Direct bandgap group IV semiconductors, like strained Ge, GeSn, or hexagonal SiGe, are considered promising for photonic integration on silicon. For group IV semiconductor lasers, it is crucial to understand the carrier cooling efficiency toward the band edges. From a fundamental perspective, a study of carrier cooling within the Γ-valley of direct bandgap group IV semiconductors is particularly interesting since the Fröhlich interaction is expected to be very weak or even absent in these materials due to the nonpolar lattice. Intravalley carrier relaxation within the Γ-valley of a nonpolar semiconductor has not been experimentally accessible before since it has always been overshadowed by intervalley processes between energetically close indirect conduction band minima. Here, we study carrier cooling in direct bandgap hexagonal silicon-germanium (hex-SiGe) nanowires, allowing us to study carrier cooling in an isolated Γ-valley that is sufficiently separated from the indirect minima. We obtain a hot carrier cooling time of 180 ps in the Γ-valley of hex-SiGe. Although the cooling is much slower than in bulk polar group III/V materials due to the absence of Fröhlich interaction, it is comparable to the cooling time in an InGaAs MQW laser structure. We conclude that carrier cooling does not inherently limit hex-SiGe to serve as a laser gain material. This result is an important insight into the field of group IV semiconductor lasers.

The monolithic integration of a silicon-based laser onto a silicon electronic chip has been a longstanding research subject. Unfortunately, silicon has an indirect bandgap, impeding efficient light emission and lasing. To circumvent the lack of a silicon laser, a lot of work has been directed toward integrating direct bandgap group III–V materials on silicon.1 An emerging technology is to exploit metastable, direct bandgap group IV materials, such as strained germanium (Ge),2 germanium tin (GeSn),3–9 or hexagonal (lonsdaleite) silicon-germanium (hex-SiGe),10 as the light-emitting active material. However, efficient light emission and lasing also require efficient carrier cooling toward the band edges by inelastic carrier-optical phonon scattering.9,11

In polar III/V semiconductors, like gallium arsenide, carrier relaxation is dominated by the Fröhlich interaction through emission of small wave vector optical phonons, as displayed in Fig. 1(a). The Fröhlich interaction drives carrier cooling within a conduction band (CB) or valence band (VB) valley, referred to as intravalley scattering. It has been studied extensively in group III/V semiconductors and is found to be a very efficient carrier cooling process.12–14 Conventional group IV semiconductors like cubic germanium (Ge) are nonpolar.11,15,16 As such, due to the absence of Fröhlich interaction, carrier relaxation in the CB is primarily governed by intervalley deformation potential scattering (intervalley DPS), as displayed in Fig. 1(b).12 After the electrons have scattered from the Γ-valley to the indirect X- and L-valleys, large wavevector intervalley DPS between the different X- and L-valleys results in efficient electron cooling on a timescale of hundreds of femtoseconds to a few picoseconds,17–20 where the hole temperature follows the electron temperature by fast electron–hole scattering.13,17 For indirect bandgap semiconductors, the small amount of intravalley carrier relaxation within the Γ-valley has been overshadowed by ultrafast intervalley DPS and could not be experimentally assessed.

FIG. 1.

Carrier relaxation processes in (a) polar direct bandgap semiconductors where polar optical phonon intravalley scattering (green arrows) through the Fröhlich interaction is dominant, (b) nonpolar indirect bandgap semiconductors where deformation potential intervalley scattering is most important, (c) nonpolar direct bandgap semiconductors with a directness of ΔEΓU. The relaxation starts by intervalley processes (yellow arrows), but it should be followed by intravalley cooling within the Γ-valley (light green arrows), which is the subject of this paper.

FIG. 1.

Carrier relaxation processes in (a) polar direct bandgap semiconductors where polar optical phonon intravalley scattering (green arrows) through the Fröhlich interaction is dominant, (b) nonpolar indirect bandgap semiconductors where deformation potential intervalley scattering is most important, (c) nonpolar direct bandgap semiconductors with a directness of ΔEΓU. The relaxation starts by intervalley processes (yellow arrows), but it should be followed by intravalley cooling within the Γ-valley (light green arrows), which is the subject of this paper.

Close modal

In direct bandgap group IV semiconductors like GeSn and hex-SiGe, as displayed in Fig. 1(c), the carriers are first very rapidly cooled by intervalley scattering by the emission of zone edge phonons. In these semiconductors, intervalley scattering is however no longer adequate for carrier cooling toward the bottom of the CB. Here, intravalley scattering within an isolated Γ-valley is the only option to cool down from the minimum of the lowest indirect valley toward the global minimum in the Γ-valley, and likewise in the VB.21 The intravalley cooling is expected to occur by emission of long-wavelength optical phonons. In hex-SiGe alloys, the phonon energies have been investigated using Raman spectroscopy, resulting in phonon energies between 35 and 57 meV for hex-Ge and hex-Si, respectively.22 Carrier cooling in such an isolated Γ-valley in a group IV semiconductor has not been studied yet. Hex-SiGe is unique as a direct bandgap group IV semiconductor with a large directness (ΔEΓ-U), which is defined as the energy separation between the lowest indirect minimum and the global minimum. As an added advantage, its directness can be tuned by the alloy composition (up to 315 meV for hex-Ge), without using external strain.10 

Experimental study on the carrier relaxation dynamics in group IV semiconductor nanowires emitting in the mid-infrared requires measurements of the hot carrier distributions as a function of time. This can be performed using either two-wavelength time-resolved pump–probe techniques or picosecond time-resolved measurement of the temporal evolution of the complete photoluminescence (PL) spectrum.20,23 Although PL lifetime measurements on GeSn have been reported as a function of emission energy,9,24,25 simultaneous measurements of the time-resolved and spectrally resolved PL, allowing to measure the carrier cooling time, are currently lacking.

We performed time-resolved micro-PL measurements on the as-grown core shell hex-SiGe nanowire arrays. Specifically, we study the carrier cooling time in a hex-Si0.32Ge0.68 nanowire (NW) sample with a NW diameter (length) of 2.4 μm (8.4 μm) and in a hex-Si0.2Ge0.8 NW sample with a NW diameter (length) of 1.3 μm (9.2 μm). The nanowires are spaced with a pitch of 4μm in a square pattern, as shown in the Scanning Electron Microscopy (SEM) images shown in the supplementary material A. The samples are grown using the crystal transfer method with wurtzite (WZ) GaAs NWs with a diameter of 180 nm, as a template by metal-organic vapor phase epitaxy (MOVPE).26,27 The details of the gold-catalyzed WZ-GaAs NW growth are given in Fadaly et al.10 The hex-SiGe shells are grown around the core NWs by MOVPE at a substrate temperature of 595 °C. The diameters of both nanowire samples are suitable to strongly confine the fundamental modes, as shown in Appendix E, resulting in a similar absorbed power density per area for both samples.

The time-resolved micro-PL setup combines a Fourier Transform Infrared (FTIR) spectrometer with a superconducting nanowire single photon detector (SNSPD). This method allows for experimental observation of the time evolution of the complete PL spectrum, which is essential to measure the carrier cooling dynamics. Our experimental setup thus features the same functionality as a streak camera but in the experimentally difficult-to-access near-to-mid-infrared (NMIR) spectral region. An NKT photonics Origami pulsed fiber laser at 1030 nm (1.2 eV), with a pulse length of 200 fs at a repetition frequency of 40 MHz, is used as an optical pump source. A reflective Cassegrain objective (36x, NA = 0.52) focuses the laser beam on the sample to an excitation spot with a diameter of approximately 4 μm, exciting only a single standing nanowire from the top. The objective is also used for PL collection.

The sample is mounted on a helium-cooled cryostat and all measurements are performed at a temperature of 6 K. To spectrally resolve the emission, the light is directed through a Nireos GEMINI interferometer to perform FTIR spectroscopy with a spectral resolution of 3 meV. The light is subsequently focused into a single-mode fiber that guides the light to the Single Quantum Eos single SNSPD with >15% quantum efficiency at a wavelength of 2000 nm. The detector is sensitive up to a wavelength of 2300 nm (0.54 eV) with a time jitter of 22 ps, limiting our temporal resolution. Both the excitation laser and SNSPD are connected to a Picoharp 300 time-correlated single photon counter (TCSPC) with a timing jitter of only 4 ps, allowing the recording of the time-resolved PL spectral map as shown in Fig. 2(a). By taking line cuts along the energy axis of the time-resolved PL map, we obtain the temporal response of the hex-Si0.32Ge0.68 sample, which is plotted in Fig. 2(b). We observe a shortening of the lifetime at increasing energy above the bandgap (0.56 eV), which is a first indication for hot carrier cooling toward the material bandgap. The time-resolved spectral map shown in Fig. 2(a) also allows us to obtain the time-resolved PL (TRPL) spectra by line cuts along the time axis. The TRPL spectra for hex-Si0.2Ge0.8 and hex-Si0.32Ge0.68 nanowires are shown in Figs. 3(a) and 3(b), respectively, as a function of the elapsed time. The narrowing of the high-energy tails of these PL spectra with time directly shows the cooling of carrier temperature. To quantitatively obtain the carrier temperature as a function of time, we fit the PL spectral line shape with a Lasher–Stern–Würfel (LSW) model (see supplementary material B1),28–31 which reads
(1)
FIG. 2.

(a) Time-resolved spectral PL map from the hex-Si0.32Ge0.68 sample. (b) Line cuts (dashed lines) along the energy axis provide the PL time decay curves at a specific emission energy. Here, t = 0 is defined as the time at which the PL intensity is maximum. The instrument response function (IRF) is indicated by a gray line.

FIG. 2.

(a) Time-resolved spectral PL map from the hex-Si0.32Ge0.68 sample. (b) Line cuts (dashed lines) along the energy axis provide the PL time decay curves at a specific emission energy. Here, t = 0 is defined as the time at which the PL intensity is maximum. The instrument response function (IRF) is indicated by a gray line.

Close modal
FIG. 3.

(a) and (b) Normalized TRPL spectra as a function of time (dark to light) recorded from an ensemble of nanowires with two distinct alloy compositions of hex-Si0.2Ge0.8 (blue) in panel (a), and hex-Si0.32Ge0.68 (green) in panel (b), at a laser excitation of 3.54mJ/cm2 and at a lattice temperature of 6 K. The LSW fits are indicated by red dashed lines and provide the carrier temperature (TC) at every time delay. (c) Schematic band structure of hex-Si0.32Ge0.68 (green) and hex-Si0.2Ge0.8 (blue) highlighting the indirect minima around the M and U symmetry points, based on Borlido et al.32 After excitation at 1.2 eV (yellow dashed arrow), the carriers relax to a state ΔEΓ-U above the Γ-minimum by intervalley carrier scattering (yellow arrows). Hex-Si0.2Ge0.8 features a larger directness, ΔEΓ-U, than hex-Si0.32Ge0.68.

FIG. 3.

(a) and (b) Normalized TRPL spectra as a function of time (dark to light) recorded from an ensemble of nanowires with two distinct alloy compositions of hex-Si0.2Ge0.8 (blue) in panel (a), and hex-Si0.32Ge0.68 (green) in panel (b), at a laser excitation of 3.54mJ/cm2 and at a lattice temperature of 6 K. The LSW fits are indicated by red dashed lines and provide the carrier temperature (TC) at every time delay. (c) Schematic band structure of hex-Si0.32Ge0.68 (green) and hex-Si0.2Ge0.8 (blue) highlighting the indirect minima around the M and U symmetry points, based on Borlido et al.32 After excitation at 1.2 eV (yellow dashed arrow), the carriers relax to a state ΔEΓ-U above the Γ-minimum by intervalley carrier scattering (yellow arrows). Hex-Si0.2Ge0.8 features a larger directness, ΔEΓ-U, than hex-Si0.32Ge0.68.

Close modal

In this equation, IPL(E) is the PL spectrum, Eg is the bandgap energy, χ is a scaling factor, and Δμ is the chemical potential describing the shift in quasi-Fermi levels by the photoexcited carrier densities. By fitting the LSW model to the time-resolved PL spectra, as shown by the red dashed curves in Figs. 3(a) and 3(b), we obtain the carrier temperature TC as a function of the elapsed time after optical excitation. Since we are interested in the carrier cooling dynamics after excitation, we define the initial time t = 0 at the maximum of the integrated PL-decay curve.

We focus on two different hex-SiGe alloy compositions, featuring different ΔEΓ-U. A schematic of the approximate band structure and possible carrier cooling pathways is shown in Fig. 3(c), showing the three lowest CB minima, at the Γ-, M-, and U-points, for both alloy compositions, based on Borlido et al.32 The hex-Si0.2Ge0.8 alloy (blue) has a large directness of ∼140 meV (Ref. 32) and has a bandgap of 0.56 eV. Intervalley scattering will initially cool down the carriers toward (1 optical phonon energy above) the bottom of the U-valley.17–20 This relaxation is too fast to be measured in our setup and thus takes place before the onset of our measurement. The carriers are subsequently scattered by an optical phonon emission toward high energy Γ-band states (yellow arrow) at a net excess energy of approximately ΔEΓ-U above the CB minimum. The carrier distribution within the Γ-valley is experimentally observed by the broad initial PL emission spectrum in Fig. 3(a). Its width of >300 meV corresponds to a carrier temperature of 680 K.

The hex-Si0.32Ge0.68 alloy (green) has a low directness of approximately 70 meV (Ref. 32) and a bandgap of 0.61 eV. Likewise, the carriers are rapidly cooled down by intervalley scattering between the indirect valleys.17–20 The fast intervalley scattering processes [Fig. 3(c), yellow arrow] are expected to subsequently scatter the carriers into the central Γ-valley with a net energy ΔEΓ-U, only slightly above the CB minimum, requiring a smaller amount of intra-Γ-valley cooling. This is experimentally confirmed by the much lower initial carrier temperature Tc(0) of 280 K, as shown in Fig. 3(b). We thus observe a large difference in the initial Tc(0) of roughly 400 K between the different hex-SiGe alloys.

By fitting the TRPL spectra in Figs. 3(a) and 3(b) with the LSW model, we obtain carrier cooling curves Tc(t) for both hex-Si0.2Ge0.8 and hex-Si0.32Ge0.68, which are displayed in Figs. 4(a) and 4(b). To obtain relevant carrier cooling times for a hex-SiGe nanowire laser, we used high excitation fluences of 5.31, 3.54, and 1.77 mJ/cm2, reasonably above the gain threshold for typical semiconductor nanowire lasers.33–36 At an elapsed time shorter than 0.15 ns, the emission spectrum of hex-Si0.2Ge0.8 shows a second peak centered around  0.82 eV [visible at t = 0 ns in Fig. 3(a)]. As this peak affects the Fermi–Dirac shape of the high energy side of the spectrum, it results in an increase in uncertainty of Tc, as displayed in Fig. 4(a). The energy of this peak is in the vicinity of the second valence band transition. However, the appearance of the peak is more pronounced at low excitation powers and is not evident in the spectra of hex-Si0.32Ge0.68. The time-resolved spectral PL maps of all measurements represented in Figs. 4(a) and 4(b) are displayed in supplementary material C.

FIG. 4.

(a) and (b) Carrier cooling curves for hex-Si0.2Ge0.8 and for hex-Si0.32Ge0.68 at laser excitation fluences of 5.31, 3.54, and 1.77mJ/cm2 for both alloy compositions. The fit [Eq. (2)], indicated by orange dashed lines, provides the carrier cooling time (τ0) for each measurement. (c) Observed carrier cooling times (τ0) as a function of the excitation fluence. We added the carrier cooling times of an InGaAs/InP multi-quantum well (see supplementary material D).

FIG. 4.

(a) and (b) Carrier cooling curves for hex-Si0.2Ge0.8 and for hex-Si0.32Ge0.68 at laser excitation fluences of 5.31, 3.54, and 1.77mJ/cm2 for both alloy compositions. The fit [Eq. (2)], indicated by orange dashed lines, provides the carrier cooling time (τ0) for each measurement. (c) Observed carrier cooling times (τ0) as a function of the excitation fluence. We added the carrier cooling times of an InGaAs/InP multi-quantum well (see supplementary material D).

Close modal
The carrier cooling curves Tc(t), which are governed by slow intra-Γ-valley cooling, are subsequently fitted with a mono-exponential decay function (see supplementary material B2),
(2)
indicated by orange dashed lines in Figs. 4(a) and 4(b), to extract the carrier cooling time (τ0).14  TL represents the lattice temperature. After approximately 1 ns, Tc(t) reaches a saturation value of 250 K in hex-Si0.2Ge0.8 and 125 K in hex-Si0.32Ge0.68, respectively, which we will refer to as the plateau temperature. In all measurements, similar plateau temperatures are observed without a strong dependence on excitation fluence or lattice temperature. We attribute this plateau to a combination of inhomogeneous broadening of the PL spectrum originating from alloy broadening, lattice heating, or material defects. These defects, which presumably give rise to the plateau temperature, are the I3 defects.37,38 In this case, quantum confinement due to type I band alignment in the vertical direction between the cubic insertion layers and the hexagonal lattice is expected to broaden PL emission toward the high energy side. As the plateau temperature originates from an inhomogeneous broadening mechanism unrelated to the actual carrier temperature, we leave the TL as a free parameter in the fit and restrict the fitting range to the times before the temperature reaches the plateau temperature. We emphasize that, although these effects might influence the carrier temperature when approaching the plateau temperature, these effects do not affect the cooling during the first tens of picoseconds and are thus outside the scope of this research. Detailed results of the fitting parameters from Eq. (2) are given in Appendix B.

The measured carrier cooling times (τ0) are plotted in Fig. 4(c) as a function of the excitation fluence. For both alloys, we obtain a carrier cooling time of roughly 180 ps at an excitation density of 3.5 mJ/cm2. The similar carrier cooling time observed for both alloy compositions provides evidence that we indeed measure the fundamental intravalley carrier cooling mechanism in hex-SiGe. As expected from the absence of Fröhlich interaction, the intravalley carrier relaxation in the direct Γ-valley hex-SiGe is clearly slower than in III–V semiconductors.14 

The influence of hot phonon effects can be experimentally assessed by observing whether τ0 increases with excitation fluence due to the presence of a hot phonon population. This heats the carriers by phonon absorption processes and thus limits the cooling rate of the carriers.39,40 As shown in Fig. 4(c), we only observe a very moderate increase in the carrier cooling time in our hex-SiGe nanowire alloys, suggesting that phonon bottleneck effects in hex-SiGe are less important compared to, e.g., III/V multiple-quantum well (MQW) structures where hot phonon effects are stronger.41–43 This is consistent with our theoretical expectations that, due to the larger number of optical modes in hexagonal crystals,22 a strong phonon bottleneck effect is absent in hex-SiGe.42,44

To establish whether the measured carrier cooling time in hex-SiGe is still acceptable for lasing, we compare our results with an InGaAs/InGaAsP MQW laser structure in Fig. 4(c) (see supplementary material D). These MQW samples are emitting in the same spectral region as our hex-SiGe samples and are measured with the same experimental setup. Importantly, we observe similar order of magnitude carrier cooling times for the InGaAs MQW reference sample as obtained for hex-SiGe. This can be understood by the large susceptibility of MQWs for hot phonon effect,41–43 as observed in supplementary material Sec. D. The similarity of the carrier cooling time in our nonpolar hex-SiGe nanowires as compared to a conventional III/V InGaAs MQW laser structure provides evidence that, from the carrier cooling perspective, the nonpolar nature of our hex-SiGe nanowires does not provide a major obstacle for lasing.

The mechanism for carrier cooling in an isolated Γ-valley, with large ΔEΓ-U, is still unknown. Using the current experimental method, no distinction can be made between specific scattering processes. Three potential mechanisms are postulated: (i) deformation potential scattering (DPS) in the VB, (ii) DPS in the CB, and (iii) polar optical phonon scattering due to the very small polarity of the Si-Ge bond.

(i) Initially, the dominant cooling mechanism is due to intervalley scattering of electrons that cool down in the indirect valleys and subsequently scatter toward the central Γ-valley. Once the electrons have arrived in the central Γ-valley, optical phonon emission by DPS in the VB probably becomes the dominant cooling mechanism.45 If this mechanism is dominant, electron cooling will follow the cooling of holes by electron–hole scattering. We note that this relaxation mechanism is expected to be equally valid in other direct bandgap group IV semiconductors like GeSn or SiGeSn.

(ii) For both strained Ge and GeSn, the intravalley-Γ-valley DPS scattering is forbidden due to the s-like characteristic of the CB.45 For hex-SiGe, the intra-Γ-valley DPS is not symmetry forbidden, as the lowest CB of hex-SiGe in the Γ-valley is of hybridized sp-type character.46 

(iii) We cannot completely exclude polar optical phonon scattering due to the small remaining polarity in either GeSn or hex-SiGe. While the Si-Si and the Ge-Ge bonds in hex-SiGe are completely nonpolar, the randomly distributed Si-Ge bonds in the hex-SiGe alloy still give rise to a slight polarity due to their small electronegativity difference (Δχr). The effective average electronegativity difference between the individual atoms within the alloy is Δχr,eff = 0.02 and Δχr,eff = 0.03 for hex-Si0.2Ge0.8 and hex-Si0.32Ge0.68, respectively.47 So, we cannot completely exclude a small amount of polar optical phonon emission by the Fröhlich interaction. However, even though there is a difference in polarity, there is no significant difference in carrier cooling rate between the two alloys, suggesting that the Fröhlich interaction is not dominant. Consequently, nonpolar intra-Γ-valley DPS within the VB Γ-valley, or possibly also within the CB Γ-valley, is the most likely mechanism responsible for the observed carrier cooling time of 180 ps, but definite conclusions can only be drawn by a future theoretical investigation, which is beyond the scope of the present paper. Further experimental research could be performed comparing the carrier cooling behavior of hex-SiGe to other direct bandgap group IV materials to determine the effect of sp hybridization in the CB and possible effects on their technological relevance for Si photonics. To exclude the contribution of any remaining Fröhlich interaction, pure hex-Ge can be investigated, which is identical to hex-SiGe in regard to sp character and the large directness of the CB. This is presently not possible using the technique presented in this work due to limitations of the detector's measurement window.

We finally note that the intravalley scattering in hex-SiGe might be different from its cubic counterparts since the hexagonal unit cell is two times larger than the cubic unit cell. Consequently, the reciprocal unit cell of hex-SiGe is twice as small as that of its cubic counterpart, resulting in backfolding of phonon modes.48 The backfolding introduces several additional phonon modes at the Γ-point. Furthermore, in hex-SiGe, the Si-Si, the Si-Ge, and the Ge-Ge bonds result in slightly different phonon energies ( 40 meV).22 A theoretical calculation might provide more insight into the role of these folded phonon modes for carrier cooling dynamics.

In conclusion, our hex-SiGe nanowires feature a large directness, allowing us to measure the carrier cooling dynamics in an isolated Γ-valley. We experimentally observe a hot carrier cooling time within the Γ-valley of approximately 180 ps. The slow cooling time implies that the Fröhlich interaction is either absent or strongly reduced. In conventional III/V semiconductors, carrier relaxation is efficient due to polar optical phonon scattering, mediated by the Fröhlich interaction. The absence of efficient Fröhlich and intervalley scattering within an isolated Γ-valley of a nonpolar direct bandgap semiconductor might provide a bottleneck for efficient group IV semiconductor lasers. However, we emphasize that the carrier cooling time in our hex-SiGe nanowires is of comparable magnitude to the carrier cooling time in an InGaAs multi-quantum well laser structure. We thus conclude that the near absence of Fröhlich interaction in direct bandgap hex-SiGe does not inherently limit its application potential for hex-SiGe semiconductor lasers or hex-SiGe semiconductor optical amplifiers.

See the supplementary material for SEM images of the nanowire samples, LSW and carrier lifetime fits, time-resolved spectral PL maps, and carrier cooling in an InGaAs/InGaAsP multi-quantum well laser structure.

This project received funding from the European Union's Horizon 2020 research and innovation program under Grant Agreement No. 964191 (Opto Silicon), the Dutch Organization for Scientific Research (NWO) in the Zwaartekracht Project (Grant No. 024.002.033), and the Mat4Sus project (Grant No.739.017.002).

The authors have no conflicts to disclose.

M. F. Schouten and M. A. J. van Tilburg contributed equally to this work.

M. F. Schouten: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (supporting); Writing – original draft (equal); Writing – review & editing (equal). M. A. J. van Tilburg: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (lead); Writing – original draft (equal); Writing – review & editing (equal). V. T. van Lange: Formal analysis (supporting); Investigation (supporting); Methodology (supporting). W. H. J. Peeters: Resources (lead). R. Farina: Investigation (supporting); Methodology (supporting); Resources (supporting). M. M. Jansen: Resources (supporting); Supervision (supporting); Writing – review & editing (supporting). M. Vettori: Resources (supporting); Supervision (supporting); Writing – review & editing (supporting). E. P. A. M. Bakkers: Investigation (supporting); Methodology (supporting); Supervision (supporting); Writing – review & editing (supporting). J. E. M. Haverkort: Conceptualization (equal); Resources (supporting); Supervision (lead); Writing – review & editing (equal).

The data that support the findings of this study are openly available in Zenodo at https://doi.org/10.5281/zenodo.10891385, Ref. 49.

1.
Z.
Wang
,
B.
Tian
,
M.
Pantouvaki
,
W.
Guo
,
P.
Absil
,
J.
Van Campenhout
,
C.
Merckling
, and
D.
Van Thourhout
,
Nat. Photonics
9
,
837
(
2015
).
2.
F. T.
Armand Pilon
,
A.
Lyasota
,
Y. M.
Niquet
,
V.
Reboud
,
V.
Calvo
,
N.
Pauc
,
J.
Widiez
,
C.
Bonzon
,
J. M.
Hartmann
,
A.
Chelnokov
,
J.
Faist
, and
H.
Sigg
,
Nat. Commun.
10
,
2724
(
2019
).
3.
S.
Wirths
,
R.
Geiger
,
N.
von den Driesch
et al,
Nat. Photon
9
,
88
92
(
2015
).
4.
R.
Soref
,
D.
Buca
, and
S.-Q.
Yu
,
Opt. Photonics News
27
,
32
(
2016
).
5.
A.
Elbaz
,
D.
Buca
,
N.
von den Driesch
,
K.
Pantzas
,
G.
Patriarche
,
N.
Zerounian
,
E.
Herth
,
X.
Checoury
,
S.
Sauvage
,
I.
Sagnes
,
A.
Foti
,
R.
Ossikovski
,
J. M.
Hartmann
,
F.
Boeuf
,
Z.
Ikonic
,
P.
Boucaud
,
D.
Grützmacher
, and
M.
El Kurdi
,
Nat. Photonics
14
,
375
(
2020
).
6.
S.
Amoah
,
S.
Ojo
,
H.
Tran
,
G.
Abernathy
,
Y.
Zhou
,
W.
Du
,
J.
Margetis
,
J.
Tolle
,
B.
Li
, and
S. Q.
Yu
, in
Conference on Lasers and Electro-Optics (CLEO 2021)
(
2021
).
7.
D.
Buca
,
A.
Bjelajac
,
D.
Spirito
,
O.
Concepción
,
M.
Gromovyi
,
E.
Sakat
,
X.
Lafosse
,
L.
Ferlazzo
,
N.
von den Driesch
,
Z.
Ikonic
,
D.
Grützmacher
,
G.
Capellini
, and
M. E.
Kurdi
,
Adv. Opt. Mater.
10
,
2201024
(
2022
).
8.
D.
Stange
,
N.
Von Den Driesch
,
T.
Zabel
,
F.
Armand-Pilon
,
D.
Rainko
,
B.
Marzban
,
P.
Zaumseil
,
J. M.
Hartmann
,
Z.
Ikonic
,
G.
Capellini
,
S.
Mantl
,
H.
Sigg
,
J.
Witzens
,
D.
Grützmacher
, and
D.
Buca
,
ACS Photonics
5
,
4628
(
2018
).
9.
E.
Rogowicz
,
J.
Kopaczek
,
J.
Kutrowska-Girzycka
,
M.
Myronov
,
R.
Kudrawiec
, and
M.
Syperek
,
ACS Appl. Electron. Mater.
3
,
344
(
2021
).
10.
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
,
M. A.
Verheijen
,
D.
Busse
,
C.
Rödl
,
J.
Furthmüller
,
F.
Bechstedt
,
J.
Stangl
,
J. J.
Finley
,
S.
Botti
,
J. E. M.
Haverkort
, and
E. P. A. M.
Bakkers
,
Nature
580
,
205
(
2020
).
11.
C.
Ciano
,
L.
Persichetti
,
M.
Montanari
,
L.
Di Gaspare
,
G.
Capellini
,
L.
Baldassarre
,
M.
Ortolani
,
A.
Pashkin
,
M.
Helm
,
S.
Winnerl
,
M.
Virgilio
, and
M.
De Seta
,
Phys. Rev. B
102
,
205302
(
2020
).
12.
J.
Shah
,
Ultrafast Spectroscopy of Semiconductors and Semiconductor Nanostructures
(
Gmbh, Springer-Verlag
,
Berlin Heidelberg
,
1999
).
13.
D. K.
Ferry
,
Appl. Phys. Rev.
8
,
021324
(
2021
).
14.
Y.
Zhang
,
M. J. Y.
Tayebjee
,
S.
Smyth
,
M.
Dvořák
,
X.
Wen
,
H.
Xia
,
M.
Heilmann
,
Y.
Liao
,
Z.
Zhang
,
T.
Williamson
,
J.
Williams
,
S.
Bremner
,
S.
Shrestha
,
S.
Huang
,
T. W.
Schmidt
, and
G. J.
Conibeer
,
Appl. Phys. Lett.
108
,
131904
(
2016
).
15.
G.
Mak
and
W. W.
Rühle
,
Phys. Rev. B
52
,
584
(
1995
).
16.
M.
Zürch
,
H. T.
Chang
,
L. J.
Borja
,
P. M.
Kraus
,
S. K.
Cushing
,
A.
Gandman
,
C. J.
Kaplan
,
M. H.
Oh
,
J. S.
Prell
,
D.
Prendergast
,
C. D.
Pemmaraju
,
D. M.
Neumark
, and
S. R.
Leone
,
Nat. Commun.
8
,
15734
(
2017
).
17.
H.
Roskos
,
B.
Rieck
,
A.
Seilmeier
, and
W.
Kaiser
,
Appl. Phys. Lett.
53
,
2406
(
1988
).
18.
X. Q.
Zhou
,
H. M.
Van Driel
, and
G.
Mak
,
Phys. Rev. B
50
,
5226
(
1994
).
19.
G.
Mak
and
H. M.
Van Driel
,
Phys. Rev. B
49
,
16817
(
1994
).
20.
K.
Tanaka
,
H.
Ohtake
, and
T.
Suemoto
,
Phys. Rev. Lett.
71
,
1935
(
1993
).
21.
S.
Sadasivam
,
M. K. Y.
Chan
, and
P.
Darancet
,
Phys. Rev. Lett.
119
,
136602
(
2017
).
22.
D.
De Matteis
,
M.
De Luca
,
E. M. T.
Fadaly
,
M. A.
Verheijen
,
M.
López-Suárez
,
R.
Rurali
,
E. P. A. M.
Bakkers
, and
I.
Zardo
,
ACS Nano
14
,
6845
(
2020
).
23.
E.
Gatti
,
E.
Grilli
,
M.
Guzzi
,
D.
Chrastina
,
G.
Isella
,
A.
Chernikov
,
V.
Bornwasser
,
N.
Köster
,
R.
Woscholski
, and
S.
Chatterjee
,
Phys. Rev. B
84
,
245319
(
2011
).
24.
S.
De Cesari
,
A.
Balocchi
,
E.
Vitiello
,
P.
Jahandar
,
E.
Grilli
,
T.
Amand
,
X.
Marie
,
M.
Myronov
, and
F.
Pezzoli
,
Phys. Rev. B
99
,
035202
(
2019
).
25.
B.
Julsgaard
,
N.
von den Driesch
,
P.
Tidemand-Lichtenberg
,
C.
Pedersen
,
Z.
Ikonic
, and
D.
Buca
,
Photonics Res.
8
,
788
(
2020
).
26.
H. I. T.
Hauge
,
M. A.
Verheijen
,
S.
Conesa-Boj
,
T.
Etzelstorfer
,
M.
Watzinger
,
D.
Kriegner
,
I.
Zardo
,
C.
Fasolato
,
F.
Capitani
,
P.
Postorino
,
S.
Kölling
,
A.
Li
,
S.
Assali
,
J.
Stangl
, and
E. P. A. M.
Bakkers
,
Nano Lett.
15
,
5855
(
2015
).
27.
H. I. T.
Hauge
,
S.
Conesa-Boj
,
M. A.
Verheijen
,
S.
Koelling
, and
E. P. A. M.
Bakkers
,
Nano Lett.
17
,
85
(
2017
).
28.
G.
Lasher
and
F.
Stern
,
Phys. Rev.
133
,
A553
(
1964
).
30.
J. K.
Katahara
and
H. W.
Hillhouse
,
J. Appl. Phys.
116
,
173504
(
2014
).
31.
H. L.
Chen
,
A.
Scaccabarozzi
,
R.
De Lépinau
,
F.
Oehler
,
A.
Lemaître
,
J. C.
Harmand
,
A.
Cattoni
, and
S.
Collin
,
Phys. Rev. Appl.
15
,
024006
(
2021
).
32.
P.
Borlido
,
J. R.
Suckert
,
J.
Furthmüller
,
F.
Bechstedt
,
S.
Botti
, and
C.
Rödl
,
Phys. Rev. Mater.
5
,
114604
(
2021
).
33.
D.
Saxena
,
S.
Mokkapati
,
P.
Parkinson
,
N.
Jiang
,
Q.
Gao
,
H. H.
Tan
, and
C.
Jagadish
,
Nat. Photonics
7
,
963
(
2013
).
34.
W. Z.
Xu
,
F. F.
Ren
,
D.
Jevtics
,
A.
Hurtado
,
L.
Li
,
Q.
Gao
,
J.
Ye
,
F.
Wang
,
B.
Guilhabert
,
L.
Fu
,
H.
Lu
,
R.
Zhang
,
H. H.
Tan
,
M. D.
Dawson
, and
C.
Jagadish
,
Nano Lett.
18
,
3414
(
2018
).
35.
S. A.
Church
,
N.
Patel
,
R.
Al-Abri
,
N.
Al-Amairi
,
Y.
Zhang
,
H.
Liu
, and
P.
Parkinson
,
Adv. Opt. Mater.
11
,
2202476
(
2023
).
36.
H.
Sumikura
,
G.
Zhang
,
M.
Takiguchi
,
N.
Takemura
,
A.
Shinya
,
H.
Gotoh
, and
M.
Notomi
,
Nano Letters
19
(
11
),
8059
8065
(
2019
).
37.
L.
Vincent
,
E. M. T.
Fadaly
,
C.
Renard
,
W. H. J.
Peeters
,
M.
Vettori
,
F.
Panciera
,
D.
Bouchier
,
E. P. A. M.
Bakkers
, and
M. A.
Verheijen
,
Adv. Mater. Interfaces
9
,
2102340
(
2022
).
38.
W. H. J.
Peeters
,
V. T.
van Lange
,
A.
Belabbes
,
M. C.
van Hemert
,
M. M.
Jansen
,
R.
Farina
,
M. A. J.
van Tilburg
,
M. A.
Verheijen
,
S.
Botti
,
F.
Bechstedt
,
J. E. M.
Haverkort
, and
E. P. A. M.
Bakkers
,
Nat. Commun.
15
,
5252
(
2024
).
39.
J.
Fu
,
Q.
Xu
,
G.
Han
,
B.
Wu
,
C. H. A.
Huan
,
M. L.
Leek
, and
T. C.
Sum
,
Nat. Commun
8
,
1300
(
2017
).
40.
Y.
Harada
,
N.
Kasamatsu
,
D.
Watanabe
, and
T.
Kita
,
Phys. Rev. B
93
,
115303
(
2016
).
41.
Y.
Zhang
,
G.
Conibeer
,
S.
Liu
,
J.
Zhang
, and
J. F.
Guillemoles
,
Prog. Photovoltaics Res. Appl.
30
,
581
(
2022
).
42.
G.
Conibeer
,
Y.
Zhang
,
S. P.
Bremner
, and
S.
Shrestha
,
Jpn. J. Appl. Phys., Part 1
56
,
091201
(
2017
).
43.
R.
Clady
,
M. J. Y.
Tayebjee
,
P.
Aliberti
,
D.
König
,
N. J.
Ekins-Daukes
,
G. J.
Conibeer
,
T. W.
Schmidt
, and
M. A.
Green
,
Prog. Photovoltaics Res. Appl.
20
,
82
(
2012
).
44.
S.
Barman
and
G. P.
Srivastava
,
Phys. Rev. B
69
,
235208
(
2004
).
45.
P. Y.
Yu
and
M.
Cardona
,
Fundamentals of Semiconductors
, 4th ed. (
Springer
,
Heidelberg Dordrecht London New York
,
2010
).
46.
C.
Rödl
,
J.
Furthmüller
,
J. R.
Suckert
,
V.
Armuzza
,
F.
Bechstedt
, and
S.
Botti
,
Phys. Rev. Mater.
3
,
034602
(
2019
).
47.
J. E.
Huheey
,
E. A.
Keiter
, and
R. L.
Keiter
,
Inorganic Chemistry: Principles of Structure and Reactivity
, 4th ed. (
HarperCollins
,
New York
,
1993
).
48.
C.
Fasolato
,
I.
Zardo
, and
M.
De Luca
,
Fundamental Properties of Semiconductor Nanowires
(
Springer Nature
,
Singapore
,
2021
).
49.
M. A. J.
van Tilburg
,
M. F.
Schouten
,
V. T.
van Lange
, and
R.
Farina
(
2024
). “Observation of carrier cooling in direct band gap hexagonal silicon-germanium,”
Zenodo
.