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

Single-mode GaN nanowire lasers

Open Access Open Access

Abstract

We demonstrate stable, single-frequency output from single, as-fabricated GaN nanowire lasers operating far above lasing threshold. Each laser is a linear, double-facet GaN nanowire functioning as gain medium and optical resonator, fabricated by a top-down technique that exploits a tunable dry etch plus anisotropic wet etch for precise control of the nanowire dimensions and high material gain. A single-mode linewidth of ~0.12 nm and >18dB side-mode suppression ratio are measured. Numerical simulations indicate that single-mode lasing arises from strong mode competition and narrow gain bandwidth.

©2012 Optical Society of America

1. Introduction

Advances in nanowire lasers are occurring at a rapid pace, with optically pumped lasers demonstrated in a variety of semiconductor material systems, including Group III-nitrides [16], Group III-V [7, 8], and Group II-VI [9, 10]. To achieve single-mode lasing, Xiao et al. used a CdSe active medium and a complex resonator comprising of coupled loops of long, flexible nanowires [11]. Recently, Scofield et al. reported single-mode lasing from a defect mode in an InGaAs photonic crystal consisting of a 2-dimensional array of InGaAs nanowires [12]. To circumvent the need for fabrication and manipulation of complicated optical structures, we concentrated our effort towards a simple, linear, Fabry-Perot nanowire laser, where the factors enabling single-mode lasing are short cavity length, small cross section and very high material gain. By reducing the nanowire size, the number of cavity modes within the gain bandwidth is dramatically reduced. This in turn requires high material gain, necessary to compensate for the reduced gain length. Low defect density and high sample uniformity are also necessary to reduce carrier losses and inhomogeneous broadening, in order to achieve high carrier density and increase gain competition, respectively. By satisfying these requirements, we are able to demonstrate single-mode lasing with a linewidth of ~0.12 nm and >18dB side-mode suppression ratio, in a 135 nm wide, 4.7 μm long GaN nanowire, under optical pumping. Single-mode operation is maintained far above lasing threshold. With GaN (or InGaN), the payoff is single-mode laser operation at ultraviolet (or green) wavelengths, which is an important spectral region not covered well by present semiconductor lasers.

2. Experiment

To achieve single mode lasing, precise control over the nanowire geometry is required. This requirement is satisfied by a top-down fabrication technique which produces uniform and vertically aligned GaN nanowire arrays from c-plane GaN epilayers on sapphire with low defect density and smooth sidewalls. We previously demonstrated this technique to fabricate nanowire LED structures with an axial GaN/InGaN multi-quantum structure [13]. Figure 1 shows SEM images of top-down GaN nanowires during the fabrication process. Starting from Si-doped planar GaN epilayers grown on 2” c-plane sapphire wafers in a Veeco D-125 metal organic chemical vapor deposition reactor, a 2-step etching process is used: a lithographic dry etch followed by an anisotropic wet etch. Following a process reported by Reculusa and Ravaine [14], a monolayer of 3 μm diameter silica colloids was self-assembled on the GaN surface in a Langmuir-Blodgett trough prior to etching to serve as a semi-periodic lithographic etch mask. GaN posts are subsequently formed by a plasma (dry) etch. As shown in Fig. 1(a), the resulting posts are tapered with large cross-section areas, and therefore, are unsuitable for single-mode nanowire lasers. Moreover, ion bombardment during the plasma etch damages the nanowire surface, as evidenced by significant increase in yellow luminescence [13]. These issues are resolved during the second (wet) etch step [Figs. 1(b)1(d)]. With the anisotropic wet-etch step, non-tapered GaN nanowires are created with damage-free surfaces, hexagonal cross sections with m-plane {10-10} sidewall facets, and a top c-plane (0001) resonator (end) facet. The GaN nanowire length is determined by the original GaN epilayers thickness and its width is determined by the duration of the wet-etch. Similar axial GaN/InGaN nanowire LED structures prepared by this top-down approach were largely dislocation free (~94% of the nanowires) [13].

 figure: Fig. 1

Fig. 1 Cross sectional SEM images showing GaN posts morphology transiting into GaN nanowires (a) before wet etch, (b) after 2 hours, (c) after 6 hours and (d) after 9 hours from start of wet etch. All images have the same magnification. Scale bars are 2 μm.

Download Full Size | PDF

The nanowires were characterized by scanning electron microscopy (SEM), x-ray diffraction (XRD) and transmission electron microscopy (TEM). For the optical pumping measurements, the nanowires were removed from their sapphire growth substrate and transferred to clean Si3N4 surfaces on TEM grids. Each nanowire was optically pumped at room temperature with a 10 kHz, 100 ps pulsed quadrupled YAG laser emitting at 267 nm. The intensity incident on the nanowire was varied using neutral density filters. A 50 × ultraviolet objective lens is used to image the pump laser output to an approximately 5 µm diameter spot on the GaN nanowire. Optical emission from the nanowire was collected with the same objective lens. The collected light was analyzed by a cooled CCD detector and a 300 mm spectrometer with a 2400 groove/mm holographic grating. The schematic of the optical pumping measurement set up is given in Fig. 2 .

 figure: Fig. 2

Fig. 2 Schematic of the GaN nanowire optical pumping and measurement setup.

Download Full Size | PDF

3. Theory

First, the nanowire passive-cavity eigenmodes are determined using a finite-difference time-domain (FDTD) solver. To determine which of the passive-cavity modes will be lasing, the FDTD information is fed into a laser model. We use a semiclassical multimode laser model, where the time evolution of the intensity In in a passive-cavity mode n is given by [15]

In˙=(αnβnInmnθnmIm)In+F
In the above equation, αn is the net modal gain, βn is the self-saturation coefficient, θnm is the cross-saturation coefficient, F accounts for spontaneous emission and the summation Σm≠n is over all passive-cavity modes in the nanowire. The equation resembles closely the Lotka-Votterra equations, which is widely used for describing predator-prey dynamics in biological systems [16]. Modifications to the typical multimode laser analysis include a more accurate treatment of outcoupling effects by coupling the nanowire cavity to free space [17]. Doing so allows the description of line narrowing of each passive cavity model as the device makes the transition from below to above lasing threshold. The GaN gain medium parameters for different frequencies and carrier densities, such as the net modal gain αn, are determined from a many-body gain theory [18].

The set of coupled mode intensity equations (numbering roughly 300) is solved numerically for the steady-state solution for a given excitation level, α11. Plotting the steady-state In versus the passive cavity-mode frequency Ωn gives the emission spectrum.

4. Results and discussion

The optical properties of the GaN nanowires were characterized by optical pumping using a micro-photoluminescence setup as sketched in Fig. 2(a). For comparison, two nanowires of similar diameters but different lengths were fabricated and characterized. A 135 nm diameter, 4.7 μm long nanowire and a 145 nm diameter, 7.2 μm long nanowire were positioned on SiN substrates. Figures 3(a) and 3(b) show CCD images of optical emission from the shorter GaN nanowire at two pump intensities. At low pump intensity, the CCD image shows roughly uniform optical emission from the entire nanowire length, indicating that unguided (out-of-plane) and guided (on-axis) emission intensities are basically equal [see Fig. 3(a)]. The inference is that the optical emission is entirely from spontaneous emission. At a higher pump intensity, a distinctly different emission pattern emerges. The on-axis emission grows significantly, appearing as two bright spots at the ends of the nanowire indicative of Fabry-Perot lasing, as shown in Fig. 3(b). The light collected by the objective lens is directly from the highly diverging output beam and indirectly from scattering of the output beam by the SiN surface. The large divergence angle of output beam is because of diffraction from a sub-wavelength nanowire aperture. At the same time, there is a clamping of the unguided spontaneous emission, which appears as a darkening (relative to the bright ends and with filters in place) of the entire length of the nanowire, also shown in Fig. 3(b). Also present in the CCD image are concentric rings of optical interference fringes, which is an indication of the spectral coherence of the emitted light.

 figure: Fig. 3

Fig. 3 CCD images of a GaN nanowire pumped below (a) and above (b) lasing threshold, respectively. The nanowire laser emits a highly divergent beam from the facets, some of which is collected by the objective lens. The objective lens also collects radiation emitted from the facets that is scattered by the SiN substrate surface, as well as spontaneous emission exiting perpendicular to the nanowire axis. Scale bars are 2 µm.

Download Full Size | PDF

Figure 4 plots the light out versus pump power and output intensity spectra for the two nanowire lasers of different lengths. The experiment was performed at T = 300 K. For both nanowires, the light out versus pump power curves show abrupt slope changes indicative of lasing threshold. Also, the peak intensity grew linearly with no sign of saturation throughout the excitation range. Emission spectra were measured at various pump intensities. The right column shows spectra for the two nanowire lasers at four optical pumping intensities. For the 4.7 μm long nanowire laser, the emission spectrum (lowest curve, Fig. 4(c)) is featureless and broad, with a full-width at half-maximum (FWHM) of approximately 6 nm at a pump intensity of 150kW/cm2. At the next higher pump intensity of 268kW/cm2, the emission spectrum starts to exhibit many sharp peaks, suggesting the onset of amplified spontaneous emission. For a pump intensity of 323kW/cm2, the spectrum shows a single sharp peak with FWHM < 0.12 nm over a broad background. The single-mode behavior remains (see highest intensity spectrum, Fig. 4(c)) as the pump intensity continues to increase. For this laser, the mode spacing was measured to be 1.9 nm. Side-mode suppression ratio is also maintained at greater than 18dB. An increase in the nanowire length results in a significant reduction in side-mode suppression, leading to emission characteristic of multimode lasing, as evidenced by the 7.2 μm long nanowire shown in Fig. 4(f). For this laser, the mode spacing was measured to be 1.3 nm. We note that multimode lasing was typical for the larger diameter nanowires (e.g. > ~180 nm), with <1.5 nm measured mode spacing. Thus, both the length and the diameter play an important role in the ability to fabricate a single-mode nanowire laser. The reduction of transverse modes is achieved by limiting the diameter and the reduction of longitudinal modes is similarly achieved by limiting the length.

 figure: Fig. 4

Fig. 4 (a, d) Nanowire laser intensity versus pump laser intensity, for two different nanowires with lengths of 4.7 µm and 7.2 µm (top and bottom, respectively). (c,f) Photoluminescence spectra from the nanowire lasers under uniform excitation for pump intensities as indicated in the figures. (b,e) Scanning electron micrographs of the GaN nanowire lasers. (b) shows the smaller dimensioned nanowire with a width of 135 nm and length of 4.7 µm (e) shows the larger dimensioned nanowire with a width of 145 nm and length of 7.2 µm.

Download Full Size | PDF

Summarizing the experimental results at this point, Figs. 3 and 4 together provide strong evidence of single-mode lasing. Figure 3(b) shows the clamping of spontaneous emission and interference fringes indicating coherence in optical emission [19]. There is the presence of a lasing threshold [Fig. 4 (left column, top)] and spectral narrowing to single frequency [Fig. 4 (right column, top)].

The experiments were modeled to better understand the mechanisms leading to single-mode lasing and to develop an analytical tool for designing future nanowire lasers. A combination of mode simulations, multimode-laser modeling and many-body laser-gain calculation were employed. First, the nanowire passive-cavity eigenmodes were determined using a full vectorial commercial mode solver from Lumerical Inc. Then, multimode semiclassical laser theory [15] and many-body gain [18] calculations were used to determine which of the passive-cavity eigenmodes will be above lasing threshold for given experimental conditions. Emission spectra at different excitations were computed for nanowire lasers with dimensions approximating those experimentally measured. Figure 5(a) shows the intensity spectra for increasing excitation for a 140 nm diameter, 4.5 µm long nanowire. The spectra show multiple resonances initially, which condense to only one mode operating above lasing threshold. A criterion for lasing is spectral line narrowing, as clearly exhibited by the mode at 363 nm. On the other hand, each side mode remains broader than the central mode, with width determined by the passive Fabry-Perot resonance. Amplified spontaneous emission gives rise to the side-mode intensity. Figure 5(b) shows the case of a longer 7.3 μm long (140 nm diameter) nanowire laser. Side-mode suppression is decreased by roughly a factor of 3 when compared to the shorter nanowire laser for α11 > 1, where α1 is the net modal gain and β1 is the self-saturation coefficient. Convincing evidence for multimode operation is from side-mode spectral narrowing to width comparable to the central mode, indicating that they are above lasing threshold. Figure 5 is qualitatively consistent with the experimental results described in Fig. 4. Simulations were also performed for longer nanowire devices. We found multimode operation for lasers longer than ~7 μm, consistent with earlier reports and experiments performed at Sandia.

 figure: Fig. 5

Fig. 5 Emission spectra for 4.5 μm (a) and 7.3 μm (b) long nanowire lasers and different excitation, α11 as indicated. The results are obtained using F = 0.05 and θnm1 = 1 for all n and m. Laser intensity is in dimensionless units, I = (µE/ħ)2T1T2 where µ,E,T1, and T2 are the dipole matrix element, electric field amplitude, population lifetime and dephasing decay time, respectively

Download Full Size | PDF

The value of the above modeling exercise is three-fold. First is a better understanding of the mechanisms giving rise to single-mode lasing. According to the multimode laser model, single-mode lasing depends on two factors: (1) suppression of modes (e.g. transverse modes) close to the single lasing mode via mode competition, and (2) absence of net gain in the further lying modes because of large longitudinal mode spacing and finite gain bandwidth. The second purpose of the model is to determine the laser parameters necessary for using the multimode laser model to design future lasers. Most important is the determination of the mode-coupling parameter [11], C≡θnmθmn/(βnβm), which is extremely difficult to obtain from first-principles because of complicated many-body correlation effects. Theory/experiment comparisons with different length nanowire lasers gives 0.7 < C < 1.0, putting the GaN active medium in the upper end of the weak-coupling regime. Consequently, suppression of side modes comes entirely from the effective net gain αnnmαmm<0 [8] which gives rise to the condition of longitudinal mode space > 10 meV and ΓT/ ΓL < 0.9 for single-mode lasing, where ΓT and ΓL are transverse and longitudinal mode confinement factors, respectively. Further study is necessary to confirm these conditions. The third value of the modeling is to provide information on the laser gain and carrier density achieved in our experiments. We estimated material gain and carrier density created under optical pumping to be ≈5.8 × 103 cm−1 and 1.4 × 1019 cm−3, respectively, which are appreciably higher than found in conventional lasers and is evidence for high material quality.

5. Conclusion

We have demonstrated stable, single-frequency output from single, as fabricated nanowire lasers operating far above lasing threshold. Each laser consists of a linear, double-facet GaN nanowire functioning as gain medium and optical resonator. A single-mode linewidth of ~0.12 nm and >18dB side-mode suppression ratio are measured. Crucial to achieving single-mode lasing is reducing the number of cavity modes within the gain bandwidth. This requires significant reduction and precise control of nanowire dimensions, as well as high material gain necessary to compensate for the reduced gain length. These challenges are met using a top-down technique that exploits a tunable dry etch plus anisotropic wet etch. Numerical simulations based on a multimode laser theory indicate that single-mode lasing arises from strong mode competition and narrow gain bandwidth.

Acknowledgments

This work was funded by Sandia’s Solid-State-Lighting Science Energy Frontier Research Center, funded by the U. S. Department of Energy, Office of Science, and Office of Basic Energy Sciences. This work was performed, in part, at the Center for Integrated Nanotechnologies, a U.S. Department of Energy, Office of Basic Energy Sciences user facility at Los Alamos National Laboratory (Contract DE-AC52-06NA25396) and Sandia National Laboratories (Contract DE-AC04-94AL85000). We thank Daniel Koleske for assistance with the GaN film growth. Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy’s National Nuclear Security Administration under contract DE-AC04-94AL85000.

References and links

1. J. C. Johnson, H. J. Choi, K. P. Knutsen, R. D. Schaller, P. D. Yang, and R. J. Saykally, “Single gallium nitride nanowire lasers,” Nat. Mater. 1(2), 106–110 (2002). [CrossRef]   [PubMed]  

2. S. Gradecak, F. Qian, Y. Li, H. G. Park, and C. M. Lieber, “GaN nanowire lasers with low lasing thresholds,” Appl. Phys. Lett. 87(17), 173111 (2005). [CrossRef]  

3. F. Qian, Y. Li, S. Gradecak, H.-G. Park, Y. Dong, Y. Ding, Z. L. Wang, and C. M. Lieber, “Multi-quantum-well nanowire heterostructures for wavelength-controlled lasers,” Nat. Mater. 7(9), 701–706 (2008). [CrossRef]   [PubMed]  

4. J. Heo, W. Guo, and P. Bhattacharya, “Monolithic single GaN nanowire laser with photonic crystal microcavity on silicon,” Appl. Phys. Lett. 98(2), 021110 (2011). [CrossRef]  

5. J. B. Schlager, N. A. Sanford, K. A. Bertness, and A. Roshko, “Injection-level-dependent internal quantum efficiency and lasing in low-defect GaN nanowires,” J. Appl. Phys. 109(4), 044312 (2011). [CrossRef]  

6. P. C. Upadhya, Q. M. Li, G. T. Wang, A. J. Fischer, A. J. Taylor, and R. P. Prasankumar, “The influence of defect states on non-equilibrium carrier dynamics in GaN nanowires,” Semicond. Sci. Technol. 25(2), 024017 (2010). [CrossRef]  

7. B. Hua, J. Motohisa, Y. Kobayashi, S. Hara, and T. Fukui, “Single GaAs/GaAsP coaxial core-shell nanowire lasers,” Nano Lett. 9(1), 112–116 (2009). [CrossRef]   [PubMed]  

8. R. Chen, T. T. D. Tran, K. W. Ng, W. S. Ko, L. C. Chuang, F. G. Sedgwick, and C. Chang-Hasnain, “Nanolasers grown on silicon,” Nat. Photonics 5(3), 170–175 (2011). [CrossRef]  

9. X. F. Duan, Y. Huang, R. Agarwal, and C. M. Lieber, “Single-nanowire electrically driven lasers,” Nature 421(6920), 241–245 (2003). [CrossRef]   [PubMed]  

10. M. A. Zimmler, F. Capasso, S. Muller, and C. Ronning, “Optically pumped nanowire lasers: invited review,” Semicond. Sci. Technol. 25(2), 024001 (2010). [CrossRef]  

11. Y. Xiao, C. Meng, P. Wang, Y. Ye, H. K. Yu, S. S. Wang, F. X. Gu, L. Dai, and L. M. Tong, “Single-Nanowire Single-Mode Laser,” Nano Lett. 11(3), 1122–1126 (2011). [CrossRef]   [PubMed]  

12. A. C. Scofield, S. H. Kim, J. N. Shapiro, A. Lin, B. L. Liang, A. Scherer, and D. L. Huffaker, “Bottom-up Photonic Crystal Lasers,” Nano Lett. 11(12), 5387–5390 (2011). [CrossRef]   [PubMed]  

13. Q. M. Li, K. R. Westlake, M. H. Crawford, S. R. Lee, D. D. Koleske, J. J. Figiel, K. C. Cross, S. Fathololoumi, Z. T. Mi, and G. T. Wang, “Optical performance of top-down fabricated InGaN/GaN nanorod light emitting diode arrays,” Opt. Express 19(25), 25528–25534 (2011). [CrossRef]   [PubMed]  

14. S. Reculusa and S. Ravaine, “Synthesis of colloidal crystals of controllable thickness through the Langmuir-Blodgett technique,” Chem. Mater. 15(2), 598–605 (2003). [CrossRef]  

15. M. Sargent, W. E. Lamb, and R. L. Fork, “Theory of a Zeeman Laser. I,” Phys. Rev. 164(2), 436–449 (1967). [CrossRef]  

16. A. J. Lotka, “Contribution to the theory of periodic reactions,” J. Phys. Chem. 14(3), 271–274 (1910). [CrossRef]  

17. W. W. Chow, “Theory of emission from an active photonic lattice,” Phys. Rev. A 73(1), 013821 (2006). [CrossRef]  

18. W. W. Chow, A. Knorr, and S. W. Koch, “Theory of Laser Gain in Group-III Nitrides,” Appl. Phys. Lett. 67(6), 754–756 (1995). [CrossRef]  

19. L. K. van Vugt, S. Rühle, and D. Vanmaekelbergh, “Phase-correlated nondirectional laser emission from the end facets of a ZnO nanowire,” Nano Lett. 6(12), 2707–2711 (2006). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Cross sectional SEM images showing GaN posts morphology transiting into GaN nanowires (a) before wet etch, (b) after 2 hours, (c) after 6 hours and (d) after 9 hours from start of wet etch. All images have the same magnification. Scale bars are 2 μm.
Fig. 2
Fig. 2 Schematic of the GaN nanowire optical pumping and measurement setup.
Fig. 3
Fig. 3 CCD images of a GaN nanowire pumped below (a) and above (b) lasing threshold, respectively. The nanowire laser emits a highly divergent beam from the facets, some of which is collected by the objective lens. The objective lens also collects radiation emitted from the facets that is scattered by the SiN substrate surface, as well as spontaneous emission exiting perpendicular to the nanowire axis. Scale bars are 2 µm.
Fig. 4
Fig. 4 (a, d) Nanowire laser intensity versus pump laser intensity, for two different nanowires with lengths of 4.7 µm and 7.2 µm (top and bottom, respectively). (c,f) Photoluminescence spectra from the nanowire lasers under uniform excitation for pump intensities as indicated in the figures. (b,e) Scanning electron micrographs of the GaN nanowire lasers. (b) shows the smaller dimensioned nanowire with a width of 135 nm and length of 4.7 µm (e) shows the larger dimensioned nanowire with a width of 145 nm and length of 7.2 µm.
Fig. 5
Fig. 5 Emission spectra for 4.5 μm (a) and 7.3 μm (b) long nanowire lasers and different excitation, α11 as indicated. The results are obtained using F = 0.05 and θnm1 = 1 for all n and m. Laser intensity is in dimensionless units, I = (µE/ħ)2T1T2 where µ,E,T1, and T2 are the dipole matrix element, electric field amplitude, population lifetime and dephasing decay time, respectively

Equations (1)

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

I n ˙ =( α n β n I n mn θ nm I m ) I n +F
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.