You are on page 1of 9

Flexible nanofiber-coupled hybrid plasmonic

Bragg grating
Sheng Liu,1 Linjie Zhou,1,* Jian Xu,2 Xinyi Wang,1 and Jianping Chen1
1
State Key Laboratory of Advanced Optical Communication Systems and Networks, Department of Electronic
Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
2
Center for Advanced Electronic Materials and Devices, Shanghai Jiao Tong University, Shanghai 200240, China
*
ljzhou@sjtu.edu.cn

Abstract: We report a hybrid plasmonic Bragg grating composed of a


nanofiber coupled with orthogonally oriented metal strips. Numerical
simulations are performed to study the transmission and reflection spectra
of the grating. It shows that the TM polarization has much stronger Bragg
reflection due to the excitation of hybrid plasmonic modes. The dependence
of reflection peaks on several key device parameters is analyzed. Light
propagation simulation further reveals that both fundamental and first-order
TM modes are excited upon Bragg reflection, leading to two separate peaks
in the spectrum. We implement the prototype device by attaching a
nanofiber onto the surface of an array of sub-micrometer-wide metal strips.
The main reflection peak is measured to have a 3-dB bandwidth of 15 nm
and out-of-band rejection of more than 30 dB. The effects of nanofiber
radius, alignment angle and coupling length on the device performance are
also experimentally investigated.
©2016 Optical Society of America
OCIS codes: (250.5300) Photonic integrated circuits; (250.5403) Plasmonics.

References and links


1. R. F. Oulton, V. J. Sorger, D. A. Genov, D. F. P. Pile, and X. Zhang, “A hybrid plasmonic waveguide for
subwavelength confinement and long-range propagation,” Nat. Photonics 2(8), 496–500 (2008).
2. C. Delacour, S. Blaize, P. Grosse, J. M. Fedeli, A. Bruyant, R. Salas-Montiel, G. Lerondel, and A. Chelnokov,
“Efficient directional coupling between silicon and copper plasmonic nanoslot waveguides: toward metal-oxide-
silicon nanophotonics,” Nano Lett. 10(8), 2922–2926 (2010).
3. M. Février, P. Gogol, G. Barbillon, A. Aassime, R. Mégy, B. Bartenlian, J. M. Lourtioz, and B. Dagens,
“Integration of short gold nanoparticles chain on SOI waveguide toward compact integrated bio-sensors,” Opt.
Express 20(16), 17402–17410 (2012).
4. I. Goykhman, B. Desiatov, and U. Levy, “Experimental demonstration of locally oxidized hybrid silicon-
plasmonic waveguide,” Appl. Phys. Lett. 97(14), 141106 (2010).
5. Z. Han, A. Y. Elezzabi, and V. Van, “Experimental realization of subwavelength plasmonic slot waveguides on a
silicon platform,” Opt. Lett. 35(4), 502–504 (2010).
6. C. C. Huang, “Ultra-long-range symmetric plasmonic waveguide for high-density and compact photonic
devices,” Opt. Express 21(24), 29544–29557 (2013).
7. X. Sun, L. Zhou, X. Li, Z. Hong, and J. Chen, “Design and analysis of a phase modulator based on a metal-
polymer-silicon hybrid plasmonic waveguide,” Appl. Opt. 50(20), 3428–3434 (2011).
8. R. A. Wahsheh, Z. Lu, and M. A. Abushagur, “Nanoplasmonic couplers and splitters,” Opt. Express 17(21),
19033–19040 (2009).
9. Q. Gan and F. J. Bartoli, “Bidirectional surface wave splitter at visible frequencies,” Opt. Lett. 35(24), 4181–
4183 (2010).
10. Y. Luo, M. Chamanzar, and A. Adibi, “Compact on-chip plasmonic light concentration based on a hybrid
photonic-plasmonic structure,” Opt. Express 21(2), 1898–1910 (2013).
11. M. Z. Alam, J. S. Aitchison, and M. Mojahedi, “A marriage of convenience: Hybridization of surface plasmon
and dielectric waveguide modes,” Laser Photonics Rev. 8(3), 394–408 (2014).
12. J. M. M. Z. Alam, J. S. Aitchison, and M. Mojahedi, “Super Mode Propagation in Low Index Medium,”
Conference on Lasers and Electro-Optics/Quantum Eletronics and Laser Science Conference and Photonic
Applications System Technologies, OSA Technical Digest Series (CD) (Optical Society of America, 2007), paper
JThD112(2007).

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9316
13. M. N. Abbas, Y.-C. Chang, and M.-H. Shih, “Plasmon-polariton band structures of asymmetric T-shaped
plasmonic gratings,” Opt. Express 18(3), 2509–2514 (2010).
14. A. Dhawan, M. Canva, and T. Vo-Dinh, “Narrow groove plasmonic nano-gratings for surface plasmon resonance
sensing,” Opt. Express 19(2), 787–813 (2011).
15. M. P. Nielsen and A. Y. Elezzabi, “Nanoplasmonic distributed Bragg reflector resonators for monolithic
integration on a complementary metal-oxide-semiconductor platform,” Appl. Phys. Lett. 103(5), 051107 (2013).
16. J. Zhang, S. Zhang, D. Li, A. Neumann, C. Hains, A. Frauenglass, and S. R. Brueck, “Infrared transmission
resonances in double-layered, complementary-structure metallic gratings,” Opt. Express 15(14), 8737–8744
(2007).
17. N. Mattiucci, G. D’Aguanno, H. O. Everitt, J. V. Foreman, J. M. Callahan, M. C. Buncick, and M. J. Bloemer,
“Ultraviolet surface-enhanced Raman scattering at the plasmonic band edge of a metallic grating,” Opt. Express
20(2), 1868–1877 (2012).
18. Z. Yu, P. Deshpande, W. Wu, J. Wang, and S. Y. Chou, “Reflective polarizer based on a stacked double-layer
subwavelength metal grating structure fabricated using nanoimprint lithography,” Appl. Phys. Lett. 77(7), 927–
929 (2000).
19. W. Zhou, K. Li, C. Song, P. Hao, M. Chi, M. Yu, and Y. Wu, “Polarization-independent and omnidirectional
nearly perfect absorber with ultra-thin 2D subwavelength metal grating in the visible region,” Opt. Express
23(11), A413–A418 (2015).
20. Y. Yu, C. Sun, J. Li, and X. Deng, “A plasmonic metal grating wavelength splitter,” J. Phys. D Appl. Phys.
48(1), 015102 (2015).
21. Y. Ye, Y. Zhou, and L. Chen, “Color filter based on a two-dimensional submicrometer metal grating,” Appl.
Opt. 48(27), 5035–5039 (2009).
22. H.-S. Lee, Y.-T. Yoon, S. S. Lee, S.-H. Kim, and K.-D. Lee, “Color filter based on a subwavelength patterned
metal grating,” Opt. Express 15(23), 15457–15463 (2007).
23. Y. Wu, B. Yao, Y. Cheng, Y. Rao, Y. Gong, X. Zhou, B. Wu, and K. S. Chiang, “Four-wave mixing in a
microfiber attached onto a graphene film,” IEEE Photonics Technol. Lett. 26(3), 249–252 (2014).
24. Z. Hong, L. Zhou, X. Li, W. Zou, X. Sun, S. Li, J. Shen, H. Luo, and J. Chen, “Design and analysis of a highly
efficient coupler between a micro/nano optical fiber and an SOI waveguide,” Appl. Opt. 51(16), 3410–3415
(2012).
25. L. Tong, F. Zi, X. Guo, and J. Lou, “Optical microfibers and nanofibers: A tutorial,” Opt. Commun. 285(23),
4641–4647 (2012).
26. L. Tong, R. R. Gattass, J. B. Ashcom, S. He, J. Lou, M. Shen, I. Maxwell, and E. Mazur, “Subwavelength-
diameter silica wires for low-loss optical wave guiding,” Nature 426(6968), 816–819 (2003).
27. C.-L. Zou, F.-W. Sun, C.-H. Dong, Y.-F. Xiao, X.-F. Ren, L. Lv, X.-D. Chen, J.-M. Cui, Z.-F. Han, and G.-C.
Guo, “Movable fiber-integrated hybrid plasmonic waveguide on metal film,” IEEE Photonics Technol. Lett.
24(6), 434–436 (2012).
28. K. O. Hill and G. Meltz, “Fiber Bragg grating technology fundamentals and overview,” J. Lightwave Technol.
15(8), 1263–1276 (1997).
29. K. S. Thyagarajan and A. Ghatak, Fiber Optic Essentials(John Wiley & Sons, 2007), Chap 11.
30. A. Othonos, “Fiber Bragg gratings,” Rev. Sci. Instrum. 68(12), 4309–4341 (1997).
31. A. D. Rakić, A. B. Djurišic, J. M. Elazar, and M. L. Majewski, “Optical properties of metallic films for vertical-
cavity optoelectronic devices,” Appl. Opt. 37(22), 5271–5283 (1998).
32. J. Chen, X. Shen, Z. Hong, and X. Li, “Nanostructure optic-figer-based devices for optical signal processing,” in
OptoElectronics and Communications Conference (IEEE, 2010), 550–551 (2010).

1. Introduction
Surface plasmon-polaritons (SPP) [1–7] are attracting considerable attentions due to the field
intensity enhancement at the interface and the light confinement ability beyond the diffraction
limit. SPPs can be excited in a subwavelength scale and guided at the interface between metal
and dielectric, providing a promising solution to realize an optical device unbounded by the
diffraction limit [8–12]. Sub-wavelength metal gratings with the SPP or hybrid SPP-dielectric
modes excited close to the metal surfaces [13–15] have many potential applications in making
functional optical components. The metal gratings have been widely used in band-pass filters
[16], chemical and biological sensors [17], reflective polarizers [18], polarization-independent
and omnidirectional absorbers [19], and color filters [20–22] etc.
In this paper, we report a Bragg grating composed of a nanofiber situated on an array of
gold (Au) strips. A hybrid plasmonic mode is excited due to the coupling between the
nanofiber dielectric mode and the metal SPP mode. The nanofiber can be freely moved on the
metal grating surface, making it flexible to adjust the coupling position, length and angle. We
investigate the device performance both theoretically and experimentally. The paper is
organized as follows. In section 2, we introduce the device design and simulation results.

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9317
Section 3 describes the device fabrication and experimental results. Section 4 concludes our
work.
2. Device structure and simulation results
As the nanofiber is very gentle and flexible, it could be bent to rest on top of the metal
grating. The metal grating is made of an alternative array of metal strips and air slits as
schematically shown in Fig. 1. Once the nanofiber touches the metal surface, it will firmly
adhere to the surface through Van der Waals and electrostatic forces [23]. A good merit of
such a grating structure lies in its easy mobility and re-configurability. The nanofiber can be
drawn from a standard single mode fiber (SMF) [24-26], resulting in low loss connection to
other fiber-based devices and systems. With an adiabatic taper, the conversation efficiency
between the SMF and the nanofiber can exceed 99% [27]. The metal grating is formed on top
of the SiO2 substrate. We denote the width, length and thickness of the metal strip as d, L and
h, respectively, the width of the air slit as a, the radius of the nanofiber as R and its alignment
angle with respect to the grating longitudinal direction as θ. Therefore, the grating period
along the nanofiber is Λ = (a + d) /cos(θ) with a filling factor of f = d/(a + d). The coupling
length between nanofiber and metal grating is denoted as Lc, which can be expressed by the
number of grating period N as Lc = N·Λ.

Fig. 1. Schematic diagrams of the hybrid plasmonic Bragg grating. (a) Top view; (b) Side
view; (c) Perspective view.

The coupled system is essentially a Bragg grating with the waveguide effective index
periodically perturbed by the underlying metal strips. The Bragg wavelength λΒ given by [28]

λB = 2neff Λ = 2 ( dneff 1 + aneff 2 ) cos (θ ) (1)

where neff = (dneff1 + aneff2)/(d + a) is the average effective index, neff1 and neff2 are the modal
effective indices of the nanofiber loaded with and without the metal strip, respectively. The
effective index is dependent on the nanofiber radius. The reflection bandwidth Δλ is given by
[29]
λ2
(κ ac Lc ) + π 2
2
Δλ = (2)
π neff Lc

where κac is the coupling coefficient. If κacLc << π, then the bandwidth is an inverse function
of the grating length as

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9318
λ2
Δλ ≈ (3)
neff Lc

The expression for the peak reflectivity is given by [30]


R ( Lc ) = tan h 2 (κ ac Lc ) (4)

The reflectivity increases with the coupling length. When the coupling length is long enough
(i.e., Lc > π/κac), it gives nearly total reflection.
We first use the 3D finite-difference time-domain (FDTD) simulation to numerically study
the device performance. The permittivity of Au is assumed to be εAu = −95.92 + 10.97i around
1550 nm wavelength [31]. The silica fiber has a permittivity of εsilica = 2.08. The height of the
Au film is set as h = 500 nm. Figure 2 shows the normalized transmission and reflection
spectra of the device with Λ = 650 nm, R = 0.7 μm, f = 0.4, N = 50 and θ = 0 for the TM and
TE polarizations. Here, the TM polarization is defined as the one whose major electric field is
perpendicular to the Au film, and the TE polarization is the one with its major electric field
parallel to the metal film. Both TM and TE spectra show two Bragg reflection bands, and yet
the TM one has much stronger peak reflection (<3 dB loss) and slightly longer central
wavelengths than the TE one. The difference could be attributed to the hybrid plasmonic
mode excitation for TM mode, which leads to the much stronger interaction between the light
field and the metal grating. In our following simulation study, we only focus on the TM mode
as it provides a superior performance than the TE mode.

Fig. 2. Simulated transmission and reflection spectra of the nanofiber-attached hybrid


plasmonic grating for (a) TM and (b) TE polarizations.

The performance of the grating depends on multiple device parameters, including the
grating period, the filling factor, the fiber radius, and the coupling length, etc. We further
study the effect of these design parameters on the grating reflection spectrum. The results are
shown in Fig. 3. We only vary one parameter each time while the others are kept the same as
those used above. The reflection peaks shift to longer wavelengths when the grating period
increases, which is expectable from the Bragg wavelength formula given by Eq. (1). The
change of filling factor also shifts the two peaks. In particular, the left one is shifted by 30 nm
and right one 20 nm when the filling factor increases from 0.3 to 0.6. The nanofiber radius
affects both of the Bragg reflection wavelengths and the peak profiles (width and height).
When the radius is small, e.g., R = 0.4 μm, the left peak is very low and broad. It gradually
grows up with the increasing radius. Finally, the number of grating periods or the coupling
length also affects the reflection peak profile. A longer coupling length results in stronger
peak reflection and a narrower bandwidth, which is consistent with Eqs. (2) and (4). It is
worth noting that the coupling length has little effect on the reflection spectrum when it
becomes long enough (N > 65), as light is fully reflected by the front part of the grating. The
simulation reveals that a very short metal grating (10’s microns) can induce a strong reflection

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9319
owning to the excitation of the hybrid plasmonic mode, which is favorable in making an ultra-
compact device.

Fig. 3. Simulated reflection spectra in response to various design parameters: (a) grating
period, (b) filling factor, (c) nanofiber radius, and (d) number of grating periods.

To better understand the Bragg reflection behavior of the device, we simulate the steady-
state light electric-field intensity pattern in the device at the two Bragg wavelengths for the set
of parameters used in Fig. 2. The results are shown in Fig. 4. Clear standing wave patterns are
observed with its period consistent with the metal grating period. Light intensity reduces
along the nanofiber axial direction with partial light absorbed by the metal and partial light
reflected back to the incident port. A difference is also discernable from the field patterns at
the two Bragg wavelengths. At the longer wavelength of 1650 nm, the standing wave node is
located above the metal strip while the antinode is located in the trench. The reflected light is
the fundamental TM mode of the nanofiber. In contrast, at the shorter wavelength of 1525 nm,
the field pattern exhibits a zig-zag path where the light wave node is pushed outwards above
the metal strip and pulled inwards the substrate in the trench. Examination of the cross-
sectional field pattern of the reflected light beyond the light launch point reveals that a first-
order TM mode is excited by the metal grating. As the first-order TM mode is close to cut-off,
its reflection peak therefore gradually disappears when the nanofiber radius reduces.

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9320
Fig. 4. (a) Light electric-field intensity pattern (|E|2) in the x-y plane at the 1650 nm
wavelength. Nanofiber fundamental mode is launched from the intersection plane at x = −17
μm. The Inset shows the cross-sectional distribution of the reflected light intensity. (b)
Magnified view showing the standing wave light field pattern at 1650 nm. The cross-sectional
field patterns along the dashed lines (i) and (ii) are also illustrated. (c) Light electric-field
intensity pattern in the x-y plane at the 1525 nm wavelength. (d) Magnified view of the light
field pattern at 1525 nm.

3. Device preparation and experiments


Figure 5(a) shows the fabrication process flow of the metal grating. We started with a silicon
wafer coated with a thick layer of silicon dioxide. It was first deposited with a 500-nm-thick
Au film using e-beam evaporation. A layer of 440-nm-thick e-beam resist (Zep 520) was spin-
coated on the gold film and baked on a hotplate at 180°C for 90s. The grating was patterned
on the resist layer by electron-beam lithography (EBL). The metal layer was then dry etched
by ion-beam etching (IBE) with the resist serving as the etching soft-mask. The ion-beam
energy was 550 eV and the chamber pressure was 5 × 10−4 Pa. The etch rates of the gold and
resist films are 45 nm/min and 22 nm/min, respectively. After dry etch, the remained e-beam
resist was subsequently removed by N-Methyl-2-Pyrrolidinone (NMP) and acetone. The
entire fabrication was done at the Center for Advanced Electronic Materials and Devices
(AEMD) platform of our university.
Figure 5(b) shows the scanning electron micrograph (SEM) image of the metal grating.
The white and dark lines are the Au strips and air slits, respectively. The grating period was

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9321
measured to be 651.3 nm and the air slit width 260.5 nm. Thus the filling factor is 0.6. In
order to make it more convenient to attach the nanofiber to the surface of metal grating, the
grating pattern was designed to be large enough with an area of 1 mm × 1.5 mm.

Fig. 5. (a) Fabrication process flow of the metal grating. (b) SEM image of the metal grating.

The nanofiber was drawn from a standard SMF using the custom nanofiber drawing
platform [32]. The parameters, such as the drawing speed and the resulting nanofiber section
length, were set in advance. The radius of the nanofiber was measured from the SEM images.
A flat spectrum with no ripples was measured, indicating that the mode transition from the
SMF to the nanofiber was smooth without excitation of back reflection. The measured
insertion loss of the nanofiber was less than 1 dB.
The coupling of the nanofiber to the metal grating was realized as follows. The fiber was
first clamped and pulled straight by two high-precision translation stages. The metal grating
chip was held by another translation stage and positioned beneath the nanofiber. We then
moved the fiber clamp stages closer so that the nanofiber gradually bent downward. Next, we
moved the chip up to approach the nanofiber. The nanofiber can tightly cling to the metal
grating surface once they were touched together due to the Van der Waals and electrostatic
forces. The contact length was decided by the vertical position of the chip and the distance
between the fiber clamp stages. The actual contact length was measured from the images
taken by the top and backside CCD cameras. The nanofiber alignment angle was adjustable
by rotating the chip.
Figure 6 shows the experimental setup to characterize the device. A broadband light
emitting diode (LED) source was used and the output light was polarized by using a
polarization beam splitter (PBS). A polarization controller was followed to control the light to
TM polarization before entering the grating. The transmission power was monitored by an
optical power meter. The reflection spectra were recorded by an optical spectrum analyzer
(OSA). A 3-dB coupler was put in front of the device in order to separate the reflected light.
Although it has a 3-dB extra loss, it has very high isolation of 63 dB (much higher than that of
a typical optical circulator with 40 dB isolation).

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9322
Fig. 6. Experimental setup to characterize the device. PC: polarization controller; PBS:
polarization beam splitter; OSA: optical spectrum analyzer.

Figure 7(a) shows the measured reflection spectra for different nanofiber radii ranging
from 0.45 μm to 0.73 μm. The coupling length is around 0.8 mm. With the increase of the
nanofiber radius, the filter central wavelength experiences red-shift as the larger nanofiber
radius increases the effective index. The peak has a 3-dB bandwidth of 15 nm and out-of-band
rejection of >30 dB. The change trend is consistent with the simulation results shown in Fig.
3(c). The peak reflection loss is about 10 dB, caused by the metal absorption and scattering. It
should be noted that a second reflection peak is also discernable for R = 0.7 μm and 0.73 μm.
Compared to the simulation, it is much weaker perhaps because of the high bending loss. We
also studied the influence of the nanofiber alignment angle on the filter performance as shown
in Fig. 7(b). It can be seen that the central wavelength of the reflection band slightly increases
with the angle and meanwhile the peak broadens with a reduced height. Such a change trend
could be understood as follows. A larger angle is equivalent to an increased grating period
and a stronger coupling coefficient. According to Eqs. (2) and (4), the reflection bandwidth
increases with a reduced peak power. Finally, we also investigated the effect of coupling
length. As stated above, the coupling length can be changed by moving the chip vertically.
Figure 7(c) shows the reflection spectra for four coupling lengths. The nanofiber radius is
0.45 μm. The coupling length is an appropriate value estimated from the microscope image.
The measurement reveals that the reflection spectrum is relatively stable and almost not
changeable with the coupling length. It hence suggests that the reflection mainly occurs at the
front of the metal grating. The effective interaction length is less than the shortest contact
length in the experiment. Due to the strong attractive force existing between the nanofiber and
the metal grating, we were unable to further shorten the contact length.

Fig. 7. Measured input-normalized TM reflection spectra for various device parameters: (a)
nanofiber radius, (b) nanofiber alignment angle and (c) coupling length.

The nanofiber coupled Bragg grating can work as a bandpass optical filter for
multiplexing/de-multiplexing wavelength division multiplexing (WDM) optical signals. In
this regard, its function is similar to a conventional fiber Bragg grating, but the grating length
is much shorter and the passband is much broader in our device because of the strong
coupling enabled by the metal grating. Another key feature that we would like to highlight is

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9323
that strong optical field is localized outside the nanofiber, especially in the air between the
nanofiber and metal grating, which makes it quite attractive in optical sensing applications.
The enhancement of optical field at the metal interface can also be exploited to study the
light-matter interaction when a thin layer of material is coated on the metal. For example, a
graphene can be sandwiched between the nanofiber and the metal grating to facilitate the
excitation of four-wave mixing in graphene.
4. Conclusions
We have presented a flexible hybrid plasmonic grating composed of a nanofiber situated on
an array of metal strips. The nanofiber is bent to touch and couple with the metal grating. The
TM mode has much strong reflection than the TE mode. Numerical simulations show that the
reflection band is strongly dependent on the nanofiber radius and metal grating period. The
interaction length affects the peak reflectivity and bandwidth. A higher order mode nanofiber
mode could be excited upon reflection by the grating even when the incident light is the
fundamental mode. We experimentally realized such a hybrid plasmonic grating device and
characterized the device with different nanofiber radii, alignment angles and coupling lengths.
A typical device exhibits a strong reflection peak with a 3-dB bandwidth of 15 nm and out-of-
band rejection of more than 30 dB. The reflection spectrum is almost unchangeable when the
coupling length is long enough, facilitating to form a stable and repeatable device. The
flexibility in nanofiber attachment makes it cost-effective to build such a hybrid plasmonic
grating device. It could find potential applications in optical filters and biochemical sensors.
Acknowledgment
This work was supported in part by the National Natural Science Foundation of China
(NSFC) (61422508), the Shanghai Rising-Star Program (14QA1402600).

#258679 Received 1 Feb 2016; revised 8 Apr 2016; accepted 15 Apr 2016; published 20 Apr 2016
© 2016 OSA 2 May 2016 | Vol. 24, No. 9 | DOI:10.1364/OE.24.009316 | OPTICS EXPRESS 9324

You might also like