0}=6.030 Å, and employ an empirical fit cBR(T)c_{\scalebox{0.7}{\text{BR}}}(T) to the experimental TT-dependent fundamental Γ6c\Gamma_{6c}-Γ8v\Gamma_{8v} band gap. Shen et al. (2019); Vurgaftman and Meyer (2001)

Radiative recombination:

Using the TB-mBJ-calculated band structure and interband momentum matrix elements, Gajdoš et al. (2006) we compute the SE (rate) spectrum in the quasi-equilibrium approximation as Chang et al. (1995)

rsp(ω)\displaystyle r_{\scalebox{0.7}{\text{sp}}}(\hbar\omega) =\displaystyle= e2nrωπϵ0m02c3nc,nvdk(2π)3|p~ncnvk|2f(ϵnck,Fe)\displaystyle\frac{e^{2}n_{r}\omega}{\pi\epsilon_{0}m_{0}^{2}\hbar c^{3}}\sum_{n_{c},n_{v}}\int\frac{\text{d}\textbf{k}}{(2\pi)^{3}}\,|\widetilde{p}_{n_{c}n_{v}\scalebox{0.7}{{k}}}|^{2}\,f(\epsilon_{n_{c}\scalebox{0.7}{{k}}},F_{e}) (2)
×\displaystyle\times (1f(ϵnvk,Fh))δ(ϵnckϵnvkω),\displaystyle(1-f(\epsilon_{n_{v}\scalebox{0.7}{{k}}},F_{h}))\,\delta(\epsilon_{n_{c}\scalebox{0.7}{{k}}}-\epsilon_{n_{v}\scalebox{0.7}{{k}}}-\hbar\omega)\,,

where nc(v)n_{c(v)} indexes the CBs (VBs), ϵnc(v)k\epsilon_{n_{c(v)}\scalebox{0.7}{{k}}} is the CB (VB) energy at wave vector k, ff is the Fermi-Dirac distribution function at temperature TT, Fe(h)F_{e(h)} is the electron (hole) quasi-Fermi level at temperature TT and electron (hole) carrier density nn (pp), and |p~ncnvk|2|\widetilde{p}_{n_{c}n_{v}\scalebox{0.7}{{k}}}|^{2} is the (squared) polarization-averaged momentum matrix element between conduction and valence bands ncn_{c} and nvn_{v} at wave vector k. The Dirac delta distribution in Eq. (2), which imposes conservation of energy, is replaced by a hyperbolic secant lineshape. Marko et al. (2016) BB is computed via the integrated SE rate Landsberg (1991); Murphy et al. (2024)

B=1nprsp(ω)d(ω),B=\frac{1}{np}\int r_{\scalebox{0.7}{\text{sp}}}(\hbar\omega)\,\text{d}(\hbar\omega)\,, (3)

where we impose net charge neutrality, n=pn=p.

Accurate evaluation of FeF_{e}, FhF_{h} and rsp(ω)r_{\scalebox{0.7}{\text{sp}}}(\hbar\omega) mandates dense sampling of BZ regions occupied by electrons and holes; we employ the adaptive k-point sampling approach of Ref. Zhang et al., 2018. In unstrained 2H-Ge carriers occupy states in the immediate vicinity of Γ\Gamma only. We replace a Γ\Gamma-centred portion of the coarse MP grid employed in the TB-mBJ electronic structure calculations – extending to 25% (50%) of the BZ extent parallel (perpendicular) to the [0001] plane – by dense MP grids containing up to 55×55×3355\times 55\times 33 k-points. Under [0001] tension, the downward shift of the U valleys can result in an appreciable fraction of injected electrons occupying states at and close to the BZ edge (cf. Fig. 3). We augment the adaptive grid employed for unstrained 2H-Ge by also replacing k points – lying within 50% of the hexagonal BZ side length of the U direction in the (0001) plane, and along the full extent of the BZ parallel to [0001] – by the k points lying within that region from Γ\Gamma-centred MP grids containing up to 72×72×4472\times 72\times 44 k-points. Using the Kohn-Sham wave functions computed on the coarse MP grid as input, ϵnk\epsilon_{{n}\scalebox{0.7}{{k}}} and |p~ncnvk|2|\widetilde{p}_{n_{c}n_{v}\scalebox{0.7}{{k}}}|^{2} are calculated non-self-consistently on the adaptive grid. The integral in Eq. (2) is then evaluated via a weighted sum over the adaptive k-point grid. Zhang et al. (2018)

Author contributions

This study was devised by C.A.B., with input from C.G.VdW. and X.Z. All calculations were performed by C.A.B, with technical support from X.Z. and M.E.T., and supervised by C.G.VdW. The writing of the manuscript was led by C.A.B., with contributions from all authors.

Acknowledgements

C.A.B. was supported by the European Commission’s Horizon 2020 research and innovation program via a Marie Skłodowska-Curie Actions Individual Fellowship (H2020-MSCA-IF; Global Fellowship “SATORI”, grant agreement no. 101030927). M.E.T. and C.G.VdW. were supported by the U.S. Department of Energy (DOE; grant no. DESC0010689). X.Z. was supported by the National Natural Science Foundation of China (NSFC; grant no. 52172136). This work used resources of the National Energy Research Scientific Computing Center (NERSC; award no. BES-ERCAP0021021), a DOE Office of Science User Facility supported by the Office of Science of the U.S. DOE (contract no. DE-AC02-05CH11231). C.A.B. thanks Prof. Jos E. M. Haverkort (TU Eindhoven, Netherlands) for useful discussions.

Data access statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • Zhou et al. (2023) Z. Zhou, X. Ou, Y. Fang, E. Alkhazraji, R. Xu, Y. Wan, and J. E. Bowers, eLight 3, 1 (2023).
  • Geiger et al. (2015) R. Geiger, T. Zabel, and H. Sigg, Front. Mater. 2, 52 (2015).
  • Hauge et al. (2017) H. I. T. Hauge, S. Conesa-Boj, M. A. Verheijen, S. Koelling, and E. P. A. M. Bakkers, Nano Lett. 17, 85 (2017).
  • De and Pryor (2014) A. De and C. E. Pryor, J. Phys.: Condens. Matter 26, 145801 (2014).
  • Rödl et al. (2019) C. Rödl, J. Furthmüller, J. R. Suckert, V. Armuzza, F. Bechstedt, and S. Botti, Phys. Rev. Materials 3, 034602 (2019).
  • Fadaly et al. (2020) E. M. T. Fadaly, A. Dijkstra, J. R. Suckert, D. Ziss, M. A. J. van Tilburg, C. Mao, Y. Ren, V. T. van Lange, K. Korzun, S. Kölling, et al., Nature 580, 205 (2020).
  • van Tilburg et al. (2024) M. A. J. van Tilburg, R. Farina, V. T. van Lange, W. H. J. Peeters, S. Meder, M. M. Jansen, M. A. Verheijen, M. Vettori, J. J. Finley, E. P. A. M. Bakkers, et al., Commun. Phys. 7, 328 (2024).
  • Suckert et al. (2021) J. R. Suckert, C. Rödl, J. Furthmüller, F. Bechstedt, and S. Botti, Phys. Rev. Materials 5, 024602 (2021).
  • Thurmond (1975) C. D. Thurmond, J. Electrochem. Soc. 122, 1133 (1975).
  • Landsberg (1991) P. T. Landsberg, Recombination in Semiconductors (Cambridge University Press, Cambridge, 1991).
  • Vurgaftman and Meyer (2001) I. Vurgaftman and J. R. Meyer, J. Appl. Phys. 89, 5815 (2001).
  • Hader et al. (2022) J. Hader, S. C. Liebscher, J. V. Moloney, and S. W. Koch, Appl. Phys. Lett. 121, 192103 (2022).
  • van Tilburg et al. (2023) M. A. J. van Tilburg, W. H. J. Peeters, M. Vettori, V. T. van Lange, E. P. A. M. Bakkers, and J. E. M. Haverkort, J. Appl. Phys. 133, 065702 (2023).
  • Hader et al. (2005) J. Hader, J. V. Moloney, and S. W. Koch, Appl. Phys. Lett. 87, 201112 (2005).
  • Hader and Moloney (2006) J. Hader and J. V. Moloney, Proc. SPIE 6115, 61151T (2006).
  • Grein et al. (2002) C. H. Grein, M. E. Flatté, J. T. Olesberg, S. A. Anson, L. Zhang, and T. F. Boggess, J. Appl. Phys. 92, 7311 (2002).
  • van Lange et al. (2024) V. T. van Lange, A. Dijkstra, E. M. T. Fadaly, W. H. J. Peeters, M. A. J. van Tilburg, E. P. A. M. Bakkers, F. Bechstedt, J. J. Finley, and J. E. M. Haverkort, ACS Photonics 11, 4258 (2024).
  • Willem-Jan Berghuis et al. (2024) W. J. H. Willem-Jan Berghuis, M. A. J. van Tilburg, W. H. J. Peeters, V. T. van Lange, R. Farina, E. M. T. Fadaly, E. C. M. Renirie, R. J. Theeuwes, M. A. Verheijen, B. Macco, et al., ACS Appl. Nano Mater. 7, 2343 (2024).
  • Fadaly et al. (2021) E. M. T. Fadaly, A. Marzegalli, Y. Ren, L. Sun, A. Dijkstra, D. de Matteis, E. Scalise, A. Sarikov, M. De Luca, R. Rurali, et al., Nano Lett. 21, 3619 (2021).
  • Imhof et al. (2010) S. Imhof, A. Thränhardt, A. Chernikov, M. Koch, N. S. Köster, K. Kolata, S. Chatterjee, S. W. Koch, X. Lu, S. R. Johnson, et al., Appl. Phys. Lett. 96, 131115 (2010).
  • Zhang et al. (2009) F. Zhang, V. H. Crespi, and P. Zhang, Phys. Rev. Lett. 102, 156401 (2009).
  • Blöchl (1994) P. E. Blöchl, Phys. Rev. B 50, 17953 (1994).
  • Kresse and Joubert (1999) G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
  • Tran and Blaha (2009) F. Tran and P. Blaha, Phys. Rev. Lett. 226401, 102 (2009).
  • Shen et al. (2019) J.-X. Shen, D. Steiauf, A. McAllister, G. Shi, E. Kioupakis, A. Janotti, and C. G. Van de Walle, Phys. Rev. B 100, 155202 (2019).
  • Gajdoš et al. (2006) M. Gajdoš, K. Hummer, G. Kresse, J. Furthmüller, and F. Bechstedt, Phys. Rev. B 73, 045112 (2006).
  • Chang et al. (1995) C.-S. Chang, S. L. Chuang, J. R. Minch, W.-C. W. Fang, Y. K. Chen, and T. Tanbun-Ek, IEEE J. Sel. Topics Quantum Electron. 1, 1100 (1995).
  • Marko et al. (2016) I. P. Marko, C. A. Broderick, S.-R. Jin, P. Ludewig, W. Stolz, K. Volz, J. M. Rorison, E. P. O’Reilly, and S. J. Sweeney, Sci. Rep. 6, 28863 (2016).
  • Murphy et al. (2024) C. Murphy, E. P. O’Reilly, and C. A. Broderick, J. Phys. D: Appl. Phys. 57, 035103 (2024).
  • Zhang et al. (2018) X. Zhang, J.-X. Shen, W. Wang, and C. G. Van de Walle, ACS Energy Lett. 3, 2329 (2018).