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

Enhanced spontaneous emission observed at one-dimensional photonic band edges

Open Access Open Access

Abstract

We recently reported experimental evidence for double enhancement of spontaneous emission due to increased photon density of states and small group velocity at photonic band-edge frequencies by observing angle-resolved emission and excitation spectra of photoluminescence [ K. Kuroda et al., Opt. Express 17, 13168 (2009) ]. The specimen we used was a one-dimensional photonic crystal composed of periodic multilayers of Ta2O5 and SiO2 with oxygen vacancies as light emitters. In the present study, we report on the lack of any excitation intensity dependence of the emission peak height and width, which excludes possibilities of nonlinear effects, the polarized emission spectra, and their comparison with theoretical calculations to further confirm our finding.

© 2009 Optical Society of America

1. INTRODUCTION

Photonic structures that have periodic variation of refractive index, which are called photonic crystals (PCs), on the order of a few hundred nanometers have been attracting much attention [1, 2, 3]. The periodic variation gives rise to a photonic bandgap (PBG) in the wavelength range from visible to near infrared (IR), where photon density of states (DOS) is vanishing. In contrast, light emission in the vicinity of the PBG is expected to be enhanced due to an increased DOS. Many studies have been conducted on modifying and designing the electromagnetic field by PCs so as to control and improve the optical properties of light emitters. The interplay between resonant electronic excitation and electromagnetic field in periodic structures, known as optical lattices, has also been studied [4, 5].

As far as photoluminescence in regular PCs is concerned, there are two enhancement mechanisms for spontaneous emission. One is increased photon DOS at the emission frequency, which accelerates the optical transition according to Fermi’s golden rule, and the other is small-photon velocity at the photoexcitation frequency, which results in a large amplitude of the local electric field due to conservation of Poynting’s vector [6]. As described in [6], the former mechanism is valid only for one-dimensional PCs or for photon emission in particular directions in two- and three-dimensional cases. So, although a change in the emission lifetime of CdSe quantum dots in inverse opal was observed in [7], it is not relevant to the divergent band-edge DOS. Reference [8] reported peculiar spontaneous emission spectra of GaAs quantum wells sandwiched by distributed Bragg reflector (DBR) structures. However, in this experimental study, the light emitter was located in the vicinity but outside of the one-dimensional PC. As for theoretical studies, Dowling et al. and Scalora et al. presented detailed calculations on finite one-dimensional PCs [9, 10].

As for the second mechanism, the enhanced local electric field increases the efficiency of photoexcitation of fluorescent species. This technique is called the band-edge excitation and was used for lasing [11] and control of the refractive index [12].

Although these two enhancement mechanisms are expected in the case of spontaneous emission of photons in regular PCs, they had not been observed because of the difficulty of fabricating sufficiently regular and thick periodic structures to create distinct peaks in the photon DOS and that of attaining a uniform distribution of light emitters in the PCs. However, we recently reported a clear observation of the double enhancement of spontaneous emission for a one-dimensional PC composed of periodic multilayers of Ta2O5 and SiO2 with oxygen vacancies as light emitters [6]. We presented angle-resolved excitation and emission spectra of photoluminescence to show the double enhancement due to the increased photon DOS at the emission wavelength and small group velocity at the excitation wavelength. The light emitters were distributed uniformly and the regularity of the specimen was good enough to observe the band-edge enhancement. We also presented a result of calculations of the enhanced luminescence spectra that supported the experimental observations.

In this paper, we report on the excitation intensity dependence of the emission peak height and width, the polarized emission spectra, and their comparison with theoretical calculations to further confirm our finding. In Section 2, the photonic band structure for two independent polarizations is calculated analytically in the weak modulation approximation. Then the enhancement factor of spontaneous emission in certain directions determined by the experimental conditions is derived, and it is shown that the enhancement factor is divergent at the band-edge frequencies. The experimental setup is briefly described in Section 3. In Section 4, experimental results on the excitation intensity dependence of emission peak height and width are presented, and it is concluded that what we observed was nothing but enhanced spontaneous emission. The polarization dependence of the emission spectra is also presented and compared with the theoretical calculation. Finally, a summary is given in Section 5.

2. THEORY

2A. Photonic Band Calculation

In this section, we calculate the dispersion relation of the electromagnetic eigenmodes and the photon DOS. Figure 1 schematically shows the experimental configuration. The specimen used in this study was a commercial optical notch filter made of dielectric multilayers designed to eliminate the 633nm line of the He–Ne laser (Semrock, NF01-633U). This notch filter gives a stop band from 620to645nm. SEM (scanning electron microscope) and EPMA (electron probe microanalyzer) analyses showed that the notch filter consisted of Ta2O5 and SiO2 multilayers on an SiO2 substrate. In Fig. 1, a is the lattice constant of the one-dimensional PC (the dielectric multilayers), θ is the observation angle, and ω and k are the angular frequency and wave number of emitted photons.

We have to consider two independent polarizations of the electromagnetic field. One is the p polarization for which the electric field is pointing in the xy plane. The other is the s polarization for which the electric field is pointing in the z direction, which is perpendicular to the xy plane. It is convenient to calculate the z component of the magnetic field, Hz, for the p polarization and that of the electric field, Ez, for the s polarization. From the Maxwell equations, we obtain the following eigen equations:

{x1ϵ(x)x+y1ϵ(x)y}Hz=ω2c2Hz,
1ϵ(x){2x2+2y2}Ez=ω2c2Ez,
where c is the velocity of light in free space and ϵ(x) is the spatially periodic dielectric constant. Here we assumed that the permeability of the specimen is equal to unity, since it does not contain magnetic materials.

The ratio of the width of the first PBG to its center wavelength is 25nm633nm4%, which is small enough to justify the weak modulation approximation. That is, it is a good approximation to truncate the expansion of ϵ(x) up to the first-order Fourier components provided we calculate the dispersion relation in the vicinity of the first PBG:

1ϵ(x)=η0+η1eiGx+η1*eiGx,
where G=2πa is the elementary reciprocal lattice constant. Because the structure is uniform in the y direction, spatial variation of the eigenfunctions in the y direction is simply described by a phase factor eikyy with the y component of the wave vector, ky. On the other hand, spatial variation in the x direction is described by the Bloch theorem because of the periodicity of the structure. Thus the eigenfunctions of the p and s polarizations should have the following forms, respectively:
Hz=m=Amei(kx+Gm)x+ikyy,
Ez=m=Bmei(kx+Gm)x+ikyy,
where
Gm=2mπa
is the reciprocal lattice constant and kx is the x component of the wave vector.

Let us consider the p polarization first. By substituting Eqs. (3, 4) into Eq. (1) and comparing the Fourier component of the same order, we obtain

Am[η0{(kx+Gm)2+ky2}ω2c2]+Am1η1{(kx+Gm1)(kx+Gm)+ky2}+Am+1η1*{(kx+Gm+1)(kx+Gm)+ky2}=0.
From this equation, we can find that A0 and A1 are large when both
ω2c2η0(kx2+ky2)andkxπa
are satisfied. Thus, under these conditions, we neglect all other Fourier components and obtain the following coupled equations.
A0[c2η0{(kx+G0)2+ky2}ω2]+A1c2η1{(kx+G1)(kx+G0)+ky2}=0,
A1[c2η0{(kx+G1)2+ky2}ω2]+A0c2η1*{(kx+G0)(kx+G1)+ky2}=0.

From the secular equation, we finally obtain

ω±caπ2(η0±|η1|)+a2ky2(η0|η1|)+ach22η0|η1|±2π2η02|η1|(π2a2ky2)π2(η0±|η1|)+a2ky2(η0|η1|),
where h=kxπa is the distance from the Brillouin zone boundary and ω+ (ω) gives the dispersion relation of the upper (lower) branch. This dispersion relation is accurate in the vicinity of the first PBG and for |h|πa.

For the s polarization, we substitute Eqs. (3, 5) into Eq. (2) and obtain

Bm[η0{(kx+Gm)2+ky2}ω2c2]+Bm1η1{(kx+Gm1)2+ky2}+Bm+1η1*{(kx+Gm+1)2+ky2}=0.
In the same conditions as given in Eq. (8), we obtain the following coupled equations for B0 and B1:
B0[c2η0{(kx+G0)2+ky2}ω2]+B1c2η1{(kx+G1)(kx+G0)+ky2}=0,
B1[c2η0{(kx+G1)2+ky2}ω2]+B0c2η1*{(kx+G0)(kx+G1)+ky2}=0.
From the secular equation, we finally obtain
ω±ca(η0±|η1|)(π2+a2ky2)+ach2η0±|η1|2π2+a2ky2{1±2π2(η0|η1|)|η1|(π2+a2ky2)}.

For ky=0, we obtain the following dispersion relation common to the p and s polarizations:

ω±πcη0±|η1|a{1±a2h2(2η0|η1|)2π2|η1|}.
So, there is no eigenmode in the frequency range of
πcaη0|η1|<ω<πcaη0+|η1|.
If |η1| is sufficiently smaller than η0, which justifies the weak modulation approximation, we obtain the center frequency, ωc, and the width, Δωg, of the gap by the Taylor expansion:
ωcπcη0aandΔωgπc|η1|aη0.
The three sample parameters (a, η0, η1) were determined from the refractive indices of the constituent materials (2.08 for Ta2O5 and 1.46 for SiO2), and the center frequency and width of the first bandgap found in the transmission spectrum. Thus we obtained a=155nm, η0=0.240, and η1=0.00877. η1η0 is sufficiently small as we expected.

2B. Enhancement Factor

In the present specimen, the luminescent species are distributed uniformly in each dielectric layer. Thus the effect of the spatial variation of the electric field intensity on the luminescence spectra is averaged out and can be neglected in the lowest-order approximation. So, the angle-resolved light emission spectrum is given by the genuine spectrum in free space times the enhancement factor of the photon DOS for the designated emission angle.

The enhancement factor of the photon DOS is obtained by calculating the ratio of the volume in the phase space in the PC to that in free space. In the present experiment, the wave vector is located in the xy plane, and its direction is specified by the observation angle θ. Thus the two dimensions are reduced by these two experimental conditions, and it is sufficient to take into consideration the one-dimensional volume ds=dkx2+dky2 in the PC and dk=dωc in free space. Since dky=dksinθ and kx is implicitly determined by Eq. (10) or Eq. (13),

c=dωdk=ωkxdkxdk+ωkysinθ.
Thus we obtain the enhancement factor as
dsdk=(dkxdk)2+(dkydk)2=(ωkx)2(cωkysinθ)2+sin2θ.
Note that this factor is divergent at the Brillouin zone boundary for any ky, since ωkx=0 there.

The theoretical emission intensity at each observation angle was obtained as a product of the genuine emission spectrum and the enhancement factor calculated according to Eq. (18) and is plotted in Fig. 2 , where thick lines denote the s polarization and thin lines denote the p polarization. Taking into account the lifetime of the eigenmodes in the actual specimen of a finite thickness (30μm), a small constant imaginary part was added to ωkx. In Fig. 2, we can clearly observe the following features. First, there is a frequency range in each spectrum where the emission intensity is equal to zero, which originates from the PBG and shifts to the shorter wavelength with increasing observation angle. Second, on each side of the PBG, there is one distinct emission peak due to a divergent photon DOS. Thirdly, the center of the PBG is nearly the same for the s and p polarizations, whereas its width is wider for the s polarization. In Section 4, we show that these features are actually observed in the experiment.

3. EXPERIMENTAL

The experimental setup for the angle-resolved photoluminescence measurement was essentially the same as in [6]. In brief, the 532nm emission line of a frequency-doubled continuous Nd:YAG laser was used as an excitation light source. The laser beam was weakly focused onto the specimen by a lens with a focal length of 100mm. The beam diameter on the sample surface was about 15μm. Luminescence light from the specimen was collected by another lens with the same focal length. A pinhole was placed before the collection lens to decrease the uncertainty of the detection angle down to 4° (full width). The dependence of the emission intensity on the detection angle (θ) was measured by rotating the sample with a rotation stage. To measure the polarization of the emission, a polarizer was placed behind the collection lens. Emission intensity was measured by a triple-grating spectrometer with a charge-coupled device (CCD) detector.

4. RESULTS AND DISCUSSION

4A. Excitation Power Dependence

First, we should confirm that all emission spectra were measured in the linear region and that lasing, or any other nonlinear emission process, was not involved. Figure 3 shows emission spectra in the normal direction (θ=0) measured with various excitation powers. The scale of the vertical axis of each spectrum is proportional to the excitation power.

It is clearly seen in Fig. 3 that there is one distinct peak on both edges of an opaque range from 620to645nm caused by the PBG. These peaks originate from enhanced spontaneous emission due to the increased photon DOS discussed in Subsection 2B and coincide well with Fig. 2. The line shapes of the emission spectra with five different excitation powers in Fig. 3 are quite similar to each other. Particularly, the peak intensity is proportional to the excitation intensity and the peak width is independent of the excitation intensity.

To analyze these two features further, we examined the peak height and width more accurately by a curve fitting analysis. As shown in Figs. 4a, 4b , each peak was fitted with a Lorentzian function, and height and width were estimated. As we can see, the line shape on the gap frequency side is reproduced well by the Lorentzian curve, because the emission probability in the gap is mainly determined by the photon escape time through the specimen surface, whereas there is an appreciable deviation in the opposite frequency side where spontaneous emission takes place according to the photon DOS in the specimen, which cannot be reproduced by a Lorentzian function.

The peak intensity and width are plotted as functions of excitation intensity in Figs. 4c, 4d, respectively. In Fig. 4c, open (solid) circles denote the peak intensity of the longer (shorter) wavelength peak and open squares denote the mean value of the intensity between 680 and 700nm, where the enhanced spontaneous emission by an increased photon DOS is not relevant. Solid lines are the best fit by the power-law dependence. As we can see, the observed emission intensity is well reproduced by the power law, and the exponents of all three data were close to unity (1.00±0.08). So, we can conclude that no nonlinear process such as stimulated emission, amplified spontaneous emission, or lasing was involved in our experiment. This fact is further corroborated by the constant width of the emission peaks given in Fig. 4d. Thus we can conclude that the two emission peaks on both ends of the PBG originate from spontaneous emission.

4B. Polarization Dependence

Figure 5 depicts the polarized emission spectra observed for detection angles θ from 0° to 45°, where thick and thin lines denote s- and p-polarized emissions. The overall features of this figure agree well with Fig. 2. First, the PBG and two emission peaks shift to the shorter wavelength with increasing detection angle. Second, the center wavelength of the PBG is nearly the same for the s and p polarizations. Third, the distance between the two peaks is larger for the s polarization than for the p polarization.

On the other hand, there are two main differences between the calculation and the experiment. One difference is the nonzero emission intensity in the gap frequency range observed in the experiment, which is a consequence of the finite number of the periodic dielectric layers of the specimen. The photon DOS in the gap frequency range is exactly equal to zero in the calculation, because we assumed the dispersion of an ideal infinite PC. On the other hand, those emitters located in the vicinity of the sample surface can emit photons even in the gap frequency range because the influence of the PC environment is imperfect. So, we observed emission intensity in the gap range as shown in Fig. 5. As for the height of the two peaks, it is proportional to the photon DOS in the PC. In an ideal one-dimensional PC with an infinite number of layers, the photon DOS diverges on both ends of the PBG. However, in the actual PC, the number of periodic layers is finite, which leads to a finite lifetime of electromagnetic modes, and hence, to the decrease of the photon DOS. So, generally speaking, we can expect a larger peak height for the larger number of periodic layers.

The second difference is the larger peak width for the larger detection angle. This is caused by a constraint in the experiment. We used a pinhole to decrease the detection angle uncertainty down to 4° (FWHM). So, for example, the spectrum for θ=40° in Fig. 5 represents a weighted average of emission spectra from θ=38° to 42°. Because the variation of the peak position with the detection angle is larger for larger θ, the observed linewidth in the averaged spectrum becomes larger with increasing θ.

To further confirm the agreement between the calculation and observation, we compared the difference of the peak positions between the s and p polarizations as a function of the detection angle. Figures 6a, 6b show a magnified view of the two peaks of the s-(thick curve) and p-polarized (thin curve) emissions at θ=20°, from which we obtain the difference of the peak wavelengths. Then, the experimental observation was compared with the calculation in Fig. 6c, where it is clearly seen that the experiment and calculation agree well with each other.

5. CONCLUSION

We investigated both theoretically and experimentally the photoluminescence of a one-dimensional PC composed of periodic multilayers of Ta2O5 and SiO2 with oxygen vacancies as light emitters. The photonic band structures for s and p polarizations were calculated analytically in the weak modulation approximation, which was justified by the small PBG width of the specimen. The enhancement factor of spontaneous emission derived from the band structure was shown to be divergent at the band-edge frequencies.

Then we presented the excitation intensity dependence of the emission peak height and width to exclude possibilities of nonlinear processes such as amplified spontaneous emission and lasing. Thus we confirmed that what we observed was nothing but enhanced spontaneous emission due to increased photon DOS. We also compared polarized emission spectra measured at various detection angles with theoretical calculations to confirm their good agreement.

ACKNOWLEDGMENTS

The authors appreciate the SEM and EPMA analyses by Toray Research Center. This work was supported by KAKENHI (20340080).

 figure: Fig. 1

Fig. 1 Schematic illustration of the experimental configuration. θ is the observation angle. ω and k denote the angular frequency and wave vector of emitted light.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 Calculated emission spectra of s (thick curve) and p (thin curve) polarizations for various observation angles.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Emission spectra in the normal direction (θ=0) measured with various excitation powers from 0.45mWto36mW. The scale of the vertical axis is proportional to the excitation power.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Curve fitting for the shorter wavelength peak with a Lorentzian function and (b) that for the longer wavelength peak. (c) Excitation intensity dependence of the emission intensity. Open circles, solid circles, and open squares denote the longer wavelength peak, the shorter wavelength peak, and the mean value between 680 and 700nm. Solid lines are the best fit by the power law. Exponents of the power-law dependence obtained by the best fit are all close to unity (1.00±0.08). (d) Linewidths of the shorter (solid circles) and longer (open circles) wavelength peaks.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Polarized emission spectra for θ=0 to 45°. Thick and thin curves show s- and p-polarized emissions, respectively.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Magnified view of the longer wavelength peaks of the s-(thick curve) and p-polarized (thin curve) emissions at θ=20°, and (b) that of the shorter wavelength peaks. (c) Difference of peak wavelengths between the s- and p-polarized emissions as a function of detection angle θ. Solid circles, longer wavelength peak; Open squares, shorter wavelength peak; Solid line, theory.

Download Full Size | PDF

1. J. D. Joannopoulos, R. D. Meade, and J. N. Winn, Photonic Crystals: Molding the Flow of Light (Princeton Univ. Press, 1995).

2. K. Sakoda, Optical Properties of Photonic Crystals, 2nd ed. (Springer, 2004).

3. K. Sakoda, “Optics of photonic crystals,” Opt. Rev. 6, 381–392 (1999). [CrossRef]  

4. M. Hübner, J. P. Prineas, C. Ell, P. Brick, E. S. Lee, G. Khitrova, H. M. Gibbs, and S. W. Koch, “Optical lattices achieved by excitons in periodic quantum well structures,” Phys. Rev. Lett. 83, 2841–2844 (1999). [CrossRef]  

5. E. L. Ivchenko, M. M. Voronov, M. V. Erementchouk, L. I. Deych, and A. A. Lisyansky, “Multiple-quantum-well-based photonic crystals with simple and compound elementary supercells,” Phys. Rev. B 70, 195106 (2004). [CrossRef]  

6. K. Kuroda, T. Sawada, T. Kuroda, K. Watanabe, and K. Sakoda, “Doubly enhanced spontaneous emission due to increased photon density of states at photonic band edge frequencies,” Opt. Express 17, 13168–13177 (2009). [CrossRef]   [PubMed]  

7. P. Lodahl, A. F. van Driel, I. S. Nikolaev, A. Irman, K. Overgaag, D. Vanmaekelbergh, and W. L. Vos, “Controlling the dynamics of spontaneous emission from quantum dots by photonic crystals,” Nature 430, 654–657 (2004). [CrossRef]   [PubMed]  

8. M. D. Tocci, M. Scalora, M. J. Bloemer, J. P. Dowling, and C. M. Bowden, “Measurement of spontaneous-emission enhancement near the one-dimensional photonic band edge of semiconductor heterostructures,” Phys. Rev. A 53, 2799–2803 (1996). [CrossRef]   [PubMed]  

9. J. P. Dowling, M. Scalora, M. J. Bloemer, and C. M. Bowden, “The photonic band edge laser: A new approach to gain enhancement,” J. Appl. Phys. 75, 1896–1899 (1994). [CrossRef]  

10. M. Scalora, J. P. Dowling, M. Tocci, M. J. Bloemer, C. M. Bowden, and J. W. Haus, “Dipole emission rates in one-dimensional photonic band-gap materials,” Appl. Phys. B 60, S57–S61 (1995).

11. Y. Matsuhisa, Y. Huang, Y. Zhou, S.-T. Wu, Y. Takao, A. Fujii, and M. Ozaki, “Cholesteric liquid crystal laser in a dielectric mirror cavity upon band-edge excitation,” Opt. Express 15, 616–622 (2007). [CrossRef]   [PubMed]  

12. M. Astic, Ph. Delaye, R. Frey, G. Roosen, R. André, N. Belabas, I. Sagnes, and R. Raj, “Time resolved nonlinear spectroscopy at the band edge of 1D photonic crystals,” J. Phys. D 41, 224005 (2008). [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 (6)

Fig. 1
Fig. 1 Schematic illustration of the experimental configuration. θ is the observation angle. ω and k denote the angular frequency and wave vector of emitted light.
Fig. 2
Fig. 2 Calculated emission spectra of s (thick curve) and p (thin curve) polarizations for various observation angles.
Fig. 3
Fig. 3 Emission spectra in the normal direction ( θ = 0 ) measured with various excitation powers from 0.45 mW to 36 mW . The scale of the vertical axis is proportional to the excitation power.
Fig. 4
Fig. 4 Curve fitting for the shorter wavelength peak with a Lorentzian function and (b) that for the longer wavelength peak. (c) Excitation intensity dependence of the emission intensity. Open circles, solid circles, and open squares denote the longer wavelength peak, the shorter wavelength peak, and the mean value between 680 and 700 nm . Solid lines are the best fit by the power law. Exponents of the power-law dependence obtained by the best fit are all close to unity ( 1.00 ± 0.08 ) . (d) Linewidths of the shorter (solid circles) and longer (open circles) wavelength peaks.
Fig. 5
Fig. 5 Polarized emission spectra for θ = 0 to 45°. Thick and thin curves show s- and p-polarized emissions, respectively.
Fig. 6
Fig. 6 Magnified view of the longer wavelength peaks of the s-(thick curve) and p-polarized (thin curve) emissions at θ = 20 ° , and (b) that of the shorter wavelength peaks. (c) Difference of peak wavelengths between the s- and p-polarized emissions as a function of detection angle θ. Solid circles, longer wavelength peak; Open squares, shorter wavelength peak; Solid line, theory.

Equations (20)

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

{ x 1 ϵ ( x ) x + y 1 ϵ ( x ) y } H z = ω 2 c 2 H z ,
1 ϵ ( x ) { 2 x 2 + 2 y 2 } E z = ω 2 c 2 E z ,
1 ϵ ( x ) = η 0 + η 1 e i G x + η 1 * e i G x ,
H z = m = A m e i ( k x + G m ) x + i k y y ,
E z = m = B m e i ( k x + G m ) x + i k y y ,
G m = 2 m π a
A m [ η 0 { ( k x + G m ) 2 + k y 2 } ω 2 c 2 ] + A m 1 η 1 { ( k x + G m 1 ) ( k x + G m ) + k y 2 } + A m + 1 η 1 * { ( k x + G m + 1 ) ( k x + G m ) + k y 2 } = 0 .
ω 2 c 2 η 0 ( k x 2 + k y 2 ) and k x π a
A 0 [ c 2 η 0 { ( k x + G 0 ) 2 + k y 2 } ω 2 ] + A 1 c 2 η 1 { ( k x + G 1 ) ( k x + G 0 ) + k y 2 } = 0 ,
A 1 [ c 2 η 0 { ( k x + G 1 ) 2 + k y 2 } ω 2 ] + A 0 c 2 η 1 * { ( k x + G 0 ) ( k x + G 1 ) + k y 2 } = 0 .
ω ± c a π 2 ( η 0 ± | η 1 | ) + a 2 k y 2 ( η 0 | η 1 | ) + a c h 2 2 η 0 | η 1 | ± 2 π 2 η 0 2 | η 1 | ( π 2 a 2 k y 2 ) π 2 ( η 0 ± | η 1 | ) + a 2 k y 2 ( η 0 | η 1 | ) ,
B m [ η 0 { ( k x + G m ) 2 + k y 2 } ω 2 c 2 ] + B m 1 η 1 { ( k x + G m 1 ) 2 + k y 2 } + B m + 1 η 1 * { ( k x + G m + 1 ) 2 + k y 2 } = 0 .
B 0 [ c 2 η 0 { ( k x + G 0 ) 2 + k y 2 } ω 2 ] + B 1 c 2 η 1 { ( k x + G 1 ) ( k x + G 0 ) + k y 2 } = 0 ,
B 1 [ c 2 η 0 { ( k x + G 1 ) 2 + k y 2 } ω 2 ] + B 0 c 2 η 1 * { ( k x + G 0 ) ( k x + G 1 ) + k y 2 } = 0 .
ω ± c a ( η 0 ± | η 1 | ) ( π 2 + a 2 k y 2 ) + a c h 2 η 0 ± | η 1 | 2 π 2 + a 2 k y 2 { 1 ± 2 π 2 ( η 0 | η 1 | ) | η 1 | ( π 2 + a 2 k y 2 ) } .
ω ± π c η 0 ± | η 1 | a { 1 ± a 2 h 2 ( 2 η 0 | η 1 | ) 2 π 2 | η 1 | } .
π c a η 0 | η 1 | < ω < π c a η 0 + | η 1 | .
ω c π c η 0 a and Δ ω g π c | η 1 | a η 0 .
c = d ω d k = ω k x d k x d k + ω k y sin θ .
d s d k = ( d k x d k ) 2 + ( d k y d k ) 2 = ( ω k x ) 2 ( c ω k y sin θ ) 2 + sin 2 θ .
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.