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

Optical field and attractive force at the subwavelength slit

Open Access Open Access

Abstract

In recent works, a novel light-induced attractive force was predicted between two metal plates. This force arises by the interaction of surface plasmons which are excited at the metal when a transverse magnetic mode propagates through a subwavelength slit between two metal bodies. In this paper, the analytical and numerical calculations of this magnetic field are presented for the perfect metal and for gold. The amplitude and the phase transient curves between the known limiting cases of narrow and wide slits compared to the wavelength are found. The curve is shown to oscillate due to the emergence of new waveguide modes. The analytic solution for the perfect metal is in agreement with the computation for gold by means of the finite element method. The simple asymptotic formula for the light-induced attractive force is found in the limit of a narrow slit.

© 2016 Optical Society of America

1. Introduction

Light-induced forces are of great interest not only for research, but also for modern technologies, e.g. in order to develop manipulation techniques for macro-, micro- and nano-objects. In laser tweezers [1–4], these forces are used to displace dielectric particles. As for metal particles, these are moved by optical tweezers if their size is smaller than 100 nm. It is, however, practically impossible to handle metal particles having a size of more than 100 nanometers. In recent works [5, 6], a light-induced force was predicted in a subwavelength slit between two metal planes. In contrast to the pressure of light, the force described is attractive if the magnetic field is parallel to the planes. This is due to the interaction of surface plasmons which are excited at the conducting surfaces. This force could be exploited for metal particle manipulation and is promising for laser microswitches.

For real metals with their complex permittivity ε = ε1 + 2, the light-induced force for a slit with the dimension L in the z-direction (see Fig. 1) is given by

F=μ0H02|H(0,0)|2L4πλ|ε1|ε2,
where H0 is the magnetic field strength amplitude of the incident wave, H(0, 0) is the normalized magnetic field at the coordinate origin, μ0 is the magnetic permeability of vacuum, and λ is the wavelength of the incident wave. To determine the magnetic field strength amplitude in the center at the entrance of the subwavelength slit, it is necessary to solve the Maxwell equations under the boundary conditions at the conductor surfaces. Maxwell’s equations were solved in the slit geometry in connection with the problem of the extraordinary optical transmission through a single slit [7, 8]. However, the entering field has not been studied.

 figure: Fig. 1

Fig. 1 Scheme of the slit geometry. The yellow areas denote the perfect metal.

Download Full Size | PDF

The aim of this paper is to improve the theoretical model and to study the width dependence of the entering field. We normalize the incident wave by unity. If the width of slit 2l approaches zero, the field can be found by using the Fresnel formulas. In this limiting case, the amplitude at the metal surface equals 2, since the field is formed by two waves of equal amplitudes, the incident one and the reflected one. If the width is much greater than the wavelength 2lλ, we pass into geometric optics [9] and the entering amplitude approaches 1, as in free space. For experimental purposes, the intermediate case 2lλ is of interest. An explicit analytical expression for the estimation of the light-induced attractive force in the limiting case of a narrow slit (a = 2πl/λ ≪ 1) is derived.

This calculation is carried out below. In Section 2, the equations are solved for the slit in the perfect metal. We find the solution in the form of the Fourier series. In Section 3, the set for Fourier coefficients is solved and explicit analytical formulas are derived for a narrow slit. Section 4 describes the finite element method computation for the gold film covering the slit surface. Section 5 summarizes the conclusions.

2. Analytical solution

Following paper [8], we examine the perfect metallic half-space with a slit having the half-width l. Figure 1 shows the cross section of the metal and the slit. The normal incidence of a wave running upwards is indicated by the arrow at the bottom. We choose the y-axis normal to the half-space boundary, the x-axis normal to the slit walls, the z-axis normal to the xy-plane, and the magnetic field vector H⃗ along the z-axis. In this geometry, the Maxwell equations for a monochromatic wave are reduced to the 2D Helmholtz equation for the magnetic field amplitude Hz = H(x, y)eiωt:

2Hx2+2Hy2+k02H=0,
where k0 = ω/c is the wavenumber, ω is the frequency, and c is the speed of light in vacuum.

The boundary conditions for the chosen polarization are reduced to continuity of the tangential magnetic field and continuity of the normal electric displacement vector field. The conditions at the perfect conductor surface are:

H|x=±(l0)=H|x=±(l+0),Hx|x=±l=0,y>0;H|y=0=H|y=+0,Hy|y=0=0,x2>l2.
At y > 0, we seek the solution of the form h(x)eiβy, where β is the propagation constant in the slit. For function h(x), the ordinary differential equation is valid:
h+(k02β2)h=0.
The first boundary condition (3) yields the slit eigen-modes x2 < l2, y > 0. For normal incidence, only the even modes hm(x) = cos(πmx/l), m = 0, 1,... are excited with
βm2=k02π2m2/l2.
In a sufficiently narrow slit with l < π/k0 = λ/2, only the zeroth mode propagates. If the slit is wider, the propagation turns to the multimode regime. The number m of each mode coincides with the number of nodes within the slit half-width 0 < x < l.

The solution at y > 0 can be found as a Fourier series, whereas at y < 0, it is a Fourier integral:

H(x,y)={2cosk0y+akeikxiϰkydk,y<0,m=0cmbmhm(x)eiβmy,y>0,
where cm = 2 − δm0, ϰk=k02k2. The coefficients ak, bm in the discrete and continuous spectra are found from the second boundary condition (3). Excluding ak, we find the following infinite set for bm:
bm+m=0Tmmbm=2δm,0,
Tmm=lBmcm4πfm,kfm,kϰkdk,fm,k=sinc(kl+πm)+sinc(klπm).
We first find the Fourier coefficients bm from set (7), and then the field (6) as a series:
H(0,0)=b0+2m=1bm.

3. Asymptotic formulas

At k0l ≪ 1, the higher spatial modes are almost negligible: |bm| ≪ |b0|, m ≠ 0. Set (7) reduces to one equation, which yields

H(0,0)=b0=21+T00.
In the limiting case of a narrow slit k0l ≪ 1, the field (10) only depends on the element T00.

As the width increases, the next modes emerge. Equation (5) shows that value βm2 is positive for mode numbers mk0l/π, and with increasing l, the number of propagating modes increases. The rest of the modes m > k0l/π has the imaginary propagation constant. They are then evanescent and decay along the y-axis. In particular near the threshold of the first mode 2l = λ, set (7) reduces as follows to a pair of equations as follows:

b0+T00b0+T01b1=2,b1+T10b0+T11b1=0,
hence
H(0,0)=21+T112T101+T00+T11+T00T11T01T10.

Figure 2 shows the square |H(0, 0)|2 within the zeroth (10) and the first (12) approximations as functions of the dimensionless width a = k0l. The solid curve and the dots coincide in the subwavelength region; then, the zeroth approximation is valid at a ≪ 1. The first approximation works up to aπ. It is clear that a breaking point (the dip with a jump of the first derivative) is formed in the curve at the first mode threshold, whereas the zeroth approximation is a smooth function. The explicit expression (10) with the matrix element T00 involves integral (8). Since it still works well for the small parameter a, we may replace it by a simpler expansion in the parameter:

T00=TiT,Taa3/6+,Tβ1a+β2a3+,β1=0.590.64lna,β2=0.11+0.11lna.
In Fig. 2, the asymptotic expression shown by crosses at small a almost coincides with Eq. (10). In the case of a ≪ 1, equation (1) takes an especially simple analytical form. Using (10) and (13), the expression for the light-induced attractive force (1) can be described by the following equation:
F=μ0H02(12a)Lπλ|ε1|ε2,a=2πλl1.

The solving of a truncated set confirms the behavior of the curve for a narrow slit. Figure 3(a) shows the absolute value of the field at the coordinate origin as a function of a for m, m′M = 50. At k0l > 1, the numerical solution has many saw-tooth-like oscillations. Each breaking point corresponds to the birth of a new mode at k0l = πm, m = 1, 2,... In the phase characteristics in Fig. 3(b), we see similar oscillations in the saw-tooth curve with a slow damping.

 figure: Fig. 2

Fig. 2 Square of the normalized field amplitude |H(0, 0)|2 as a function of the dimensionless slit width a = k0l: (a) calculated by formula (10) (solid curve), by (12) (dots), and by the simple asymptotic expression (13) at small a (crosses).

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 The square of absolute value (a) and phase (b) of the normalized field H(0,0) as a function of the dimensionless slit half-width a = k0l: the perfect metal at M = 50 (solid line) and JCMsuite simulations for gold (dots).

Download Full Size | PDF

4. Numerical modeling

The calculation for real metal by eigen-mode expansion is more complicated since it includes the numerical solution of a transcendental equation for the slit modes. In addition, as shown in paper [10], the zeroth mode vanishes when one takes the damping in real metal into consideration. The damping influences the field, as well as the surface plasmon excited at the edges of the slit [11]. The plasmon excitation is taken into account by the adaptive solver JCMsuite [12]. This program solves the Maxwell equations in a defined calculation area by means of the finite element method. A special feature of the problem is the semi-infinite domain (see Fig. 1). JCMsuite offers, however, the possibility to truncate the computational domain by applying virtual, perfectly matched absorbing layers [13]. With these boundary conditions implemented, the calculation area is free of feedback from the rim. Moreover, the area can be kept sufficiently small.

The thickness of the gold layer covering the silicon substrate is 100 nm in order to practically prevent the penetration of the electromagnetic wave. The gold film at the silicon substrate is a set-up to be employed in the forthcoming experiment. The refractive index of the gold used is n = 0.3226 + 10.6200i at the wavelength λ = 1.55 μm [14]. A number of quality laser sources has been developed for telecommunication carrier wave then this wavelength is chosen. The magnetic field and the associated phase of the wave are calculated numerically as a function of parameter a. The results are shown in Fig. 3. As the plot shows, the result for gold is in good agreement with the eigen-mode calculation for the perfect metal.

5. Conclusions

The field has been studied at the entrance to the slit between two parallel metal planes as a function of the slit width. At zero and infinite slit width, the field amplitude is well known: 2 and 1, respectively, whilst the incident wave is normalized by unity. The problem is to calculate the transient curves between these limiting cases. The entering field is obtained as a series (9); its coefficients are found by the solving of set (7), where the matrix elements are integrals (8). The number of terms of the series which needs to be considered is approximately equal to the number of modes propagating trough the slit.

While the frequency is less than the critical value for the slit considered as a planar waveguide, i.e. in the subwavelength regime, the field variation is monotonic and smooth. In the case of a wider slit, the next modes must be accounted for. The amplitude and the phase characteristics are found. Both curves turn out to be non-monotonic and involve saw-tooth-like oscillations. Each bend of the curves corresponds to a new mode excitation. The finite element method was applied for the slit covered with gold taking into account the Joule energy losses. The computation is in agreement with the perfect metal solution. The simple analytical expression for estimation of the light-induced attractive force in the limiting case of a narrow slit (a = 2πl/λ ≪ 1) has been derived.

Acknowledgments

This work is supported by the Deutsche Forschungsgemeinschaft (DFG NE 1550/2-1 and DFG NE 1550/2-2) and by the Russian Russian foundation for Basic Research (# 16-52-12026).

References and links

1. S. Chu, “Nobel lecture: The manipulation of neutral particles,” Rev. Mod. Phys. 70, 685–706 (1998). [CrossRef]  

2. Y. Liu and M. Yu, “Investigation of inclined dual-fiber optical tweezers for 3d manipulation and force sensing,” Opt. Express 17, 13624–13638 (2009). [CrossRef]   [PubMed]  

3. C. Pacoret, R. Bowman, G. Gibson, S. Haliyo, D. Carberry, A. Bergander, S. Régnier, and M. Padgett, “Touching the microworld with force-feedback optical tweezers,” Opt. Express 17, 10259–10264 (2009). [CrossRef]   [PubMed]  

4. Y. Gong, A.-Y. Ye, Y. Wu, Y.-J. Rao, Y. Yao, and S. Xiao, “Graded-index fiber tip optical tweezers: Numerical simulation and trapping experiment,” Opt. Express 21, 16181–16190 (2013). [CrossRef]   [PubMed]  

5. V. Nesterov, L. Frumin, and E. Podivilov, “Negative light pressure force between two metal bodies separated by a subwavelength slit,” EPL (Europhysics Letters) 94, 64002 (2011). [CrossRef]  

6. V. Nesterov and L. Frumin, “Light-induced attractive force between two metal bodies separated by a subwavelength slit,” Measurement Science and Technology 22, 094008 (2011). [CrossRef]  

7. R. Gordon, “Light in a subwavelength slit in a metal: Propagation and reflection,” Phys. Rev. B 73, 153405 (2006). [CrossRef]  

8. B. Sturman, E. Podivilov, and M. Gorkunov, “Transmission and diffraction properties of a narrow slit in a perfect metal,” Physical Review B 82, 115419 (2010). [CrossRef]  

9. J. Bravo-Abad, L. Martín-Moreno, and F. J. García-Vidal, “Transmission properties of a single metallic slit: From the subwavelength regime to the geometrical-optics limit,” Phys. Rev. E 69, 026601 (2004). [CrossRef]  

10. B. Sturman, E. Podivilov, and M. Gorkunov, “Optical properties of periodic arrays of subwavelength slits in a perfect metal,” Physical Review B 84, 205439 (2011). [CrossRef]  

11. L. Aigouy, P. Lalanne, J. P. Hugonin, G. Julié, V. Mathet, and M. Mortier, “Near-field analysis of surface waves launched at nanoslit apertures,” Phys. Rev. Lett. 98, 153902 (2007). [CrossRef]   [PubMed]  

12. S. Burger, L. Zschiedrich, J. Pomplun, and F. Schmidt, “JCMsuite: An adaptive FEM solver or precise simulations in nano-optics,” in “Integrated Photonics and Nanophotonics Research and Applications,” (Optical Society of America, 2008), p. ITuE4. [CrossRef]  

13. J.-P. Berenger, “A perfectly matched layer for the absorption of electromagnetic waves,” Journal of Computational Physics 114, 185–200 (1994). [CrossRef]  

14. R. L. Olmon, B. Slovick, T. W. Johnson, D. Shelton, S.-H. Oh, G. D. Boreman, and M. B. Raschke, “Optical dielectric function of gold,” Phys. Rev. B 86, 235147 (2012). [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 (3)

Fig. 1
Fig. 1 Scheme of the slit geometry. The yellow areas denote the perfect metal.
Fig. 2
Fig. 2 Square of the normalized field amplitude |H(0, 0)|2 as a function of the dimensionless slit width a = k0l: (a) calculated by formula (10) (solid curve), by (12) (dots), and by the simple asymptotic expression (13) at small a (crosses).
Fig. 3
Fig. 3 The square of absolute value (a) and phase (b) of the normalized field H(0,0) as a function of the dimensionless slit half-width a = k0l: the perfect metal at M = 50 (solid line) and JCMsuite simulations for gold (dots).

Equations (14)

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

F = μ 0 H 0 2 | H ( 0 , 0 ) | 2 L 4 π λ | ε 1 | ε 2 ,
2 H x 2 + 2 H y 2 + k 0 2 H = 0 ,
H | x = ± ( l 0 ) = H | x = ± ( l + 0 ) , H x | x = ± l = 0 , y > 0 ; H | y = 0 = H | y = + 0 , H y | y = 0 = 0 , x 2 > l 2 .
h + ( k 0 2 β 2 ) h = 0 .
β m 2 = k 0 2 π 2 m 2 / l 2 .
H ( x , y ) = { 2 cos k 0 y + a k e i k x i ϰ k y d k , y < 0 , m = 0 c m b m h m ( x ) e i β m y , y > 0 ,
b m + m = 0 T m m b m = 2 δ m , 0 ,
T m m = l B m c m 4 π f m , k f m , k ϰ k d k , f m , k = sinc ( k l + π m ) + sinc ( k l π m ) .
H ( 0 , 0 ) = b 0 + 2 m = 1 b m .
H ( 0 , 0 ) = b 0 = 2 1 + T 00 .
b 0 + T 00 b 0 + T 01 b 1 = 2 , b 1 + T 10 b 0 + T 11 b 1 = 0 ,
H ( 0 , 0 ) = 2 1 + T 11 2 T 10 1 + T 00 + T 11 + T 00 T 11 T 01 T 10 .
T 00 = T i T , T a a 3 / 6 + , T β 1 a + β 2 a 3 + , β 1 = 0.59 0.64 ln a , β 2 = 0.11 + 0.11 ln a .
F = μ 0 H 0 2 ( 1 2 a ) L π λ | ε 1 | ε 2 , a = 2 π λ l 1 .
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.