Sie sind auf Seite 1von 3

September 1, 2007 / Vol. 32, No.

17 / OPTICS LETTERS

2499

Surface corrugation Bragg gratings on optical fiber


tapers created via plasma etch postprocessing
W. Ding,* S. R. Andrews, and S. A. Maier
Department of Physics, University of Bath, Bath BA2 7AY, UK
*Corresponding author: pypwd@bath.ac.uk
Received March 26, 2007; revised June 11, 2007; accepted June 27, 2007;
posted July 24, 2007 (Doc. ID 81500); published August 16, 2007
We describe a reflection-based fiber filter fabricated by plasma etching a surface corrugation Bragg grating
on a tapered single-mode fiber. The taper waist with the grating forms the functional part of the filter, and
the adiabatic taper transition removes unwanted higher-order modes. The spectral response is controlled by
varying the taper diameter while maintaining a constant grating period. Reflection spectra have been investigated theoretically and experimentally and found to be in good agreement. 2007 Optical Society of
America
OCIS codes: 060.2340, 230.1480, 230.4000.

With the extensive use of optical fibers in telecommunication systems, in-line fiber devices are receiving
more attention than free-space optical devices due to
their intrinsic compatibility with fiber links. One
such example is a tapered fiber in a directional coupler [1]. With decreasing taper diameter, the evanescent field tail of the guided mode spreads out across
the silicaair interface, and the buried waveguide of
the untapered fiber is converted to an open waveguide. Although open waveguides can also be created
by fiber side polishing [2] or drawing D-shaped fibers
[3], fiber tapering provides an ideal transition with
low excess loss [4]. The conservation of cylindrical
symmetry also retains the mode structure of the untapered fiber.
The functionality of fiber tapers can be enhanced
by incorporating longitudinal Bragg gratings. However, the standard use of photosensitive fiber core
and UV writing is inconvenient for a micrometerscale fiber taper [5] due to the disappearance of the
germanium core [6] and the decrease of the field overlap with the core. Alternative techniques, such as
metal deposition [7], focused ion beam milling or implantation [8], and femtosecond infrared irradiation
[9], have been reported, although the last two methods managed to produce only third-order gratings. In
this Letter, we present a method based on creating a
surface corrugation grating with a submicrometer period by interference lithography and reactive ion
etching. At the current stage of development, the etch
depth is limited by the mechanical robustness of the
fiber taper. The shallow etch depth that we have
achieved, together with the modest permittivity contrast between silica and air and relatively short grating length presently precludes transmission-filter applications. In this Letter, we concentrate on studying
the experimental and theoretical reflection spectra.
In addition, we present a way to flexibly manipulate
the reflection spectral properties by writing a uniform grating on a taper with a variable diameter profile. The wavelength dependence of the reflectivity of
this device is discussed.
Figure 1 shows schematically the working principle of a surface corrugation grating fiber taper. In
0146-9592/07/172499-3/$15.00

the waist part, the forward-propagating fundamental


mode is converted from the core mode of the untapered, single-mode fiber via the adiabatic transition
[4]. This mode is coupled to a variety of backwardpropagating modes by the grating in the waist.
Among those backward-propagating waves, only the
fundamental mode is coupled back into the core of
the input fiber due to the adiabatic transition, while
the higher-order modes are converted into cladding
modes of the untapered fiber and are strongly
damped at the lossy interface between the silica and
polymer coating. In this way, the grating taper waist
acts as the functional part of a reflection device and
the adiabatic transition as a filter of unwanted
higher-order modes.
Considering only the backward-propagating fundamental mode, the resonant condition for a Bragg
grating is
R = 2n0D,R,

where R is the resonant wavelength, is the grating


period, D is the taper diameter, and n0 is the effective
modal index of the fundamental mode.
Equation (1) suggests that the spectral properties
of the grating fiber taper can be modified by introducing spatial variations in either the grating period or
the effective modal index. In this work, we adopt the
latter approach and exploit the dependence of the effective modal index on taper diameter. A uniform period grating was therefore etched into a varying diameter taper waist produced using a precise fiber
tapering station [10]. This method is simpler and
more flexible than producing variable period gratings
[11].

Fig. 1. Schematic showing the working principle of the


surface grating fiber taper.
2007 Optical Society of America

2500

OPTICS LETTERS / Vol. 32, No. 17 / September 1, 2007

Figures 2(a) and 2(b) show the dependence of the


effective index of the fundamental mode and the coupling coefficient of the surface corrugation grating on
taper diameter. The grating consists of an array of
semicircular grooves, as shown in Fig. 2(c). In these
calculations, the wavelength is 1.5 m, and the etch
depth is 100 nm. The coupling coefficient between the
forward- and backward-propagating fundamental
modes is given by [12]

= ic/4

* x,y a dA,
x,yefx,y h
z
b

2a
1/2

*x,y a dA = 1,
efx,y h
z
f

2b

where is the permittivity perturbation caused by


the grating, c and are, respectively, the speed of
light and the permeability in vacuum, is the angular frequency, az is the unit vector along the fiber
axis, the subscripts f and b represent the forwardand backward-propagating fundamental modes, and
the caret denotes unit field normalization.
From Figures 2(a) and 2(b), it is clear that only
when the taper diameter decreases to the micrometer
scale does the diameter dependence become significant. For example, when the taper diameter is reduced from 10 to 6 m, the change in effective modal
index, n0, is 0.007, which is bigger than the refractive
index difference between the core and the cladding of
a conventional Corning SMF-28 fiber. Such a large
adjustment range provides freedom to manipulate
the resonant condition in Eq. (1).
In the fabrication process, we used interference lithography and plasma etching. A 25 nm thick NiCr
layer was deposited on an approximately hemicylindrical section of a 10 m diameter fiber taper by
thermal evaporation. A layer of photoresist was then
applied by dip coating, and the metal side was exposed to the interference pattern produced by two
beams of 406.7 nm light [7]. During exposure, the
NiCr layer prevents light from penetrating the fiber,
which would otherwise degrade the quality of the
grating by multiple reflections. The resist thickness
depended on fiber diameter but was typically
200 nm. After resist development, a solution of cerium ammonium nitrate and acetic acid was used to
etch away NiCr between the resist strips. All but a
central portion of the patterned fiber was then
masked by further dip coatings. The grating was

Fig. 2. Taper diameter dependence of (a) the effective


modal index and (b) the coupling coefficient [Eq. (2a)]. (c)
Schematic cross section of the grating fiber taper.

transferred to this section of fiber by reactive ion


etching in an inductively coupled CHF3 plasma. After
that, the resist and the residual NiCr were removed.
We first discuss a taper with constant waist diameter (denoted as type I). Figure 3(a) shows a scanning
electron micrograph of the etched side of the grating
taper. The taper waist and taper transitions are both
20 mm long. The 10 mm long grating is located in
the waist. The taper diameter and grating period
were measured using a calibrated scanning electron
microscope (SEM) to be 9.7 m and 522 nm, respectively. From knowledge of the large area etch rate,
the etch depth was estimated to be between 150 and
200 nm. Figure 3(b) shows the reflectivity and transmittance spectra measured with a resolution of
0.5 nm using an optical circulator and an ultrabroadband supercontinuum source from an endlessly
single-mode photonic crystal fiber.
The top curve in Fig. 3(b) shows the transmission
spectrum. There is no discernable dip, and the transmittance is 2 dB throughout the measured range.
This weak coupling is partly due to the fact that the
fundamental-mode field is strongly confined along
the taper axis because of the relatively large diameter and is therefore significantly reduced at the
silicaair interface. In the reflection spectrum, i.e.,
the bottom curve in Fig. 3(b), there are multiple
peaks arising from coupling of the forwardpropagating fundamental mode to different
backward-propagating modes. The resonant wavelengths of the mode couplings [12], calculated using
the measured values D and , are indicated in Fig.
3(b). The reflection peaks agree quite well with the
calculation. The full width at half-maximum of the
main peak LP01 is 0.6 nm. There is evidence that
the main peak is split due to the polarization dependence associated with the asymmetry of the grating.
Compared with the main peak, the strength of the
coupling to higher-order modes is weaker by at least
10 dB because of the filtering effect of the adiabatic
taper transition, as described previously. An even
higher suppression of 20 dB was obtained in another
sample prepared with a longer, therefore more adiabatic, taper transition of length 25 mm.

Fig. 3. (a) Scanning electron micrograph of the surface


corrugation Bragg grating fiber taper. (b) Reflectivity and
transmittance spectra of the sample of type I. The calculated resonant wavelengths are marked with arrows.

September 1, 2007 / Vol. 32, No. 17 / OPTICS LETTERS

We now consider tapers with spatially varying


waist diameter Dz, and derive the reflectivity spectra as follows. The forward- and backwardpropagating fundamental modes along the taper can
be expressed as

Ef,b = Af,bzexp i

0,L

0zdz ef,bx,y,z, 3

where ef,b represent the normalized mode electric


field, 0z is the propagation constant, and Af,bz is
the slowly varying field amplitude. The grating exists
from z = 0 to L. In the weak coupling approximation,
Afz Afz = 0 A0 and Abz  A0, as verified by
Fig. 3(b).
The coupling equation is given by [12]
dAbz
dz

A0 z
n0z

ei0 0zdz exp i2

0zdz

where is from Eq. 2. The reflectivity of the grating


taper is then calculated by R = Abz = 0 / A02 and is
a function of the taper diameter profile, Dz.
We fabricated two variable diameter tapers, one
with a V-shaped profile (type II) and one with a linearly varying diameter (type III). Uniform period
gratings were created with an estimated etch depth
of 100 nm and grating periods of 522 nm (type II)
and 534 nm (type III). A series of SEM images were
taken along the length of the grating tapers to determine the diameter profiles, which are shown in Figs.
4(a) and 4(b). Figures. 4(c) and 4(d) show the measured (Mea.) and calculated (Cal.) reflectivity spectra, which qualitatively agree quite well. The discrepancy in the long wavelength cutoffs is probably due to
underestimating the value of L. In practice, the grating begins and ends gradually because of variations
in the thickness of the resist used to define the grating region. Outside the nominal grating region, the
taper diameter and modal index are larger, thus leading to a longer resonant wavelength. The short wavelength cutoff, on the other hand, is simply deter-

Fig. 4. Diameter profiles of the taper samples of (a) type II


and (b) type III in their grating regions. (c) and (d) show the
measured (Mea.) and calculated (Cal.) reflectivity spectra
for samples II and III, respectively. The calculated curves
have been moved downward by 20 dB for clarity.

2501

mined by the minimum taper diameter. Figure 4 also


shows that the nonuniform taper waists lead to a
substantial broadening of the reflection peaks from
0.6 (type I) to 5.2 (type II) and 3.7 nm (type III).
Figure 4(c) shows that the V-shaped taper diameter profile of sample II produces multiple dips in the
spectrum. The reflection of light from the two sides of
the V-shaped grating waist result in strong interference fringes in the spectrum that strongly depend on
the precise fiber profile and are thus difficult to reproduce exactly in the calculation. In the calculation
shown in Fig. 4(c), we used the measured grating period = 522 nm and taper diameter profile [Fig.
4(a)] and took the etch depth to be 100 nm. The
etched grooves were assumed to extend over half the
fiber surface with a width equal to half of the grating
period.
For sample III, interference between reflections
from the two sides of the grating region disappears.
However, undulations still exist in the spectrum in
Fig. 4(d). This is because the effective reflection region along the grating, and thus the reflection intensity, depends on wavelength. The calculation in Fig.
4(d) used a grating period of = 534 nm, an etch
depth of 100 nm, and the taper diameter profile
shown in Fig. 4(b).
In conclusion, we have created surface corrugation
Bragg gratings on micrometer-scale fiber tapers using interference lithography and plasma etching. By
writing a uniform grating on a varying diameter
taper, we can manipulate the reflectivity spectrum in
a predictable way by controlling the geometry. This
wavelength dependence of the reflectivity can be further enhanced by increasing the grating length and
etch depth, and this may lead to applications in point
probe evanescent field sensing of, for example, refractive index or temperature.
This work was supported by the Engineering and
Physical Sciences Research Council.
References
1. T. Bricheno and A. Fielding, Electron. Lett. 20, 230
(1984).
2. R. A. Bergh, G. Kotler, and H. J. Shaw, Electron. Lett.
16, 260 (1980).
3. R. B. Dyott and P. F. Schrank, Electron. Lett. 18, 980
(1982).
4. J. D. Love and W. M. Henry, Electron. Lett. 22, 912
(1986).
5. D. M. Hernandez, J. Mora, P. P. Millan, A. Diez, J. L.
Cruz, and M. V. Andres, Appl. Opt. 43, 2393 (2004).
6. K. O. Hill, Y. Fujii, D. C. Johnson, and B. S. Kawasaki,
Appl. Phys. Lett. 32, 647 (1978).
7. W. Ding, S. R. Andrews, T. A. Birks, and S. A. Maier,
Opt. Lett. 31, 2556 (2006).
8. V. Hodzic, J. Orloff, and C. C. Davis, J. Lightwave
Technol. 22, 1610 (2004).
9. D. Grobnic, S. J. Mihailov, H. Ding, and C. W. Smelser,
IEEE Photon. Technol. Lett. 18, 160 (2006).
10. T. A. Birks and Y. W. Li, J. Lightwave Technol. 10, 432
(1992).
11. M. G. Xu, L. Dong, L. Reekie, J. A. Tucknott, and J. L.
Cruz, Electron. Lett. 31, 823 (1995).
12. T. Erdogan, J. Lightwave Technol. 15, 1277 (1997).

Das könnte Ihnen auch gefallen