Vous êtes sur la page 1sur 5

THE EFFECT OF DISC DEGENERATION ON MECHANICAL

PROPERTIES OF INTERVERTEBRAL DISCS UNDER SHEAR


LOADING
By Ryan Quarrington

1. Introduction
Degeneration of the intervertebral disc (IVD) is
related to increased age and has a great effect on
its load carrying capacity. After initiation of
degeneration, the mechanics of the intervertebral
joint will be affected and further degeneration may
occur [1, 2]. A feature of IVD is annular lesions,
which are characterized by concentric tears, radial
tears and rim lesions. The effects of these lesions
on the biomechanical properties of the IVD have
been studied previously; both in-vivo and in-vitro
in sheep, as well as in cadaveric human specimens
[1-4]. It has been established that sheep spines are
an appropriate model for the human spine in
biomechanical testing [5].
Thompson et al. conducted experiments on invitro intervertebral joints from the lumbar region
of sheep to assess the effect of annular lesions on
mechanical properties under flexion/extension,
lateral bending and axial rotation loading [Error!
Reference source not found.]. They found that
the presence of rim lesions reduced the stiffness of
the intervertebral joint and radial tears caused
reduction in hysteresis loss coefficient (HLC).
Interestingly, this study did not apply a
compressive pre-load or shear loading conditions
to the specimen. Resistance to shear loading is an
important function of the cervical spine [6].
Therefore, this paper addresses this by analysing
the effect of annular lesions on the mechanical
properties of sheep IVDs under shear loading.

2. Methods
2.1. Specimen Preparation
Functional spinal units (FSU) (N=2) were
disarticulated from the spine of a sheep and
included the C5/C6 and C7/T1 vertebrae. The
spine was thawed prior to removal of excess tissue
and dissection, leaving only the relevant vertebrae
and IVDs. The anterior-posterior (AP) depth (D)
and mediolateral (ML) width (W) of the IVD was
measured at the centre of the disc using Vernier
callipers, and from this the approximate area of the

disc was calculated. These values were then used


to calculate the compressive load required for a
0.2MPa pre-load.
Each specimen was potted to the approximate
sagittal centre of each of the superior and inferior
vertebral bodies in dental stone (Investo White
Diestone, Ainsworth Dental, NSW). Measures
were taken to ensure each specimen was potted to
an appropriate depth to maintain a consistent
moment arm distance. Prior to potting, screws
were inserted into the discs and lengths of chicken
wire were looped around the transverse processes.
These screws and the wire ends were immersed
within the dental stone to assist in securing the
FSU in the potting. Hydration of the specimens
was maintained throughout preparation and testing
by wrapping the exposed areas in damp gauze.
2.2. Loading Protocol
A photo of the testing setup can be seen in
Figure 1. The potted specimen ends were aligned
and fixed in an Instron (Coronation Road, High
Wycombe, Bucks HP12 3SY, United Kingdom)
mechanical testing machine using specially
designed aluminium fixation cups. A 100N
compressive load was applied to each specimen
throughout testing to induce a physiologicallyrelevant 0.2MPa pre-load, calculated from the
average area of each specimen [7].
5 continuous shear displacement cycles
between 3.5mm (positive: anterior shear,
negative: posterior shear; where anterior shear is
anterior translation of the superior vertebral body)
at a rate of 0.1Hz were performed to the initially
undamaged specimens. Annular lesions were then
introduced to the IVD of each specimen by the
method described in [3] (Appendix A), and testing
was repeated. A second defect was then
introduced, increasing the width of the lesion to
10mm and another test was performed to assess
the effect of this further degeneration. Final testing
was also conducted on the second FSU after an

incision was performed which spanned the entire

length of the IVD.

Upper load cell

Anterior surface of
potted FSU specimen

Fixation cups

Lower load cell

Figure 1: Photo of testing setup with key aspects indicated.

Tests were conducted twice for each degenerative


condition, except for the second lesion of the first
FSU (C7/T1), which was only performed once.
2.3. Data Analysis
Force-displacement data from the final cycle of
each test were analysed to determine stiffness,
hysteresis loss coefficient (HLC) and maximum
load in both the anterior and posterior shear
loading directions. The first four cycles of each
test were considered pre-conditioning cycles and
were omitted from analysis. The loaddisplacement plot of the final cycle from each test
can be seen in Appendix B.
Values for the derived properties were
calculated using a custom program written in
MATLAB (The Mathworks, Mattick, MA), which
is documented in Appendix C. Stiffness was
calculated from the slope of the linear region of
the load-displacement plot by fitting a linear
regression between
0.5mm
and 1.5mm
displacement (positive for anterior and negative
for posterior) (Figure 2). HLC was defined as the
hysteresis area divided by the total area under the
loading portion of the curve, and was calculated
using numerical integration. For those conditions
where multiple tests were performed the values
were averaged to obtain a single result for each
condition.

3. Results
Stiffness values ranged from 49N/mm up to
99N/mm (Table 1), which appears consistent with
current literature [3, 6]. Stiffness in the anterior
shear loading direction appeared to decrease as the
severity of the lesion was increased, particularly in
the C7/T1 FSU (Figure 3 A and B). No clear
trends were observed in the posterior loading
direction and there didnt appear to be any distinct
variation between loading directions.
Hysteresis loss coefficient values ranged from
0.148 to 0.341 (Table 1). Increasing severity of the
induced annular lesion looked to decrease HLC in
the anterior direction for C7/T1 decreased it in the
posterior direction for C5/C6 (Figure 3 C and D).
The maximum peak load value of 389N was
experienced in the anterior loading direction by
C7/T1, while the minimum of 263N was
experienced in the same direction by C5/C6 (Table
1). These values are within the same order of
magnitude as those recorded in similar studies and
therefore appear valid [6]. As with stiffness, these
values appeared to decrease with increasing lesion
severity in the anterior loading direction; however,
no obvious trends were observed in the posterior
direction (Figure 3 E and F).

Table 1: Derived mechanical properties for the C7/T1 and C5/C6 FSUs in the anterior and posterior shear loading directions for the various
degenerative conditions. Stiffness in N/mm, peak load in N, HLC = hysteresis loss coefficient, ant. = anterior, post. = posterior.
C7/T1
No Lesion

Stiffness

Lesion 1

C5/C6
Lesion 2

No Lesion

Lesion 1

Lesion 2

Lesion 3

Ant.

Post.

Ant.

Post.

Ant.

Post.

Ant.

Post.

Ant.

Post.

Ant.

Post.

Ant.

Post.

84.29

80.41

79.72

89.61

74.36

81.13

61.87

53.71

70.61

61.66

64.16

60.49

57.60

68.50

HLC

0.212

0.158

0.227

0.148

0.238

0.170

0.341

0.242

0.316

0.233

0.304

0.228

0.321

0.218

Peak Load

374.06

375.80

360.80

413.40

343.33

389.31

291.18

313.95

305.33

339.64

297.56

327.64

263.02

340.07

Shear Load vs Displacement


400

Peak Load
300

200

Hysteresis Area

Load [N]

100

Slope Stiffness (N/mm)

-100

-200

HLC = Hysteresis Area/Total Loading Area


Peak Load

-300

-400
-4

-3

-1.5

-2

-1

-0.5

Displacement [mm]

Posterior Shear

Anterior Shear

Figure 2: An example shear loading cycle from which mechanical properties were derived, as shown.

A: C7/T1 Stiffness Values

100.00

Stiffness (N/mm)

100.00

B: C5/C6 Stiffness Values

80.00

80.00

Stiffness (N/mm)

120.00

60.00

No Lesion

60.00
40.00

N=2 N=2
N=1

N=2

N=2

Lesion 1
N=1
Lesion 2

No Lesion
Lesion 1

40.00
N=2N=2
N=2
N=2

N=2

N=2N=2

N=2

Lesion 3

20.00

20.00
0.00

Lesion 2

0.00
Anterior

Posterior

Anterior

C: C7/T1 HLC

Posterior

D: C5/C6 HLC
0.400

0.250

0.350
0.200
0.300
0.250

HLC

No Lesion

0.100

N=1
N=2 N=2

Lesion 1
N=2 N=2 N=1

Lesion 2

HLC

0.150

No Lesion

0.200
0.150

Lesion 1
N=2 N=2

N=2 N=1

Lesion 2
N=2 N=2 N=2 N=1

0.100

Lesion 3

0.050
0.050
0.000

0.000
Anterior
500.00

Posterior

Anterior

E: C7/T1 Peak Load

Posterior

F: C7/T1 Peak Load


450.00

450.00

400.00

400.00

350.00

350.00
300.00
No Lesion
No Lesion
250.00
200.00

N=2 N=2
N=1

N=2 N=1
N=2

150.00

Lesion 1
Lesion 2

Load (N)

Load (N)

300.00
250.00

Lesion 1
200.00
150.00

100.00

100.00

50.00

50.00

0.00

N=2 N=2 N=2

N=2 N=2 N=2 N=2

Lesion 2
Lesion 3

N=2

0.00
Anterior

Posterior

Anterior

Posterior

Figure 3: Bar graphs of the stiffness (A/B), hysteresis loss coefficient (C/D) and peak load (E/F) values obtained from shear loading of the C7/T1
and C5/C6 FSUs in the various degenerative states. Error bars indicate 1 standard deviation from the mean.

4. Discussion
The results from testing of the C7/T1 vertebrae
seem to suggest that the presence of an annular
lesion in the IVD will decrease stiffness and load
resistance in anterior shear. These findings appear
to be consistent with the findings in [3], which
observed a decrease in stiffness in flexion and
extension after a lesion was introduce to the IVD.
This was the only clear trend observed in the
initial analysis of the data; however, a full twoand three-way analysis of variance (ANOVA),
including factors such as vertebral level and disc
geometry, should be conducted (time-permitting)
to determine any significant predictors of the
properties observed for each degenerative
condition of the specimens.
The considerable variation in values obtained
between some consecutive tests of the same
specimen may be due to limitations of the testing
procedure.
These limitations
may
have
significantly altered the properties derived and
obscured any potential trends in the data. There are
a number of examples in the plots in Figure 3
where a single data point does not fit the trend of
the rest of the data, and it is hypothesised that
these trends may be more obvious if these
limitations were minimised.
The extremely small sample size used in this
project is a significant limitation. Only two
specimens and only two test repetitions for each
degenerative condition were tested. This sample
size was chosen due to time restrictions and
limited access to specimens. The specimen
preparation and potting procedures took
approximately 25 hours to complete, with an
additional 10+ hours of testing time and data
analysis. Thus, due to the limited amount of time
available to devote to this project, a small sample
size was required. It is suggested that testing of a
larger number of specimens may provide a better
representative of the mechanical properties of the
FSU at each degenerative condition.
The considerable spread of data may also be
due to limitations of the testing apparatus and
procedures. Specimens were fixed in the
mechanical testing machine using custom
aluminium cups which gripped the potted ends
using fasteners. Any slip or loosening of these
fasteners within the cups could significantly alter
the mechanical properties derived and, although
the fasteners were checked every few tests, this
would go unnoticed during testing.
Hydration and equilibrium conditions of the
specimen may also have varied between
specimens. Though efforts were made to maintain

hydration during preparation and testing, these


methods do not appropriately replicate hydration
of the IVD and entire FSU in-vivo. Current
literature recommends that specimens are
equilibrated prior to testing with an axial
compressive preload of 300N in a hypotonic bath
for 15h [8]; however, this was not possible due to
time restrictions. This may have resulted in
varying hydration and equilibrium levels of
specimens throughout testing which could
significantly alter the properties observed. Testing
could also have been conducted in a saline bath so
maintain the hydration and equilibrium levels
throughout testing.
A compressive preload of 0.2MPa was applied
throughout testing to replicate in-vivo loads
experienced by the IVD [7]. However, the
compressive load required to induce this stress was
calculated using the average geometry of the
specimens. This reduced the complexity of the
testing procedure but may also have compromised
its consistency. The area of each IVD was also
assumed to be consistent along their length, so
measurements were only taken at one location.
However, this assumption will have overestimated the cross-sectional area of the disc, and
thus underestimated the compressive stresses
experienced. More accurate and precise
measurements of disc geometry could have been
performed using image analysis from X-Ray or CT
images of each specimen. These could then have
been used to calculate the required load to induce a
0.2MPa compressive stress in each specimen.
The findings of this project were largely limited
by time and available resources; however, the
stiffness values obtained from anterior shear
loading of the C7/T1 vertebrae are consistent with
current literature.
References
1.
Thompson, RE, Pearcy, MJ, and Barker, TM, The
mechanical effects of intervertebral disc lesions. Clin
Biomech (Bristol, Avon), 2004. 19(5): p. 448-55.
2.
Hansson, T and Roos, B, The relation between bone
mineral content, experimental compression fractures, and
disc degeneration in lumbar vertebrae. Spine (Phila Pa
1976), 1981. 6(2): p. 147-53.
3.
Latham, J, Pearcy, M., Costi, J., Moore, R., Fraser,
R., Vernon-Roberts, B., Mechanical consequences of
annular tears and subsequent intervertebral disc
degeneration. Clinical Biomechanics, 1994. 9: p. 211-219.
4.
Fazzalari, NL, Costi, JJ, Hearn, TC, Fraser, RD,
Vernon-Roberts, B, Hutchinson, J, Manthey, BA,
Parkinson, IH, and Sinclair, C, Mechanical and pathologic
consequences of induced concentric anular tears in an
ovine model. Spine (Phila Pa 1976), 2001. 26(23): p. 257581.

5.
Wilke, HJ, Kettler, A, and Claes, LE, Are sheep
spines a valid biomechanical model for human spines?
Spine (Phila Pa 1976), 1997. 22(20): p. 2365-74.
6.
Howarth, SJ, Gallagher, KM, and Callaghan, JP,
Postural influence on the neutral zone of the porcine
cervical spine under anterior-posterior shear load. Med
Eng Phys, 2012.
7.
Korecki, CL, MacLean, JJ, and Iatridis, JC,
Characterization of an in vitro intervertebral disc organ
culture system. Eur Spine J, 2007. 16(7): p. 1029-37.
8.
Stokes, IA, Laible, JP, Gardner-Morse, MG, Costi,
JJ, and Iatridis, JC, Refinement of elastic, poroelastic, and
osmotic tissue properties of intervertebral disks to analyze
behavior in compression. Ann Biomed Eng, 2011. 39(1): p.
122-31.

Vous aimerez peut-être aussi