Download A minimally invasive in-fibre Bragg grating sensor

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Optical tweezers wikipedia , lookup

Ultraviolet–visible spectroscopy wikipedia , lookup

Astronomical spectroscopy wikipedia , lookup

Optical attached cable wikipedia , lookup

Diffraction grating wikipedia , lookup

Birefringence wikipedia , lookup

Optical amplifier wikipedia , lookup

Opto-isolator wikipedia , lookup

Fiber Bragg grating wikipedia , lookup

Harold Hopkins (physicist) wikipedia , lookup

Transcript
Super-structured D-shape fibre Bragg grating contact force sensor
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
Superstructured DShape sensor
Christopher R. Dennison and Peter M. Wild
Department of Mechanical Engineering, University of Victoria, British Columbia, Canada
Corresponding Author:
Chris R. Dennison
University of Victoria
Department of Mechanical Engineering
P.O. Box 3055
Victoria, B.C.
V8W 3P6
Ph: (250) 853-3198
Fax: (250) 721-6051
e-mail: [email protected]
Abstract
1
Super-structured D-shape fibre Bragg grating contact force sensor
48
49
1. Introduction
In-fibre Bragg gratings (FBGs) can be configured as sensors for various parameters
50
including displacement [1], strain [2], temperature [3], pressure [3], contact force [4],
51
humidity [5], and radiation dose [6] among others. FBGs are an attractive alternative to
52
other piezoelectric, resistive or other solid-state sensing technologies because they are:
53
small (125 μm diameter and smaller), biocompatible, mechanically compliant, chemically
54
inert, resistant to corrosive environments, immune to electromagnetic interference, and
55
are capable of simultaneous multi-parameter sensing when suitably configured [1, 7-11].
56
Moreover, multiple FBG sensors can be multiplexed along a single optical fibre thereby
57
allowing spatially distributed measurements [12]. FBG-based contact force sensors
58
typically comprise Bragg gratings in birefringent optical fibre.
59
Birefringent fibre has different refractive indices along two orthogonal, or
60
principal, directions that are commonly referred to as the slow (higher index) and fast
61
(relatively lower index) axes. The difference in the refractive indices along the slow and
62
fast axis is termed the birefringence. A Bragg grating in a birefringent fibre reflects two
63
spectra, one polarized along the fast axis and one along the slow axis, with centre
64
wavelength of each spectrum given by the Bragg condition for reflection [12]. The centre
65
wavelength of each spectrum is a function of, among other parameters, the refractive
66
index of the fast and slow axis.
67
The physical mechanism causing sensitivity to contact force is stress-induced
68
changes in birefringence [13] that are governed by the stress-optic effect [14]. Contact
69
forces on the fibre induce stresses (or strains), the principal directions of which are
70
uniform in the region of the fibre core. However, the magnitudes of the principal stresses
71
are a function of the magnitude of the contact force [4]. Therefore, when the contact
2
Super-structured D-shape fibre Bragg grating contact force sensor
72
force magnitude changes, a predictable change in the principal stress magnitudes and,
73
therefore, birefringence is induced. These force-induced changes in birefringence cause
74
predictable changes in the spectrum reflected by the Bragg grating, including shifts in the
75
Bragg spectra corresponding to the fast and slow axes of the fibre. Conversely, applied
76
axial forces/strains and changes in fibre temperature cause relatively lower, in the case of
77
certain fibres insignificant, changes in fibre birefringence [12]. Therefore, one advantage
78
of Bragg grating sensors in birefringent fibres is that force measurements are not
79
confounded by changes in axial strain and temperature. More specifically, D-shaped
80
optical fibre has been shown to experience virtually no change in fibre birefringence with
81
changes in fibre temperature [4] or changes in axial strain that can result from, among
82
other influences, fibre bending [15].
83
Bragg gratings in birefringent fibres possess higher sensitivity to contact force
84
than gratings in non-birefringent fibres because they are typically constructed with stress
85
concentrating features embedded in the fibre clad, or clad geometries, that increase force-
86
induced changes in birefringence [12]. These features are generally referred to as stress-
87
applying parts. Each commercially available birefringent fibre possesses unique stress-
88
applying parts or clad geometry. Therefore, the contact force sensitivity of each type of
89
birefringent fibre is unique. Contact force sensitivity is also a function of the orientation
90
of the stress-applying parts relative to the direction of load [4]. Therefore, one limitation
91
of birefringence-based contact force sensors is the necessity to control fibre orientation
92
with respect to the contacting surface applying the force.
93
A number of researchers have studied the relationship between contact force and
94
birefringence for several types of birefringent fibres. Udd et al. (1996) [16] applied Bragg
3
Super-structured D-shape fibre Bragg grating contact force sensor
95
gratings in 3M, Fujikura and Corning birefringent fibres to simultaneously measure
96
temperature and force-induced axial and transverse strain. Wierzba and Kosmowski
97
(2003) [17] applied Side-Hole birefringent fibre to force measurements. Chehura et al.
98
(2004) [4] investigated the force sensitivity of D-shape, Elliptical core, TruePhase,
99
Panda, Bow Tie and Elliptical clad birefringent fibres as a function of fibre orientation.
100
More recently, Abe et al. (2006) [18] measured force using chemically-etched Bow Tie
101
and Elliptical clad birefringent fibres to understand the influence that reduction in fibre
102
diameter has on fibre birefringence and force sensitivity.
103
In all of the studies outlined above, fibre orientation relative to the contact force
104
was controlled. However, there are applications for FBG-based contact force
105
measurements where fibre orientation is difficult or impossible to control. One example
106
application from the authors’ previous work is contact force measurements between the
107
articulating surfaces of cartilage within intact human joints such as the hip [19]. Contact
108
force measurements over the cartilage surfaces of the hip, and other articular joints, can
109
indicate the mechanics of the joint, which is closely linked to the etiology of joint
110
degeneration and osteoarthritis [20]. In an intact human hip, the contoured cartilage
111
surfaces of the femoral head and acetabulum are covered by a synovial fluid layer and
112
sealed by a fibrous joint capsule. The presence of synovial fluid within the joint prevents
113
permanent fixation of sensors to cartilage surfaces with conventional adhesives or
114
mechanical fixation. Because fixation is impossible, the orientation of fibre-based sensors
115
with respect to the cartilage surfaces cannot be controlled. Moreover, because the joint
116
capsule surrounds the entire joint and prevents optical access, it is impossible to orient
4
Super-structured D-shape fibre Bragg grating contact force sensor
117
sensors relative to cartilage surfaces using the techniques applied in the above
118
applications of birefringent fibre.
119
To avoid the challenge of controlling sensor orientation, a non-birefringent Bragg
120
grating contact force sensor, with orientation-independent force sensitivity, was
121
developed and applied ex vivo to intact cadaveric hips. This sensor addresses limitations
122
of the current standard film-based sensors [21, 22] for contact force measurements in
123
joints because it can be implanted while leaving the joint capsule and synovial fluid
124
layers intact.
125
The contact force sensor comprises a non-birefringent optical fibre containing a 1
126
mm FBG that is coaxially encased within a silicone annulus and polyimide sheath
127
(outside diameter 240 microns) [19]. Contact forces applied to the outside of the sheath
128
cause compressive strain in both the sheath and silicone annulus surrounding the optical
129
fibre. Because the silicone annulus is approximately incompressible, the annulus volume
130
is conserved by tensile Poisson strain along the axis of the optical fibre containing the
131
FBG. Furthermore, because the magnitude of Poisson strain is directly proportional to the
132
contact force, Bragg wavelength shifts of the FBG within the non-birefringent fibre are
133
directly proportional to contact force.
134
This sensor exhibits orientation independence of sensitivity because of its coaxial
135
configuration. However, like other sensors based on FBGs written in non-birefringent
136
fibre, this sensor is also sensitive to both applied axial strain and temperature changes,
137
which are two mechanical parameters that can confound force measurements on the
138
contoured surfaces of articular joints. More recently, contact force sensors based on tilted
139
Bragg gratings have been proposed that also exhibit orientation independence of
5
Super-structured D-shape fibre Bragg grating contact force sensor
140
sensitivity [23]. However, these approaches are also susceptible to confounding errors
141
associated with axial strain, bending and temperature. (mention limited in multiplexing
142
due to high spectral width?)
143
The objective of the work reported here is to develop a birefringent Bragg
144
grating-based contact force sensor hat addresses limitations of non-birefringent Bragg
145
grating sensors, namely co-sensitivity to axial strain and temperature, and one limitation
146
of Bragg grating sensors in birefringent fibres: the necessity to control sensor orientation.
147
D-shape birefringent fibre is used in the sensor presented because it possesses negligible
148
co-sensitivity to temperature and axial strain.
149
2. Materials and methods
150
151
2.1 Principles of FBGs, birefringence and D-shape optical fibre
FBGs are formed in optical fibres by creating a periodic variation in the refractive
152
index of the fibre core [12, 24]. The length of the FBG and the magnitude and period of
153
the variation in the refractive index determine the optical spectrum that is reflected by the
154
FBG [12, 25]. When light spanning a broad range of wavelengths travels along the core
155
of a non-birefringent fibre and encounters a Bragg grating (Figure 1a), a single-peaked
156
spectrum of wavelengths is reflected. This spectrum is centered at the Bragg wavelength,
157
B , which is given by:
158
B  2n0
(1)
159
where, as shown in Figure 1a,  is the spatial-period of the variation in the refractive
160
index and, n0 , is the effective refractive index of the fibre core [12]. However, when light
161
travels along the core of a birefringent optical fibre and encounters a grating, two single
162
peaked spectra are reflected. One spectrum corresponds to light polarized along the slow
163
axis (s) and the other corresponds to light polarized along the fast axis (f) (Figure 1b):
6
Super-structured D-shape fibre Bragg grating contact force sensor
s  2ns
 f  2 n f
164
(2)
165
where the subscripts s and f denote the slow and fast axis, respectively. In general,
166
because ns and n f differ in a birefringent fibre, the slow and fast axis Bragg wavelengths
167
will be different, and the magnitude of the difference is termed the spectral separation.
168
The spectral separation between the slow and fast axis Bragg wavelengths, s   f is
169
given by:
170
s   f  2(ns  n f )  2B
171
where the refractive index difference between the slow and fast axis is referred to as the
172
birefringence, B .
173
174
175
(3)
For any optical fibre, the birefringence in the fibre core can be expressed as the
sum of three contributions [26]:
B  ns  n f  BG  BIS  BE
(4)
176
where BG is the geometric contribution that, typically, is found only in optical fibres with
177
asymmetric or elliptical cores; BIS is the internal stress contribution that, typically, is
178
found in optical fibres with internal stress applying parts; and BE is the external
179
contribution to fibre birefringence that is typically caused by externally applied contact
180
forces. There are two main classes of birefringent fibre that are distinguished by the
181
predominant birefringence contribution that exists in the fibre core: either internal stress-
182
induced birefringence or geometric birefringence.
183
In fibres with internal stress-induced birefringence, the difference in refractive
184
index is created primarily by thermal residual stresses that are created within the fibre
7
Super-structured D-shape fibre Bragg grating contact force sensor
185
when the optical fibre cools from its drawing temperature to ambient temperature.
186
Typically, there is no geometric contribution to birefringence. The thermal residual
187
stresses are created by different thermal expansion coefficients of the fibre clad, stress
188
applying parts which are embedded in the clad, and core. When the fibre, which initially
189
has uniform temperature throughout its cross section, cools from its drawing temperature
190
to ambient temperature thermal residual stresses evolve because the clad, stress applying
191
parts and core undergo differing volume contractions. The configuration of the stress
192
applying parts results in a uniform principal stress field in the region of the core, which
193
through the stress-optic effect [14] results in a uniform difference in refractive index, or,
194
alternatively, uniform BIS throughout the core. In the context of contact force sensing,
195
applied contact forces cause changes in the birefringence contribution BE and, therefore,
196
the Bragg wavelengths associated with the fast and slow axis of the fibre. The
197
mechanisms underlying changes in Bragg wavelengths will be detailed after the
198
principles of D-shape fibre are discussed. Examples of fibres with stress-induced
199
birefringence include Side-hole, PANDA, and Bow Tie. (Need more refs for this
200
paragraph).
8
Super-structured D-shape fibre Bragg grating contact force sensor
201
202
203
204
205
206
207
208
Figure 1: a) schematic of D-shape fibre showing the clad, core and Bragg grating
comprising regions of modified refractive index, n spaced with period, Λ. b) schematic of
D-shape fibre showing variable names assigned to nominal major and minor outside
diameters of fibre (Dmajor and Dminor) and minimum core offset from clad, r; major and
minor diameters of the elliptical core (dmajor and dminor); and the directions of the fibre
axis, slow axis and fast axis.
209
D-shape optical fibre (Figure 1a and 1b) is one example of the second class of
210
birefringent fibre that is based on geometric birefringence. While the physics and
211
mathematics behind geometric birefringence is beyond the scope of this work, a
212
qualitative discussion of birefringence in D-shaped fibres with elliptical cores will be
213
included.
214
The majority of birefringence in D-shaped fibres is caused by the geometry of the
215
fibre clad and elliptical core [27]. More specifically, the geometric birefringence of an
9
Super-structured D-shape fibre Bragg grating contact force sensor
216
optical fibre with an elliptical core is a function of the core location relative to the air
217
surrounding the fibre, r (Figure 1b); the major and minor diameters of the elliptical core,
218
dmajor and
219
through the core; and the refractive indices of the clad, core and air surrounding the fibre
220
[28]. The internal stress contribution to birefringence is negligible at the wavelengths
221
used for Bragg grating sensing applications [29]. When the major axis of the fibre core is
222
aligned as shown in Figure 1b, the fast axis of the fibre is normal to the flat-side of the D-
223
shape clad (Figure 1b). The slow axis is orthogonal to the fast axis and is aligned with the
224
major axis of the elliptical core (Figure 1b). In the context of force sensing, applied
225
forces cause changes in the birefringence contribution, BE and, therefore, the Bragg
226
wavelengths of the fast and slow axes.
dminor, respectively (Figure 1b); the wavelength of the light propagating
227
There are two mathematical formulations that can be used to calculate changes in
228
the Bragg wavelengths of the fast and slow axes as a function of applied contact forces.
229
The first, or stress-optic, formulation relates changes in contact force-induced principal
230
stresses to changes in fibre birefringence. The changes in birefringence can then be used
231
to calculate spectral separation (Equation 3).
232
The second, or strain-optic, formulation can calculate both spectral separation and
233
the magnitudes of the Bragg wavelengths for the fast and slow axes. This formulation
234
relates changes in force-induced strains to changes in the Bragg wavelength of the fast
235
axis and slow axis, which can then be used to calculate spectral separation. Because
236
experimental data for the D-shape sensor will comprise fast axis, slow axis and spectral
237
separation data, we will apply the strain-optic formulation because it can be used to
238
calculate Bragg wavelengths for the fast and slow axes and spectral separation.
10
Super-structured D-shape fibre Bragg grating contact force sensor
239
240
2.2 Strain-optic principles
As mentioned above, the centre wavelength of a Bragg grating changes as a function of
241
the mechanical strains in the core of the optical fibre. In the case of birefringent optical
242
fibre, the changes in the Bragg wavelength of the light polarized along the slow axis (x-
243
axis) and fast axis (y-axis) are given by:


n2
s  s  z  s  pxz z  pxx  x  pxy y  
2


244


n 2f
 f   f  z   pyz z  pyx x  pyy y  
2


(5)
245
where  denotes a change in Bragg wavelength;  denotes mechanical strain with the
246
subscripts referencing the fibre coordinate system shown in Figure 1b; and
247
pxz  pxy  p yx  p yz  0.252 and pxx  p yy  0.113 are strain-optic constants [12]. Note
248
that the subscripts on the photoelastic constants are referred to the fibre coordinate
249
system. To calculate the Bragg wavelength of the fast and slow axis for any state of strain
250
the changes in Bragg wavelength (Equations 5) are added to the initial Bragg
251
wavelengths (i.e. the Bragg wavelengths when the fibre is unstrained).
252
To predict the contact force-induced changes in Bragg wavelength for the sensor
253
presented in this work the equations relating contact force and mechanical strain must be
254
established. To establish the equations, the sensor geometry and orientation to contact
255
force is presented.
256
257
2.3 Super-structured D-shape fibre sensor and force/strain models
As shown in Figure 2a, the contact force sensor consists of a D-shaped birefringent fibre
258
around which there is a super-structure that both transmits contact forces to the fibre and
259
maintains the orientation of the sensor with respect to the contact force (Figure 2b). The
260
super-structured D-shape fibre sensor, hereafter referred to as the sensor, comprises an
11
Super-structured D-shape fibre Bragg grating contact force sensor
261
outside Polyimide© sheath (SmallParts Inc., Miami FL, Ds = 165 micron, tW = 18 micron
262
nominally) that contains a Nitinol© wire (Need the manuf. Details, Dw = 51 micron
263
nominally) and a D-shape birefringent fibre (KVH Industries Inc., Middletown RI, Dmajor
264
= 125 micron, Dminor = 76 micron, r = 14 micron, dmajor = 8 micron, dminor = 5 micron)
265
with a 5 mm Bragg grating photo-inscribed in the fibre core (TechnicaSA, Beijing China,
266
the grating details). The length, along the fibre axis (Figure 1b), of the Polyimide©
267
sheath and Nitinol© wire is 12 mm at the center of which is the 5 mm Bragg. The
268
clearance space (Figure 2a) between the sheath, wire and fibre is filled with a compliant
269
adhesive (Dow Corning®, Midland MI, silicone 3-1753) that serves to maintain the
270
alignment of the features comprising the super-structure. A Polyimide© alignment
271
feature (Figure 2a) is affixed to the outside of the Polyimide© sheath by submerging the
272
sheath in a heat curable Polyimide© solution (HD MicroSystems™, Parlin NJ, solution
273
PI-2525) and curing it into the geometry shown. The length of the alignment feature is
274
nominally the same as the length of the sheath (i.e. 12 mm). The width of the alignment
275
feature was XXX (I actually made a bunch of them, how can I convey this).
276
Contact forces, F, are applied to the outside of the sheath and are aligned with the
277
internal features (e.g. wire and fibre) as shown in Figure 2b (alignment feature omitted
278
for clarity). The upper force is transmitted through the sheath, wire, and onto the fibre at
279
contact 1. The lower force is transmitted through the sheath and onto the fibre at contact
280
2.
281
The key difference between the sensor shown in Figure 2 and force sensors based
282
on bare (i.e. not superstructured) D-shape fibre [4] is that the contact force transmitted to
283
the flat side of the D-shape fibre of the super-structured sensor (Figure 2b) is
12
Super-structured D-shape fibre Bragg grating contact force sensor
284
concentrated by the wire. Conversely, force sensors using bare D-shape fibre distribute
285
forces over the entire flat side of the D-shape clad (Figure 1). As will be shown,
286
concentrating the contact force onto the fibre through the wire will result in increased
287
force sensitivity of the super-structured D-shape sensor relative to force sensors based on
288
bare D-shape fibre. This increase in force sensitivity is predicted by the equations,
289
developed below, relating contact force to strain in the fibre core.
290
13
Super-structured D-shape fibre Bragg grating contact force sensor
291
292
293
294
295
296
297
Figure 2: a) schematic of super-structured D-shape sensor cross section showing
Polyimide© sheath, Nitinol© wire, D-shape fibre and Polyimide© alignment feature. b)
contact forces, F are applied to the outside of the sheath and are aligned with the internal
features of the sensor as shown. The upper contact force is transmitted through the
sheath, to the wire and to the fibre at contact 1. The lower force is transmitted through the
sheath to the fibre at contact 2.
298
A plane elasticity model is applied to calculate the stresses and strains created in
299
the D-shape fibre core as a result of applied contact forces. To calculate the stresses and
300
strains, the D-shape fibre is approximated as an elastic half space [30] that is subjected to
301
contact forces of magnitude F (N/mm, normalized to sensor length along the fibre axis).
302
One of the key assumptions associated with elastic half space theory is that
303
stresses and strains vanish at distances infinitely far from the point of force application.
304
In the case of D-shape fibre, which is finite in size, this assumption cannot be satisfied.
305
To address this seemingly irreconcilable difference between the boundary conditions of
306
half space theory and the D-shape fibre sensor a composite half space model is proposed
307
that accounts for the shape and boundary conditions of the super-structured D-shape fibre
308
sensor.
309
As shown in Figure 3a, elastic half space formulations of contact consist of a
310
contact force applied at the origin of a large elastic continuum [30]. To satisfy boundary
311
conditions, when x and z are large all stresses and strains vanish and static equilibrium is
312
achieved by internal stresses that balance the external force. The stresses at any location
313
in the elastic continuum are:
14
Super-structured D-shape fibre Bragg grating contact force sensor
x2y
x  
 (x 2 +y 2 ) 2
2F
314
y3
 (x 2 +y 2 ) 2
y  
2F
 xy  
2F
(6)
xy 2
 (x 2 +y 2 ) 2
315
where  x ,  y ,  xy are the normal stresses aligned with the x and y directions and the
316
shear stress that shears the xy plane [30].
317
To address the difference in boundary conditions between the elastic half space
318
(Figure 3a) and the D-shape fibre sensor (Figure 3b) Equations 6 are adapted to formulate
319
the stresses for the composite half space.
320
321
322
323
324
325
326
327
328
329
Figure 3: a) schematic showing contact force, F applied to elastic half space with zero
stress boundary condition at regions far from co-ordinate system origin. b) schematic
showing D-shape fibre cross section as a composite elastic half space. The core of the
fibre is located relative to two co-ordinate systems with origins at contact 1 and contact 2.
The co-ordinate of the core relative to contact 1 and contact 2 is x1, y1 and x’2, y’2,
respectively. c) schematic representation of half space model for a bare D-shape fibre
subjected to distributed contact force over the flat side of the clad (pressure, P) and
concentrated contact force on the round side of the clad.
330
The sensor (Figure 2) is approximated using the boundary conditions shown in
331
Figure 3b. We assume that the contact forces applied to the top and bottom of the sheath
15
Super-structured D-shape fibre Bragg grating contact force sensor
332
(Figure 2b) are transmitted to the D-shape fibre. To calculate the stresses at any location
333
in the D-shape fibre, we assume that the state of stress in the fibre is due to a stress field,
334
given by Equations 6 due to the contact 1 and a stress field, also given by equations like
335
6, due to contact 2. Because we assume linear elasticity, we apply superposition to find
336
the total state of stress due to these contacts as:
337
x  
x 22 y'2 
2 F  x12 y1
+
  (x12 +y12 ) 2 (x 22 +y'22 ) 2 
y  

y'32
2 F  y13
+
 2 2 2
2
2 2 
  (x1 +y1 ) (x 2 +y'2 ) 
 xy  
x 2 y'22 
2 F  x1 y12
+
  (x12 +y12 ) 2 (x 22 +y'22 ) 2 
(7)
338
In the region of the centre of the fibre core, where the majority of light is transmitted, the
339
dimensions x are small relative to the dimensions y and Equations 7 can be simplified as:
x  0
y  
340
2F  1
1 
  
  y1 y'2 
(8)
 xy  0
341
Based on Equations 8, in the region of the centre of the fibre core there is negligible shear
342
and there is one principal stress. The simple state of stress can be used with Hooke’s Law
343
for plane strain [30] to obtain strains:
 (1  ) y
1
(1  2 ) x  (1  ) y 
E
E
(1  2 ) y
1
y 
(1  2 ) y  (1  ) x 
E
E
2(1  )
 xy 
 xy  0
E
x 
344




(9)
16
Super-structured D-shape fibre Bragg grating contact force sensor
345
where E and  are the Young’s modulus and Poisson ratio of the silica glass of the fibre,
346
77 GPa and 0.17, respectively [29] (check these values and the reference). When
347
Equations 7 are substituted into Equations 9, the resulting expressions for strain are
348
functions of contact force, F, and the co-ordinates of the core relative to contact 1 and
349
contact 2. For the sensor presented in Figure 2, y1 and y’2 are constant: 14 microns and 62
350
microns, respectively.
351
For the purpose of eventual comparison of the contact force sensor presented in
352
this work and contact force sensors based on bare D-shape fibre, we have also developed
353
a half-space model to estimate the stresses and strains in a bare D-shape fibre (Figure 3c).
354
In this model, contact force on the flat side of the clad is distributed, shown as contact
355
pressure, P, while contact force on the opposite side of the fibre remains concentrated.
356
We apply similar reasoning as that described for Figure 3b to obtain the stresses in the
357
region of the fibre core as:
x 
358
y 
P

P

 xy  0
  sin  
  sin   
2F  1 
  y'2 
(10)
359
360
where P is the applied pressure (N/mm2) and α is the included angle (i.e. 2.73 radians for
361
D-shape fibre) that is subtended by the pressure. The boundary conditions shown in
362
Figure 3b do not completely satisfy those of half-space theory because the pressure
363
boundary condition extends to dimensions comparable to the fibre diameter.
Nevertheless, Equations 10 can serve to approximate the state of stress and be used to
17
Super-structured D-shape fibre Bragg grating contact force sensor
364
365
gain insight into increases/decreases in sensitivity between the sensor presented here and
366
those based on bare D-shape fibre. The state of stress given by Equations 10 can be
367
substituted into Hooke’s Law for plane strain to obtain the strains in a similar form to
368
those presented as Equations 9.
369
The force-induced strains for the sensor presented here (Equations 9), and those
370
for a bare D-shape sensor, can then be substituted into Equations 5 to calculate the Bragg
371
wavelength shifts for the fast and slow axis for the sensor presented in this work and that
372
of a bare D-shape fibre subjected to contact force.
373
The process of Bragg wavelength shift calculation described above was
374
completed for contact forces ranging from 0 N/mm to 2 N/mm to predict force
375
sensitivity. Equations 5 were also used to calculate the axial strain sensitivity, for the
376
contact force sensor, as described below.
377
As will be described in detail below, the contact force sensor was tested
378
experimentally for sensitivity to axial strain. The experimental set-up for this test
379
consisted of the force sensor affixed to the top of an aluminum cantilever. Prescribed
380
deflections were applied to the cantilever, and the fast and slow axis Bragg wavelengths
381
were monitored. To estimate axial strain at the sensor, equations from classical beam
382
theory [31] were used to calculate the axial strain at the location along the beam of the
383
force sensor. To verify that results from this experiment agreed with strain-optic theory
384
and the known behavior of D-shape fibres subjected to axial strain [15], the axial strains,
385
and corresponding transverse Poisson strains, for the cantilever beam were used with
18
Super-structured D-shape fibre Bragg grating contact force sensor
386
Equations 5 to predict the Bragg wavelength shifts. Verification involved comparison of
387
the experimentally measured wavelength shifts to the strain-optic model predicted
388
wavelength shifts.
389
2.4 Sensor calibration protocols
390
A single force sensor was constructed and calibrated for sensitivity to contact-force, axial
391
strain and temperature in three separate calibration protocols.
392
The apparatus used for force calibration was identical to that used in a previous
393
study (Figure 4a) [19]. The force sensor and a support cylinder of identical outside
394
diameter were subjected to contact force between two metrology gauge blocks (Class 0,
395
24.1 mm × 24.1 mm, steel, Mitutoyo Can., Toronto, ON). The bottom gauge block
396
supported the force sensor and support cylinder and the top block applied contact forces.
397
Both gauge blocks were constrained by guide blocks allowing motion only in the
398
direction of load application.
399
Contact forces were applied by compressing a calibrated spring with a manual
400
screw-follower (not shown in Figure 4a). These forces were transmitted through a pre-
401
calibrated load cell (445 N capacity, ±0.1% FS non-repeatability, Futek Inc., Irvine, CA),
402
fixed to the top gauge block, connected to a data-acquisition system implemented in
403
LabView™ (version 8, National Instruments Inc., Austin, TX). The force was transmitted
404
through the top gauge block, through the sensor and support cylinder and finally to the
405
bottom gauge block. Applied contact forces in all calibrations ranged from 0 N mm−1 to,
406
nominally, 2 N mm−1. This range of forces was chosen because it brackets the range of
407
forces experienced in the hip during walking [57]
19
Super-structured D-shape fibre Bragg grating contact force sensor
408
409
410
411
412
413
414
Figure 4: a) schematic showing relevant features of contact force calibration apparatus.
Force is applied by compressing a calibrated spring that is in contact with a load cell and
top gauge block. Forces are transmitted through the force sensor and support cylinder,
and to the bottom gauge block. b) schematic showing forces applied to force sensor
without alignment feature, and c) with alignment feature.
415
To allow comparison of sensor sensitivity for a sensor without and with an
416
alignment feature, calibration was completed first for a sensor without an alignment
417
feature (Figure 4b), and subsequently with the alignment feature (Figure 4c). For the
418
sensor without the alignment feature, sensor orientation was controlled with an external
419
fibre rotator while contact forces were applied. The alignment feature was then added to
420
the sensor and calibration was repeated without external orientation control. One
421
calibration was performed for the sensor without an alignment feature, while three were
422
performed for the sensor with the alignment feature. Bragg wavelength shifts of the fast
20
Super-structured D-shape fibre Bragg grating contact force sensor
423
and slow axis were measured using an optical spectrum analyzer (ANDO AQ6331,
424
Tokyo, Japan) as described below.
425
Bragg wavelengths were demodulated by directing light from a broad C-band
426
light source (AFC-BBS1550, Milpitas, California) into a linear polarizer (PR 2000, JDS
427
Uniphase, Milpitas, California) and then into one of the input channels of a 3 dB optical
428
coupler (Blue Road Research). The light was then directed via the coupler to the FBG in
429
the test fiber, and the reflected spectrum was directed back through the optical coupler
430
and into the optical spectrum analyzer. The polarization of the light was adjusted to
431
illuminate either the fast or the slow axis of the fiber by using the tuning facilities of the
432
linear polarizer. The demodulation scheme described is similar to that presented by
433
numerous researchers [3, 32, 33].
434
The sensitivity to force, for both the slow and the fast axes, was calculated by
435
using linear regression as the slope in the recorded data (slope ± standard deviation for
436
slope), in terms of wavelength shift versus applied force. Sensitivity was also calculated
437
based on the spectral separation of the slow and fast axes (i.e., the difference between the
438
slow and the fast axis sensitivities).
439
In the axial strain protocol, the force sensor was affixed to the top surface of an
440
aluminum cantilever (Figure 5, w = 25 mm, h = 0.25 mm, L = 45 mm, and x”s = 10 mm)
441
using a UV curable adhesive (Norland Products Inc., Cranbury NJ, Blocking adhesive
442
107). To create axial strains in the cantilever, prescribed deflections, δ, were created at
443
the end of the cantilever. The prescribed deflections were measured using a mechanical
444
indicator (Mitutoyo Corp., JP, Indicator 2046F, 10 mm range), while Bragg wavelengths
445
of the fast and slow axis were recorded. As described in Section 2.3, the axial strains
21
Super-structured D-shape fibre Bragg grating contact force sensor
446
transmitted to the sensor from the cantilever were estimated from the deflections using
447
classical beam equations [31] by assuming that the tensile strains at the sensor/beam
448
interface (Section A-A, Figure 5) were transmitted to the sensor. The estimated axial
449
strains ranged from 0 με to 70 με. The protocol was conducted twice. The sensitivity to
450
strain, for both the slow and the fast axes and based on spectral separation, was calculated
451
by using linear regression as the slope in the recorded data, in terms of wavelength shift
452
versus strain.
453
454
455
456
457
458
Figure 5: schematic showing relevant feature axial strain calibration apparatus. The force
sensor was affixed to the top of the cantilever (Section A-A). Prescribed deflections were
applied to the end of the cantilever (x” = L).
459
Controls Inc., Winona MN, Model CascadeTEK, controller series 982). The force sensor
460
and a pre-calibrated thermocouple (Omega Engineering Inc., Stamford CT, Type-T),
461
interrogated with a thermocouple amplifier (Omega Engineering, Super MCJ), were fixed
462
to a glass slide (VWR International™, 76 mm by 26 mm by 1 mm) and placed within the
Temperature calibration was performed within a controllable oven (Watlow
22
Super-structured D-shape fibre Bragg grating contact force sensor
463
oven. The temperature, as reported by the thermocouple, of the force sensor was
464
increased from 30 ˚C to 60 ˚C, in 5 ˚C increments, while the Bragg wavelengths of the
465
fast and slow axis were recorded. This temperature range was chosen because it spans a
466
larger range than that experienced in most ex vivo experiments with cadaveric material.
467
temperature The protocol was repeated three times. The sensitivity to temperature, for
468
both the slow and the fast axes and based on spectral separation, was calculated by using
469
linear regression as the slope in the recorded data, in terms of wavelength shift versus
470
temperature.
471
472
473
3. Results
Figure 6 shows spectra recorded using the optical spectrum analyzer for both the fast and
474
slow axis. As shown, the Bragg spectra of both the fast and slow axis are symmetric
475
about their initial Bragg wavelength when the sensor is not subjected to contact force. As
476
force is applied to the sensor, both the fast and slow axis Bragg wavelengths shift to
477
longer wavelengths while retaining symmetry. The symmetry exhibited by the fast and
478
slow axis spectrums, when the sensor is subjected to force, indicates that the strains
479
experienced along the Bragg grating are uniform along the grating length. Conversely,
480
non-uniform strains over the grating length would manifest in increased width, along the
481
wavelength axis, of the Bragg spectrums and decreased peak light intensity [25]. As
482
shown in Figure 6, the slow axis Bragg wavelength experiences greater shift than the fast
483
axis. Therefore, as described below, the slow axis Bragg wavelength exhibits greater
484
sensitivity to contact force than the fast axis (Figure 7 and Table 1).
485
23
Super-structured D-shape fibre Bragg grating contact force sensor
486
487
488
489
490
491
492
493
Figure 6: Bragg spectra for the fast and slow axis recorded using the optical spectrum
analyzer. Contact forces applied to sensor cause shifts to longer Bragg wavelengths. The
symmetry of the spectra indicate that strains along grating are uniform while the relative
shifts of the slow and fast axis indicate that the slow axis has greater force sensitivity
than the fast axis.
494
wavelengths obtained from the half-space/strain-optic model, while Figure 7b shows the
495
measured Bragg wavelength shifts for the contact force sensor without the alignment
496
feature. As shown, for the super-structured sensor, the model predicted (Figure 7a) and
497
measured (Figure 7b) slopes, or sensitivities to contact force, match to within 1.8 % and
Figure 7a shows the predicted wavelength shifts of the fast and slow axis Bragg
24
Super-structured D-shape fibre Bragg grating contact force sensor
498
13.2 % (relative to experimental slope) for the slow and fast axis, respectively. Relative
499
to a bare D-shape sensor, the modeled slow and fast axis sensitivities for the super-
500
structured sensor are 125.4 % and 3.5% greater than those modeled for a bare D-shape
501
sensor.
502
25
Super-structured D-shape fibre Bragg grating contact force sensor
503
504
505
506
507
508
509
510
511
512
513
514
515
Figure 7: a) Predicted sensitivities of the fast and slow axis Bragg wavelengths obtained
from the half-space/strain-optic model for both the super-structured sensor and a bare Dshape sensor. b) measured wavelength shifts of the fast and slow axis Bragg wavelengths
for the sensor without the alignment feature. Sensitivities are reported as slope from
regressions (dashed lines). Error bars shown are not clearly visible at given scale but
convey non-repeatability of contact force measurements (±0.02 N/mm) and observed
non-repeatability of wavelength shift measurements (mean of ±1 pm for all data points
presented) from optical spectrum analyzer.
Table 1: Summary of experimental results for slow and fast axis sensitivity to contact
force, axial strain and temperature of contact force sensor. Sensitivities reported obtained
using linear regression and are reported as the best-fit slope ± the standard deviation on
slope. Correlation coefficients, r2, also obtained from regression calculations.
Sensitivity (pm/measurand unit)
Calibration protocol
Measurand unit
Trial
Slow axis
r†
2
Fast axis
r†
2
Spectral separation
force, without
alignment feature
(N/mm)
1
255.9±21.1
0.96
61.66±7.98
0.91
194.2
force, with alignment
feature
(N/mm)
1
2
3
mean
228.6±3.98
232.1±4.30
234.3±4.06
231.7
0.99
0.99
0.99
0.99
62.42±3.71
62.48±4.05
61.57±5.36
62.16
0.98
0.98
0.96
0.97
166.2
169.6
172.7
169.5
axial strain, with
alignment feature
(micro-strain)
1
2
mean
1.100±0.0786
1.130±0.0827
1.115
0.97
0.96
0.97
1.091±0.0996
1.129±0.0240
1.11
0.95
0.99
0.97
0.009‡
0.001‡
0.005‡
degree Celsius
1
2
3
mean
8.435±0.156
8.451±0.126
8.408±0.144
8.431
0.99
0.99
0.99
0.99
9.577±0.310
9.527±0.309
9.548±0.310
9.551
0.99
0.99
0.99
0.99
-1.142
-1.076
-1.140
-1.119
temperature, with
alignment feature
notes:
516
517
518
† r2 is abbreviated notation for linear correlation coefficient from least squares regression
‡ only one significant digit reported because of relatively low magnitude compared to slow and fast axis sensitivity
Figure 8 shows typical measured Bragg wavelength shifts for the contact force
519
sensor with the alignment feature (trial 1, results also noted in Table 1). As shown by the
520
results in Table 1 (force, with alignment feature), both the slow and fast axis sensitivities
521
are nearly constant over the three trials in which the sensor aligns itself (experimental
522
configuration shown in Figure 4c. The lowest and highest sensitivities over the three
523
trials deviate from the tabulated mean sensitivities for the slow and fast axis by only
524
-1.3% and 1.1% for the slow axis and -0.9% and 0.5% for the fast axis (Table 1). The
26
Super-structured D-shape fibre Bragg grating contact force sensor
525
lowest and highest spectral separation sensitivities deviate from the mean (i.e. 169.5
526
pm/(N/mm)) by only -1.9% and 1.9% (Table 1).
527
528
529
530
531
532
533
534
Figure 8: measured wavelength shifts of fast and slow axis Bragg wavelengths for contact
force sensor with alignment feature (trial 1 in Table 1). Error bars not clearly visible at
given scale but convey non-repeatability of measurements conveyed in caption for Figure
7b.
Figure 9a shows the strain/strain-optic model predicted wavelength shifts for the
535
fast and slow axis as a function of axial strain along the Bragg grating. As shown, the
536
model predicted sensitivities of the fast and slow axis differ by only 0.008% of the mean
537
sensitivity of the fast and slow axis (i.e.1.24475 pm/microstrain). The measured changes
538
in Bragg wavelength versus axial strain (Figure 9b) also show nearly identical sensitivity
539
of the fast and slow axis (also shown in Table 1, axial strain, with alignment feature). The
540
mean difference between the fast and slow axis sensitivities reported for trial 1 and 2,
541
shown in Table 1, is only 0.18%.
542
27
Super-structured D-shape fibre Bragg grating contact force sensor
543
544
545
546
547
548
549
Figure 9: a) Predicted sensitivities to axial strain of the fast and slow axis Bragg
wavelengths obtained from the strain/strain-optic model. b) measured (trial 2) wavelength
shifts of the fast and slow axis Bragg wavelengths for the sensor with the alignment
feature. Sensitivities are reported as slope from regressions (dashed lines). Error bars
shown are not clearly visible at given scale but convey non-repeatability of axial strain
values (±1.69 micro-strain) and observed non-repeatability of wavelength shift
28
Super-structured D-shape fibre Bragg grating contact force sensor
550
551
552
553
measurements (mean of ±1 pm for all data points presented) from optical spectrum
analyzer.
554
protocol (Table 1, temperature, with alignment feature). Like the results for axial strain
555
calibration, the fast and slow axis sensitivities to temperature were nearly identical. For
556
example, as shown by the results based on spectral separation (Table 1), the spectral
557
separation sensitivity was only 12.4 % of the mean of the fast and slow axis sensitivities
558
(i.e. 8.99 pm/degree Celsius).
559
560
561
562
563
564
565
566
567
568
569
Figure 10 shows typical results obtained during the temperature calibration
Figure 10: measured (trial 1) fast and slow axis Bragg wavelengths versus temperature
for the contact force sensor with the alignment feature. Sensitivities are reported as slope
from regressions (dashed lines). Error bars shown are not clearly visible at given scale
but convey non-repeatability of thermocouple measurements (±0.1 degrees Celsius) and
observed non-repeatability of wavelength shift measurements (mean of ±1 pm for all data
points presented) from optical spectrum analyzer.
29
Super-structured D-shape fibre Bragg grating contact force sensor
570
571
4. Discussion
The contact force sensor presented in this work has desirable characteristics when
572
compared to other birefringence-based force sensors and the authors’ previous work [19]
573
including: increased sensitivity to contact force relative to bare-Dshape fibre and other
574
birefringent fibres; near-constant force sensitivity without external orientation control;
575
and negligible co-sensitivity to extraneous axial strain and temperature changes. Insights
576
gained from the half-space/strain-optic model, with reference to sensor design features,
577
can be used to elucidate the causes of these benefits.
578
As shown by the results plotted in Figure 7, the slow axis Bragg wavelength of
579
the super-structured sensor has sensitivity 125.4 % greater than the slow axis of a bare D-
580
shape fibre. Conversely, the fast axis sensitivities of the super-structured and bare D-
581
shape sensors are nearly identical, and differ by only 3.5%. The principal cause of the
582
increase in slow axis sensitivity is a 273% increase, relative to the case of bare D-shape
583
fibre, in compressive stresses,  y , that result from concentrating contact forces at contact
584
1 with the Nitinol© wire. Conversely, fast axis sensitivities of the super-structured and
585
bare D-shape sensors match to within 3.5% because the state of stresses in the two
586
sensors (Equations 8 for super-structured and Equations 10 for bare fibre) lead to
587
mechanical strains in the core that cause nearly identical fast axis wavelength shifts for a
588
given contact force.
589
Experimental measurements reinforce the model-predicted increases in sensitivity
590
for the slow axis of the super-structured sensor. The half-space/strain-optic model
591
predicted slow axis sensitivity is 260.6 pm/(N/mm), which is 45% higher than previously
592
reported maximum slow axis sensitivity for bare D-shape fibre (i.e. 180 pm/(N/mm) [4]).
593
The model-predicted slow axis sensitivity of the bare D-shape fibre is 115.6 pm/(N/mm),
30
Super-structured D-shape fibre Bragg grating contact force sensor
594
which is much lower than the previously sensitivity (180 pm/(N/mm)). We suspect the
595
primary cause for the discrepancy between the model predicted and previously reported
596
slow axis sensitivity is differences in boundary conditions for compressive stresses,  y ,
597
between the half-space model and actual bare D-shape sensors. As shown in Figure 3c,
598
compressive stresses are equivalent to the contact pressure, P, along the entire flat of the
599
D-shape clad, which contradicts one of the assumptions of half-space theory: stresses
600
vanish at regions remote from the center of contact.
601
602
Experimental measurements also reinforce model predicted values for the fast
axis Bragg wavelength for both the super-structured and bare D-shape sensors.
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
5. Conclusions
References
[1]
Pieter, L. S., Beatrys, M. L., and Anatoli, A. C., 2005, "Chirped fiber Bragg
grating sensor for pressure and position sensing," SPIE, p. 054402.
[2]
Udd, E., Lawrence, C., and Nelson, D., 1997, "Development of a Three Axis
Strain and Temperature Fiber Optic Grating Sensor," Proceedings of SPIE, 3042, pp.
229-236.
[3]
Xu, M. G., Reekie, L., Chow, Y. T., and Dakin, J. P., 1993, "Optical in-fibre
grating high pressure sensor," Electronics Letters, 29, pp. 398-399.
[4]
Chehura, E., Ye, C.-C., Staines, S. E., James, S. W., and Tatam, R. P., 2004,
"Characterization of the response of fibre Bragg gratings fabricated in stress and
geometrically induced high birefringence fibres to temperature and transverse load,"
Smart Materials and Structures, 13, pp. 888-895.
[5]
Yeo, T. L., Sun, T., Grattan, K. T. V., Parry, D., Lade, R., and Powell, B. D.,
2005, "Characterisation of a polymer-coated fibre Bragg grating sensor for relative
humidity sensing," Sensors and Actuators B: Chemical, 110, pp. 148-155.
[6]
Fernandez, A. F., Brichard, B., Berghmans, F., and Decreton, M., 2002, "DoseRate Dependencies in Gamma-Irradiated Fiber Bragg Grating Filters," IEEE Transactions
on Nuclear Science, 49(6), pp. 2874-2878.
[7]
Lawrence, C. M., Nelson, D. V., and Udd, E., 1996, "Multi-Parameter Sensing
with Fiber Bragg Gratings," Proceedings of SPIE, 2872, pp. 24-31.
[8]
Udd, E., 1991, Fibre Optic Sensors, An Introduction for Engineers and Scientists,
Wiley InterScience.
31
Super-structured D-shape fibre Bragg grating contact force sensor
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
[9]
Liu, Y., Guo, Z., Zhang, Y., Seng, K., Dong, C., and Dong, X., 2000,
"Simultaneous pressure and temperature measurement with polymer-coated fibre Bragg
grating," Electronics Letters, 36(6), pp. 564-566.
[10] Nunes, L. C. S., Valente, L. C. G., Llerena, R. W. A., Braga, A. M. B., and
Triques, A. L. C., 2004, "Simultaneous measurement of temperature and pressure using
single fiber Bragg grating and fixed filter demodulation technique," Proceedings of SPIE,
5622, pp. 906-911.
[11] Sun, A., Qiao, X. G., Jia, Z. A., Li, M., and Zhao, D. Z., 2005, "Study of
simultaneous measurement of temperature and pressure using double fiber Bragg gratings
with polymer package," SPIE, p. 034402.
[12] Measures, R. M., 2001, Structural Health Monitoring with Fiber Optic
Technology, Academic Press.
[13] Okamoto, K., Hosaka, T., and Edahiro, T., 1981, "Stress analysis of optical fibers
by a finite element method," IEEE Journal of Quantum Electronics, QE-17(10), pp. 21232129.
[14] Barlow, A. J., and Payne, D., 1983, "The stress-optic effect in optical fibers,"
IEEE Journal of Quantum Electronics, QE-19(5), pp. 834-839.
[15] Zhao, D., Zhou, K., Chen, X., Zhang, L., Bennion, I., Flockhart, G., MacPherson,
W. N., Barton, J. S., and Jone, J. D. C., 2004, "Implementation of vectorial bend sensors
using long-period gratings UV-inscribed in special shape fibres," Measurement Science
and Technology, 15, pp. 1647-1650.
[16] Udd, E., Nelson, D., and Lawrence, C., 1996, "Three Axis Strain and
Temperature Fiber Optic Grating Sensor," Proceedings of SPIE, 2718, pp. 104-109.
[17] Wierzba, P., and Kosmowski, B. B., 2003, "Application of polarisationmaintaining side-hole fibres to direct force measurement," Opto-Electronics Review,
11(4), pp. 305-312.
[18] Abe, I., Frazao, O., Schiller, M. W., Noqueira, R. N., Kalinowski, H. J., and
Pinto, J. L., 2006, "Bragg gratings in normal and reduced diameter high birefringence
fibre optics," Measurement Science and Technology, 17, pp. 1477-1484.
[19] Dennison, C. R., Wild, P. M., Wilson, D. R., and Gilbart, M. K., 2010, "An infiber Bragg grating sensor for contact force and stress measurements in articular joints,"
Measurement Science and Technology, 21, p. 115803.
[20] Wilson, D. R., McWalter, E. J., and Johnston, J. D., 2008, "The measurement of
joint mechanics and their role in osteoarthritis genesis and progression," Rheumatic
Disease Clinics of North America, 34, pp. 605-622.
[21] Anderson, A. E., Ellis, B. J., Maas, S. A., Peters, C. L., and Weiss, J. A., 2008,
"Validation of finite element predictions of cartilage contact pressure in the human hip
joint," Journal of Biomechanical Engineering, 130, p. 10pp.
[22] Cottrell, J. M., Scholten, P., Wanich, T., Warren, R. F., Wright, T. M., and Maher,
S. A., 2008, "A new technique to measure the dynamic contact pressures on the tibial
plateau," Journal of Biomechanics, 41, pp. 2324-2329.
[23] Shao, L.-Y., Jiang, Q., and Albert, J., 2010, "Fiber optic pressure sensing with
conforming elastomers," Applied Optics, 49(35), pp. 6784-6788.
[24] Hill, K. O., Fujii, Y., Johnson, D. C., and Kawasaki, B. S., 1978, "Photosensitivity
in optical fiber waveguides: Application to reflection filter fabrication," Applied Physics
Letters, 32(10), pp. 647-649.
32
Super-structured D-shape fibre Bragg grating contact force sensor
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
[25] Huang, S., LeBlanc, M., Ohn, M. M., and Measures, R. M., 1995, "Bragg
intragrating structural sensing," Applied Optics, 34(22), pp. 5003-5009.
[26] Noda, J., Okamoto, K., and Sasaki, Y., 1986, "Polarization maintaining fibers and
their applications," Journal of Lightwave TEchnology, 4(8), pp. 1071-1089.
[27] Mendez, A., and Morse, T. F., 2007, Specialty optical fibers handbook, Academic
Press.
[28] Kumar, A., Gupta, V., and Thyagarajan, K., 1987, "Geometrical birefringence of
polished and D-shape fibers," Optics Communications, 61(3), pp. 195-198.
[29] Urbanczyk, W., Martynkien, T., and Bock, W. J., 2001, "Dispersion effects in
elliptical-core highly birefringent fibers," Applied Optics, 40(12), pp. 1911-1920.
[30] Johnson, K. L., 1987, Contact mechanics, Cambridge University Press.
[31] Norton, R. L., 2000, Machine Design, An integrated approach, Prentice Hall.
[32] Xu, M. G., Geiger, H., and Dakin, J. P., 1996, "Fibre grating pressure sensor with
enhanced sensitivity using a glass-bubble housing," Electronics Letters, 32, pp. 128-129.
[33] Yamate, T., Ramos, R. T., Schroeder, R. J., and Udd, E., 2000, "Thermally
insensitive pressure measurements up to 300 degree C using fiber Bragg gratings written
onto side hole single mode fiber," Proceedings of SPIE, 4185, pp. 628-632.
33