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

Optimization of the electroluminescence from SiNx-based light-emitting devices by modulating the size and morphology of silver nanostructures

Open Access Open Access

Abstract

A maximal enhancement of ~6.5 times of the external quantum efficiency (EQE) for SiNx-based light-emitting devices (LEDs) is achieved by magnetron sputtering a silver nanostructures layer onto the active matrix. The enhancement of EQE is affected by the dimension and morphology of silver nanostructures, which can be controlled by the sputtering time and the post treatment of rapid thermal annealing. The optimal size of silver nanostructures is about 100 nm in diameter by comparing the integrated electroluminescence intensity under the same input power. The optimization of EQE for SiNx-based LEDs is discussed by considering the contributions of the enhancement of light-extraction efficiency induced by the surface roughening of the front electrode, internal quantum efficiency due to the coupling between excitons and localized surface plasmons, and carrier injection efficiency. Our work may provide an alternative approach for the fabrication of Si-based light sources with promising luminescence efficiency.

©2013 Optical Society of America

1. Introduction

In the last decades, silicon-rich silicon nitride (SiNx, x<1.33) films have attracted a great research interest in a wild range of fields, including the memory medium in nonvolatile metal-insulator-semiconductor (MIS) devices [1], passivation layers and antireflection coatings for current Si solar cells [2, 3], optical waveguides [4], and light-emitting devices (LEDs) [57]. For the application on LEDs as a candidate of Si-based light sources, enormous efforts have been devoted to SiNx films fabricated by plasma-enhanced chemical vapor deposition (PECVD) technique due to its conspicuous and tunable room-temperature photoluminescence (PL) with a promising electroluminescence (EL) efficiency [5, 8, 9]. Besides the SiNx matrix, recent progress had also been focused on the Si quantum dots (QDs) based LEDs [1012] and the Ge-on-Si emitters with the band of Ge engineered by tensile strain and n-type doping [13, 14] as the alternative Si-based light sources. Generally, the formation of Si QDs should experience a high-temperature thermal annealing procedure (usually >1000 °C) [11, 12], and the Ge-on-Si lasers would suffer an extremely high threshold current density and poor efficiency [14]. Moreover, compared with its congener silicon-rich silicon oxide (SiOx) which is also commonly used for Si-based light sources [15], SiNx provides a lower carrier injection barrier that may realize a higher external quantum efficiency (EQE) [16]. Meanwhile, the fabrication of SiNx-based LEDs was a complementary metal-oxide semiconductor (CMOS) compatible process. By employing a SiNx/SiNy multilayer structure, Huh et al. achieved an EQE of 10−3% for SiNx-based LEDs [17]. This value is still too low to achieve the demand of Si-based light sources for the application in optical interconnections. The partial reason for this insufficient EL efficiency may result from its unbalanced carrier injection barrier heights. The injection barrier height for electrons is ~3.0 eV (from ITO electrode to SiNx matrix), which is much higher than that for holes on the p-type Si side (1.9 eV) [5]. Besides, the strong nonradiative recombination derived from the band tails and defect states in SiNx due to the structural disorder also constrains its EL efficiency [5].

The inherent problems mentioned above in SiNx films prevent it from being used in actual applications as Si-based light sources. Consequently, various methods were proposed to improve the EL performance of SiNx-based LEDs in recent years. For the purpose of improving the current injection, Kim et al. used the Ni/Au contact placed on SiNx layer, obtaining an EQE of (~3.3 x 10−3)% [18]. Meanwhile, by employing the n-type SiC layer to balance the carrier injection levels accompanied with a periodic micron-scale rugged surface pattern produced by optical lithographic and a standard Si etching technique to increase the light extraction from SiNx film, Kim et al. achieved a 2.8 times enhancement of the light output power [19]. On the other hand, the nonradiative recombination can be decreased significantly by the reduction of SiNx/Si interfacial state density or the passivation of nonradiative defects in the SiNx films using NH3 plasma pretreatments, which can further increase the EL efficiency of SiNx-based LEDs [20]. Besides, the EL efficiency can be enhanced by depositing Ag or Au nanostructures underneath the luminescence layer, which is originated from the coupling between the excitons and localized surface plasmons (LSPs, the collective oscillations of excited free electrons occurring at the interface between a metal and a dielectric layer, which is confined to the surrounding of metallic nanostructures) [7, 21].

Recently, the application of LSPs on LEDs has been received a striking attention not only for the improvement of carrier injection resulted from the enhancement of the inhomogeneous local electric fields at the interface between metals and luminescence matrixes [22], but also for the increase of radiative efficiency originated from the strong coupling between exciton dipole moments and local electric field of LSPs [7]. The use of surface plasmon coupled to the excitons in active matrixes had been reported in (In)GaN-based LEDs for achieving the improved spontaneous emission rate from Purcell factor enhancement [2325]. As have been investigated, this coupling is stronger at the frequency near its plasmon resonance affected by the size, shape, and dielectric environment of metal particles [21]. Among all the metals with plasmon resonances at visible frequencies, which are near the wavelength of EL peak for SiNx-based LEDs, Ag provides the lowest absorption losses [26] and the strongest resonances [27].

In our previous work, a Ag nanostructures layer was deposited onto or underneath the SiNx matrix to improve the light extraction efficiency (LEE) via the roughening of the ITO electrode, and an almost one order of magnitude enhancement of EQE for SiNx-based LEDs was achieved [28, 29]. Besides, we found that the improvement of the LEE via embedding the Ag nanostructures underneath the luminescence layer was not only originated from the surface roughening of the front electrode, but also originated from the increased back-scattering due to the LSP resonance coupling [30]. Moreover, we fabricated the Ag nanostructures with various sizes and surface coverage onto the luminescence matrix by the modulation of the thermal annealing procedure, and obtained that the density of Ag nanostructures has a significant influence on the improvement of the PL intensity for the SiNx films with different N/Si ratios [31]. Furthermore, the coupling between LSPs and excitons as well as the determination of the average position of excitons in SiNx matrix was demonstrated via the Purcell effect [32]. Further investigations of the other effects of LSP resonance on the improvement of the EQE and/or the evolution of EL from SiNx-based LEDs are also in progress. In this paper, we demonstrate the optimization process of the EL efficiency from SiNx-based LEDs via depositing a Ag nanostructures layer with different dimensions and morphologies onto the SiNx matrix by employing a simple and COMS compatible fabrication process. The separation analysis of the three contributions of the LEE, carrier injection efficiency, and radiative efficiency enhancement on the improvement of the EL efficiency is provided. The metal nanostructures with highly-uniform close-packed two-dimensional arrays or with highly ordered patterns can be fabricated by using the diblock-copolymer lithography templates [33, 34] or using a rapid convective deposition method [3537]. In this work, the Ag nanostructures layer was formed by rapid thermal annealing (RTA) the sputtered Ag film, and the dimension distribution of Ag nanostructures is relative random and large.

2. Experimental

Boron-doped p/p+-Si (100) wafer (with the thickness and resistivity of the epitaxial layer were 18.5 ± 1 µm and 9 ± 1.5 Ω·cm, respectively) was employed as a substrate for the fabrication of SiNx-based LED. A PECVD system was used to deposit SiNx films, in which N2-diluted 10% SiH4 and NH3 were used as the reactant gas sources. The flow rate ratio of SiH4 and NH3, substrate temperature, chamber pressure, and r.f. power density for the deposition were maintained at 1:1, 400 °C, 0.1 Torr, and ~1 kW/m2, respectively. After the deposition of SiNx films, the samples were annealed at 800 °C for 1 hour in N2 ambience. To investigate the effect of Ag nanostructures on the EL efficiency of SiNx-based LEDs, a Ag layer was deposited by magnetron sputtering onto the SiNx films. The thickness of Ag was regulated by the sputtering time, which was also used to label the samples, e.g. Ag20 refers to the sample with the sputtering time of 20 s. After the post-treatment of rapid thermal annealing (RTA) in argon at 500 °C for 1 min, Ag nanostructures with various dimensions and surface morphologies were formed. For the investigation of the electrical properties of SiNx-based LEDs with and without Ag nanostructures, a top transparent ITO circular electrode (~100 nm) with the diameter of 1.0 cm was deposited by magnetron sputtering using a circular mask and an Al metal contact of about 200 nm was evaporated on the backside of the Si substrates. The fabrication process and the final device structures of the SiNx-based LEDs are shown in Fig. 1. A SiNx-based LED without the addition of Ag nanostructures was also fabricated for comparison, which was labeled as Ag0. Meanwhile, we also fabricated the quartz/SiNx/Ag structures for the measurement of the extinction spectra of the Ag nanostructures with different sizes and morphologies.

 figure: Fig. 1

Fig. 1 Fabrication process diagrams of the SiNx-based LEDs.

Download Full Size | PDF

The thickness of SiNx films was around 50 nm measured by ellipsometry, J. A. Woollam M-2000D. The scanning electron microscopy (SEM, Hitachi S-4800) was used to characterize the Ag nanostructure size and the surface morphology. A Veeco Digital Instruments Innova atomic force microscopy (AFM) was employed for the characterization of the surface morphology of the ITO electrode with and without the addition of Ag nanostructures. Both the photoluminescence (PL) and EL signals of SiNx-based LEDs were recorded by an Acton SpectraPro-2500i monochrometer coupled to a photomultiplier tube (PMT). The PL signals were excited by a 325 nm He-Cd laser, while the EL ones were driven by a DC power source. The current-voltage (I-V) characteristics of the devices were measured by a Keithley 4200 SCS semiconductor parameter analyzer. A HITACHI U-4100 spectrophotometer was employed to measure the extinction spectra of the SiNx films with and without Ag nanostructures.

3. Results and discussion

Figure 2 shows the SEM images and the size distributions of Ag nanostructures formed by the post-treatment of RTA with the sputtering time ranging from 20 s to 80 s. The Ag nanostructures reveal a good dispersion with the average dimension (D) from ~40 nm to ~240 nm in diameter for the sputtering time from 20 s to 80 s. The aggregation of particles becomes more and more serious with the increase of the average sizes. The primary shapes are circles with a little trend of being ellipses, which is gradually noticeable as the average diameter increases. Moreover, with the increase of the average diameter of Ag nanostructures, the dipolar resonance peak in extinction spectra gradually red-shifts, as shown in Fig. 3. The intensity of the extinction spectra for Ag nanostructures is significantly larger than that for the SiNx films [28], and the relative intensity of the dipolar resonance peak is gradually decreased with the increasing size of Ag nanostructures (not shown here). An obvious extinction valley can be observed at short wavelength, which also gradually red-shifts with the increasing size of Ag nanostructures. Interestingly, both the extinction peak and valley for Ag80 blue-shift a little compared with that for Ag60, which may result from the Ag nanostructures with smaller sizes due to the nonuniformed distribution of the dimensions for Ag80 or from the minor axis of ellipsoidal-like Ag nanostructures, as shown in Fig. 2(d). The quadrupolar resonance peak can be observed clearly in the extinction spectra of Ag80, as shown in Fig. 3.

 figure: Fig. 2

Fig. 2 (a)-(d) SEM images and (e)-(h) the size distributions of Ag nanostructures with sputtering time (a) and (e) 20 s, (b) and (f) 40 s, (c) and (g) 60 s, and (d) and (h) 80 s.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Extinction spectra of Ag nanostructures with different sputtering time.

Download Full Size | PDF

The effect of Ag nanostructures with different sizes on the EL efficiency of SiNx-based LEDs was investigated at room temperature for further improving the luminescence performance of the devices. As can be observed in Figs. 4(c)-4(f), two EL peaks can be resolved for the devices with the addition of Ag nanostructures, with the longer wavelength peaks in the range of 620-780 nm and the shorter ones at 500-600 nm. While for the sample without the addition of Ag nanostructures (Ag0), only one shorter wavelength peak (~600 nm) can be observed, as shown in Figs. 4(a) and 4(b). Meanwhile, the peak with longer wavelength can be resolved hardly for the sample with the addition of Ag nanostructures of larger sizes (Ag80), as shown in Figs. 4(e) and 4(f). We speculate that this phenomenon may be due to its poor injection of electrons as the barrier height of electrons on ITO side (3.0 eV) is much higher than that of holes on p-Si side (1.9 eV) [5]. We attributed all these peaks to the recombination of electrons and holes at defect levels or tail states lie in SiNx bandgap in the range from ~400 nm to ~780 nm [5, 38, 39]. The peak at the shorter wavelength (~500-600 nm) is due to the radiative recombination between the K center (≡Si-) and the valence band tail levels composed by N dangling bands ( = N-) and Si-Si bonding states, which are in the range of ~350-650 nm [38, 39]. Meanwhile, we attribute the one at the longer wavelength to the recombination of electrons at conduction band tail states and holes trapped in the center of ≡Si0 [38, 39]. The weakening of the longer wavelength EL peak, as shown in Fig. 4(f), may be originated from the insignificantly improved electron injection on ITO side, where only an extremely small amount of electrons can be injected into the conduction band and contribute to the EL.

 figure: Fig. 4

Fig. 4 EL spectra of SiNx-based LEDs for (a) Ag0, (c) Ag40, and (e) Ag80 measured at different injected current. Their Gauss fittings are also provided on the right side, (b) for Ag0, (d) for Ag40, and (f) for Ag80.

Download Full Size | PDF

For the investigation of the size and morphology modulation effect of Ag nanostructures on the improvement of EL efficiency, the integrated EL intensity of SiNx-based LEDs with and without the addition of Ag nanostructures is compared, as shown in Fig. 5. Obviously, a maximal enhancement of EL efficiency of about 6.5 times is achieved for the device with the average size of Ag nanostructures ~100 nm in diameter by comparing the integrated EL intensity of Ag40 to that of Ag0 under the almost same input power. Three aspects including the increase of LEE, internal quantum efficiency (IQE), and carrier injection efficiency should be considered to determine the reasons for the optimal EL performance of Ag40.

 figure: Fig. 5

Fig. 5 Integrated EL intensity of the devices with different sputtering times of Ag nanostructures layer vs. the input power.

Download Full Size | PDF

The improvements of IQE for Ag20-Ag80 can be estimated semiquantitatively by simply comparing the enhancement factor of integrated PL intensity [28], which is defined by the ratio of the integrated PL intensity of the devices with Ag nanostructures to that of Ag0, as shown in Fig. 6(a). A maximum improvement of ~1.8 times is achieved for Ag40. We attribute this improvement mainly to the enhancement of the spontaneous emission rate due to the coupling between the excitons in SiNx and LSPs [7, 32, 40], which is confirmed by the consistence between the enhancement factor of EL intensities and that of PL ones, as shown in Fig. 6(b). The enhancement of the flow of light emitted from the active layer due to the minimum absorption of Ag nanostructures also has a contribution to this improvement, as shown in Fig. 2 and Fig. 6(b). Obviously, the improvement of IQE is not the main contribution to the enhancement of EL efficiency in our SiNx-based LEDs by further considering the surface coverage of Ag nanostructures (ηs), as shown in Fig. 6(c) left. The serious backscattering induced by the high ηs (>40%) of Ag nanostructures will make it difficult for light to escape from the active layer. The contribution of this improvement to the EL efficiency will be more insignificantly by considering the absorption of Ag nanostructures further.

 figure: Fig. 6

Fig. 6 (a) Enhancement factor of the integrated PL intensity for Ag20-Ag80 comparing to Ag0, (b) EL and PL enhancement factor for Ag40, and (c) The surface coverage of Ag nanostructures (left axis) and the average distance between Ag nanostructures (right axis) for Ag20-Ag80 with the model for this calculation inset.

Download Full Size | PDF

Interestingly, the same tendencies of the enhancement factor of integrated PL intensity and the ηs with the increase of D are observed, as shown in Figs. 6(a) and 6(c). The improvements of IQE for Ag60 and Ag80 are higher than that for Ag40 theoretically by simply comparing the enhancement factors of integrated PL intensity and the ηs. Considering the origin of the enhancement of IQE, the distance between Ag nanostructures should also be taken account of, which will affect the spontaneous emission rate significantly [21]. A simple model is adopted for the calculation of the average distance (d) between Ag nanostructures, as shown in the inset of Fig. 6(c). The value of d can be obtained from the formula:

ηs=π[r/(d+2r)]2
where r = D/2 is the average radii of Ag nanostructures. This value is raised quickly with the increase of D, as shown in Fig. 6(c) (right axis). The contribution of the interaction between dipoles to the enhancement of the spontaneous emission rate will decrease significantly as the intensity of the local electromagnetic fields is decayed exponentially away from the metal surface [7, 40]. Consequently, the maximum enhancement of IQE may be obtained for Ag40 among the devices fabricated here. Further improvement of IQE can be achieved by the optimization of ηs and D of Ag nanostructures. The abnormal phenomenon for the enhancement factors of integrated PL intensity of Ag60 and Ag80, which have the similar values of d, may result from the improved light-extraction efficiency of Ag80 comparing with that of Ag60.

The roughening of ITO electrode surface via the introduction of Ag nanostructures, as shown in Fig. 7, can decrease the inner-reflection of light and improve the LEE of SiNx-based LEDs significantly [30, 41]. The surface of ITO electrode for the device Ag0 is rather smooth with the area root mean square (RMS) roughness (δ) only ~2.0 nm, as shown in Fig. 7(a). After the addition of Ag nanostructures, this value of δ increases monotonically with the increased D of Ag nanostructures. The contribution of this surface roughening on the improvement of the LEE for our LEDs can be estimated semiquantitatively based on the total integrated scatter (TIS), which is defined as the total power scattered into a hemisphere divided by the incident power [42]. This value of TIS is proportional to the square of δ under the assumption that the scattering mainly occurs at the regular reflection direction [42]. An obvious enhancement of LEE can be achieved by this roughening of ITO electrode surface even considering the surface coverage and the absorption of Ag nanostructures on the suppression of light extraction. The significant increase of LEE also confirms that the improvement of IQE is not the main contribution to the enhancement of EL efficiency of our SiNx-based LEDs, as have been mentioned above. Besides, the increased δ of ITO electrode would also enhance the nonradiative recombination due to its increased interface state between SiNx matrix and ITO. Consequently, there is a tradeoff between the improvement of LEE and the enhancement of nonradiative recombination via the roughening of ITO electrode surface, and it is difficult to differentiate them. Due to the extremely large δ of ITO electrode for the sample Ag80, the overall effect of this roughening on the improvement of EL efficiency might be limited. Better improvement may be achieved by the addition of Ag nanostructures with smaller size and suitable roughness. Hence, the maximum enhancement of LEE via the surface roughening of ITO electrode may also be achieved for Ag40 among our devices. Further improvement of EL efficiency can be achieved by the optimization of the surface roughness of ITO electrode.

 figure: Fig. 7

Fig. 7 AFM images of ITO for (a) Ag0, (b) Ag40, and (c) Ag80.

Download Full Size | PDF

The effect of the improvement of carrier injection efficiency on the optimization of EL efficiency for our devices is also checked via J-V characteristics, as shown in Fig. 8. A significant enhancement of carrier injection for the devices Ag20 and Ag40 is achieved comparing to the device Ag0. While for the device Ag60 and Ag80, this enhancement is not so much obviously, as shown in Fig. 8. We attribute the improvement of the carrier injection to the increased surface roughness due to the addition of Ag nanostructures, by which a large enhancement of the localized electric field at the interface can be obtained [21]. Due to the rather short distance between Ag nanostructures and high surface coverage of Ag nanostructures for Ag20 and Ag40, as shown in Fig. 2 and Fig. 6(c), the transport of carriers between Ag nanostructures may be more serious for Ag20 and Ag40 comparing to that for Ag60 and Ag80, which have a longer particles’ distance and a lower surface coverage. Better conductance is achieved for the devices Ag20 and Ag40 than that for Ag60 and Ag80. However, the accumulation of carriers at the interface states between the contact electrode and the active material will also be more serious for Ag20 and Ag40 than that for Ag60 and Ag80, from which the limited effective injection will be more obvious due to the screened applied electric field [43]. Interestingly, there is a similar relation between the relative intensity of EL peak with longer wavelength and the improvement of carrier injection. For the sample Ag40, the improvement of carrier injection, especially for electrons, is significant, as shown in Fig. 8, and the EL peak with longer wavelength can be resolved distinctly, as shown in Figs. 4(c) and 4(d). While for the sample Ag80, this improvement of electron injection is not so much obvious and the longer wavelength EL peak can hardly be resolved, as shown in Figs. 4(e) and 4(f). Consequently, the evolution of the EL peak with longer wavelength is originated from the changes of the carrier injection levels, as have been speculated above. From the discussion above, the improvement of the carrier injection may also have an instructive contribution to the enhancement of EL efficiency.

 figure: Fig. 8

Fig. 8 Current density (J) vs. applied voltage (V) plot for Ag0-Ag80 with the values of J for Ag0, Ag60, and Ag80 multiplied by 10 times.

Download Full Size | PDF

4. Conclusion

In summary, an enhanced EL efficiency of SiNx-based LED is observed by depositing a Ag nanostructures layer on its luminescence layer. By optimizing the size and morphology of Ag nanostructures, a maximal 6.5 times enhancement of integrated EL intensity is achieved at the same input power. Two distinct EL Gauss peaks originated from the radiative recombination between its corresponding band tails and trapped levels can be resolved after the addition of Ag nanostructures. We attribute the enhancement of EL efficiency mainly to the improvement of light-extraction efficiency caused by the surface roughening of ITO electrode, as well as the improved carrier injection efficiency. The improvement of internal quantum efficiency might also have an instructive contribution to this enhancement. Further detailed studies are required to quantify the various contributions of the individual efficiency enhancements. Nevertheless, the current work pointed out the importance of the dimensions of the Ag nanostructures on the optimized EL efficiency of the SiNx-based LEDs. Moreover, our results may provide an alternative approach towards the fabrication of SiNx-based or Si-based LEDs with high EL efficiency. Meanwhile, we also provide the possible methods on further improving the EL efficiency of SiNx-based LEDs, which can be achieved by the optimization of the surface coverage (ηs) of Ag nanostructures, the average distance (d) between Ag nanostructures, and/or the RMS roughness (δ) of ITO surface.

Acknowledgment

This work is funded by the 863 Program (No. 2011AA050517), the National Natural Science Foundation of China (No.61176117), and the Innovation Team Project of Zhejiang Province (No. 2009R50005).

Reference and links

1. H. J. Stein, “Thermally annealed silicon nitride films: Electrical characteristics and radiation effects,” J. Appl. Phys. 57(6), 2040–2047 (1985). [CrossRef]  

2. G. Aberle, “Surface passivation of crystalline silicon solar cells: a review,” Prog. Photovolt. Res. Appl. 8(5), 473–487 (2000). [CrossRef]  

3. P. Doshi, G. E. Jellison Jr, and A. Rohatgi, “Characterization and optimization of absorbing plasma-enhanced chemical vapor deposited antireflection coatings for silicon photovoltaics,” Appl. Opt. 36(30), 7826–7837 (1997). [CrossRef]   [PubMed]  

4. L. Zhuang, D. Marpaung, M. Burla, W. Beeker, A. Leinse, and C. Roeloffzen, “Low-loss, high-index-contrast Si₃N₄/SiO₂ optical waveguides for optical delay lines in microwave photonics signal processing,” Opt. Express 19(23), 23162–23170 (2011). [CrossRef]   [PubMed]  

5. M. Wang, J. Huang, Z. Yuan, A. Anopchenko, D. Li, D. Yang, and L. Pavesi, “Light emission properties and mechanism of low-temperature prepared amorphous SiNX film. II. Defect states electroluminescence,” J. Appl. Phys. 104(8), 083505 (2008). [CrossRef]  

6. Z. H. Cen, T. P. Chen, Z. Liu, Y. Liu, L. Ding, M. Yang, J. I. Wong, S. F. Yu, and W. P. Goh, “Electrically tunable white-color electroluminescence from Si-implanted silicon nitride thin film,” Opt. Express 18(19), 20439–20444 (2010). [CrossRef]   [PubMed]  

7. B.-H. Kim, C.-H. Cho, J.-S. Mun, M.-K. Kwon, T.-Y. Park, J.-S. Kim, C.-C. Byeon, J. Lee, and S.-J. Park, “Enhancement of the external quantum efficiency of a silicon quantum dot light-emitting diode by localized surface plasmons,” Adv. Mater. (Deerfield Beach Fla.) 20(16), 3100–3104 (2008). [CrossRef]  

8. N.-M. Park, C.-J. Choi, T.-Y. Seong, and S.-J. Park, “Quantum confinement in amorphous silicon quantum dots embedded in silicon nitride,” Phys. Rev. Lett. 86(7), 1355–1357 (2001). [CrossRef]   [PubMed]  

9. M. Wang, M. Xie, L. Ferraioli, Z. Yuan, D. Li, D. Yang, and L. Pavesi, “Light emission properties and mechanism of low-temperature prepared amorphous SiNX films. I. Room temperature band tail states photoluminescence,” J. Appl. Phys. 104(8), 083504 (2008). [CrossRef]  

10. G.-R. Lin, C.-J. Lin, and C.-K. Lin, “Enhanced Fowler-Nordheim tunneling effect in nanocrystallite Si based LED with interfacial Si nano-pyramids,” Opt. Express 15(5), 2555–2563 (2007). [CrossRef]   [PubMed]  

11. T. Creazzo, B. Redding, E. Marchena, J. Murakowski, and D. W. Prather, “Pulsed pumping of silicon nanocrystal light emitting devices,” Opt. Express 18(11), 10924–10930 (2010). [CrossRef]   [PubMed]  

12. C.-H. Cheng, C.-L. Wu, C.-C. Chen, L.-H. Tsai, Y.-H. Lin, and G.-R. Lin, “Si-rich SixC1-x light-emitting diodes with buried Si quantum dots,” IEEE Photon. J 4(5), 1762–1775 (2012). [CrossRef]  

13. J. Liu, X. Sun, L. C. Kimerling, and J. Michel, “Direct-gap optical gain of Ge on Si at room temperature,” Opt. Lett. 34(11), 1738–1740 (2009). [CrossRef]   [PubMed]  

14. B. Dutt, D. S. Sukhdeo, D. Nam, B. M. Vulovic, Ze Yuan, and K. C. Saraswat, “Roadmap to an efficient germanium-on-silicon laser: strain vs. n-type doping,” IEEE Photon. J 4(5), 2002–2009 (2012). [CrossRef]  

15. R. J. Walters, G. I. Bourianoff, and H. A. Atwater, “Field-effect electroluminescence in silicon nanocrystals,” Nat. Mater. 4(2), 143–146 (2005). [CrossRef]   [PubMed]  

16. G. R. Lin, Y. H. Pai, C. T. Lin, and C. C. Chen, “Comparison on the electroluminescence of Si-rich SiNx and SiOx based light-emitting diodes,” Appl. Phys. Lett. 96(26), 263514 (2010). [CrossRef]  

17. C. Huh, K.-H. Kim, B. K. Kim, W. Kim, H. Ko, C.-J. Choi, and G. Y. Sung, “Enhancement in light emission efficiency of a silicon nanocrystal light-emitting diode by multiple-luminescent structures,” Adv. Mater. (Deerfield Beach Fla.) 22(44), 5058–5062 (2010). [CrossRef]   [PubMed]  

18. B.-H. Kim, C.-H. Cho, S.-J. Park, N.-M. Park, and G. Y. Sung, “Ni/Au contact to silicon quantum dot light-emitting diodes for the enhancement of carrier injection and light extraction efficiency,” Appl. Phys. Lett. 89(6), 063509 (2006). [CrossRef]  

19. K.-H. Kim, J.-H. Shin, N.-M. Park, C. Huh, T.-Y. Kim, K.-S. Cho, J. C. Hong, and G. Y. Sung, “Enhancement of light extraction from a silicon quantum dot light-emitting diode containing a rugged surface pattern,” Appl. Phys. Lett. 89(19), 191120 (2006). [CrossRef]  

20. D. Li, J. Huang, and D. Yang, “Enhanced electroluminescence of silicon-rich silicon nitride light-emitting devices by NH3 plasma and annealing treatment,” Physica E 41(6), 920–922 (2009). [CrossRef]  

21. J. S. Biteen, D. Pacifici, N. S. Lewis, and H. A. Atwater, “Enhanced radiative emission rate and quantum efficiency in coupled silicon nanocrystal-nanostructured gold emitters,” Nano Lett. 5(9), 1768–1773 (2005). [CrossRef]   [PubMed]  

22. R. T. Tung, “Electron transport at metal-semiconductor interfaces: General theory,” Phys. Rev. B Condens. Matter 45(23), 13509–13523 (1992). [CrossRef]   [PubMed]  

23. K. Okamoto, I. Niki, A. Shvartser, Y. Narukawa, T. Mukai, and A. Scherer, “Surface-plasmon-enhanced light emitters based on InGaN quantum wells,” Nat. Mater. 3(9), 601–605 (2004). [CrossRef]   [PubMed]  

24. H. Zhao, J. Zhang, G. Liu, and N. Tansu, “Surface plasmon dispersion engineering via double-metallic Au/Ag layers for III-nitride based light-emitting diodes,” Appl. Phys. Lett. 98(15), 151115 (2011). [CrossRef]  

25. C.-H. Lu, C.-C. Lan, Y.-L. Lai, Y.-L. Li, and C.-P. Liu, “Enhancement of green emission from InGaN/GaN multiple quantum wells via coupling to surface plasmons in a two-dimensional silver array,” Adv. Funct. Mater. 21(24), 4719–4723 (2011). [CrossRef]  

26. S. Pillai, K. R. Catchpole, T. Trupke, and M. A. Green, “Surface plasmon enhanced silicon solar cells,” J. Appl. Phys. 101(9), 093105 (2007). [CrossRef]  

27. J. S. Biteen, N. S. Lewis, H. A. Atwater, H. Mertens, and A. Polman, “Spectral tuning of plasmon-enhanced silicon quantum dot luminescence,” Appl. Phys. Lett. 88(13), 131109 (2006). [CrossRef]  

28. F. Wang, D. Li, D. Yang, and D. Que, “Enhancement of light-extraction efficiency of SiNx light emitting devices through a rough Ag island film,” Appl. Phys. Lett. 100(3), 031113 (2012). [CrossRef]  

29. D. Li, F. Wang, C. Ren, and D. Yang, “Improved electroluminescence from silicon nitride light emitting devices by localized surface plasmons,” Opt. Mater. Express 2(6), 872–877 (2012). [CrossRef]  

30. F. Wang, D. Li, D. Yang, and D. Que, “Enhancement of orange-yellow electroluminescence extraction from SiNx light-emitting devices by silver nanostructures,” Opt. Express . in progress.

31. F. Wang, M. Wang, D. Li, and D. Yang, “Localized surface plasmon resonance enhanced photoluminescence from SiNx with different N/Si ratios,” Opt. Mater. Express 2(10), 1437–1448 (2012). [CrossRef]  

32. F. Wang, D. Li, D. Yang, and D. Que, “The coupling between localized surface plasmons and excitons via Purcell effect,” Nanoscale Res. Lett. 7(1), 669 (2012). [CrossRef]   [PubMed]  

33. T. F. Kuech and L. J. Mawst, “Nanofabrication of III–V semiconductors employing diblock copolymer lithography,” J. Phys. D Appl. Phys. 43(18), 183001 (2010). [CrossRef]  

34. G. Liu, H. Zhao, J. Zhang, J. H. Park, L. J. Mawst, and N. Tansu, “Selective area epitaxy of ultra-high density InGaN quantum dots by diblock copolymer lithography,” Nanoscale Res. Lett. 6(1), 342 (2011). [CrossRef]   [PubMed]  

35. Y.-K. Ee, R. A. Arif, N. Tansu, P. Kumnorkaew, and J. F. Gilchrist, “Enhancement of light extraction efficiency of InGaN quantum wells light emitting diodes using SiO2/polystyrene microlens arrays,” Appl. Phys. Lett. 91(22), 221107 (2007). [CrossRef]  

36. Y. K. Ee, P. Kumnorkaew, R. A. Arif, H. Tong, J. F. Gilchrist, and N. Tansu, “Light extraction efficiency enhancement of InGaN quantum wells light-emitting diodes with polydimethylsiloxane concave microstructures,” Opt. Express 17(16), 13747–13757 (2009). [CrossRef]   [PubMed]  

37. X.-H. Li, R. Song, Y.-K. Ee, P. Kumnorkaew, J. F. Gilchrist, and N. Tansu, “Light extraction efficiency and radiation patterns of III-nitride light-emitting diodes with colloidal microlens arrays with various aspect ratios,” IEEE Photon. J. 3(3), 489–499 (2011). [CrossRef]  

38. J. Robertson and M. J. Powell, “Gap states in silicon-nitride,” Appl. Phys. Lett. 44(4), 415–417 (1984). [CrossRef]  

39. D. Li, F. Wang, D. Yang, and D. Que, “Electrically tunable electroluminescence from SiNx-based light-emitting devices,” Opt. Express 20(16), 17359–17366 (2012). [CrossRef]   [PubMed]  

40. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [CrossRef]   [PubMed]  

41. Y. Liu, J. Xu, H. Sun, S. Sun, W. Xu, L. Xu, and K. Chen, “Depth-dependent anti-reflection and enhancement of luminescence from Si quantum dots-based multilayer on nano-patterned Si substrates,” Opt. Express 19(4), 3347–3352 (2011). [CrossRef]   [PubMed]  

42. A. I. Zhmakin, “Enhancement of light extraction from light emitting diodes,” Phys. Rep. 498(4–5), 189–241 (2011). [CrossRef]  

43. A. Marconi, A. Anopchenko, G. Pucker, and L. Pavesi, “Power efficiency estimation of silicon nanocrystals based light emitting devices in alternating current regime,” Appl. Phys. Lett. 98(20), 201103 (2011). [CrossRef]  

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 (8)

Fig. 1
Fig. 1 Fabrication process diagrams of the SiNx-based LEDs.
Fig. 2
Fig. 2 (a)-(d) SEM images and (e)-(h) the size distributions of Ag nanostructures with sputtering time (a) and (e) 20 s, (b) and (f) 40 s, (c) and (g) 60 s, and (d) and (h) 80 s.
Fig. 3
Fig. 3 Extinction spectra of Ag nanostructures with different sputtering time.
Fig. 4
Fig. 4 EL spectra of SiNx-based LEDs for (a) Ag0, (c) Ag40, and (e) Ag80 measured at different injected current. Their Gauss fittings are also provided on the right side, (b) for Ag0, (d) for Ag40, and (f) for Ag80.
Fig. 5
Fig. 5 Integrated EL intensity of the devices with different sputtering times of Ag nanostructures layer vs. the input power.
Fig. 6
Fig. 6 (a) Enhancement factor of the integrated PL intensity for Ag20-Ag80 comparing to Ag0, (b) EL and PL enhancement factor for Ag40, and (c) The surface coverage of Ag nanostructures (left axis) and the average distance between Ag nanostructures (right axis) for Ag20-Ag80 with the model for this calculation inset.
Fig. 7
Fig. 7 AFM images of ITO for (a) Ag0, (b) Ag40, and (c) Ag80.
Fig. 8
Fig. 8 Current density (J) vs. applied voltage (V) plot for Ag0-Ag80 with the values of J for Ag0, Ag60, and Ag80 multiplied by 10 times.

Equations (1)

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

η s =π [ r/(d+2r) ] 2
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.