• Aucun résultat trouvé

4000-enhancement of difference frequency generation in a mode-matching metamaterial

N/A
N/A
Protected

Academic year: 2021

Partager "4000-enhancement of difference frequency generation in a mode-matching metamaterial"

Copied!
14
0
0

Texte intégral

(1)

HAL Id: hal-02976580

https://hal.archives-ouvertes.fr/hal-02976580

Submitted on 4 Dec 2020

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Léna Soun, Sébastien Héron, Hasnaa El Ouazzani, Baptiste Fix, Riad Haïdar, Patrick Bouchon

To cite this version:

Léna Soun, Sébastien Héron, Hasnaa El Ouazzani, Baptiste Fix, Riad Haïdar, et al.. 4000-

enhancement of difference frequency generation in a mode-matching metamaterial. Optics Express,

Optical Society of America - OSA Publishing, 2020, 28 (19), pp.27210. �10.1364/OE.398733�. �hal-

02976580�

(2)

4000-enhancement of difference frequency generation in a mode-matching metamaterial

L ÉNA S OUN , 1 S ÉBASTIEN H ÉRON , 1 H ASNAA E L O UAZZANI , 1 B APTISTE F IX , 1 R IAD H AÏDAR , 1,2 AND P ATRICK B OUCHON 1,*

1

DOTA, ONERA, Université Paris-Saclay, F-91123 Palaiseau, France

2

École Polytechnique, Département de Physique, 91128 Palaiseau, France

*patrick.bouchon@onera.fr

Abstract: In the wake of the control of light at the sub-wavelength scale by nanoresonators, metasurfaces allowing strong field exaltations are an attractive platform to enhance nonlinear processes. Recently, high efficiency second harmonic and difference frequency generations were demonstrated in metasurfaces that generate a nonlinear polarization normal to the surface.

Here, we introduce a mode matched resonator that is able to produce this particular nonlinear polarization in a layer of gallium arsenide associated with a gold metasurface. The nonlinear conversion mechanism is described as a two-step process in which efficiency is shown to yield a good colocalization and a strong enhancement of the pump fields, as well as a high extraction efficiency of the generated field. This mode-matched metasurface is able to reach a difference frequency generation (DFG) efficiency of 10

−2

W/W

2

. This opens a new paradigm where alternative nonlinear materials could be reintroduced in metasurfaces and yields even higher efficiency than high effective χ

(2)

structures.

© 2020 Optical Society of America under the terms of theOSA Open Access Publishing Agreement

1. Introduction

Nanoantennas have the ability to concentrate light in subwavelength volume, which is particulary appealing for exalting light-matter interactions [1,2]. Taking advantage of this capacity, many works have been focused on developing nonlinear metasurfaces [3–7]. They exhibit effective nonlinearities several orders of magnitude higher than in bulk materials, and can be tailored to tune the wavelength, polarization or phase of the emitted wave.

Second order nonlinear conversion have been studied in all-dielectric metasurfaces based on nanocylinders [8–10], gratings [11] or waveguides [12], as well as in metallic metamaterials [13–17] and photonic crystals [18]. While they can be used to enhance the nonlinearities of metal [13,19], III-V semiconductors like gallium arsenide are particularly interesting since they have a high nonlinear susceptibility χ

(2)

' 150 pm/V in the infrared [20]. It has been included in various antennas to enhance further the nonlinear SHG efficiency, for instance GaAs was included in a subwavelength metallic hole [21] or a subwavelength metallic slit [22,23], or associated with metallic split-ring resonators [24]. More recently, GaAs nanopatterned layer were shown to support Mie-type modes that enhanced further the SHG efficiency [25]. Difference frequency generation (DFG) was also enhanced by using split ring resonators [26]. DFG is a very promising way to produce sources of light spanning the infrared and terahertz regions. These processes provide high enhancement ratio of the nonlinear conversion, still, they have maximum nonlinear conversion effiencies around 10

−5

W/W

2

.

Another way to enhance the effective nonlinear susceptibility of materials is to engineer multiple quantum wells [27]. Naturally, with the uprising of metasurfaces, both concepts were associated and demonstrated to provide very high efficiency for the SHG [28,29]. High efficiency DFG has also been demonstrated for generating light in the infrared [30] and in the Terahertz [31].

These processes have demonstrated conversion effiencies up to 10

−2

W / W

2

for SHG and up to 10

−3

W/W

2

for DFG. One particularity of these multiple quantum wells coupled to a metasurface

#398733 https://doi.org/10.1364/OE.398733

Journal © 2020 Received 27 May 2020; revised 29 Jul 2020; accepted 29 Jul 2020; published 31 Aug 2020

(3)

is that the nonlinear polarization is enhanced in the longitudinal direction (P

z

), while in most metasurfaces, the nonlinear polarization is enhanced in the lateral directions.

Here, we introduce a multi-resonant nanostructure that includes a single epitaxial GaAs layer able to provide an even higher DFG efficiency (up to 10

−2

W/W

2

) than multiple quantum wells structures. Since the generation of P

z

is very rarely studied, we introduce a theoretical framework to compute it in the case of modal methods. It takes into account the specific boundaries conditions encountered by the generated propagating waves in the nonlinear medium. Besides, the high efficiency is obtained thanks to a careful mode-matching optimization so that each pump and signal wavelengths is paired with a resonance. We also show how to describe the DFG as a two-steps process, each impacting its conversion efficiency. In the first step, the nonlinear polarization is obtained by a strong co-localization of the pump fields, then the emitted light depends on the extraction efficiency of the structure.

2. Modal method for nonlinear effects 2.1. Waveguide resonator structure

The structure studied in the whole manuscript is a waveguide resonator, which is depicted in Fig. 1. The waveguide structure is composed of a layer of material with a χ

(2)

(thickness h) encapsulated between a continuous gold layer, which is thick enough to be optically opaque and is considered as a semi-infinite layer, and a gold grating (period d, thickness t and width of the gold ribbons w). The model of gold used here is a Drude model in agreement with experimental data in the infrared range:

Au

(λ) = 1 − 1/(λ

p

/λ(λ

p

/λ + iγ)) with λ

p

= 159 nm and γ = 0.0075 [32].

Fig. 1. Waveguide resonator structure made of one layer of gold, one layer of nonlinear material (thickness h, nonlinear susceptibility χ

(2)

) and a gold grating (period d, thickness t and width w). Two pump waves are incoming on the structure with a normal incidence (ω

1

and ω

2

).

Computing the field generated by the nonlinear polarization along the z axis is pretty unusual,

and we introduce a theoretical framework for modal methods that takes into account the specific

boundary conditions encountered by the generated field in this case. Noteworthily, the simulation

method described in this section can be applied to any grating structures that includes nonlinear

materials.

(4)

2.2. Nonlinear polarization computation

The nonlinear complex polarization generated by second order nonlinear effects is given by the equation:

P

(2)

3

= ω

1

− ω

2

) =

0

χ

(2)

E

1

1

)E

2

2

) , (1) where E

1

1

) and E

2

2

) are the complex linear pump fields .

These linear pump fields have to be computed inside the structure in each area that exhibits a nonlinear susceptibility. In the waveguide resonator structure, the pump fields are computed in the gallium arsenide layer thanks to a B-spline modal method (BMM) [33]. In this method, the electric and magnetic fields are interpolated on a basis of non uniform B-splines basis along the x axis in each layer. Eigenmodes of the electromagnetic fields are computed by solving Maxwell equations at each sampling coordinate x.

The xyz components of the electric field E, at the sampling coordinates x and fixed height z, can be represented as the superposition of two waves, one is up-going, while the other is down-going in the structure:

E(z) = E

+

(z) + E

(z) . (2)

The fields are written at time frequency ω thanks to their projections on the eigenmodes basis E

Cω

(see Fig. 2):

E

±

(z) = E

ωC

diag(e

±ikm(zz0)

)U

±

(z

0

), (3) where U

±

are the vectors of the modes amplitudes, which are given relatively to a height z

0

. In order to compute them, the scattering matrices at each interface are determined thanks to the continuity of the tangential electric and magnetic fields [34]. Block sub-matrices are then used in the following of the manuscript. For instance, S

αBA

stands for the transmission scattering matrix from layer B to medium A, and S

αBB

for the reflection scattering matrix on interface α in layer B (see Fig. 2) [35].

Fig. 2. The nonlinear source dΠ(z) is generated in a GaAs sub-layer of height dz. Then we

calculate the propagation in the GaAs layer (dU). After a integration of all the dΠ(z) in the

GaAs layer, the field is propagated in the other layers.

(5)

2.3. Construction of the nonlinear field

Once the nonlinear polarization is computed, each area of the nonlinear layer generates a source term that contributes to the generated field. The total generated field is the sum of each of these elementary contributions, and in practice, the nonlinear layer is discretized along the z axis. The sublayer located at height z and of thickness dz generates a source term d Π(z) , and the scattering matrices and propagations undergone by the generated fields dU

±

are shown in Fig. 2.

Due to our choice of the pump fields and the orientation of the crystal in layer C, the source term is oriented along the z axis. This orientation determines the boundary conditions encountered by the elementary generated fields. For other pump field choices, the source term can be oriented according to x or y, in which case, boundary conditions have been described by Heron et al. [36].

First, the Maxwell-Ampere and Maxwell-Faraday equations are integrated on a sub-layer (thickness dz) and a period d:

∂E

x

(z

+

)

∂x dU

+

− ∂E

x

(z

)

∂x dU

= µ

0

ω

2

(P

z

+

0

E

z

)dz, (4) where z

±

= z ±

dz2

.

The following equations describe the continuity of the field across the sub-layer:

dU

+

(z

+

) + dU

(z

+

) = dU

+

(z

) + dU

(z

). (5) The upward field at the lower interface of the sublayer corresponds to the downward field that has propagated in the GaAs layer and been reflected by lower gold interface γ :

dU

+

(z

) = M

γ

dU

(z

) , (6)

The downward field at the upper interface of the sublayer corresponds to the upward field that has propagated in the GaAs layer and been reflected by upper gold interface β :

dU

(z

+

) = M

β

dU

+

(z

+

) , (7)

with M

i

= Q

iz

S

iCC

Q

zi

, and Q

zi

is the propagation vector from the height z to the interface i.

Equations (5)–(7) can be written as:

M

γ

dU

(z

) + dΠ(z) = dU

+

(z

+

), (8) and

M

β

dU

+

(z

+

) + dΠ(z) = dU

(z

), (9) with d Π(z) = µ

0

ω

2

(P

z

+

0

E

z

).

Then the contribution of each sub-layer is computed as:

dU

+

(z

+

) = [ I − M

γ

M

β

]

−1

[M

γ

+ I ] d Π(z) , (10) dU

(z

) = [ I − M

β

M

γ

]

−1

[M

β

+ I ] d Π(z) , (11) where I stands for the unity matrix.

Finally, the total generated fields are computed from E

ωC3

the linear field exalted at ω

3

and by integrating dU

+

and dU

on all the sub-layers:

E

+

(z) = E

ωC3

z

Q

zz0

dU

+

(z

0

)dz

0

, (12)

E

(z) = E

ωC3

z

Q

zz0

dU

(z

0

)dz

0

. (13)

This total generated field, once computed in the nonlinear layer, can be propagated and computed

in all the layers by straightforward scattering matrices computations.

(6)

3. Waveguide resonator 3.1. Resonance behavior

We now consider the resonator with gallium arsenide as the nonlinear material. The gallium arsenide layer has a second-order nonlinear susceptibility χ

xyz(2)

, only these xyz coefficients are non zero due to its crystal symmetries (zinc blende type, 43m class). The nonlinear properties of gold are not considered in the following, as the nonlinear susceptibility of gold is far smaller than for gallium arsenide [37]. The structure is illuminated with two pump waves, the first one is transverse electric (TE) polarized at a ω

1

frequency while the second one is transverse magnetic (TM) polarized at a ω

2

frequency. They generate a nonlinear polarization that is particularly strong along the z axis that gives birth to different second order nonlinear effects, like SHG, SFG (sum-frequency generation) and DFG. Here we present a design that is exalting a generated signal by difference of frequencies at ω

3

= ω

1

− ω

2

. The structure can be engineered so that it supports Fabry-Perot and guided modes in the GaAs layer [38] that subsequently improve the efficiency of the DFG. The DFG can be described as a two-steps process that will drive the conversion efficiency. First, the nonlinear polarization is obtained by a good co-localization of the pump fields, then the emitted light depends on the extraction efficiency of the structure. The importance of these two steps are illustrated on a grating with dimensions d = 458 nm, w = 220 nm, t = 40 nm, h = 145 nm. Figure 3 shows the computed reflectivity spectra in both polarizations for this structure. It exhibits two resonances in TE polarization, which are due to a guided mode resonance [38] and a Fabry-Perot resonance, and three in TM polarization which are three harmonics of a Fabry-Perot resonance in the metal-insulator-metal (MIM) cavity [39].

These different types of resonances do not depend on the same geometrical parameters, as shown in Fig. 4.

Fig. 3. Reflectivity spectra of the waveguide resonator in TM polarization (red curve) and in TE polarization (green curve). Two scenarios of DFG are pointed out with the phase matching

λ1

1

λ1

2

=

λ13

. The parameters of the grating are d = 458 nm, w = 220 nm, t = 40 nm, h = 145 nm.

3.1.1. Vertical Fabry-Perot resonance

In this resonance, the light undergoes successive reflections between the gold miroir and the gold

grating, which acts as a partially reflective mirror (see Fig. 4(a)). This resonance appears in TE

(7)

Fig. 4. Resonance wavlength fonction of some parameters :(a) Vertical Fabry-Perot

resonance fonction of the GaAs layer height h (in TE polarization), (b) Guided mode

resonance fonction of the periodicity of the grating d (in TE polarization), (c) Horizontal

Fabry-Perot resonance fonction of the ribbons width w (in TM polarization).

(8)

polarization, and is ruled by this equation [40]:

λ

r

= 2nh + λ

φ

, (14)

where h is the height of the cavity (here the thickness of the GaAs layer), n the refractive index of the GaAs, and λ

φ

describes the reflecting condidions of the cavity. The dependance in h is shown in Fig. 4(a).

3.1.2. Guided modes resonance

In this resonance, the light propagates horizontally in the GaAs layer, by successive reflections.

Here the light comes from the top of the system, and is injected in the GaAs layer thanks to the diffraction created by the grating (see Fig. 4(b)). This brings to some rules on the geometrical parameters: First, in order to optimize the injection in the structure, we exclude the reflection diffraction [41]:

sin θ = sin θ

i

d >1, (15)

with θ

i

the incident angle on the grating, θ the angle coming out of the grating (coming in the GaAs layer), n GaAs refractive index, m the order of diffraction and d the periodicity.

In normal incidence it remains to:

d<λ<nd. (16)

The dependance in d is shown in Fig. 4(b).

Secondly, to prevent a cut-off phenomenon, we can write the phase condition:

2k

z

(h + 2δ) = 2π. (17)

Which brings to a cut-off thickness of GaAs:

h> λ

2n − 2δ. (18)

3.1.3. Horizontal Fabry-Perot resonance

This resonance results from the coupling between the surface plasmons propagating on both dielectric/metal interfaces on either side of the dielectric, which must be thin enough to allow the coupling. The coupling will change the index under the bar, which will create a horizontal Fabry-Perot cavity (Fig. 4(c)) [39]. The resonance wavelength is ruled by this equation:

λ

r

(m) ' 2n

eff

(w + 2δ)

m − 1/2 , (19)

where δ si the skin depth of the metal (δ = 25 nm), m is the considered harmonic, and n

eff

is the effective index : n

eff

= q

1 +

h

d

m

, with

d

the dielectric permitivity (here GaAs) and

m

the metal permitivity (here gold). Figure 4(c) shows the dependance on w.

These different rules of conception allow us to drastically reduce the range of geometrical parameters on which we realize the optimisation with a brute force scanning. Actually, all resonances depends on the thickness h of the GaAs layer. Nevertheless, the guided modes resonances depends on the periodicity d, regardless the width of the ribbons w. On the contrary, the horizontal Fabry-Perot resonance is extremely dependent on w. These nearly independent behaviors provide us design rules for a triple-resonance matching for DFG generation. Two scenarios for generations of DFG that takes advantage of the concentration of the fields at the pump frequencies are introduced. In both cases, the DFG is given by:

λ11

λ1

2

=

λ13

. In the first

case (pointed out in blue in the figures), there are resonances at each wavelengths λ

1

= 1.064 µ m,

λ

2

= 1.6 µm and λ

3

= 3.17 µm. In the second case (pointed out in orange in the figures), there is

no resonance at the signal wavelength, and the wavelengths values are λ

10

= 1.2 µm, λ

20

= 1.55 µm

and λ

30

= 5.3 µm.

(9)

3.2. Co-localization of the pump fields

Figure 5 represents the field maps of pumps and the nonlinear polarization computed from Eq. (1) in both scenarios. The nonlinear polarization in the blue scenario (Fig. 5(a)) is much higher than in the orange scenario (Fig. 5(b)), because the resonances at pump wavelengths leads to higher enhancement of the electric field, and there is a better overlap between the two pump fields. The enhancement of the field can be related to the quality factors [42,43], which are higher in the blue scenario (Q = 96 for TE and Q = 29 for TM in the blue scenario, Q = 14 for TE and Q = 54 for TM in the orange scenario). The effective nonlinear susceptibility can be computed from the following expression [30]:

χ

eff(2)

=

x,z

χ

(2)

E

1

E

2

E

3

dxdz

|E

inc

|

3

dh . (20)

The integral is defined on the cross-section of the GaAs layer. In the blue (resp. orange) scenario, it yields a value χ

eff(2)

= 13 nm/V (resp. χ

eff(2)

= 3 nm/V). Thus, only the blue case DFG scenario is studied hereafter:

1 1.064 µ m

| {z }

TE

− 1

1.6 µ m

| {z }

TM

= 1 3.17 µ m

| {z }

TM

. (21)

Fig. 5. Field maps of pump and idler fields (in one period, with gold represented in gray) and the nonlinear polarization resulting in both cases (P

z(2)

=

0

χ

xyz(2)

E

pump

E

idler

). (a) Blue scenario: the pump wave is TE-polarized at λ

1

= 1.064 µm and the idler wave is TM-polarized at λ

2

= 1.6 µm generate a nonlinear polarization at λ

3

= 3.17 µm (

1.0641

1

1.6

=

3.171

). (b) Orange scenario: the pump wave is TM-polarized at λ

1

= 1.2 µm and the idler wave is TE-polarized at λ

2

= 1.55 µm generate a nonlinear polarization at λ

3

= 5.3 µm (

1.21

1

1.55

=

5.31

).

Scenarios with other wavelengths could be studied that exploit the same resonances, by tuning

the geometrical parameters of the structure. In these cases, a similar nonlinear efficiency is

obtained.

(10)

3.3. Extraction of the nonlinear field

The nonlinear field generated in the layer has to be extracted from the structure through the grating. A structure that is efficiently trapping an incoming wave, is reciprocally able to efficiently extract a generated wave in the layer. This behavior is not fully intuitive, and it can be understood by writing the reflectivity and extraction ratio expressions from S-matrices. The reflectivity and extraction ratio amplitudes are written as:

Reflectivity = S

αAA

+ S

αAB

( I − Q

βα

(ΓQ

αβ

S

αBB

)

−1

Q

βα

ΓQ

βα

S

αBA

), (22) Extraction = ( Reflectivity − S

αAA

) S

αAB1

, (23) where Γ = S

βBB

+ S

βCB

( I − Q

γβ

(S

γCC

PrS

αBB

)

1

Q

γβ

S

βCC

Q

γβ

S

αBA

and Q

ij

is the propagation vector from the interface i to the interface j.

The matrices S

αAA

and S

αAB

are the scattering matrices at the the gold grating interfaces, so are not depending on the GaAs layer thickness h. Thus, the derivative against h of the extraction amplitude is the same than the reflectivity multiplied by the matrix S

αAB−1

. It means that they have the same extrema and that their variations are linked through the S

αAB−1

matrix that depends only on the metallic grating geometry. Besides, when there is no resonance, the reflectivity amplitude is close to S

αAA

, and therefore the extraction ratio is close to zero.

The schematics of S-matrices involved in the reflectivity and in the extraction are shown in Fig. 6. The reflectivity (orange curve) and extraction (purple curve) efficiencies are plotted at

Fig. 6. Scheme of the S-matrices involved in (a) the reflectivity computation and (b) in the

extraction of a nonlinear polarization in the GaAs layer. (c) Absolute values of reflectivity

(orange curve) and extraction efficiency (purple curve) as a function of the GaAs layer height

h at the signal wavelength λ

3

= 3.17 µm.

(11)

the signal wavelength λ

3

= 3.17 µ m as a function of the gallium arsenide layer thickness. Both curves are exhibiting resonances with an alternance of peaks and dips, but they are in phase opposition. The extraction yield is maximum when the reflectivity reaches a minimum value. In the orange scenario, since there is no resonance at λ = 5.3 µ m in the structure, there is a high reflectivity (mirror behavior) and thus a weak extraction. This low extraction ratio, combined to a poorer co-localization, leads to a 30 times smaller efficiency of conversion. The signals generated by other effects (SHG and SFG) present also a low extraction ratio, because the structure does not present resonances in the reflectivity spectrum at their respective signal wavelengths.

3.4. Efficiency

Figure 7 represents the components of the nonlinear electric field generated by the P

z

nonlinear polarization. The generated electric field is mainly enhanced inside the grating layer with hot spots on the sidewalls of the metallic ribbons. The distribution of the electric field is quite different from that of the nonlinear polarization since it accounts for the effects of the resonance in the structure at λ

3

. In fact, this distribution is very close to the one obtained for an incoming wave at λ

3

(data not shown). This resonance behavior is paramount for the extraction and efficiency of the structure, and to evaluate this latter, the efficiency of the DFG process is defined as η

sample

=

PPDFG2

inc

.

Fig. 7. Normalized electric field maps E

x

and E

z

generated by the resonator at λ = 3.17 µm (blue scenario). The generated field is TM-polarized.

For the sake of comparison, the efficiency of the structure is compared to a GaAs layer deposited on a gold mirror without the gold metasurface with an efficiency η

ref

. This layer acts as a cavity, which resonance wavelength is tuned by the thickness of the layer. In order to have the same resonance wavelength as the sample, the height is set at h = 200 nm.

Figure 8 shows the effiencies η

sample

and η

ref

as a function of the pump wavelength λ

2

and with the other pump wavelength set at λ

1

= 1.064 µ m. As expected, the efficiencies are maximum near the resonance wavelength of the structure and are tremendously reduced away from this resonance.

The efficiency of the DFG process can be increased by three orders of magnitude compared to

the reference (η

sample

= 14 × 10

−3

W/W

2

and η

ref

= 35 × 10

−6

W/W

2

at λ

2

= 1.608 µm). This is

one order of magnitude higher than metasurfaces associated to multiple quantum wells, albeit it

was also demonstrated experimentally. To emphasize the key contributions of the colocalization

and the extraction, the maximum efficiency was computed in the orange ratio and in this case

η

sample

= 10

5

W/W

2

at λ

2

= 1.608 µm).

(12)

Fig. 8. Efficiency spectra of the resonator and a reference, as a function of the pump wavelength λ

2

. The reference is composed of a layer of GaAs (thickness h = 200 nm) on a gold mirror. At the pump wavelength λ

2

= 1.6 µm, the resonator is 4000 times more efficient than the reference.

4. Conclusion

In summary, we have demonstrated nonlinear metasurfaces that exhibit a very high efficiency of difference frequency generation in the infrared. By cautiously engineering the Fabry-Perot resonances and the guided mode resonances that takes place in the GaAs waveguide layer, it is possible to give birth to a 20-fold enhanced nonlinear polarization. In order for the signal wavelength to escape the layer where it was originally generated, the resonator has to act as a trap for the incoming linear light, this is achieved by matching an additional resonance. Our work demonstrates the possibility of achieving such high efficiency without using multiple quantum wells included in metasurfaces. This new paradigm can be applied to other nonlinear materials with greatly relaxed phase-matching conditions in order to conceive infrared or THz sources.

Obviously, applying these design rules to quantum heterostructures could also yield even higher efficiencies, as well as increasing the quality factors of the resonances in MIM structures [42], even if it can be at the expense of a more complex fabrication.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. J. A. Schuller, E. S. Barnard, W. Cai, Y. C. Jun, J. S. White, and M. L. Brongersma, “Plasmonics for extreme light concentration and manipulation,” Nat. Mater.

9(3), 193–204 (2010).

2. P. Bouchon, F. Pardo, B. Portier, L. Ferlazzo, P. Ghenuche, G. Dagher, C. Dupuis, N. Bardou, R. Haïdar, and J.-L.

Pelouard, “Total funneling of light in high aspect ratio plasmonic nanoresonators,” Appl. Phys. Lett.

98(19), 191109

(2011).

3. S. Xiao, Q. He, X. Huang, S. Tang, and L. Zhou, “Enhancement of light-matter interactions in slow-wave metasurfaces,”

Phys. Rev. B

85(8), 085125 (2012).

4. M. Kauranen and A. V. Zayats, “Nonlinear plasmonics,” Nat. Photonics

6(11), 737–748 (2012).

5. S. Keren-Zur, L. Michaeli, H. Suchowski, and T. Ellenbogen, “Shaping light with nonlinear metasurfaces,” Adv. Opt.

Photonics

10(1), 309–353 (2018).

6. N. Panoiu, W. Sha, D. Lei, and G. Li, “Nonlinear optics in plasmonic nanostructures,” J. Opt.

20(8), 083001 (2018).

7. A. Krasnok, M. Tymchenko, and A. Alu, “Nonlinear metasurfaces: a paradigm shift in nonlinear optics,” Mater.

Today

21(1), 8–21 (2018).

(13)

8. S. Liu, M. B. Sinclair, S. Saravi, G. A. Keeler, Y. Yang, J. Reno, G. M. Peake, F. Setzpfandt, I. Staude, T. Pertsch, and I. Brener, “Resonantly enhanced second-harmonic generation using iii–v semiconductor all-dielectric metasurfaces,”

Nano Lett.

16(9), 5426–5432 (2016).

9. V. F. Gili, L. Carletti, A. Locatelli, D. Rocco, M. Finazzi, L. Ghirardini, I. Favero, C. Gomez, A. Lemaître, M.

Celebrano, C. D. Angelis, and G. Leo, “Monolithic algaas second-harmonic nanoantennas,” Opt. Express

24(14),

15965–15971 (2016).

10. D. Rocco, V. F. Gili, L. Ghirardini, L. Carletti, I. Favero, A. Locatelli, G. Marino, D. N. Neshev, M. Celebrano, M.

Finazzi, G. Leo, and C. D. Angelis, “Tuning the second-harmonic generation in algaas nanodimers via non-radiative state optimization,” Photonics Res.

6(5), B6–B12 (2018).

11. D. De Ceglia, G. D’Aguanno, N. Mattiucci, M. Vincenti, and M. Scalora, “Enhanced second-harmonic generation from resonant gaas gratings,” Opt. Lett.

36(5), 704–706 (2011).

12. M. Gromovyi, J. Brault, A. Courville, S. Rennesson, F. Semond, G. Feuillet, P. Baldi, P. Boucaud, J.-Y. Duboz, and M. P. De Micheli, “Efficient second harmonic generation in low-loss planar gan waveguides,” Opt. Express

25(19),

23035–23044 (2017).

13. E. Kim, F. Wang, W. Wu, Z. Yu, and Y. R. Shen, “Nonlinear optical spectroscopy of photonic metamaterials,” Phys.

Rev. B

78(11), 113102 (2008).

14. V. Valev, N. Smisdom, A. Silhanek, B. De Clercq, W. Gillijns, M. Ameloot, V. Moshchalkov, and T. Verbiest,

“Plasmonic ratchet wheels: switching circular dichroism by arranging chiral nanostructures,” Nano Lett.

9(11),

3945–3948 (2009).

15. P.-Y. Chen, C. Argyropoulos, and A. Alù, “Enhanced nonlinearities using plasmonic nanoantennas,” Nanophotonics

1(3-4), 221–233 (2012).

16. H. Husu, R. Siikanen, J. Makitalo, J. Lehtolahti, J. Laukkanen, M. Kuittinen, and M. Kauranen, “Metamaterials with tailored nonlinear optical response,” Nano Lett.

12(2), 673–677 (2012).

17. M. Celebrano, X. Wu, M. Baselli, S. Großmann, P. Biagioni, A. Locatelli, C. De Angelis, G. Cerullo, R. Osellame, B.

Hecht, L. Duò, F. Ciccacci, and M. Finazzi, “Mode matching in multiresonant plasmonic nanoantennas for enhanced second harmonic generation,” Nat. Nanotechnol.

10(5), 412–417 (2015).

18. N. Segal, S. Keren-Zur, N. Hendler, and T. Ellenbogen, “Controlling light with metamaterial-based nonlinear photonic crystals,” Nat. Photonics

9(3), 180–184 (2015).

19. S. Tang, D. J. Cho, H. Xu, W. Wu, Y. R. Shen, and L. Zhou, “Nonlinear responses in optical metamaterials: theory and experiment,” Opt. Express

19(19), 18283–18293 (2011).

20. T. Skauli, K. L. Vodopyanov, T. J. Pinguet, A. Schober, O. Levi, L. A. Eyres, M. M. Fejer, J. S. Harris, B. Gerard, L.

Becouarn, E. Lallier, and G. Arisholm, “Measurement of the nonlinear coefficient of orientation-patterned gaas and demonstration of highly efficient second-harmonic generation,” Opt. Lett.

27(8), 628–630 (2002).

21. W. Fan, S. Zhang, K. Malloy, S. Brueck, N. Panoiu, and R. Osgood, “Second harmonic generation from patterned gaas inside a subwavelength metallic hole array,” Opt. Express

14(21), 9570–9575 (2006).

22. M. Scalora, M. Vincenti, D. De Ceglia, V. Roppo, M. Centini, N. Akozbek, and M. Bloemer, “Second-and third-harmonic generation in metal-based structures,” Phys. Rev. A

82(4), 043828 (2010).

23. S. Héron, P. Bouchon, and R. Haïdar, “Mode matching in second-order-susceptibility metal-dielectric structures,”

Phys. Rev. A

94(3), 033831 (2016).

24. F. Niesler, N. Feth, S. Linden, J. Niegemann, J. Gieseler, K. Busch, and M. Wegener, “Second-harmonic generation from split-ring resonators on a gaas substrate,” Opt. Lett.

34(13), 1997–1999 (2009).

25. P. P. Vabishchevich, S. Liu, M. B. Sinclair, G. A. Keeler, G. M. Peake, and I. Brener, “Enhanced second-harmonic generation using broken symmetry iii–v semiconductor fano metasurfaces,” ACS Photonics

5(5), 1685–1690 (2018).

26. A. Feng, Z. Yu, and X. Sun, “Giant enhancement of nonlinear optical processes with split-ring resonators for thz applications,” IEEE Photonics Technol. Lett.

31(21), 1681–1684 (2019).

27. E. Rosencher, A. Fiore, B. Vinter, V. Berger, P. Bois, and J. Nagle, “Quantum engineering of optical nonlinearities,”

Science

271(5246), 168–173 (1996).

28. J. Lee, M. Tymchenko, C. Argyropoulos, P.-Y. Chen, F. Lu, F. Demmerle, G. Boehm, M.-C. Amann, A. Alu, and M.

A. Belkin, “Giant nonlinear response from plasmonic metasurfaces coupled to intersubband transitions,” Nature

511(7507), 65–69 (2014).

29. S. Campione, A. Benz, M. B. Sinclair, F. Capolino, and I. Brener, “Second harmonic generation from metamaterials strongly coupled to intersubband transitions in quantum wells,” Appl. Phys. Lett.

104(13), 131104 (2014).

30. Y. Liu, J. Lee, S. March, N. Nookala, D. Palaferri, J. F. Klem, S. R. Bank, I. Brener, and M. A. Belkin, “Difference- frequency generation in polaritonic intersubband nonlinear metasurfaces,” Adv. Opt. Mater.

6(20), 1800681 (2018).

31. M. Tymchenko, J. Gomez-Diaz, J. Lee, M. Belkin, and A. Alù, “Highly-efficient thz generation using nonlinear plasmonic metasurfaces,” J. Opt.

19(10), 104001 (2017).

32. E. D. Palik,

Handbook of optical constants of solids, vol. 3 (Academic, 1998).

33. P. Bouchon, F. Pardo, R. Haïdar, and J.-L. Pelouard, “Fast modal method for subwavelength gratings based on b-spline formulation,” J. Opt. Soc. Am. A

27(4), 696–702 (2010).

34. L. Li and C. Haggans, “Convergence of the coupled-wave method for metallic lamellar diffraction gratings,” J. Opt.

Soc. Am. A

10(6), 1184–1189 (1993).

35. L. Li, “Formulation and comparison of two recursive matrix algorithms for modeling layered diffraction gratings,” J.

Opt. Soc. Am. A

13(5), 1024–1035 (1996).

(14)

36. S. Héron, F. Pardo, P. Bouchon, J.-L. Pelouard, and R. Haïdar, “Modal method for second harmonic generation in nanostructures,” J. Opt. Soc. Am. B

32(2), 275–280 (2015).

37. F. X. Wang, F. J. Rodríguez, W. M. Albers, R. Ahorinta, J. Sipe, and M. Kauranen, “Surface and bulk contributions to the second-order nonlinear optical response of a gold film,” Phys. Rev. B

80(23), 233402 (2009).

38. C. Wei, S. Liu, D. Deng, J. Shen, J. Shao, and Z. Fan, “Electric field enhancement in guided-mode resonance filters,”

Opt. Lett.

31(9), 1223–1225 (2006).

39. C. Koechlin, P. Bouchon, F. Pardo, J. Jaeck, X. Lafosse, J. Pelouard, and R. Haïdar, “Photon sorting in plasmonic multiresonators,” Appl. Phys. Lett.

99(24), 241104 (2011).

40. M. Makhsiyan, “Nano-émetteurs thermiques multi-spectraux,” Ph.D. thesis, Paris Saclay (2017).

41. E. Sakat, G. Vincent, P. Ghenuche, N. Bardou, C. Dupuis, S. Collin, F. Pardo, R. Haïdar, and J.-L. Pelouard, “Free- standing guided-mode resonance band-pass filters: from 1d to 2d structures,” Opt. Express

20(12), 13082–13090

(2012).

42. C. Qu, S. Ma, J. Hao, M. Qiu, X. Li, S. Xiao, Z. Miao, N. Dai, Q. He, S. Sun, and L. Zhou, “Tailor the functionalities of metasurfaces based on a complete phase diagram,” Phys. Rev. Lett.

115(23), 235503 (2015).

43. L. Cong, P. Pitchappa, C. Lee, and R. Singh, “Active phase transition via loss engineering in a terahertz mems

metamaterial,” Adv. Mater.

29(26), 1700733 (2017).

Références

Documents relatifs

In this section we study the structure of the space of tangent vector fields TVect(M ) viewed as a CVect(M )-module.. 4.1 A

À la suite de notre recherche nous avons pu déterminer qu’une intervention précoce, comprenant des manipulations, des facilitations ainsi que des positionnements, chez les

CHF 58 million over 6 years, the Strategy relied on three priority domains: (i) Rural Income and Em- ployment through the creation of market and job opportunities for the rural

In the Falck case, the far-sighted family champion of change Alberto Falck—with crucial support of the external CEO Achille Colombo—was able to de-escalate the family business

Une voiture abandonnée par son occupant, dévale les flots avec une célérité impressionnante. Que devient le propriétaire de ce véhicule ? Que sont devenus les

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

l’utilisation d’un remède autre que le médicament, le mélange de miel et citron était le remède le plus utilisé, ce remède était efficace dans 75% des cas, le

L’hépatite aigue fulminante a été initialement définie comme la survenue brutale d’une insuffisance hépatocellulaire sévère à l’origine d’un ictère et d’une