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

Virtual cavity in distributed Bragg reflectors

Open Access Open Access

Abstract

We show theoretically and experimentally that distributed Bragg reflector (DBR) supports a surface electromagnetic wave exhibiting evanescent decay in the air and oscillatory decay in the DBR. The wave exists in TM polarization only. The field extension in the air may reach several wavelengths of light. Once gain medium is introduced into the DBR a novel class of diode lasers, semiconductor optical amplifiers, light-emitting diodes, etc. can be developed allowing a new type of in-plane or near-field light outcoupling. To improve the wavelength stability of the laser diode, a resonant cavity structure can be coupled to the DBR, allowing a coupled state of the cavity mode and the near-field mode. A GaAlAs-based epitaxial structure of a vertical-cavity surface-emitting laser (VCSEL) having an antiwaveguiding cavity and multiple GaInAs quantum wells as an active region was grown and processed as an in–plane Fabry-Pérot resonator with cleaved facets. Windows in the top stripe contact were made to facilitate monitoring of the optical modes. Three types of the optical modes were observed in electroluminescence (EL) studies under high current densities > 1 kA/cm2. Mode A with the longest wavelength is a VCSEL–like mode emitting normal to the surface. Mode B has a shorter wavelength, emitting light at two symmetric lobes tilted with respect to the normal to the surface in the direction parallel to the stripe. Mode C has the shortest wavelength and shifts with a temperature at a rate 0.06 nm/K. Polarization studies reveal predominantly TE emission for modes A and B and purely TM for mode C in agreement with the theory. Spectral position, thermal shift and polarization of mode C confirm it to be a coupled state of the cavity mode and near-field DBR surface-trapped mode.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Various types of surface or interface electromagnetic waves propagating along a boundary between two media and decaying in both directions away from the boundary attract a strong interest, both fundamental and practical. In most cases very special criteria are required to enable such waves. Thus an interface wave on a boundary between two homogeneous media is possible only if the dielectric permittivity ε of one media is negative [1], which can be provided by the plasma contribution to ε [2], e. g. in metals. Photonic crystal [3] may allow propagation of resonant surface waves on its boundary [4]. Structuring of a surface using gratings or ridges can also support surface waves [5]. Other examples include interface waves in chiral materials [6], on a boundary of a gyrotropic material and a negative phase velocity medium [7] or at a non–linear interface [8]. An optically anisotropic medium with optical axes tilted with respect to the interface may support interface waves in a narrow interval of in–plane angles [9].

Arising new applications of semiconductor diode lasers require devices with substantially advanced performance over presently available ones. Thus recently proposed design concepts [10–35] have led to devices with novel functionality. In this regard, there arises a question of existence of surface waves beyond complex systems of [1–9], i. e., waves bounded to a surface of conventional semiconductor materials, and of a possibility to employ such waves in practical devices.

A surface of a DBR is likely to be the simplest structure of this type. In our earlier work [36] we studied DBR–containing devices, namely those based on VCSEL epiwafers, capable to high performance operation once fabricated as oxide–confined VCSELs. The structures in [36] were grown in line with the antiwaveguiding VCSEL (A–VCSEL) design [37, 38] where the in–plane waveguide mode is prohibited due to the reduced overlap with the active medium. Electroluminescence (EL) studies revealed three different types of modes. The longest wavelength mode (mode A) is a VCSEL–like mode at ~854 nm emitting normal to the surface with a full width at half maximum (FWHM) of the far field ~10°. Accordingly the lasing wavelength demonstrates a thermal shift at a rate 0.07 nm/K. Mode B was at a shorter wavelength of ~840 nm at room temperature, emitting light in two symmetric lobes tilted with respect to the normal to the surface in the directions parallel to the stripe. The emission wavelength of this mode shifts at a rate 0.22 nm/K according to the GaAs bandgap shift. The tilt angle of mode B with respect to the normal reduces as the wavelength approaches the vertical cavity etalon wavelength and this mode finally merges with the VCSEL mode. Mode B hops between different lateral modes of the VCSEL forming a dense spectrum due to significant longitudinal cavity length. Mode C was at ~820 nm, shifting upon temperature at the rate 0.06 nm/K. The wavelength of mode C matches the wavelength of the tilted mode propagating in the semiconductor structure at the critical angle for total internal reflection for light impinging from the semiconductor chip on the semiconductor/air interface. This suggests that mode C could be related to a specific surface mode. Indeed, modeling performed in Ref [36] suggested that the lasing mode C represents a coupled state between the TM–polarized surface–trapped optical mode and the VCSEL cavity mode. However, polarization of the light emission was not measured.

In the present paper we extend the studies of Ref [36]. We address theoretically a fundamental phenomenon of the surface electromagnetic waves in a generic structure of a DBR terminated by a free surface, without mixing with a resonant cavity. Further, we investigate experimentally the electroluminescence and lasing properties of a device formed of a similar epitaxial structure as in Ref [36] but with a smaller detuning between the gain spectrum and the cavity resonance. The second structure reveals a similar behavior emitting in the same three types of the modes. We measure the polarization of all the modes and confirm the theoretical predictions.

2. Electromagnetic waves on a surface of a DBR

Let a DBR be formed of 25 pairs of layers alternating in the z direction, namely layers of a high (n1) and a low (n2) refractive indices with thickness of one quarter of the material wavelength, i. e. with d1 = λ0/(4n1) and d2 = λ0/(4n2), where λ0 is the wavelength of the peak reflectivity. The structure is formed on a substrate with the refractive index n1 and is terminated by air from the top at the surface z = z0. The topmost layer is the high index layer, which is typically used in VCSELs to provide in–phase reflection at the semiconductor/air interface and maximize the photon lifetime. To specify, we choose n1 = 3.5 and n2 = 3.0 which mimics typical refractive indices of Ga1–xAlxAs alloy with low and high Al composition, respectively.

Consider a TM wave propagating along the x–direction and having only one component of the magnetic field in the y–direction,

H(t,r)=[Hx(t,r),Hy(t,r),Hz(t,r)]=[0,Hy(z)exp(iωt)exp(ineffk0x),0],
where k0 = 2π/λ is the wavenumber of the light, and neff is the effective refractive index to be found as the eigenvalue of the wave equation:
ε(z)ddz[1ε(z)ddzHy(z)]ε(z)k02Hy(z)=neff2k02Hy(z).
Here ε(z) = n2(z) is the dielectric permittivity. Targeted surface modes should have neff >1, whereas the mode is evanescent decaying exponentially Hy(z)~exp[βk0(zz0)] in the air away from the surface where the decay is characterized by the mode confinement coefficient β=neff21. If neff <1, the mode is radiative.

The second criterion for the existence of the surface mode reads that the mode should decay within the DBR, i. e. the mode wavelength should be within the reflectivity stopband of the DBR corresponding to the effective mode index neff. The effective mode angle defined for the DBR material with a high refractive index n1 = 3.5 equals ϑeff=sin1(neff/n1). Since we seek surface modes with neff close to 1, it is of particular interest to consider the DBR properties for the light impinging from material #1 at the tilt angle ϑeff=sin1(1/3.5)16.60, which corresponds to the onset of the total internal reflectance at the surface of material #1. Figure 1(a) shows the optical power reflectance (OR) spectrum at normal incidence, equal for both TE and TM light. Figures 1(b) and 1(c) display OR spectra for the oblique incidence at the angle 16.6° for TE and TM light, respectfully. Both spectra reveal stopband shifted to shorter wavelengths with respect to λ0, the spectrum for TE mode having a broader stopband, and the spectrum for TM mode having a more narrow stopband that the one for the normal incidence, as noticed also in [39]. A drastic difference between TE– and TM–polarized light appears once the OR spectra for gliding light impinging from the air at the angle 89.9° are considered. Once the OR spectrum for TE light (Fig. 1(d)) shows no extra features within the stopband, the TM light reveals a clear reflectivity dip (Fig. 1(e)) indicating presence of an effective virtual cavity at the surface capable to localize an optical mode.

 figure: Fig. 1

Fig. 1 Optical power reflectance (OR) spectra of the model DBR. (a)–(c) Modeled spectra for light impinging from material #1 onto the DBR. (d), (e) Modeled spectra for light impinging from the air at ϑ = 89.9°.

Download Full Size | PDF

Figures 2(a) and 2(b) display modeled OR spectra for light impinging from the air at different glancing angles of incidence. One can see that the dip position pointing at the transparency wavelength, in the interval of glancing angles from 86° to 90°, where the dip is pronounced, is not a function of the angle allowing angle–insensitive filters. The dip narrows once the glancing angle approaches 90 degrees, which is in line with the general trend of increasing the DBR reflectivity at glancing angles pointed out in [39].

 figure: Fig. 2

Fig. 2 (a, b) Modeled optical power reflectance spectra for light impinging from the air at different glancing angles of incidence. One can see that the within the interval of angles from 86° to 90° the spectral position of the dip, or the transparency wavelength is not a function of the glancing angle allowing angle–insensitive filters.

Download Full Size | PDF

Results of the solution of Eq. (2) for TM polarized modes are presented in Fig. 3, wherein the generic case valid for an arbitrary wavelength λ0 is presented for the particular value of λ0 = 850 nm. At the wavelength 0.96λ0 the solution yields multiple modes, in which the mode decaying in the DBR is coupled to some of the modes of the air (not shown). At λ = 0.95λ0 the mode becomes localized but decays very slowly in the air and is extended over a few tens of micrometers. As this wavelength is close to the center of the DBR reflectivity stopband in Figs. 1(c) and 1(e), the mode decays rapidly in the DBR. Upon decrease of the mode wavelength, the extension length of the surface mode in the air decreases. Further, as the mode wavelength shits away from the center of the stopband to its short wavelength boundary (in Figs. 1(c) and 1(e)), it decays weaker in the DBR. The mode corresponding to λ = 0.91λ0 is close to the stopband edge and the exponential decay of the field oscillation amplitude in the DBR nearly vanishes. The wavelength λ = 0.90λ0 is beyond the DBR stopband, and the mode is no longer a surface mode.

 figure: Fig. 3

Fig. 3 Refractive index profile of the model DBR and magnetic field strength profile in surface TM modes at different wavelengths.

Download Full Size | PDF

A large extension of the TM mode in the air implies that the effective refractive index of the mode is close to 1, and the effective angle of the mode propagation in the material layer is close to the onset of the total internal reflection at the semiconductor/air interface. Upon wavelength decrease as in Fig. 3, the extension of the mode in the air also decreases, and both neff and ϑeff=sin1(neff/n1) increase. As the localized mode does not propagate in the air, it does not manifest itself in the optical reflectance spectra of the structure for the light impinging from the air (as in Figs. 1(e), 2(a) and 2(b)). Some features of these modes in part of the mode decay within the DBR can be addressed, once the transmittance or reflectance of the DBR for the light impinging from semiconductor material (e. g., material #1, like in Fig. 1(c)) is considered. As ϑeff increases, the DBR reflectivity stopband shifts towards shorter wavelengths, but slower than the mode wavelength, and at 0.91λ0 the mode wavelength comes out of the stopband. Thus, the spectral region of the existing TM surface modes is bound from the long wavelength side, once the mode no longer decays in the air and is a radiative mode, and is bound from the short wavelength side once the wavelength comes out of the DBR reflectivity stopband.

Once a near–field device, i. e., a device with a certain mode extension in the air is targeted, the important features of the mode are the mode extension length in the air lext, the mode intensity in the active region, and the mode leakage losses to the substrate. Figure 3 suggests that, depending on a particularly sought application, an optimum mode wavelength can be found within the spectral range, in which the surface mode exists.

The spectral range, in which the surface TM mode exists, can be tuned, e. g., by varying the thickness of the topmost layer of the DBR. Such a change can be implemented by epitaxial growth of a thicker top layer or by partial etching of the grown wafer, which is also known for VCSELs and can be applied to tune the photon lifetime [40]. The dispersion curves in Fig. 4 display the mode confinement coefficient of the surface mode which is inverse proportional to lext, β=λ/(4πlext). Plots are presented for the nominal thickness of the topmost layer λ0/(4n1) as well as for a smaller and a larger thickness (5/6 and 7/6 of the nominal thickness, respectively). The surface mode exists in a broader spectral range for a thicker topmost layer.

 figure: Fig. 4

Fig. 4 Mode confinement coefficient of the surface TM mode which is inverse proportional to the mode extension length in the air, β=λ/(4πlext), versus mode wavelength for different thickness of the topmost layer.

Download Full Size | PDF

It should be noted that a TE surface mode does not exist for the structures considered. A near–field TE mode could be allowed only if an artificial cavity is formed on top of the DBR. The detailed analysis of the properties of near–field modes on the boundary of a DBR will be presented and possible applications will be discussed elsewhere [41].

3. Structure

A VCSEL–type GaAlAs–based structure was grown in a multiwafer industrial reactor by metalorganic vapor phase epitaxy. The structure contained a resonant λ/2 cavity with five compressively strained GaInAs quantum wells as active region, the cavity being sandwiched between a 34–periods GaAlAs–based n–doped bottom DBR and a 21.5–periods p–doped top DBR. The structure was cleaved in rectangular–shaped ~350 × 400 µm pieces with perpendicular facets. The contact grid region has a total width of w~70 µm. d = 7 µm–wide metal stripes serving as non–alloyed metal contact form periodic rectangular openings of a size b × a = 10 × 40 µm. The chosen dimensions of the contact grid allow a good electric contact and efficient in–plane current spreading from the contact region to ensure laterally–uniform current injection into the active region, on the one hand, and a sufficient open area, from which light can be emitted, on the other hand. Copper wire was bonded directly on top of the contact grid. Surface emission through the windows on top of the chip was measured at temperatures from 90 to 380 K. The schematics of the chip is presented in Fig. 5(a) and the image is shown in Fig. 5(b). Unlike typical VCSELs, where the current path is contracted by oxide–confined aperture or by proton bombardment area, no such electric confinement occurs in the structures studied. Current flowing from the top contact grid through the ~2.5 µm–thick top DBR is spread across the entire active region, and the resistance of the device is defined by the chip area and not by the grid area. In this connection it is worth noting that the emission occurs only in the region of the contact grid presumably due to in–plane localization of the optical modes by the grid.

 figure: Fig. 5

Fig. 5 (a) Perspective view of a chip with a metal grid contact on top surface. The chip is a rectangular–shaped ~350 × 400 µm piece with perpendicular facets. The contact grid region has a total width of w~70 µm. d = 7 µm–wide metal stripes serving as non–alloyed metal contact forms periodic rectangular openings having a size of b × a = 10 × 40 µm. Numbers 1, 2, 3 indicate various positions of the photodetector measuring electroluminescence. (b) Infrared image of the lasing device. Gray structure in the center is the copper wire bonded on top of the grid.

Download Full Size | PDF

4. Electroluminescence studies

Figure 5(a) displays several positions of the photodetector used to measure electroluminescence (EL) spectra from the device from different observation points to address various optical modes. In Ref [36]. EL spectra of the chip recorded from the chip facet (by detector #1 in Fig. 5(a)) in a broad range of temperatures are displayed. The spectra were measured by using a lens collecting light from an angle of ~20°. Typically the spectra contain three different peaks (see, e. g., Fig. 6).

 figure: Fig. 6

Fig. 6 Electroluminescence spectrum measured from the edge of the device in the stripe direction.

Download Full Size | PDF

To elucidate the physical features of each of the modes we consider the temperature behavior of each of the three peaks (Fig. 7(a)) as well as the far field profile of the emission (Figs. 8(a) and 8(b)). The longest wavelength mode (mode A) is a VCSEL–like mode at ~854 nm emitting predominantly normal to the surface (Figs. 8(a) and 8(b)). This mode exhibits a thermal shift of the wavelength of ~0.07 nm/K.

 figure: Fig. 7

Fig. 7 (a) Shift of three electroluminescence (EL) peaks versus temperature. (b) Evolution of the EL spectrum versus temperatures while passing the resonance between mode B and mode C.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 (a) Far field profile of the electroluminescence (EL). (b) Spectrally–resolved far field profile of EL.

Download Full Size | PDF

Mode B is at shorter wavelengths of ~840 nm at room temperature, emitting light at two symmetric lobes tilted with respect to the normal to the surface in the directions parallel to the stripe (Fig. 8(a)). The mode lies within the continuum spectrum of tilted modes, having the vertical profile of the optical field same as the VCSEL mode and propagating forth and back in the rectangular chip acting as an in–plane Fabry–Pérot resonator. Such modes are also known to represent slow light in VCSEL amplifiers (see, e. g., [42] are references therein). The emission wavelength of mode B shifts at a rate 0.22 nm/K according to GaAs bandgap shift. Upon temperature increase the angle of the mode with respect to the normal reduces (Fig. 8(a)) as the wavelength approaches the vertical cavity etalon wavelength and mode B finally merges with the VCSEL mode A.

It is important to note a strong resonance effect once the wavelengths of mode B and mode C match (Fig. 7(b)). At this temperature the intensity of the related feature increases more than by an order of magnitude with respect to that at non–resonant wavelengths. It was necessary to reduce the drive current 6–fold to reduce the intensity of the emitted light to avoid saturation of the detector. The resonant enhancement of the light emission occurs once the gain spectrum matches the wavelength of mode C and indicates a possibility to channel all the power to the surface–trapped mode, achieve light amplification and lasing in the surface–trapped TM– polarized near–field mode.

Far–field emission profiles were measured with the spectral resolution by scanning an angle ϑx at which the optical fiber is directed towards the sample in the (xz)–plane as noted in Fig. 5(a). The light from the exit of the fiber was directed to a monochromator. The spectrally–resolved far–field profiles (Fig. 8(b)) show that the angle for each particular mode is fixed, the far field of each mode is narrow, and the broadening of the far field in Fig. 8(a) is related to the spectral broadening of the tilted emission. The far field measured at 852 nm reveals the full width at half maximum of the VCSEL mode ~10°.

As the theoretical modeling demonstrates that the mode C combining the VCSEL cavity mode and the surface–trapped mode [36], exists only in TM polarization, it is important to measure polarization–resolved EL spectra. Figure 9 displays polarization–resolved spectra measured from position 1 in Fig. 5(a). In the long–wavelength part of the spectrum, close to 854 nm related to mode A and close to 840 nm related to mode B, TE–polarized emission dominates, as should be expected for a device based on III–V semiconductor materials. Indeed, as pointed out in [43], because of microscopic selection rules associated with the unit cell wavefunctions in III–Vs, for the TM polarization, the conduction–to–heavy–hole transitions are forbidden, and only conduction–to–light–hole transitions contribute to the emission of light. Since there are more heavy holes than light holes in thermal equilibrium due to a higher density of states, the gain associated with heavy holes is larger, and hence the gain is larger for the TE polarization. The feature persists in compressively strained quantum wells, where the splitting of the degenerated valence band pushes the light hole band towards higher energies of holes keeping the valence band more populated with heavy holes. For this reason, the stimulated emission in modes A and B is predominantly TE–polarized. On the contrary, since mode C exists in TM polarization only, the emission at ~821 nm is completely TM–polarized, in full agreement with the modeling.

 figure: Fig. 9

Fig. 9 Polarization–resolved electroluminescence (EL) spectra of the device.

Download Full Size | PDF

For a further comparison with the modeling we note that the topmost layer of the structure has an extended thickness of (7/6)λ0/(4n1). According to the corresponding dispersion curve of Fig. 4, the spectral range of existing surface modes is extended up to 0.97λ0 and includes the value 0.96λ0 equal to the ratio of the wavelengths of modes C and A.

It should be noted that the observed EL spectrum of the device is governed by the competition of these three types of the modes. It is possible to influence this competition by intentionally scratching the rear facet of the chip and destroying its mirror properties. Figures 10(a) and 10(b) compare EL spectra of a chip with as–cleaved mirror–like facets (Fig. 10(a)) and of a chip with the rear facet scratched (Fig. 10(b)), the rear facet being marked in Fig. 8(a). The EL spectrum in a conventional in–plane Fabry-Pérot resonator at a moderate current reveals all three modes (Fig. 10(a)), whereas an increase in current results in mode B strongly dominating the spectrum. Once the rear facet is scratched, the EL spectrum reveals two strong features related to mode A and mode C, mode C dominating the spectrum, mode B being suppressed. Thus, the scratched facet affects the tilted mode B stronger than mode C, in which a significant portion of the electromagnetic field propagates in the air in the near–field zone close to the surface. Further, Fig. 10(b) demonstrates a superlinear increase of the EL intensity upon injection current, indicating a possibility to reach lasing.

 figure: Fig. 10

Fig. 10 Effect of the facet scratching on the mode competition in electroluminescence (EL) spectra. (a) Device with cleaved facets emits light in all three modes. Increase in drive current leads to lasing in the tilted mode B. (b) In a device with the rear facet scratched, the tilted mode B is suppressed, and mode C coupled with the surface mode dominates the EL spectrum.

Download Full Size | PDF

Apart from scratching chip facets as addressed in Fig. 10(b), it is possible to support mode C by designing the device such that the gain spectrum will match the wavelength of the surface mode, allowing lasing in the mode significantly extended to the air. Further suppression of the vertical mode A can be achieved by reducing the number of the top DBR pairs thus increasing the mode losses. The combined mode C will be formed at a stronger coupling and, hence, at a broader resonance between the cavity mode and the surface mode, will be less wavelength–stabilized, but will have a larger intensity of the near field in the air. Continuing this trend will bring the active region to the top of the structure removing both the top DBR and the resonant VCSEL cavity, returning to the structure considered above in Section 2. The operation of such device will be considered in detail in [41].

Thus, by optimizing the thickness of the top DBR it is possible to suppress undesired VCSEL mode and to fabricate a device operating exclusively in the wavelength–stabilized near–field mode, including a near field laser. Such a device will then be advantageous for direct coupling into an optical fiber or a waveguide [19] for generation of a high–brightness beam. Further, a temperature stable distributed feedback (DFB) laser can be fabricated if the top layer of the DBR is formed of a dielectric material with nearly temperature–independent refractive index.

5. Summary

To conclude, we have presented modeling results demonstrating the existence of a near–field TM–polarized electromagnetic mode on a surface of a DBR, wherein the mode propagates along the surface, decays evanescently in the air and decays in an oscillating manner in the DBR. The mode allows near–field optoelectronic devices like diode lasers, superluminescent light–emitting diodes, novel types of wavelength–stabilized filters or couplers. A VCSEL structure containing a resonant cavity enables resonant coupling between the cavity mode and the surface mode rendering the resulting near–field mode wavelength–stabilized. A VCSEL structure fabricated in the form of an in–plane Fabry–Pérot resonator shows EL spectra containing three types of modes, namely a vertical VCSEL–like mode, a tilted mode, and a combined near–field mode with the cavity mode. The dominating lasing mechanism at room temperature is the tilted wave lasing. The tilt angle of the VCSEL mode is governed by the wavelength of the maximum gain. At small detuning of the VCSEL etalon wavelength from the gain towards the longer wavelengths the mode in the air is tilted with respect to the normal to the surface. At large detuning the mode angle in the crystal reaches the critical angle of the total internal reflection at semiconductor/air interface and becomes coupled to a surface–trapped in–plane mode with significant extension of the optical field to the air at the edge through in–plane emission lobe. By relative positioning the gain spectrum, by changing the number of the top DBR pairs (or DBR reflectivity), or by destroying the facet mirrors one can select the desirable lasing mechanism. Modeling shows that the near–field surface mode is necessarily TM–polarized. Polarization studies indeed reveal the reversal of polarization across the spectrum from TE–polarized tilted modes to purely TM–polarized near–field surface–trapped mode.

Funding

Ministry of Education and Science of the Russian Federation (Project 3.9787.2017/8.9).

References

1. L. D. Landau and E. M. Lifshits, “Electrodynamics of continuous media,” Pergamon Press, Bristol, 1963, §68.

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

3. E. Yablonovitch, “Photonic crystals,” J. Mod. Opt. 41(2), 173–194 (1994). [CrossRef]  

4. B. Wang, W. Dai, A. Fang, L. Zhang, G. Tuttle, T. Koschny, and C. M. Soukoulis, “Surface waves in photonic crystal slabs,” Phys. Rev. B 74(19), 195104 (2006). [CrossRef]  

5. S. Knauer, M. López-García, and J. G. Rarity, “Structured polymer waveguides on distributed Bragg reflector coupling to solid state emitter,” J. Opt. 19(6), 065203 (2017). [CrossRef]  

6. N. Engheta and P. Pelet, “Surface waves in chiral layers,” Opt. Lett. 16(10), 723–725 (1991). [CrossRef]   [PubMed]  

7. H. Y. Yang, J. A. Castaneda, and N. G. Alexopoulos, “Surface waves in gyrotropic substrates,” AP–S International Symposium (Digest) (IEEE Antennas and Propagation Society), IEEE, 1651–1654 (1991).

8. W. J. Tomlinson, “Surface wave at a nonlinear interface,” Opt. Lett. 5(7), 323–325 (1980). [CrossRef]   [PubMed]  

9. M. I. Dyakonov, “New type of electromagnetic wave propagating at an interface,” Sov. Phys. JETP 67(4), 714–716 (1988).

10. L. A. Coldren, S. W. Corzine, and M. L. Mašanović, “Diode lasers and photonic integrated circuits,” Wiley (2012).

11. N. N. Ledentsov and V. A. Shchukin, “Novel concepts for injection lasers,” Opt. Eng. 41(12), 3193–3203 (2002). [CrossRef]  

12. C. W. Lee, G. Singh, and Q. Wang, “Light extraction--a practical consideration for a plasmonic nano-ring laser,” Nanoscale 5(22), 10835–10838 (2013). [CrossRef]   [PubMed]  

13. T. Komljenovic, L. Liang, R.-L. Chao, J. Hulme, S. Srinivasan, M. Davenport, and J. E. Bowers, “Widely tunable ring–resonator semiconductor lasers (Review),” Appl. Sci. 7(7), 732 (2017). [CrossRef]  

14. L. A. Coldren, G. A. Fish, Y. Akulova, J. S. Barton, L. Johansson, and C. W. Coldren, “Tunable semiconductor lasers: A tutorial,” J. Lightwave Technol. 22(1), 193–202 (2004). [CrossRef]  

15. A. Taghizadeh, J. Mørk, and I.-S. Chung, “Vertical–cavity in–plane heterostructures: physics and applications,” Appl. Phys. Lett. 107(18), 181107 (2015). [CrossRef]  

16. Z. Wang, B. Zhang, and H. Deng, “Dispersion engineering for vertical microcavities using subwavelength gratings,” Phys. Rev. Lett. 114(7), 073601 (2015). [CrossRef]   [PubMed]  

17. M. C. Y. Huang, Y. Zhou, and C. J. Chang-Hasnain, “A surface–emitting laser incorporating a high–index–contrast subwavelength grating,” Nat. Photonics 1(2), 119–122 (2007). [CrossRef]  

18. I.-S. Chung and J. Mørk, “Silicon–photonics light source realized by III–V/Si–grating–mirror laser,” Appl. Phys. Lett. 97(15), 151113 (2010). [CrossRef]  

19. V. A. Shchukin, N. N. Ledentsov, J.-R. Kropp, G. Steinle, N. Ledentsov, Jr., K. D. Choquette, S. Burger, and F. Schmidt, “Engineering of optical modes in vertical–cavity microresonators by aperture placement: applications to single–mode and near–field lasers,” Proc. SPIE 9381, Vertical–Cavity Surface–Emitting Lasers XIX, C. Lei and K. D. Choquette, eds., 93810V (March 20, 2015).

20. V. A. Shchukin, N. N. Ledentsov, S. S. Mikhrin, I. L. Krestnikov, A. V. Kozhukhov, A. R. Kovsh, L. Ya. Karachinsky, M. V. Maximov, I. I. Novikov, and Yu. M. Shernyakov, “Tilted cavity laser (Critical review lecture),” Proc. SPIE Int. Soc. Opt. Eng. 5509, 61–71 (2004). [CrossRef]  

21. N. N. Ledentsov, V. A. Shchukin, M. V. Maximov, N. Yu. Gordeev, N. A. Kaluzhniy, S. A. Mintairov, A. S. Payusov, and Yu. M. Shernyakov, “Optical mode engineering and high power density per facet length (8.4 kW/cm) in tilted wave laser diodes,” Proc. SPIE 9733, 97330P (2016).

22. N. N. Ledentsov, V. A. Shchukin, M. V. Maximov, N. Yu. Gordeev, N. A. Kalyuzhnyi, S. A. Mintairov, A. S. Payusov, Yu. M. Shernyakov, K. A. Vashanova, M. M. Kulagina, and N. Y. Shmidt, “Passive cavity laser and tilted wave laser for Bessel–like beam coherently–coupled bars and stacks,” Proc. SPIE 9357, Physics and Simulation of Optoelectronic Devices XXIII, B. Witzigmann, M. Osiński, F. Henneberger and Y. Arakawa, eds., 93570X, (March 16, 2015).

23. M. V. Maximov, Yu. M. Shernyakov, I. I. Novikov, S. M. Kuznetsov, L. Ya. Karachinsky, N. Yu. Gordeev, V. P. Kalosha, V. A. Shchukin, and N. N. Ledentsov, “High–performance 640–nm–range GaInP–AlGaInP lasers based on the longitudinal photonic bandgap crystal with narrow vertical beam divergence,” IEEE J. Quantum Electron. 41(11), 1341–1348 (2005). [CrossRef]  

24. J. A. Lott, V. A. Shchukin, N. N. Ledentsov, A. M. Kasten, and K. D. Choquette, “Passive cavity surface emitting laser,” Electron. Lett. 47(12), 717–718 (2011). [CrossRef]  

25. V. Shchukin, N. Ledentsov, K. Posilovic, V. Kalosha, Th. Kettler, D. Seidlitz, M. Winterfeldt, D. Bimberg, N. Yu. Gordeev, L. Ya. Karachinsky, I. I. Novikov, Yu. M. Shernyakov, A. V. Chunareva, M. V. Maximov, F. Bugge, and M. Weyers, “Tilted wave lasers: A way to high brightness sources of light,” IEEE J. Quantum Electron. 47(7), 1014–1027 (2011). [CrossRef]  

26. D. McGloin and K. Dholakia, “Bessel beams: diffraction in a new light,” Contemp. Phys. 46(1), 15–28 (2005). [CrossRef]  

27. M. Mazilu, D. J. Stevenson, F. Gunn-Moore, and K. Dholakia, “Light beats the spread: non–diffracting beams,” Laser Photonics Rev. 4(4), 529–547 (2010). [CrossRef]  

28. M. Duocastella and C. B. Arnold, “Bessel and annular beams for materials processing,” Laser Photonics Rev. 6(5), 607–621 (2012). [CrossRef]  

29. M. McLaren, M. Agnew, J. Leach, F. S. Roux, M. J. Padgett, R. W. Boyd, and A. Forbes, “Entangled Bessel-Gaussian beams,” Opt. Express 20(21), 23589–23597 (2012). [CrossRef]   [PubMed]  

30. C. López-Mariscal and K. Helmerson, “Shaped nondiffracting beams,” Opt. Lett. 35(8), 1215–1217 (2010). [CrossRef]   [PubMed]  

31. N. N. Ledentsov, V. A. Shchukin, and J. A. Lott, “Ultrafast nanophotonic devices for optical interconnects” Proc. 2012 Advanced Research Workshop (FTM–7), June 25–29, 2012: Corsica, France, in “Future Trends in Microelectronics: Into the Cross Currents,” S. Luryi, J. Xu and A. Zaslavsky, eds., Wiley (2013).

32. V. Shchukin, N. Ledentsov, J. Kropp, G. Steinle, N. Ledentsov Jr, S. Burger, and F. Schmidt, “Single–mode vertical cavity surface emitting laser via oxide–aperture–engineering of leakage of high–order transverse modes,” IEEE J. Quantum Electron. 50(12), 990–995 (2014). [CrossRef]  

33. N. Ledentsov Jr, V. A. Shchukin, N. N. Ledentsov, J. R. Kropp, S. Burger, and F. Schmidt, “Direct evidence of the leaky emission in oxide–confined vertical cavity lasers,” IEEE J. Quantum Electron. 52(3), 2400207 (2016). [CrossRef]  

34. N. N. Ledentsov, N. Ledentsov Jr, M. Agustin, J.-R. Kropp, and V. A. Shchukin, “Application of nanophotonics to the next generation of surface–emitting lasers,” Nanophotonics 6(5), 813–829 (2017). [CrossRef]  

35. D. G. Deppe, M. Li, X. Yang, and M. Bayat, “Advanced VCSEL technology: self–heating and intrinsic modulation response,” IEEE J. Quantum Electron. 54(3), 2400209 (2018). [CrossRef]  

36. V. A. Shchukin, N. N. Ledentsov, V. P. Kalosha, N. Ledentsov, Jr., M. Agustin, J. R. Kropp, M. V. Maximov, F. I. Zubov, Yu. M. Shernyakov, A. S. Payusov, N. Yu. Gordeev, M. M. Kulagina, and A. E. Zhukov, “Thremally stable surface–emitting tilted wave laser,” Proc. SPIE 10552, Vertical–Cavity Surface–Emitting Lasers XXII, C. Lei and K. D. Choquette, eds., 1055207 (2018).

37. N. Ledentsov and V. Shchukin, “Optoelectronic device based on an antiwaveguiding cavity,” United States Patent 7,339,965, issued March 4, 2008, priority date April 7, 2004.

38. N. N. Ledentsov, V. A. Shchukin, V. P. Kalosha, N. N. Ledentsov, J.-R. Kropp, M. Agustin, Ł. Chorchos, G. Stępniak, J. P. Turkiewicz, and J.-W. Shi, “Anti-waveguiding vertical-cavity surface-emitting laser at 850 nm: From concept to advances in high-speed data transmission,” Opt. Express 26(1), 445–453 (2018). [CrossRef]   [PubMed]  

39. A. Yariv and P. Yeh, Optical Waves in Crystals (Wiley, 1984), section 6.3.

40. P. Westbergh, J. S. Gustavsson, B. Kögel, Å. Haglund, and A. Larsson, “Impact of photon lifetime on high–speed VCSEL performance,” IEEE J. Sel. Top. Quantum Electron. 17(6), 1603–1613 (2011). [CrossRef]  

41. V. A. Shchukin, N. N. Ledentsov, and A. Yu. Egorov, “Surface electromagnetic waves at the boundary of a distributed Bragg reflector,” Sci. Rep. (to be published).

42. X. Zhao, P. Palinginis, B. Pesala, C. Chang-Hasnain, and P. Hemmer, “Tunable ultraslow light in vertical-cavity surface-emitting laser amplifier,” Opt. Express 13(20), 7899–7904 (2005). [CrossRef]   [PubMed]  

43. D. A. B. Miller, “Optical physics of quantum wells,” (1996), https://www.researchgate.net/publication/260403078_Optical_Physics_of_Quantum_Wells

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 (10)

Fig. 1
Fig. 1 Optical power reflectance (OR) spectra of the model DBR. (a)–(c) Modeled spectra for light impinging from material #1 onto the DBR. (d), (e) Modeled spectra for light impinging from the air at ϑ = 89.9°.
Fig. 2
Fig. 2 (a, b) Modeled optical power reflectance spectra for light impinging from the air at different glancing angles of incidence. One can see that the within the interval of angles from 86° to 90° the spectral position of the dip, or the transparency wavelength is not a function of the glancing angle allowing angle–insensitive filters.
Fig. 3
Fig. 3 Refractive index profile of the model DBR and magnetic field strength profile in surface TM modes at different wavelengths.
Fig. 4
Fig. 4 Mode confinement coefficient of the surface TM mode which is inverse proportional to the mode extension length in the air, β = λ / ( 4 π l e x t ) , versus mode wavelength for different thickness of the topmost layer.
Fig. 5
Fig. 5 (a) Perspective view of a chip with a metal grid contact on top surface. The chip is a rectangular–shaped ~350 × 400 µm piece with perpendicular facets. The contact grid region has a total width of w~70 µm. d = 7 µm–wide metal stripes serving as non–alloyed metal contact forms periodic rectangular openings having a size of b × a = 10 × 40 µm. Numbers 1, 2, 3 indicate various positions of the photodetector measuring electroluminescence. (b) Infrared image of the lasing device. Gray structure in the center is the copper wire bonded on top of the grid.
Fig. 6
Fig. 6 Electroluminescence spectrum measured from the edge of the device in the stripe direction.
Fig. 7
Fig. 7 (a) Shift of three electroluminescence (EL) peaks versus temperature. (b) Evolution of the EL spectrum versus temperatures while passing the resonance between mode B and mode C.
Fig. 8
Fig. 8 (a) Far field profile of the electroluminescence (EL). (b) Spectrally–resolved far field profile of EL.
Fig. 9
Fig. 9 Polarization–resolved electroluminescence (EL) spectra of the device.
Fig. 10
Fig. 10 Effect of the facet scratching on the mode competition in electroluminescence (EL) spectra. (a) Device with cleaved facets emits light in all three modes. Increase in drive current leads to lasing in the tilted mode B. (b) In a device with the rear facet scratched, the tilted mode B is suppressed, and mode C coupled with the surface mode dominates the EL spectrum.

Equations (2)

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

H ( t , r ) = [ H x ( t , r ) , H y ( t , r ) , H z ( t , r ) ] = [ 0 , H y ( z ) exp ( i ω t ) exp ( i n e f f k 0 x ) , 0 ] ,
ε ( z ) d d z [ 1 ε ( z ) d d z H y ( z ) ] ε ( z ) k 0 2 H y ( z ) = n e f f 2 k 0 2 H y ( z ) .
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.