You are on page 1of 14

Annals of Biomedical Engineering ( 2018)

https://doi.org/10.1007/s10439-018-2078-7

Trunk Hybrid Passive–Active Musculoskeletal Modeling to Determine


the Detailed T12–S1 Response Under In Vivo Loads
P. KHODDAM-KHORASANI,1 N. ARJMAND ,1 and A. SHIRAZI-ADL2
1
Department of Mechanical Engineering, Sharif University of Technology, Tehran 11155-9567, Iran; and 2Division of Applied
Mechanics, Department of Mechanical Engineering, École Polytechnique, Montreal, QC, Canada
(Received 9 December 2017; accepted 18 June 2018)

Associate Editor Thurmon E. Lockhart oversaw the review of this article.

Abstract—Biomechanical models of the spine either simplify spine loads are invasive, complex, costly, and limited
intervertebral joints (using spherical joints or deformable with low number of volunteers. For instance, loads
beams) in musculoskeletal (MS) or overlook musculature in transmitted through instrumented vertebral cages have
geometrically-detailed passive finite element (FE) models.
These distinct active and passive models therefore fail to been measured in patients.54,56 Intradiscal pressure
determine in vivo stresses and strains within and load-sharing (IDP) at lumbar34,35,42,62,75,76 and thoracic50,61 inter-
among the joint structures (discs, ligaments, and facets). A vertebral discs has also been measured as an indirect
novel hybrid active–passive spine model is therefore devel- estimate of the spine compressive loads. Such pressure
oped in which estimated trunk muscle forces from a MS measurements remain, however, sensitive to pressure
model for in vivo activities drive a mechanically-equivalent
passive FE model to quantify in vivo T12–S1 compression/ transducer technology (e.g., liquid-filled or piezo-re-
shear loads, intradiscal pressures (IDP), centers of rotation sistive transducers),21 needle/transducer size,10 and its
(CoR), ligament/facet forces, and annulus fiber strains. The self-heating.32,33 On the other hand, there exists no
predicted and in vivo L4–L5 IDP and L1–S1 CoRs showed direct in vivo measurement of forces in ligaments,
satisfactory agreements. The FE model under commonly- contact facet joints, and disc collagen fibers. Similarly,
used in vitro loading (pure moments and follower loads)
predicted different kinetics from those of the hybrid model muscle forces can only be indirectly estimated via
under in vivo loads (muscle exertions and gravity loads) electromyography (EMG); an approach that suffers
contributing to suggest the inadequacy of such in vitro loads from several simplifications/assumptions in the collec-
when simulating in vivo tasks. For an improved assessment of tion and post-processing of EMG signals and the
the injury risk, evaluation of the internal loads, and design of muscle EMG–force relationship.21
implants, such hybrid models should therefore be used.
In the absence of satisfactory non-invasive mea-
surement techniques, computational biomechanical
Keywords—Spine, Finite element, Musculoskeletal, Model,
models have emerged to evaluate internal loads within
Disc fiber strain, Center of rotation, Intradiscal pressure,
the passive–active structures of the spine. These mod-
Muscles, Facets, Ligaments.
els, however, often tend to focus on and more accu-
rately represent either active or passive structures.
Musculoskeletal (MS) models, that are primary
INTRODUCTION
developed to compute forces in trunk muscles, incor-
Quantification of load sharing among different porate often a rather detailed musculature but a sim-
passive (discs, ligaments and facets) and active (mus- plified passive (ligamentous) spine.3,4,6,16,17,25,41,70
cles) structures of the spine during in vivo regular, sport Intervertebral joints are regularly simulated as (fric-
and workplace activities is essential for the improved tional or frictionless) pivots16,17,25 or deformable
development of injury prevention, treatment strategies beams3,4,6 (with no facets or ligaments) that allow for
and implants of the spine. In vivo measurements of the the estimation of the overall joint loads but fail to
evaluate internal stresses and strains within and load
sharing between the joint structures. Such MS models
Address correspondence to N. Arjmand, Department of Mechanical do neither allow for the direct estimation of IDP nor
Engineering, Sharif University of Technology, Tehran 11155-9567,
Iran. Electronic mail: arjmand@sharif.edu

 2018 Biomedical Engineering Society


KHODDAM-KHORASANI et al.

do they account for the load-dependant motion of the IDP, ligament forces, facet joint forces (FJFs) and
centers of rotation (CoRs) at different levels. CoRs. For validation, predictions of the hybrid model
On the other hand, finite element (FE) models of the for L4–L5 IDP and L1–S1 CoRs are compared with
passive spine incorporate accurate detailed geometries available in vivo data. It is hypothesized that (1) a MS
and nonlinear mechanical properties of the joint model (with motion segments as deformable beams)
components (i.e., ligaments, facet articulations, disc with nonlinear passive resistance and kinematics
annulus and nucleus, vertebrae, and endplates). identical to those in the detailed FE model can itera-
Therefore, these FE models allow for the prediction of tively be developed, and (2) in vitro loading (pure
stresses and strains within and load sharing among moment plus FL) in passive FE models cannot gen-
these structures under given load or displacement erate accurate in vivo load-sharing among different
conditions.22 Excluding muscles, these models are of- joint structures.
ten used as complementary tools along with in vitro
measurements. Consequently, the important role of
muscle exertions is either fully overlooked or simplified MATERIALS AND METHODS
by a single follower load (FL)37,55,65,67; a virtual
T12–S1 Passive FE Model
compression force with varying line of action that
follows the lumbar lordosis curve while passing This model is an updated version of a previously-
through the vertebral body centroids. As a pure com- developed L1–S1 lumbar spine model12,64 based on CT
pressive force, this FL fails to mimic the shear com- images of the lumbar spine of a cadaver (65 years-old
ponents of trunk muscle forces and gravity/inertia male). Six vertebrae (L1 through S1 as rigid bodies),
loads.26 Moreover, a single constant FL cannot rep- five discs (annulus and nucleus), 10 contact facet sur-
resent the varying compressive force, due to distributed faces, and nine sets of ligaments (anterior longitudinal
gravity as well as the complex muscle recruitments, (ALL), posterior longitudinal (PLL), capsular (CL),
along the spinal height.67 These passive models are intertransverse (ITL), ligamentum flavum (LF), su-
therefore limited when simulating in vivo activities. praspinous (SSL), interspinous (ISL), fascia (L4 and
To circumvent these shortcomings, a coupled hybrid L5 to ilium) and iliolumbar (IL, L5 to ilium) liga-
model simulating in sufficient details both active ments) were included. Annulus layers and disc nucle-
(muscles) and passive components is required. Due to uses were considered as, respectively, isotropic annulus
the computational effort necessary in such coupled bulk (reinforced by membranes of collagen fibres) and
models, we recently proposed a staggered solution in fluid-filled cavities. For the current study, this model
which gravity loads and muscle forces computed by a was updated (Figs. 1a and 1b) by the (1) addition of
validated MS model during in vivo activities were the T12–L1 motion segment (assuming a segmental
prescribed into a passive L4–L5 FE model.9 This study lordosis similar to that of the L1–L2 motion segment
demonstrated the feasibility of combining two distinct and a mean disc height identical to that of the T12–L1
(MS and FE) models to simulate in vivo activities. The in the MS model), (2) substantial refinement of the
passive model however was limited only to one lumbar mesh (~ 1730 solid brick elements for each disc annu-
segment (L4–L5). Moreover, due to different passive lus) to improve model predictions, (3) use of hypere-
joint properties used in the MS and passive FE models, lastic Mooney–Rivlin material properties (c1 = 0.42,
dissimilar kinematics between the two models were c2 = 0.10515) for the ground substance of the annulus
computed under identical tasks. In addition, the MS bulk, (4) representation of 14 distinct collagen fiber
model, unlike the passive FE model, neglected the layers using membrane layers reinforced by rebar ele-
passive disc stiffening under increasing compressive ments (with criss-cross patterns at ± 3031 and non-
loads.44,65 linear stress–strain data for material properties66,68)
The present study therefore aims to perform a distributed in the foregoing annulus ground substance,
combined MS-FE model study of the whole T12–S1 and (5) simulation of facet articulations by frictionless
spine in which the in vivo active–passive load sharing, surface-to-surface contacts with average gap limit of
internal stresses and strains, and vertebral CoRs are 1.25 mm (simulating compliant articular cartilage
determined in an iterative procedure that guarantees layers). Similar to the original model, bodies and
identical spine kinematics and load-dependent nonlin- posterior bony elements of the vertebrae were distinct
ear passive properties in both models. Trunk gravity rigid elements attached together via two deformable
loads and predicted muscle forces by our MS beams at pedicles.64 Ligaments were modeled by mul-
model3,4,6 during in vivo activities are iteratively used tiple axial tension-only elements with nonlinear prop-
to drive a mechanically-equivalent passive FE model of erties.66 Each nucleus was an incompressible fluid-filled
the T12–S1 that computes disc stresses and strains, cavity.43
Trunk Hybrid Passive–Active Model for T12–S1 Responses

FIGURE 1. The T12–S1 passive finite element (FE) model: (a) 3D mesh generation and (b) detailed view of the L4–L5 disc
including its ground substance, nucleus cavity and 14 circumferential rebar layers of disc annulus. The musculoskeletal (MS)
model and its musculature in (c) sagittal and (d) frontal planes that include 46 local (originating from the pelvis and inserting into
lumbar vertebrae) and 10 global (originating from the pelvis and inserting into the thoracic spine) fascicles of muscles (ICPL:
iliocostalis lumborum pars lumborum, ICPT: iliocostalis lumborum pars thoracic; IP, iliopsoas; LGPL, longissimus thoracis pars
lumborum; LGPT, longissimus thoracis; MF, multifidus; QL, quadratus lumborum; IO, internal oblique; EO, external oblique; RA,
rectus abdominus). (e) Sagittal and (f) frontal view of the FE model with the axial connector elements used to apply muscle forces.

MS Model incorporated (Figs. 1c and 1d).71 Curved lines of ac-


tion of back muscles (i.e., their wrapping around and
Our previously-developed MS model3,4,6,8 evaluated
contact forces at bony vertebrae) were considered.8
forces in trunk muscles during eight in vivo activities
Based on our earlier in vivo measurements for each
using an optimization algorithm described below. The
task,3,4 the rotation of thorax (T), pelvis (P), and
geometry of the spine (dimensions and lumbar lordo-
lumbar (L = T 2 P; partitioned between T12 and L5
sis) in the MS and FE models was identical. The pelvis,
vertebrae6) as well as gravity/hand loads were pre-
T1–T12 thorax and lumbar vertebrae were rigid
scribed into the model. To determine muscle forces, an
(Figs. 1c and 1d). T12–L1 through L5–S1 segments
optimization algorithm minimizing sum of cubed
were simulated by 3-node nonlinear shear-deformable
muscle stresses was used. Overall joint loads (local
beams located 4 mm posterior to the disc centers to
compression and shear loads at the beams) were sub-
account for the shift in the disc CoR observed in the
sequently evaluated.
passive FE model under different loads.66 The trunk
weight of ~ 344 N (~ 52%51 of the total body weight of
the male subject in our earlier studies with a body mass
Distinct Validation of the MS and FE Models
of ~ 68 kg) was partitioned among upper arms
(~ 36 N), forearms/hands (~ 29 N), head (46 N) (ap- The MS model was previously validated by com-
plied via rigid elements at their centers of mass on the paring its predicted muscle forces and compressive
rigid thorax recorded in our in vivo measurements),3,4 loads vs. measured muscle EMG and IDPs, respec-
and T1–L5 segments (~ 233 N between the T1 through tively.3,4,6,52 As to the passive FE model, range of
L5 segments and applied via rigid elements at their motions (RoMs), IDPs, and FJFs predicted under
centers of mass located off center at each vertebra. single/combined moments and FL38 (in vitro loading,
Segmental gravity forces and their corresponding lever Table 1) were compared with in vitro and in silico data.
arms in the sagittal plane are provided in supplemen- The L5 vertebra was fixed in these simulations.
tary data).5,49 For tasks with a hand load, a force in the Moreover, a loading condition suggested elsewhere to
gravity direction was also applied via a rigid element simulate forward flexion (i.e., 7.5 Nm flexion plus a FL
attached to the rigid thorax at the center of mass of the of 1175 N)22,60 was also considered in the FE model
hand load measured in vivo. In total, 56 muscles were and its predictions were compared with those of the
KHODDAM-KHORASANI et al.

TABLE 1. Combined moment and follower load (FL) analysis of MS model was repeated after adjustments
conditions considered for the passive FE model.
in its passive, compression-dependent stiffness (mo-
Loading FL (N) Moment (Nm) References ment-rotation) properties based on those in the passive
FE model under gravity loads, weight held in hands
Flexion 1175 7.5 Rohlmann et al.60 and muscle exertions. The updated muscle forces were
Extension 500 7.5 Rohlmann et al.60
Lateral bending 700 7.8 Dreischarf et al.20
again prescribed into the passive FE model and this
Axial rotation 720 5.5 Dreischarf et al.19 iterative procedure was continued until convergence,
i.e., the two models had almost similar kinematics for
the simulated task (average difference of < 1 mm
hybrid model under in vivo loading (muscle exertions). between the MS and FE models for the final positions
For this latter simulation, the S1 translational degrees- of the T12–S1 centroids). In all simulations, the S1
of-freedom were constrained. translational degrees-of-freedom were constrained,
pelvis rotation was prescribed based on the measured
value in vivo while remaining rotational and transla-
Simulated Tasks tional degrees-of-freedom at the vertebral levels were
Eight symmetric static tasks including the neutral left unconstrained (i.e., predicted by the hybrid model).
upright standing posture, forward flexion of 20, 40, A flowchart representing the method was provided in
60, 80 and 107 (full flexion) with arms in the gravity supplementary data.
direction, and upright two-handed lifting of 19.8 kg at
25 cm (close) and 55 cm (away) anterior distances to Model Outputs
the L5–S1 disc were considered.3,4,24 These tasks were
chosen due to the availability of the measured L4–L5 T12–S1 IDPs, CoR of each vertebra with respect to
IDP76 for subsequent comparisons and the required its lower vertebra,37,46 ligament forces, FJFs, and
kinematics3,4 to drive simulations. maximal disc fiber strains were computed for each
in vivo activity. Moreover, based on the free body
diagram at the disc mid-height, ligaments, and facet
Hybrid Model joints at each level of the passive FE model, and con-
For each task, the associated pelvic rotation mea- sideration of the force/moment equilibrium require-
sured in vivo and muscle forces estimated in the MS ments at the disc centroid, the resultant spinal loads
model (via connector elements in between insertion (local compression, shear, and passive moment) as well
points as in Figs. 1e and 1f) at the deformed configu- as force/moment sharing among discs, ligaments and
rations were prescribed into the passive FE model. facet joint were determined37 for each activity. For this
Each uniaxial connector element ran between the up- purpose, force-moment vectors in ligaments and con-
per insertion point of a muscle to its lower insertion tacting facet joints were directly calculated from nodal
point. These insertion points were either located on forces in the FE model while those at the disc were
bony vertebrae in the FE model or attached via rigid computed by solving the force equilibrium equations in
elements to the bony structures (Figs. 1e and 1f). The axial and shear directions as well as the moment
force–elongation property of each connector element equilibrium equations in the sagittal plane.
was defined so that it generated a constant axial force
(equal to the muscle force estimated from the MS
Validations of the Hybrid Model
model) regardless of the elongation in the connector
element. As for the curved (wrapping) global thoracic The predicted L4–L5 IDP in each task and L1–S1
muscles (LGPT and ICPT), multi-segment connectors CoRs in flexion tasks were compared, respectively,
with a unique force in between insertions and wrapping with the corresponding in vivo IDP in one healthy male
points on vertebrae in contact were considered. In subject76 and in vivo image-based (upright-to-flexion)
addition, gravity loads and weight in hands were CoR in ten male subjects.46 The body weight (~ 68 kg)
applied at their locations measured in vivo. Despite considered in the hybrid model and that of the subject
identical loading conditions, different deformed con- (70 kg) who participated in the IDP measurements76
figurations (between MS and passive FE models) were were similar thus making such comparisons more rel-
computed due to their distinct simulations (e.g., com- evant (body weight has a great impact on spinal
pression-dependent stiffness) of motion segments. To loads2,24,25). Moreover, L4–L5 IDP was also estimated
assure identical final deformed T12–S1 kinematics in the MS model using an equation that relates IDP to
between the MS and FE models (the T1-T12 segment is axial compression force and intersegmental flexion
considered as a single rigid body in both models), the rotation.24
Trunk Hybrid Passive–Active Model for T12–S1 Responses

RESULTS

Validations of the Passive FE Model


RoM of the L4–L5 disc-vertebra unit model alone
(without ligaments and facets) loaded by a 10 Nm pure
moment in different planes fell within the in vitro range
for the human disc specimens27 (Fig. 2a). The L1-L5
intact model (with ligaments and facets) under a 7.5
Nm moment in different directions, yielded RoMs in
agreement with in vitro57 and in silico22 data (Fig. 2b).
The model loaded with a FL up to 1000 N predicted
L4–L5 IDPs within the in vitro13 and other FE model22
ranges (Fig. 2c). In combined FL and moment in dif-
ferent anatomical planes,58–60 the model also predicted
intervertebral rotations, IDP, and FJFs (Figs. 3a–3k)
within the range of numerical predictions22 though
generally in disagreement with in vivo data for IDP76
and intervertebral rotations.45,47,48

Hybrid Model Predictions


Convergence
Depending on the simulated task and after 3-4
iterations (20-30 min computational time), similar
kinematics were found between the FE and MS models
(differences of < 1 mm and 0.6 in, respectively, dis-
placements and rotations of the T12–S1 vertebral
centroids) under identical passive properties and
muscle forces (Fig. 4). For different tasks, the com-
pression-induced increase in disc stiffness remained, in
agreement with in vitro measurements,44 below 20%.

L5–S1 Load Sharing


Resultant spinal loads (total joint compression and
shear forces) and passive joint moments were almost
identical in both models. Contributions of various
joint structures to resist the overall spinal forces/mo-
ments could however be estimated only by the hybrid FIGURE 2. Comparisons between the passive FE model
predictions and in vitro/numerical data obtained under (a)
model, i.e., the passive FE model under muscle forces moment of up to 10 Nm to the L4–L5 disc-alone model
and gravity loads (Fig. 5). Generally, the disc had the (without ligaments and facets), (b) moment of up to 7.5 Nm
greatest contribution to resist compression and shear to the L1–L5 intact model (with ligaments and facets), and (c)
a follower load (FL) of up to 1000 N. Numerical ranges are
loads (Figs. 5a and 5b). Ligaments resisted greater provided based on eight different validated FE models
shear forces at larger trunk flexion angles while facets available in the literature.22 In vitro data for the disc-alone
resisted greater shear and compression loads in tasks moment-RoM, intact L1–L5 moment-RoM, and FL-IDP graphs
are based on the work of Heuer et al.,27 Rohlmann et al.57 and
involving larger extension (or less flexion) angles Brinckmann and Grootenboer,13 respectively. For the L1–L5
(Figs. 5a and 5b). Predicted IDPs, mean (left and RoM–moment curve in lateral bending, our model being in
right) FJFs, and maximal disc fiber strains were found agreement with several others22 shows a rather linear
behavior in contrast to in vitro data. The moment–RoM
to be disc-level and task dependent (Fig. 6). Regardless response in lateral bending is influenced by the facet
of the simulated task, the maximal IDP occurred either contact surfaces that vary from an individual to another.
at the upper (L1–L2) or lowermost (L5–S1) disc level
(Fig. 6a). Moreover, IDP increased at all lumbar levels right and flexed postures and considerably increased,
with the trunk flexion and extension from the relaxed especially at lower lumbar levels, in tasks involving
upright posture. FJFs were relatively small in the up- trunk extension (Fig. 6b). Disc fiber strains were lar-
KHODDAM-KHORASANI et al.

FIGURE 3. Passive FE model predictions for intervertebral rotations, intradiscal pressure (IDP), and facet joint forces (FJF)
(average of left and right contact forces) are compared with experimental/numerical data obtained under (a and b) 7.5 Nm flexion
moment and 1175 N follower load, (c–e) 7.5 Nm extension moment and 500 N follower load, (f–h) 7.8 Nm lateral bending moment
and 700 N follower load, and (i–k) 5.5 Nm axial moment and 720 N follower load. Numerical ranges are provided based on eight
different validated FE models available in the literature.22 Experimental data for the intervertebral rotations are based on the in vivo
three-dimensional radiographic images of lumbar spine (Pearcy and Tibrewal46; Pearcy et al.45; Pearcy47) while those for the IDP
data are based on the measured in vivo values using a pressure transducer inserted into the L4–L5 disc.75 Note that these loading
conditions (combined moment and follower load) do not accurately represent any specific in vivo loading task but, similar to
previous model investigations, such crude comparisons with in vivo data are nevertheless made here. Moreover, the geometry of
our FE model does not necessarily match that of the subjects participated in the foregoing experimental studies. In vivo IDPs were
collected in only one subject thus neglecting the inter-individual variabilities.

gest either at the upper (L1–L2) or lowermost (L5–S1) outermost layer occurred in the anterolateral regions
disc and increased with the trunk flexion or extension (at disc mid-height) for all the simulated tasks. Fiber
from the upright posture (Fig. 6c). Radial location of strains in the innermost layers were mainly influenced
the maximum fiber strain was generally in the inner- by intervertebral rotations while in the outermost
most layer while its circumferential position changed layers were primarily affected by the disc bulging (i.e.,
depending on the task; in posterolateral regions (at the IDP). For the two load-holding tasks, segments in
upper annulus-endplate junction) in neutral upright extension (L2–S1) had their maximum fiber strains at
and flexed tasks63 and in lateral regions (at the lower the innermost layer in lateral regions (at the lower
annulus-endplate junction) in extended (i.e., load- annulus-endplate junction) and for the outermost layer
holding) tasks (Fig. 7). Maximal fiber strain for the in the anterolateral regions (at the disc mid-height).
Trunk Hybrid Passive–Active Model for T12–S1 Responses

FIGURE 4. Schematic of the passive FE model in the undeformed (supine posture without any loading) and deformed (full flexion
under gravity loads and muscle exertions) configurations at different iterations. Symbols indicate position of geometric center of
vertebrae.

In Vivo vs. In Vitro Loading Conditions activities. To do this, trunk muscle forces predicted by
a musculoskeletal (MS) model of the thoracolumbar
The in vitro loading proposed elsewhere to simulate
spine were used to drive a mechanically-equivalent,
most realistically in vivo forward flexion (7.5 Nm
geometrically-detailed, passive FE model of the T12–
flexion plus a FL of 1175 N),22,60 when applied to the
S1 spine. For a given simulated task and after an
passive FE model resulted in different patterns of IDPs
iterative approach, the MS and passive FE models
(e.g., decreasing from upper toward lower levels) and
reached similar geometries (height, lordosis, and ver-
disc fiber strains as compared to those predicted by the
tebral orientations), deformed configurations, gravity
hybrid model under the in vivo full flexion task (Fig. 6).
and muscle loadings, boundary conditions, and passive
stiffness thus confirming our first hypothesis that an
Validation of the Hybrid Model for the CoR and IDP
equivalent detailed passive FE model of a MS model
For the tasks involving trunk flexion and in general can be used to predict joint load sharing under similar
agreement with in vivo measurements,46 the T12–S1 kinematics, gravity and muscle forces. It is to be noted
CoRs for both FE and MS models were located ~ 6– that the iterative interaction between the two models
8 mm posterior to the disc centroid (Fig. 8a). In cor- that results in the accurate estimation of muscle forces
roboration, location of the L2–L3 and L3–L4 CoRs in accounting for the complex nonlinear passive response
trunk flexion of 45 (with pelvic being constrained) has and of detailed passive structure load-sharing under
also been measured (radiographic imaging) to be at realistic external loads is original. Predictions of the
~ 5.0 mm posterior to vertebral body centroid.77 hybrid model for the L4–L5 IDPs and L1–S1 CoRs
Moreover, predicted and in vivo76 L4–L5 IDPs were in showed adequate agreements with measured in vivo
satisfactory agreement (R2 = 0.99 and RMSE= 0.38 data. Results also indicated that pure moments com-
MPa) (Fig. 8b). bined with a single-valued FL fail to mimic the in vivo
loading generated by muscle exertions and gravity
loads thus confirming our second hypothesis.
DISCUSSION
Limitations
This study presents a novel hybrid passive–active
model of the T12–S1 spine to determine tissue stresses The optimization algorithm used in the MS model
and strains, joint load sharing, and CoRs during in vivo to estimate muscle forces, i.e., minimizing sum of cu-
KHODDAM-KHORASANI et al.

FIGURE 5. Comparison of (a) compression, (b) shear and (c) moment sharing of different L5–S1 joint structures (disc, ligaments
and facets) under the simulated in vivo tasks by the hybrid-model (for the sake of clarity, compression sharing on the vertical axis
was shown from 75% to 100%). For facets, sum of bilateral contact forces is considered. For all the simulated tasks facets resist
extension moment (i.e., generate flexion moment). As ALL and PLL are attached to the disc, their forces were included in
calculation of disc forces.

bed muscle stresses, overlooked antagonistic (mainly unloaded tasks.7 The other cost functions that also
abdominal) muscle forces. It has been shown that this estimate muscle activities in qualitative agreement with
cost function generates back muscle activities in best EMG data generate similar lumbar compression and
qualitative agreement, as compared to seven other cost shear loads with differences less than 10%. Moreover,
functions, with EMG data for a number of loaded and only major local and global muscles attaching the
Trunk Hybrid Passive–Active Model for T12–S1 Responses

FIGURE 6. (a) Intradiscal pressure values (IDP), (b) mean (left and right) facet joint forces (FJFs), and (c) maximal disc fiber
strains at all lumbar levels during different simulated in vivo tasks by the hybrid model. Results of the passive FE model under the
proposed in vitro loading condition for flexion (7.5 Nm and 1175 N FL)22,60 are also depicted.

pelvis-sacrum, respectively, to the lumbar and thoracic agreements could have also been obtained with some-
spines were considered in the present study while what alternative sets of material properties for the discs
neglecting the smaller intersegmental muscles; a sim- and ligaments.22,36 In a recent study, we showed that
plification whose effect on the predicted joint load changes in the stiffness of passive tissues (ligament
sharing remains to be investigated. Neglecting antag- properties and their initial laxities) could affect the
onistic muscle forces in the present study is also sug- predicted joint load sharing under moment/FL loading
gested to underestimate spine loads.72 Predicted muscle conditions.36 For this reason, a model should prefer-
activities in agonistic (back) muscles, however, re- ably be validated as much as possible in different
mained in qualitative agreement with our measured loading conditions (magnitudes, directions and com-
EMG data3,4,6 even after the iterative procedure that binations). It is also to be noted that while wrapping of
resulted in small to moderate changes (< 20%) in the back muscles around bony structures was accounted
estimated muscle forces. for in the FE model, the inter-muscular forces were
Although the passive FE model was validated by neglected. Muscles were simulated by 1D, rather than
comparing its predictions for RoMs, IDPs, and FJFs 3D, elements in the current FE model. Moreover,
under various loadings with available measured data estimation of annulus fiber strains were made assuming
(Figs. 2 and 3), it is however to be noted that such constant annulus fibrosus criss-cross angles at ± 30,
KHODDAM-KHORASANI et al.

FIGURE 7. Radial and circumferential locations (shown by thick solid red and blue lines for, respectively, the innermost and
outermost layer) and magnitudes of the maximal fiber strain in different levels and tasks.

FIGURE 8. Predictions of the hybrid FE/MS model as compared with experimental data obtained under in vivo loading conditions:
(a) sagittal T12–S1 center of rotations (CoRs) for tasks involving trunk flexion and (b) L4–L5 intradiscal pressure (IDP) during all
simulated tasks. IDP predictions based on the MS model for two different L4–L5 disc cross-sectional areas, CSA, (1455 mm2 similar
to the disc CSA in the MS and passive FE models, R2 = 0.88 and RMSE= 0.35 MPa, or 1800 mm2 similar to the disc CSA of the
subject that participated in the measurement study,75 R2 = 0.91 and RMSE= 0.34 MPa) are also depicted. Center of rotation of each
vertebra, relative to its inferior vertebra in different tasks was depicted on the schematic of the FE model in upright posture and
was compared with radiographic in vivo measurements48 (in vivo CoR for the T12–L1 segment was not available). In vivo
measurements of the L4–L5 IDPs are based on data reported by Wilke et al.75
Trunk Hybrid Passive–Active Model for T12–S1 Responses

while the neglected additional complexity of the disc Analyses of Findings


fiber organization and likely inhomogeneous distri-
The in vitro loading proposed elsewhere to most
bution of these angles18,28,79 could influence predic-
realistically represent in vivo forward flexion22,60 failed,
tions.40
when applied to the passive FE model, to yield pat-
All three models (MS, FE, and hybrid) were vali-
terns of IDPs and disc fiber strains in agreement with
dated as much as possible by comparing their pre-
those predicted by the hybrid model under forward
dictions with available experimental data. For MS
flexion tasks. For instance, the suggested in vitro
model, estimated muscle activities and L4–L5 com-
loading (7.5 Nm flexion plus 1175 N FL) resulted in a
pressive loads were compared with, respectively,
monotonically decreasing IDP from the upper
EMG data3,4,6 and measured IDPs (Fig. 8b). The
(~ 1.4 MPa at the L1–L2) toward the lower
passive FE model was validated by the comparison of
(~ 1.0 MPa at the L5–S1) discs (Fig. 6a). This
its predictions under pure moment in different planes,
decreasing trend was in agreement with predictions of
pure follower load, and combinations thereof with
five other passive FE models under the proposed
measurements (Figs. 2 and 3). Finally, for the hybrid
in vitro loading condition.22 Our hybrid model under
model, that is built from these validated MS and FE
muscle exertions and gravity loads, however, predicted
models, predictions for the T12–S1 CoRs and L4–L5
relatively larger IDPs at the lower discs for all flexion
IDPs were compared with in vivo data (Fig. 8).
activities is expected under the additional segmental
Comparisons of the predicted and measured L4–L5
masses and local lumbar muscle forces acting from the
IDPs (Fig. 8b) should be carried out in the light of
upper toward the lower levels. The same discrepancy is
few facts. Our MS and FE models were different
found when comparing the maximum disc fiber strains
from the single subject that participated in the IDP
(Fig. 6c).
measurement study76 in terms of size, lordosis, mus-
The disagreement between predictions of the FE
culature, as well as geometry and structural-material
model (under in vitro loading) and measured in vivo
properties of the passive tissues. A more accurate,
data for IDP and intervertebral rotations is evident
though currently unavailable, validation procedure
under all loads considered (Fig. 3). Such findings fur-
involves modeling and measurements on the same
ther demonstrate that an in vivo loading condition
subjects. Nevertheless, a number of parameters
cannot accurately be represented by a constant varia-
including mechanical properties of the passive tissues,
tion of both moment and FL (i.e., in vitro loading)
accurate geometry of ligaments and discs, as well as
throughout the lumbar spine. In other words, the
detailed musculature cannot yet accurately be made
complex recruitment patterns of trunk muscles pre-
subject-specific. Note that the size of the subject
dicted by our MS model cannot be simply mimicked by
based on which the MS and FE models are devel-
a constant FL. In the full flexion task, for instance,
oped (68.4 kg and 174.5 cm) match closely that of
segmental loads at the T12–L1 through L5–S1 levels
the subject participated in the IDP measurement
were, respectively, ~ 1283, 1346, 1357, 1459, 1558,
study (70 kg and 173.9 cm).76 This concordance
1508 N in compression and ~ 100, 92, 39, 70, 79 and
nevertheless makes presented comparisons more rel-
304 N in anterior shear forces. Such proximal to distal
evant. The L4–L5 disc cross-sectional areas were
increases in compression forces, even when neglecting
however quite different between the FE model
accompanying shear forces, demonstrates the need for
(1455 mm2) and the subject that participated in the
a variable FL instead of a constant one.30,60 Multiple
in vivo measurement study (1800 mm2).76 Therefore,
FL applications67 along with flexion rotations or mo-
comparisons were made (Fig. 8b) for both values in
ments that both vary from a level to another are hence
the MS model. Differences in the disc heights also
essential when attempting to accurately simulate an
affect model predictions.22,39,78 Furthermore, in vivo
in vivo task.
IDPs recorded on one subject overlook the inherent
With the maximum fiber strains remained below
inter-individual variabilities. The equation relating
their failure tensile strain, i.e., ~ 15–25%,63,69 no disc
the L4–L5 IDP to joint compression and flexion in
failure is expected in the currently simulated tasks.
the MS model,24 that was used in this work, was
However, the posterolateral region (near annulus–
developed for rather small extension intersegmental
endplate junctions) could be at the risk of failure ini-
rotations (< 2) and thus may not be as accurate for
tiation under greater compression and flexion rota-
IDP predictions in tasks involving larger extension
tions.1,63,74 Our results are also corroborated by
rotations (load holding away from the chest). Finally,
findings of both in vivo (imaging and intraoperative
changes in the applied L1–L5 lumbar rotations in the
examinations) and in vitro studies that indicate lumbar
MS model when simulating a specific in vivo task
disc herniation commonly occurs at the endplate or
could also influence predictions and hence associated
annulus-endplate junctions.53,73,74
comparisons.
KHODDAM-KHORASANI et al.

The MS model with T12–S1 segments simulated by subject-specific modeling investigation. J. Biomech. 70:
shear-deformable beam elements located at fixed pos- 102–112, 2018.
3
Arjmand, N., D. Gagnon, A. Plamondon, A. Shirazi-Adl,
terior distances (4 mm) to the disc geometric centers
and C. Lariviere. Comparison of trunk muscle forces and
predicted CoRs in flexion tasks that differed from those spinal loads estimated by two biomechanical models. Clin.
of the FE model (under in vivo loading conditions) by Biomech. 24:533–541, 2009.
4
~ 3 mm in average for all levels/tasks. Compared to the Arjmand, N., D. Gagnon, A. Plamondon, A. Shirazi-Adl,
FE model, the MS model generally predicted more and C. Lariviere. A comparative study of two trunk
biomechanical models under symmetric and asymmetric
posterior CoRs at the upper levels whereas more
loadings. J. Biomech. 43:485–491, 2010.
anterior CoRs at the lower vertebrae. Nevertheless, 5
Arjmand N. Computational biomechanics of the human
both MS and FE models predicted CoRs to be within spine in static lifting tasks. PhD Thesis, Ecole Polytech-
the range of those recorded in vivo. Our recent modeling nique de Montreal, pp. 218-220, 2006.
6
investigation has demonstrated the substantial effects Arjmand, N., and A. Shirazi-Adl. Model and in vivo
studies on human trunk load partitioning and stability in
of the joint (simulated as beams or spherical pivots)
isometric forward flexions. J. Biomech. 39:510–521,
positioning on muscle forces and spinal loads.23 The 2006.
alterations in CoR position with loading at different 7
Arjmand, N., and A. Shirazi-Adl. Sensitivity of kinematics-
levels and their effects on spinal loads hence caution based model predictions to optimization criteria in static
against the use of fixed spherical joints regularly placed lifting tasks. Med. Eng. Phys. 28:504–514, 2006.
8
Arjmand, N., A. Shirazi-Adl, and B. Bazrgari. Wrapping
at the joint geometric center.11,14,25,29
of trunk thoracic extensor muscles influences muscle forces
In conclusion, this study proposed a novel approach and spinal loads in lifting tasks. Clin. Biomech. 21:668–675,
to estimate the lumbar joint load sharing, internal 2006.
9
stresses and strains, and CoRs during various in vivo Azari, F., N. Arjmand, A. Shirazi-Adl, and T. Rahimi-
activities by using a combined hybrid (validated MS Moghaddam. A combined passive and active muscu-
loskeletal model study to estimate L4–L5 load sharing. J.
and passive FE) model. Predictions of this hybrid
Biomech. 70:157–165, 2018.
model for the L4–L5 IDPs and L1–S1 CoRs compared 10
Bashkuev, M., P.-P. A. Vergroesen, M. Dreischarf, C.
well with the measured in vivo values. It was also Schilling, A. J. van der Veen, H. Schmidt, and I. Kingma.
shown that the loading conditions proposed in the Intradiscal pressure measurements: a challenge or a rou-
literature with fixed moment and FL magnitudes tine? J. Biomech. 49:864–868, 2016.
11
Bassani, T., E. Stucovitz, Z. Qian, M. Briguglio, and F.
throughout the lumbar spine cannot adequately rep-
Galbusera. Validation of the AnyBody full body muscu-
resent in vivo loading conditions. Passive FE models, loskeletal model in computing lumbar spine loads at L4L5
when applied to assess the risk of injury or to design level. J. Biomech. 58:89–96, 2017.
12
implants, should therefore be used under loads closely Breau, C., A. Shirazi-Adl, and J. De Guise. Reconstruction
replicating in vivo loading conditions. of a human ligamentous lumbar spine using CT images—a
three-dimensional finite element mesh generation. Ann.
Biomed. Eng. 19:291–302, 1991.
ELECTRONIC SUPPLEMENTARY MATERIAL 13
Brinckmann, P., and H. Grootenboer. Change of disc
height, radial disc bulge, and intradiscal pressure from
The online version of this article (https://doi.org/10. discectomy an in vitro investigation on human lumbar
1007/s10439-018-2078-7) contains supplementary discs. Spine 16:641–646, 1991.
14
Bruno, A. G., M. L. Bouxsein, and D. E. Anderson.
material, which is available to authorized users. Development and validation of a musculoskeletal model of
the fully articulated thoracolumbar spine and rib cage. J.
Biomech. Eng. 137:081003, 2015.
ACKNOWLEDGMENTS 15
Chen, C.-S., C.-K. Cheng, C.-L. Liu, and W.-H. Lo. Stress
analysis of the disc adjacent to interbody fusion in lumbar
This work was supported by grants from Sharif spine. Med. Eng. Phys. 23:485–493, 2001.
University of Technology (Tehran, Iran). Assistance of 16
Cholewicki, J., and S. M. McGill. Mechanical stability of
Mr. H. Asadi and Dr. S. Naserkhaki in the finite ele- the in vivo lumbar spine: implications for injury and
ment modeling and Mr. A.H. Eskandari in the mus- chronic low back pain. Clin. Biomech. 11:1–15, 1996.
17
culoskeletal modeling is greatly appreciated. Damsgaard, M., J. Rasmussen, S. T. Christensen, E. Sur-
ma, and M. de Zee. Analysis of musculoskeletal systems in
the anybody modeling system. Simul. Model. Pract. Theory
14:1100–1111, 2006.
18
REFERENCES Dittmar, R., M. M. van Rijsbergen, and K. Ito. Moder-
ately degenerated human intervertebral disks exhibit a less
1
geometrically specific collagen fiber orientation distribu-
Adams, M., and W. Hutton. Prolapsed intervertebral disc: tion. Global Spine J 6:439–446, 2016.
a hyperflexion injury. Spine 7:184–191, 1982. 19
2
Dreischarf, M., A. Rohlmann, G. Bergmann, and T. Zan-
Akhavanfar, M., H. Kazemi, A. Eskandari, and N. Arjmand. der. Optimised loads for the simulation of axial rotation in
Obesity and spinal loads: a combined MR imaging and the lumbar spine. J. Biomech. 44:2323–2327, 2011.
Trunk Hybrid Passive–Active Model for T12–S1 Responses
20
Dreischarf, M., A. Rohlmann, G. Bergmann, and T. Zan- property datasets on biomechanics of a lumbar L4–L5 fi-
der. Optimised in vitro applicable loads for the simulation nite element model. J. Biomech. 70:33–42, 2018.
37
of lateral bending in the lumbar spine. Med. Eng. Phys. Naserkhaki, S., and M. El-Rich. Sensitivity of lumbar spine
34:777–780, 2012. response to follower load and flexion moment: finite ele-
21
Dreischarf, M., A. Shirazi-Adl, N. Arjmand, A. Rohl- ment study. Comput. Methods Biomech. Biomed. 20:550–
mann, and H. Schmidt. Estimation of loads on human 557, 2017.
38
lumbar spine: a review of in vivo and computational model Naserkhaki, S., J. L. Jaremko, and M. El-Rich. Effects of
studies. J. Biomech. 49:833–845, 2016. inter-individual lumbar spine geometry variation on load-
22
Dreischarf, M., T. Zander, A. Shirazi-Adl, C. Puttlitz, C. sharing: geometrically personalized finite element study. J.
Adam, C. Chen, V. Goel, A. Kiapour, Y. Kim, and K. Biomech. 49:2909–2917, 2016.
39
Labus. Comparison of eight published static finite element Natarajan, R. N., and G. B. Andersson. The influence of
models of the intact lumbar spine: Predictive power of lumbar disc height and cross-sectional area on the
models improves when combined together. J. Biomech. mechanical response of the disc to physiologic loading.
47(8):1757–1766, 2014. Spine 24:1873, 1999.
23 40
Ghezelbash, F., A. Eskandari, A. Shirazi-Adl, N. Arjmand, Noailly, J., J. A. Planell, and D. Lacroix. On the collagen
Z. El-Ouaaid, and A. Plamondon. Effects of motion seg- criss-cross angles in the annuli fibrosi of lumbar spine finite
ment simulation and joint positioning on spinal loads in element models. Biomech. Model. Mechanobiol. 10:203–
trunk musculoskeletal models. J. Biomech. 70:149–156, 219, 2011.
41
2018. Noailly, J., H.-J. Wilke, J. A. Planell, and D. Lacroix. How
24
Ghezelbash, F., A. Shirazi-Adl, N. Arjmand, Z. El-Ouaaid, does the geometry affect the internal biomechanics of a
and A. Plamondon. Subject-specific biomechanics of trunk: lumbar spine bi-segment finite element model? Conse-
musculoskeletal scaling, internal loads and intradiscal quences on the validation process. J. Biomech. 40:2414–
pressure estimation. Biomech. Model. Mechanobiol. 2425, 2007.
42
15:1699–1712, 2016. Okushima, H. Study on hydrodynamic pressure of lumbar
25
Hajihosseinali, M., N. Arjmand, A. Shirazi-Adl, F. intervertebral disc. Nihon geka hokan 39:45, 1970.
43
Farahmand, and M. Ghiasi. A novel stability and kine- Park, W. M., K. Kim, and Y. H. Kim. Effects of degen-
matics-driven trunk biomechanical model to estimate erated intervertebral discs on intersegmental rotations,
muscle and spinal forces. Med. Eng. Phys. 36:1296–1304, intradiscal pressures, and facet joint forces of the whole
2014. lumbar spine. Comput. Biol. 43:1234–1240, 2013.
26 44
Han, K.-S., A. Rohlmann, S.-J. Yang, B. S. Kim, and T.- Patwardhan, A. G., R. M. Havey, G. Carandang, J. Si-
H. Lim. Spinal muscles can create compressive follower monds, L. I. Voronov, A. J. Ghanayem, K. P. Meade, T.
loads in the lumbar spine in a neutral standing posture. M. Gavin, and O. Paxinos. Effect of compressive follower
Med. Eng. Phys. 33:472–478, 2011. preload on the flexion–extension response of the human
27
Heuer, F., H. Schmidt, L. Claes, and H.-J. Wilke. Stepwise lumbar spine. J. Orthop. Res. 21:540–546, 2003.
45
reduction of functional spinal structures increase vertebral Pearcy, M. J. Stereo radiography of lumbar spine motion.
translation and intradiscal pressure. J. Biomech. 40:795– Acta Orthop. Scand. 56:1–45, 1985.
46
803, 2007. Pearcy, M. J., and N. Bogduk. Instantaneous axes of
28
Holzapfel, G. A., C. Schulze-Bauer, G. Feigl, and P. rotation of the lumbar intervertebral joints. Spine 13:1033–
Regitnig. Single lamellar mechanics of the human lumbar 1041, 1988.
47
anulus fibrosus. Biomech. Model. Mechanobiol. 3:125–140, Pearcy, M., I. Portek, and J. Shepherd. Three-dimensional
2005. x-ray analysis of normal movement in the lumbar spine.
29
Ignasiak, D., S. J. Ferguson, and N. Arjmand. A rigid Spine 9:294–297, 1984.
48
thorax assumption affects model loading predictions at the Pearcy, M., and S. Tibrewal. Axial rotation and lateral
upper but not lower lumbar levels. J. Biomech. 49:3074– bending in the normal lumbar spine measured by three-
3078, 2016. dimensional radiography. Spine 9:582–587, 1984.
30 49
Kim, K., Y. H. Kim, and S. Lee. Investigation of optimal Pearsall, D. J., J. G. Reid, and L. A. Livingston. Segmental
follower load path generated by trunk muscle coordination. inertial parameters of the human trunk as determined from
J. Biomech. 44:1614–1617, 2011. computed tomography. Ann. Biomed. Eng. 24:198–210,
31
Little, J. P. Finite element modelling of anular lesions in the 1996.
50
lumbar intervertebral disc. Brisbane: Queensland Univer- Polga, D. J., B. P. Beaubien, P. M. Kallemeier, K. P.
sity of Technology, 2004. Schellhas, W. D. Lew, G. R. Buttermann, and K. B. Wood.
32
Nachemson, A. Towards a better understanding of low- Measurement of in vivo intradiscal pressure in healthy
back pain: a review of the mechanics of the lumbar disc. thoracic intervertebral discs. Spine 29:1320–1324, 2004.
51
Rheumatology 14:129–143, 1975. Potvin, J. Use of NIOSH equation inputs to calculate
33
Nachemson, A. L. Disc pressure measurements. Spine lumbosacral compression forces. Ergonomics 40:691–707,
6:93–97, 1981. 1997.
34 52
Nachemson, A., and G. Elfstrom. Intravital dynamic Rajaee, M. A., N. Arjmand, A. Shirazi-Adl, A. Plamon-
pressure measurements in lumbar discs. Scand J Rehabil don, and H. Schmidt. Comparative evaluation of six
2:1–40, 1970. quantitative lifting tools to estimate spine loads during
35
Nachemson, A., and J. M. Morris. In vivo measurements static activities. Appl. Ergon. 48:22–32, 2015.
53
of intradiscal pressure. J. Bone Joint Surg. 46:1077–1092, Rajasekaran, S., N. Bajaj, V. Tubaki, R. M. Kanna, and A.
1964. P. Shetty. ISSLS prize winner: the anatomy of failure in
36
Naserkhaki, S., N. Arjmand, A. Shirazi-Adl, F. Farah- lumbar disc herniation: an in vivo, multimodal, prospective
mand, and M. El-Rich. Effects of eight different ligament study of 181 subjects. Spine 38:1491–1500, 2013.
KHODDAM-KHORASANI et al.
54
Rohlmann, A., U. Arntz, F. Graichen, and G. Bergmann. a three-dimensional nonlinear finite element study. Spine
Loads on an internal spinal fixation device during sitting. J. 9:120–134, 1984.
69
Biomech. 34:989–993, 2001. Stemper, B. D., J. L. Baisden, N. Yoganandan, B. S.
55
Rohlmann, A., L. Bauer, T. Zander, G. Bergmann, and H.- Shender, and D. J. Maiman. Mechanical yield of the lum-
J. Wilke. Determination of trunk muscle forces for flexion bar annulus: a possible contributor to instability. J. Neu-
and extension by using a validated finite element model of rosurg. Spine 21:608–613, 2014.
70
the lumbar spine and measured in vivo data. J. Biomech. Stokes, I. A., and M. Gardner-Morse. Lumbar spine
39:981–989, 2006. maximum efforts and muscle recruitment patterns pre-
56
Rohlmann, A., G. Bergmann, and F. Graichen. Loads on dicted by a model with multijoint muscles and joints with
internal spinal fixators measured in different body posi- stiffness. J. Biomech. 28:173177–175186, 1995.
71
tions. Eur. Spine J. 8:354–359, 1999. Stokes, I. A., and M. Gardner-Morse. Quantitative anat-
57
Rohlmann, A., S. Neller, L. Claes, G. Bergmann, and H.-J. omy of the lumbar musculature. J. Biomech. 32:311–316,
Wilke. Influence of a follower load on intradiscal pressure 1999.
72
and intersegmental rotation of the lumbar spine. Spine Van Dieen, J., and I. Kingma. Effects of antagonistic co-
26:E557–E561, 2001. contraction on differences between electromyography
58
Rohlmann, A., R. Petersen, V. Schwachmeyer, F. Grai- based and optimization based estimates of spinal forces.
chen, and G. Bergmann. Spinal loads during position Ergonomics 48:411–426, 2005.
73
changes. Clin. Biomech. 27:754–758, 2012. Wade, K. R., P. A. Robertson, A. Thambyah, and N. D.
59
Rohlmann, A., T. Zander, F. Graichen, M. Dreischarf, and Broom. How healthy discs herniate: a biomechanical and
G. Bergmann. Measured loads on a vertebral body microstructural study investigating the combined effects of
replacement during sitting. Spine J. 11:870–875, 2011. compression rate and flexion. Spine 39:1018–1028, 2014.
60 74
Rohlmann, A., T. Zander, M. Rao, and G. Bergmann. Wade, K., P. Robertson, A. Thambyah, and N. Broom.
Realistic loading conditions for upper body bending. J. ‘‘Surprise’’ loading in flexion increases the risk of disc
Biomech. 42:884–890, 2009. herniation due to annulus–endplate junction failure: a
61
Sato, K., S. Kikuchi, and T. Yonezawa. In vivo intradiscal mechanical and microstructural investigation. Spine
pressure measurement in healthy individuals and in 40(12):891–901, 2015.
75
patients with ongoing back problems. Spine 24:2468, 1999. Wilke, H. J., P. Neef, M. Caimi, T. Hoogland, and L. E.
62
Schultz, A., G. Andersson, R. Ortengren, K. Haderspeck, Claes. New in vivo measurements of pressures in the
and A. Nachemson. Loads on the lumbar spine: validation intervertebral disc in daily life. Spine 24:755–762, 1999.
76
of a biomechanical analysis by measurements of intradiscal Wilke, H.-J., P. Neef, B. Hinz, H. Seidel, and L. Claes.
pressures and myoelectric signals. JBJS 64:713–720, 1982. Intradiscal pressure together with anthropometric data–a
63
Shiraz-Adl, A. Strain in fibers of a lumbar disc: analysis of data set for the validation of models. Clin. Biomech.
the role of lifting in producing disc prolapse. Spine 14:96– 16:S111–S126, 2001.
77
103, 1989. Xia, Q., S. Wang, M. Kozanek, P. Passias, K. Wood, and
64
Shirazi-Adl, A. Biomechanics of the lumbar spine in G. Li. In-vivo motion characteristics of lumbar vertebrae in
sagittal/lateral moments. Spine 19:2407–2414, 1994. sagittal and transverse planes. J. Biomech. 43:1905–1909,
65
Shirazi-Adl, A. Analysis of large compression loads on 2010.
78
lumbar spine in flexion and in torsion using a novel Zander, T., M. Dreischarf, A.-K. Timm, W. W. Baumann,
wrapping element. J. Biomech. 39:267–275, 2006. and H. Schmidt. Impact of material and morphological
66
Shirazi-Adl, A., A. M. Ahmed, and S. C. Shrivastava. parameters on the mechanical response of the lumbar spi-
Mechanical response of a lumbar motion segment in axial ne—a finite element sensitivity study. J. Biomech. 53:185–
torque alone and combined with compression. Spine 190, 2017.
79
11:914–927, 1986. Zhu, D., G. Gu, W. Wu, H. Gong, W. Zhu, T. Jiang, and
67
Shirazi-Adl, A., and M. Parnianpour. Load-bearing and Z. Cao. Micro-structure and mechanical properties of
stress analysis of the human spine under a novel wrapping annulus fibrous of the L4–5 and L5–S1 intervertebral discs.
compression loading. Clin. Biomech. 15:718–725, 2000. Clin. Biomech. 23:S74–S82, 2008.
68
Shirazi-adl, S. A., S. C. Shrivastava, and A. M. Ahmed.
Stress analysis of the lumbar disc-body unit in compression

You might also like