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

Semipolar III-nitride laser diodes with zinc oxide cladding

Open Access Open Access

Abstract

Incorporating transparent conducting oxide (TCO) top cladding layers into III-nitride laser diodes (LDs) improves device design by reducing the growth time and temperature of the p-type layers. We investigate using ZnO instead of ITO as the top cladding TCO of a semipolar (202¯1) III-nitride LD. Numerical modeling indicates that replacing ITO with ZnO reduces the internal loss in a TCO clad LD due to the lower optical absorption in ZnO. Lasing was achieved at 453 nm with a threshold current density of 8.6 kA/cm2 and a threshold voltage of 10.3 V in a semipolar (202¯1) III-nitride LD with ZnO top cladding.

© 2017 Optical Society of America

1. Introduction

Following the development of p-type GaN and the demonstration of the blue light emitting diode (LED) [1,2], III-nitride light emitting devices have been aggressively pursued for commercial applications. Most commercially available III-nitride devices are grown on the c-plane of GaN and are affected by the presence of inherent spontaneous and piezoelectric polarization. This has led to active research on nonpolar and semipolar orientations of GaN to reduce the effects of polarization [3–6]. Reducing the polarization leads to a reduction in the quantum confined stark effect, higher radiative recombination rates, and increased differential gain [7].

The top cladding of a conventional III-nitride laser diode (LD) consists of thick p-(Al,Ga,In)N layers which are grown at temperatures much higher than the quantum wells (QWs) in the active region (AR). This can lead to thermal damage to the QWs, resulting in material degradation and reduced performance [8–11]. To mitigate this issue, p-GaN can be grown at colder temperatures of around 850-950 °C instead of the optimal 1000-1100 °C. One approach is to use non-epitaxial layers to replace or reduce the thickness of III-nitride p-cladding. Violet and blue c-plane LDs that omit the top cladding and rely on the exceptionally low refractive index of a Ag or APC (Ag-Pd-Cu) p-contact have been demonstrated [12–14]. However, this approach is limited by the loss in the metallic layers. Another approach to solving this design limitation is to replace a part of the III-nitride top cladding with a relatively low loss transparent conducting oxide (TCO). This design was demonstrated using indium-tin-oxide (ITO) as the top TCO layer in (202¯1) III-nitride LD, with threshold current density (Jth) of 6.2 kA/cm2 and threshold voltage of 8.9 V [15].

As an alternative to ITO, ZnO can be used as the TCO layer for III-nitride LDs. ZnO is a wurtzite semiconductor, with a direct bandgap of 3.3 eV at room temperature [16]. The c-plane lattice mismatch of ZnO to GaN is only ~2%, which facilitates high quality single-crystal growth of ZnO on GaN. ZnO is an attractive alternative to ITO because its optical absorption coefficient [17], is more than one order of magnitude lower than ITO in the blue and green regions of the spectrum [15]. Recent studies found that external quantum efficiency (EQE), luminous efficacy, and series resistance were improved in LEDs with ZnO current spreading layers relative to ITO, indicating that ZnO layers had both lower absorption and lower contact resistance to p-GaN [18]. Introducing dopants has also been used to demonstrate ZnO ohmic contacts to p-GaN [19]. In this letter, we use numerical modeling to investigate the impact of using a lower loss cladding design with ZnO as the TCO layer instead of ITO and demonstrate a blue 202¯1) III-nitride LD with a ZnO top cladding design.

2. Modeling

One benefit of using ZnO as a TCO layer is that it avoids the use of thick p-AlGaN layers and allows the use of thinner p-GaN cladding layers. It is important to understand how the thickness of the p-GaN layers and the choice of TCO affects device performance. To investigate the effect of changing the thickness of the p-GaN layers, 2D optical modes of the proposed LD structure were simulated using Fimmwave, a vectorial mode solver for 2D waveguides (WG)s [20]. GaN, InGaN and AlGaN refractive indices were taken from Goldhahn et. al. [21], and the absorption coefficients were taken from Kioupakis et. al. [22]. The ITO refractive index and absorption coefficient were taken from Hardy et. al. [15]. The ZnO refractive index and absorption coefficient were taken from Reading et. al. [17]. The confinement factor (Γ) and internal loss (αi) were calculated using the Fimmwave real film mode matching method (FMM) and then used to calculate threshold material gain (gth) and differential efficiency (ηd) assuming uncoated laser facets and injection efficiency (ηi) of 65% [23].

The simulated LD structure consisted of 1 µm n-GaN cladding layer, an n-In0.08Ga0.92N WG layer with a thickness between 0 and 100 nm, an active region consisting of three 4.8 nm In0.20Ga0.80N QWs and four 7.6 nm GaN quantum barriers (QBs), a 10 nm Al0.21Ga0.79N electron blocking layer (EBL), a p-In0.08Ga0.92N WG layer with a thickness between 0 and 100 nm, a p-GaN layer with a thickness between 10 to 900 nm, a 10 nm p + GaN contact layer, and 250 nm of ITO or ZnO. The total thickness of the In0.08Ga0.92N WG was fixed at 100 nm and the ridge WG was etched to a depth of 100 nm above the top interface of the last QB.

The use of a TCO in the top cladding creates an asymmetric WG [15,24]. As the p-GaN gets thinner, the modal overlap with the TCO increases and the peak of the mode shifts towards the n-side. For a given p-GaN thickness and a fixed total InGaN WG thickness, the best balance of high Γ and low αi can be achieved by placing a larger fraction of the InGaN WG on the n-side. Our simulations showed that for a fixed p-GaN thickness both Γ and αi decrease as the fraction of the InGaN WG on the n-side increases. This happens due to the reduced overlap of the mode with the QWs, the EBL and the p-GaN layers, all of which have relatively high absorption. Placing the QWs symmetrically in the waveguiding layer (i.e. 50/50 nm n-/p-In0.08Ga0.92N) results in the highest optical confinement, but results in a relatively high internal loss. However, as a higher fraction of the InGaN waveguiding layer is placed on the n-side, internal loss decreases more rapidly than optical confinement. Therefore, Γ can be minimally sacrificed to achieve considerably lower αi. The LD design demonstrated in this work had a 325 nm p-GaN layer and a 70/30 nm n-/p-In0.08Ga0.92N WG, which resulted in a 3% lower Γ and 12% lower αi, compared to a LD structure with a symmetric 50/50 n-/p-In0.08Ga0.92N WG. The demonstrated LD, whose structure and simulated optical mode profile are shown in Fig. 1, had a Γ and αi, of 5.2% and 12.8 cm−1, respectively.

 figure: Fig. 1

Fig. 1 (a) Cross sectional schematic of the LD with ZnO cladding. The structure consisted of a 1μm n-GaN cladding layer, a 70 nm n-In0.08Ga0.92N waveguide (WG), an active region consisting of three 4.8 nm In0.2Ga0.8N quantum wells and four 7.6 nm GaN quantum barriers, a 30 nm p-In0.08Ga0.92N WG, a 10 nm Al0.21Ga0.79N electron blocking layer (EBL), a 325 nm p-GaN layer, a 10 nm p+ GaN contact layer, and 780 nm of ZnO. The thin green layer indicates the location of the p-Al0.21Ga0.79N EBL. The ridge WG was etched to a depth of 100 nm above the top interface of the last quantum barrier. (b) Transverse mode profile and refractive index as a function of distance in the growth direction.

Download Full Size | PDF

The dependence of Γ and αi on p-GaN thickness and TCO composition is shown in Figs. 2(a) and 2(b), respectively. Γ does not depend on the TCO composition, which is expected as ZnO and ITO have almost identical refractive indices in the blue and green regions of the spectrum. For p-GaN thicknesses of less than 100 nm, the optical mode has poor lateral confinement due to the shallow etched ridge WG and Γ decreases significantly [24]. For p-GaN thicknesses of greater than 500 nm, the ZnO and ITO cladding layers have a negligible effect on the optical mode and the Γ and αi show little dependence on p-GaN thickness.

 figure: Fig. 2

Fig. 2 Dependence of (a) confinement factor and (b) internal mode loss on p-GaN thickness for a structure with 250 nm of ITO or ZnO in the top cladding layer.

Download Full Size | PDF

For p-GaN thicknesses between 100 and 500 nm, Γ increases as the p-GaN thickness decreases because the peak intensity of the mode increases in the vicinity of the QWs as the mode is compressed by the thinner p-GaN. For the device with ITO cladding, αi increases as the p-GaN thickness decreases because the overlap of the mode with the high loss ITO cladding increases. The mode overlap with the TCO layer is the same for both ITO and ZnO for a given p-GaN thickness. However, since the absorption coefficient for ZnO (100 cm−1) [17], is much lower than the absorption coefficient for ITO (2000 cm−1) [15], the contribution of the ZnO layer to the overall device loss is negligible for devices with ZnO cladding, even for structures with thin p-GaN layers. This means that devices with ZnO cladding layers havelower overall αi than devices with ITO cladding layers, especially in devices with thin p-GaN layers. This reduction in αi is a key advantage of replacing ITO with ZnO in a TCO clad LD design.

The dependence of gth and ηd on p-GaN thickness and TCO composition are shown in Figs. 3(a) and 3(b), respectively. For these calculations, the cavity length and mean mirror reflection coefficient (R) [25], were assumed to be 1200 μm and 0.18, respectively, corresponding to a mirror loss of 14.3 cm−1. For p-GaN thicknesses of greater than 500 nm, the ZnO and ITO cladding layers have a negligible effect on the optical mode, so gth and ηd show little dependence on p-GaN thickness and are equal for both TCOs. For p-GaN thicknesses between 100 and 500 nm, for the device with ITO cladding, gth increases slightly and ηd decreases significantly as the p-GaN thickness decreases because Γ only increases slight while αi increases significantly with decreasing p-GaN thickness. In contrast, for the device with ZnO cladding, both gth and ηd do not decrease as the p-GaN thickness decreases because Γ increases slightly and αi actually decreases with decreasing p-GaN thickness. The effects of the cladding layer TCO material on optical power output and ηd are particularlymarked and should be even more apparent for state-of-the-art low-loss high-power LD designs [26].

 figure: Fig. 3

Fig. 3 Dependence of (a) threshold material gain and (b) differential efficiency on p-GaN thickness for a structure with 250 nm of ITO or ZnO in the top cladding layer.

Download Full Size | PDF

3. Experimental

LDs were grown on free standing (202¯1) semipolar GaN substrates provided by Mitsubishi Chemical Corporation using atmospheric pressure metalorganic chemical vapor deposition (MOCVD). The LD structure had a 1 μm n-GaN cladding layer, a 70 nm n-In0.08Ga0.92N WG, an AR consisting of three 4.8 nm In0.2Ga0.8N QWs and four 7.6 nm GaN QBs, a 30 nm p- In0.08Ga0.92N WG, a 10 nm Al0.21Ga0.79N EBL, a 325 nm p-GaN layer, and a 10 nm p + -GaN contact layer. A thick n-GaN layer was used for the bottom cladding to avoid the potential relaxation of a strained n-AlGaN cladding layer.

The LDs were fabricated using a self-aligned ridge WG process. Ridge WGs were defined using a BCl3/Cl2 reactive ion etch (RIE) and were etched to a depth of 100 nm above the top interface of the last QB. Following the ridge WG etch, SiO2 was deposited via sputtering on the sidewalls of the ridge WG and in the field for electrical isolation. The samples were then cleaned in a two-step process to remove organic and oxide residue prior to the ZnO deposition. The organic residue was removed with a 10 min RCA standard clean (1:1:5 NH4OH:H2O2) solution heated to 80 °C, which leaves an oxide residue [27]. The oxide residue was removed with a 10 min etch in a 1:1 HCl:H2O solution. A 780 nm thick undoped ZnO layer was then deposited on the exposed p-GaN surface using aqueous solution deposition, the details of which are discussed elsewhere [28,29]. The first step is the growth of a 200 nm thick partially coalesced ZnO seed layer. The seed layer growth is catalyzed at 90 °C by adding ammonium hydroxide (NH4OH) to a zinc nitrate hexa-hydrate (Zn(NO3)2:6H2O) and ammonium nitrate (NH4NO3) solution. To ensure good electrical contact and adhesion, the sample was then annealed at 500 °C in N2 for 15 minutes in a rapid thermal anneal (RTA) system. A second 580 nm thick fully-coalesced ZnO layer was deposited using a 90 °C solution of zinc nitrate hexa-hydroxide (Zn(NO3)2:6H2O), trisodium citrate (NaC6H5O7), and ammonium hydrate (NH4OH). The relatively small chip size (5 mm x 7 mm) ensured uniform concentrations over the deposition area.

After the ZnO deposition, 30 nm of Ti was deposited by electron beam evaporation along the entire length of the laser ridge and then 30/1000 nm Ti/Au was deposited also by electron beam evaporation to form a p-contact pad. The thin Ti layer extended beyond the p-contact pad to ensure uniform current injection. The thin Ti layer also serves to avoid peeling the metal off the ZnO during the facet polishing process. The sample was then diced to form laser bars with a cavity length of 1200 μm. Each facet was polished using diamond polishing films and the grit size was graded from 6.0 μm to 0.1 μm. Facet polishing was completed using 0.05 μm alumina and 0.03 μm silica polishing films, resulting in smooth and featureless facets, as depicted in Fig. 4, which shows a scanning electron microscope (SEM) image of the facet of a fully processed laser, where the epitaxial layers, SiO2, ZnO bilayer and p-contact pad are clearly visible. The laser fabrication was completed with the deposition of a 50/3000 nm Al/Au common back side contact via electron beam evaporation.

 figure: Fig. 4

Fig. 4 Scanning electron microscope (SEM) image of the polished facet of a fully processed laser, where the epitaxial layers, SiO2, ZnO bilayer and p-contact pad are clearly visible.

Download Full Size | PDF

4. Results and discussion

Figure 5(a) shows the light-current-voltage (L-I-V) characteristic for a 1.6 μm wide by 1200 μm long blue LD. The device was tested under pulsed electrical injection with a pulse width of 1.0 μs and a duty cycle of 1%. Lasing was achieved at a threshold current (Ith) of 166 mA, a threshold current density (Jth) of 8.6 kA/cm2 and threshold voltage (Vth) of 10.3 V. The single facet slope efficiency (ηsl) was 0.37 W/A which corresponds to a single facet ηd of 13.5%. Assuming equal power output from both facets, the experimental ηd for both facets was 27%. This value is close to the 32% ηd predicted by optical modeling for this p-GaN thickness (see Fig. 3(b)). Compared to a similar (202¯1) LD structure with ITO top cladding [15], this LD showed a marked improvement in slope efficiency and output power.

 figure: Fig. 5

Fig. 5 (a) Pulsed L-I-V characteristics, (b) current dependent spectra, and (c) the far field pattern of a 1.6 μm wide by 1200 μm long LD.

Download Full Size | PDF

The differential resistance, Rd, was 9 Ω at currents greater than 166 mA. Assuming that the ZnO resistivity is 4.50 x 10-3 Ω-cm [30], the average resistivity in the p-type (Al,In,Ga)N layers is 1 Ω-cm, the average resistivity in the n-type (Al,In,Ga)N layers is 0.001 Ω-cm, and that the resistance contributions of the n-contact, the substrate, the Ti/ZnO interface, and the heterointerfaces are negligible, we estimate that the specific contact resistivity of the ZnO/p-GaN interface is ~1.35 x 10−4 Ω-cm2. Optimizing the aqueous solution deposition methodology and introducing dopants during the deposition should also improve the electrical properties of the ZnO [19], reduce the specific contact resistivity at the ZnO/p-GaN interface, and lower the overall device operating voltage. The current dependent spectra, which exhibit a peak wavelength of 453 nm above threshold, are depicted in Fig. 5(b). Figure 5(c) shows the far field pattern above threshold.

5. Conclusions

In conclusion, we have investigated using ZnO instead of ITO as the top cladding layer in semipolar (202¯1) III-nitride LDs. Numerical modeling indicated that replacing ITO with ZnO reduces the internal loss and threshold gain and increases the differential efficiency for a TCO clad LD design due to the relatively low optical absorption in ZnO. Lasing was achieved at 453 nm with a threshold current density of 8.6 kA/cm2 and a threshold voltage of 10.3 V in a (202¯1) III-nitride LD with ZnO cladding.

Funding

Solid State Lighting and Energy Electronics Center (SSLEEC) at the University of California Santa Barbara (UCSB); Solid State Lighting Program (SSLP), a collaboration between King Abdulaziz City for Science and Technology (KACST), King Abdullah University of Science and Technology (KAUST), and UCSB; National Science Foundation (NSF) National Nanotechnology Infrastructure Network (NNIN) (ECS-0335765); NSF Materials Research Science and Engineering Centers (MRSEC) Program (DMR-1121053).

Acknowledgements

The authors would like to thank A. Hopkins of the UCSB nanofabrication facility for his help with developing the dicing process used to fabricate these devices.

References and links

1. S. Nakamura, N. Iwasa, T. Mukai, and M. Senoh, “Hole compensation mechanism of p-type GaN films,” Jpn. J. Appl. Phys. 31(Part 1, No. 5A), 1258–1266 (1992). [CrossRef]  

2. S. Nakamura, T. Mukai, and M. Senoh, “High-power GaN P-N junction blue-light-emitting diodes,” Jpn. J. Appl. Phys. 30(Part 2, No. 12A), L1998 (1991). [CrossRef]  

3. J. S. Speck and S. F. Chichibu, “Nonpolar and semipolar group III-nitride-based materials,” MRS Bull. 34(05), 304–312 (2009). [CrossRef]  

4. L. Y. Kuritzky and J. S. Speck, “Lighting for the 21st century with laser diodes based on non-basal plane orientations of GaN,” MRS Commun. 5(03), 463–473 (2015). [CrossRef]  

5. J. W. Raring, M. C. Schmidt, C. Poblenz, Y. Chang, M. J. Mondry, B. Li, J. Iveland, B. Walters, M. R. Krames, R. Craig, P. Rudy, J. S. Speck, S. P. Denbaars, and S. Nakamura, “High-efficiency blue and true-green-emitting laser diodes based on non-c-plane oriented GaN substrates,” Appl. Phys. Express 3(11), 112101 (2010). [CrossRef]  

6. D. Sizov, R. Bhat, J. Wang, D. Allen, B. Paddock, and C. E. Zah, “Development of semipolar laser diode,” Phys. Status Solidi Appl. Mater. Sci. 210(3), 459–465 (2013). [CrossRef]  

7. Y. Zhao, R. M. Farrell, Y. Wu, and J. S. Speck, “Valence band states and polarized optical emissiono from nonpolar and semipolar III-nitride quantum well optoelectronic devices,” Jpn. J. Appl. Phys. 53(10), 100206 (2014). [CrossRef]  

8. J. Kim, H. Kim, and S. N. Lee, “Thermal degradation in InGaN quantum wells in violet and blue GaN-based laser diodes,” Curr. Appl. Phys. 11(4), S167–S170 (2011). [CrossRef]  

9. S.-N. Lee, J. Son, H. Paek, T. Sakong, W. Lee, K. Kim, E. Yoon, J. Kim, Y.-H. Cho, Y. Lee, S. Kim, D. Noh, O. Nam, and Y. Park, “Effect of thermal damage on optical and structural properties of In0.08Ga0.92N/In0.02Ga0.98N multi-quantum wells grown by MOCVD,” J. Cryst. Growth 275(1-2), 1041–1045 (2005). [CrossRef]  

10. S. Keller, R. M. Farrell, M. Iza, Y. Terao, N. Young, U. K. Mishra, S. Nakamura, S. P. Denbaars, and J. S. Speck, “Influence of the structure parameters on the relaxation of semipolar InGaN/GaN multi quantum wells,” Jpn. J. Appl. Phys. 52(8S), 08JC10 (2013). [CrossRef]  

11. D. H. Yoon, K. S. Lee, J. B. Yoo, and T. Y. Seong, “Reduction of threading dislocations in InGaN/GaN double heterostructure through the introduction of low-temperature GaN intermediate layer,” Jpn. J. Appl. Phys. 41(Part 1, No. 3A), 1253–1258 (2002). [CrossRef]  

12. D. Bour, C. Chua, Z. Yang, M. Teepe, and N. Johnson, “Silver-clad nitride semiconductor laser diode,” Appl. Phys. Lett. 94(4), 041124 (2009). [CrossRef]  

13. B. Cheng, C. L. Chua, Z. Yang, M. Teepe, C. Knollenberg, A. Strittmatter, and N. Johnson, “Nitride laser diodes with nonepitaxial cladding layers,” IEEE Photonics Technol. Lett. 22(5), 329–331 (2010). [CrossRef]  

14. C. Chua, Z. Yang, C. Knollenberg, M. Teepe, B. Cheng, A. Strittmatter, D. Bour, and N. M. Johnson, “InAlGaN optical emitters-laser diodes with non-epitaxial cladding layers and ultraviolet light-emitting diodes,” Proc. SPIE 7939, 793918 (2011). [CrossRef]  

15. M. T. Hardy, C. O. Holder, D. F. Feezell, S. Nakamura, J. S. Speck, D. A. Cohen, and S. P. Denbaars, “Indium-tin-oxide clad blue and true green semipolar InGaN/GaN laser diodes,” Appl. Phys. Lett. 103(8), 081103 (2013). [CrossRef]  

16. V. Srikant and D. Clarke, “On the optical band gap of zinc oxide,” J. Appl. Phys. 83(10), 5447–5451 (1998). [CrossRef]  

17. A. H. Reading, J. J. Richardson, C.-C. Pan, S. Nakamura, and S. P. DenBaars, “High efficiency white LEDs with single-crystal ZnO current spreading layers deposited by aqueous solution epitaxy,” Opt. Express 20(S1), A13–A19 (2012). [CrossRef]   [PubMed]  

18. A. J. Mughal, S. Oh, A. Myzaferi, S. Nakamura, J. S. Speck, and S. P. DenBaars, “High-power LEDs using Ga-doped ZnO current-spreading layers,” Electron. Lett. 52(4), 304–306 (2016). [CrossRef]  

19. E. Kaminska, A. Piotrowska, K. Golaszewska, M. Guziewicz, R. Kruszka, A. Kudla, T. Ochalski, A. Barcz, T. Dietl, F. Matsukura, M. Sawicki, A. Wawro, M. Zielinski, J. Jasinski, and Z. Liliental-Weber, “Transparent ZnO-based ohmic contact to p-GaN,” Proc. MRS 693, 831 (2002).

20. Fimmwave version 5.3.2, Photon Design (2011).

21. R. Goldhahn, A. T. Winzer, A. Dadgar, A. Krost, O. Weidemann, and M. Eickhoff, “Modulation spectrocsopy of AlGaN/GaN heterostructures: The influence of electron-hole interaction,” Phys. Status Solidi Appl. Mater. Sci. 204(2), 447–458 (2007). [CrossRef]  

22. E. Kioupakis, P. Rinke, and C. G. Van De Walle, “Determination of internal loss in nitride lasers from first principles,” Appl. Phys. Express 3(8), 082101 (2010). [CrossRef]  

23. R. M. Farrell, D. A. Haeger, P. S. Hsu, K. Fujito, D. F. Feezell, S. P. Denbaars, J. S. Speck, and S. Nakamura, “Determination of internal paramters for AlGaN-cladding-free m-plane InGaN/GaN laser diodes,” Appl. Phys. Lett. 99, 17 (2011).

24. A. Myzaferi, A. H. Reading, D. A. Cohen, R. M. Farrell, S. Nakamura, J. S. Speck, and S. P. DenBaars, “Transparent conducting oxide clad limited area epitaxy semipolar III-nitride laser diodes,” Appl. Phys. Lett. 109(6), 061109 (2016). [CrossRef]  

25. L. A. Coldren, S. W. Corzine, and M. L. Masanovic, Diode Lasers and Photonic Integrated Circuits (John Wiley & Sons, Inc., 2012).

26. M. Kawaguchi, O. Imafuji, S. Nozaki, H. Hagino, S. Takigawa, T. Katayama, and T. Tanaka, “Optical-loss supressed InGaN laser diodes using undoped thick waveguide structure,” Proc. SPIE 9748, 974818 (2016). [CrossRef]  

27. W. Kern, Handbook of Semiconductor Wafer Cleaning Technology (Noyes Publications, 1993).

28. J. J. Richardson and F. F. Lange, “Controlling low temperature aqueous synthesis of ZnO. 1. Thermodynamic analysis,” Cryst. Growth Des. 9(6), 2570–2575 (2009). [CrossRef]  

29. J. J. Richardson and F. F. Lange, “Controlling low termpature aqueous synthesis of ZnO. 2. A novel continuous circulation reactor,” Cryst. Growth Des. 9(6), 2576–2581 (2009). [CrossRef]  

30. A. H. Reading, Aqueous Synthesis of Zinc Oxide Films for GaN Optoelectronic Devices (University of California Santa Barbara, 2014).

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 (a) Cross sectional schematic of the LD with ZnO cladding. The structure consisted of a 1μm n-GaN cladding layer, a 70 nm n-In0.08Ga0.92N waveguide (WG), an active region consisting of three 4.8 nm In0.2Ga0.8N quantum wells and four 7.6 nm GaN quantum barriers, a 30 nm p-In0.08Ga0.92N WG, a 10 nm Al0.21Ga0.79N electron blocking layer (EBL), a 325 nm p-GaN layer, a 10 nm p+ GaN contact layer, and 780 nm of ZnO. The thin green layer indicates the location of the p-Al0.21Ga0.79N EBL. The ridge WG was etched to a depth of 100 nm above the top interface of the last quantum barrier. (b) Transverse mode profile and refractive index as a function of distance in the growth direction.
Fig. 2
Fig. 2 Dependence of (a) confinement factor and (b) internal mode loss on p-GaN thickness for a structure with 250 nm of ITO or ZnO in the top cladding layer.
Fig. 3
Fig. 3 Dependence of (a) threshold material gain and (b) differential efficiency on p-GaN thickness for a structure with 250 nm of ITO or ZnO in the top cladding layer.
Fig. 4
Fig. 4 Scanning electron microscope (SEM) image of the polished facet of a fully processed laser, where the epitaxial layers, SiO2, ZnO bilayer and p-contact pad are clearly visible.
Fig. 5
Fig. 5 (a) Pulsed L-I-V characteristics, (b) current dependent spectra, and (c) the far field pattern of a 1.6 μm wide by 1200 μm long LD.
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.