Compact optical interconnects require efficient lasers and modulators compatible with silicon. Ab initio modeling of Ge1−xCx (x = 0.78%) using density functional theory with HSE06 hybrid functionals predicts a splitting of the conduction band at Γ and a strongly direct bandgap, consistent with band anticrossing. Photoreflectance of Ge0.998C0.002 shows a bandgap reduction supporting these results. Growth of Ge0.998C0.002 using tetrakis(germyl)methane as the C source shows no signs of C-C bonds, C clusters, or extended defects, suggesting highly substitutional incorporation of C. Optical gain and modulation are predicted to rival III–V materials due to a larger electron population in the direct valley, reduced intervalley scattering, suppressed Auger recombination, and increased overlap integral for a stronger fundamental optical transition.
I. INTRODUCTION
Germanium and its alloys have recently gathered attention as laser materials for silicon photonics. Not only is Ge compatible with silicon IC fabrication, with dislocation density below 105 cm−2,1 but the direct (Γ) conduction band (CB) valley of Ge is only 140 meV higher in energy than its indirect (L) valley. Adding strain or alloying with the appropriate material can reduce the energy of the Γ valley, allowing increased light emission and even a direct bandgap. Lasers have been demonstrated using GeSn2 and tensile-strained Ge.3,4 This paper investigates another potential direct bandgap group IV alloy, Ge1−xCx.
Ge1−xCx is a highly mismatched alloy (HMA) similar to dilute nitrides such as InGaAs:N. Carbon and N are much more electronegative than their respective hosts Ge and As: 2.55 and 3.04 vs. 2.01 and 2.18 Pauling units, respectively. This difference introduces a strong local perturbation in the host crystal. In the dilute nitrides, the result is a splitting of the conduction band at Γ leading to a reduced CB minimum and decrease in bandgap, and a large electron effective mass. These effects are often described by a band anticrossing (BAC) model in which N introduces a localized state within the CB that interacts with and repels the nearby CB states to lower energies.5
However, like other HMAs, Ge1−xCx can readily form defects, especially from typical growth techniques. C-C bonds are stronger than C-Ge bonds,6 forming CCGe split interstitial defects and other carbon clusters, all of which form trap states within the bandgap.7 C triplets are the most common molecular species evaporated thermally from graphite sources,8 with heavier clusters up to C14 detected.9 Atomic sources of C9 offer no panacea because typical Ge epitaxial techniques favor surface segregation of C,10 leading to C-C bonds as well. Indeed, Park et al. showed it would be virtually impossible to achieve fully substitutional growth of Ge1−xCx alloys by conventional molecular beam epitaxy (MBE) techniques.10
The tendency to C-C bonds and carbon clusters has led to a large range of reported characteristics, particularly in reports on the effect of C on the bandgap. Theoretical analysis has been inconclusive, with some techniques reporting a linear increase in bandgap with %C,11 others claiming band bowing12 to various extents, and a few predicting a composition range with a direct bandgap.8
This work seeks to resolve these discrepancies using improved ab initio modeling and growth techniques to extract the intrinsic properties of Ge1−xCx in the absence of C-C defects. Hybrid density functionals are used for numerical modeling with high accuracy despite the large difference in atom size and electronegativity between Ge and C. These results are coupled with material growth using a technique that preferentially incorporates substitutional C atoms rather than clustering and interstitial defects. All alloy percentages below are given as mole fractions (at. %).
II. COMPUTATIONAL METHODS
A. Computational details
Density functional theory (DFT) calculations were performed using the Vienna ab initio Simulation Package (VASP),13–16 which implements DFT using plane waves. The projector-augmented wave (PAW) core electron method was used with the generalized gradient approximation (GGA) and Perdew-Burke-Ernzerhof (PBE) functional,17–20 with a high cutoff energy of 400 eV to account for the "hard" C atom core. The PBE functional alone underestimates the Ge bandgap to the point of predicting it to be a semimetal, so the Heyd-Scuseria-Ernzerhof (HSE06) range separated hybrid exchange-correlation functional was also used in order to give a bandgap close to experimental values.21 When feasible, spin-orbit coupling (SOC) was also included in order to accurately represent the splitting of the valence bands. Ion locations were first relaxed within the GGA, then the lattice constant was varied using fixed fractional coordinates using the HSE06 potential to minimize system energy to a tolerance of 10−5 eV. Due to computational limits and the large size of the supercells used here (128 atoms), inclusion of PBE0 hybrid functionals22 and d electrons were impractical, although these made only minor shifts to the band structures of small Ge supercells. Further simulation details can be found in Ref. 22.
B. Folded band structures and semiempirical correction
DFT of supercells larger than the primitive crystal unit cell leads to folded band structures, as discussed below. Fig. 1 shows the folded band structure for Ge0.9922C0.0078 using HSE06 and spin-orbit coupling. No attempt was made to fit salient features such as bandgaps to experimental data for Ge1−xCx alloys due to the lack of consistent data to fit to. However, pure Ge offers a first order correction since its properties are well known, and HSE slightly underestimates the direct bandgap. By decreasing the lattice constant by 1.33% and changing the fraction of exact exchange (mixing parameter) in the hybrid functional from 25% to 18%, the band alignments and bandgaps of Ge can be accurately reproduced.23,24 Fig. 2 shows the Ge band structure with and without these corrections.
Folded band structure of Ge0.9922C0.0078 calculated including spin-orbit coupling, showing a splitting of the CB (red) leading to a sharp decrease in the minimum CB energy at Γ. Artificially folded band structure is apparent in the CB in the X direction ⟨100⟩, as well as light and heavy hole valence bands (blue).
Folded band structure of Ge0.9922C0.0078 calculated including spin-orbit coupling, showing a splitting of the CB (red) leading to a sharp decrease in the minimum CB energy at Γ. Artificially folded band structure is apparent in the CB in the X direction ⟨100⟩, as well as light and heavy hole valence bands (blue).
Comparison of 2-atom Ge band structure with SOC before (a) and after (b) adjusting lattice constant and HSE mixing parameter to accurately reproduce the band alignments and bandgaps at L and Γ. Dashed lines are provided as a guide to the eye.
Comparison of 2-atom Ge band structure with SOC before (a) and after (b) adjusting lattice constant and HSE mixing parameter to accurately reproduce the band alignments and bandgaps at L and Γ. Dashed lines are provided as a guide to the eye.
With or without these semi-empirical corrections, the change in the direct bandgap energy with the addition of dilute C is expected to be correct. We find the Γ conduction band valley decreases by approximately 170 ± 50 meV/%C for the first percent C, which is very close to the observed change in dilute nitrides (150–200 meV/%N for the first percent N).25,26
C. Unfolded band structures
Modeling of dilute alloys requires large supercell sizes, which leads to a folding of the Brillouin zone (BZ) and apparent band structure. The supercell used here consisted of 128 atoms: the 2 atom Ge primitive unit cell repeated four times along each of its basis vectors ½[110], ½[101], and ½[011]. Without unfolding, the bands are folded 4× in each direction, which overlays the Γ, X, and L points. Unfolding helps clarify changes in the band structure away from the Brillouin zone center at Γ. Unfolding is essential when attempting to distinguish between multiple minima and maxima in a single band, such as L and Γ valleys in the Ge CB, to establish whether the bandgap is direct or indirect.
To validate the interpretation of band splitting and effective masses from the folded band structures, bands were unfolded using the method described by Tomić et al.27 Fig. 3 shows the unfolded band structure for Ge0.9922C0.0078 without spin-orbit coupling, but including the semiempirical corrections from Section II B. For comparison, the band structure without these corrections is plotted in Fig. 4, i.e., the same uncorrected computational lattice constant as in Figure 1, and a mixing parameter of 25%. The splitting of the CB minimum at Γ into E+ and E− bands is clearly apparent, and the direct gap is clearly the lowest CB minimum. The origin of several noisy data points near 2/3⟨111⟩ remains unclear. These points do not appear to significantly affect the bands at Γ or L, and they do not appear to be attached to any band, so we speculate that they result from either insufficient k-points during the band structure extraction or insufficient convergence. Band structure calculations with additional k-points are currently underway.
Unfolded band structure of Ge0.9922C0.0078 with the same corrections used to correctly align the Ge band structure as in Fig. 2(b), albeit without spin-orbit coupling. E− and E+ bands are clearly visible at Γ, as is a direct bandgap. Radius of each data point represents the projection of the wavefunction onto p orbitals, allowing facile tracing of different bands across the Brillouin zone.
Unfolded band structure of Ge0.9922C0.0078 with the same corrections used to correctly align the Ge band structure as in Fig. 2(b), albeit without spin-orbit coupling. E− and E+ bands are clearly visible at Γ, as is a direct bandgap. Radius of each data point represents the projection of the wavefunction onto p orbitals, allowing facile tracing of different bands across the Brillouin zone.
Unfolded Ge0.9922C0.0078 band structure without SOC, similar to Figure 3, but without semiempirical corrections.
Unfolded Ge0.9922C0.0078 band structure without SOC, similar to Figure 3, but without semiempirical corrections.
III. GROWTH OF DILUTE GERMANIUM CARBIDE
As mentioned earlier, in order to grow high quality Ge1−xCx, the C must be sourced by a method that preferentially incorporates the atoms in substitutional sites. This can be accomplished through gas chemistry, by building a molecule that has one carbon atom bonded to four Ge atoms.28 When coupled with kinetically limited growth, this prevents C atoms bonding with other C atoms as well as C incorporating in interstitial sites.
Preliminary growth was completed using a hybrid gas/solid-source molecular beam epitaxy (MBE) chamber using tetrakis(germyl)methane (4GeMe) as the carbon source, diluted by solid-source Ge. A smooth epitaxial GaAs surface was prepared on a GaAs(100) wafer in a Veeco Gen 930 III-V MBE chamber by desorbing the native oxide, then growing a buffer of 100 nm GaAs, 50 nm AlAs for a back barrier, and 20 nm GaAs on top. The wafer was then transferred under ultrahigh vacuum to an Intevac Mod Gen II hybrid source MBE with a background pressure of 6 × 10−10 Torr without liquid nitrogen. Next, 10 nm of Ge was deposited at a 60 nm/h growth rate using solid-source Ge at a wafer temperature of 400 °C, calibrated by pyrometer. The Ge shutter remained open for the remaining growth.
Ge1−xCx quantum well (QW) growth began by adding 4GeMe in cycles as follows. To maximize 4GeMe incorporation efficiency, the gate valve was closed to the MBE turbomolecular pump. 4GeMe was injected from a nozzle ∼10 cm from the wafer surface at a foreline reading of 30 mTorr for 20 s until the growth chamber pressure reached 5 × 10−5 Torr. 4GeMe gas flow was stopped for a 40 s pause to permit additional residence time in the chamber. Then the gate valve was reopened for 20 s to pump out residual gases. The cycle was repeated for 20 min or 15 cycles, corresponding to 20 nm of equivalent solid-source Ge growth. Then the wafer temperature was increased by 25 °C, and a 50 nm Ge barrier layer was grown. This process was then repeated for 5 more wells at successively higher temperatures, up to 525 °C.
IV. EXPERIMENTAL RESULTS
A. Composition and morphology
To measure the carbon concentration vs. depth, secondary ion mass spectroscopy (SIMS) was performed by Evans Analytical Group, shown in Fig. 5. The C composition shows two plateaus at approximately 1019 cm−3 (0.03%) at ∼250 nm depth and 2 × 1019 cm−3 (0.05%) at 100–200 nm depth. These results are an average composition over the QWs and barriers, corresponding to QW compositions of 0.1% and 0.2%, respectively. The resolution of the SIMS measurement was poor (∼100 nm) due to the large surface roughness, as shown by atomic force microscopy (AFM) in Fig. 6 and transmission electron microscopy (TEM) in Fig. 7. The higher C concentration is found closer to the surface (at a depth of 100–200 nm, compared to the concentration at 250 nm depth), suggesting increased incorporation efficiency of the 4GeMe precursor and increased C incorporation, presumably due to increased hydrogen desorption with higher growth temperature. An increase in C composition with depth beyond 300 nm is attributed to carbon contamination at the original GaAs wafer surface. Deconvolution of the As concentration using height profiles derived from AFM and TEM show that the As tail toward the surface can be largely explained by the large surface roughness; during SIMS, GaAs is exposed in the deepest valleys before the Ge and Ge1−xCx are consumed.
SIMS from MBE-grown Ge1−xCx sample showing two plateaus at approximately 0.03% and 0.05% C incorporation. This equates to approximately 0.1% and 0.2% C in the quantum wells. Dotted lines for two C concentrations are shown as guides to the eye, using profile extracted from Ge at the Ge-GaAs interface.
SIMS from MBE-grown Ge1−xCx sample showing two plateaus at approximately 0.03% and 0.05% C incorporation. This equates to approximately 0.1% and 0.2% C in the quantum wells. Dotted lines for two C concentrations are shown as guides to the eye, using profile extracted from Ge at the Ge-GaAs interface.
10 × 10 μm AFM image of surface of MBE-grown Ge1−xCx sample showing very rough surface with flat valleys at equal height. Height profile along dashed line is shown at bottom.
10 × 10 μm AFM image of surface of MBE-grown Ge1−xCx sample showing very rough surface with flat valleys at equal height. Height profile along dashed line is shown at bottom.
Bright field TEM image along the zone center of Ge1−xCx sample. The first four layers appear to have grown smoothly up to dashed line (guide to eye), with roughening occurring as the growth temperature increased above 475 °C in layers 5 and 6. Inset: high resolution TEM.
Bright field TEM image along the zone center of Ge1−xCx sample. The first four layers appear to have grown smoothly up to dashed line (guide to eye), with roughening occurring as the growth temperature increased above 475 °C in layers 5 and 6. Inset: high resolution TEM.
Fig. 6 shows AFM of the finished growth over a 10 × 10 μm area, with an RMS roughness of 68 nm. However, as shown in the TEM in Fig. 7, this roughness appears to originate at a specific layer, with flat valleys between the peaks.
The first few layers appear to have grown smoothly, with roughening increasing sharply at or after the fourth QW, corresponding to a growth temperature of 475 °C. These observations suggest the optimal growth temperature is close to 450 °C to allow high C incorporation but still maintain a smooth surface. Despite the rough surface, no defects or dislocations were visible by high resolution TEM (not shown), with neither twins nor stacking faults.8 The C fraction was too low to detect by either image contrast or by energy-dispersive x-ray spectroscopy (EDX).
Raman spectroscopy (Fig. 8) shows no sign of graphitic C clusters; however, the substitutional C percentage is too low to be directly detectable, so there is no visible peak for the Ge-C bond at 530 cm−1.
Raman spectroscopy showing neither C clusters nor graphitic carbon in expected ranges. The substitutional C percentage is too low to unambiguously detect the Ge-C bond at 530 cm−1. Figure shows multiple scans from different regions of the sample.
Raman spectroscopy showing neither C clusters nor graphitic carbon in expected ranges. The substitutional C percentage is too low to unambiguously detect the Ge-C bond at 530 cm−1. Figure shows multiple scans from different regions of the sample.
Nuclear reaction analysis Rutherford backscattering spectroscopy (NRA-RBS) was performed to measure the crystallinity of the Ge1−xCx sample, as well as that of a control sample that consisted of a similar growth of Ge using digermane as a gas source. Fig. 9 shows very low back scattering signal along the ⟨001⟩ channels in the Ge1−xCx sample, but, again, the C percentage is below detectable limits to confirm the fraction of substitutional C.
NRA-RBS comparing Ge0.998C0.002 grown using 4GeMe (left) with pure Ge grown using digermane (right). The low backscattering signal along ⟨001⟩ channels shows high crystallinity in the Ge0.998C0.002 sample. Simulation fits are shown as solid black for total backscattering and dashed lines for component elements.
NRA-RBS comparing Ge0.998C0.002 grown using 4GeMe (left) with pure Ge grown using digermane (right). The low backscattering signal along ⟨001⟩ channels shows high crystallinity in the Ge0.998C0.002 sample. Simulation fits are shown as solid black for total backscattering and dashed lines for component elements.
B. Photoreflectance
In order to see the effect on the band edge, photoreflectance (PR) was performed on the Ge1−xCx sample. PR is an experimentally performed derivative of reflectivity, dR/R, accomplished by modulating the internal electric field while measuring reflectivity (R). The result eliminates the undesirable background reflectivity while enhancing weak features corresponding to specific electronic transitions. It is relatively insensitive to indirect gap transitions. Fig. 10 shows room temperature and 20 K PR spectra from Ge1−xCx and a bare Ge wafer for reference. A single resonance related to a direct optical transition at the Γ point is visible for the reference sample.
PR comparing a Ge substrate to MBE-grown Ge0.998C0.002, at room temperature and 20 K. At both temperatures, Ge0.998C0.002 shows a clear band-to-band absorption followed by Franz-Keldysh oscillation, whose extent is shown by dashed arrows. Numbered maxima and minima are used in Figure 11.
PR comparing a Ge substrate to MBE-grown Ge0.998C0.002, at room temperature and 20 K. At both temperatures, Ge0.998C0.002 shows a clear band-to-band absorption followed by Franz-Keldysh oscillation, whose extent is shown by dashed arrows. Numbered maxima and minima are used in Figure 11.
For the Ge1−xCx sample, a strong PR resonance followed by Franz Keldysh oscillations (FKO) is visible, rather than the single PR resonance typical of a single energy gap transition. Such a shape of PR resonance means that a built-in electric field exists in the Ge1−xCx layer. This field is estimated from the FKO period29 to be ∼80 kV/cm. The origin of this built-in field may be due to the gradient of C concentration along the growth direction, dipping the CB deeper in layers with higher %C. Alternately, C may introduce traps whose concentration varies with %C, causing the Fermi level to vary along the growth direction, and a built-in field. The observation of a clear PR signal represents good crystal quality and a well-defined electronic band structure. But a clear redshift of the PR resonance in comparison with the reference sample is difficult to see in this case because of the presence of FKO. On the other hand, for such low C concentration, the expected red shift of the PR resonance should be very small (25 meV), making it difficult to observe.
It is worth noting that the spectral position of the direct optical transition in Ge grown on GaAs can be different than in Ge bulk due to the slight residual strain in Ge layers grown on non-native substrates. This means that Ge1−xCx samples with larger C concentrations will be needed for further studies of C-related changes in the electronic band structure of Ge1−xCx alloy. For this sample, we can estimate the energy of the direct optical transition at the Γ point from the analysis of FKO period as shown in Figure 11. The accuracy of such estimation is limited, but in this case the transition energy (0.78 eV, corresponding to a 20 meV shift in E−) is consistent with our computational model.
Analysis of FKO period for Ge1−xCx sample at 20 K and room temperature. Arrows indicate extrapolated position of (top-bottom): E0 at 20 K, direct gap at 20 K, direct gap at room temperature.
Analysis of FKO period for Ge1−xCx sample at 20 K and room temperature. Arrows indicate extrapolated position of (top-bottom): E0 at 20 K, direct gap at 20 K, direct gap at room temperature.
V. DISCUSSION
Using parameters extracted from the HSE modeling, we predict direct bandgap lasing from Ge1−xCx at photon energies below 0.5 eV, corresponding to wavelengths of 2.5–5 μm. A Γ valley decrease of 170 ± 50 meV/%C suggests a strongly direct bandgap for 1%C, which would increase electron population in the Γ valley for increased optical gain. It would also suppress intervalley scattering, which would increase modulation by the quantum confined Stark effect (QCSE) by increasing the exciton lifetime.30
This is the first demonstration of Ge1−xCx growth in which defects are undetectable yet the bandgap has shifted. Point defects such as C-C clusters, split interstitials, and graphitic bonds were not detectable by Raman, EDX, or NRA-RBS, permitting a clear PR signal even though the C concentration itself was too low to verify Ge-C bonding. Furthermore, the lack of extended defects such as dislocations and stacking faults in both scanning and high resolution TEM indicates an absence of carbon clusters, which had been shown to nucleate extended defects.10
Although the 0.2% C concentration in the grown sample was insufficient for a direct bandgap, these results nevertheless show significant promise for direct bandgap devices. Strong optical transitions and the decrease in direct bandgap energy validate the DFT model results, which in turn show a direct bandgap at much smaller concentrations than previous reports. This is similar to dilute nitrides such as GaInNAsSb, in which the new E− band enabled even stronger optical transitions than Ga(In)As, thus increasing laser gain and modulation.31,32
The electron effective mass predicted by our DFT work, 0.045m0, is 1.5× that of the Γ CB valley in Ge (0.030m0 by DFT for a 128-atom supercell of pure Ge without semiempirical corrections). This increase is consistent with the band anticrossing model and dilute nitrides, and it suggests further increases in gain and modulation due to an increased electron-hole overlap integral in QWs.32 More importantly, it also improves on the small density of states in the Γ valley of Ge, tensile Ge, and GeSn, in which the heavy L valleys capture most of the electrons even if a direct bandgap is achieved.33,34 Thus, Ge1−xCx offers a route to lasers and modulators comparable with III-V materials.
The multi-QW structure grown in this experiment and the roughness of the sample make interpretation of CV and other electrical measurements problematic. However, this work does provide a target growth temperature of 450 °C for subsequent growths of bulk Ge1−xCx for electrical characterization of trap energies and densities. Future efforts will focus on increasing the C concentration to demonstrate a direct bandgap and lasing in Ge1−xCx.
Finally, we note that the calculated bandgap of Ge0.9922C0.0078 is approaching the energy of the split-off valence band (Fig. 1). A slightly higher C concentration in the alloy would suppress CHHS Auger recombination if EG < ESOH. Likewise, the similar wavefunction character in the E+ and E− bands (Fig. 4) suggests that free carrier absorption (FCA) from E− to E+ may be insignificant, although such calculations are beyond the scope of this work. Suppressed Auger and intraband FCA may make Ge1−xCx a compelling laser source for wavelengths from 3 to 5 μm.
VI. CONCLUSION
We have provided evidence from ab initio modeling with hybrid functionals for a direct bandgap in Ge1−xCx for x < 1%, supported by photoreflectance showing a 20 meV reduction in bandgap even for 0.2% C. The numerical results are generally consistent with a band anticrossing model, in which the carbon atom causes a new, localized state above the conduction band minimum. As a result, the conduction band is split at Γ into E+ and E− bands, as it is in dilute nitrides. Because the interaction potential is stronger at Γ than L, the CB minimum at Γ decreases faster than the minimum at L, turning the alloy into a direct bandgap semiconductor.
This model is supported by experimental measurement of a strong optical transition with built-in field. Although the C mole fraction (0.2%) was too low to provide definitive measurements, no evidence of graphitic or split-interstitial C was detectable by Raman, NRA-RBS, or HRTEM, yet a 20 meV decrease in the direct bandgap was apparent in PR, consistent with our ab initio results. Lack of extended defects suggests that C surface segregation leading to C-C clustering was suppressed. From this we infer that the 4GeMe carbon precursor led to substitutional incorporation.
The experimental growth of Ge1−xCx without detectable carbon-carbon defects was made possible using a precursor gas that deposited C atoms on the surface pre-bonded to Ge, thus bypassing the thermodynamically favored surface segregation and carbon clusters. These are the first experimental evidence of bandgap reduction and anticrossing in Ge1−xCx material that is not dominated by C-C bonds or other carbon cluster defects.
Efforts to grow Ge1−xCx with higher concentrations of C for a strongly direct bandgap are currently underway.
ACKNOWLEDGMENTS
This work was supported in part by the National Science Foundation under Grant No. DMR-1508646, the Extreme Science and Engineering Discovery Environment (XSEDE) supported by NSF Grant No. ACI-1053575, and a Notre Dame Energy Center postdoctoral fellowship.
The authors thank the Notre Dame Center for Research Computing for calculation time with technical assistance from Dodi Heryadi, and Vincenzo Lordi and Eoin P. O'Reilly for helpful discussions.