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

Metal-halide perovskite-based edge emitting lasers

Open Access Open Access

Abstract

Edge-emitting metal halide perovskites-based 1st order distributed feedback lasers are realized and studied for the first time. The properties and performances of these devices are compared in details to those of the well-studied 2nd order DFB perovskite lasers. As expected, 1st order lasers exhibit superior properties in terms of edge emission, rendering them highly attractive for applications such as photonic integrated circuits. In addition, it is found that, unexpectedly, the threshold levels of the 1st order lasers are higher than those of the 2nd order devices. We show that this phenomenon stems from the efficiency of the optical excitation which depends on the incidence angle and the gratings period.

© 2022 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Metal halide perovskite semiconductors have been receiving much interest during the last decade. Recently, perovskite based solar cells reached an efficiency level of 29% [1]. Perovskites have also found applications in optoelectronics including light emitting devices [2,3], photodetectors [4], and lasers [5,6]. These materials exhibit high absorption coefficients [7], direct band gap [8] and a sharp optical band edge. They can be deposited by several types of techniques some of these are, one-step deposition from a solution processable precursor solution [7], or a two-step process in which a metal halide is deposited and then converted to perovskite by exposure to a volatile organic compound [9]. These fabrication techniques which provide an opportunity to tune the band gap over the range 1.1- 3.3 eV simply by varying the anion (X = Cl-, Br-, I-), metal (Pb, Cs etc.) and cation (MA, FA etc.) composition and ratios [10]. This is undoubtedly an advantage over conventional semiconductor materials, which are limited by the lattice mismatch and strained layer epitaxy. Furthermore, using perovskite lasers [11], the so called “green-gap” [12] in the existing semiconductor gain media can be overcome. Another important advantage of perovskites over conventional semiconductor materials is the presence of predominant Van der Waals interactions among the halide atoms and hydrogen bonding, which render them “soft materials”. Consequently, perovskites can be directly patterned by thermal nanoimprint lithography (NIL), a technique which cannot be directly used with conventional inorganic semiconductor (AlAs, GaAs, GaN etc.). The use of NIL not only improves the film quality and reduces scattering but also offers a high throughput for the device fabrication [13].

In virtue of the high optical gain of perovskites, various configurations of lasers have been demonstrated. The resonators used for feedback include Fabry-Perot cavities [14], spherical resonators [15], random laser cavities [16], photonic crystal cavities [17], and more. Compared to all other structures, distributed feedback (DFB) cavities received much attention due to their lasing wavelength flexibility and narrow linewidth properties. Particularly, the wavelength can be tuned by varying several factors such as the gratings periodicity, the thickness and refractive index of the gain material, the surrounding refractive index etc. [18]. Such lasers also exhibit very low threshold levels, high-quality factors and mirror free design.

DFB lasers are realized in two main common configurations, utilizing 1st or 2nd order Bragg reflectors. 2nd order lasers (which are based on 2nd order Bragg grating) exhibit strong vertical emission (perpendicular to the surface) and are often used as surface emitting laser. In contrast, 1st order DFB lasers (which are based on 1st order Bragg grating) only emit along the surface plane, exhibiting edge emission properties, as well as superior quality factor [19]. In fact, only for a very narrow range of duty cycles, the 2nd order grating exhibit dominant edge emission properties. This is in contrast to 1st order grating structures which emit only broadside for a large range of duty cycles. Finally, 1st order lasers are more suitable for integration in optical devices such as waveguides, modulators, etc. Despite the abovementioned advantages, 1st order DFB edge emitting lasers based on perovskite have not been demonstrated yet. It should be noted that the Bragg mirrors used for perovskite based VCSELs are quarter-wavelength thick. Thus, such VCSELs can be considered as 1st order lasers. Nonetheless, the mirrors used in such devices where either dielectric [20] or consist of conventional semiconductor materials GaN/NP-GaN [21], thus requiring additional fabrication and processing steps. Furthermore, VCSELs are inherently surface emitting devices and unsuitable of edge emitting applications.

2. Results and discussion

In this paper, we design and fabricate, using nano-imprint lithography (NIL) 1st and 2nd order perovskite lasers and characterize them experimentally under pulsed optical pumping. The lasing properties of 1st (pitch of 180 nm) order and 2nd (pitch of 360 nm) order lasers are compared, showing that 1st order lasers exhibit mainly edge emission characteristics. The dependence of the properties of both 1st and 2nd order DFB lasers on the excitation angle is also studied in details. The nano-imprint lithography used to fabricate the lasers enables a facile and inexpensive way to fabricate DFB resonators, which is feasible to upscaling and mass production.

DFB lasers are formed by exploiting the high reflectivity of a periodic structure in the vicinity of their photonic bandgap edges. These cavities can be realized by precise nano-structuring of the gain medium or by deposition of that medium on a pre-corrugated surface [22]. If the corrugations are shallow (weak index perturbations) lasing occurs very close the wavelength satisfying the Bragg condition, ${{\lambda} _{Bragg}}$: $2{n_{eff}}\Lambda = m{{\lambda} _{Bragg}}$, where ${n_{eff}}$ is the effective refractive index of the waveguide optical mode, $\Lambda $ is the periodicity of the corrugated structure, and m is the diffraction order. For a waveguide structure consisting of a 200 nm perovskite core, SiO2 substrate and air clad, the Bragg condition for $m = 2$ and grating period of $\Lambda = 360\; \textrm{nm}$ leads to ${\lambda _{Bragg}} = 784.4\; nm$ [23]. Clearly, a first order ($m = 1$) grating with half that period (i.e., $\Lambda = 180\; \textrm{nm}$) yields the same Bragg wavelength.

The thickness of the perovskite layer must be chosen carefully. If the thickness is too large, the device might support many modes of the perovskite slab waveguide. Consequently, higher order modes have lower ${n_{eff}}$ compared to that of lower order modes and, hence, exhibit different lasing wavelengths. A multimode laser can result in either broadening of the lasing peak or lasing with multiple wavelengths. These additional lasing modes peaks will also compete over the optical gain and may lead to higher threshold and mode switching effects. On the other hand, a thin layer below the cutoff condition will not support any modes. Figure S1.2b (see supporting info.) depicts the dependence of the effective index of the two fundamental TE and TM mode on the waveguide thickness. This calculation is made for a perovskite slab waveguide on a SiO2 substrate and air clad. Our aim is to find a thickness, which supports a single guided mode in the vertical direction while sufficiently away from the cutoff region. From figure S1.1 (see supporting info) it is clear that the confinement factor for 100 nm thick waveguide is very poor and expect to yield low modal gain. On the other hand, 300 nm thick perovskite core supports a second, undesired, TE mode (TE1, see S1.1 with ${n_{eff}} = 1.7256$). A waveguide with a 200 nm thick core supports only the two fundamental TE0 and TM0 modes where the ${n_{eff}}$ of the TE0 mode is 2.173. Nevertheless, only the reflectivity bandwidth of the TE0 exhibits good overlap with the gain curve and expected to lase in practice. Thus, core thickness of 200 nm seems to be a good compromise between spatial confinement and single mode operation.

It should be noted that there is no lateral confinement in the structure and, in principle, many transverse modes can be supported. Nevertheless, the lasing spectrum exhibits a single peak and the linewidth of wide DFB lasers is identical to that of a laterally confined one [23]. We attribute this to the combination of mode suppression in the laser (due to the homogeneously broadened asymmetric gain profile) and the resolution of the spectrometer. Thus, it is possible that the laser emits several spatial modes of almost identical wavelength which are not resolved by the spectrometer, though the fact that laterally confined lasers exhibit similar linewidth does not support this possibility. While lateral confinement can be readily introduced (see e.g., Ref. [23]), it is unnecessary as the main objective of this work is to compare the impact of the grating order (1st or 2nd) on the laser properties.

We performed FDTD simulations to identify the optimal etch-depth of the grating in the gain (perovskite) medium. We found that 1st order grating shows a sharp reflectivity peak for etch depth thicker than 10 nm whereas 2nd order grating shows good reflectivity (>60%) for etch depth larger than 60 nm (Supporting info Sec. 2). Consequently, in our optimized design we use an etch depth in the range of 60–70 nm.

In order to compare the lasing characteristics of 1st and 2nd order DFB perovskite lasers, there are several design parameters that should be considered. In particular, varying the gratings duty-cycle can substantially modify the value and bandwidth of the reflectivity and hence, the lasing wavelength and threshold. Generally speaking, the coupling coefficient between the forwards and backwards propagating waves exhibit double hump curve dependence on the duty cycle for a 2nd order grating where the two local maxima are at ∼20% and at ∼70–80% (see Figure S3.1b, c) [24]. On the other hand, the coupling coefficient of the 1st order Bragg grating exhibits a single maximum (as a function of the duty-cycle – see Figure S3.1a, c). Referring to Figure S3.1, the 1st order laser shows a single humped curve with a maximum reflectivity at duty cycle of 60% and lasing at 790 nm whereas the peak reflectivity for the 2nd order devices is obtained at duty-cycle of ∼85%. Consequently, we choose duty cycle of 60% and 85% for the 1st and 2nd order lasers respectively.

In order to obtain lasing at 790 nm we choose to use gratings periodicity of 180 nm and 360 nm for the 1st and 2nd order devices respectively. The gratings are imprinted into the perovskite layer using a set of lithographic techniques (Fig. 1). First, E-beam lithography and reactive ion etching (RIE) are used for fabricating a master. Then, sol-gel lithography is utilized for fabricating a stamp from the master and finally, nano imprint lithography (NIL) is used for realizing the final device. The etch depth of the final DFB lasers are ∼60 nm And their lateral dimension are either 200×500 µm2 or 500×500 µm2 . The patterned length along the grating vector (Λ) is kept fixed at 500 µm. The number of periods for the 1st and 2nd order periods are 2777 and 1388 respectively, corresponding to the angular dispersion of 5.56 mrad/nm and 2.78 mrad/nm respectively. The duty cycles (DC) are 59% and 85% for 1st order and 2nd order lasers respectively. The complete details of the fabrication steps are provided in the Supporting info.

 figure: Fig. 1.

Fig. 1. Fabrication process flow and characterization of perovskite DFB lasers.

Download Full Size | PDF

To account for potential fabrication errors and tolerances we realized four perovskite based 1st order DFB laser designs having dimension of 200 µm × 500 µm and periodicities of 170 nm, 175 nm, 180 nm, and 185 nm. To ensure compatibility with our measurement set-up we cut the chip with a scriber along the grating period (Fig. 1) and measured the emission characteristics in two configurations. First, excitation and emission collection from the top. Second, excitation from the top and emission collection from the edge of the device (Fig. 1). The DFB lasers with gratings periods of 170 and 185 nm did not lase. In contrast, the devices with gratings periods of 175 nm and 180 nm do show lasing under pulsed optical pumping (5 ns pulses, 20 Hz rep-rate, λp = 532 nm). The spectrum of the DFB laser with periodicity $\Lambda = 175{\; }nm$ exhibits two competing peaks at 780.19 nm and at 794.15 nm (Figure S4a, b) and is unsuitable for single mode lasing operation.

Figure 2(a) shows the evolution of lasing in a 1st order laser ($\Lambda $ = 180 nm, top excitation, side collection geometry). The excitation scheme included a 20X objective with an additional lens (before the objective) in order to increase the spot size, leading to a spot size of ∼100 µm in radius. Figure 2(b) shows the FWHM as a function of excitation flux corresponding to Fig. 2(a). The threshold is found to be 300 µJ/cm2 and FWHM at threshold is 1.2 nm. Figures 2(c) and 2(d) depicts similar plots for the top excitation - top collection geometry. The threshold in this case is found to be around 539 µJ/cm2 and the FWHM at threshold is also 1.2 nm. The lasing wavelength of the 1st order DFB laser is 794 nm, very close to the desired wavelength of 790 nm.

 figure: Fig. 2.

Fig. 2. (a) Emission from a 1st order DFB ($\varLambda = 180{\; }nm)$ when collected from side with an optical fiber of comparable mode field diameter. b) Peak intensity/FWHM vs excitation fluences for a side collection set-up. c) Same as (a) but when collected from top with a 20X objective. d) Same as (b) when collected from top.

Download Full Size | PDF

We note, that although the fabricated devices are standard DFB lasers (i.e., without a λ/4 phase-shift), the observed lasing spectrum includes only a single peak. We attribute this to the strong asymmetry of the material gain profile which highly overlaps with its the absorption profile. This complex gain/absorption spectra results in an asymmetric net gain profile which favors one of the solutions and suppresses the other.

In order to compare the threshold characteristics between 1st and 2nd order lasers, the 1st order and 2nd order DFB lasers are fabricated on the same chip with the same process to minimize variations that may arise from parameters tolerances (e.g., thickness, etch depth, defect density etc.).

Figure 3 shows experimentally obtained collected output power and FWHM as a function of the pump level for 1st and 2nd order lasers. All measurements were obtained from devices fabricated simultaneously on the same chip (note that the 1st order lasers of Fig. 3(a) and Fig. 2 are of the same design but from different batches). The lasers were excited from the top with 20X (NA 0.42) objective. When the emission is collected from the edge of the chip, the emission collected from the edge of the 1st order lasers is substantially larger than that collected from the top (approximately by a factor of 10). This is expected as the 1st order gratings do not scatter light efficiently in the vertical direction. It should be noted that the side collection is less efficient than the top collection (the top collection utilizes a 20X objective with NA of 0.42 whereas the side collection utilizes an optical fiber with NA of 0.22), which means that the difference is even larger. Consequently, the difference between the intensity emitted broadside and that emitted vertically is considerably larger and the 1st order laser is indeed mainly edge-emitting. In contrast, the emission collected from the edge of the 2nd order lasers is lower than that collected from the top (by a factor of 5). However, as the difference is not as dramatic as in the case of the 1st order DFB lasers and considering the higher efficiency of the top collection, it seems that the edge and top emissions of the 2nd order DFB lasers are of similar order of magnitude. To evaluate the threshold levels, we consider the measurements with the best SNR, i.e., side collection for the 1st order DFB lasers and top collection for the 2nd order devices. The thresholds are found to be 239 µJ/cm2 and 310 µJ/cm2 for the 2nd and 1st order devices respectively. Thus, in contrast to what one may expect, the threshold level of the 2nd order lasers is lower than that of the 1st ones.

 figure: Fig. 3.

Fig. 3. (a) Emission intensity maxima versus excitation energy (532 nm, 5 ns. 20 Hz) gray stars and black spheres (left axis), and FWHM (red stars and red spheres) versus pump energy for a 1st order (stars) and spheres when excited from (top) and collected from side. (b) Similar to (a) for 2nd order lasers.

Download Full Size | PDF

The fact that the threshold levels of the 2nd order DFB lasers is lower than that of 1st order ones is rather surprising as the later are expected to exhibit higher quality factor or at least similar one to that of the former. A possible explanation for that discrepancy is that the gain in the 2nd order DFB lasers is larger than that of the 1st order devices. Such difference could stem from differences in the coupling efficiency of the pump to the guided mode of the slab due to the gratings. In particular, if the 2nd order grating couple the pump into the waveguide mode more efficiently than the 1st order grating then the threshold of the former would be lower.

Figure 4(a) depicts FDTD-based calculations of the coupling efficiency of the pump into the waveguide mode as a function of the incidence angle for the 1st and the 2nd order devices. We note that the coupling efficiency of the pump into the 2nd order devices is substantially larger than that of the 1st order ones for incidence angle smaller than ∼12°. For incidence angle smaller than 5°, the difference between the coupling efficiency is of several orders of magnitudes. These calculations support the hypothesis that the lower threshold levels of the 2nd order DFB laser originate from better coupling into the waveguide mode. Intuitively, it does not seem to be surprising that the 2nd order gratings provide better input efficiency than that of 1st order ones. 2nd order gratings exhibit efficient vertical emission and thus, due to reciprocity, should facilitate efficient coupling into the waveguide mode. However, as the grating periodicity is designed according to the wavelength of emission (and not of the pump), it is not that obvious that the gratings will exhibit efficient in-coupling at the pump wavelength. Furthermore, one may intuitively expect to see enhanced in-coupling efficiency at discrete angles (corresponding to the Bragg order), which is not the case.

 figure: Fig. 4.

Fig. 4. a) FDTD simulation showing the coupled power in the waveguide (for λ = 532 nm) as a function of incidence angle. b) Emission intensity maxima versus excitation energy (532 nm, 5 ns. 20 Hz), white circles and gray stars (left axis), and FWHM (red stars and red circles) versus pump energy for a 1st order (stars) and 2nd order (circles) when excited from top and collected from top with a 5X, 0.21 NA objective. (c) Similar to (b) but when excited and collected from top with a 20X, 0.42 NA objective.

Download Full Size | PDF

In order to further verify that the pump coupling efficiency is indeed the reason for the threshold levels properties, we take advantage of the large difference in the coupling efficiencies at small angles. When the DFB lasers are excited with a 20X objective lens, the pump beam consists of components with incidence angle of up to ±25°. Thus, significant part of the pump beam can couple efficiently to both type of devices. However, if the devices are excited with an objective with lower NA, then the pump coupling efficiency into the 2nd order devices is expected to be substantially larger and their lasing threshold should be much lower than that of the 1st order devices. To verify that, we replaced the 20X objective lens with a 5X one (NA ∼0.21). This lens supports excitation angles of up to ±12°. Figure 4(b) compares the corresponding Pin-Pout curves of the two types of DFB lasers. For the experiment, we chose to use 500 × 500 µm2 devices, with the optimal grating configuration for each laser type (i.e., duty cycle of 60% and 85% for the 1st order and 2nd order devices respectively). The threshold of the 2nd order devices has reduced to 92.45 µJ/cm2 while the 1st order ones do not exhibit lasing even at pumping levels as high as 350 µJ/cm2. The reduction of threshold for the 2nd order is attributed to a larger pumping spot size which increases the reflectivity of the grating.

The observed threshold levels with the 5X objective lens pumping scheme are in full agreement with the pump beam coupling efficiency hypothesis. The lower NA of the lens enhances the pump beam coupling efficiency into the waveguide mode of the 2nd order devices, thus yielding a lower threshold that observe with the 20X lens. In conjunction, the use of the 5X lens reduces dramatically the pump beam coupling efficiency into the waveguide mode of the 1st order devices, leading to complete extinguishing of lasing action.

3. Conclusion

We have designed, fabricated, and characterized 1st and 2nd order DFB lasers based on nanoimprint lithography of MAPbI3. To our knowledge, this is the first demonstration of perovskites-based 1st order devices. The 1st order lasers exhibit strong edge emitting properties while the 2nd order one’s exhibit both edge and surface emission. We have studied the lasing thresholds and output power characteristics of the two types of lasers and found that, in contrast to what can be expected, the 2nd order DFB lasers exhibit lower threshold levels under optical pumping. We showed that this phenomenon is attributed to better coupling of the pump beam into the laser waveguide mode, which yields substantially larger modal gain for the 2nd order devices. It should be noted, however, that the lower lasing threshold of the 2nd order DFB lasers originates from the use of optical pumping and that under electrical pumping conditions the 1st order devices are expected to exhibit lower threshold levels than those of 2nd order devices. The edge emitting properties of 1st order DFB lasers render them of great potential for integrated optics in conjunction with waveguide configuration. Furthermore, the ability to tune continuously the emission from 1.1 to 3.1 eV render such devices highly valuable for telecommunication, medicine, and numerous other applications.

Funding

Israel Ministry of Science and Technology.

Acknowledgments

This research was partially supported by the Israel Ministry of Science and technology. S.B. and O.B.-O. acknowledges the support from the Israeli Ministry of Science for the Postdoc Fellowship.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. Best Research-Cell Efficiency Chart, https://www.nrel.gov/pv/cell-efficiency.html (accessed: Sep 2020).

2. S. Adjokatse, H.-H. Fang, and M. A. Loi, “Broadly tunable metal halide perovskites for solid-state light-emission applications,” Mater. Today 20(8), 413–424 (2017). [CrossRef]  

3. Z. K. Tan, R. S. Moghaddam, M. L. Lai, P. Docampo, R. Higler, F. Deschler, M. Price, A. Sadhanala, L. M. Pazos, D. Credgangton, F. Hanusch, T. Bein, H. J. Snaith, and R. H. Friend, “Bright light-emitting diodes based on organometal halide perovskite,” Nat. Nanotechnol. 9(9), 687–692 (2014). [CrossRef]  

4. L. Dou, Y. Yang, J. You, Z. Hong, W. H. Chang, and Y. Yang, “Solution-processed hybrid perovskite photodetectors with high detectivity,” Nat. Commun. 5(1), 5404 (2014). [CrossRef]  

5. B. R. Sutherland and E. H. Sargent, “Perovskite photonic sources,” Nat. Photonics 10(5), 295–302 (2016). [CrossRef]  

6. G. Xing, N. Mathews, S. S. Lim, N. Yantara, X. Liu, D. Sabba, M. Gratzel, S. Mhaisalkar, and T. C. Sum, “Low-temperature solution-processed wavelength-tunable perovskites for lasing,” Nat. Mater. 13(5), 476–480 (2014). [CrossRef]  

7. M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, and H. J. Snaith, “Efficient hybrid solar cells based on meso-superstructured organometal halide perovskites,” Science 338(6107), 643–647 (2012). [CrossRef]  

8. J. Even, L. Pedesseau, and C. Katan, “Analysis of multivalley and multibandgap absorption and enhancement of free carriers related to exciton screening in hybrid perovskites,” J. Phys. Chem. C 118(22), 11566–11572 (2014). [CrossRef]  

9. J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin, and M. Gratzel, “Sequential deposition as a route to high-performance perovskite-sensitized solar cells,” Nature 499(7458), 316–319 (2013). [CrossRef]  

10. J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal, and S. I. Seok, “Chemical management for colorful, efficient, and stable inorganic− organic hybrid nanostructured solar cells,” Nano Lett. 13(4), 1764–1769 (2013). [CrossRef]  

11. J. R. Harwell, G. L. Whitworth, G. A. Turnbull, and I. D. W. Samuel, “Green perovskite distributed feedback lasers,” Sci. Rep. 7(1), 11727 (2017). [CrossRef]  

12. T. Langer, A. Kruse, F. A. Ketzer, A. Schwiegel, L. Hoffmann, H. Jönen, H. Bremers, U. Rossow, and A. Hangleiter, “Origin of the “green-gap”: Increasing non-radiative recombination in indium-rich GaInN/GaN quantum well structures,” Phys. Status Solidi C 8(7-8), 2170–2172 (2011). [CrossRef]  

13. P. Brenner, O. Bar-On, M. Jakoby, I. Allegro, B. S. Richards, U. W. Paetzold, I. A. Howard, J. Scheuer, and U. Lemmer, “Continuous wave amplified spontaneous emission in phase-stable lead halide perovskites,” Nat. Commun. 10(1), 988 (2019). [CrossRef]  

14. F. Deschler, M. Price, S. Pathak, L. E. Klintberg, D.-D. Jarausch, R. Higler, S. Hüttner, T. Leijtens, S. D. Stranks, H. J. Snaith, M. Atatüre, R. T. Phillips, and R. H. Friend, “High photoluminescence efficiency and optically pumped lasing in solution-processed mixed halide perovskite semiconductors,” J. Phys. Chem. Lett. 5(8), 1421–1426 (2014). [CrossRef]  

15. B. R. Sutherland, S. Hoogland, M. M. Adachi, C. T. O. Wong, and E. H. Sargent, “Conformal organohalide perovskites enable lasing on spherical resonators,” ACS Nano 8(10), 10947–10952 (2014). [CrossRef]  

16. R. Dhanker, A. N. Brigeman, A. V. Larsen, R. J. Stewart, J. B. Asbury, and N. C. Giebink, “Random lasing in organo-lead halide perovskite microcrystal networks,” Appl. Phys. Lett. 105(15), 151112 (2014). [CrossRef]  

17. S. Chen, K. Roh, J. Lee, W. K. Chong, Y. Lu, N. Mathews, T. C. Sum, and A. Nurmikko, “A photonic crystal laser from solution based organo-lead Iodide perovskite thin films,” ACS Nano 10(4), 3959–3967 (2016). [CrossRef]  

18. V. Bonal, J. A. Quintana, J. M. Villalvilla, and M. A. Díaz-García, “Controlling the emission properties of solution-processed organic distributed feedback lasers through resonator design,” Sci. Rep. 9(1), 11159 (2019). [CrossRef]  

19. L. A. Coldren, S. W. Corzine, and M. L. Mašanović, Diode Lasers and Photonic Integrated Circuits. Series in Microwave and Optical Engineering (John Wiley & Sons, Inc., 2012).

20. N. Pourdavoud, T. Haeger, A. Mayer, P. J. Cegielski, A. L. Giesecke, R. Heiderhoff, S. Olthof, S. Zaefferer, I. Shutsko, A. Henkel, D. Becker-Koch, M. Stein, M. Cehovski, O. Charfi, H.-H. Johannes, D. Rogalla, M. C. Lemme, M. Koch, Y. Vaynzof, K. Meerholz, W. Kowalsky, H.-C. Scheer, P. Görrn, and T. Riedl, “Room-temperature stimulated emission and lasing in recrystallized cesium lead bromide perovskite thin films,” Adv. Mater. 31(39), 1903717 (2019). [CrossRef]  

21. S. Chen, C. Zhang, J. Lee, J. Han, and A. Nurmikko, “High-q, low-threshold monolithic perovskite thin-film vertical-cavity lasers,” Adv. Mater. 29(16), 1604781 (2017). [CrossRef]  

22. M. Saliba, S. M. Wood, J. B. Patel, P. K. Nayak, J. Huang, J. A. Alexander-Webber, B. Wenger, S. D. Stranks, M. T. Hörantner, J. T.-W. Wang, R. J. Nicholas, L. M. Herz, M. B. Johnston, S. M. Morris, H. J. Snaith, and M. K. Riede, “Structured organic–inorganic perovskite toward a distributed feedback laser,” Adv. Mater. 28(5), 923–929 (2016). [CrossRef]  

23. O. Bar-On, P. Brenner, U. Lemmer, and J. Scheuer, “Micro lasers by scalable lithography of metal-halide perovskites,” Adv. Mater. Technol. 3(12), 1800212 (2018). [CrossRef]  

24. W. Streifer, D. R. Scifres, and R. D. Bumham, “Coupling coefficients for distributed feedback single- and double-heterostructure diode lasers,” IEEE J. Quantum Electron. 11(11), 867–873 (1975). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental 1

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (4)

Fig. 1.
Fig. 1. Fabrication process flow and characterization of perovskite DFB lasers.
Fig. 2.
Fig. 2. (a) Emission from a 1st order DFB ($\varLambda = 180{\; }nm)$ when collected from side with an optical fiber of comparable mode field diameter. b) Peak intensity/FWHM vs excitation fluences for a side collection set-up. c) Same as (a) but when collected from top with a 20X objective. d) Same as (b) when collected from top.
Fig. 3.
Fig. 3. (a) Emission intensity maxima versus excitation energy (532 nm, 5 ns. 20 Hz) gray stars and black spheres (left axis), and FWHM (red stars and red spheres) versus pump energy for a 1st order (stars) and spheres when excited from (top) and collected from side. (b) Similar to (a) for 2nd order lasers.
Fig. 4.
Fig. 4. a) FDTD simulation showing the coupled power in the waveguide (for λ = 532 nm) as a function of incidence angle. b) Emission intensity maxima versus excitation energy (532 nm, 5 ns. 20 Hz), white circles and gray stars (left axis), and FWHM (red stars and red circles) versus pump energy for a 1st order (stars) and 2nd order (circles) when excited from top and collected from top with a 5X, 0.21 NA objective. (c) Similar to (b) but when excited and collected from top with a 20X, 0.42 NA objective.
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.