Вы находитесь на странице: 1из 23

doi: 10.1098/rspa.2006.

1685
, 2481-2502 462 2006 Proc. R. Soc. A

Pablo D Ruiz, Jonathan M Huntley and Alejandro Maranon

media
displacement fields within optically scattering
mapping structure and three-dimensional
Tilt scanning interferometry: a novel technique for

References
html#ref-list-1
http://rspa.royalsocietypublishing.org/content/462/2072/2481.full.
This article cites 19 articles
Email alerting service
here the box at the top right-hand corner of the article or click
Receive free email alerts when new articles cite this article - sign up in
http://rspa.royalsocietypublishing.org/subscriptions go to: Proc. R. Soc. A To subscribe to
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
Tilt scanning interferometry: a novel
technique for mapping structure and
three-dimensional displacement elds within
optically scattering media
BY PABLO D. RUIZ, JONATHAN M. HUNTLEY* AND ALEJANDRO MARANON
Wolfson School of Mechanical and Manufacturing Engineering, Loughborough
University, Loughborough, Leicestershire LE11 3TU, UK
We describe a novel technique that we call tilt scanning interferometry to measure
depth-resolved structure and displacement elds within semi-transparent scattering
materials. The method differs signicantly from conventional optical coherence
tomography, in that only one wavelength is used throughout the whole measurement
process. Temporal sequences of speckle interferograms are recorded while the
illumination angle is tilted at constant rate. Fourier transformation of the resulting
three-dimensional intensity distribution along the time axis reconstructs the scattering
potential within the medium. Repeating the measurements with the object wave at
equal and opposite angles about the observation direction results in two three-
dimensional phase-change volumes, the sum of which gives the out-of-plane-sensitive
phase volume and the difference between which gives the in-plane phase volume. From
these phase-change volumes the in-plane and out-of-plane depth-resolved displacement
elds are obtained. The theoretical framework for the technique is explained in detail
and a practical optical implementation is described. Finally, results from proof-of-
principle experiments involving a semi-transparent beam undergoing bending are
presented.
Keywords: speckle interferometry; depth-resolved displacements; whole-eld;
displacement measurement; optical coherence tomography;
tilt scanning interferometry
1. Introduction
The ability to measure internal displacements and strains within a material has
many potential applications in engineering and medical sciences. A broad range
of methods has been developed in the last few decades, such as neutron
diffraction (ND; Fitzpatrick & Lodini 2003; Hutchings et al. 2005), photoelastic
tomography (Abe et al. 1986; Aben et al. 2004), phase contrast magnetic
resonance imaging (PCMRI; Steele et al. 2000; Draney et al. 2002) and three-
dimensional digital image correlation (DIC) using data acquired with X-ray
computed tomography (Bay et al. 1999) and optical coherence tomography
Proc. R. Soc. A (2006) 462, 24812502
doi:10.1098/rspa.2006.1685
Published online 30 March 2006
* Author for correspondence (j.m.huntley@lboro.ac.uk).
Received 26 October 2005
Accepted 27 January 2006 2481 q 2006 The Royal Society
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
(OCT; Schmitt 1998; Fercher et al. 2003). Each technique has a restricted range
of materials to which it can be applied: ND, for example, is suitable for
polycrystalline metals; PCMRI requires signicant water or fat content in the
sample. For many technologically and medically important materials (optically
scattering polymers, composites and biological tissues), the existing techniques
are often either not applicable or else are too insensitive. The lack of sensitivity
arises from the fact that in DIC the sensitivity is coupled to the spatial resolution
of the technique and is therefore limited by it.
Recent promising developments in OCT technology sidestep this problem by
making use of the phase information within the OCT signal, thereby decoupling
the displacement sensitivity and depth resolution (Gastinger et al. 2003; Gulker &
Kraft 2003; Ruiz et al. 2004, 2005). As a result, displacement sensitivity is
improved by 23 orders of magnitude compared to the depth resolution of state-of-
the-art OCT systems.
The resolution of depth in OCT is achieved through the use of multiple
wavelengths which are present either simultaneously (Born & Wolf 1959; Dresel
et al. 1992), leading to a narrow-gated region of modulating scatterers, or else
sequentially. The latter case is known as wavelength scanning interferometry
(WSI; Takeda et al. 1982; Fercher et al. 1995; Lexer et al. 1997; Kuwamura &
Yamaguchi 1997; de Groot 2000) and relies on a depth-encoding frequency shift
in the interference signal as the wavelength of the light source is swept through a
narrow range (see gure 1a). While the WSI form of OCT offers superior signal-
to-noise ratio to standard OCT for displacement sensing, it suffers from a
signicant practical drawback that tunable light sources are expensive and
difcult to operate without mode-hopping.
In this paper we present a different approach to measuring depth-resolved
displacements within optically scattering materials, which is based on tilting the
illuminating beam during the acquisition of the image sequences. This provides
the necessary depth-dependent phase shifts that allow the reconstruction of the
l(t)
l
R
S
1
S
2
f
2
f
1
S
j
(a)
R
S
1
S
2
f
2
f
1
S
j
(b)
Figure 1. (a) Generation of depth-encoding frequency shifts by: (a) tuning the wavelength of the
light source through a narrow range (wavelength scanning interferometry) and (b) by tilting a
monochromatic wavefront (tilt scanning interferometry).
P. D. Ruiz and others 2482
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
object structure and its internal displacements. The method is quite distinct from
all the previous OCT literature, in that only a single wavelength is present
throughout the entire data recording process, thereby considerably reducing the
expense and complexity of the light source. The depth-encoding frequency shift
can now be regarded as coming from the Doppler shift of the photons reected
from the tilting mirror in the object illumination beam path. This is shown
schematically and in a simplied way in gure 1b, where different heterodyne
frequencies are produced from scattering points lying at different depths within
the material. As with the WSI version of OCT, displacement sensitivity of this
new technique that we denote here by tilt scanning interferometry (TSI) is
decoupled from the intrinsic depth resolution of the technique and is a few tens of
nanometres at optical wavelengths. Extraction of depth and displacement
information of non-transparent object surfaces has been previously reported
using speckle contouring techniques based on source displacement (Rodr guez-
Vera et al. 1992). TSI extends this capability to study depth-resolved
displacement elds inside semi-transparent materials.
In 2, we provide a mathematical description of the technique following a ray-
tracing approach to model the problem. We also explain how to obtain the
scattering potential of the sample as well as depth-resolved displacement elds
and introduce some important parameters of the technique, namely the depth
resolution, gauge volume, depth range and displacement sensitivity. In 3, we
present a proof-of-principle experiment based on a TSI approach and show
encouraging results of depth-resolved displacements measured within a beam
under three-point bending up to a depth of over 5 mm. We compare the
measurements with nite-element predictions, discuss some error sources, and
nally in 4 draw some conclusions.
2. Tilt scanning interferometry
(a ) Depth-dependent phase shift introduced by a tilting wavefront
Figure 2a shows a semi-transparent scattering material of refractive index n
1
immersed in a medium of refractive index n
0
, illuminated by a collimated
beam of wavelength l at an angle q to the optical axis of the system. We
assume that the surface is at, though this restriction can be relaxed through a
suitable extension of the analysis presented here. It is convenient to place a
at but microscopically rough opaque surface over the region y%0 of the plane
zZ0, where the coordinate system (x, y, z) is as dened in gure 2a. This
reference surface serves two purposes: rst, allowing for correction of the
nonlinearity of the tilting device and second (since it does not strain during
the loading of the sample), enabling registration of the before- and after-load
scans of the sample. We should emphasize, however, that it does not provide
the reference wave for the interferometer, which rather is introduced by means
of a separate beam splitter, BS. While the reference surface has the obvious
drawback of obscuring the sample for some of the pixels in the eld of view
(FOV), it could, in principle, be dispensed with given a sufciently well-
calibrated tilting device.
The illumination beam is refracted at the object surface z
1
(x, y) and reaches
point F with coordinates (x, y, z) at an angle q
r
Zsin
K1
n
0
sin q=n
1
to the optical
2483 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
axis. Some of the light scattered at F travels vertically downwards, and is
recombined by BS with a reference wave derived from the same laser light source
as the object beam, and is imaged onto a pixel lying within a two-dimensional
photodetector array.
The phase difference between light scattered at F and a reference wavefront
with phase f
r
can be expressed relative to the phase difference at point G, which
lies on a rough opaque reference surface R at the origin (0, 0, 0), as
fx; y; z Zf0; 0; 0 C
2p
l
n
0
x sin q Cn
0
z
1
1 Ccos q Cn
1
zKz
1
1 Ccos q
r
;
2:1
where we assumed that zRz
1
R0. The phase differences due to the rst, second
and third terms between square brackets in equation (2.1) account for the optical
paths AB, CDCDB and EFCFD, respectively. The random distribution of
scattering centres within the material gives rise to a speckle phase distribution
along x in this two-dimensional representation.
If the illumination angle q changes linearly with time about the centre angle
q
c
, i.e.
qt Zq
c
C
Dq
T
t; 2:2
T being the time it takes to scan through the tilt angle Dq and t the time
variable (KT/2%t%T/2), then the phase f(x, y, z) will vary as
vfx; y; z; t
vt
Z
vf0; 0; 0; t
vt
C
2pn
0
l
Dq
T
x cos qKz
1
sin qKzKz
1
x: 2:3
z
1
z
x
y
z
max
F
(x, y, z)
E
f
z
(y, zz
1
)
f(x, y, z
1
)
f
z1
(y, z
1
)
f ( x, y, 0)
f
x
( x, y)
f (0, y, 0)
(0, y, 0)
(x, y, 0) x
m
a
g
n
i
t
u
d
e


I
r
(
x
,
y
)
f (x, y, z)
fro
m
in
s
id
e
th
e
o
b
je
c
t
fro
m
re
fe
re
n
c
e
s
u
rfa
c
e
R
(x, y, z
1
)
(x, y, 0)
(0, 0, 0)
R o
p
t
i
c
a
l

a
x
i
s
I(x, y, t)
(a) (b) (c)
n
1
n
0
q
r
f
r
q
q
l
D
C
B R
G
A
BS
z
for y>0
for y<0
f f
z
1
Figure 2. (a) Ray diagram for tilt scanning interferometry. A variation in the illumination angle
modulates the interference signal with a depth-dependent Doppler shift. (b) Spectrum of
interference signal associated with points along BF. (c) Shift of the spectrum for different
positions in the horizontal direction.
P. D. Ruiz and others 2484
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
In the last term between the square brackets the following relationship, derived
from Snells law of refraction, was used:
vq
r
vq
Z
c cos q

1Kc
2
sin
2
q
_ ; 2:4
with cZn
0
=n
1
. For convenience, we dene a parameter x as
x Z
c cos q sin q

1Kc
2
sin
2
q
_ : 2:5
Equation (2.3) can be expressed in terms of temporal frequencies as
f x; y; z Zf 0; 0; 0 C
n
0
Dq
lT
x cos qKz
1
sin qKzKz
1
x; 2:6
Zf 0; 0; 0 Cf
x
x; y Cf
z1
y; z
1
Cf
z
y; zKz
1
: 2:7
f (0, 0, 0) is a carrier frequency due to the rigid body translation or piston
term of wavefront GA as it tilts around an axis perpendicular to the plane of
gure 2a, and is zero if that axis passes through point G. The frequency of
the second term in equation (2.7) varies linearly with x, whereas f
z1
(y, z
1
)
accounts for the distance from the object surface to the reference surface,
z
1
, and f
z
(y, zKz
1
) to the depth zKz
1
of scattering points within the sample.
The last two terms can be interpreted as depth-encoding heterodyne frequency
shifts due to the Doppler shift of the tilting beam. f (x, y, z) in equations (2.6)
and (2.7) represents the absolute modulation frequency produced by
interference between light coming from point F at position (x, y, z) within the
object and the plane reference wavefront with phase f
r
coming from the
recombining beam splitter.
At zZz
1
, f
z
y; zKz
1
Z0 and f x; y; z
1
Zf 0; 0; 0Cf
x
x; yCf
z1
y; z
1
is the
frequency associated with point D at position (x, y, z
1
) on the object surface in
gure 2a. If we put zZz
1
Z0, f
z
Zy; zKz
1
Zf
z1
y; z
1
Z0 and f x; y; 0Z
f 0; 0; 0Cf
x
x; y is the frequency associated with point B lying at (x, 0, 0) on
the reference surface R in gure 2a.
(b ) Extraction of the scattering potential
The scattering potential of the sample can be extracted from the Doppler
shifts by effectively mapping depth from frequency. In what follows, we will
assume, as in standard OCT, that the contributions from multiple scattering
within the material can be neglected. The interference between light coming from
all the scattering points along DF in gure 2a and the reference wavefront f
r
gives rise to an intensity signal modulated with multiple frequencies:
I x; y; t ZI
r
x; y C2
_
z
max
z
1
I
r
x; yI x; y; z
1=2
cos2pf x; y; ztdz
C2
_
z
max
z
1
_
z
max
z
1
I x; y; zI x; y; z
0

1=2
cosf2pf x; y; zKf x; y; z
0
tgdz dz
0
:
2:8
While the rst term in the right-hand side of equation (2.8) represents the dc
component of the reference beam, the second term corresponds to the
2485 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
modulation due to interference between the reference beam and light scattered
within the material. The integration limit z
max
represents either the object
back surface, the maximum penetration depth which depends on the
absorbance of the material and the laser power, or the depth range of the
system (as discussed in 2d), whichever is the minimum. The double integral
in the third term is due to cross-interference between light coming from within
the object and contributes to the dc component and low-frequency components
in the interference signal.
The frequency of each term on the right-hand side of equations (2.6) and (2.7)
is, in general, dependent on q and therefore changes during the course of the scan.
We assume for now that Dq is small enough so that the resulting frequency shifts
can be neglected.
A one-dimensional Fourier transform of signal I(x, y, t) (the light intensity
measured at a single pixel) along the time axis gives rise to a spectrum as shown
schematically at gure 2b. Any given pixel either sees the reference surface or the
sample, but not both; this gure can therefore be interpreted as a top view of the
spectrum over all y onto the (f, x) plane, i.e. as a superposition of the spectra
corresponding to the reference surface R and the interior of the sample. There is
a dc peak at fZ0 for all x, another peak at f (x, 0) corresponding to the reference
surface R, and a band associated with the object. The leading edge at f (x, z
1
)
generally shows the highest amplitude of the band due to the refractive index
change at the surface of the sample. For higher frequencies, the amplitude
decreases due to absorption and multiple scattering within the material. The
position of the peaks is therefore associated with the internal structure of the
object and its position relative to the reference surface, whereas their amplitudes
are related to the degree of scattering or reection coefcient at each point within
the object or at the reference surface.
Usually, the spectrum
~
I x; f (where w indicates a Fourier transformed
variable) is obtained through the Fourier transform of the product of the
intensity signal I(x, t) with a window function W(t). In this way, the Fourier
transform of I(x, t) is convolved in the frequency domain with the Fourier
transform of the window function,
~
Wf . A Hanning window WtZ1=2K
1=2cos2ptCT=2=T was used in this work in order to reduce the cross-talk
between secondary peaks that would be present if a rectangular window, or
equivalently no window at all, were used.
The position of the object surface relative to the reference surface R, z
1
, is
proportional to f
z1
, which is the frequency difference between the spectral peak
due to the scattering at the sample surface and that due to the reference surface.
This difference is not present in a spectrum from a single pixel, but rather is
calculated from two or more pixels imaging the reference and sample at the same
x-value (i.e. from the same column of the photodetector array), so that all other
terms in the right-hand side of equation (2.7) are common to both. From the
third term of equation (2.7), we have
z
1
ZK
lTf
z1
n
0
Dq sin q
: 2:9
For simplicity, the following analysis will be limited to the case where z
1
(x)
is a constant. For objects with a surface of arbitrary shape, the refracted angle
q
r
at point F would depend on the coordinates x, y, z and the angle between
P. D. Ruiz and others 2486
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
the incident illumination beam and the normal of the object surface to the
optical axis at the point where the ray that ends in F intersects the surface. In
this case, the expression for the optical phase difference f(x, y, z) in equation
(2.1) would include extra terms. This case can be simplied in two different
ways:
(i) Experimentally, by immersing the object in a cell lled with index
matching uid. This approach is equivalent to having a at object
surface which is parallel to the reference surface R.
(ii) Numerically, by solving the refraction problem in two stages. The
problem of an object with a at surface parallel to the reference surface
is solved rst and then the refraction effects due to a curved surface (or
arbitrary surface) are evaluated to correct the rst approximation.
For the initially at samples used in the experiments presented here, such
corrections were relatively minor and neglected for the rest of the paper.
Once z
1
has been evaluated through equation (2.9), the position z of a
scattering point underneath the object surface can be obtained as
zx; y Zz
1
K
lTf x; y; zKf x; y; z
1

xn
0
Dq
: 2:10
The difference between the terms in square brackets is simply f
z
y; zKz
1

(as seen in gure 2b) and depends on the refractive indices of the material and
surrounding medium and on the illumination angle through the parameter x.
The spectral bandwidth Df associated with a thickness Dz within the object
can be obtained using equation (2.10) as
Df ZK
n
0
jxjDq
lT
Dz: 2:11
z < z
1
<z
2
f (x
1
, y, z)
f (x
1
, y, 0)
f (x
2
, y, z)
f (x
2
, y, z
1
)
f (x
2
, y, 0)
(0, 0, 0)
R
(0, 0, 0) (0, 0, 0)
n
1
x
1
x
y
x
2
n
0
z
1
n
2
z
n
0
< n
1
<n
2
z
1
z
2
z
for y>0
for y<0
f (x
1
, y, z
1
)
Figure 3. Effect of the refractive index on the depth resolution and the spectral width of an object
of constant thickness.
2487 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
This is illustrated in gure 3, with an object of constant thickness and different
refractive index for different values of x. The spectrum looks narrower for the
material with higher refractive index n
2
, for which x is lower.
The scattering potential is obtained simply by a mapping of the modulation
amplitude j
~
I x; y; f j
2
from spectrum coordinates (x, y, f ) into spatial coordinates
(x, y, z) through the relationship between frequency and position shown in
equation (2.10).
While a single illumination direction is sufcient to obtain the scattering
potential, multiple directions can be used to improve its estimation.
A symmetric lateral illumination conguration has the advantage of providing
spectra with the same bandwidth for a given object thickness, assuming a
uniform refractive index distribution. If the left and right illumination angles
q
c
Zq
L
and q
c
Zq
R
are set so that jxj is a maximum, then small differences
between q
L
and q
R
will produce negligible differences between jxq
L
j and
jxq
R
j, thus relaxing alignment requirements. A more important consequence
of this choice of q
L
and q
R
is that the depth resolution of the system is also
optimized. Figure 4 shows the variation of jxqj with q for a range of typical c
values. The variation with refractive index means that the depth resolution
outside the object is better than that within it.
(c ) Depth-resolved displacements
The position of scattering points within the object will change according to the
mechanical properties of the material and the loads applied to it. The value of f
at a particular voxel (x, y, z) in the specimen will reect that change and can be
calculated from the real and imaginary parts of the spectrum
~
I x; y; f at a
frequency given by equation (2.10).
For the system shown in gure 5, displacements in the y direction cause no
phase change due to the illumination geometry. If point F at (x, y, z) moves to
a new position F
0
with coordinates (xCu, y, zCw) after deformation (see
gure 6), then from equation (2.1) the phase difference after displacement can
20 40 60 80
incidence angle q (deg)
0
0.2
0.4
0.6
0.8
1.0
x
1.0
1.4
1.2
1.3
1.1
Figure 4. Variation of refraction parameter x with the illumination angle q for ve different values
of cZn
0
=n
1
Z1; 1:1; 1:2; 1:3 and 1:4.
P. D. Ruiz and others 2488
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
be written as
fxCu;y;z CwZf0;0;0C
2p
l
fxCux;y;zn
0
sinqCz
1
Cwx;y;z
1
n
0
!1CcosqCz Cwx;y;zKz
1
Kwx;y;z
1
n
1
1Ccosq
r
g:
2:12
While a single illumination direction is sufcient to extract the scattering
potential, at least two illumination directions are essential to determine the in-
plane (x) and the out-of-plane (z) displacement components. For right and left
lateral illumination, where q
R
O0 and q
L
!0, respectively, the phase difference
Dfx; y; zZfxCu; y; zCwKfx; y; z due to object deformation is
Df
R
x; y; z Z
2p
l
fux; y; zn
0
sin q
R
Cwx; y; z
1
n
0
1 Ccos q
R

Cwx; y; zKwx; y; z
1
n
1
1 Ccos q
Rr
g;
Df
L
x; y; z Z
2p
l
fKux; y; zn
0
sin q
L
Cwx; y; z
1
n
0
1 Ccos q
L

Cwx; y; zKwx; y; z
1
n
1
1 Ccos q
Lr
g;
_

_
2:13
where w(x, y, z
1
) is the out-of-plane displacement of point D at (x, y, z
1
) to D
0
at (x, y, z
1
Cw(z
1
)). By choosing q
R
ZKq
L
Zq, the in-plane (x) and out-of-plane
(z) phase difference components Df
x
x; y; zZDf
R
x; y; zKDf
L
x; y; z and Df
z
x; y; zZDf
R
x; y; zCDf
L
x; y; z are obtained as
Df
x
x; y; z Z
4p
l
n
0
ux; y; zsin q; 2:14
L
1
L
2
L
3
WBS
C
O
M
2
M
1
CBS
L
4
TM
S
OF
2
OF
1
R
RG
q
q
Figure 5. Optical setup of a tilt scanning interferometer showing: optical bres OF
1
and OF
2
; lenses
L
1
L
4
; tilting mirror TM; cube beam splitter CBS; mirrors M
1
, M
2
; alignment screen S; wedge
beam splitter WBS; camera C; ramp generator RG and object O.
2489 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
Df
z
x; y; z Z
4p
l
fwx; y; z
1
n
0
1 Ccos qKn
1
1 Ccos q
r

Cwx; y; zn
1
1 Ccos q
r
g: 2:15
The rst term within the {.} in equation (2.15) corresponds to the phase change
due to a displacement of the object surface relative to the reference surface, while
the second one is due to the displacement of point F to F
0
along the optical axis.
The in-plane and out-of-plane displacements can be extracted from equations
(2.14) and (2.15) as
ux; y; z Z
lDf
x
x; y; z
4pn
0
sin q
; 2:16
wx; y; z Z
lDf
z
x; y; z
4pn
1
1 Ccos q
r

Kwx; y; z
1

n
0
1 Ccos q
n
1
1 Ccos q
r

K1
_ _
: 2:17
Putting zZz
1
into equation (2.17) gives
wx; y; z
1
Z
lDf
z
x; y; z
1

4pn
0
1 Ccos q
: 2:18
Substitution of equation (2.18) into equation (2.17) then leads to
wx; y; z Z
lDf
z
x; y; z
4pn
1
1 Ccos q
r

K
lDf
z
x; y; z
1

4p
1
n
1
1 Ccos q
r

K
1
n
0
1 Ccos q
_ _
: 2:19
This suggests that in the case of a sample with a lay-up of different refractive
indexes, or index gradient in the depth dimension, the displacements evaluated
for shallow slices should be used to correct for the displacements at deeper ones.
In the case of rigid body translation of the sample along Cz, then wx; y; zZ
wx; y; z
1
and equation (2.15) reduces to the standard expression for out-of-plane
sensitivity (Huntley 2001).
x
y
z
dx
dz
d
w(x,y)
u(x,y)
d
F
F'
Figure 6. A slice through the gauge volume in the tilt scanning interferometry system showing
depth resolution dz, lateral resolution dx and average displacement d of scatterers.
P. D. Ruiz and others 2490
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
(d ) Gauge volume, depth range and displacement sensitivity
In a symmetric lateral illumination system setup, the measurement volume is
dened by the intersection of the illumination beams during the whole tilt scan
(see gure 7). The depth resolution can be dened as the minimum distance
between surfaces inside the measuring volume whose corresponding interference
signals can be fully resolved in the frequency domain. The usual resolution
criterion is that the frequency difference df between two neighbouring peaks has
to be at least twice the distance from their centres to their rst zero, which is
equivalent to the width of the central lobe. If a window function W(t) is used,
apart from reducing the leakage through secondary lobes, it has the effect of
broadening the spectral lines. A rectangular window of duration T, for example,
results in a sinc function of width dfZ2/T, while a Hanning window has a
spectral width of dfZ4/T. From equation (2.11), the depth resolution is therefore
dz ZK
gl
n
0
jxjDq
; 2:20
where gZ2, 4 for rectangular or Hanning windows, respectively. Although the
rectangular window has superior depth resolution, this is accompanied by the
undesirable presence of large secondary lobes in
~
Wf , which may interfere with
other peaks, leading to phase errors. This effect is signicantly reduced if a
Hanning window is used instead.
If the layout in gure 1 is imaged by a telecentric optical system with
magnication M, then the lateral resolution dx and dy will depend on M and the
spatial resolution of the sensor used in the same way as a conventional camera.
The lateral and depth resolution dene a gauge volume of size dx!dy!dz within
the measurement volume, as shown in gure 6. All scattering points within the
gauge volume centred at (x, y, z) contribute to the interference signal at position
(Mx, My) in the image plane, modulated with frequencies within the range
f Kdf =2; f Cdf =2. After deformation, the scattering points initially inside the
gauge volume centred at (x, y, z) move to a new position xCu; y; zCw,
indicated with a dashed line rectangle in gure 6. w(x, y, z) then represents the
out-of-plane component of displacement vector d, which is an average
measurement
volume
O
R
q q
q+q q+q
Figure 7. The measurement volume is dened by the intersection of the illumination beams during
the tilt scan.
2491 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
displacement of all scattering points inside the intersection of the solid line and
the dashed line rectangles.
From equation (2.3), the total phase change Df introduced in the wavefront
coming from point F, by a tilt angle Dq introduced in the illumination angle in a
time T, is
Dfx; y; z ZDf0; 0; 0 C
2pn
0
Dq
l
x cos qKz
1
sin qKzKz
1
x: 2:21
This phase change introduces CZjDfj=2p modulation cycles in the interference
intensity signal. In order to comply with the Shannon sampling condition, the
number of samples required has to be at least twice the number of cycles:
N
f
R
Df0; 0; 0
p
C
2n
0
Dq
l
x cos qKz
1
sin qKzKz
1
x: 2:22
It follows that for a given central illumination angle and tilting range, the
number of samples N
f
determines the maximum depth a slice can be within the
material in order to be able to determine its displacements, assuming the
scattering, absorption and laser coherence length do not set a lower limit. This
distance is known as the depth range, and is given by
Dz zK
l
2n
0
xDq
N
f
K
Df0; y; 0
p
_ _
Kz
1
sin q
x
K1
_ _
C
x cos q
x
: 2:23
If we think in terms of frequencies, it is easy to realize that the offset frequency
f x; y; 0Zf 0; 0; 0Cf
x
x; y; z ((2.6) and (2.7)) pushes the spectrum towards
the Nyquist frequency, thus reducing the depth range. In order to reduce this
wasted bandwidth, f (x, y, 0) should be as low as possible, but without interfering
with the cross-interference peaks close to the dc.
The displacement resolution s
z
(sometimes called sensitivity of the technique)
is decoupled from the depth resolution dz and only depends on the wavelength of
the laser source and the degree of speckle decorrelation. For out-of-plane
sensitivity, as is the case in gures 2a and 5, it is typically better than l
c
/30.
3. Proof-of-principle experiments
(a ) Experimental setup
In order to test the proposed technique described in 2, we carried out proof-of-
principle experiments, in which the tilt scanning technique was used to measure
displacements elds within a partially transparent scattering sample. The optical
setup is shown in gure 5. A collimated continuous wave (CW) beam of
wavelength lZ532 nm and power of approximately 100 mW is steered by mirror
TM mounted on a tilting stage. Left and right illumination beams are obtained
with the aid of a cube beam splitter CBS and mirrors M
1
and M
2
. TM is tilted by
means of a piezoelectric leadzirconatetitanate (PZT) actuator controlled by a
ramp generator RG.
The imaging system consists of an imaging lens L
1
, eld and relay lenses L
2
and L
3
, respectively, wedge beam splitter WBS and CMOS high speed camera C
(HCC-1000 Vosskuhler). WBS is used with the reective surface closer to the
P. D. Ruiz and others 2492
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
object, to avoid multiple reections of the object beam within the glass thickness.
This is crucial in order to avoid multiple peaks in the spectrum corresponding to
the same depths in the sample. Preliminary experiments with an external BS in
front of L
1
resulted in multiple reections of the object beam, causing mixing of
the spatial information from the interference intensity signal. WBS serves to
recombine reference and object beams onto the camera sensor. A smooth on-axis
reference wave was used rather than a second speckled wave in order to
maximize the signal-to-noise ratio. The purpose of lenses L
2
and L
3
is to place the
WBS between the imaging lens and the camera sensor. A screen S by the CBS
serves for alignment purposes: if the object is replaced by an optical at aligned
with its normal parallel to the optical axis, then the left and right illumination
beams interfere at S after reection (transmission) on the optical at and onto
(through) M
2
, M
1
and (CBS). When the left and right illumination beams
subtend the same angle to the optical axis, the fringes are nulled at S. This is
important because equations (2.14) and (2.15) assume that the angle is the same
for both illumination directions.
Figure 7 shows a close up of the illumination beams in the region surrounding
the object, before and after tilt of the beams by an angle Dq. It can be seen that
the measurement volume is the intersection of the beams at the beginning and
end of the scan range of the tilt angle. Scattering points within this volume give
rise to a continuously modulated interference signal throughout the whole tilt
scanning sequence.
The test object was a beam manufactured in-house with water clear casting
epoxy resin (from CFS Fibreglass Partnership), seeded with a small amount of
titanium oxide white pigment to increase the scattering within the material. The
front and back surfaces of the beam were polished to reduce supercial scattering
and the back face was painted black to eliminate the scattering produced by the
reected component due to internal reection. Under a three-point bending test,
the beam was loaded with a ball-tipped (6 mm diameter) micrometer against two
cylindrical rods (12 mm diameter), as shown in gures 7 and 8. A rough reference
surface R was placed just in front of the object, so as to cover approximately 20%
of the lower portion of the area imaged. As described earlier, this served to
compensate for the shift of the peaks along the horizontal axis x and to allow for
correction of the nonlinear response of the tilting stage at TM.
field of view
reference surface
o
p
t
i
c
a
l
a
x
i
s
x
y
z
loading pin
Figure 8. Schematic of the epoxy resin beam under three-point bending load. In-plane and
out-of-plane depth-resolved displacements were measured through the thickness of the beam.
2493 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
The measurement sequence was as follows:
(i) The beam was preloaded by displacing the spherical indenter at the point
of contact by K0.500 mm to bring the team into position against the
support rods.
(ii) Two reference-state three-dimensional data volumes I
L1
(x, y, t) and
I
R1
(x, y, t) were recorded sequentially with the left and right illumination
beams (with right and left beams blocked, respectively, making use of a
screen between CBS and M
1
and CBS and M
2
, respectively).
(iii) The pin was displaced 40 mm along the Kz-axis to bend the beam.
(iv) As in (ii), two loaded-state three-dimensional data volumes I
L2
(x, y, t)
and I
R2
(x, y, t) were recorded with the left and right illumination beams.
The main measurement parameters were set as follows: camera exposure time
T
exp
Z0:1397 s; framing rate F
R
Z7.16 fps; acquired frames N
f
Z480 frames;
acquisition time TZN
f
T
exp
Z68:6 s; spatial resolution of FOV: 256!256 pixels;
size of FOV: 7.2!7.2 mm
2
; driver ramp voltage range V
pp
Z080 V; tilt
angle scanning range DqZ0.0048 rad; illumination angle qZ458; refractive
indices n
0
Z1 and n
1
Z1.4; laser wavelength lZ532 nm; laser power per beam:
w35 mW CW.
(b ) Results and discussion
From equation (2.5) the refraction parameter is xZ0:4138 close to the
maximum for the corresponding illumination angle and refractive index. The
resulting depth resolution of the system was therefore dzw1.1 mm and the size of
the gauge volume was dx !dy!dz w0:028 mm!0:028 mm!1:1 mm. The
number of frames N
f
was adjusted to guarantee a depth range Dz bigger than
the object depth dZ7.8 mm.
Figure 9 shows the interference intensity signal from a pixel imaging part of
the reference surface (bottom), and from another pixel imaging part of the epoxy
resin beam (top). The former shows a single frequency and corresponds to the
100 200 300 400
0
40
80
120
160
p
i
x
e
l

i
n
t
e
n
s
i
t
y

(
g
r
a
y

l
e
v
e
l
)
frame number
Figure 9. Interference intensity signal due to tilt scanning from a pixel on the epoxy resin beam
(top), and from a pixel on the reference surface (bottom).
P. D. Ruiz and others 2494
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
signal coming from a single depth, while the latter shows a more complex wave
train due to a mixing of frequencies coming from within the thickness of the
sample. Figure 10 shows the magnitude spectrum (also scattering potential) for
left and right illumination obtained along the horizontal axis x, averaged along
the columns of
~
I
L1
x; y; f and
~
I
R1
x; y; f to reduce the noise content. These
correspond closely to the scheme shown in gure 2c. The peak due to the
reference surface and the band corresponding to scattering points through the
whole thickness d of the beam can be clearly seen. Figure 10 should be
interpreted as a top view where both the reference surface and beam cross-
section are visible.
The noisy appearance in the right-hand side of gure 10a is an artefact of the
linearization routine we had to use to correct for the nonlinearity of the tilting
stage. The effect of the linearization routine on the reference surface peak is
shown in gure 11. The reference surface has no appreciable thickness, so it
should appear in the spectrum as a narrow peak. The nonlinear tilt scan causes
an asymmetric broadening of the spectral lines. We have numerically corrected
for this effect by using the signal from the reference surface, as it does not change
Figure 10. Magnitude spectrum, or scattering potential, along the horizontal axis x averaged along
the columns of (a)
~
I
L1
x; y; f for left illumination and (b)
~
I
R1
x; y; f for right illumination.
0
50 100 150 200
500
1000
1500
2000
2500
3000
FFT frequency index
a
m
p
l
i
t
u
d
e

(
a
.
u
.
)
AL
BL
Figure 11. A linearization routine that operates on the temporal axis of the I(x, y, t) volumes
reduces the peak width, and therefore the spectral leakage and depth resolution. The reference
surface peak is shown here before (BL) and after (AL) linearization.
2495 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
after sample deformation. First, the intensity signal I(x, y, t), with x xed and y
corresponding to coordinates in the reference surface, is evaluated along the time
axis. Using the Takeda method of phase evaluation (Takeda et al. 1982), the
phase shift f(x, y, t) introduced by the tilting mirror is obtained from I(x, y, t).
In an ideal system this should be a linear function, but it looks like a slanted S
due to our nonlinear tilting stage. A polynomial is then tted to the inverse
function t(x, y, f) to get an analytical expression used to interpolate t(x, y, f)
and resample it at equally spaced phase shift values f
0
. These new phase shift
values f
0
lead to non-equally spaced sampling times t
0
Ztx; y; f
0
. Finally, the
original signal I(x, y, t) is interpolated and resampled at times t
0
to get a linear
phase shift between consecutive samples. Owing to the frequency dependence on
x (2.6), resampling times t
0
are evaluated for each position x along the horizontal
axis and used to linearize all signals I(x, y, t) in the corresponding column. This
linearization considerably reduces the leakage between spectral components and
improves the depth resolution.
The bandwidth associated with the beam thickness (marked with arrows in
gure 10a) was Df Zf
zmax
Kf
z1
Z0:4 Hz, which from equation (2.11) corresponds
to a beam thickness DzZ7.46 mm. This is within a 4% error of the actual beam
thickness dZ7.8 mm.
A one-dimensional Fourier transform was performed along the time axis of
the four data volumes I
L1
, I
R1
, I
L2
and I
R2
, allowing the calculation of the
optical phase for each (x, y, f ) coordinate in the conjugate spectrum-volumes.
This resulted in two three-dimensional phase-change volumes Df
L
(x, y, f ) and
Df
R
(x, y, f ) corresponding to each illumination direction, the sum of which
gave the out-of-plane-sensitive phase-change volume and the difference between
them gave the in-plane phase-change volume. The top row of gure 12 shows
the wrapped in-plane phase-change distribution for different slices within the
epoxy resin beam starting at the object surface zKz
1
Z0 mm (left) in steps of
1.74 mm down to zKz
1
Z5.22 mm (right). The wrapped out-of-plane phase-
change distribution for the same depth slices are shown in the bottom row of
gure 12.
y

(
m
m
)
2
0
2
2 0 2
y

(
m
m
)
2
0
2
x (mm)
2 0 2
x (mm)
2 0 2
x (mm)
2 0 2
x (mm)
Figure 12. In-plane (top row) and out-of-plane (bottom row) wrapped phase-change distribution
for different slices within the beam. Black represents Kp and white Cp.
P. D. Ruiz and others 2496
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
The depth-resolved in-plane and out-of-plane displacement elds within the
beam corresponding to the wrapped phase-change slices shown in gure 12 are
presented in the top and bottom rows of gure 13, respectively. Phase
unwrapping of each slice was carried out using the algorithm described in
Buckland et al. (1995) as implemented in the Phase Vision Ltd MATLAB phase
unwrapping toolbox. It can be seen that the gradient of the in-plane
displacements reduces as we move from the front to the back surface of the
beam, indicating the expected variation of the tensile state for the rst front
slices. However, the gradient does not reverse to show a compressive state for the
slices behind the neutral axis at zKz
1
Z3.8 mm. As discussed later, this
discrepancy is likely caused by phase changes due to refraction in the curved
surface of the beam and possibly also due to stress-optic coupling. The out-of-
plane displacements show different levels of bending as we approach the back
surface from the front surface. The asymmetry of the distribution is produced by
the position of the point of contact between the loading pin and the beam, which
was approximately 2 mm below the horizontal symmetry axis of the beam. The
last slice at zKz
1
Z5.22 mm starts to reveal detail of the local deformation
around the point of contact, marked with a cross in gure 13.
It is worth noting that the problems with the nonlinear scans described earlier
arose from the use of a home-made open-loop tilting stage. Closed loop PZT tilt
stages are now available commercially with sub-microradian resolution that can
scan a beam through 100 mrad in a fraction of 1 s. At an angle of incidence q of
458, a wavelength of 532 nm, n
0
Z1 and jxjZ0:414, the effective depth resolution
is given by equation (2.20) as 50 mm, equivalent to that in WSI provided by a
tuning range of approximately 20 nm for a material with the same index of
refraction. External cavity diode lasers in the visible or near IR are unable to
provide such a wide range in a single sweep, requiring the use of expensive dye or
Ti : sapphire lasers.
y

(
m
m
)
2
0
2
2 0 2 2 0 2 2 0 2 2 0 2
y

(
m
m
)
2
0
2
x (mm) x (mm) x (mm) x (mm)
(m)
1.5
1.0
0.5
0
0.5
(m)
1.5
1.0
0.5
0
0.5
Figure 13. In-plane and out-of-plane depth-resolved displacements calculated from the
phase-change distributions shown in gure 12.
2497 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
To study objects of arbitrary shape, index matching uid can be used, as
suggested in 2b, to eliminate distortions in the illumination angle due to
refraction at the material interface. The matching uid would also eliminate the
problem due to interface deformation between a reference and a loaded state of
the object.
(c ) Finite-element modelling
In order to validate the results from TSI, we performed a three-dimensional
nite-element analysis of the geometry measured experimentally. Simulation of
the displacement eld distributions resulting from the contact interaction
between two bodies requires, in general, the nite-element discretization of both
the target surface (in this case the spherical indenter) and the contact surface
(epoxy beam). In addition, it is necessary to include contact elements between the
target and contact surfaces to adjust the level of interpenetration between the
target and the contact surfaces. Even though this modelling technique provides
good predictions about displacements and stresses, it is computationally
expensive, as it requires high levels of mesh renement and nonlinear solution
algorithms that increase dramatically the solution time required for a given
problem. In this work a simpler, yet effective, technique was used. Given the large
ratio between the modulus of elasticity of the steel indenter and epoxy beam
(E
s
Z270 GPa; E
g
Z2.41 GPa; E
s
/E
g
O100), the interaction between them can be
simulated using the rigid-to-exible contact approximation (Johnson 1985),
where the indenter is modelled as a single rigid element and the beam is considered
as an elastic medium. This technique reduced dramatically the solution times.
The general purpose nite-element software ANSYS was used. The target
surface and the spherical indenter were treated in the original geometric
conguration and the motion of the entire surface was then controlled by the
imposed displacements on a pilot node that represents the kinematics of the
surface. Therefore, the steel indenter was modelled using the 1-node element
TARGE170 with standard spherical shape of radius 3 mm. The contact surface,
surface of epoxy beam, was modelled using the 8-node element CONTA174, and
the contact kinematics was governed using the pure Lagrange multiplier method
(Bathe 1996) that prescribes zero penetration and zero slip when the contact
occurs. The epoxy beam was modelled using the 20-node isoparametric element
SOLID95, which has three degrees of freedom per node: translations in the nodal
x, y and z directions. The elastic modulus of the epoxy beam was measured using
the three-point bending test (result reported above).
The main objective of the nite-element modelling was to extract the
displacements occurring in the same FOV used during the experiments. In order
to minimize the discretization error (the discrepancy between the discretized FE
model and the mathematical model, considering the latter as exact), the FE
mesh was rened inside the FOV using the regular renement method described
in Cook et al. (1989). The number of elements in this area was incremented
iteratively until the absolute difference between the maximum out-of-plane
displacements of two successive iterations was less than 1%.
The nite-element representation of the indenterbeam system is illustrated in
gure 14. The epoxy beam had working volume dimensions of 75!22.8!
7.6 mm
3
between the two cylindrical supports shown in gure 8. It is convenient
P. D. Ruiz and others 2498
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
to dene a local coordinate system (x
0
, y
0
, z
0
) whose origin is at the bottom of the
left-hand cylinder. In this system, the boundary conditions corresponding to the
cylindrical supports are specied as follows: uZwZ0 along the line formed by
the intersection of the planes x
0
Z0 and z
0
Z0; and wZ0 along the line formed by
the intersection of the planes x
0
Z75 mm and z
0
Z0. The initial preload of the
sample was sufciently large to cause large deformations. Under these conditions
a structures changing geometric conguration can cause the structure to respond
nonlinearly, and this is referred to as geometrical nonlinearities. Two static
nite-element models were therefore solved to simulate the experimentally
measured incremental displacement elds. In the rst, the indenter displacement
was specied as wZ0.5 mm, and for the second wZ0.504 mm. Both models were
x
y
z
Figure 14. Finite-element representation of the contact between the spherical indenter and beam.
y

(
m
m
)
2
0
2
2 0 2 2 0 2 2 0 2 2 0 2
y

(
m
m
)
2
0
2
x (mm) x (mm) x (mm) x (mm)
(m)
1.5
1.0
0.5
0
0.5
(m)
1.2
0.8
0.4
0
Figure 15. In-plane and out-of-plane depth-resolved displacements predicted by a nite-element
model of the bending beam under the loading conditions used during the experiment.
2499 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
solved consecutively (Dell Precision M70, 1.86 GHz, 512 Mb RAM).
The difference in the displacement elds between the two models, (u, v, w),
were compared with the experimental incremental displacement elds over
the experimental observation window (7.2!7.2 mm
2
, centred at x
0
Z36 mm,
y
0
Z8.85 mm). Figure 15 shows the in-plane and out-of-plane displacements
within the beam as predicted by nite-element analysis. This gure can be
compared directly with gure 13, and it is observed that the predicted behaviour
is consistent with the measured results, even to the extent of showing the
asymmetry of the out-of-plane displacement eld with respect to the horizontal
axis of the beam. Figure 16 shows four horizontal proles of the in-plane and out-
of-plane displacementsmeasured experimentally and predicted with the nite-
element modelat a height equal to that of the pin loading contact point. The
experimental two-dimensional phase maps were unwrapped independently of one
another and as a result the phase offset between slices is unknown. For clarity,
arbitrary offsets were therefore added to both the experimental and numerical
results at different depths within the beam. Agreement between the experimental
and predicted out-of-plane displacement elds is excellent. The agreement is
somewhat less convincing for the in-plane displacements, a result that we
attribute to the load-induced curvature of the sample surface. A relatively simple
algorithm to correct for this effect was developed, which produced signicantly
closer correspondence between the experimental and numerical in-plane proles,
but degraded somewhat the agreement between the corresponding out-of-plane
proles. No account is taken of load-induced spatial variations in refractive index
within the beam, and this may be a worthwhile avenue for further development
of the technique.
4. Conclusions
We have proposed a novel technique that we call tilt scanning interferometry to
measure three-dimensional depth-resolved displacement elds within semi-
transparent scattering materials. A simple proof of principle experiment showed
promising results even using a home-made open-loop PZT-driven tilting stage.
3 2 1 0 1 2 3
0
1
2
3
4
5
6
7
8
x (mm)
surface
zz
1
=1.74mm
zz
1
=3.48mm
zz
1
=5.22mm
FEA
exp
0
1
2
3
4
5
6
7
8
3 2 1 0 2
x (mm)
FEA
surface
zz
1
=1.74mm
zz
1
=3.48mm
zz
1
=5.22mm
exp
(a) (b)
o
u
t
-
o
f
-
p
l
a
n
e

d
i
s
p
.

(

m
)
i
n
-
p
l
a
n
e

d
i
s
p
.

(

m
)
3 1
Figure 16. Proles of (a) the in-plane and (b) the out-of-plane displacements measured
experimentally and predicted with a nite-element model.
P. D. Ruiz and others 2500
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
Among the system parameters, the depth resolution is the most important from a
practical point of view, and in our experiment it was dzw1.1 mm for a tilting
range of 0.0048 rad. By means of TSI, the scattering potential within the sample
can be reconstructed in a three-dimensional data volume as in scanning OCT.
Most importantly, in-plane and out-of-plane displacements can be measured
within the object under study with sub-wavelength sensitivity (decoupled from
the depth resolution) and up to a depth exceeding 5 mm. This sensitivity is some
four to ve orders of magnitude better than the intrinsic depth resolution and
could not therefore be achieved by image correlation techniques based on the
magnitude of the spectrum.
We are grateful to the Engineering and Physical Sciences Research Council for nancial support.
J.M.H. also acknowledges support from The Royal Society and Wolfson Foundation in the form of
a Royal Society-Wolfson Research Merit Award.
References
Abe, T., Mitsunaga, Y. & Koga, H. 1986 Photoelastic computer tomography: a novel measurement
method for axial residual-stress prole in optical bers. J. Opt. Soc. Am. A. 3, 133138.
Aben, H., Errapart, A., Ainola, L. & Anton, J. 2004 Photoelastic tomography for residual stress
measurement in glass. In Optical metrology in production engineering, vol. 5457 (ed. K.
Gastinger, O. J. Lkberg & S. Winther), pp. 111. Bellingham, WA: SPIE.
Bathe, K. J. 1996 Finite element procedures. Englewood Cliffs, NJ: Prentice-Hall.
Bay, B. K., Smith, T. S., Fyhrie, D. P. & Saad, M. 1999 Digital volume correlation: three-
dimensional strain mapping using X-ray tomography. Exp. Mech. 39, 217226. (doi:10.1007/
BF02323555)
Born, M. & Wolf, E. 1959 Principles of optics: electromagnetic theory of propagation, interference
and diffraction of light. London: Pergamon Press.
Buckland, J. R., Huntley, J. M. & Turner, S. R. E. 1995 Unwrapping noisy phase maps by use of a
minimum-cost-matching algorithm. Appl. Opt. 34, 51005108.
Cook, R. D., Malkus, D. S. & Plesha, M. E. 1989 Concepts and applications of nite element
analysis. New York: Wiley.
de Groot, P. 2000 Measurement of transparent plates with wavelength-tuned phase-shifting
interferometry. Appl. Opt. 39, 26582663.
Draney, M. T., Herfkens, R. J., Hughes, T. J. R., Pelc, N. J., Wedding, K. L., Zarins, C. K. &
Taylor, C. A. 2002 Quantication of vessel wall cyclic strain using cine phase contrast magnetic
resonance imaging. Ann. Biomed. Eng. 30, 10331045. (doi:10.1114/1.1513566)
Dresel, T., Hausler, G. & Venzke, H. 1992 3-dimensional sensing of rough surfaces by coherence
radar. Appl. Opt. 31, 919925.
Fercher, A. F., Hitzenberger, C. K., Kamp, G. & El-Zaiat, S. Y. 1995 Measurement of intraocular
distances by backscattering spectral interferometry. Opt. Commun. 117, 4348. (doi:10.1016/
0030-4018(95)00119-S)
Fercher, A. F., Drexler, W., Hitzenberger, C. K. & Lasser, T. 2003 Optical coherence
tomographyprinciples and applications. Rep. Prog. Phys. 66, 239303. (doi:10.1088/0034-
4885/66/2/204)
Fitzpatrick, M. E. & Lodini, A. (eds) 2003. Analysis of residual stress by diffraction using neutron
and synchrotron radiation. London: Taylor & Francis.
Gastinger, K., Winther, S. & Hinsch, K. D. 2003 Low-coherence speckle interferometer (LCSI)
for characterisation of adhesion in adhesive-bonded joints. In Speckle metrology 2003, vol. 4933
(ed. K. Gastinger, O. J. Lkberg & S. Winther), pp. 5965. Bellingham, WA: SPIE.
2501 Tilt scanning interferometry
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from
Gulker, G. & Kraft, A. 2003 Low-coherence ESPI in the investigation of ancient terracotta
warriors. In Speckle metrology 2003, vol. 4933 (ed. K. Gastinger, O. J. Lkberg & S. Winther),
pp. 5358. Bellingham, WA: SPIE.
Huntley, J. M. 2001 Automated analysis of speckle interferograms. In Digital speckle pattern
interferometry and related techinques (ed. P. K. Rastogi), pp. 59139. Chichester, West Sussex:
Wiley.
Hutchings, M. T., Withers, P. J., Holden, T. M. & Lorentzen, T. 2005 Introduction to the
characterization of residual stress by neutron diffraction. London: Taylor & Francis.
Johnson, K. L. 1985 Contact mechanics. Cambridge: Cambridge University Press.
Kuwamura, S. & Yamaguchi, I. 1997 Wavelength scanning prolometry for real-time surface shape
measurement. Appl. Opt. 36, 44734482.
Lexer, F., Hitzenberger, C. K., Fercher, A. F. & Kulhavy, M. 1997 Wavelength-tuning
interferometry of intraocular distances. Appl. Opt. 36, 65486553.
Rodr guez-Vera, R., Kerr, D. & Mendoza-Santoyo, F. 1992 Electronic speckle contouring. J. Opt.
Soc. Am. A 9, 20002008.
Ruiz, P. D., Zhou, Y., Huntley, J. M. & Wildman, R. D. 2004 Depth resolved whole eld
displacement measurement using wavelength scanning interferometry. J. Opt. A: Pure Appl.
Opt. 6, 679683. (doi:10.1088/1464-4258/6/7/004)
Ruiz, P. D., Huntley, J. M. & Wildman, R. D. 2005 Depth-resolved whole-eld displacement
measurement by wavelength-scanning electronic speckle pattern interferometry. Appl. Opt. 44,
39453953. (doi:10.1364/AO.44.003945)
Schmitt, J. M. 1998 OCT elastography: imaging microscopic deformation and strain of tissue. Opt.
Express 3, 199211.
Steele, D. D., Chenevert, T. L., Skovoroda, A. R. & Emelianov, S. Y. 2000 Three-dimensional
static displacement, stimulated echo NMR elasticity imaging. Phys. Med. Biol. 45, 16331648.
(doi:10.1088/0031-9155/45/6/316)
Takeda, M., Ina, H. & Kobayashi, S. 1982 Fourier-transform method of fringe-pattern analysis for
computer-based topography and interferometry. J. Opt. Soc. Am. 72, 156160.
P. D. Ruiz and others 2502
Proc. R. Soc. A (2006)
on February 27, 2012 rspa.royalsocietypublishing.org Downloaded from

Вам также может понравиться