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

Determination of the excitation rate of quantum dots mediated by momentum-resolved Bloch-like surface plasmon polaritons

Open Access Open Access

Abstract

When light emitters are being placed in close proximity to a plasmonic system, not only the emission but also the excitation can be strongly enhanced and both yield the surface plasmon polariton (SPP) mediated fluorescence enhancement. Here, we combine the rate equation model and coupled mode theory to formulate the excitation rate of light emitters located on a periodic metallic array. The rate is expressed in terms of quantities that can be measured by angle- and polarization-resolved reflectivity and photoluminescence spectroscopy. As a demonstration, we have studied the excitation rate of CdSeTe quantum dots deposited on a 2D Au nanohole array as a function of the propagation direction of the (−1,0) Bloch-like SPPs. At the excitation wavelength of 633 nm, we find the rate remains almost constant at ~44 ps−1 regardless of the propagation direction of SPPs, which move from the Γ-X towards the Γ-M direction in the first Brillouin zone, and the crossing of the (−1,0) and (0,-1) SPPs along the Γ-M direction where two bright and dark modes are formed. The results are supported by the finite-difference time-domain simulations. We conclude the excitation rate is an intrinsic parameter and the enhanced excitation of the quantum dots arises entirely from field enhancement.

© 2017 Optical Society of America

1. Introduction

The interaction between light emitters and surface plasmon polaritons (SPPs) arising from metal nanostructures has aroused much interest recently. Its knowledge plays a major role in understanding the fundamental of light-matter interaction as well as in the advancement of fluorescence enhancement [1], energy harvesting [2], biosensing [3], and nonlinear optics [4]. For the fluorescence, SPPs enhance not only the spontaneous emission but also the excitation of the emitters [5,6]. The excitation and emission enhancements together yield the SPP mediated fluorescence [5,6]. To understand the entire process completely, differentiating the enhancements one by one is necessary for quantification. While studies on extracting the enhanced emission rate from the well-known Purcell factor have been conducted for years [7,8], little is known about the enhanced absorption. In fact, understanding how the absorption is affected by SPPs is of great importance not only in fluorescence but also in photovoltaics where an increase of the excited carrier concentration is always favorable for photocurrent generation [2,9]. Unfortunately, the exact determination of the excitation rate is not yet available and most of the studies reveal only the excitation efficiency ratio, which simply compares the emission with and without SPPs, in the photoluminescence excitation spectroscopy (PLE) and this method sometimes is very system dependent for quantification [10,11].

In this work, we combine the rate equation model and coupled mode theory (CMT) to determine the excitation rate of light emitters positioned on a metallic array which generates Bloch-like SPPs. The rate is analytically expressed in measurable quantities. By angle- and polarization-resolved reflectivity and photoluminescence (PL) spectroscopy, we have determined the excitation rate of CdSeTe quantum dots (QDs) on 2D Au nanohole array mediated by the (−1,0) SPPs propagating in different directions. At λ = 633 nm, it is found the rate is ~44 ps−1 and is independent of the propagation direction when the SPPs move away from the Γ-X direction as well as along the Γ-M direction where two degenerate SPPs couple with each other to form the bright and dark modes. Our results are consistent with the finite-difference time-domain (FDTD) simulations and they indicate the enhanced absorption of the QDs depends only on the plasmonic field.

2. Excitation rate formulation

2.1 Rate equation model

When light emitters are placed near a periodic plasmonic system such as a 2D square lattice array, SPP mediated fluorescence occurs [12]. Physically, the emitters can be described as a simple three level system as given in Fig. 1(a) such that the excitation and emission wavelengths, λex and λem, are slightly different. This difference occurs in many organic and inorganic emitters where the Stokes shift prevails [13]. The incident light at λex with power Po excites the electrons in the emitters from the ground state (S0) to the second excited state (S2) with excitation efficiency η and at the same time generates the incoming Block-like propagating SPPsin if the phase-matching condition is fulfilled [12]:

2πλexεaεmεm+εa=(2πλexsinθcosφ+2mπP)2+(2πλexsinθsinφ+2nπP)2
where εa and εm are the dielectric constants of the dielectrics and metal, θ and φ are the incident polar and azimuthal angles, P is the period of the array, and m and n are the integers specifying the order of SPPs. These SPPs can either transfer their energy to the emitters with the excitation rate Γex to yield the excitation enhancement or decay via Ohmic absorption and radiation damping at Γabsin and Γradin [14]. Noted η is wavelength dependent and is expected to follow the absorption and the thickness of the emitters. The electrons then fall to the first excited state (S1) with negligible energy loss and the excited emitters with energy [EM] will decay at λem via direct radiative emission, direct nonradiative absorption, and coupling of energy to the outgoing SPPsout specified by the rates Γr, Γnr, and Γc. Likewise, the outgoing SPPs decay via the absorption and radiative decay rates denoted as Γabsout and Γradout. The coupling Γc and the subsequent Γradout together define the emission enhancement [12,15]. It is noted that the propagation directions of the incoming and outgoing SPPs are different. While the direction of incoming SPPs is determined completely by the phase-matching equation given in Eq. (1), the outgoing SPPs travel in all directions as the interaction between the emitters and the outgoing SPPs is primarily near field in nature [12]. Nevertheless, the radiation damping of both SPPs follows the reverse of Eq. (1). As a result, Γex and Γc together play an important role in giving rise to the SPP mediated fluorescence. The determination of Γc has been conducted earlier in [12], we will focus on Γex here.

 figure: Fig. 1

Fig. 1 (a) The schematic of surface plasmon mediated emission fluorescence. The emitters contain a ground state (S0) and two excited states (S1 and S2). Excitation occurs from S0 to S2 by light source at λex with power Po whereas emission arises from the decay from S1 to S0 at λem. Both incoming and outgoing SPPs, SPPsin and SPPsout, can take part in the process. They have their respective absorption and radiative decay rates, Γabs and Γrad. Direct excitation has efficiency η and direction emission has nonradiative and radiative emission rates Γnr and Γr. (b) The schematic for the best coupling condition. Inset: the SEM image of the array. (c) The absorption and emission spectra of the CdSeTe quantum dots.

Download Full Size | PDF

By using the rate equation model [12,16], the transient of [EM] in the presence of incoming SPPs can be expressed as:

d[EM]dt=Γex|a|2[EM](Γr+Γnr+BZΓc(kSPPout)dkSPPout)
where |a|2is the energy of the incoming SPPs [17–19] and kSPPout is wavevector of the outgoing SPPs. We assume the absorption is dominantly plasmonic and the direct excitation is insignificant i.e. Poη0. Since the outgoing SPPs propagate in all directions, Γc is a function of the propagation vector, kSPP, and the total coupling rate should be the integral over the entire Brillouin zone (BZ) [12]. In steady state where d[EM]/dt=0, Eq. (2) becomesΓex|a|2=[EM](Γr+Γnr+BZΓc(kSPPout)dkSPPout). Similarly, for the direct excitation without involving the incoming SPPs, we have Poη=[EM]w/o(Γr+Γnr+BZΓc(kSPPout)dkSPPout), where [EM]w/o is the energy of the emitters without the excitation of the incoming SPPs. The items within the bracket on the right-hand side of two cases are the same, indicating the decays of the outgoing SPPs are identical with or without the incoming SPPs. Dividing one from another and knowing the direct emission powers are Pdw=Γr[EM] and Pdw/o=Γr[EM]w/o, we obtain: Γex|a|2Poη=PdwPdw/o, or
Γex=η(PdwPdw/o)Po|a|2.
As we will show later, Pdw/Pdw/o can be measured by using angle-resolved PL spectroscopy whereas η and Po/|a|2will be discussed by CMT given in the following.

2.2 Coupled mode theory

If only one input and output port is available, the incoming SPPsin can be described by the CMT as dadt=iωexaΓtotin2a+κΓradinsp+ [14,17], where a is the SPPs mode amplitude, ωex is the resonant excitation frequency, Γtotin is the total decay rate of SPPs =Γabsin+Γradin+Γex, and |sp+|2=Po is the p-polarized incident power. κ is the complex coupling constant, which depends on the in-coupling phase shift [14,17] and the relationship between the incident polarization and the field of SPPs [20], as will be explained later. At ωex, |a|2Po=4κ2ΓradinΓtotin2. Therefore, Eq. (3) becomes:

Γex=η(PdwPdw/o)Γtotin24κ2Γradin.

2.3 Parametric determination for nondegenerate SPPs

We need to determine κ, η, and the decay rates in prior to solving Γex. κ is expected to depend on both the incident angle and the polarization [20]. In fact, to elucidate the dependence completely, the concept of the best coupling polarization angle needs to be considered here. The best polarization angle accounts for how much incident energy can be channeled to SPPs when the electric field component of the incident polarization is not aligned with the field of SPPs [20,21]. In other words, the best condition for energy input occurs when the incident polarization and the SPP field is alike. For nondegenerate SPPs where the electric field ESPPis on the same plane as the propagation direction k^SPP, this condition can be mathematically simplified as: e^(k^SPP×z^)=0, where e^ is the polarization unit vector and z^is the unit vector normal to the sample surface [20]. Following Fig. 1(b), these unit vectors can be written as e^=(cosαcosθ,sinα,cosαsinθ), k^SPP=(cosβ,sinβ,0), and z^=(0,0,1), where α is the best polarization angle α between e^ and the incident plane and β is the angle between the SPPs propagation and the incident plane that can be estimated from the phase matching equation in Eq. (1). Substituting them into the vector product, we have:

tanα=tanβcosθ.
From this equation, the best angle α for inputting energy can be determined once the sample orientation and the SPP order are fixed for a given wavelength. For example, for 2D square lattice array, along the Γ-X direction where φ = 0°, for ( ± 1,0) SPP modes so that β = 0°, we have α = 0° or a pure p-polarized light is the best polarization for exciting SPPs. Any α deviates from the best condition will reduce the amount of energy being input to the SPPs. However, it is noted that Eq. (5) only holds for nondegenerate SPP modes. In the Γ-M direction where two degenerate SPP modes couple with each other to form two hybridized dark and bright modes, the plasmonic field is no longer aligned with the propagation direction. Two SPP modes interfere to form standing waves with ESPP1+ESPP2 and ESPP1ESPP2 for the dark and bright modes, respectively [20,22,23] and the resulting fields are parallel with and perpendicular to the incident plane. The α for two cases are therefore 0° and 90°, which are consistent with the fact that p- and s-polarized lights are required to excite the dark and bright modes, respectively [24].

Following [20], κ can now be written as κ'cosα, which contains a complex constant κ' for in-coupling phase shift and cosα for the dependence of energy channeling on α. On the other hand, for out-coupling, under reciprocal theorem, the polarization of the radiation damping for SPPs will also follow Eq. (5) [20,25]. As a result, for any arbitrary incident polarization angle, the reflectivity contains two components. While the first is the non-resonant reflection that has the same polarization as the incidence, the second is the radiation damping from the SPPs with the polarization angle defined by Eq. (5). Therefore, upon p-incidence, the p- and s-polarized (orthogonal) reflectivity, Rpp and Rps, can be written as [14,20]:

[RppRps]=[|rp+κ'2cos2αΓradini(ωωex)+Γtotin/2|2η|κ'2cosαsinαΓradini(ωωex)+Γtotin/2|2].
where η is assumed as a constant around ωex. By fitting the Rpp and Rps spectra, one can determine rp, κ, η, Γtotin, and Γradin accordingly.

2.4 Parametric determination for two coupled degenerate SPPs

Equation (6) is only valid for one traveling SPPs. For the coupling of two degenerate SPP modes along the Γ-M direction, the analysis is different but has been studied earlier for 1D and 2D arrays by CMT [23,24,26]. Under coordinate transformation, the dynamics of the bright and dark mode amplitudes, a+ and a, can be written as ddt(a+a)=(iω˜+00iω˜)(a+a)+(0κ+'Γ+radκ'Γrad0)(sp+ss+), whereω˜+,=ω+,+iΓ+,tot/2are the complex frequencies for two modes, Γ+,radare the radiative decay rates, κ+,' are the modified coupling constants, a+ and a are excited solely by s- and p-polarized lights [23]. The mode amplitudes are a+=κ+'Γ+radi(ωω+)+Γ+tot/2ss+and a=κ'Γradi(ωω)+Γtot/2sp+. Therefore, at ω+,, the excitation rates are Γex,bright=η((PdwPdw/o)s)(Γ+tot)24κ+'2Γ+rad and Γex,dark=η((PdwPdw/o)p)(Γtot)24κ'2Γrad, where (PdwPdw/o)s,pare the power ratios under s- and p-excitations. The parameters κ'+,, Γ+,rad, and Γ+,tot again can be determined by fitting the s- and p-polarized reflectivity spectra withRss=|rs+κ+'2Γ+radi(ωω+)+Γ+tot/2|2η, and Rpp=|rp+κ'2Γradi(ωω)+Γtot/2|2η.

3. Experiment

3.1 Method

We will study the dependence of Γex on φ, which reflects different SPP propagation directions. CdSeTe QDs with diameter of 21 nm (Life Technologies) are spin-coated on a 2D Au nanohole array prepared by interference lithography [12]. The array has period P, hole depth and radius = 450, 80, and 60 nm as illustrated in the scanning electron microscopy (SEM) image in the inset of Fig. 1(b). Before spin-coating, the concentration of QDs is further diluted in methanol from 2 µM to 50 nM to reduce dot-dot interaction. The absorption and emission spectra of the QDs in methanol are shown in Fig. 1(c). The sample is then mounted onto a computer-controlled goniometer as shown in Fig. 2(a) for angle- and polarization-resolved reflectivity and PL measurements. The goniometer has three rotation stages to independently control θ and φ [12,27]. A pair of incident polarizer and detection analyzer is placed after the light source and the sample for polarization measurements. A collimated quartz lamp is used for the reflectivity measurements whereas a 633 nm HeNe laser is used for the PL spectroscopy. The signals are collected by a spectrometer attached with a CCD detector.

 figure: Fig. 2

Fig. 2 (a) The schematic of the angle- and polarization resolved reflectivity and photoluminescence spectroscopy. (b) The p-polarized specular reflectivity mapping of the array taken at φ = 0°. The dash line, obtained from the phase matching equation, indicates the excitation of (−1,0) SPPs. (c) The reflectivity spectrum (dot) extracted from θ = 18.5° and φ = 0° together with the best fit (solid line) for parametric determination.

Download Full Size | PDF

3.2 Reflectivity and photoluminescence measurements in one SPP direction

We present the reflectivity and PL results at one φ as an example to show how the parameters are extracted. By setting φ = 0°, the θ-dependent p-polarized (Rpp) reflectivity mapping, also known as the dispersion relation, of the sample is shown in Fig. 2(b) for mode identification. The orthogonal (Rps) mapping for φ = 0° is zero and thus is not shown. Apparently, the dispersive reflection dips match well with the (−1,0) SPP mode calculated by Eq. (1) [12]. In contrast, the absorption band of QDs is not seen owing to the low concentration. We extract the reflectivity spectrum taken at θ = 18.5° where the λex = 633 nm incoming SPPs are excited in Fig. 2(c) and then fit it with Eq. (6) to extract the parameters. The results, rp, κ', α, η, Γtotin, and Γradin, are tabulated in Table. 1. For determining Pdw/Pdw/o, angle-resolved PL scans are then performed. Two types of measurements are conducted and they are the detection and incident scans [12,27,28]. For the detection scan, the incident angle is kept fixed so that the excitation condition does not change but the detection angle is scanned for screening the angular emission at a certain emission plane [12,27,28]. As the radiation damping of the outgoing SPPs is highly emission angle dependent, detection scan probes the emission enhancement [12]. On the other hand, for the incident scan, while the detection angle is fixed with respect to the sample normal, the incident angle is varied continuously. This configuration allows us to selective excite the incoming SPPs while monitoring the emission at the same angle. In other words, it probes the excitation enhancement. As we will see, the interplay between the incident and detection scans assist us in identifying Pdw and Pdw/o and determining their ratio. A p-incident polarizer is used without the analyzer so that full emission is detected in these two scans. For illustration, Fig. 3(a) & (b) show the PL detection scans with φ = 0° and θ = 0° and 18.5°. An almost one to one correspondence is seen between the reflectivity mapping in Fig. 2(b) and the detection scans, confirming the emission enhancement from (−1,0) SPPs. One also sees a non-dispersive emission band at λ ~690 nm, which is attributed to the direct emission of QDs when comparing with Fig. 1(c). Except the intensities, two detection scans are almost the same, indicating the emission enhancement is identical with and without the incoming SPPs. We then perform the incident scan by keeping the detection angle at 0° in Fig. 3(c). Strong direct emission occurs at 18.5° is due to the excitation of the 633 nm incoming SPPs. The normalized emission intensity at λem = 690 nm is extracted in Fig. 3(d) for determining Pdw/Pdw/o. The emission shows an asymmetric “Fano-like” profile and the power ratio Pdw/Pdw/ois determined to be 6.9 in Table. 1. Once all the parameters are available, Γex is then calculated to be 40.3 ps−1 along the Γ-X direction.

Tables Icon

Table 1. Parametric determination of nondegenerate and degenerate SPP modes for different φ.

 figure: Fig. 3

Fig. 3 The detection scans of the array taken at φ = 0° and θ = (a) 0° without the incoming 633 nm SPPs and (b) 18.5° with the incoming SPPs. (c) The incident scan of the array taken at φ = 0° and a fixed detection angle = 0°. The band at 690 nm is from the direct emission. (d) The line incident scan taken from the direct emission at 690 nm.

Download Full Size | PDF

3.3 Dependence of Γex on propagation direction

The above procedures then apply to different φ. We measure the Rpp and Rps reflectivity spectra and the PL incident scans at λem = 690 nm from φ = 5° to 35°, where only one single 633 nm incoming SPP mode is excited, in Fig. 4. The best fits yield rp, κ, α, η, Γtotin, Γradin, and Pdw/Pdw/o in Table 1. We compare the fitted α with the values deduced from Eq. (5) in Fig. 5(a) and they agree with each other. Γex are then determined in Fig. 5(b), showing it is weakly dependent on φ.

 figure: Fig. 4

Fig. 4 The (a) p-polarized (Rpp) and (b) orthogonal (Rps) reflectivity spectra with the nondegenerate incoming SPPs located at 633 nm taken from φ = 5° – 35°. The best fits are indicated by the dash lines for parametric determination. The spectra are vertically shifted for visualization. (c) The line incident scans taken from the direct emission at 690 nm at detection angle = 0° and the corresponding φ.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 (a) The comparison between the fitted α and the values deduced from Eq. (5). (b) The plot of Γex with φ, including the bright and dark modes.

Download Full Size | PDF

For φ = 45° where two (−1,0) and (0,-1) SPPs cross, the p- and s-polarized reflectivity spectra are taken at θ = 40° and 29.5° in Fig. 6(a) & (b) so that the dark and bright modes are both excited at 633 nm. The PL incident scans taken under p- and s-polarizations are plotted in Fig. 6(c) for λem = 690 nm. Consistent with the reflectivity, strong excitation enhancements are observed at 29.5° and 38° for the bright and dark modes. The emission from the bright mode is stronger than that of the dark mode. The parameters are extracted from the reflectivity spectra as well as the PL incident scans and they are then tabulated in Table 1. Γex for the dark and bright modes are plotted in Fig. 5(b), showing they are almost the same as the nondegenerate counterparts.

 figure: Fig. 6

Fig. 6 The (a) Rpp and (b) Rss reflectivity spectra for the dark and bright modes taken at φ = 45° together with the best fits (dash lines). (c) The line incident scans taken from the direct emission at 690 nm for the dark and bright modes.

Download Full Size | PDF

4. Numerical simulation

The weak dependence of Γex on φ shows the excitation rate is an intrinsic parameter and is not affected by the environment. It depends only on the dipole moment of the emitters [29,30]. The plasmonic absorption therefore is caused by |a|2/Po, which is related to the |ESPP|2. As a result, if Γex/η is a constant, we expect PdwPdw/o=[EM][EM]w/o=Γex|a|2Poη|ESPP|2|Ew/o|2, where |Ew/o|2is the field strength without the incoming SPPs. To verify it, we perform FDTD simulations. The unit cell of the QDs/array for FDTD simulations is shown in the inset of Fig. 7(a). Following the SEM image, the nanohole array has P = 450 nm, hole depth and radius = 80 and 60 nm. The QDs layer is modeled as a 21 nm thick dielectric layer with the complex refractive index of 1.12 + 0.05i capped on the optically thick Au array. While the thickness of the QDs layer is measured by a profilometer, the refractive index is obtained by fitting the experimental p-polarized reflectivity spectrum taken at θ = 18.5° and φ = 0° with the FDTD data, as shown in Fig. 7(a). Bloch boundary condition is used at four sides whereas perfectly matched layer is used at the top and the bottom. A power monitor is placed at 5 nm above the surface to record the electric field intensity, which is then averaged across the entire simulation cell. The field intensity at different wavelengths are then simulated and plotted in Fig. 7(b) at different θ and φ under p-incidence for exciting the (−1,0) SPPs at 633 nm. For the bright and dark modes, s- and p-incidences are used at φ = 45° and θ = 29.5° and 38°. An intensity spectrum taken at both θ and φ = 0° so that no 633 nm SPPs is excited is also simulated in Fig. 7(b) for comparison. The intensity |ESPP|2/|Ew/o|2 and the experimental Pdw/Pdw/o ratios are then plotted in Fig. 7(c) for comparison. Their overall trend including the dark and bright modes agrees well between the simulation and experiment, verifying that Γex is a constant. Importantly, the decrease of the intensity of the nondegenerate SPPs as a function of φ is mostly due to the mismatch between the incident and the best polarizations, thereby reducing the energy coupling from the incidence to the SPPs, i.e. |ESPP|2. By plotting cos2α in Fig. 7(c), all are found to be consistent with each other.

 figure: Fig. 7

Fig. 7 (a) The FDTD simulated p-polarized reflectivity spectra with and without the dielectric layer calculated at θ = 18.5° and φ = 0° together with the experimental spectrum. Inset: the unit cell for the FDTD simulation. (b) The electric field intensity spectra calculated at 633 nm for φ = 0° - 45°. (c) The comparison between the direct emission power ratio Pdw/Pdw/o(solid squares), the calculated electric field intensity enhancement |ESPP|2/|Ew/o|2 (open circles), and cos2α (open triangles).

Download Full Size | PDF

5. Conclusion

In summary, we have determined φ-dependent Γex of CdSeTe QDs deposited on 2D Au nanohole array by using angle- and polarization-resolved reflectivity and photoluminescence spectroscopy. Based on the rate equation model and coupled mode theory, we express Γex in terms of measurable quantities that can be determined sequentially. We find Γex at 633 nm ~44 ps−1 and is independent of the propagation direction and the coupling of SPPs. Our results indicate Γex is an intrinsic parameter and depends only on the emitters themselves but not the environment. The enhanced absorption of the emitters is due solely to the plasmonic field enhancement. The experiment is found to agree very well with the FDTD simulations.

Funding

RGC Competitive Earmarked Research Grants (402812 and 14304314); Collaborative Research Fund CUHK1/CRF/12G; Area of Excellence AoE/P-02/12.

References and links

1. K. Okamoto, I. Niki, A. Shvartser, Y. Narukawa, T. Mukai, and A. Scherer, “Surface-plasmon-enhanced light emitters based on InGaN quantum wells,” Nat. Mater. 3(9), 601–605 (2004). [CrossRef]   [PubMed]  

2. H. A. Atwater and A. Polman, “Plasmonics for improved photovoltaic devices,” Nat. Mater. 9(3), 205–213 (2010). [CrossRef]   [PubMed]  

3. J. N. Anker, W. P. Hall, O. Lyandres, N. C. Shah, J. Zhao, and R. P. Van Duyne, “Biosensing with plasmonic nanosensors,” Nat. Mater. 7(6), 442–453 (2008). [CrossRef]   [PubMed]  

4. M. Kauranen and A. V. Zayats, “Nonlinear plasmonics,” Nat. Photonics 6(11), 737–748 (2012). [CrossRef]  

5. A. Kinkhabwala, Z. Yu, S. Fan, Y. Avlasevich, K. Mullen, and W. E. Moerner, “Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna,” Nat. Photonics 3(11), 654–657 (2009). [CrossRef]  

6. G. M. Akselrod, T. Ming, C. Argyropoulos, T. B. Hoang, Y. Lin, X. Ling, D. R. Smith, J. Kong, and M. H. Mikkelsen, “Leveraging Nanocavity Harmonics for Control of Optical Processes in 2D Semiconductors,” Nano Lett. 15(5), 3578–3584 (2015). [CrossRef]   [PubMed]  

7. I. Gontijo, M. Boroditsky, E. Yablonovitch, S. Keller, U. K. Mishra, and S. P. DenBaars, “Coupling of InGaN quantum-well photoluminescence to silver surface plasmons,” Phys. Rev. B 60(16), 11564–11567 (1999). [CrossRef]  

8. D. Y. Lei, J. Li, and H. C. Ong, “Tunable surface plasmon mediated emission from semiconductors by using metal alloys,” Appl. Phys. Lett. 91(2), 021112 (2007). [CrossRef]  

9. K. R. Catchpole and A. Polman, “Plasmonic solar cells,” Opt. Express 16(26), 21793–21800 (2008). [CrossRef]   [PubMed]  

10. Y. Chen, K. Munechika, I. Jen-La Plante, A. M. Munro, S. E. Skrabalak, Y. Xia, and D. S. Ginger, “Excitation enhancement of CdSe quantum dots by single metal nanoparticles,” Appl. Phys. Lett. 93(5), 053106 (2008). [CrossRef]  

11. L. Fornasari, F. Floris, M. Patrini, G. Canazza, G. Guizzetti, D. Comoretto, and F. Marabelli, “Fluorescence excitation enhancement by Bloch surface wave in all-polymer one-dimensional photonic structure,” Appl. Phys. Lett. 105(5), 053303 (2014). [CrossRef]  

12. Z. L. Cao and H. C. Ong, “Determination of coupling rate of light emitter to surface plasmon polaritons supported on nanohole array,” Appl. Phys. Lett. 102(24), 241109 (2013). [CrossRef]  

13. J. R. Lakowicz, Principles of Fluorescence Spectroscopy (Springer, 2011).

14. Z. Cao, H. Y. Lo, and H. C. Ong, “Determination of absorption and radiative decay rates of surface plasmon polaritons from nanohole array,” Opt. Lett. 37(24), 5166–5168 (2012). [CrossRef]   [PubMed]  

15. M. Pelton, “Modified spontaneous emission in nanophotonic structures,” Nat. Photonics 9(7), 427–435 (2015). [CrossRef]  

16. C. Y. Chen, D. M. Yeh, Y. C. Lu, and C. C. Yang, “Dependence of resonant coupling between surface plasmons and an InGaN quantum well on metallic structure,” Appl. Phys. Lett. 89(20), 203113 (2006). [CrossRef]  

17. H. A. Haus, Waves and Fields in Optoelectronics (Prentice-Hall, 1984).

18. T. J. Seok, A. Jamshidi, M. Kim, S. Dhuey, A. Lakhani, H. Choo, P. J. Schuck, S. Cabrini, A. M. Schwartzberg, J. Bokor, E. Yablonovitch, and M. C. Wu, “Radiation Engineering of Optical Antennas for Maximum Field Enhancement,” Nano Lett. 11(7), 2606–2610 (2011). [CrossRef]   [PubMed]  

19. Z. Cao, L. Zhang, C. Y. Chan, and H. C. Ong, “Interplay between absorption and radiative decay rates of surface plasmon polaritons for field enhancement in periodic arrays,” Opt. Lett. 39(3), 501–504 (2014). [CrossRef]   [PubMed]  

20. Z. L. Cao, L. Y. Yiu, Z. Q. Zhang, C. T. Chan, and H. C. Ong, “Understanding the role of surface plasmon polaritons in two-dimensional achiral nanohole arrays for polarization conversion,” https://arxiv.org/abs/1608.07027.

21. J. Elliott, I. I. Smolyaninov, N. I. Zheludev, and A. V. Zayats, “Wavelength dependent birefringence of surface plasmon polaritonic crystals,” Phys. Rev. B 70(23), 233403 (2004). [CrossRef]  

22. C. Sauvan, C. Billaudeau, S. Collin, N. Bardou, F. Pardo, J.-L. Pelouard, and P. Lalanne, “Surface plasmon coupling on metallic film perforated by two-dimensional rectangular hole array,” Appl. Phys. Lett. 92(1), 011125 (2008). [CrossRef]  

23. Z. L. Cao and H. C. Ong, “Determination of the absorption and radiative decay rates of dark and bright plasmonic modes,” Opt. Express 22(13), 16112–16129 (2014). [CrossRef]   [PubMed]  

24. Z. L. Cao and H. C. Ong, “Momentum-dependent group velocity of surface plasmon polaritons in two-dimensional metallic nanohole array,” Opt. Express 24(12), 12489–12500 (2016). [CrossRef]   [PubMed]  

25. J. D. Jackson, Calssical Electrodynamics (John Wiley & Sons, 1999).

26. C. Ropers, D. J. Park, G. Stibenz, G. Steinmeyer, J. Kim, D. S. Kim, and C. Lienau, “Femtosecond light transmission and subradiant damping in plasmonic crystals,” Phys. Rev. Lett. 94(11), 113901 (2005). [CrossRef]   [PubMed]  

27. C. Y. Chan, Z. L. Cao, and H. C. Ong, “Study of coupling efficiency of molecules to surface plasmon polaritons in surface-enhanced Raman scattering (SERS),” Opt. Express 21(12), 14674–14682 (2013). [CrossRef]   [PubMed]  

28. J. J. Baumberg, T. A. Kelf, Y. Sugawara, S. Cintra, M. E. Abdelsalam, P. N. Bartlett, and A. E. Russell, “Angle-Resolved Surface-Enhanced Raman Scattering on Metallic Nanostructured Plasmonic Crystals,” Nano Lett. 5(11), 2262–2267 (2005). [CrossRef]   [PubMed]  

29. G. Lozano, T. Barten, G. Grzela, and J. G. Rivas, “Directional absorption by phased arrays of plasmonic nanoantennae probed with time-reversed Fourier microscopy,” New J. Phys. 16(1), 013040 (2014). [CrossRef]  

30. G. Grzela, R. Paniagua-Domínguez, T. Barten, D. van Dam, J. A. Sánchez-Gil, and J. G. Rivas, “Nanowire Antenna Absorption Probed with Time-Reversed Fourier Microscopy,” Nano Lett. 14(6), 3227–3234 (2014). [CrossRef]   [PubMed]  

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 (a) The schematic of surface plasmon mediated emission fluorescence. The emitters contain a ground state (S0) and two excited states (S1 and S2). Excitation occurs from S0 to S2 by light source at λex with power Po whereas emission arises from the decay from S1 to S0 at λem. Both incoming and outgoing SPPs, SPPsin and SPPsout, can take part in the process. They have their respective absorption and radiative decay rates, Γabs and Γrad. Direct excitation has efficiency η and direction emission has nonradiative and radiative emission rates Γnr and Γr. (b) The schematic for the best coupling condition. Inset: the SEM image of the array. (c) The absorption and emission spectra of the CdSeTe quantum dots.
Fig. 2
Fig. 2 (a) The schematic of the angle- and polarization resolved reflectivity and photoluminescence spectroscopy. (b) The p-polarized specular reflectivity mapping of the array taken at φ = 0°. The dash line, obtained from the phase matching equation, indicates the excitation of (−1,0) SPPs. (c) The reflectivity spectrum (dot) extracted from θ = 18.5° and φ = 0° together with the best fit (solid line) for parametric determination.
Fig. 3
Fig. 3 The detection scans of the array taken at φ = 0° and θ = (a) 0° without the incoming 633 nm SPPs and (b) 18.5° with the incoming SPPs. (c) The incident scan of the array taken at φ = 0° and a fixed detection angle = 0°. The band at 690 nm is from the direct emission. (d) The line incident scan taken from the direct emission at 690 nm.
Fig. 4
Fig. 4 The (a) p-polarized (Rpp) and (b) orthogonal (Rps) reflectivity spectra with the nondegenerate incoming SPPs located at 633 nm taken from φ = 5° – 35°. The best fits are indicated by the dash lines for parametric determination. The spectra are vertically shifted for visualization. (c) The line incident scans taken from the direct emission at 690 nm at detection angle = 0° and the corresponding φ.
Fig. 5
Fig. 5 (a) The comparison between the fitted α and the values deduced from Eq. (5). (b) The plot of Γex with φ, including the bright and dark modes.
Fig. 6
Fig. 6 The (a) Rpp and (b) Rss reflectivity spectra for the dark and bright modes taken at φ = 45° together with the best fits (dash lines). (c) The line incident scans taken from the direct emission at 690 nm for the dark and bright modes.
Fig. 7
Fig. 7 (a) The FDTD simulated p-polarized reflectivity spectra with and without the dielectric layer calculated at θ = 18.5° and φ = 0° together with the experimental spectrum. Inset: the unit cell for the FDTD simulation. (b) The electric field intensity spectra calculated at 633 nm for φ = 0° - 45°. (c) The comparison between the direct emission power ratio P d w / P d w/o (solid squares), the calculated electric field intensity enhancement | E SPP | 2 / | E w/o | 2 (open circles), and cos2α (open triangles).

Tables (1)

Tables Icon

Table 1 Parametric determination of nondegenerate and degenerate SPP modes for different φ.

Equations (6)

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

2π λ ex ε a ε m ε m + ε a = ( 2π λ ex sinθcosφ+ 2mπ P ) 2 + ( 2π λ ex sinθsinφ+ 2nπ P ) 2
d[ EM ] dt = Γ ex | a | 2 [ EM ]( Γ r + Γ nr + BZ Γ c ( k SPP out ) d k SPP out )
Γ ex =η( P d w P d w/o ) P o | a | 2 .
Γ ex =η( P d w P d w/o ) Γ tot in 2 4 κ 2 Γ rad in .
tanα=tanβcosθ.
[ R pp R ps ]=[ | r p + κ ' 2 cos 2 α Γ rad in i( ω ω ex )+ Γ tot in /2 | 2 η | κ ' 2 cosαsinα Γ rad in i( ω ω ex )+ Γ tot in /2 | 2 ].
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.