Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

IN-VIVO MEASUREMENT OF ELASTO-MECHANICAL PROPERTIES OF

SOFT BIOLOGICAL TISSUES


V. Vuskovic1, M. Kauer2
1 Institute of Robotics, ETH-Zurich, Zurich, Switzerland
2 Institute of Mechanics, ETH-Zurich, Zurich, Switzerland

Abstract. We present a method to determine elasto-mechanical properties of soft biological tissues, and a device able
to perform the required measurements in-vivo. The device permits the controlled application of vacuum to small
spots of organic tissue and registers the small deformation caused, during the whole measurement process. Deforma-
tion is measured with a vision based technique and the grabbed images are processed in real-time to avoid storage
problems. We model biological tissue with a hyperelastic quasilinear viscoelastic material law and determine the
unknown material parameters via inverse finite element methods.

1. Introduction

Mechanical properties of soft tissues give us valuable information about pathological changes and
provide also the basic quantitative input for mechanical models. One example of application of
precise mechanical models of soft tissues can be found in the growing field of surgical simulation.
Particularly important seems to be the development of virtual reality based endoscopic surgical
simulators, e.g. [10, 16, 20]. The need for endoscopic surgery trainers is due to the fact that mini-
mal-invasive surgical techniques require very special skills from the surgeon, which can only be
acquired with extensive training. Force feedback devices of such systems, simulating the contact
of operating instruments with living organic tissue, should ideally provide realistic haptic sensa-
tions resulting from a computer simulation of tissue deformation. Therefore, the selection of an
appropriate constitutive equation for modeling a specific organic material should be followed by
the determination of the actual numerical values of material parameters.
Since the mechanical properties of biological materials are subject to significant changes when
the tissue is prepared for experimental testing, we aim to perform in-vivo measurements. In the
past, a number of experiments were performed on a variety of prepared tissues under different
loading conditions, [2, 8], but there exists very little quantitative data from in-vivo measurements,
[9, 12].

2. Method

One of the main difficulties in performing in-vivo measurements is clearly that it is impossible to
use the classical experimental setups. In-vivo measurements require high safety standards and
must be performed under sterile conditions. A further problem is presented by the difficulty of
keeping track of the boundary conditions during the measurements, because any mechanical inter-
action with the organ in situ, whose parameters are to be determined, causes some unquantifiable
motion of the whole organ.

1
Pipette aspiration of tissue is one experimental setup that fulfils all the above requirements. It has
the advantage to set well defined boundary conditions for the experiment (contact surface
between tissue and pipette). Pipette aspiration has already been used to determine the mechanical
properties of endothelial cells, [6], and been presented as an option to measure soft tissue parame-
ters, [15]. In both references linear elastic models are used and the tip displacement under load of
the deformed tissue is used to determine the elastic material parameters. In this work the identifi-
cation of the tissue parameters is done with inverse finite element routines. We use a quasilinear-
viscoelastic model combined with an isotropic hyperelastic constitutive law. In the process of
parameter determination not only the tip displacement but a two dimensional description of the
deformed tissue surface is used and the deformation is tracked over the whole measurement
period.
The method consists in putting a tube against the target tissue and applying a weak vacuum. The
organ remains fixed to the tube, specifying well defined boundary conditions, and a small defor-
mation of the tissue is caused inside the tube. The measurement process consists in varying the
applied vacuum over time and determining the functions d ( t ) and P ( t ) , where d(t) is a measure
of the surface deformation under the tube; t, the time from the start of the measurement; and P ( t ) ,
the negative relative pressure in the tube at time t. Thus, we track over time the applied negative
pressure and the resulting deformation. Measurement of more than just one-dimensional force-
displacement information leads to more realistic parameter determination, whereas measuring the
deformation over time permits estimation of time-dependent properties such as viscoelasticity.
Assuming axisymmetry and homogenous tissue in the portion covered by the aspiration tube, a
complete description of the deformation is given by a profile of the aspirated tissue (or more
exactly by one half of a profile). The aspiration setup and the profile to be determined are illus-
trated in Figure 1.

Profile to be measured
z
z
r

(side-view)
r

Figure 1. Deformation profile

Therefore, the measuring process reduces to the determination of the functions z ( r, t ) and P ( t ) ,
where z is the tube symmetry axis, r the radial distance from the center, and P ( t ) the relative neg-
ative pressure measured in the tube at time t . Figure 2 shows the result of an experiment per-
formed ex-vivo on pig kidney cortex. It is possible to observe how the profile of the aspirated
tissue follows the pressure curve. The measurements are thereafter passed off-line to the parame-
ter estimator, which determines the parameters of the finite-element model of the investigated tis-
sue via the inverse finite element method by minimizing the squared differences between
measured and simulated load-displacement relations.

2
One of the major problems in the identification of parameters of biological materials is to estimate
the initial state of stress. Here, a stress free initial configuration is assumed.

negative relative pressure [mbar]


120

100

of profile [mm]
z-coordinate
80

60

40

20

0
0 2 4 6 8 10 12 14 16

time [s] r [mm] time [s]

Figure 2. Acquired data from one experiment

3. Modelling of tissue

3.1. Mechanical description of soft tissues

The mechanical behavior of soft tissues is very complex. Anisotropy, viscoelasticity and non-lin-
ear behavior are some of the important aspects in modelling tissues. Anisotropy is very difficult to
determine under in-vivo conditions. Without any knowledge of the directions of anisotropy the
determination of all parameters seems almost impossible. Therefore we concentrate on the
description of the non-linear and viscoelastic behavior. We use two different material laws for soft
tissues: one is similar to the Neo-Hookean material law and the other one a description similar to
the one of Veronda-Westmann, [1, 8]. The strain energy function for the first is:
1 2
W = µ ( J 1 – 3 ) + --- κ ( J 3 – 1 ) , (1)
2
whereas for the Veronda-Westmann like material:
µ γ (J – 3) 1 2
W = --- e 1 + --- κ ( J 3 – 1 ) . (2)
γ 2
µ [N/m2] and γ [] are material parameters; the bulk modulus κ [N/m2] allows for slight compress-
ibility and is not included in the identification process but set large compared to the other material
constants to account for near incompressibility. J 1 is the first reduced invariant of the right
Cauchy deformation tensor C. Using reduced invariants is equivalent to separating the deforma-
tion in an isochoric part:
* –1 ⁄ 3 *
C = I3 C , with det ( C ) = 1 , (3)
and a pure dilatational part with:
1⁄3
I 3 I , because (4)
* –1 ⁄ 3 1⁄2
I 3 = det ( C ) and I 1 ( C ) = I 3 I 1 ( C ) = J 1 , whereas J 3 is defined as J 3 = [ det ( C ) ] .

3
The 2-nd Piola Kirchhoff stresses Se can be derived from eq. (1) and eq. (2) by:
∂W e
, S ij = (5)
∂ E ij
where Eij denotes the components of the Green-Lagrange strain tensor. Viscoelasticity can be
described by means of a relaxation function G(t):
t
∂ e
S(t ) = S (t ) + ∫ G(t – s)
e
S ( s ) ds , (6)
∂s
0
where eq. (6) corresponds to a quasilinear viscoelastic formulation. Choosing G(t) as a series of
exponentials, allows to approximate the convolution integral and requires only the stresses of the
last time step, [18], and not of the whole load history to be saved. Soft tissues show small sensitiv-
ity to the rate of deformation over the range of some decades. This property can be modelled with
a continuous relaxation spectrum, which could also be approximated by a series of exponential
relaxation functions. In order to keep the number of parameters in the identification small, we
chose:
–( t – s ) ⁄ τ
G ( t – s ) = Ce , (7)
where C [] and τ [s] again are material parameters.

3.2. Finite Element Calculations

To solve the geometrically and materially nonlinear problem of tissue deformation, we use the
standard finite element methods described by Sussman and Bathe, [3]. Since the optimization
algorithm used in the inverse method needs the forward solution of the finite element model for
the estimation of the gradient of the residual function with respect to the unknown parameters, we
restricted our model to axial symmetry. This strongly reduces the computational cost of the
inverse solution. The governing dynamic equations for the system are:
Mu̇˙ + Cu̇ = R – F , (8)
where u denotes the nodal displacement vector, M the mass-matrix, C the damping matrix, R the
externally applied forces and F the equivalent nodal forces. If we diagonalize the mass-matrix
with a lumping process and assume the damping matrix to be proportional to the mass-matrix a
central difference approximation of eq. (8) yields:
u ( t + ∆t ) = ----------------------------------  ( R – F ) + u ( t )M  -------2- + -------2- – u ( t – ∆t ) -------2- , where
1 2 2ξ M
(9)
 -------- + -------- M
1 2ξ    
∆t ∆t ∆t
 2 2
∆t ∆t
∆tC = 2ξM. (10)
∆t stands for the time step of the central difference method and ξ is the damping factor. The time
step is restricted by stability conditions and must be smaller than the theoretically admissible
value of ∆t crit :
κ
∆t ≤ ∆t crit = minlength ⁄ --- , (11)
ρ

4
where minlength is the minimal length of all edges of the finite element mesh, κ stands for the
bulk modulus of the material and ρ is the density of the material.
We use a Total Lagrangian approach with separate pressure interpolation (u/p-formulation) for the
element formulation. The domain is discretized with 4 node bilinear elements having an addi-
tional degree of freedom for the separate pressure interpolation at the centre of the element. Con-
tact is treated as deformable-undeformable contact between the tissue and the aspiration pipette.
Friction is modelled with a radial return approach, which accounts for elastic sticking and non-
elastic sliding, [17], as illustrated in Figure 3.

As long as the absolute value |∆urel| of the relative dis-


τfr/N
placement of a node in contact from its first point of con-
tact with the rigid surface is smaller than utol, it is pulled
back to this point by a force proportional to this distance,
µfr to the friction coefficient µfr and to the normal force N:

∆u rel ∆u rel ∆u rel


- ---------------- µ N = – ------------
τ fr = – --------------- - µ N for ∆u rel < u tol
u tol ∆u rel fr u tol fr
utol |∆urel|
If the node moves outside this region of elastic sticking, a
elastic sticking non-elastic sliding force proportional to the normal force in the opposite
direction of the relative motion is applied:

∆u rel
- µ N for ∆u rel ≥ u tol
τ fr = – ---------------
∆u rel fr
Figure 3. Radial return model for friction

4. Measurement device

We developed a vision technique to measure the profile illustrated in Figure 1. Slight modification
of the tube’s geometry makes it possible to place a small mirror beside the aspiration hole and to
track the silhouette of the deformed tissue with a camera placed at the other end of the tube. This
periscope-geometry is illustrated in Figure 4.
Camera Connection to the
pneumatic system

Optical fiber
with connector Lens

Silicon tube

Aspiration hole

Mirror

Figure 4. Design and photograph of the instrument

5
The method allows to measure the desired profile very accurately, rapidly and without contact
with the tissue. An optical fiber fixed in the tube, illuminates the aspirated tissue. The camera with
close-up lenses, the tube, the light-fiber, the pressure sensor and the control switches are mounted
onto an aluminium body, which also gives maneuverability to the instrument.The pneumatic sys-
tem is connected to the aspiration tube via a flexible silicon tube and is designed to allow easy and
secure control of the pressure in the tube and to measure the pressure very accurately. All the
components of the pneumatic system are inert and can be sterilized.

4.1. Supporting hardware and software

Magnetic valves of the pneumatic system, the vision system, the pressure sensor and all other
instrument controls are connected to a 300 MHz PowerPC based computer system, running under
the hard real time operating system XOberon, [19]. The grabbed images are processed in real time
and the acquired curves, pressures and times collected during a measurement are stored in the
RAM of this target computer-system. By pushing a button on the instrument, the data from the
whole measurement period are transferred from the RAM-based file system of the target-system,
via an ethernet connection, to a remote data storage, residing on the host system (laptop PC). A
graphical feed-back from the instrument is given through a web-based interface on the host-sys-
tem, provided by a web-server running on the PowerPC board, which calls a visualizing applica-
tion as a CGI process. That way it is possible to monitor the functionality of the instrument during
measurements, observing the acquired images and the extracted profiles in real-time, with a 6 Hz
rate on an intranet. This graphical feed-back makes it possible to interact better with the instru-
ment, to acknowledge when a measurement period is finished, and to briefly check the correct-
ness of measurements. The whole system is shown in Figure 5.

Pneumatic system, analog


electronics and light source

Host system with program-


ming environment, visual inter-
Target system face and mass data storage
(PowerPC VME Board)
with Frame-Grabber

Ethernet
(Http, Ftp - TCP/IP)

Figure 5. Hardware

6
5. Image processing

If the deformation of tissue is to be tracked over a period of time, a very large number of images
must be taken. To reduce the quantity of information, a real-time contour extraction of measured
profiles is of great advantage. In our setup we achieve a processing frequency of 25 Hz, i.e. we
grab and process 25 images per second. A typical image grabbed during the measurement process
(ex-vivo experiment on bovine liver) and a short scheme of the image processing algorithm are
shown in Figure 6.
Grabbed image

Thresholding Edge-Extraction

Profile-Extraction

Figure 6. Algorithm for profile extraction

The approach chosen here combines the precision of edge extraction with the ease of interpreta-
tion of thresholded images. The thresholded image gives an approximation of the searched con-
tour, while the edge image gives precise positions of candidate points.

7
5.1. Thresholding

One of the advantages of the processing method in Figure 6 is that the thresholding block must
not provide exact information, but just approximate the silhouettes. However it is not possible to
use a simple single-value thresholding, because the illumination of the aspirated tissue is not
homogeneous and because different organic tissues can have completely different reflection prop-
erties. The method used to evaluate rapidly the silhouettes is presented in Figure 7. In a first step
the image is thresholded with the level calculated by applying the SIS-algorithm (Simple Image
Statistics algorithm, [4, 5, 7]) to the whole image. Then, the left and right extremes of the bina-
rized image are determined and the same procedure is repeated for the two rectangles around the
extremes. This is repeated until the extremes are found at the hole level or until a maximum num-
ber of steps is reached.

Figure 7. Thresholding algorithm

8
5.2. Edge-extraction

The fact that edge extraction is decoupled from the rest of the process, makes it possible to select
the edge detection algorithm making a trade-off between the execution time and the achieved
accuracy. We implemented several simple and fast classical edge detectors, as Roberts and Sobel
operators, and several optimized derivation filters, discussed in [13] and [14] and decided to use a
derivative filter optimized with a weighted least-squares method, whose mask is shown for the x-
direction in eq. (12), [13]. Same coefficients are used to build the derivation filter for the y direc-
tion, Dy.
D x = 0.0420264 – 0.229945 0.833812 0 – 0.833812 0.229945 – 0.0420264 (12)
Subsequently, a thresholding is performed on the modulus of the gradient field (with non maxi-
mum suppression in the gradient direction). The obtained edge image is used by the profile-
extraction block to identify the pixels to be considered candidates to belong to the silhouette con-
tour. Binarizing the gradient field, helps to limit the number of candidate pixels, reducing drasti-
cally the computation time needed to choose the contour points among them.

5.3. Profile-extraction

In order to describe how the contour is extracted, let us first define the pixels coordinates of the
images as p = ( x, y ) , with 1 ≤ x ≤ n and 1 ≤ y ≤ m , where n and m are the image dimensions.
Let also the functions X ( p ) = x and Y ( p ) = y be the x and y coordinates of a pixel p . We can
now define the profile-extraction block as a discrete function y = f ( x ) , that assigns to every
coordinate x exactly one coordinate y .
In a first step, the information from the edge and from the thresholded images is elaborated:
• The edge image provides a set of candidate contour points, that can be sorted in sets relative to
their x-coordinate, as:
D ( x ) = { p 1, …, p k ( x ) } , where (13)
p i stands for p i ( x ) , and k(x) is the number of candidate points for x.
• The thresholded image on the other hand allows an easy extraction of the approximated sil-
houette contour:
yt = f t ( x ) (14)
The selection of the y coordinate for a given x is done by minimizing a cost function C ( p ) ,
which is minimized over the set D ( x ) ∪ { ( x, f t ( x ) ) }, which contains the k ( x ) candidate values
and the pixel given by the approximated contour y t = f t ( x ) . We define therefore the f(.) function
as:
f ( x ) = Y ( p m ), with p m ∈ D ( x ) ∪ { ( x, f t ( x ) ) }, so that

 
C ( p m ) = min  min { C ( p ) }, C ( ( x, f t ( x ) ) ) + Const  (15)
 p ∈ D( x) 
In the minimization process we also consider the points of the approximated contour obtained
through thresholding, but we penalize them with a constant factor, in order to give preference to

9
points from the edge image. The cost function C ( p ) takes into account the vertical distance of the
candidate pixel p from the silhouette contour given by the thresholded image, the vertical dis-
tance from the last extracted contour point (at the coordinate x-1), and the modulus of the gradient
field ( ∇im ( x, y ) ]):
( p ) = α 1 ⋅ Y ( p ) – f t ( X ( p ) ) + α 2 ⋅ Y ( p ) – f ( X ( p ) – 1 ) – α 3 ⋅ ∇im ( X ( p ), Y ( p ) (16)
The constants α 1, α 2 and α 3 are positive weight factors and have been determined experimen-
tally. The superposition of an original image and the extracted contour is given in Figure 8.

Figure 8. Contour superimposed to image


With some modifications this method could also be applied to contour extraction for objects with
a free form.
The extracted silhouette y = f ( x ) can subsequently be transformed in the aspirated tissue profile
z = f asp ( r ), with y corresponding to z and x corresponding to r , with a linear transformation
described in subsection 5.4.

5.4. Calibration and resolution

We performed the calibration of the vision system with the simple device illustrated in Figure 9.
The method consists in introducing a tip of known length relative to the tube base and measuring
it with the vision system.

Mirror Instrument tube

Fixation
screw

Conus
tip

Build-in Micrometer

Figure 9. Calibration device


The obtained calibration curve is almost linear for the region of interest, with 1 pixel in z-direc-
tion corresponding to 31.75 µm . Introducing a cylinder with 1 cm diameter in the aspiration hole,
we also determined the pixel-width in the r-image-direction to be 62.9 µm . This fixes the sam-
pling space frequency in the r-direction and the resolution for the deformation measurements in z-

10
direction, and defines the transformation between the contours in image space y = f ( x ) and the
deformation curves z = f asp ( r ).

5.5. Execution times

For the part of interest of images of 180 per 180 pixels the processing times for some of the imple-
mented algorithms are given in Table 1. With the chosen edge-extraction algorithm, we need
39 ms (the sum of terms with a ‘*’) to process an image, achieving so the 25 Hz performance.

Algorithm time

DMA-Frame Grabber to Computer RAM 7 ms*


Thresholding 7 ms*
Sobel Operator 10 ms
Derivative filter 3rd grade optimized with Taylor series method ([13]) 15 ms
Least-squares optimized filter from eq. (12) 17 ms*
Thresholding of the edge image with non-maximal suppression in the gradi- 3 ms*
ent direction
Profile-Extraction block 5 ms*
Table 1: Execution times

6. Parameter estimation and results

In this section we show first results of the evaluation of tissue aspiration experiments performed
ex-vivo on pig kidney cortex.

6.1. Mesh generation for the finite element model

In the here presented experiments we used dead pig kidneys. Though the experiments were per-
formed within a few hours after the purchase, it is not exactly known how long the kidneys have
been kept refrigerated before. Thus, the obtained data cannot be compared to data gained from
fresh kidneys or in-vivo experiments. The aim of this experiments is to demonstrate the ability of
the process to perform in-vivo measurements in the future.
As already stated, the deformation of the tissue is tracked with a mirror, thus resulting in a side-
view of the tissue surface. Due to the finite thickness of the base of the aspiration tube (0.5 mm),
not all of the surface of the deformed and undeformed tissue is visible with this tracking system,
as shown in Figure 10. The contour of the visible part of the undeformed tissue surface is used to
generate the finite element mesh for the identification model. The non-visible part of the mesh at

11
the edges of the aspiration hole must be guessed. The quality of the guess can be controlled since
the non-visible part of the tissue partially moves into the visible range with the load being applied.

z [mm]
2 line of visibility
base of the aspiration
1.5
tube
1
0.5
0
r [mm]
−0.5
−1
−5 −4 −3 −2 −1 0 1 2 3 4 5 Axisymmetric Finite Element Mesh
pixel data initial guess
smoothed data

Figure 10. Generation of undeformed mesh from tissue data


The contour data is extracted from pixel information and therefore doesn’t represent smooth
curves. Smoothing of this data is done with a least squares routine. The smoothed data is then
used to generate the finite element mesh. This process of mesh generation is shown in Figure 10.
In the identification only the part of the surface above the line of visibility is used to calculate the
objective function res(.) (light grey surface in Figure 11)

z-coordinate of profile [mm]


z-coordinate of profile [mm]

radius [mm] time [s]


time [s] radius [mm]

Figure 11. Two views of same deformation profile tracked over time
Figure 11 shows the deformed profile of one aspiration experiment over time. Since rotational
symmetry (with z as axis of rotation) is assumed, only one half of the profile is used in the identi-
fication of the material parameters. The black mesh shows the measurement. The light part of the
underlying surface is the resulting visible part of the profile from the identification, whereas the
dark part shows the initially guessed part of the profile.
The measurements are characterized by a steep initial deformation, which shows the very small
initial stiffness of the tissue, Figure 12. After an almost constant pressure period of about 3 sec-
onds the tissue is released again.

12
The surface of the kidneys showed a strong tendency to stick to the surface of the aspiration tube.
For this reason the friction coefficient can be set to the maximum value of 1.0 in the identifica-
tion. The parameter utol in the friction model was set to 1e-4 [mm].

negative relative pressure [mbar]


120

negative relative pressure [mbar]


120

100 100

80 80

60 60

40 40

20 20

0 0
0 2 4 6 8 10 12 14 16 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

time [s] tip displacement [mm]


Figure 12. Pressure measurement over time and pressure versus tip displacement hysteresis

6.2. Parameter Identification

In the parameter identification simulated data are compared to experimentally obtained data. The
simulation of the experiment gives the deformation in a finite number of nodes, the element
nodes. In a first step, the nodes defining the deformation profile are extracted and a polynomial fit
is performed to get continuous simulated profiles. Sampling the continuous simulated data at the
same times t1....tk and radial positions r1....rl, as in the experiment, we can define the objective
vector function as:
res [ m ] ( p ) = [ z sim ( r j, t i, p ) – z meas ( r j, t i, p ) ] . (17)
m=i⋅ j
Objective of the parameter identification algorithm is to minimize the function:
k⋅l k l
1 1 1
∑ res [ m ] ( p ) = --- ∑ ∑ [ zsim ( r j, t i, p ) – zmeas ( r j, t i, p ) ] .
T 2 2
- res ( p ) res ( p ) = --- (18)
2 2 2
m=1 i=1j=1
The parameter vector p corresponding to the minimum is assumed to represent the real tissue
parameters.
In the identification we use the Levenberg-Marquardt Algorithm from the IMSL Math-Library,
which has already been used in other finite strain applications, [9, 11].
From a current point p c of parameters, the algorithm uses the trust region approach
min r ( p c ) + J ( p c ) ( p n – p c ) 2 , (19)
subject to pn – pc 2
≤ δc
to get to a new point p n :
T –1 T
p n = p c – ( J ( p c ) J ( p c ) + µ c I ) J ( p c ) res ( p c ) (20)
where:

13
T –1 T
µ c = 0 ifδ c ≥ ( J ( p c ) J ( p c ) ) J ( p c ) res ( p c ) 2 (full Newton step without step restriction),
and µ c > 0 , otherwise. (21)
The values µc and δc are set by the algorithm. res ( p c ) and J ( p c ) are the function values and the
Jacobian at the current point p c , respectively.
Since J ( p c ) is not known analytically it is approximated by finite differences. The optimization
starts with an initial guess for the parameter vector p c and is repeated until one of the stopping
criteria is satisfied (different absolute or relative convergence criteria are available).

6.3. Results

Due to the strong non-linearity of the pig kidney cortex the Neo-Hookean Material law failed to
fit the experimentally obtained data in an adequate manner. Therefore we only show results for
the Veronda-Westmann-like material law.
As the right side of Figure 13 shows, the major principal stretch in the tissue during the aspiration
experiment with maximum relative pressure of 100 mbar is about 25%. Therefore the stress-strain
curve is assumed to be reliable only up to this value. The cortex of the kidney shows a very strong
non-linearity in the stress-strain curve. Up to stretches of about 20% the tissue doesn’t show
nearly any stiffness. This flat part of the stress-strain curve is then followed by a strong rise in the
material stiffness.
The bulk modulus κ plays a very important role in the identification process. It has to ensure near
incompressibility. Since the material shows a strong non-linearity, κ must be set very high to
ensure that it is much larger than the other elastic constants (linearized elastic modulus) even
when the material behaves very stiff at large strains.

κ 105 106 107


µ [Pa] 517 728 693
γ [] 23.1 11.9 12.0
C [] 0.44 0.44 0.41
τ [s] 3.4 5.1 4.7
Table 2: Influence of the bulk modulus κ

Table 2 shows the results for three identifications performed on the same measurement with dif-
ferent bulk moduli. The bulk modulus has the strongest influence on the parameter γ, which
describes the non-linearity of the material behaviour. The larger the bulk modulus the smaller the
theoretically admissible time step. A factor of 100 in the bulk modulus means a factor of 10 in
computation time.
Another important aspect is friction between the base of the aspiration tube and the tissue. The
influence of the friction coefficient on the resulting parameters is however not as strong as the one
of the bulk modulus, as demonstrated by Table 3. The ‘sticking’ in Table 3 means, that a node
sticks to the contact surface and doesn’t move as long as the contact force normal to the surface is
positive. As soon as the contact force changes sign the node is released again. Again, the influ-

14
ence of friction on the identification is the strongest on the parameter γ.

µfr [] µ [Pa] γ [] C [] τ [s]

0.0 693 12.0 0.41 4.7


sticking 700 10.5 0.44 4.8
Table 3: Influence of the friction coefficient µfr

In order to validate the results of the identification, we performed a series of uniaxial classical ten-
sile experiments on stripes of the same pig kidney. Subsequently we compared the data from the
tensile experiments with theoretically predicted data obtained with parameters from Table 3. As
the left part of Figure 13 shows, the predicted stress-strain curves with the parameters from Table
3 compare well with stress-strain data measured on the same organ with uniaxial experimental
tensile setups (up to stretches of 25%).

0.016
z [mm]
Nominal stress [N/mm2]

2
3 different uniaxial 0.2
0.014
tensile experiments 0

0.012
........... Prediction for −2
0.1

µfr = 0
−4
0.01
−6
0
0.008 Prediction for
µfr = “sticking”
−8

0.006 −10
−0.1

−12
0.004

−14 −0.2
0.002
−16

0 −0.3
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 −20 −15 −10 −5 0

Nominal stretch [ ] r [mm]


Left hand side: Theoretical stress-strain curve for uniaxial tension at a loading speed of 10
[mm/min] for the parameters in Table 3 and data from tensile tests.
Right hand side: Plot of the major principal strain in the tissue during the simulation of the
aspiration experiment
Figure 13.

7. Conclusions

The presented tissue aspiration method in combination with the inverse finite element character-
ization showed to give good results up to stretches that are equivalent to the maximum positive
major strain in the aspiration experiment. The quantization of friction between the aspiration tube
and the tissue hasn’t been done yet and thus remains one of the major targets in our future work.
Also, measurements on dead animal tissue, and on animals and humans in-vivo will be soon per-
formed and the identified material parameters will be compared.

15
References

1970
[1] Veronda D. R., Westmann R. A. : “Mechanical Characterization of Skin-Finite Deforma-
tions”, J. Biomechanics, Vol. 3, pp. 111-124.
1973
[2] Yamada H. : “Strength of Biological Materials”, Robert E. Krieger Publishing Company
Huntington, New York.
1987
[3] Sussman T., Bathe K. J. : “A Finite Element Formulation for Nonlinear Incompressible Elas-
tic and Inelastic Analysis,” Computers and Structures, Vol. 26, No. 1/2, pp. 357-409.
1985
[4] Kittler J., Illingworth J., Foglein J. : “Thresholding selection based on a simple image statis-
tics”, Computer Vision, Graphics and Image Processing, Vol. 30, pp. 125-147.
1988
[5] Sahoo P. K., Soltani S., Wong A. C., Chen Y. C. : “A survey of thresholding techniques”,
Computer Vision, Graphics and Image Processing, Vol. 41, pp. 233-260.
[6] Theret D. P., Levesque M. J., Sato M., Nerem R. M., Wheeler L. T. : “The Application of a
Homogenous Half-Space Model in the Analysis of Endothelial Cell Micropipette Measure-
ments”, ASME J. Biomech. Eng., Vol. 110, pp 190-199.
1990
[7] Lee S. U., Chung S. Y., Park R. H. : “A comparative performance study of several global
thresholding techniques for segmentation”, Computer Vision, Graphics and Image Process-
ing, Vol. 52, pp. 171-190.
1993
[8] Fung, Y. C. : “Biomechanics: Mechanical Properties of Living Tissues,” Springer-Verlag,
New York.
1995
[9] Moulton, M. J. et al. : “An Inverse Approach to Determining Myocardial Material Proper-
ties,” J. Biomechanics, Vol. 28, pp. 935-948.
[10] R. Ziegler, W. Mueller, G. Fischer, M. Goebel: “A Virtual Reality Medical Training System”,
Proc. First International Conference on Computer Vision, Virtual Reality and Robotics in
Medicine, CVRMed’95, Nice, Lecture in Comp. Sci., 905, Springer Verlag, pp. 282-286.
1996
[11] Kyriacou S. K., Schwab C., Humphrey J. D. : “Finite Element Analysis of Nonlinear Ortho-
tropic Hyperelastic Membranes,” Computational Mechanics, Vol. 18, pp. 269-278.
[12] Carter F. J., Davis P.J. : “Biomechanical and physical modelling of the handling characteris-
tics of tissues and organs during surgical interventions”, DARPA progress reports.
1997
[13] Jaehne B. : “Practical handbook on image processing for scientific applications”, CRC Press,
Boca Raton, New York, pp. 401-413.

16
[14] Jaehne B. : “Digital image processing”, Springer, Berlin.
[15] Aoki T. et al. : “The Pipette Aspiration Applied to the Local Stiffness Measurement of Soft
Tissues”, Annals of Biomedical Engineering, Vol. 25, pp. 581-587.
[16] S. Gibson et al. : “Simulating Arthroscopic Knee Surgery using Volumetric Object Represen-
tation, Real-Time Volume Rendering and Haptic Feedback”, Proc. CVRMed’97, Springer-
Verlag, pp. 369-378.
1998
[17] Berg H. : “Prozessoptimierte numerische Verfahren zur Auslegung von wirkmedienunte
stützten Umformvorgängen”, Diss. Nr. 12394, ETH Zürich.
[18] Puso, M. A., Weiss, J. A. : “Finite Element Implementation of Anisotropic Quasi-Linear Vis-
coelasticity Using a Discrete Spectrum Approximation”, J. Biomech. Eng., Vol. 120, pp 62-
70.
[19] Brega R. : “A Real Time Operating System Designed for Predictability and Run-Time
Safety”, MOVIC ‘98, Zurich, Switzerland, Vol. 1, pp. 379-384.
[20] G. Szekely et al. : “Virtual Reality Based Simulation of Endoscopic Surgery”, Presence in
press.

17

You might also like