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

Partial coherence and polarization of a two-mode surface-plasmon polariton field at a metallic nanoslab

Open Access Open Access

Abstract

Rigorous electromagnetic theory is utilized to characterize the partial spatial coherence and partial polarization of a two-mode field consisting of the long-range and the short-range surface-plasmon polariton at a metallic nanofilm. By employing appropriate formulations for the spectral degrees of coherence and polarization, we examine the fundamental limits for these quantities associated with such a superposition field and explore how the degrees are influenced when the media, frequency, and slab thickness are varied. It is in particular shown that coherence lengths extending from subwavelength scales up to thousands of wavelengths are possible and their physical origins are elucidated. In addition, we demonstrate that for ultra-thin films the generally highly polarized two-mode field can be partially polarized in close vicinity of the polariton excitation region. The results could benefit cross-disciplinary electro-optical applications in which near-field interactions between plasmons and nanoparticles are exploited.

© 2015 Optical Society of America

1. Introduction

Plasmonics [1–3 ], based on the utilization of electromagnetic evanescent near fields coupled to collective electron oscillations at a metal-dielectric interface, known as surface-plasmon polaritons (SPPs) [4, 5], has in recent years attracted extensive research activity within the realms of nanophotonics [6–8 ]. Due to their unique features, such as strong enhancement and localization of light energy within subwavelength domains, SPPs occupy a pivotal position in a wide range of optoelectronic and nanophotonic applications. Besides the rapid technological progress, research concerning the fundamental nature of SPPs has very recently suggested the existence of entirely new SPP classes, both at single boundaries [9] and at metallic films [10, 11], which further broadens the possibilities for novel polaritonic devices.

While a single boundary allows only one plasmon mode, the SPPs supported by the individual surfaces of a thin metal film can couple and form two bounded modes with symmetric or antisymmetric field profiles and different dispersion characteristics [10, 12–14 ]. In particular, as the thickness of the film becomes sufficiently small, the symmetric mode may acquire a propagation length which is several hundreds of times greater than that of the single-interface SPP [15, 16], whereupon it is frequently called the ‘long-range surface-plasmon polariton’ (LRSPP) [17]. The antisymmetric mode, on the other hand, becomes increasingly confined as the slab thickness decreases, resulting in a much smaller surface propagation distance than those of the LRSPP and the respective single-interface SPP, and therefore the antisymmetric mode is termed the ‘short-range surface-plasmon polariton’ (SRSPP) [16]. Due to its unique capability of long-range guidance the LRSPP has over the years received considerably more attention and practical significance than the SRSPP. However, several phenomena have been reported in which the SRSPP plays a crucial role, including plasmonic focusing [18], stimulated-amplification supported SPP propagation [19], extraordinarily low transmission through nanopatterned films [20], and plasmon-waveguide sensing [21]. Methods which allow efficient excitation of the SRSPP, either simultaneously with [22] or without [23] the LRSPP, have also been presented.

The electromagnetic coherence, the partial polarization, and the spectrum of random evanescent near fields, with or without plasmon excitations, can differ substantially from those associated with the far field. It has been demonstrated that subwavelength coherence lengths, as well as polarization states not allowed by the traditional beam-field formalism, may occur in purely evanescent waves generated at an interface between two lossless dielectric media [24]. Likewise, the correlation length in a fluctuating thermal near field can be significantly shorter than the wavelength of light, or it may extend over several tens of wavelengths under SPP excitation [25, 26]. The broadband near field radiated by a thermal half-space source may become essentially quasi-monochromatic [27], and highly polarized [28], when surface plasmons are involved. Owing to SPP interactions, bodies in thermal equilibrium may emit spectrally dependent, directional beams of radiation if gratings are fabricated on their surface [29]. These results illustrate that depending on the specific circumstances, the presence of surface plasmons may significantly alter, not just the intensity and the spectrum, but also the spatial coherence and polarization characteristics of the electromagnetic field near the surface.

Motivated by the conclusion above, in this work we analyze the spatial coherence and polarization properties of a two-mode surface-plasmon polariton field consisting of the LRSPP and the SRSPP at a metallic nanofilm. These are the two modes supported by a thin metal slab, and the SRSPP has a pivotal role as regards the coherence and polarization modifications. More specifically, on the basis of exact formulations of SPP modes and appropriate spectral electromagnetic coherence theory, we investigate the fundamental ranges in which the degrees of spatial coherence and polarization may vary in practice in such a two-mode configuration, regardless of the excitation method. The degrees of coherence and polarization are the basic measures of any partially coherent and partially polarized field, but no studies employing the rigorous definitions of these quantities, which take into account the complete vector-valued nature of the electromagnetic field, have been carried out for SPPs. We introduce a local and a global coherence length and show that the coherence may range from subwavelength scales up to hundreds or thousands of wavelengths depending on the choice of the media, frequency, and slab thickness. Both coherence lengths, describing different features of the SPP field, as we explain later, can be measured by use of leakage radiation microscopy [30]. Physically, the coherence and polarization states influence the interaction of SPPs with the surrounding and collections of nanoparticles located in close proximity of the metal slab. For example, SPPs propagating on the surfaces of a metallic nanolayer can form a highly sensitive interferometric biosensor [31]. Likewise, a coherent plasmon field may excite a random set of molecules to radiate coherently from different emitters [32]. Our results further indicate that the two-mode SPP field is generally highly polarized, but can be partially polarized in and near the plasmon excitation region. Nanoparticle scattering is known to depend on the polarization properties of the field, to the extent that the field’s polarization state can be fully deduced from measurements of the scattered radiation [33].

2. Long-range and short-range SPPs

We consider SPPs, represented as monochromatic realizations of electromagnetic plane waves of angular frequency ω, at two planar interfaces lying in the xy plane between linear, homogeneous, isotropic, non-magnetic, stationary, and spatially non-dispersive media. The geometry, as illustrated in Fig. 1, includes an absorptive metal slab (region |z| < d/2) characterized by a complex (ω-dependent) relative permittivity ε r1 = ε′ r1 + iε″ r1 (the prime ′ and double prime ″ denote real and imaginary parts, respectively), with ε′ r1 < 0 and ε″ r1 > 0, surrounded on both sides (regions |z| ≥ d/2) by a lossless dielectric medium having a real and positive permittivity ε r2. We consider, as is customary, a system with two (transversally counter-propagating) waves within and one on each side of the film.

 figure: Fig. 1

Fig. 1 Illustration of the directions of phase-front propagation (black arrows) and field attenuation (solid-red curves) for SPPs at a lossy metal film of thickness d in a non-absorptive dielectric surrounding, possessing relative permittivities ε r1 (complex) and ε r2 (real), respectively. The boundaries of the slab are at z = ±d/2.

Download Full Size | PDF

The spatial part of the p-polarized electric field outside the slab can be written as

E2(±)(r,ω)=E2eikxx|k2|{e+ik2zz(k2ze^x+kxe^z),zd/2,eik2zz(±k2ze^x±kxe^z),zd/2,
while inside the film one has [10]
E1(±)(r,ω)=E1ei(kxx+k1zz)|k1|[k1z(1e2ik1zz)e^x+kx(1±e2ik1zz)e^z],|z|<d/2,
where Eα with α ∈ {1, 2} is a complex (ω-dependent) field amplitude, |kα|=[kα*kα]1/2 (the asterisk denotes complex conjugation) is the magnitude of the complex (ω-dependent) wave vector k α = kαx ê x + kαz ê z, and ê x and ê z are the unit vectors in the positive x and z directions, respectively. The tangential wave-vector component is continuous across the boundaries, i.e., kαx = kx = k′x + ik″x, whereas the normal component kαz = k′αz + ik″αz satisfies kαz=(k02εrαkx2)1/2 (k 0 is the free-space wave number). In this work we examine only such fields which outside the slab are non-radiative (k′ 2z ≤ 0) and bound to the slab interfaces (k″ 2z > 0), and whose phase-propagation and amplitude-attenuation directions along the x axis are positive (k′x, k″x > 0).

The superscript (±) in Eqs. (1) and (2), where the + (−) sign corresponds to the upper (lower) signs inside the parentheses, identifies two different kinds of modes, Eα(+)(r,ω) and Eα()(r,ω), representing electric fields whose normal components, or magnetic fields, are symmetric and antisymmetric with respect to z = 0, respectively. It follows from Maxwell’s equations and the boundary conditions that Eqs. (1) and (2) have to satisfy [4, 5, 7, 10, 11, 13, 14, 16]

Symmetric(+):εr1εr2k2zk1z=tanh(12ik1zd),
Antisymmetric():εr1εr2k2zk1z=coth(12ik1zd),
which are transcendental equations for kx and generally require numerical methods to solve. It can be verified, however, that when d → 0, Eq. (3) has a solution for which [10, 11, 14, 16]
kxk0εrz,k1zk0εr1εr2,k2z0,
corresponding to the LRSPP. Equation (4), on the other hand, representing the SRSPP, has no such analogous solution [10] but instead the wave-vector components of the antisymmetric mode diverge for d → 0.

It should be emphasized, though, that in fact both Eqs. (3) and (4) stand for an infinite number of modes for any chosen media, frequency, and film thickness [10, 11, 14]. All of these modes can be divided into two main classes: fundamental modes (FMs) and higher-order modes (HOMs). The former are the two familiar ones (one symmetric and one antisymmetric) which become degenerate and reduce to the respective single-interface SPP as the film thickness gets large, while the latter instead are infinite in number and have no correspondence at a single boundary. Traditionally only the FMs have acquired attention and physical importance, since normally they possess the largest surface propagation lengths (the symmetric FM evolves into the LRSPP as d becomes small), whereas the HOMs have been rejected due to their extremely small propagation range [14]. We remark, however, that the existence of HOMs which assume the function of long-range guidance (together with even enhanced field confinement), in circumstances where the propagation distances of the FMs and the single-interface SPP are negligible, has been suggested very recently [11]. Nevertheless, in this work we analyze the FMs and refer to [10, 11, 14] for a more elaborate discussion concerning the HOMs.

3. Electromagnetic coherence and polarization in the space-frequency domain

All the information of second-order spatial coherence of a random electromagnetic field is in the space-frequency domain included in the cross-spectral density matrices [34]. For a stationary (vector) field, the elements of the electric cross-spectral density matrix, at a pair of points r 1 and r 2 (and at frequency ω), can be expressed as a correlation function over an ensemble of strictly monochromatic field realizations [35], i.e.,

W(r1,r2,ω)=E(r1,ω)E(r2,ω).
Here the angle brackets and the dagger stand for ensemble averaging and conjugate transposing, respectively. At a single point, the cross-spectral density matrix reduces to the spectral polarization matrix,
Φ(r,ω)W(r,r,ω),
which contains all the information about the polarization state of the fluctuating electric field.

The strength of correlations that exist between the orthogonal components of the electric field at two points is described by the (electromagnetic) degree of coherence for vectorial fields [36], which is defined as [37]

μ(r1,r2,ω)W(r1,r2,ω)FtrΦ(r1,ω)trΦ(r2,ω),
where ‖ · ‖F is the Frobenius norm and tr denotes the trace. Physically, for beam-like fields, μ(r 1, r 2, ω) characterizes the modulation of the four Stokes parameters in Young’s interference experiment, i.e., not only the intensity variation but also the polarization-state modulation [38, 39], and is therefore a generalization of the traditional degree of coherence for scalar fields (which only involves the visibility of intensity fringes). Analogously to its scalar counterpart it is bounded as [36, 37]
0μ(r1,r2,ω)1,
where the upper and lower limits, respectively, correspond to complete spatial correlation and lack of correlation between the Cartesian field components, while the intermediate values stand for partially coherent fields. In the following we will only be interested in the degree of coherence for vectorial fields (SPPs are vector valued), so henceforth the term degree of coherence is used to refer to the degree of coherence for vectorial fields defined in Eq. (8).

The degree of polarization is a measure of the correlations that prevail between the orthogonal electric-field components at a point. In general, an electric field is composed of three orthogonal components and the concept of the degree of polarization must take into account such three-dimensional fields with wave fronts of arbitrary form [40, 41]. Nevertheless, since the electric field of the SPP modes only has two components, the traditional two-dimensional (2D) treatment of the polarization state is formally sufficient in this analysis. In our case, however, the degree of polarization is defined in the plane parallel to the wave vector (xz plane, where the modes are generally elliptically polarized) and thus the situation differs entirely from that of beam-like fields for which the electric field is perpendicular to the wave vector.

Mathematically, the (2D) degree of polarization is defined as [34]

P(r,ω)2trΦ2(r,ω)tr2Φ(r,ω)1,
and it has a physical interpretation as the ratio of the intensity of the polarized constituent of the field to the total intensity. In analogy to Eq. (22), the degree of polarization obeys
0P(r,ω)1.
For a fully unpolarized field the intensities of the two components are equal and no correlations exist between them, leading to P(r, ω) = 0. The highest value, P(r, ω) = 1, corresponds to the case when the field is completely polarized and this occurs if, and only if, the components are fully correlated. Any other value of P(r, ω) represents partial polarization.

4. Degree of coherence in the LRSPP-SRSPP superposition

As explained in Sec. 2, the slab supports both the long-range and the short-range SPPs, which individually are fully spatially coherent and completely polarized. We therefore consider a superposition of these two modes, which allows for the possibility of partial coherence and partial polarization. The two modes can be excited, for example, by employing the well-known Otto [42] or Kretschmann [43] configurations based on attenuated total reflection, or, alternatively, by end-fire coupling [12, 14] (see also [22, 23]). We will not examine in this work the fields inside the slab any further, but focus on the electric field outside the film. So, in order to keep the notation simpler, henceforward the subscript 2 in the field amplitudes, wave vectors, etc., referring to medium 2 (surrounding) is dropped off. In this case, by defining the unit polarization vector ≡ |k|−1(−kz ê x + kx ê z), and according to Eq. (1), for the region above the slab the cross-spectral density matrix given by Eq. (6) reads as

W(r1,r2,ω)=|E(+)|2ei[k(+)r2k(+)*r1]p^(+)p^(+)+|E()|2ei[k()r2k()*r1]p^()p^()+E(+)*E()ei[k()r2k(+)*r1]p^(+)p^()+E()*E(+)ei[k(+)r2k()*r1]p^()p^(+),
where
p^(a)p^(b)=1|k(a)||k(b)|(kz(a)*kz(b)kz(a)*kx(b)kx(a)*kz(b)kx(a)*kx(b)),a,b{+,}.
The superscripts (+) and (−) refer to the symmetric LRSPP and antisymmetric SRSPP (cf. Sec. 2), respectively.

Equation (12) is general since the mutual correlation between the two modes can be anything. In a practical situation the mode correlation depends on the excitation process. From now on, however, the LRSPP and SRSPP are taken to be mutually uncorrelated, i.e., 〈E (+)* E (−)〉 = 〈E (−)* E (+)〉 = 0. This choice is expected to give the fundamental lower boundary for the coherence degree of such a superposition, since any mutual correlation between the constituents leads to a more coherent field. Furthermore, for the same reason, the modes are considered to have equal intensities at r 1 = x 1 ê x +(d/2)ê z which we define as the excitation point. With these decisions, and by using Eq. (12), the degree of coherence given by Eq. (8) becomes

μ(r1,r2,ω)=μ(Δr,ω)=121+κcos(ΔkΔr)cos(ΔkΔr),
where Δrr 2r 1 is the separation between the observation and the excitation points,
κ[p^(+)*p^()]2=|kx(+)kx()*+kz(+)kz()*|2|k(+)|2|k()|2,
and
Δk[kx(+)kx()]e^x+[kz(+)kz()]e^z,
Δk[kx(+)kx()]e^x+[kz(+)kz()]e^z.
From Eqs. (3) and (4) one numerically solves the components of the wave vectors k (+) and k (−), and thereby one can verify that κ given by Eq. (15) satisfies
1/2κ1,
where the lower and upper limits correspond to d → 0 and d → ∞, respectively. We note that while k (+)′ ·k (+)″ = k (−)′ ·k (−)″ = 0 [10], Eqs. (16) and (17) imply that generally Δk′ ·Δk″ ≠ 0 (cf. Sec. 4.1 below).

4.1. Spatial behavior

The cosine and the hyperbolic cosine in Eq. (14) indicate, respectively, that generally the degree of coherence both oscillates and decays [nevertheless, there exists one particular direction for which μr, ω) is exclusively oscillating and another one for which μr, ω) is solely decaying, namely, perpendicularly to Δk″ and perpendicularly to Δk′, respectively]. The oscillation originates from the fact that at certain periodic distances the two-mode surface-plasmon polariton field is electromagnetically similar [44] to the total field at the excitation point (we neglect the decay of the two modes for the moment). For these positions the field shows a high degree of coherence [however, as the modes have different polarization states, μr, ω) < 1] even though the two modes are mutually uncorrelated and hence do not interfere. Thus the situation in this context is completely different from the conventional (wave-field) beating of two modes excited coherently. The effect is instead akin to the customary oscillatory behavior of the (scalar) degree of coherence for two uncorrelated modes in a gas laser [45]. The decaying behavior of the degree of coherence, which arises from the decay of the two modes, implies that, as long as Δr ⊥̸ Δk″, for a given slab thickness

μ(Δr,ω){1/2,|Δr|,(1+κ)/2,|Δr|0.
The upper value in Eq. (19), which we label as μ , is valid regardless of the material parameters, the frequency of the field, or the thickness of the slab. This is a consequence of the different decay rates of the two modes: for a sufficiently large |Δr|, the mode with the lower decay rate (LRSPP) dominates the superposition and the mode with a higher decay rate (SRSPP) can be neglected. Thus, at large distances, the field can practically be considered as a single LRSPP. For a small |Δr|, on the other hand, the amplitudes of both modes are strong, whereupon the lower value in Eq. (19) depends on ε r1, ε r2, ω, and d according to Eqs. (3), (4), and (15). Furthermore, the lower value is in fact the maximum for the degree of coherence (cf. Sec. 4.3).

As an example, Fig. 2 illustrates the spatial dependence of μr, ω) given in Eq. (14) for Ag films of two different thicknesses in vacuum. We observe from Fig. 2 that, as discussed above, the degree of coherence shows an oscillatory behavior, and both the amplitude and the period become smaller as d decreases. In addition μr, ω) decays, starting from its maximum at |Δr| = 0, towards the constant value μ when |Δr| increases, in accordance with Eq. (19), and the decay becomes stronger as the film gets thinner. Further one notices the differences along the x and z axes: there is no oscillation of μr, ω) perpendicular to the interface, in contrast to the situation along the surface, and the decay is notably stronger (around 30 times) in the z direction than in the x direction. The reason for the absence of the oscillatory behavior perpendicularly to the slab is that Δk″ 2z ≫ Δk′ 2z, making the degree of coherence ‘overdamped’, and the decay of μr, ω) in the z direction is much stronger than along the x axis because Δk″ 2z ≫ Δk″x [cf. Eqs. (14) and (17)]. To illustrate further the directions in which μr, ω) in Fig. 2 is purely oscillating and purely decaying, as pointed out in the paragraph above, we show in Fig. 3 the corresponding density plots.

 figure: Fig. 2

Fig. 2 Degree of coherence μr, ω) for the LRSPP-SRSPP superposition above a 50 nm (left) and 20 nm (right) thick Ag slab in vacuum, as a function of the longitudinal and transverse distances Δx/λ 0 and Δz/λ 0, respectively, at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point Δx = Δz = 0. The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Density plots corresponding to the 3D graphs of μr, ω) in Fig. 2. The dashed and dotted lines illustrate the directions in which the degree of coherence is solely oscillating and solely decaying, determined by Δk″ and Δk′, respectively. The dotted lines are located at the positions of local coherence minima. The vectors are drawn with correct relative magnitudes in each panel (but not among the panels).

Download Full Size | PDF

4.2. Slab-thickness dependence

By keeping Δr fixed (0 < |Δr| < ∞), it can be verified from Eqs. (14) and (18) that

μ(Δr,ω){1,d,1/2,d0,
where this time both limits are independent of material parameters and frequency. The physical origin behind the upper value in Eq. (20), representing full coherence, is that for large film thicknesses the solutions of the symmetric and antisymmetric modes approach each other [10], resulting in a fully coherent field consisting of two modes with the same degenerate wave vectors. The lower value, corresponding to partial coherence, is a consequence of the diverging decay rate of the SRSPP as d decreases towards zero, whereupon the field contains only the LRSPP for any |r| > 0 [however, for |Δr| = 0 the SRSPP is always present, no matter how small the slab thickness becomes, and therefore, according to Eqs. (18) and (19), in this specific case μ(Δr,ω)3/2 when d → 0].

Figure 4 illustrates the d-dependent behavior of μr, ω) given in Eq. (14) along the x and z axes for a Ag slab in vacuum. Figure 4 confirms the features given by Eq. (20): for a given |Δr|, the degree of coherence starts from 1 when d is large and approaches 1/2 when the slab thickness goes to zero. Furthermore, the larger |Δr| is the sooner μ(Δr,ω)1/2 as d decreases. This behavior has its origin in the significantly higher decay rate of the SRSPP than of the LRSPP at small film thicknesses. For a sufficiently small |Δr|, on the other hand, the amplitude of the SRSPP is strong even though d is small and hence also the SRSPP will contribute to the field’s coherence properties at that |Δr|. Figure 4 further shows that along the z axis the degree of coherence decays monotonically from 1 to 1/2 as d decreases, implying that a thinner slab results in a less coherent field for a given Δz, while in the x direction μr, ω) oscillates within a particular Δx-dependent slab-thickness region before reaching 1/2. When Δx decreases the oscillation region shifts towards smaller film thicknesses and the separation between two nearby extrema (which is not constant for a fixed Δx) becomes smaller.

 figure: Fig. 4

Fig. 4 Degree of coherence μr, ω) for the LRSPP-SRSPP field along the x (left) and z (right) axes above Ag slabs of thickness d in vacuum at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point Δx = Δz = 0. The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].

Download Full Size | PDF

4.3. Maxima and minima

The oscillation of μr, ω) implies that the field may show a high or a low degree of coherence at certain locations. Equation (14) indicates that such local maxima and minima can be found for points where Δk′ ·Δr = 2 and Δk′ ·Δr = (2m+1)π, respectively, with m being an integer. The collection of these specific points, which we denote as Δrmax(m) and Δrmin(m), form an infinite number of straight lines in the xz plane (cf. Fig. 2) which are separated by π/|Δk′| [cf. Eq. (24) in Sec. 5]. The respective extrema, μmax(m) and μmin(m), generally depend on m and the direction of Δrmax(m) and Δrmin(m). In particular, the case Δr ⊥ Δk″, whereby cosh(Δk″ · Δr) = 1, is the most interesting, because, according to Eq. (14), it gives the (m-independent) global extrema,

μmax=(1+κ)/2,μmin=(1κ)/2,
and thus sets the limits for the domain in which μr, ω) is restricted.

Equation (21) shows that μmax2+μmin2=1, indicating that an increase of one is accompanied with a decrease of the other. Furthermore, from Eqs. (18) and (21) one finds that the global extrema are bounded as

3/2μmax1,0μmin1/2,
where the lower and upper limits of μ max (μ min) correspond to d → 0 (d → ∞) and d → ∞ (d → 0), respectively. Equation (22) particularly demonstrates that, irrespective of the chosen materials, frequency, and slab thickness, there are always regions for which the total field displays a quite high or a rather low degree of coherence.

In Fig. 5 we have plotted the d-dependent behavior of μ max and μ min for a Ag slab at different wavelengths (left panel) and in various surroundings (right panel). Concerning the right panel in Fig. 5, we see that, practically, neither μ max nor μ min is affected at all when ε r2 is varied. Neither is μ max in the left panel affected significantly as λ 0 is altered, whereas μ min slightly increases when the wavelength is reduced. Furthermore, in both panels the maxima are close to unity for d ≳ 50 nm and smoothly decrease towards 3/2 when d → 0, while the minima approach 1/2 as the film thickness goes to zero and 0 for d → ∞, in accordance with Eq. (22).

 figure: Fig. 5

Fig. 5 Maxima μ max and minima μ min of the degree of coherence for the LRSPP-SRSPP superposition as a function of the film thickness d for a Ag slab in vacuum at various free-space wavelengths λ 0 (left) and in different surroundings at λ 0 = 632.8 nm (right). The solid-blue, dashed-green, and dotted-red lines in the left (right) panel correspond to λ 0 = 400 nm (vacuum), λ 0 = 550 nm (SiO2), and λ 0 = 700 nm (ZnO), respectively. The modes are mutually uncorrelated and have the same intensities at the excitation point. The relative permittivity for Ag is ε r1 = −3.77 + i0.67 (λ 0 = 400 nm), ε r1 = −11.13 + i0.83 (λ 0 = 550 nm), ε r1 = −15.87+ i1.07 (λ 0 = 632.8 nm), and ε r1 = −20.44+ i1.29 (λ 0 = 700 nm) [46], and those for SiO2 and ZnO are ε r2 = 2.12 [46] and ε r2 = 3.96 [47], respectively.

Download Full Size | PDF

5. Global and local coherence length

The definition of the degree of coherence in Eq. (8) enables us to introduce, in analogy with the traditional scalar theory, a coherence length, as a distance between r 1 and r 2 over which μ(r 1, r 2, ω) drops from its maximum value at r 1 = r 2 to a particular number (chosen appropriately for each situation). It was shown in the previous section that μr, ω) generally decays towards μ=1/2 when |Δr| increases, making it natural to define a coherence length with the ‘particular number’ to be close to μ . Therefore we define such a length, L coh, which we call the global coherence length, as the distance from the excitation point after which μr, ω) differs from μ at most by a small value ξ, i.e.,

|Δr|Lcoh|μ(Δr,ω)μ|ξ,ξ1.
The global coherence length is then a measure for the range outside which μr, ω) does not essentially change anymore. We recall from Sec. 4 that the degree of coherence generally oscillates, starting from μ max that according to Eq. (22) is close to unity. The quantity L coh therefore marks the distance up to which μr, ω) fluctuates, physically meaning that within it there are regions in which the electromagnetic SPP field at the two points may be highly correlated and regions where the field may be rather uncorrelated.

Owing to the oscillations the degree of coherence shows local maxima and minima at certain positions which we in Sec. 4.3 labeled as Δrmax(m) and Δrmin(m), respectively. Thus it is reasonable to further introduce an additional, local coherence length, l coh, as the shortest distance between two such nearby points, i.e.,

lcoh=π/|Δk|.
We emphasize, though, that the local coherence length is not meaningful if L coh < l coh. Equation (24) indicates that l coh → 0 for d → 0 and l coh → ∞ as d → ∞. Further Eq. (24) shows that l coh is independent of m and thus, for any chosen media, frequency, and slab thickness, the distances between all the nearby extrema are the same. Nevertheless, as was already pointed out in Sec. 4.3, the values of the extrema themselves generally depend on m. The global extrema, μ max and μ min given by Eq. (21), are the most interesting ones, and therefore we refer specifically to the distance between μ max and μ min when addressing l coh. Thus the local coherence length gives a rough estimation for the domain at the excitation region in which the field is electromagnetically coherent (meaning that all field components are fully correlated). Such a field interacts with nearby nanoscatterers in a coherent manner as in [32].

We next analyze the behavior of L coh and l coh along the x and z axes separately. When discussing the surface propagation length and the penetration depth into the surrounding of the field, lx and lz, respectively, we refer to those of the LRSPP, i.e., lx ≡ 1/kx (+)″ and lz ≡ 1/kz (+)″.

5.1. Along the slab

Let us first consider the local coherence length, l coh,x = π/|Δk′x|, whose behavior along a Ag layer in vacuum is illustrated in Fig. 6 as a function of d for some optical wavelengths. We find from Fig. 6 that increasing λ 0 results in a larger l coh,x in terms of the wavelength (left panel), but a smaller one with respect to the propagation distance (right panel), especially when λ 0 approaches the near-ultraviolet regime. One further notices that, for slab thicknesses where l coh,xlx, the coherence length can be several tens of λ 0 at larger wavelengths, while for very thin films, where l coh,x is only a fraction of the propagation range, a smaller wavelength may lead to subwavelength coherence lengths. Figure 7, on the other hand, illustrates the d-dependent behavior of l coh,x at λ 0 = 632.8 nm along a Ag slab in different surroundings, from which we observe that a higher ε r2 results in a smaller coherence length compared to the wavelength in the surrounding medium ( λ2=λ0εr2), but a larger one with respect to lx.

 figure: Fig. 6

Fig. 6 Local coherence length l coh,x with respect to the free-space wavelength λ 0 (left) and the surface propagation distance lx (right) for the LRSPP-SRSPP superposition along (Δz = 0) a Ag slab in vacuum, as a function of the film thickness d at selected wavelengths: λ 0 = 400 nm (solid-blue lines), λ 0 = 550 nm (dashed-green lines), and λ 0 = 700 nm (dotted-red lines). The modes are mutually uncorrelated and have the same intensities at the excitation point. The insets demonstrate the behavior of l coh,x at very thin films. The relative permittivity of Ag is ε r1 = −3.77 + i0.67 (λ 0 = 400 nm), ε r1 = −11.13 + i0.83 (λ 0 = 550 nm), and ε r1 = −20.44 + i1.29 (λ 0 = 700 nm) [46].

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Local coherence length l coh,x with respect to the wavelength in the surrounding medium λ 2 (left) and the surface propagation distance lx (right) for the LRSPP-SRSPP field along (Δz = 0) a Ag slab as a function of the film thickness d at λ 0 = 632.8 nm in different surroundings: vacuum (solid-blue lines), SiO2 (dashed-green lines), and ZnO (dotted-red lines). The modes are mutually uncorrelated and have the same intensities at the excitation point. The insets demonstrate the behavior of l coh,x at very thin films. The relative permittivities of Ag, SiO2, and ZnO are ε r1 = −15.87 + i1.07 [46], ε r2 = 2.12 [46], and ε r2 = 3.96 [47], respectively.

Download Full Size | PDF

Figures 8 and 9 depict the global coherence length introduced in Eq. (23) when ξ = 0.01 for the same fields considered in Figs. 6 and 7, from which one observes that the λ 0- and ε r2-dependent behavior of L coh,x is somewhat similar to that of l coh,x. However, the order of magnitude of L coh,x differs from that of l coh,x. Particularly one finds that for thicker slabs the global coherence length can be many times larger than the actual propagation distance of the field, even several hundreds or thousands of wavelengths, indicating that μr, ω) may oscillate over the field’s whole propagation regime. On the other hand, for very thin films L coh,x may be just a fraction of lx, which signifies that, apart from a small |Δr| with respect to lx, the degree of coherence is practically constant along the surface of the slab. The small abrupt jumps in Figs. 8 and 9 arise from the formal definition of L coh in Eq. (23).

 figure: Fig. 8

Fig. 8 Global coherence length L coh,x with respect to the free-space wavelength λ 0 (left) and the surface propagation length lx (right) for the fields of Fig. 6 as a function of the film thickness d. The labeling of the curves (λ 0 values) is as in Fig. 6.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Global coherence length L coh,x with respect to the wavelength in the surrounding medium λ 2 (left) and the surface propagation distance lx (right) for the fields of Fig. 7 as a function of the film thickness d. The labeling of the curves (ε r2 values) is as in Fig. 7.

Download Full Size | PDF

5.2. Perpendicular to the slab

In this case Eq. (24) for l coh is indeed still valid (l coh,z = πk′z), but it turns out that l coh,z > L coh,z (data not shown) due to the ‘overdamped’ behavior of μr, ω) in the z direction (cf. Fig. 2), whereupon there is no meaning, at least in a practical sense, to investigate the local coherence perpendicularly to the slab. Therefore we consider only the global coherence length along the z axis.

Figures 10 and 11 illustrate the d-dependent transverse behavior of L coh,z defined by Eq. (23) for ξ = 0.01 above a Ag layer for selected vacuum wavelengths and surroundings, respectively. We see from Figs. 10 and 11 that varying the free-space wavelength or the surrounding medium has the same effect on the global coherence length in the z direction as along the x axis (cf. Figs. 8 and 9): increasing λ 0 (ε r2) gives a larger (smaller) L coh,z with respect to the wavelength, but a smaller (larger) one in terms of lz. One also finds that, unless the film gets very thin, the global coherence length is (much) larger than the actual penetration depth of the field. Consequently, since this feature implies that the degree of coherence decays slowly and monotonically [L coh,z > l coh,z and thus there is no oscillation of μr, ω) in the z direction, cf. Fig. 2], the field is highly coherent perpendicularly to the slab, even for quite small slab thicknesses. Nonetheless, as d becomes very small we see that L coh,z < lz and L coh,z < λ 2, indicating that μr, ω) can change notably at subwavelength distances when moving away from the surface of an ultra-thin film.

 figure: Fig. 10

Fig. 10 Global coherence length L coh,z with respect to the free-space wavelength λ 0 (left) and the penetration depth lz (right) for the fields of Fig. 6 and 8 as a function of the film thickness d. The labeling of the curves (λ 0 values) is as in Fig. 6.

Download Full Size | PDF

 figure: Fig. 11

Fig. 11 Global coherence length L coh,z with respect to the wavelength in the surrounding medium λ 2 (left) and the penetration depth lz (right) for the fields of Fig. 7 and 9 as a function of the film thickness d. The labeling of the curves (ε r2 values) is as in Fig. 7.

Download Full Size | PDF

6. Degree of polarization in the LRSPP-SRSPP superposition

We now turn our attention to the polarization properties and consider, as before, the LRSPP-SRSPP superposition above the slab. For such a two-mode field, according to Eqs. (12) and (16), the spectral polarization matrix in Eq. (7) reads as

Φ(r,ω)=|E(+)|2e2k(+)rp^(+)p^(+)+|E()|2e2k()rp^()p^()+2e[k(+)+k()]r[E(+)*E()eiΔkrp^(+)p^()].
In a development paralleling the steps taken in Sec. 4, we thus assume that 〈E (+)* E (−)〉 = 0 and rewrite rr 1 + r where r 1 is the excitation point at which the intensities of the two modes are equal. On the basis of Eq. (25), in this case the degree of polarization in Eq. (10) becomes
P(r,ω)=11κcosh2(Δkr),
where also Eqs. (15) and (17) have been used.

6.1. Spatial behavior

Unlike the degree of coherence in Eq. (14), one notices from Eq. (26) that the degree of polarization does not display an oscillatory term, but is characterized only by a hyperbolic cosine. Excluding the particular direction along which P(r, ω) is constant, i.e., the one that is perpendicularly to Δk″, for a given slab thickness Eq. (26) indicates that

P(r,ω){1,|r|,κ,|r|0,
which are readily verified to be, respectively, the maximum and minimum of P(r, ω). Consequently, since 1/2 ≤ κ ≤ 1 [cf. Eq. (18)], we conclude that the degree of polarization for the LRSPP-SRSPP superposition is always bounded as
1/21/κP(r,ω)1.
The physical origin behind the (ε r1-, ε r2-, ω-, and d-independent) maximum, representing a fully polarized field, is by now apparent: when |r| becomes large enough, the SRSPP vanishes and the field can practically be considered as a single LRSPP for which P(r, ω) = 1 regardless of the chosen media, frequency, film thickness, and position [cf. the behavior of μr, ω) in Eq. (19)]. By a similar reasoning, i.e., as the contribution of the SRSPP to the degree of polarization cannot be neglected when |r| → 0, the minimum, corresponding to partial polarization, depends on ε r1, ε r2, ω, and d, according to Eqs. (3), (4), and (15).

Figure 12 illustrates the spatial behavior of P(r, ω) given in Eq. (26) for Ag films of two different thicknesses in vacuum at λ 0 = 632.8 nm. It is seen from Fig. 12 that for d = 50 nm (left panel) the degree of polarization is about unity everywhere [even for the minimum at |r| = 0 one has P(r, ω) = 0.993], corresponding to a fully polarized field. However, when the slab thickness is reduced to 20 nm (right panel), we observe that the degree of polarization slightly decreases [P(r, ω) ≈ 0.9] within a small region (Δx ≲ 3λ 0 and Δz ≲ 0.15λ 0), suggesting that the field can be partially polarized for ultra-thin films within, or close to, subwavelength distances from the excitation point.

 figure: Fig. 12

Fig. 12 Degree of polarization P(r, ω) for the LRSPP-SRSPP field above a 50 nm (left) and 20 nm (right) thick Ag slab in vacuum, as a function of the longitudinal and transverse distances x/λ 0 and z/λ 0, respectively, at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point x = z = 0. The straight lines illustrate the directions in which P(r, ω) is constant, determined by Δk″ (the vectors are drawn with correct relative magnitudes). The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].

Download Full Size | PDF

6.2. Slab-thickness dependence

By choosing an arbitrary r, such that |r| > 0 (and r ⊥̸ Δk″), Eqs. (15) and (26) imply that

P(r,ω){1,d1,d0,
where both limits, accounting for a fully polarized field, are independent of material parameters and frequency. As for μr, ω) in Eq. (20), the upper value of P(r, ω) in Eq. (29) follows from the fact that, for a sufficiently large film thickness, the wave vectors for the constituents of the two-mode field become degenerate. The reason for the lower value in Eq. (29) is somewhat similar to that of μr, ω) found in Eq. (20), namely, for any |r| > 0, no matter how small, the field contains only the LRSPP because the decay rate of the SRSPP diverges as the slab thickness goes to zero [if |r| = 0, on the other hand, the SRSPP cannot be neglected and in that case Eqs. (18) and (27) imply that P(r,ω)1/2 when d → 0].

Equations (26) and (29) together indicate that for a fixed |r| > 0 there exists a specific film thickness for which the degree of polarization is at its lowest. Furthermore, the results in Sec. 6.1 allude that the field gets partially polarized inside a particular (small) region as d is very small. Therefore, in Fig. 13 we demonstrate the behavior of P(r, ω) given in Eq. (26) along the x and z axes for ultra-thin Ag slabs in vacuum when λ 0 = 632.8 nm. Figure 13 shows that, indeed, within a particular region, the superposition becomes less polarized when the film thickness is reduced, but at the same time also the regime in which the field is partially polarized gets smaller. Nevertheless, the conclusion is clear: for a sufficiently thin slab, the degree of polarization of the two-mode field can fluctuate at subwavelength distances from the excitation point even if it is fully polarized farther away.

 figure: Fig. 13

Fig. 13 Degree of polarization P(r, ω) for the LRSPP-SRSPP field along the x (left) and z (right) axes above ultra-thin Ag slabs of thickness d in vacuum at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point x = z = 0. The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].

Download Full Size | PDF

6.3. Minimum

Finally, we examine how the choice of λ 0 and ε r2 affects the degree of polarization of the superposition. Since 1/κ is the fundamental lower boundary for the polarization degree of the two-mode field [cf. Eq. (28)], labeled as P min, we focus specifically on this minimum value.

Figure 14 illustrates the d-dependent behavior of P min above a Ag layer for selected vacuum wavelengths and surroundings, from which one finds that decreasing λ 0 (left panel) reduces the degree of polarization, while varying ε r2 (right panel) has a negligible effect on P min.

 figure: Fig. 14

Fig. 14 Minimum P min of the degree of polarization for the LRSPP-SRSPP field above a Ag slab for different wavelengths in vacuum (left), and various surroundings at the free-space wavelength λ 0 = 632.8 nm (right), as a function of the film thickness d. The solid-blue, dashed-green, and dotted-red lines in the left (right) panel correspond to λ 0 = 400 nm (vacuum), λ 0 = 550 nm (SiO2), and λ 0 = 700 nm (ZnO), respectively. The modes are mutually uncorrelated and have the same intensities at the excitation point. The relative permittivity of Ag is ε r1 = −3.77 + i0.67 (λ 0 = 400 nm), ε r1 = −11.13 + i0.83 (λ 0 = 550 nm), ε r1 = −15.87+ i1.07 (λ 0 = 632.8 nm), and ε r1 = −20.44+ i1.29 (λ 0 = 700 nm) [46], and those of SiO2 and ZnO are ε r2 = 2.12 [46] and ε r2 = 3.96 [47], respectively.

Download Full Size | PDF

7. Conclusions

Within a rigorous electromagnetic framework, we have investigated the spatial coherence and polarization characteristics of a superposition field that consists of the long-range and the short-range surface-plasmon polaritons at a thin metal slab. The two polariton modes can be excited in a number of different ways and their mutual correlation depends on the details of the actual generation arrangement. The SRSPP, which has attracted little attention and practical importance as compared to the LRSPP, plays a crucial role in this study as it can be viewed to enable the creation of partial coherence and partial polarization in the two-mode SPP field. The aim of the work was to examine the fundamental limits that the degrees of electromagnetic coherence and polarization associated with such a LRSPP-SRSPP superposition field can assume. To this end, the two modes were taken uncorrelated and of equal intensity at their excitation point. Furthermore, we have explored how the two degrees vary within their extremal values when the media, frequency, and film thickness are changed.

Regarding coherence, we specifically analyzed the (spectral) degree of spatial coherence of the two-mode superposition field above the metal slab. Due to the vectorial nature of SPPs, for which the traditional degree of coherence for scalar fields is inadequate, an appropriate degree of coherence valid for vector fields was employed. It was shown that the superposition field is highly coherent and assumes the maximum value for the degree of coherence at the point where the two modes are generated. Nevertheless, away from the excitation region, the coherence degree oscillates and sufficiently far the field generally becomes partially coherent with an essentially constant degree of coherence. Because of this behavior, we introduced a global coherence length as the distance from the excitation point beyond which the degree of coherence does not essentially change anymore. Physically this means that within the global coherence length there are locations at which the field may be highly correlated or relatively uncorrelated with respect to the field at the excitation point. It was demonstrated that for thicker slabs the global coherence length can extend over ranges many times larger than the actual propagation distance or penetration depth of the field, from tens to even thousands of wavelengths, while thinner films may result in subwavelength global coherence lengths. The oscillatory behavior, particularly along the surface of the slab, led us further to introduce a local coherence length, as the distance from the excitation point to the first minimum (which is also the global minimum) of the degree of coherence. Hence the local coherence length serves as a rough measure for the regime in which the two-mode field is practically electromagnetically coherent. Similarly to the global coherence length, it was shown that the local coherence length can be greater than the field’s propagation range when the film thickness is large, or alternatively, just a fraction of it if the slab is very thin. Furthermore, we demonstrated that increasing the frequency and/or the permittivity of the surrounding medium results in larger global as well as local coherence lengths with respect to the propagation distance and the penetration depth, but smaller ones in terms of the wavelength.

Concerning polarization, we investigated the (spectral) degree of polarization of the two-mode field above the slab. Even though the usual two-dimensional treatment is adequate for describing the polarization state of such a superposition, the degree of polarization is here defined in the plane of incidence where the two modes are elliptically polarized, and thus the situation differs substantially from that for ordinary beam-like wave fields. As might be expected, since the two modes are p polarized and excited in the same plane, we found that the superposition field is generally highly polarized. Nevertheless, the analysis suggested that for ultra-thin films the field can be partially polarized and the degree of polarization may fluctuate within subwavelength distances from the excitation region. Fully polarized and partially polarized electromagnetic fields interact with nanoparticles differently, and such scattering may be used to determine the state of polarization of the field. Finally we investigated the influence of the frequency and the surrounding on the polarization and found that reducing the frequency increases the degree of polarization of the field, while varying the permittivity of the surrounding has a negligible effect on it.

The analysis in this work on the LRSPP-SRSPP superposition can be regarded quite generic, since it explores the fundamental coherence and polarization limits of such a two-mode field, irrespective of the excitation technique. Thus our results could particularly find use in various interdisciplinary plasmonic and nanophotonic applications, including optoelectronic thin-film waveguiding and near-field interactions, where controlled partial spatial coherence is desirable. Our results also indicate the possibility for achieving and tailoring genuine subwavelength polarization effects by utilizing such a configuration.

Acknowledgments

This work was partially funded by the Academy of Finland (projects 268480 and 268705). A. Norrman is thankful for support from the Jenny and Antti Wihuri Foundation.

References and links

1. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).

2. D. K. Gramotnev and S. I. Bozhevolnyi, “Plasmonics beyond the diffraction limit,” Nature Photon. 4, 83–91 (2010). [CrossRef]  

3. M. I. Stockman, “Nanoplasmonics: past, present, and glimpse into future,” Opt. Express 19, 22029–22106 (2011). [CrossRef]   [PubMed]  

4. H. Raether, Surface Plasmons on Smooth and Rough Surfaces and on Gratings (Springer, 1988).

5. J. M. Pitarke, V. M. Silkin, E. V. Chulkov, and P. M. Echenique, “Theory of surface plasmons and surface-plasmon polaritons,” Rep. Prog. Phys. 70, 1–87 (2007). [CrossRef]  

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

7. A. V. Zayats, I. I. Smolyaninov, and A. A. Maradudin, “Nano-optics of surface plasmon polaritons,” Phys. Rep. 408, 131–314 (2005). [CrossRef]  

8. L. Novotny and B. Hecht, Principles of Nano-Optics, 2nd ed. (Cambridge University, 2012). [CrossRef]  

9. A. Norrman, T. Setälä, and A. T. Friberg, “Exact surface-plasmon polariton solutions at a lossy interface,” Opt. Lett. 38, 1119–1121 (2013). [CrossRef]   [PubMed]  

10. A. Norrman, T. Setälä, and A. T. Friberg, “Surface-plasmon polariton solutions at a lossy slab in a symmetric surrounding,” Opt. Express 22, 4628–4648 (2014). [CrossRef]   [PubMed]  

11. A. Norrman, T. Setälä, and A. T. Friberg, “Long-range higher-order surface-plasmon polaritons,” Phys. Rev. A 90, 053849 (2014). [CrossRef]  

12. J. J. Burke, G. I. Stegeman, and T. Tamir, “Surface-polariton-like waves guided by thin, lossy metal films,” Phys. Rev. B 33, 5186–5201 (1986). [CrossRef]  

13. J. A. Dionne, L. A. Sweatlock, H. A. Atwater, and A. Polman, “Planar metal plasmon waveguides: frequency-dependent dispersion, propagation, localization, and loss beyond the free electron model,” Phys. Rev. B 72, 075405 (2005). [CrossRef]  

14. A. R. Zakharian, J. V. Moloney, and M. Mansuripur, “Surface plasmon polaritons on metallic surfaces,” Opt. Express 15, 183–197 (2007). [CrossRef]   [PubMed]  

15. D. Sarid, “Long-range surface-plasma waves on very thin metal films,” Phys. Rev. Lett. 47, 1927–1930 (1981). [CrossRef]  

16. F. Yang, J. R. Sambles, and G. W. Bradberry, “Long-range surface modes supported by thin films,” Phys. Rev. B 44, 5855–5872 (1991). [CrossRef]  

17. P. Berini, “Long-range surface plasmon polaritons,” Adv. Opt. Photon. 1, 484–588 (2009). [CrossRef]  

18. A. Yanai and U. Levy, “The role of short and long range surface plasmons for plasmonic focusing applications,” Opt. Express 17, 14270–14280 (2009). [CrossRef]   [PubMed]  

19. Y. Li, H. Zhang, N. Zhu, T. Mei, D. H. Zhang, and J. Teng, “Short-range surface plasmon propagation supported by stimulated amplification using electrical injection,” Opt. Express 19, 22107–22112 (2011). [CrossRef]   [PubMed]  

20. Q. Gan, W. Bai, S. Jiang, Y. Gao, W. Li, W. Wu, and F. J. Bartoli, “Short-range surface plasmon polaritons for extraordinary low transmission through ultra-thin metal films with nanopatterns,” Plasmonics 7, 47–52 (2012). [CrossRef]  

21. L. Brigo, E. Gazzola, M. Cittadini, P. Zilio, G. Zacco, F. Romanato, A. Martucci, M. Guglielmi, and G. Brusatin, “Short and long range surface plasmon polariton waveguides for xylene sensing,” Nanotechnology 24, 155502 (2013). [CrossRef]   [PubMed]  

22. R. Slavíc and J. Homola, “Simultaneous excitation of long and short range surface plasmons in an asymmetric structure,” Opt. Commun. 259, 507–512 (2006). [CrossRef]  

23. R. Wan, F. Liu, Y. Huang, S. Hu, B. Fan, Y. Miura, D. Ohnishi, Y. Li, H. Li, and Y. Xia, “Excitation of short range surface plasmon polariton mode based on integrated hybrid coupler,” Appl. Phys. Lett. 97, 141105 (2010). [CrossRef]  

24. A. Norrman, T. Setälä, and A. T. Friberg, “Partial spatial coherence and partial polarization in random evanescent fields on lossless interfaces,” J. Opt. Soc. Am. A 28, 391–400 (2011). [CrossRef]  

25. R. Carminati and J. J. Greffet, “Near-field effects in spatial coherence of thermal sources,” Phys. Rev. Lett. 82, 1660–1663 (1999). [CrossRef]  

26. K. Joulain and C. Henkel, “The near field correlation spectrum of a metallic film,” Appl. Phys. B 93, 151–158 (2008). [CrossRef]  

27. A. V. Shchegrov, K. Joulain, R. Carminati, and J. J. Greffet, “Near-field spectral effects due to electromagnetic surface exitations,” Phys. Rev. Lett. 85, 1548–1551 (2000). [CrossRef]   [PubMed]  

28. T. Setälä, M. Kaivola, and A. T. Friberg, “Degree of polarization in near fields of thermal sources: effects of surface waves,” Phys. Rev. Lett. 88, 123902 (2002). [CrossRef]   [PubMed]  

29. J. J. Greffet, R. Carminati, K. Joulain, J. P. Mulet, S. Mainguy, and Y. Chen, “Coherent emission of light by thermal sources,” Nature 416, 61–64 (2002). [CrossRef]   [PubMed]  

30. S. Aberra Guebrou, J. Laverdant, C. Symonds, S. Vignoli, and J. Bellessa, “Spatial coherence properties of surface plasmon investigated by Young’s slit experiment,” Opt. Lett. 37, 2139–2141 (2012). [CrossRef]  

31. P. Debackere, S. Scheerlinck, P. Bienstman, and R. Baets, “Surface plasmon interferometer in silicon-on-insulator: novel concept for an integrated biosensor,” Opt. Express 14, 7063–7072 (2006). [CrossRef]   [PubMed]  

32. S. Aberra Guebrou, C. Symonds, E. Homeyer, J. C. Plenet, Yu. N. Gartstein, V. M. Agranovich, and J. Bellessa, “Coherent emission from a disordered organic semiconductor induced by strong coupling with surface plasmons,” Phys. Rev. Lett. 108, 066401 (2012). [CrossRef]   [PubMed]  

33. L.-P. Leppänen, A. T. Friberg, and T. Setälä, “Partial polarization of optical beams and near fields probed with a nanoscatterer,” J. Opt. Soc. Am. A 31, 1627–1635 (2014). [CrossRef]  

34. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995). [CrossRef]  

35. J. Tervo, T. Setälä, and A. T. Friberg, “Theory of partially coherent electromagnetic fields in the space-frequency domain,” J. Opt. Soc. Am. A 21, 2205–2215 (2004). [CrossRef]  

36. J. Tervo, T. Setälä, and A. T. Friberg, “Degree of coherence for electromagnetic fields,” Opt. Express 11, 1137–1143 (2003). [CrossRef]   [PubMed]  

37. T. Setälä, J. Tervo, and A. T. Friberg, “Complete electromagnetic coherence in the space-frequency domain,” Opt. Lett. 29, 328–330 (2004). [CrossRef]   [PubMed]  

38. T. Setälä, J. Tervo, and A. T. Friberg, “Contrasts of Stokes parameters in Young’s interference experiment and electromagnetic degree of coherence,” Opt. Lett. 31, 2669–2671 (2006). [CrossRef]  

39. L.-P. Leppänen, K. Saastamoinen, A. T. Friberg, and T. Setälä, “Interferometric interpretation for the degree of polarization of classical optical beams,” New J. Phys. 16, 113059 (2014). [CrossRef]  

40. T. Setälä, A. Shevchenko, M. Kaivola, and A. T. Friberg, “Degree of polarization for optical near fields,” Phys. Rev. E 66, 016615 (2002). [CrossRef]  

41. C. Brosseau and A. Dogariu, “Symmetry properties and polarization descriptors for an arbitrary electromagnetic wavefield,” in Progress in Optics, E. Wolf, ed. (Elsevier, 2006), vol. 49, pp. 315–380. [CrossRef]  

42. A. Otto, “Excitation of nonradiative surface plasma waves in silver by the method of frustrated total reflection,” Z. Phys. 216, 398–410 (1968). [CrossRef]  

43. E. Kretschmann, “Die Bestimmung optischer Konstanten von Metallen durch Anregung von Oberflächen-plasmaschwingungen,” Z. Phys. 241, 313–324 (1971). [CrossRef]  

44. T. Voipio, T. Setälä, and A. T. Friberg, “Statistical similarity and complete coherence of electromagnetic fields in time and frequency domains,” J. Opt. Soc. Am. A 32, 741–750 (2015). [CrossRef]  

45. R. J. Collier, C. B. Burckhardt, and L. H. Lin, Optical Holography (Academic, 1971).

46. E. D. Palik, ed., Handbook of Optical Constants of Solids (Academic, 1998).

47. W. L. Bond, “Measurement of the refractive indices of several crystals,” J. Appl. Phys. 36, 1674–1677 (1965). [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 (14)

Fig. 1
Fig. 1 Illustration of the directions of phase-front propagation (black arrows) and field attenuation (solid-red curves) for SPPs at a lossy metal film of thickness d in a non-absorptive dielectric surrounding, possessing relative permittivities ε r1 (complex) and ε r2 (real), respectively. The boundaries of the slab are at z = ±d/2.
Fig. 2
Fig. 2 Degree of coherence μr, ω) for the LRSPP-SRSPP superposition above a 50 nm (left) and 20 nm (right) thick Ag slab in vacuum, as a function of the longitudinal and transverse distances Δx/λ 0 and Δz/λ 0, respectively, at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point Δx = Δz = 0. The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].
Fig. 3
Fig. 3 Density plots corresponding to the 3D graphs of μr, ω) in Fig. 2. The dashed and dotted lines illustrate the directions in which the degree of coherence is solely oscillating and solely decaying, determined by Δk″ and Δk′, respectively. The dotted lines are located at the positions of local coherence minima. The vectors are drawn with correct relative magnitudes in each panel (but not among the panels).
Fig. 4
Fig. 4 Degree of coherence μr, ω) for the LRSPP-SRSPP field along the x (left) and z (right) axes above Ag slabs of thickness d in vacuum at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point Δx = Δz = 0. The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].
Fig. 5
Fig. 5 Maxima μ max and minima μ min of the degree of coherence for the LRSPP-SRSPP superposition as a function of the film thickness d for a Ag slab in vacuum at various free-space wavelengths λ 0 (left) and in different surroundings at λ 0 = 632.8 nm (right). The solid-blue, dashed-green, and dotted-red lines in the left (right) panel correspond to λ 0 = 400 nm (vacuum), λ 0 = 550 nm (SiO2), and λ 0 = 700 nm (ZnO), respectively. The modes are mutually uncorrelated and have the same intensities at the excitation point. The relative permittivity for Ag is ε r1 = −3.77 + i0.67 (λ 0 = 400 nm), ε r1 = −11.13 + i0.83 (λ 0 = 550 nm), ε r1 = −15.87+ i1.07 (λ 0 = 632.8 nm), and ε r1 = −20.44+ i1.29 (λ 0 = 700 nm) [46], and those for SiO2 and ZnO are ε r2 = 2.12 [46] and ε r2 = 3.96 [47], respectively.
Fig. 6
Fig. 6 Local coherence length l coh,x with respect to the free-space wavelength λ 0 (left) and the surface propagation distance lx (right) for the LRSPP-SRSPP superposition along (Δz = 0) a Ag slab in vacuum, as a function of the film thickness d at selected wavelengths: λ 0 = 400 nm (solid-blue lines), λ 0 = 550 nm (dashed-green lines), and λ 0 = 700 nm (dotted-red lines). The modes are mutually uncorrelated and have the same intensities at the excitation point. The insets demonstrate the behavior of l coh,x at very thin films. The relative permittivity of Ag is ε r1 = −3.77 + i0.67 (λ 0 = 400 nm), ε r1 = −11.13 + i0.83 (λ 0 = 550 nm), and ε r1 = −20.44 + i1.29 (λ 0 = 700 nm) [46].
Fig. 7
Fig. 7 Local coherence length l coh,x with respect to the wavelength in the surrounding medium λ 2 (left) and the surface propagation distance lx (right) for the LRSPP-SRSPP field along (Δz = 0) a Ag slab as a function of the film thickness d at λ 0 = 632.8 nm in different surroundings: vacuum (solid-blue lines), SiO2 (dashed-green lines), and ZnO (dotted-red lines). The modes are mutually uncorrelated and have the same intensities at the excitation point. The insets demonstrate the behavior of l coh,x at very thin films. The relative permittivities of Ag, SiO2, and ZnO are ε r1 = −15.87 + i1.07 [46], ε r2 = 2.12 [46], and ε r2 = 3.96 [47], respectively.
Fig. 8
Fig. 8 Global coherence length L coh,x with respect to the free-space wavelength λ 0 (left) and the surface propagation length lx (right) for the fields of Fig. 6 as a function of the film thickness d. The labeling of the curves (λ 0 values) is as in Fig. 6.
Fig. 9
Fig. 9 Global coherence length L coh,x with respect to the wavelength in the surrounding medium λ 2 (left) and the surface propagation distance lx (right) for the fields of Fig. 7 as a function of the film thickness d. The labeling of the curves (ε r2 values) is as in Fig. 7.
Fig. 10
Fig. 10 Global coherence length L coh,z with respect to the free-space wavelength λ 0 (left) and the penetration depth lz (right) for the fields of Fig. 6 and 8 as a function of the film thickness d. The labeling of the curves (λ 0 values) is as in Fig. 6.
Fig. 11
Fig. 11 Global coherence length L coh,z with respect to the wavelength in the surrounding medium λ 2 (left) and the penetration depth lz (right) for the fields of Fig. 7 and 9 as a function of the film thickness d. The labeling of the curves (ε r2 values) is as in Fig. 7.
Fig. 12
Fig. 12 Degree of polarization P(r, ω) for the LRSPP-SRSPP field above a 50 nm (left) and 20 nm (right) thick Ag slab in vacuum, as a function of the longitudinal and transverse distances x/λ 0 and z/λ 0, respectively, at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point x = z = 0. The straight lines illustrate the directions in which P(r, ω) is constant, determined by Δk″ (the vectors are drawn with correct relative magnitudes). The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].
Fig. 13
Fig. 13 Degree of polarization P(r, ω) for the LRSPP-SRSPP field along the x (left) and z (right) axes above ultra-thin Ag slabs of thickness d in vacuum at the free-space wavelength λ 0 = 632.8 nm. The modes are mutually uncorrelated and have the same intensities at the excitation point x = z = 0. The relative permittivity of Ag is ε r1 = −15.87 + i1.07 [46].
Fig. 14
Fig. 14 Minimum P min of the degree of polarization for the LRSPP-SRSPP field above a Ag slab for different wavelengths in vacuum (left), and various surroundings at the free-space wavelength λ 0 = 632.8 nm (right), as a function of the film thickness d. The solid-blue, dashed-green, and dotted-red lines in the left (right) panel correspond to λ 0 = 400 nm (vacuum), λ 0 = 550 nm (SiO2), and λ 0 = 700 nm (ZnO), respectively. The modes are mutually uncorrelated and have the same intensities at the excitation point. The relative permittivity of Ag is ε r1 = −3.77 + i0.67 (λ 0 = 400 nm), ε r1 = −11.13 + i0.83 (λ 0 = 550 nm), ε r1 = −15.87+ i1.07 (λ 0 = 632.8 nm), and ε r1 = −20.44+ i1.29 (λ 0 = 700 nm) [46], and those of SiO2 and ZnO are ε r2 = 2.12 [46] and ε r2 = 3.96 [47], respectively.

Equations (29)

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

E 2 ( ± ) ( r , ω ) = E 2 e i k x x | k 2 | { e + i k 2 z z ( k 2 z e ^ x + k x e ^ z ) , z d / 2 , e i k 2 z z ( ± k 2 z e ^ x ± k x e ^ z ) , z d / 2 ,
E 1 ( ± ) ( r , ω ) = E 1 e i ( k x x + k 1 z z ) | k 1 | [ k 1 z ( 1 e 2 i k 1 z z ) e ^ x + k x ( 1 ± e 2 i k 1 z z ) e ^ z ] , | z | < d / 2 ,
Symmetric ( + ) : ε r 1 ε r 2 k 2 z k 1 z = tanh ( 1 2 i k 1 z d ) ,
Antisymmetric ( ) : ε r 1 ε r 2 k 2 z k 1 z = coth ( 1 2 i k 1 z d ) ,
k x k 0 ε r z , k 1 z k 0 ε r 1 ε r 2 , k 2 z 0 ,
W ( r 1 , r 2 , ω ) = E ( r 1 , ω ) E ( r 2 , ω ) .
Φ ( r , ω ) W ( r , r , ω ) ,
μ ( r 1 , r 2 , ω ) W ( r 1 , r 2 , ω ) F tr Φ ( r 1 , ω ) tr Φ ( r 2 , ω ) ,
0 μ ( r 1 , r 2 , ω ) 1 ,
P ( r , ω ) 2 tr Φ 2 ( r , ω ) tr 2 Φ ( r , ω ) 1 ,
0 P ( r , ω ) 1 .
W ( r 1 , r 2 , ω ) = | E ( + ) | 2 e i [ k ( + ) r 2 k ( + ) * r 1 ] p ^ ( + ) p ^ ( + ) + | E ( ) | 2 e i [ k ( ) r 2 k ( ) * r 1 ] p ^ ( ) p ^ ( ) + E ( + ) * E ( ) e i [ k ( ) r 2 k ( + ) * r 1 ] p ^ ( + ) p ^ ( ) + E ( ) * E ( + ) e i [ k ( + ) r 2 k ( ) * r 1 ] p ^ ( ) p ^ ( + ) ,
p ^ ( a ) p ^ ( b ) = 1 | k ( a ) | | k ( b ) | ( k z ( a ) * k z ( b ) k z ( a ) * k x ( b ) k x ( a ) * k z ( b ) k x ( a ) * k x ( b ) ) , a , b { + , } .
μ ( r 1 , r 2 , ω ) = μ ( Δ r , ω ) = 1 2 1 + κ cos ( Δ k Δ r ) cos ( Δ k Δ r ) ,
κ [ p ^ ( + ) * p ^ ( ) ] 2 = | k x ( + ) k x ( ) * + k z ( + ) k z ( ) * | 2 | k ( + ) | 2 | k ( ) | 2 ,
Δ k [ k x ( + ) k x ( ) ] e ^ x + [ k z ( + ) k z ( ) ] e ^ z ,
Δ k [ k x ( + ) k x ( ) ] e ^ x + [ k z ( + ) k z ( ) ] e ^ z .
1 / 2 κ 1 ,
μ ( Δ r , ω ) { 1 / 2 , | Δ r | , ( 1 + κ ) / 2 , | Δ r | 0 .
μ ( Δ r , ω ) { 1 , d , 1 / 2 , d 0 ,
μ max = ( 1 + κ ) / 2 , μ min = ( 1 κ ) / 2 ,
3 / 2 μ max 1 , 0 μ min 1 / 2 ,
| Δ r | L coh | μ ( Δ r , ω ) μ | ξ , ξ 1 .
l coh = π / | Δ k | .
Φ ( r , ω ) = | E ( + ) | 2 e 2 k ( + ) r p ^ ( + ) p ^ ( + ) + | E ( ) | 2 e 2 k ( ) r p ^ ( ) p ^ ( ) + 2 e [ k ( + ) + k ( ) ] r [ E ( + ) * E ( ) e i Δ k r p ^ ( + ) p ^ ( ) ] .
P ( r , ω ) = 1 1 κ cosh 2 ( Δ k r ) ,
P ( r , ω ) { 1 , | r | , κ , | r | 0 ,
1 / 2 1 / κ P ( r , ω ) 1 .
P ( r , ω ) { 1 , d 1 , d 0 ,
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.