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

Nonlinearity of surface-plasmon polaritons in sub-wavelength metal nanowires

Open Access Open Access

Abstract

We investigate the nonlinear propagation of surface plasmon polaritons guided on gold nanowires surrounded by silica glass. Based on the Lorentz reciprocity theorem, we derive a formula for the complex nonlinear susceptibility, and study its dependence on waveguide parameters and wavelength for the fundamental mode. Depending on these parameters both positive and negative signs of the real and imaginary parts of the nonlinear coefficient are predicted. This implies that nanowires exhibit the property of saturable absorption or optical limiting as well as positive and negative nonlinear phase shifts. The physical origin of this phenomenon is discussed.

© 2016 Optical Society of America

1. Introduction

Metal nanosystems play a central role in the emerging field of nanooptics and plasmonics and have been the subject of extensive research, both fundamental and with view to applications. In particular, the need for compact and high-performance optical devices has driven the interest in plasmonic waveguides which can be used to carry optical signals to different parts of subwavelength components of plasmonic optoelectronic circuids (see e.g. [1]). In metal nanowaveguides the nonlinear susceptibility is significantly enhanced (see e.g. [2–4]) and are related with peculiarities of intraband and interband nonlinear electronic processes in metal under the influence of electromagnetic fields [5]. In the last decades, different types of waveguides has been theoretically investigted, such as nonlinear plasmonic planar waveguides [6–9], metal nanowires [10, 11], plasmonic slot waveguides [12] and periodic plasmonic waveguides [13]. In such sub-wavelength waveguides TM modes exhibits a significant longitudinal component of the electric field, which plays an important role in nonlinear propagation (see e.g. [14]).

Among the different types of plasmonic waveguides metal nanowires have the most simplest waveguide structure and are of great interest due to their potential in several fields. Subwavelength waveguiding of plasmons in such structures has been observed in a number of experiments [15–18]. The nonlinear optical behaviour of crystalline metal nanowires has been experimetally studied in several papers (see e.g. [19–21]). The experimetal results in these papers indicate that depending on parameters crystalline metal nanowires exhibit both the property of saturable absorption and of optical limiting. Similar nonlinear effects has also been found in metal nanoparticle composites (see e.g. [22,23] and references therein). In a theoretical study of this phenomenon in [24] it was shown that the physical origin of the saturation effect is related to the intrinsic Kerr nonlinearity of the metal nanoparticles and the linear metal loss leading to a shift of the plasmon resonance and therefore to a reduction of loss with increasing intensity. The saturable loss in metal nanocomposites can be used for modelocking and Q-switching of lasers as proposed and theoretically studied in [25] and experimentally realized by several groups (see e.g. [26] and refereces therein). Although the experimental observations for metal nanowires [19–21] were done in samples of crystaline structures, one can expect that analogous phenomena occure also in single metal nanowires. Note that in previous theoretical studies of the nolinear propagation in single metal nanowires [10, 11] only positive values of the real and imaginary part of the nonlinear susceptibility were predicted.

In this paper we reexamine the problem of nonlinear propagation in plasmonic waveguides and use an alternative method for the derivation of the nonlinear propagation equation for the slowly varying field amplitude of plasmon-polaritons in metal nanowaveguides. In our approach we use the Lorentz reciprocity theorem (see e.g. [13, 27, 28] and references therein) for the derivation of the nonlinear coefficient and study its dependence on the parameters of the incident field and of the waveguide for the case of metal-air interfaces and metal nanowires. The main result of our work is the prediction that the sign of the real and imaginary part of the effective nonlinear coefficient of gold nanowires at visible wavelengths can be shifted from positive to negative values by changing wire radii and wavelengths. This means in dependence on the parameters our predictions describe both the property of saturable absorption and optical limiting as well a positive or negative nonlinear phaseshifts. A physical explanation of this phenomenon is given.

2. Derivation of the nonlinear coefficient using the reciprocity theorem

We use the mode expansion for the electric and the magnetic fields F=(E;H) in the plasmonic waveguide which can be expressed as F(r,t)=(1/2)[F(r)exp(iωt)+c.c], where F(r) is a time-independent amplitude for the mode m. In the following we restrict ourself to the fundamental mode m = 0 which can be presented by

E(r)=1/s0Ψ(z)exp(iκz)e0(r),
H(r)=(iωμ0)1×E=1/s0Ψ(z)exp(iκz)h0(r).

Here z and are the coordinate and the unit vector in the direction of propagation, r is the position vector in the transverse plane, s0 is defined as s0=(1/2)Re(e0×h0)z^dσ, where the integral is performed in the transverse plane, β = κ + iα/2 is the SPP propagation constant, Ψ is normalized so that |Ψ|2 is equal to the power flow along the z direction P(z)=(1/2)Re(E×H)z^dσ|Ψ|2, and e0(r) and h0(r) describe spatial transverse distribution of the fundamental mode.

The Lorentz reciprocity theorem states the following relation for two electromagnetic fields (E1,H1) and (E2,H2) which are solutions of the Maxwell equations corresponding to the relative permittivity ε1(r) and ε2(r), respectively, [13, 27, 28]:

z[E1(r)×H2(r)E2(r)×H1(r)]z^dσ=ik0Z0[ε2(r)ε1(r)]E1(r)E2(r)dσ.

We substitute for (E1,H1) and (E2,H2) the unperturbed backward propagating field (e,h) and the perturbed forward propagating field (E,H), correspondingly. In the first order of perturbation, Eq. (3) leads to the following equation describing the amplitude Ψ(z)

dΨdz=α2Ψ+ik0δneffΨ,
and the expression for the change of the effective refractive index:
δneff=δε(e022e0z2)dσ2Z0(e0×h0)z^dσ.

Here we use the relations e0x=e0x, e0y=e0y, e0z=e0z, h0x=h0x, h0y=h0y, h0z=h0z.

3. Nonlinear metal-air interface

To benchmark the derived formula (5), we apply it to the simple analytically solvable case of a single gold-air interface, for which (ε = εgold ·θ (−x) + εair · θ (x)) for δε = δε1 · θ (−x) + δε2 · θ (x), where θ (x) is the Heaviside step function. For the calculations we used the·measured wavelength-dependent permittivity of gold [29]. Using the formula of effective refractive index neff=ε1ε2/(ε1+ε2) for a single interface and substracting from the perturbed refractive index the unperturbed one, we can obtain the following analytical formula of δneff:

δneff=(ε1+δε1)(ε2+δε2)(ε1+δε1)+(ε2+δε2)ε1ε2ε1+ε2.

On the other hand the expression for δneff derived in [6, 14] is

δneff=δε|e0|2dσ2Z0Re(e0×h0*)z^dσ,
where Z0=μ0/ε0 is the wave impedance in vacuum. We calculate for the case of gold δneff for δε1 = 0.01, δε2 = 0 [Fig. 1(a)]. As shown in Fig. 1(a), the result calculated by the derived formula (5) coincides with that by using the analytical formula (6) in the whole spectral region as should be expected. As shown by the red crosses, it agrees also with results in [6, 14] at wavelengths larger than 600 nm, but shows significant deviations at visible wavelengths smaler than 550 nm. In particular in this spectral range the sign of the real part of δneff, given by Eqs. (5) and (6), is changed, while Eq. (7) predicts only positive values. In addition, Eq. (5) predict also an imaginary part of δneff produced by a real perturbation δε. A phase shift of the SPP mode e0(r) induced by the large loss of gold Imε/Reε gives an additional phase to δneff which is taken into account in formula (5).

 figure: Fig. 1

Fig. 1 Response of effective refractive index in a single gold-air interface to a perturbation of the permittivity in gold δε1 = 0.01, δε2 = 0. The solid curve, blue circles and red crosses represent the results by Eq. (5), the analytical formula Eq. (6), and of Eq. (7), respectively.

Download Full Size | PDF

In Fig. 1(a) we have studied the influence of a perturbation of the dielectric function on the effective refraction index without specifying the physical nature of this perturbation. To investigate an intrinsic nonlinear response of SPPs in plasmonic waveguides, we now consider a Kerr-type nonlinear contribution to the permittivity εnl=(3/4)χ(3)(ω;ω,ω,ω)|E|2 [30, 31]. Equation (5) then leads to the nonlinear propagation equation

dΨdz=α2Ψ+iγ|Ψ|2Ψ,
with
γ=k0(3/4)χ(3)|e0|2(e022e0z2)dσZ0(e0×h0)z^dσRe(e0×h0)z^dσ,
where γ is the effective nonlinear coefficient and χ(3) is the third-order nonlinear susceptibility of the corresponding material.

We apply the Eqs. (8) and (9) to a single gold-air interface. Since the intrinsic nonlinearity of gold is much larger than that of the dielectric, we neglect the nonlinear susceptibility of the dielectric and use the wavelength-depending values of the complex nonlinear susceptibility of gold χgold(3) from [11]. In Fig. 2(a) the nolinear phaseshift is presented at the wavelength of 535 nm. We can see that a negative real part of γ leads to a negative nonlinear phase shift [see Fig. 2(a)] in spite of a positive real part of χgold(3). In Fig. 2(b) the nonlinear contribution to the effctive loss of the interface is presented. As can be seen the negative imaginary part of γ for the wavelength of 502 nm leads to a negative nonlinear nonlinear absorption cefficient or to saturable loss in spite of a positive imaginary part of χgold(3). Our results are in good agreement with results of a full-dimensional simulation which were obtained by numerically solving Maxwell equations in the frequency domain [32] (blue circles in Fig. 2).

 figure: Fig. 2

Fig. 2 Power-dependence of the nonlinear phase shift for the wavelength of 535 nm and a propagation distance of 1000 nm (a) and nonlinear absorption for the wavelength of 502 nm and a propagation distance of 500 nm (b) for a single gold-air interface. The black line is the results obtained by Eq. (5), the red crosses by Eq. (7) and the blue circles correspond to a full-dimensional simulation.

Download Full Size | PDF

Let us compare the prediction provided by (9) with a result of [7] using a Green-function formalism. We obtained γ = (1.16 + i1.00) × 10−7 W−1m−1 for the single gold-air interface at λ = 796 nm using the measured value of χ(3)gold = (4.67 + i3.03) × 10−19 m2/V2 [33], which is in a good agreement with the result γ = (1.03 + i0.98) × 10−7 W−1m−1 in [7].

4. Nonlinear metal nanowires

Let us now study the effective nonlinear response of gold nanowires surrounded by silica glass. We consider a silica glass host, since this material was used in previous investigations of nanowires (e.g. [11]). The standard Bessel-function formalism for linear modes in cylindrically symmetric structures is used, with wavenumber found numerically. Here we focus on the TM mode (m = 0) because the fundamental TM mode shows the strongest nonlinear response and a well-confined spatial distribution [11]. We used the Sellmeier expansion for the permittivity of silica glass [31], which is included in the abovementioned numerical calculation. The non-linearity of silica glass is disregarded because it is several orders of magnitude smaller than that of gold. As shown in Fig. 3 for two different wire radii, in the visible region a large nonlinear coefficient |γ| > 105 W−1m−1 is attained. This is attributed to the large nonlinear susceptibility of gold in the visible region and comparably strong confinement and enhancement of the field. Note that both Reγ and Imγ changes the sign in the visible region, while in the infrared region it has only the positive sign. This means depending on frequency gold nanowires can be used for positive or negative nonlinear self-phase modulation.

 figure: Fig. 3

Fig. 3 Effective nonlinear coefficient of a gold nanowires with a wire radii 200 nm (a) and 50 nm (b) surrounded by silica glass for the fundamental TM mode (m=0). The red solid and blue dashed line shows the real and imaginary part of the effective nonlinear coefficient, respectively.

Download Full Size | PDF

In Fig. 4 the dependence of Re γ on the wire radii is plotted. As can be seen Re γ has a negative sign in the range of wire radii beetween about 30 nm and 50 nm. This gives the possibility to design the required properties of nonlinear device by simply choosing the radius of the plasmonic waveguides, and possibly to applications such as for switching or quasi-phase-matching. Fig. 5 shows that Im γ also change the sign at two values of wire radii at a fixed wavelength. This implies that gold nanowires in the visible region would exhibit saturated absorption or reverse-saturated absorption for different wire radii. Note that the dependence of the sign of the effective nonlinear coefficients on the waveguide radius is a rather surprising phenomenon because the linear and nonlinear properties of the component materials do not depend on the wire radius.

 figure: Fig. 4

Fig. 4 Real part of effective nonlinear coefficient of gold nanowires surrounded by silica glass for the fundamental TM mode (m=0) at the wavelength of 720 nm versus the wire radius. The third-order nonlinear susceptibility of gold χgold(3)=(2+i15)×1017m2/V2 from [11] is used.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Imaginary part of effective nonlinear coefficient of gold nanowires surrounded by silica glass for the fundamental TM mode (m=0) at the wavelength of 630 nm versus the wire radius. The third-order nonlinear susceptibility of gold χgold(3)=(20i5)×1017m2/V2 from [11] is used.

Download Full Size | PDF

The high intrinsic nonlinearity of SPP is naturally accompanied with a high loss of the SPP. In order to define the performance of a lossy nonlinear waveguide, we introduce a figure of merit F = |γ|Lspp, where Lspp = 1 is the SPP propagation length. The input power needed for a certain nonlinear effect is inversely proportional to this figure of merit. In Fig. 6 one can see that F exhibits a maximum which varies in position from 610 nm for 30-nm nanowires to 750 nm for 200-nm nanowires and can be positioned at either side of the resonance. The lower limit of all curves is determined by the mode cutoff. The maximum of the nonlinearity as well as its sign change is attributed to an additional phase of (E22Ez2)/|E2|, which is related to the gold loss. The accumulated additional phase is influenced by two factors, local additional phase distribution and field enhancement distribution, as one can see in Eq. (9). With increasing the radius, the local additional phase in the central part increases [Fig. 7(a)], but the local field magnitude in the central part decreases [Fig. 7(b)]. Thus the accumulated additional phase is maximum for a certain radius.

 figure: Fig. 6

Fig. 6 SPP propagation length Lspp (a) and figure of merit F = |γ| Lspp (b) of gold nanowires with wire radii 30nm (black solid), 50 nm (red dashed) and 200 nm (blue crosses) surrounded by silica glass for the fundamental TM mode (m=0).

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Additional phase ϕ and field enhancement |E|4/|E04 due to metal loss versus r in the gold nanowire with different radii R = 30, 50, 80 nm. The wavelength is 630 nm and χgold(3)=(20i5)×1017m2/V2 from [11] is used.

Download Full Size | PDF

Let us discuss the physical reason for the above decribed sign change of both the real and imaginary part of the nonlinear susceptibility of the metal nanowire. We remind a similar phenomenon of saturated absorption in metal nanocomposites where the sign change is the result of the fact that near the plasmon resonance the local field enhancement factor x = Ein/E0 has an imaginary component that arises by the linear loss of the metal nanoparticles and leads to a phase difference between the field inside the nanoparticles Ein and the external field E0 [24, 34]. A simple explanation of the sign change of the nonlinearity and of saturable loss can be provided by the dependence of the field enhancement factor for spherical nanoparticles on wavelengths, radius and intensity. This means that the sign change of the nonlinear susceptibility is a result of both the high metal nonlinearity and the high metal loss. For the nonlinear metal nanowires a qualitatively analogous effect arises. The sensitive dependence of the sign of the effective nonlinear susceptipility of the nanowire is related with the dependence of the plasmon resonance on the waveguide radius.

5. Conclusion

In summary, by using the Lorentz reciprocity theorem we reconsidered the problem to model the nonlinear response of SPPs in plasmonic waveguides and derived a formula for the effective complex SPP nonlinear susceptibility in these waveguides. For the case of gold nanowires surrounded by silica glass we calculated the nonlinear coeficient in dependence on the wavelength and the wire radius. We predicted a large nonlinear coefficient (|γ| > 105 W−1m−1) and have shown that the sign of the real and imaginary part of the nonlinear susceptibility depend on the wavelength and can also be changed for a fixed wavelength by changing the wire radius. This opens the possibility to design the nonlinear properties of metal nanowires by the choice of the nanowire radius which can realize both positive and negative phase shifts by self-phase modulation as well as saturable loss or optical limiting which could be used in future ultracompact photonic devices. We discussed the physical origin of these effects.

References and links

1. E. Ozbay, “Plasmonics: Merging Photonics and Electronics at Nanoscale Dimensions,” Science 311, 189–193 (2006). [CrossRef]   [PubMed]  

2. M. Kauranen and A. V. Zayats, “Nonlinear plasmonics,” Nature Photon. 6, 737–748 (2012) [CrossRef]  

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

4. 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, 193–204 (2010). [CrossRef]   [PubMed]  

5. F. Hache, D. Ricard, C. Flytzanis, and U. Kreibig, “The optical kerr effect in small metal particles and metal colloids: the case of gold,” Appl. Phys. 47, 347–357 (1988).

6. D. V. Skryabin, A. Gorbach, and A. Marini, “Surface-induced nonlinearity enhancement of TM modes in planar subwavelength waveguides,” J. Opt. Soc. Am. B 28, 109–114 (2011). [CrossRef]  

7. I. De Leon, J. E. Sipe, and R. W. Boyd, “Self-phase-modulation of surface plasmon polaritons,” Phys. Rev. A 89, 013855 (2014). [CrossRef]  

8. A. Baron, S. Larouche, D. J. Gauthier, and D. R. Smith, “Scaling of the nonlinear response of the surface plasmon polariton at a metal/dielectric interface,” J. Opt. Soc. Am. B 32, 9–14 (2014). [CrossRef]  

9. A. Baron, T. B. Hoang, C. Fang, M. H. Mikkelsen, and D. R. Smith, “Ultrafast self-action of surface-plasmon polaritons at an air/metal interface,” Phys. Rev. B 91, 195412 (2015). [CrossRef]  

10. A. Marini, R. Hartley, A. V. Gorbach, and D. V. Skryabin, “Surface-induced nonlinearity enhancement in sub-wavelength rod waveguides,” Phys. Rev. A 84, 063839 (2011). [CrossRef]  

11. A. Marini, M. Conforti, G. D. Valle, H. W. Lee, W. C. Tr, X. Tran, M. A. Schmidt, S. Longhi, P. S. J. Russell, and F. Biancalana, “Ultrafast nonlinear dynamics of surface plasmon polaritons in gold nanowires due to the intrinsic nonlinearity of metals,” New J. of Phys. 15, 013033 (2013). [CrossRef]  

12. A. R. Davoyan, I. V. Shadrivov, and Y. S. Kivshar, “Nonlinear plasmonic slot waveguides,” Opt. Express 26, 21209–21214 (2008). [CrossRef]  

13. A. A. Sukhorukov, A. S. Solntsev, S. S. Kruk, D. N. Neshev, and Yuri S. Kivshar, “Nonlinear coupled-mode theory for periodic plasmonic waveguides and metamaterials with loss and gain,” Opt. Lett. 39, 462–465 (2014). [CrossRef]   [PubMed]  

14. S. V. Afshar and T. M. Monro, “A full vectorial model for pulse propagation in emerging waveguides with subwavelength structures part I: Kerr nonlinearity,” Opt. Express 17, 2298–2318 (2009). [CrossRef]  

15. R. M. Dickson and L. Andrew Lyon, “Unidirectional plasmon propagation in metallic nanowires,” J. Phys. Chem. B 104, 6095–6098 (2000). [CrossRef]  

16. J. R. Krenn, B. Lamprecht, H. Ditlbacher, G. Schider, M. Salerno, A. Leitner, and F. R. Aussenegg, “Non-diffraction-limited light transport by gold nanowires,” Europhys. Lett. 60, 663–665 (2002). [CrossRef]  

17. H. Ditlbacher, A. Hohenau, D. Wagner, U. Kreibig, M. Rogers, F. Hofer, F. R. Aussenegg, and J. R. Krenn, “Silver nanowires as surface plasmon resonators,” Phys. Rev. Lett. 95, 257403 (2005). [CrossRef]   [PubMed]  

18. A. W. Sanders, D. A. Routenberg, B. J. Wiley, Y. Xia, E. R. Dufresne, and M. A. Reed, “Observation of plasmon propagation, redirection, and fan-Out in silver nanowires,” Nano Letters 6, 1822–1826 (2006). [CrossRef]   [PubMed]  

19. C. Zheng, X. Y. Ye, S. G. Cai, M. J. Wang, and X. Q. Xiao, “Observation of nonlinear saturable and reverse-saturable absorption in silver nanowires and their silica gel glass composite,” Appl. Phys. B 101, 835–840 (2010). [CrossRef]  

20. Y. P. Han, J. L. Sun, H. A. Ye, W. Z. Wu, and G. Shi, “Nonlinear refraction of silver nanowires from nanosecond to femtosecond laser excitation,” Appl. Phys. B 94, 233–239 (2009). [CrossRef]  

21. R. Udabyabhaskar, M. S. Ollakkan, and B. Karthikeyan, “Preparation, optical and non-linear optical power limiting properties of Cu, CuNi nanowires,” Appl. Phys. Lett. 104, 013107 (2014). [CrossRef]  

22. R. A. Ganeev, I. A. Ryasnyanskiy, A. L. Stepanov, and T. Usmanov, “Saturated absorption and nonlinear refraction of silicate glasses doped with silver nanoparticles at 532 nm,” Opt. and Quantum Electr. 36, 949–960 (2004). [CrossRef]  

23. U. Gurudas, E. Brooks, D. M. Bubb, S. Heiroth, T. Lippert, and A. Wokaun, “Saturable and reverse saturable absorption in silver nanodots at 532 nm using picosecond laser pulses,” J. Appl. Phys. 104, 073107 (2008). [CrossRef]  

24. K.-H. Kim, A. Husakou, and J. Herrmann, “Saturable absorption in composites doped with metal nanoparticles,” Opt. Express 18, 21918–21925 (2010). [CrossRef]   [PubMed]  

25. K.-H. Kim, U. Griebner, and J. Herrmann, “Theory of passive mode-locking of semiconductor disk lasers in the blue spectral range by metal nanocomposites,” Opt. Express 20, 16174–16179 (2012). [CrossRef]  

26. D. Wu, Jian Peng, Z. Cai, J. Weng, Z. Luo, N. Chen, and H. Xu, “Gold nanoparticles as a saturable absorber for visible 635 nm Q-switched pulse generation,” Opt. Express 23, 24071–24076 (2015). [CrossRef]   [PubMed]  

27. A.W. Snyder and J. D. Love, Optical Waveguide Theory (Chapman and Hall, 1983).

28. D. Michaelis, U. Peschel, C. Walter, and A. Bräuer, “Reciprocity theorem and perturbation theory for photonic crystal waveguides,” Phys. Rev. E 68, 065601 (2003). [CrossRef]  

29. P. B. Johnson and R. W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6, 4370–4379 (1972). [CrossRef]  

30. R. W. Boyd, Nonlinear Optics, 3rd ed (Academic, 2008).

31. G. P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic, 2007).

32. S.-J. Im and G.-S. Ho, “Plasmonic amplification and suppression in nanowaveguide coupled to gain-assisted high- quality plasmon resonances,” Laser Phys. Lett. 12, 045902 (2015). [CrossRef]  

33. I. De Leon, Z. Shi, A. C. Liapis, and R. W. Boyd, “Measurement of the complex nonlinear optical response of a surface plasmon-polariton,” Opt. Lett. 39, 8, 2274–2276 (2014). [CrossRef]   [PubMed]  

34. D. D. Smith, G. Fischer, R. W. Boyd, and D. A. Gregory, “Cancellation of photoinduced absorption in metal nanoparticle composites through a counterintuitive consequence of local field effects,” J. Opt. Soc. Am. B 14, 1625–1631 (1997). [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 (7)

Fig. 1
Fig. 1 Response of effective refractive index in a single gold-air interface to a perturbation of the permittivity in gold δε1 = 0.01, δε2 = 0. The solid curve, blue circles and red crosses represent the results by Eq. (5), the analytical formula Eq. (6), and of Eq. (7), respectively.
Fig. 2
Fig. 2 Power-dependence of the nonlinear phase shift for the wavelength of 535 nm and a propagation distance of 1000 nm (a) and nonlinear absorption for the wavelength of 502 nm and a propagation distance of 500 nm (b) for a single gold-air interface. The black line is the results obtained by Eq. (5), the red crosses by Eq. (7) and the blue circles correspond to a full-dimensional simulation.
Fig. 3
Fig. 3 Effective nonlinear coefficient of a gold nanowires with a wire radii 200 nm (a) and 50 nm (b) surrounded by silica glass for the fundamental TM mode (m=0). The red solid and blue dashed line shows the real and imaginary part of the effective nonlinear coefficient, respectively.
Fig. 4
Fig. 4 Real part of effective nonlinear coefficient of gold nanowires surrounded by silica glass for the fundamental TM mode (m=0) at the wavelength of 720 nm versus the wire radius. The third-order nonlinear susceptibility of gold χ gold ( 3 ) = ( 2 + i 15 ) × 10 17 m 2 / V 2 from [11] is used.
Fig. 5
Fig. 5 Imaginary part of effective nonlinear coefficient of gold nanowires surrounded by silica glass for the fundamental TM mode (m=0) at the wavelength of 630 nm versus the wire radius. The third-order nonlinear susceptibility of gold χ gold ( 3 ) = ( 20 i 5 ) × 10 17 m 2 / V 2 from [11] is used.
Fig. 6
Fig. 6 SPP propagation length Lspp (a) and figure of merit F = |γ| Lspp (b) of gold nanowires with wire radii 30nm (black solid), 50 nm (red dashed) and 200 nm (blue crosses) surrounded by silica glass for the fundamental TM mode (m=0).
Fig. 7
Fig. 7 Additional phase ϕ and field enhancement |E|4/|E04 due to metal loss versus r in the gold nanowire with different radii R = 30, 50, 80 nm. The wavelength is 630 nm and χ gold ( 3 ) = ( 20 i 5 ) × 10 17 m 2 / V 2 from [11] is used.

Equations (9)

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

E ( r ) = 1 / s 0 Ψ ( z ) exp ( i κ z ) e 0 ( r ) ,
H ( r ) = ( i ω μ 0 ) 1 × E = 1 / s 0 Ψ ( z ) exp ( i κ z ) h 0 ( r ) .
z [ E 1 ( r ) × H 2 ( r ) E 2 ( r ) × H 1 ( r ) ] z ^ d σ = i k 0 Z 0 [ ε 2 ( r ) ε 1 ( r ) ] E 1 ( r ) E 2 ( r ) d σ .
d Ψ d z = α 2 Ψ + i k 0 δ n eff Ψ ,
δ n eff = δ ε ( e 0 2 2 e 0 z 2 ) d σ 2 Z 0 ( e 0 × h 0 ) z ^ d σ .
δ n eff = ( ε 1 + δ ε 1 ) ( ε 2 + δ ε 2 ) ( ε 1 + δ ε 1 ) + ( ε 2 + δ ε 2 ) ε 1 ε 2 ε 1 + ε 2 .
δ n eff = δ ε | e 0 | 2 d σ 2 Z 0 Re ( e 0 × h 0 * ) z ^ d σ ,
d Ψ d z = α 2 Ψ + i γ | Ψ | 2 Ψ ,
γ = k 0 ( 3 / 4 ) χ ( 3 ) | e 0 | 2 ( e 0 2 2 e 0 z 2 ) d σ Z 0 ( e 0 × h 0 ) z ^ d σ Re ( e 0 × h 0 ) z ^ d σ ,
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.