Biomechanics Of Knee Ligaments.pdf

(694 KB) Pobierz
doi:10.1016/j.jbiomech.2004.10.025
ARTICLE IN PRESS
Journal of Biomechanics ] (]]]]) ]]]–]]]
Review
Biomechanics of knee ligaments: injury, healing, and repair
Savio L.-Y. Woo , Steven D. Abramowitch, Robert Kilger, Rui Liang
Department of Bioengineering, Musculoskeletal Research Center, University of Pittsburgh, Pittsburgh, PA, 15219, USA
Accepted 20 October 2004
Abstract
Knee ligament injuries are common, particularly in sports and sports related activities. Rupture of these ligaments upsets the
balance between knee mobility and stability, resulting in abnormal knee kinematics and damage to other tissues in and around the
joint that lead to morbidity and pain. During the past three decades, significant advances have been made in characterizing the
biomechanical and biochemical properties of knee ligaments as an individual component as well as their contribution to joint
function. Further, significant knowledge on the healing process and replacement of ligaments after rupture have helped to evaluate
the effectiveness of various treatment procedures.
This review paper provides an overview of the current biological and biomechanical knowledge on normal knee ligaments, as well
as ligament healing and reconstruction following injury. Further, it deals with new and exciting functional tissue engineering
approaches (ex. growth factors, gene transfer and gene therapy, cell therapy, mechanical factors, and the use of scaffolding
materials) aimed at improving the healing of ligaments as well as the interface between a replacement graft and bone. In addition, it
explores the anatomical, biological and functional perspectives of current reconstruction procedures. Through the utilization of
robotics technology and computational modeling, there is a better understanding of the kinematics of the knee and the in situ forces
in knee ligaments and replacement grafts.
The research summarized here is multidisciplinary and cutting edge that will ultimately help improve the treatment of ligament
injuries. The material presented should serve as an inspiration to future investigators.
r 2004 Elsevier Ltd. All rights reserved.
Keywords: Biomechanics; Knee ligaments; Tissue engineering; Healing
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Anatomy, histological appearance and biochemical constituents of normal ligaments. . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3. Tensile properties of ligaments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1. Ligament anisotropy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.2. Significant biological factors on the properties of ligaments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4. Viscoelastic properties of ligaments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.1. The quasi-linear viscoelastic theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4.2. Continuum based viscoelastic models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Corresponding author. Department of Bioengineering, Musculoskeletal Research Center, 405 Center for Bioengineering, 300 Technology Drive,
P.O. Box 71199, Pittsburgh, PA 15219, USA. Tel.: +14126482000; Fax: +14126882001.
E-mail addresses: ddecenzo@pitt.edu, slyw@pitt.edu (S.L.-Y. Woo).
0021-9290/$-see front matter r 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jbiomech.2004.10.025
60963401.022.png
2
ARTICLE IN PRESS
S.L.-Y. Woo et al. / Journal of Biomechanics ] (]]]]) ]]]–]]]
5. Healing of knee ligaments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.1. MCLhealing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.2. Phases of ligament healing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.3. New animal model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
6. New approaches to improve healing of ligaments—functional tissue engineering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
6.1. Growth factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
6.2. Gene transfer and gene therapy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
6.3. Cell therapy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
6.4. Biological scaffolds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
6.5. Mechanical factors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
7. ACLreconstruction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
7.1. Graft function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
7.2. Graft incorporation and remodeling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
8. Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Acknowledgements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1. Introduction
majority of ligament reconstructions yield good short-
term clinical results, 20–25% of patients experience
complications including instability that could progres-
sively damage other knee structures ( Aglietti et al., 1997 ;
Bach et al., 1998 ; Daniel et al., 1994 ; Jomha et al., 1999 ;
Ritchie and Parker, 1996 ; Shelbourne et al., 1995 ; Yagi
et al., 2002 ).
Thus, there has been a tremendous quest for knowl-
edge to better understand ligament injuries, healing and
remodeling in hope to develop new and improved
treatment strategies. The needs in meeting this goal
have stimulated researchers to seek new and innovative
methods of investigation. Because of the complex
biological process, it has become clear that collabora-
tions from different disciplines rather than an indivi-
dualistic approach in research must be developed. In this
review, the properties of normal ligaments, including
their anatomical, biological, biochemical and mechan-
ical properties, as well as the changes that occur
following injury will be described. The MCLwill be
used as a model because of its uniform cross-sectional
area, large aspect ratio, and propensity for healing.
Subsequently, novel functional tissue engineering meth-
odologies and some of the early findings will be
presented. The challenging problems which remain to
be solved and the potential of new treatment strategies
will be explored. In terms of ligament reconstruction, the
biomechanics of surgical reconstruction of the ACLand
the utilization of robotics technology to study some of
the key surgical parameters that affect the performance
of the replacement grafts will be reviewed. It is hoped
that these creative research approaches will inspire many
to join this course of investigation and ultimately help
improve the treatment of ligament injuries.
Injuries to knee ligaments are very common. It has
been estimated that the incidence could be at 2/1000
people per year in the general population ( Miyasaka et
al., 1991 ) and a much higher rate for those involved in
sports activities ( Bruesch and Holzach, 1993 ). Ninety
percent of knee ligament injuries involve the anterior
cruciate ligament (ACL) and the medial collateral
ligament (MCL) ( Miyasaka et al., 1991 ). In fact, recent
studies have documented that ACLinjuries in females
are reaching epidemic proportions with the frequency of
rupture more than 3 times greater than that of their male
counterparts ( Anderson et al., 2001 ; Arendt and Dick,
1995 ; Powell and Barber-Foss, 2000 ). The results of
ligament injuries can be devastating. Frequently, surgery
is required, but the outcomes are variable. Further,
post-surgical rehabilitation could require an extended
absence from work or athletic competition.
Basic science and clinical studies have revealed that a
ruptured MCLcan heal spontaneously ( Frank et al.,
1983 ; Indelicato, 1983 ; Jokl et al., 1984 ; Kannus, 1988 ).
However, laboratory studies have shown that its
ultrastructure and biochemical composition remain
significantly altered ( Frank et al., 1983 ; Niyibizi et al.,
2000 ; Weiss et al., 1991 ). Furthermore, the mechanical
properties of the ligament substance remain substan-
tially inferior to those of normal ligaments even after
years of remodeling ( Loitz-Ramage et al., 1997 ; Ohland
et al., 1991 ). On the other hand, midsubstance tears of
the ACLand posterior cruciate ligament (PCL) would
not heal spontaneously and surgical reconstruction
using a replacement graft is often required ( Hirshman
et al., 1990 ; Kannus and Jarvinen, 1987 ). While the
 
ARTICLE IN PRESS
S.L.-Y. Woo et al. / Journal of Biomechanics ] (]]]]) ]]]–]]]
3
2. Anatomy, histological appearance and biochemical
constituents of normal ligaments
bone
Ligaments are composed of closely packed collagen
fiber bundles oriented in a parallel fashion to provide for
stability of joints in the musculoskeletal system. The
major cell type is the fibroblast and they are interspersed
in the parallel bundles of collagen.
In the human knee, the MCLis approximately 80mm
long and runs from the medial femoral epicondyle
distally and anteriorly to the posteromedial margin of
the metaphysis of the tibia. The lateral collateral
ligament (LCL) originates from the lateral femoral
epicondyle and passes postero-distally to the top of the
fibular head. The cruciate ligaments, which are named
anterior and posterior according to their site of
attachment to the tibia, are located within the capsule
and cross each other obliquely. The anterior cruciate
ligament (ACL) arises from the anterior part of the
intercondylar eminentia of the tibia and extends to
the posterolateral aspect of the intercondylar fossa of
the femur. The posterior cruciate ligament (PCL) arises
from the posterior part of the intercondylar eminentia of
the tibia and passes to the anterolateral aspect of the
intercondylar fossa of the femur. Although morpholo-
gically intraarticular, the cruciate ligaments are sur-
rounded by a synovial layer. The ACLconsists of two
bundles, an anteromedial (AM) and a posterolateral
(PL) bundle. The AM bundle is thought to be important
as a restraint to anterior–posterior translation of the
knee, while the PLbundle is thought to be an important
restraint to rotational moments about the knee ( Yagi
et al., 2002 ). This anatomic division of these bundles is
based on the gross tensioning pattern of the ACLduring
passive flexion-extension of the knee, with the AM
bundle being tauter in flexion and the PLbundle tauter
in extension. The PCLis also composed of two distinct
bundles, the antero-lateral (AL) and the postero-medial
(PM) bundle. Additionally, ligaments are sometimes
found anterior and posterior to the PCLin some people.
They are the anterior meniscofemoral ligament (MFL;
i.e. ligamentum Humphrey) and the posterior menisco-
femoral ligament (i.e. ligamentum Wrisberg) ( Girgis
et al., 1975 ).
Generally, ligaments are inserted to bone in two ways;
direct and indirect ( Fig. 1 ). For direct insertions (e.g. the
femoral insertion of MCL), fibers attach directly into
the bone and the transition of ligament to bone occurs in
four zones: ligament, fibrocartilage, mineralized fibro-
cartilage and bone ( Woo et al., 1987 ). For an indirect
insertion (e.g. the tibial insertion of MCL) superficial
fibers are attached to periosteum while the deeper fibers
are directly attached to the bone at acute angles ( Woo
et al., 1987 ). The tibial insertion of the MCLcrosses the
epiphyseal plate so that it can be lengthened in
synchrony with the bone growth.
mineralized
fibrocartilage
fibrocartilage
ligament
(A)
deep fibers
bone
superficial
fibers
connects to
periosteum
(B)
Fig. 1. (A) Photomicrograph demonstrating direct insertion, i.e. the
femoral insertion of rabbit medial collateral ligament (MCL). (B)
Photomicrograph demonstrating indirect insertion, i.e. the tibial
insertion of rabbit MCL. (Hematoxylin and eosin, x50) (permission
requested from ( Woo et al., 1987 )).
Between 65 and 70% of a ligament’s total weight is
composed of water. On a fat-free basis, Type I collagen
is the major constituent (70–80% dry weight) and is
primarily responsible for a ligament’s tensile strength.
Type III collagen (8% dry weight) and Type V collagen
(12% dry weight) are other major components ( Birk and
Mayne, 1997 ; Linsenmayer et al., 1993 ). Collagen Types
II, IX, X, XI, and XII have also been found to be
present ( Fukuta et al., 1998 ; Niyibizi et al., 1996 ;
Sagarriga Visconti et al., 1996 ).
Variations in the concentrations of these basic
constituents lead to a diverse array of mechanical
behaviors of knee ligaments that are suitable for their
respective functions. A comparative study showed that
the tangent modulus and tensile strength of the rabbit
MCLis higher than the ACL( Woo et al., 1992 ) which
correlates with a larger mean fibril diameter for the
MCL( Hart et al., 1992 ). In addition, the fibroblasts of
the MCLare more spindle shaped ( Lyon et al., 1991 )
and produce a higher level of procollagen type I mRNA
( Wiig et al., 1991 ) and a lower collagen type III to type I
ratio in culture ( Ross et al., 1990 ). Further, mechanical
loading has been found to regulate the gene expression
60963401.023.png 60963401.024.png 60963401.001.png
ARTICLE IN PRESS
4
S.L.-Y. Woo et al. / Journal of Biomechanics ] (]]]]) ]]]–]]]
of collagens in ligaments ( Hsieh et al., 2002 ). Therefore,
each ligament’s composition is directly correlated with
its mechanical properties.
the human ACLwas larger than that for the PLbundle
( Butler et al., 1992 ). In a separate study, the mechanical
properties of the bundles of the human PCLwere found
to be different as well ( Harner et al., 1995 ). The tangent
modulus of the ALbundle (294
115MPa) was almost
twice that of the PM bundle (150
7
3. Tensile properties of ligaments
69MPa). The fact
that different bundles have different properties suggests
that each bundle contributes to knee joint stability
differently, which may have important ramifications on
their replacements ( Table 1 ).
7
Ligaments are best suited to transfer load from bone
to bone along the longitudinal direction of the ligament.
Thus, their properties are commonly studied via a
uniaxial tensile test of a bone–ligament–bone complex
(e.g. femur-MCL-tibia complex). These tests result in a
load–elongation curve that is non-linear and concave
upward. This enables ligaments to help to maintain
smooth movement of joints under normal, physiologic
circumstances and to restrain excessive joint displace-
ments under high loads. The parameters describing the
structural properties of the bone–ligament–bone com-
plex include stiffness, ultimate load, ultimate elongation,
and energy absorbed at failure. With cross-sectional
area and strain measurements, a stress–strain curve
representing the mechanical properties (quality) of the
ligamentous tissue can be obtained. The parameters
describing the mechanical properties of the ligament
substance include tangent modulus, ultimate tensile
strength, ultimate strain, and strain energy density. A
large number of experimental methods have been
employed by investigators to overcome some of the
technical difficulties encountered in measuring the
mechanical properties of ligaments ( Beynnon et al.,
1992 ; Ellis, 1969 ; Lam et al., 1992 ; Lee and Woo, 1988 ;
Peterson et al., 1987 ; Peterson and Woo, 1986 ; Smutz et
al., 1996 ). Furthermore, environmental factors can also
cause large differences in the experimental data obtained
( Crowninshield and Pope, 1976 ; Figgie et al., 1986 ;
Haut, 1983 ; Haut and Powlison, 1990 ; Noyes et al.,
1974 ). For more information on these methodologies
and environmental factors, the readers are encouraged
to read the provided references and study the chapter
entitled: Biology, Healing and Repair of Ligaments in
Biology and Biomechanics of the Traumatized Synovial
Joint: The Knee as a Model, 1992 by the authors ( Woo
et al., 1992 ).
An equally important consideration is the geometry of
the ligament. Unlike the MCLwhose cross-section is
relatively uniform over its length, the ACLand PCL
have two functionally distinct bundles that are loaded
non-uniformly ( Fuss, 1989 ; Girgis et al., 1975 ; Sakane et
al., 1997 ). Thus, they need to be separated in order to
have a specimen with a more uniform cross-sectional
area for tensile testing. Using this approach, a study
performed at our center showed the tangent modulus of
a section of the rabbit ACL(516
3.1. Ligament anisotropy
Ligaments are three dimensional (3-D) anisotropic
structures. To describe the 3-D mechanical behavior of
the human MCL, investigators have developed a quasi-
static hyperelastic strain energy model based on the
assumption of transverse isotropy ( Quapp and Weiss,
1998 ). The total strain energy, W, in response to a
stretch along the collagen fiber direction, l; was defined
to be equal to the sum of the strain energy resulting from
ground substance (F 1 ), collagen fibers (F 2 ), and an
interaction component (F 3 ),
W ðI 1 ; I 2 ; lÞ¼F 1 ðI 1 ; I 2 ÞþF 2 ðlÞþF 3 ðI 1 ; I 2 ; lÞ (1)
where I 1 and I 2 are invariants of the right Cauchy stretch
tensor. For a uniaxial tensile test, F 1 was described with
a two coefficient Mooney–Rivlin material model
F 1 ¼ 1=2½C 1 ðI 1 3ÞþC 2 ðI 2 3Þ; (2)
where C 1 and C 2 are constants, and F 2 was described by
separate exponential and linear functions. F 3 was
assumed to be zero.
The Cauchy stress, T, can then be written as
T ¼ 2fðW 1 þ I 1 W 2 ÞB W 2 B 2 gþlW l aaþ r1; (3)
where, B is the left deformation tensor, and W 1 , W 2 ,and
W l are the partial derivatives of strain energy with
respect to I 1 , I 2 , and l; respectively. The unit vector field,
a, represents the fiber direction in the deformed state,
and r is the hydrostatic pressure required to enforce
incompressibility.
It was found that this constitutive model can fit both
the data obtained from longitudinal and transverse
dumbbell shaped specimens cut from the human MCL
Table 1
Values for tangent modulus of the human MCL( Quapp and Weiss,
1998 ), AM and PLbundles of the human ACL( Butler et al., 1992 ),
and ALand PM bundles of the PCL( Harner et al., 1995 ).
Tangent modulus (MPa)
64MPa) was less than
half of that for the rabbit MCL(1120
7
Human MCLHuman ACL
Human PCL
153MPa) ( Woo
et al., 1992 ). Further, the tangent modulus, tensile
strength, and strain energy density of the AM bundle in
7
AM
PL
AL
PM
332
7
58
283
7
114 154
7
120 294
7
115 150
7
69
60963401.002.png
ARTICLE IN PRESS
S.L.-Y. Woo et al. / Journal of Biomechanics ] (]]]]) ]]]–]]]
5
Transverse
Decrease
Stress
Increase
Stress
40
Longitudinal
30
Immobilization
Normal
Activity
Exercise
20
10
0
0
4
8
12
16
Strain (%)
In-vivo Loads and Activity Levels
Fig. 2. Stress–strain curves for human MCLs longitudinal and
transverse to the collagen fiber direction (permission requested from
Quapp and Weiss (1998) ).
Fig. 3. A schematic diagram describing the homeostatic responses of
ligaments and tendons in response to different levels of stress and
motion (permission requested from ( Woo et al., 1987 )).
( Fig. 2 ). The longitudinal specimens displayed a tangent
modulus of 332.2
7
58.3MPa and a tensile strength of
Based on the results of these and other related studies,
a highly non-linear representation of the relationship
between different levels of stress and ligament properties
is depicted in Fig. 3 . The normal range of physiological
activities is represented by the middle of the curve.
Immobilization results in a rapid reduction in tissue
properties and mass. In contrast, long term exercise
resulted in a slight increase in mechanical properties as
compared with those observed in normal physiological
activities.
Skeletal maturity also causes significant changes
to ligaments whereby the stiffness and ultimate load
of the FMTC was shown to increased dramatically
from 6 to 12 months of age followed by insignificant
change from 1 to 4 years in the rabbit model ( Woo et al.,
1990 ). This corresponded with a change in failure
mode from the tibial insertion to the midsubstance
reflecting closure of the tibial epiphysis during matu-
ration ( Woo et al., 1986 ). On the other hand, the
human FATC demonstrated a significant decrease in
the stiffness and ultimate load with increasing age
( Noyes and Grood, 1976 ; Woo et al., 1991 ). There-
fore, each ligament is unique in its growth, development,
and aging. Investigators should be cautious when
extrapolating age related changes from one ligament
(ex. ACLto PCL) or species (ex. rabbit to human) to
another.
4.8MPa, while the transverse specimens were an
order of magnitude lower with a tangent modulus of
11.0
7
7
0.9MPa and tensile strength of 1.7
7
0.5MPa
( Quapp and Weiss, 1998 ).
3.2. Significant biological factors on the properties of
ligaments
The effects of immobilization and exercise on the
mechanical properties of ligaments has been investigated
by a number of laboratories ( Larsen et al., 1987 ;
Newton et al., 1990 ; Noyes, 1977 ; Woo et al., 1987 ).
When rabbit hind limbs were subjected to a few weeks of
immobilization, there were marked decreases in the
structural properties of the femur–MCL–tibia complex
(FMTC). These decreases occurred due to subperiosteal
bone resorption within the insertion sites, as well as
microstructural changes in the ligament substance.
Remobilization was found to reverse these negative
changes. However, up to one year of remobilization was
required for the properties of the ligament to return to
normal levels following 9 weeks of immobilization ( Woo
et al., 1987 ). Similar results were found for the
femur–ACL–tibia complex (FATC) of primates and
rabbits ( Newton et al., 1990 ; Noyes, 1977 ). Long periods
of exercise training, on the other hand, only showed
marginal increases in the structural properties of
ligaments with a 14% increase in linear stiffness of the
FMTC and a 38% increase in ultimate load/body weight
( Laros et al., 1971 ; Woo et al., 1982, 1979 ). There was
only a slight change in the mechanical properties of the
ligament substance.
4. Viscoelastic properties of ligaments
The complex interactions of collagen with elastin,
proteoglycans, ground substance, and water results in
the time- and history-dependent viscoelastic behaviors
of ligaments. In response to various tensile loading
38.6
60963401.003.png 60963401.004.png 60963401.005.png 60963401.006.png 60963401.007.png 60963401.008.png 60963401.009.png 60963401.010.png 60963401.011.png 60963401.012.png 60963401.013.png 60963401.014.png 60963401.015.png 60963401.016.png 60963401.017.png 60963401.018.png 60963401.019.png 60963401.020.png 60963401.021.png
Zgłoś jeśli naruszono regulamin