You are on page 1of 5

LETTERS

PUBLISHED ONLINE: 19 JANUARY 2015 | DOI: 10.1038/NPHOTON.2014.321

Lasing in direct-bandgap GeSn alloy grown on Si


S. Wirths1*†, R. Geiger2,3†, N. von den Driesch1, G. Mussler1, T. Stoica1, S. Mantl1, Z. Ikonic4,
M. Luysberg5, S. Chiussi6, J. M. Hartmann7,8, H. Sigg2, J. Faist3, D. Buca1* and D. Grützmacher1

Large-scale optoelectronics integration is limited by the Süess et al.10 presented a stressor-free technique that enables the
inability of Si to emit light efficiently1, because Si and the introduction of more than 5.7% (ref. 24) uniaxial tensile strain in
chemically well-matched Ge are indirect-bandgap semiconduc- Ge µ-bridges via selective wet under-etching of a prestressed layer.
tors. To overcome this drawback, several routes have been An alternative technique to achieve a direct-bandgap material is to
pursued, such as the all-optical Si Raman laser2 and the hetero- incorporate Sn atoms into a Ge lattice, which primarily reduces the
geneous integration of direct-bandgap III–V lasers on Si3–7. gap at the Г-point. With a sufficiently high fraction of Sn, the
Here, we report lasing in a direct-bandgap group IV system energy of the Г-valley decreases to below that of the L-valley. This
created by alloying Ge with Sn8 without mechanically introdu- indirect-to-direct transition for relaxed GeSn binaries has been pre-
cing strain9,10. Strong enhancement of photoluminescence dicted to occur at ∼20% Sn by Jenkins and colleagues25, but more
emerging from the direct transition with decreasing tempera- recent calculations indicate much lower Sn concentrations in the
ture is the signature of a fundamental direct-bandgap semicon- range of 6.5–11.0% (refs 26, 27). A major challenge for the realization
ductor. For T ≤ 90 K, the observation of a threshold in emitted of such GeSn alloys is the low (<1%) equilibrium solubility of Sn in
intensity with increasing incident optical power, together with Ge28 and the large lattice mismatch of ∼15% between Ge and α-Sn.
strong linewidth narrowing and a consistent longitudinal For GeSn grown on Ge substrates, this mismatch induces biaxial
cavity mode pattern, highlight unambiguous laser action11. compressive strain, causing a shift of the Г- and L-valley crossover
Direct-bandgap group IV materials may thus represent a towards higher Sn concentrations27. Hence, strategies have been
pathway towards the monolithic integration of Si-photonic adopted to obtain partially and also fully relaxed GeSn layers on
circuitry and complementary metal–oxide–semiconductor Si29 and on lattice-matched InGaAs ternary alloy30, respectively.
(CMOS) technology. Here, we adopt the partial relaxation of up to 560-nm-thick layers
Although a group IV direct-bandgap material has not been of GeSn on Ge/Si(001)-virtual substrates (Ge-VS).
demonstrated as yet, Si photonics using CMOS-compatible processes In the present study we investigated five samples (A to E) grown
has made great progress through the development of Si-based wave- using an industry-compatible 200 mm wafer reduced-pressure CVD
guides12, photodetectors13 and modulators14. The technology now AIXTRON Tricent reactor and Ge2H6 and SnCl4 as precursors31,32.
emerging is rapidly expanding the landscape of photonics appli- The GeSn layer thickness was 200–300 nm for samples A to D
cations towards tele- and data communications, as well as sensing and 560 nm for sample E. The Sn concentrations (Table 1) were
from the infrared to the mid-infrared wavelength range15–17. The determined by Rutherford backscattering spectrometry (RBS)
light sources for such systems are currently lasers made from and X-ray diffraction reciprocal space mapping (XRD-RSM;
direct-bandgap group III–V materials operated off- or on-chip, and Supplementary Fig. 1). The Ge buffer layers grown at 400/750 °C
therefore require fibre coupling or heterogeneous integration, for contained a weak biaxial tensile strain of 0.16% at room temperature
example by wafer bonding3, contact printing4,5 or direct growth6,7. due to the different thermal expansion coefficients of Si and Ge. The
Hence, a laser source made of a direct-bandgap group IV material strain levels as well as Sn concentrations in the partially relaxed
would further boost lab-on-a-chip and trace gas sensing15, as well GeSn layers are summarized in Table 1. The latter were determined
as optical interconnects18, by enabling monolithic integration. In from a modified version of Vegard’s law33. The experimentally
this context, Ge has a prominent role because its conduction band determined Sn concentration and strain were used to calculate the
minimum at the Г-point of the Brillouin zone (referred to as the electronic bandgaps at room temperature (Supplementary Fig. 2).
Г-valley) is located only ∼140 meV above the fourfold degenerate Sample A, containing ∼8% Sn, was expected to be an indirect-
indirect L-valley. To compensate for this energy difference and thus bandgap semiconductor, because its L-valley is well below its
form a laser gain medium, heavy n-type doping of slightly tensile- Г-valley in energy. In samples B and C, the difference between the
strained Ge was proposed19. Later, laser action was reported for Г- and L-valleys (c.f. Table 1) was smaller than kBT at room temp-
optically20 and electrically pumped Ge21 doped to ∼1 × 1019 cm−3 erature. According to the calculation, the −0.71% strained sample D
and 4 × 1019 cm−3, respectively. However, pump–probe measure- exhibits a fundamental direct bandgap with the Г-valley 28 meV
ments of similarly doped and strained material did not show evidence below the indirect L-valley. Sample E is a replica of sample D,
for net gain22 and, in spite of numerous attempts, researchers have except for the epilayer thickness, which has been increased to
failed to substantiate the above results. Other concepts that have improve the overlap between the optical mode and the gain material.
been investigated concern the engineering of the Ge bandstructure The cross-sectional transmission electron microscopy (XTEM)
towards a direct-bandgap semiconductor using micromechanically micrographs in Fig. 1a,b of sample D (12.6% Sn) show the high crys-
stressed Ge nanomembranes9 or Si3N4 stressor layers23. Very recently, talline quality of the GeSn layer and reveal a high density of misfit

1
Peter Grünberg Institute 9 (PGI 9) and JARA-Fundamentals of Future Information Technologies, Forschungszentrum Juelich, 52425 Juelich, Germany.
2
Laboratory for Micro- and Nanotechnology (LMN), Paul Scherrer Institut, CH-5232 Villigen, Switzerland. 3 Institute for Quantum Electronics, ETH Zürich,
CH-8093 Zürich, Switzerland. 4 Institute of Microwaves and Photonics, School of Electronic and Electrical Engineering, University of Leeds, Leeds LS2 9JT,
UK. 5 Peter Grünberg Institute 5 (PGI 5) and Ernst Ruska-Centrum, Forschungszentrum Juelich, 52425 Juelich, Germany. 6 Dpto. Física Aplicada,
E.E.Industrial, Univ. de Vigo, Campus Universitario, 36310 Vigo, Spain. 7 University of Grenoble Alpes, F-38000 Grenoble, France. 8 CEA, LETI, MINATEC
Campus, F-38054 Grenoble, France. †These authors contributed equally to this work. * e-mail: s.wirths@fz-juelich.de; d.m.buca@fz-juelich.de

88 NATURE PHOTONICS | VOL 9 | FEBRUARY 2015 | www.nature.com/naturephotonics

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.321 LETTERS
Table 1 | Layer properties, bandstructure parameters and effective masses.
Sample a par (Å) (±0.01) a perp (Å) (±0.01) x Sn,XRD (%) (±0.3) x Sn,RBS (%) (±0.5) ε XRD (%) ΔE exp (meV) ΔE calc (meV)
A 5.688 5.757 8.0 8.0 −0.70 −80 −50
B 5.712 5.765 9.6 10.3 −0.52 −10 −8
C 5.731 5.773 11.1 11.5 −0.41 −5 26
D 5.727 5.799 12.6 13.0 −0.71 25 28
E 5.735 5.776 12.6 13.0 −0.57 25 39
apar and aperp are the in-plane and out-of-plane lattice constants, ΔEexp and ΔEcalc are the band offsets between Γ- and L-valleys extracted from the JDOS model and calculated from the deformation potentials.

dislocations at the interface (orange arrows). Part of the plastic steadily increases by about two orders of magnitude with the
relaxation occurred through the creation of dislocation half-loops temperature decreasing from 300 K to 20 K. This change in behav-
(blue arrows) extending into the Ge buffer layer. High-resolution iour is consistent with the fundamental bandgap being direct. The
imaging (Fig. 1c) shows that most of the misfit dislocations at an temperature dependence of the integrated photoluminescence
average spacing of 12.5 nm are pure edge dislocations with a intensities of samples B and C is more complex. For temperatures
Burgers vector of a/2[110]. These so-called Lomer dislocations are ≥150 K the intensities remain nearly constant, and they increase
the most efficient type of dislocation to induce strain relaxation34. slightly for T ≤ 150 K. The explanation for this involves the appli-
The fact that no threading dislocation reached the sample surface cation of a joint density of states (JDOS) model (Supplementary
in any of the examined TEM samples allows an estimate of the Section 4), including calculated effective masses and valence band
upper limit of the threading dislocation density of 5 × 106 cm−2. parameters (Supplementary Tables 1 and 2). The energy difference
To prove whether the bandgap is fundamental direct or indirect, ΔE between the Г- and L-valleys is used as a fitting parameter
temperature-dependent photoluminescence measurements (see together with the optically injected carrier density nC(T ) = (n0/τ0)
Methods) were performed. Figure 2a presents photoluminescence τ(T ), where n0 represents the density of carriers at room tempera-
spectra in the range of 20–300 K for four different GeSn alloys ture, τ(T ) is the temperature-dependent recombination time and
(samples A–D). Note that the ordinate is fixed to facilitate the com- τ0 = τ(300 K). For the fit we assume the following: (1) τ(T ) is iden-
parison of peak intensities. Going from sample A at room tempera- tical for all samples and (2) τ(T ) resembles the Shockley–Read–Hall
ture to sample D at 20 K, the peak intensity increases approximately (SRH) recombination process (Supplementary Fig. 4). An excellent
350 times. This enormous gain in intensity is a combined conse- fit of the T-dependence of the photoluminescence intensity is
quence of sample cooling and inversion of the band offset presented in Fig. 2b. We also find that n0 = 4 × 1017cm−3, which
between Г- and L-valleys. The weak, broad luminescence observed agrees with the excitation density (see Methods) and a τ0 of 0.35 ns
around 0.4 eV at lower temperatures might stem from misfit dislo- corresponding to a surface recombination velocity of 570 m s–1.
cations formed at the GeSn/Ge interface (discussed above). The Similar relaxation times have been measured for elemental Ge37,
photoluminescence intensity of the main peak is linearly related supporting the high crystalline quality of our GeSn layers. In the
to the excitation power, which is characteristic for a dominant fit of sample D, the Г-valley lies 25 meV below the L-valley, in excel-
band-to-band recombination (Supplementary Fig. 3). Figure 2b pre- lent agreement with the prediction of a fundamental direct bandgap.
sents the integrated photoluminescence intensity as a function of For sample A, the experiments reveal a clear indirect bandgap with
temperature. The curves are normalized to unity at 300 K. For 8% an −80 meV offset compared to the prediction of ΔE = −50 meV. To
Sn (sample A), the Г-valley emission intensity strongly decreases extract the dependence of the conduction band offset on Sn concen-
with decreasing temperature, which is typical for Ge35 and low- tration, xSn , the measured values were extrapolated to a strain of 0%
Sn-content GeSn alloys36. The Г-valley luminescence of sample D using ΔE = 7.7 eV per unit strain following the model calculations.
The direct bandgap as revealed by experiment is therefore reached
b 50 nm for fully relaxed samples for Sn concentrations of ∼9%, which is
in fair agreement with the theoretical prediction shown by the
green line in Fig. 2c.
Ge0.874 Sn0.126
As we will show in the remainder of this Letter, sample E (which is
the thicker pendant of layer D; Supplementary Fig. 5) provides suffi-
Ge-VS cient optical gain to enable lasing. For the gain measurement, the
luminescence was collected from the edge of a several-millimetre-
a long and 5-µm-wide waveguide structure that was excited over a vari-
g220
Ge0.874 Sn0.126 able length L by a 5-ns-long laser pulse at a wavelength of 1,064 nm.
Figure 3a presents photoluminescence spectra at 20 K with an
c
b = a/2 (110) optical excitation of 595 kW cm–2 for different stripe lengths,
revealing a more than linear increase of the intensity and a substan-
Ge0.874 Sn0.126 tial decrease in linewidth for increasing stripe lengths L (plotted in
Fig. 3b). As expected, the photoluminescence emission energy of
sample E (∼0.55 eV, which corresponds to a wavenumber of
50 nm 4,435 cm−1 and wavelength of 2.25 µm) closely matches the one
Ge-VS 5 nm Ge-VS
observed for sample D. The emission shifts to the blue with increas-
ing excitation power (Fig. 3b). The inset to Fig. 3a indicates an
Figure 1 | Crystal quality and dislocation analysis. a, Cross-sectional TEM overlap of the fundamental mode with the GeSn layer of ∼60%
image of Ge0.874Sn0.126 (sample D). b, Dislocation loops (indicated by blue (c.f. Supplementary Table 3). The modal gain g, which includes
arrows) emitted below the Ge0.874Sn0.126/Ge interface (indicated by orange waveguide losses, is determined for four different excitations
arrows) penetrate only into the Ge buffer. c, High-resolution TEM image of from38 IASE = IO + ISP/g[exp( gL) – 1], as shown in the lower part
the interface used for Burgers vector calculations. Lomer dislocations with of Fig. 3b, where IASE and ISP refer to the amplified and unamplified
b = a/2[110] are identified. spontaneous emission, respectively. IO. contains contributions from

NATURE PHOTONICS | VOL 9 | FEBRUARY 2015 | www.nature.com/naturephotonics 89

© 2015 Macmillan Publishers Limited. All rights reserved


LETTERS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.321

a 30
12 300 K
300 K 250 K
200 K
Photoluminescence intensity (a.u.)

0.06 150 K
1.0
6 100 K
20 50 K
0.03 20 K
200 K

0.00 0.0 0
0.4 0.6 0.8 0.4 0.6 0.8 0.4 0.6 0.8
10

Sample A Sample B Sample C Sample D


xSn = 8.0% xSn = 9.6% xSn = 11.1% xSn = 12.6%

0
0.4 0.6 0.8 0.4 0.6 0.8 0.4 0.6 0.8 0.4 0.6 0.8
Energy (eV)
c 100
b Experiment
Experiment
Normalized integrated photoluminescence intensity

Sample A Extrapolation to ε = 0%
Theory ε = 0%
Sample B
50
Sample C
101 Sample D

Direct bandgap
EL− EΓ (meV) 0
Indirect bandgap

100 −50
JDOS model
−80 meV
−10 meV
−5 meV −100
25 meV
10−1
0 40 80 120 160 200 240 280 6 8 10 12 14
Temperature (K) Sn content (%)

Figure 2 | Temperature-dependent photoluminescence measurements and modelling. a, Temperature-dependent photoluminescence spectra for samples A
to D. b, Integrated photoluminescence intensities normalized to the corresponding intensity at room temperature. Coloured curves show the result of a
photoluminescence intensity simulation that includes the calculation of the joint density of states (JDOS), with the band offset ΔE between the minima of the
Γ- and L-valleys being the key fitting parameter. c, ΔE as a function of Sn concentration. The indirect-to-direct bandgap transition is found at ∼9% Sn for
unstrained layers using 7.7 eV per unit of strain for the extrapolation.

the excited (but not amplified) higher-order modes as well as light At 1,000kW cm–2 peak excitation, lasing was observed up to 90 K
collected from the sidewalls. We limited the gain analysis to exci- (Fig. 4a, inset). This temperature coincides with the activation
tations of <600 kW cm–2 and lengths of ≤550 µm to avoid the temperature for SRH recombination (Supplementary Fig. 4).
stimulated feedback of backwards-reflected light from the waveguide Hence, we tentatively ascribe the threshold degradation as well as
sidewalls. The obtained modal gain as a function of pump energy is the still low external differential quantum efficiency (estimated at
plotted in Fig. 3c, and a differential gain of 0.40 ± 0.04 cm kW–1 is 1.5%; Supplementary Fig. 6) to a reduced carrier lifetime due to
obtained from the slope. By extrapolation to the gain onset, we as yet unidentified extrinsic recombination centres together with
obtain a threshold excitation density of ∼325 kW cm–2. the small energy separation between Г- and L-valleys, and
By exciting over the full length of a 1-mm-long waveguide and valence interband absorption22. The operating temperature and
hence employing the multiple reflections feedback from the wave- lasing efficiency can be improved by introducing heterostructure
guide facets forming a Fabry–Perot cavity (Fig. 4c, inset), an unam- layers comprising GeSnSi/GeSn31 for carrier confinement and
biguous proof of lasing can be seen in Fig. 4a as a distinct threshold by n-doping19.
in output intensity. Once this threshold is exceeded, the full-width at Figure 4c presents a final piece of evidence for lasing11 by
half-maximum (FWHM) decreases and the emission intensity showing the Fabry–Perot oscillations observed in 250- and 500-
increases dramatically (Fig. 4b). The laser intensity increase flatten- µm-long waveguide structures. From the oscillation period, a
s at ∼650 kW cm–2, which we attribute to sample heating. Shot-to- group mode refractive index of 4.5 is deduced, which reflects the
shot fluctuations of the excitation power are the reason for the lasing dispersion of the refractive index in the pumped GeSn as well as
onset lying slightly below the gain onset, as found from the variable in the Ge substrate37.
stripe length measurement. Similarly, the modal gain, as estimated In summary, we present detailed photoluminescence studies per-
from the reflection losses using α = 1/L·ln(1/R) with reflectivity R, formed on high-quality, partially strain-relaxed GeSn layers with Sn
appears at lower average excitation values according to Fig. 3c. concentrations of up to 12.6% grown on Ge-buffered Si(001)

90 NATURE PHOTONICS | VOL 9 | FEBRUARY 2015 | www.nature.com/naturephotonics

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.321 LETTERS
a 1 μm neff = 4.03 b c

FWHM (meV)
0.30
595 kW cm−2 Г = 60% 40 120 Gain from VSL method
Cavity threshold
GeSn n = 4.2 20
Linear fit
100
Ge-VS n = 4.0 595 kW cm−2
540 kW cm−2
0.20 Si n = 3.4 0.25 485 kW cm−2 558 meV 80
430 kW cm−2

Gain (cm−1)
IASE (a.u.)

0.0 0.2 0.4 0.6 Fit using:


TE mode intensity (a.u.) IASE = IO + (ISP/g)[exp(gL) − 1] 60

IASE (a.u.)
0.15 558 meV
0.10 Gain g (cm−2): 40
L = 400 μm 110 ± 8 551 meV
94 ± 2
63 ± 2
46 ± 2 20
0.05
50 μm 551 meV
0.00 0
0.51 0.54 0.57 0.60 0.63 0.66 0 200 400 600 300 350 400 450 500 550 600 650 700
Energy (eV) Stripe length, L (μm) Excitation (kW cm−2)

Figure 3 | Optical gain determination via the variable stripe length (VSL) method. a, Amplified spontaneous emission (ASE) spectra obtained from the
560 nm Ge0.874Sn0.126 layer (sample E) excited over lengths L between 50 and 400 µm at 595 kW cm–2. Inset: calculated intensity of the fundamental
transverse electric mode (colour-coded) within a 5-µm-wide, 900 nm steep waveguide structure, revealing an overlap with the GeSn layer of 60%.
b,c, Top (b): FWHM of the spectra presented in a, decreasing with increasing stripe length, L. Bottom (b): ASE intensities for the peak energies (551 meV
and 558 meV), fit using IASE = IO + (ISP/g)[exp(gL) – 1] to determine the modal gain (plotted in c as a function of excitation). Red: modal gain, obtained from
the lasing threshold observed in homogeneously excited Fabry–Perot waveguide cavities with lengths of 250 µm, 500 µm and 1 mm.

a b c 8
Energy (eV)
40 Cavity length P = 500 kW cm−2 Pump laser GeSn cavity
0.58 0.54 0.5 20 K
photoluminescence intensity (a.u.)

Photoluminescence intensity (a.u.)


30 250 µm
Photoluminescence

x 2,000 100 K
500 µm
intensity (a.u.)

90 K
30 1 mm 6 Ge VS
80 K
20 Si(001)
60 K
Integrated

Coherent light
40 K
10 20 4
20 K
25
FWHM (meV)

2.2 2.3 2.4


Wavelength (µm) 380
360 10 2 LC = 500 µm
)
−2

340 5 LC = 250 µm
cm

neff ~ 4.5
W

x 200 320 280 320 360


(k

x 200 Excitation (kW cm−2)


300 0 0
ion

x 200
at

280 200 400 600 800 1,000 545 550 555


cit
Ex

0.48 0.50 0.52 0.54 0.56 0.58 Excitation (kW cm−2) Energy (meV)
Energy (eV)

Figure 4 | Optically pumped direct-bandgap GeSn laser. a, Power-dependent photoluminescence spectra of a 5-µm-wide and 1-mm-long Fabry–Perot
waveguide cavity fabricated from sample E (dGeSn = 560 nm, 12.6% Sn). Inset: temperature-dependent (20–100 K) photoluminescence spectra at
1,000 kW cm–2 excitation density. b, Integrated photoluminescence intensity as a function of optical excitation for waveguide lengths LC = 250 µm, 500 µm
and 1 mm. Inset: FWHM around the lasing threshold for the 1-mm-long GeSn waveguide. c, High-resolution spectra of 250- and 500-µm-long waveguides
taken at 500 kW cm–2. The mode spacings are 0.50 meV and 0.27 meV, which correspond to a group refractive index of ∼4.5 for the lasing mode.
The pump laser homogeneously excites the waveguide cavity, and the light emitted from one of the etched facets is analysed (inset).

substrates. Structural investigations show a low density of threading gain material platform for cost-effective integration of electronic
dislocations, homogeneously distributed Sn atoms, and mild com- and photonic circuits.
pressive strain levels facilitated by a particularly favourable relax-
ation mechanism. The existence of a direct-bandgap group IV Methods
GeSn layers were grown on thick Ge/Si virtual substrates using a CVD AIXTRON
semiconductor that exhibits modal gain is demonstrated. Fabry– Tricent reduced-pressure reactor. Growth temperatures were chosen between 350
Perot resonators are fabricated, permitting the demonstration of and 390 °C, at rates between 17 and 49 nm min–1.
lasing under optical pumping. Owing to the striking relation The bandstructure around the Г point was calculated by the 8-band k.p method
between the SRH recombinations and laser quenching at ∼90 K, including strain effects. Indirect conduction band valleys split with the applied
surface passivation and design optimization regarding doping, strain, as described via appropriate deformation potentials. The parameters used are
provided in Supplementary Table 1.
optical mode confinement and carrier injection will help to increase For photoluminescence spectroscopy, a continuous-wave solid-state laser
the operation temperature as well as decrease the threshold exci- emitting at a wavelength of 532 nm with a power of 2 mW was focused to a spot size
tation density. In a forthcoming development, electrical injection of ∼5 µm using a 15× Schwarzschild objective (NA = 0.4). The emitted luminescence
in optimized SiGeSn/GeSn/SiGeSn double heterostructures15,31 was collected by the same objective, spectrally analysed using a Fourier-transform
will be demonstrated. In conclusion, although lasing is achieved at infrared spectrometer, and detected using a liquid-nitrogen-cooled InSb detector
with cutoff at 0.27 eV. The samples were mounted in a helium cold-finger cryostat.
low temperatures and relatively high optical pumping, this Steep, 900-nm-deep sidewalls and facets of the waveguide were fabricated using an
demonstration of a direct-bandgap group IV laser on Si(001) rep- SF6/C4F8-based reactive ion etching process. The gain measurements and lasing
resents a promising proof of principle for a CMOS-compatible demonstration were performed using a pulsed laser (5 ns) emitting at a wavelength

NATURE PHOTONICS | VOL 9 | FEBRUARY 2015 | www.nature.com/naturephotonics 91

© 2015 Macmillan Publishers Limited. All rights reserved


LETTERS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.321

of 1,064 nm and focused via a cylindrical lens onto a variable slit imaged 1:1 24. Sukhdeo, D. S., Nam, D., Kang, J.-H., Brongersma, M. L. & Saraswat, K. C.
onto the sample by a biconvex lens. Direct bandgap germanium-on-silicon inferred from 5.7% 〈100〉 uniaxial tensile
Cross-sectional TEM specimens were prepared using a dual-beam focused ion strain. Photon. Res. 2, A8 (2014).
beam (FIB) apparatus (FEI Helios Nanolab 400S) operated at 30 and 5 kV. A 3-µm- 25. Jenkins, D. & Dow, J. Electronic properties of metastable GexSn1–x alloys.
thick Pt/C protective layer was deposited on the surface of the sample before FIB Phys. Rev. B 36, 7994–8000 (1987).
milling. Surface damage by Ga ions was reduced by low-energy (<1 kV) Ar ion 26. Lu Low, K., Yang, Y., Han, G., Fan, W. & Yeo, Y. Electronic band structure and
milling using the Fischione Instruments Model 1040 Nanomill system. effective mass parameters of Ge1–xSnx alloys. J. Appl. Phys. 112, 103715 (2012).
Conventional and high-resolution images were recorded using an aberration- 27. Gupta, S., Magyari-Köpe, B., Nishi, Y. & Saraswat, K. C. Achieving direct
corrected (fourth-order) FEI Titan 80–300 transmission electron microscope band gap in germanium through integration of Sn alloying and external strain.
operated at 300 kV. J. Appl. Phys. 113, 073707 (2013).
28. He, G. & Atwater, H. A. Interband transitions in SnxGe1–x alloys. Phys. Rev. Lett.
Received 12 September 2014; accepted 2 December 2014; 79, 1937–1940 (1997).
published online 19 January 2015 29. Grzybowski, G. et al. Next generation of Ge1–ySny ( y = 0.01–0.09) alloys
grown on Si(100) via Ge3H8 and SnD4: reaction kinetics and tunable emission.
References Appl. Phys. Lett. 101, 072105 (2012).
1. Iyer, S. S. & Xie, Y. H. Light emission from silicon. Science 260, 40–46 (1993). 30. Chen, R. et al. Increased photoluminescence of strain-reduced, high-Sn
2. Rong, H. et al. An all-silicon Raman laser. Nature 433, 292–294 (2005). composition Ge1–xSnx alloys grown by molecular beam epitaxy. Appl. Phys. Lett.
3. Fang, A. W. et al. Electrically pumped hybrid AlGaInAs–silicon evanescent 99, 181125 (2011).
laser. Opt. Express 14, 9203–9210 (2006). 31. Wirths, S. et al. Tensely strained GeSn alloys as optical gain media. Appl.
4. Justice, J. et al. Wafer-scale integration of group III–V lasers on silicon using Phys. Lett. 103, 192110 (2013).
transfer printing of epitaxial layers. Nature Photon. 6, 612–616 (2012). 32. Wirths, S. et al. Band engineering and growth of tensile strained Ge/(Si)GeSn
5. Yang, H. et al. Transfer-printed stacked nanomembrane lasers on silicon. Nature heterostructures for tunnel field effect transistors. Appl. Phys. Lett. 102,
Photon. 6, 617–622 (2012). 192103 (2013).
6. Liu, H. et al. Long-wavelength InAs/GaAs quantum-dot laser diode 33. Gencarelli, F. et al. Crystalline properties and strain relaxation mechanism of
monolithically grown on Ge substrate. Nature Photon. 5, 416–419 (2011). CVD grown GeSn. ECS J. Solid State Sci. Technol. 2, P134–P137 (2013).
7. Chen, R. et al. Nanolasers grown on silicon. Nature Photon. 5, 170–175 (2011). 34. Gerthsen, D., Biegelsen, D., Ponce, F. A. & Tramontana, J. C. Misfit dislocations
8. Chen, R. et al. Demonstration of a Ge/GeSn/Ge quantum-well microdisk in GaAs heteroepitaxy on (001) Si. J. Cryst. Growth 106, 157–165 (1990).
resonator on silicon: enabling high-quality Ge(Sn) materials for micro- and 35. Sun, X., Liu, J., Kimerling, L. C. & Michel, J. Direct gap photoluminescence of
nanophotonics. Nano Lett. 14, 37–43 (2014). n-type tensile-strained Ge-on-Si. Appl. Phys. Lett. 95, 011911 (2009).
9. Sánchez-Pérez, J. R. et al. Direct-bandgap light-emitting germanium in 36. Ryu, M.-Y., Harris, T. R., Yeo, Y. K., Beeler, R. T. & Kouvetakis, J. Temperature-
tensilely strained nanomembranes. Proc. Natl Acad. Sci. USA 108, dependent photoluminescence of Ge/Si and Ge1–ySny/Si, indicating possible
18893–18898 (2011). indirect-to-direct bandgap transition at lower Sn content. Appl. Phys. Lett.
10. Süess, M. J. et al. Analysis of enhanced light emission from highly strained 102, 171908 (2013).
germanium microbridges. Nature Photon. 7, 466–472 (2013). 37. Geiger, R. et al. Excess carrier lifetimes in Ge layers on Si. Appl. Phys. Lett.
11. Samuel, I. D. W., Namdas, E. B. & Turnbull, G. A. How to recognize lasing. 104, 062106 (2014).
Nature Photon. 3, 546–549 (2009). 38. Shaklee, K. L., Nahory, R. E. & Leheny, R. F. Optical gain in semiconductors.
12. Xia, F., Sekaric, L. & Vlasov, Y. Ultracompact optical buffers on a silicon chip. J. Lumin. 7, 284–309 (1973).
Nature Photon. 1, 65–71 (2007).
13. Assefa, S., Xia, F. & Vlasov, Y. A. Reinventing germanium avalanche Acknowledgements
photodetector for nanophotonic on-chip optical interconnects. Nature The authors acknowledge the hospitality of the IR beamline of the SLS, where the
464, 80–84 (2010). photoluminescence experiments were performed. Part of this work was funded by the Swiss
14. Xu, Q., Schmidt, B., Pradhan, S. & Lipson, M. Micrometre-scale silicon National Science Foundation (SNF). This research received funding for CVD growth
electro-optic modulator. Nature 435, 325–327 (2005). investigations from the European Community’s Seventh Framework Programme (grant
15. Soref, R. Mid-infrared photonics in silicon and germanium. Nature Photon. agreement no. 619509; project E2SWITCH) and the BMBF project UltraLowPow
4, 495–497 (2010). (16ES0060 K).
16. Roelkens, G. et al. Silicon-based photonic integration beyond the
telecommunication wavelength range. IEEE J. Sel. Top. Quantum Electron. Author contributions
20, 394–404 (2014). J.M.H. fabricated the Ge/Si substrates. S.W. and D.B. planned the GeSn epitaxial growth
17. Duan, G.-H. et al. Hybrid III–V on silicon lasers for photonic integrated experiments and S.W. and N.v.d.D. fabricated the GeSn/Ge/Si samples. M.L. and S.C.
circuits on silicon. IEEE J. Sel. Top. Quantum Electron. 20, 158–170 (2014). carried out the TEM measurements and analysis. S.W., D.B., G.M., N.v.d.D. and T.S. carried
18. Heck, M. J. R. & Bowers, J. E. Energy efficient and energy proportional optical out crystal structure analysis including strain determination via XRD and RBS. Z.I.
interconnects for multi-core processors: driving the need for on-chip sources. performed the bandstructure simulations. S.W. and R.G. performed the optical
IEEE J. Sel. Top. Quantum Electron. 20, 1–12 (2014). measurements. R.G. and H.S. performed the JDOS modelling, gain analysis and mode
19. Liu, J. et al. Tensile-strained, n-type Ge as a gain medium for monolithic laser simulations. R.G. processed the GeSn cavities. S.M., J.F., D.B., H.S. and D.G. supervised the
integration on Si. Opt. Express 15, 11272–11277 (2007). experiments and coordinated data interpretation. S.W., H.S., R.G. and D.B. wrote the paper.
20. Liu, J., Sun, X., Camacho-Aguilera, R., Kimerling, L. C. & Michel, J. Ge-on-Si All authors discussed the results and commented on the manuscript.
laser operating at room temperature. Opt. Lett. 35, 679–681 (2010).
21. Camacho-Aguilera, R. E. et al. An electrically pumped germanium laser. Additional information
Opt. Express 20, 11316–11320 (2012). Supplementary information is available in the online version of the paper. Reprints and
22. Carroll, L. et al. Direct-gap gain and optical absorption in germanium correlated permissions information is available online at www.nature.com/reprints. Correspondence and
to the density of photoexcited carriers, doping, and strain. Phys. Rev. Lett. requests for materials should be addressed to S.W. and D.B.
109, 057402 (2012).
23. De Kersauson, M. et al. Optical gain in single tensile-strained germanium Competing financial interests
photonic wire. Opt. Express 19, 17925–17934 (2011). The authors declare no competing financial interests.

92 NATURE PHOTONICS | VOL 9 | FEBRUARY 2015 | www.nature.com/naturephotonics

© 2015 Macmillan Publishers Limited. All rights reserved

You might also like