Skip to main content
Log in

A forward dynamics simulation of human lumbar spine flexion predicting the load sharing of intervertebral discs, ligaments, and muscles

  • Original Paper
  • Published:
Biomechanics and Modeling in Mechanobiology Aims and scope Submit manuscript

Abstract

Determining the internal dynamics of the human spine’s biological structure is one essential step that allows enhanced understanding of spinal degeneration processes. The unavailability of internal load figures in other methods highlights the importance of the forward dynamics approach as the most powerful approach to examine the internal degeneration of spinal structures. Consequently, a forward dynamics full-body model of the human body with a detailed lumbar spine is introduced. The aim was to determine the internal dynamics and the contribution of different spinal structures to loading. The multi-body model consists of the lower extremities, two feet, shanks and thighs, the pelvis, five lumbar vertebrae, and a lumped upper body including the head and both arms. All segments are modelled as rigid bodies. 202 muscles (legs, back, abdomen) are included as Hill-type elements. 58 nonlinear force elements are included to represent all spinal ligaments. The lumbar intervertebral discs were modelled nonlinearly. As results, internal kinematics, muscle forces, and internal loads for each biological structure are presented. A comparison between the nonlinear (new, enhanced modelling approach) and linear (standard modelling approach, bushing) modelling approaches of the intervertebral disc is presented. The model is available to all researchers as ready-to-use C/C++ code within our in-house multi-body simulation code demoa with all relevant binaries included.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Abbreviations

IVD:

Intervertebral disc

MTC:

Muscle–tendon complex

MB:

Multi-body

FE:

Finite element

DOF:

Degree of freedom

CT:

Computed tomography

CE:

Contractile element

IVFE:

Intervertebral flexion–extension

ROM:

Range of motion

ALL:

Anterior longitudinal ligament

PLL:

Posterior longitudinal ligament

LF:

Ligamentum flavum

SSL:

Supraspinal ligament

ISL:

Interspinal ligament

RA:

Rectus abdominis muscle

EO:

External oblique muscle

IO:

Internal oblique muscle

PM:

Psoas major muscle

MF:

Multifidus

IT_m:

Intertransversarii mediales

LTpL:

Longissimus thoracis pars lumborum

IL:

Iliocostalis lumborum pars lumborum

C1–C7:

Cervical vertebrae

T1–T12:

Thoracic vertebrae

L1–L5:

Lumbar vertebrae

S1–S5:

Sacral vertebrae

Sagittal axis:

Dorsal to ventral (\(=\!x\)-axis)

Longitudinal axis:

Caudal to cranial (\(=\!z\)-axis)

Graphic primitive:

Geometric primitive, describes the simplest geometric objects that the system can handle; in this model, the graphic primitives represent respective body segments

References

  • Ashton-Miller J, Schultz A (1997) Biomechanics of the human spine. In: Mow VC, Hayes WC (eds) Basic orthopaedic biomechanics. Lippincott–Raven, New York

    Google Scholar 

  • Aspden RM (1992) Review of the functional anatomy of the spinal ligaments and the lumbar erector spinae muscles. Clin Anat 5(5):372–387

    Article  Google Scholar 

  • Belytschko TB, Privitzer E (1978) Refinement and validation of a three dimensional head-spine model. Report AMRL-TR-3-78-7, Aerospace Medical Research Laboratory, Wright-Patterson Air Force Base, Ohio

  • Berkson MH, Nachemson A, Schultz AB (1979) Mechanical properties of human lumbar spine motion segments-Part II: responses in compression and shear; influence of gross morphology. J Biomech Eng Trans ASME 101(1):53–57

    Article  Google Scholar 

  • Bogduk N (2005) Clinical anatomy of the lumbar spine and sacrum, 4th edn. Elsevier, Amsterdam

    Google Scholar 

  • Bogduk N, Macintosh JE, Pearcy MJ (1992) A universal model of the lumbar back muscles in the upright position. Spine 17(8):897–913

    Article  Google Scholar 

  • Busscher I, Ploegmakers JJW, Verkerke GJ, Veldhuizen AG (2010) Comparative anatomical dimensions of the complete human and porcine spine. Eur Spine J 19(7):1104–1114

    Article  Google Scholar 

  • Calisse J, Rohlmann A, Bergmann G (1999) Estimation of trunk muscle forces using the finite element method and in vivo loads measured by telemeterized internal spinal fixation devices. J Biomech 32(7):727–731

    Article  Google Scholar 

  • Campbell-Kyureghyan N, Jorgensen M, Burr D, Marras W (2005) The prediction of lumbar spine geometry: method development and validation. Clin Biomech 20(5):455–464

    Article  Google Scholar 

  • Carman AB, Milburn PD (2005) Dynamic coordinate data for describing muscle-tendon paths: a mathematical approach. J Biomech 38(4):943–951

    Article  Google Scholar 

  • Cassinelli E, Kang J (2000) Current understanding of lumbar disc degeneration. Oper Tech Orthop 10(4):254–262

    Article  Google Scholar 

  • Chazal J, Tanguy A, Bourges M, Gaurel G, Escande G, Guillot M, Vanneuville G (1985) Biomechanical properties of spinal ligaments and a histological study of the supraspinal ligament in traction. J Biomech 18(3):167–176

    Article  Google Scholar 

  • Cholewicki J, McGill SM (1996) Mechanical stability of the in vivo lumbar spine: implications for injury and chronic low back pain. Clin Biomech 11(1):1–15

    Article  Google Scholar 

  • Christophy M, Faruk Senan NA, Lotz JC, O’Reilly OM (2011) A musculoskeletal model for the lumbar spine. Biomech Model Mechanobiol 11(1–2):19–34

    Google Scholar 

  • Daggfeldt K, Thorstensson A (2003) The mechanics of back-extensor torque production about the lumbar spine. J Biomech 36(6):815–825

    Article  Google Scholar 

  • de Zee M, Hansen L, Wong C, Rasmussen J, Simonsen EB (2007) A generic detailed rigid-body lumbar spine model. J Biomech 40(6):1219–1227

    Article  Google Scholar 

  • Delp SL, Suryanarayanan S, Murray WM, Uhlir J, Triolo RJ (2001) Architecture of the rectus abdominis, quadratus lumborum, and erector spinae. J Biomech 34(3):371–375

    Article  Google Scholar 

  • Denavit J, Hartenberg R, Razi R (1965) Velocity, acceleration, and static-force analyses of spatial linkages. J Appl Mech 32:903–910

    Article  Google Scholar 

  • Denavit R, Hartenberg R (1955) A kinematic notation for lower pairs mechanisms based on matrices. J Appl Mech 22:215–221

    MathSciNet  MATH  Google Scholar 

  • Denoth J (1985) The dynamic behaviour of a three link model of the human body during impact with the ground. In: Winter D, Norman R, Wells R, Hayes K, Patla A (eds) Biomechanics IX-A, vol 5B., International Series on BiomechanicsHuman Kinetics Publishers, Champaign, pp 102–106

    Google Scholar 

  • Denoth J (1986) Load on the locomotor system and modelling. In: Nigg B (ed) Biomechanics of running shoes, chapter 3. Human Kinetics Publishers, Champaign, pp 63–116

    Google Scholar 

  • Denoth J, Gruber K, Keppler M, Ruder H (1985) Forces and torques during sport activities with high accelerations. In: Perren S, Schneider E (eds) Biomechanics: current interdisciplinary research, developments in biomechanics. Springer, Netherlands, pp 663–668

    Chapter  Google Scholar 

  • Desailly E, Sardain P, Khouri N, Yepremian D, Lacouture P (2010) The convex wrapping algorithm: a method for identifying muscle paths using the underlying bone mesh. J Biomech 43(13):2601–2607

    Article  Google Scholar 

  • DIN (1986) Körpermaße des Menschen. Tech Rep DIN 33402-2:1986-10

  • Eberlein R, Holzapfel GA, Fröhlich M (2004) Multi-segment FEA of the human lumbar spine including the heterogeneity of the annulus fibrosus. Comput Mech 34(2):147–163

    Article  MATH  Google Scholar 

  • Ehlers W, Karajan N, Markert B (2008) An extended biphasic model for charged hydrated tissues with application to the intervertebral disc. Biomech Model Mechanobiol 8(3):233–251

    Article  Google Scholar 

  • El-Rich M, Shirazi-Adl A, Arjmand N (2004) Muscle activity, internal loads, and stability of the human spine in standing postures: combined model and in vivo studies. Spine 29(23):2633–2642

  • Elias F (1958) Roentgen findings in the asymptomatic cervical spine. N Y State J Med 58(20):3300–3303

    Google Scholar 

  • Esat V, Acar M (2008) Viscoelastic finite element analysis of the cervical intervertebral discs in conjunction with a multi-body dynamic model of the human head and neck. Proc Inst Mech Eng Part H J Eng Med 223(2):249–262

    Article  Google Scholar 

  • Feldman AG (1986) Once more on the equilibrium-point hypothesis (lambda model) for motor control. J Mot Behav 18(1):17–54

    Article  Google Scholar 

  • Feldman AG, Levin MF (2009) The equilibrium-point hypothesis—past, present and future. Prog Mot Control Multidiscip Perspect 629:699–726

    Article  Google Scholar 

  • Gardner-Morse MG, Stokes IAF (2004) Structural behavior of human lumbar spinal motion segments. J Biomech 37(2):205–212

    Article  Google Scholar 

  • Gatton ML, Pearcy MJ, Pettet GJ (2011) Computational model of the lumbar spine musculature: implications of spinal surgery. Clin Biomech 26(2):116–122

    Article  Google Scholar 

  • Gerritsen KG, van den Bogert AJ, Nigg BM (1995) Direct dynamics simulation of the impact phase in heel-toe running. J Biomech 28(6):661–668

    Article  Google Scholar 

  • Gilad I, Nissan M (1986) A study of vertebra and disc geometric relations of the human cervical and lumbar spine. Spine 11(2):154– 157

    Article  Google Scholar 

  • Gracovetsky S, Newman N, Pawlowsky M, Lanzo V, Davey B, Robinson L (1995) A database for estimating normal spinal motion derived from noninvasive measurements. Spine 20(9):1036–1046

    Article  Google Scholar 

  • Gruber K, Ruder H, Denoth J, Schneider K (1998) A comparative study of impact dynamics: wobbling mass model versus rigid body models. J Biomech 31(5):439–444

    Article  Google Scholar 

  • Günther M, Sholukha VA, Keßler D, Wank V, Blickhan R (2003) Dealing with skin motion and wobbling masses in inverse dynamics. J Mech Med Biol 3(3/4):309–335

    Article  Google Scholar 

  • Günther M, Müller O, Blickhan R (2012) What does head movement tell about the minimum number of mechanical degrees of freedom in quiet human stance? Arch Appl Mech 82(3):333–344

    Article  Google Scholar 

  • Günther M, Ruder H (2003) Synthesis of two-dimensional human walking: a test of the lambda-model. Biol Cybern 89(2):89–106

    Article  MATH  Google Scholar 

  • Günther M, Schmitt S, Wank V (2007) High-frequency oscillations as a consequence of neglected serial damping in Hill-type muscle models. Biol Cybern 97(1):63–79

    Article  MATH  Google Scholar 

  • Haeufle D, Günther M, Bayer A, Schmitt S (2014) Hill-type muscle model with serial damping and eccentric force-velocity relation. J Biomech 47(6):1531–1536

    Article  Google Scholar 

  • Hancock MJ, Maher CG, Laslett M, Hay E, Koes B (2011) Discussion paper: What happened to the ‘bio’ in the bio-psycho-social model of low back pain? Eur Spine J 20(12):2105–2110

    Article  Google Scholar 

  • Hatze H (1976) The complete optimization of a human motion. Math Biosci 28(1):99–135

    Article  MathSciNet  MATH  Google Scholar 

  • Hatze H (1981) A comprehensive model for human motion simulation and its application to the take-off phase of the long jump. J Biomech 14(3):135–142

    Article  Google Scholar 

  • Hayes MA, Howard TC, Gruel CR, Kopta JA (1989) Roentgenographic evaluation of lumbar spine flexion-extension in asymptomatic individuals. Spine 14(3):327–331

    Article  Google Scholar 

  • Henze A (2002) Dreidimensionale biomechanische Modellierung und die Entwicklung eines Reglers zur Simulation zweibeinigen Gehens. Ph.D. thesis, Eberhard-Karls-Universität zu Tübingen

  • Hill A (1938) The heat of shortening and the dynamic constants of muscle. Proc R Soc Lond Ser B 126(843):136–195

  • Hirsch C, Nachemson A (1954) New observations on the mechanical behavior of lumbar discs. Acta Orthop 23(4):254–283

    Article  Google Scholar 

  • Huynh K, Gibson I, Lu W (2010) Simulating dynamics of thoracolumbar spine derived from LifeMOD under haptic forces. World Acad Sci Eng Technol 64:278–285

  • Ibarz E, Herrera A, Más Y, Rodríguez-Vela J, Cegoñino J, Puértolas S, Gracia L (2013) Development and kinematic verification of a finite element model for the lumbar spine: application to disc degeneration. BioMed Res Int 2013(4):1–18

    Article  Google Scholar 

  • Karajan N (2012) Multiphasic intervertebral disc mechanics: theory and application. Arch Comput Methods Eng 19:261–339

  • Karajan N, Otto D, Oladyshkin S, Ehlers W (2014) Application of the polynomial chaos expansion to approximate the homogenised response of the intervertebral disc. Biomech Model Mechanobiol 13(5):1065–1080

  • Karajan N, Röhrle O, Ehlers W, Schmitt S (2013) Linking continuous and discrete intervertebral disc models through homogenisation. Biomech Model Mechanobiol 12(3):453–466

    Article  Google Scholar 

  • Kistemaker DA, van Soest AJ, Bobbert MF (2007) Equilibrium point control cannot be refuted by experimental reconstruction of equilibrium point trajectories. J Neurophysiol 98(3):1075–1082

    Article  Google Scholar 

  • Kitazaki S, Griffin M (1997) A modal analysis of whole-body vertical vibration, using a finite element model of the human body. J Sound Vib 200(1):83–103

    Article  Google Scholar 

  • Lafortune M, Lake M, Henning E (1996) Differential shock transmission response of the human body to impact severity and lower limb posture. J Biomech 29(12):1531–1537

    Article  Google Scholar 

  • Langevin HM, Sherman KJ (2007) Pathophysiological model for chronic low back pain integrating connective tissue and nervous system mechanisms. Med Hypotheses 68(1):74–80

    Article  Google Scholar 

  • Lee S, Wong KWN, Chan M, Yeung H, Chiu JLF, Leong JCY (2002) Development and validation of a new technique for assessing lumbar spine motion. Spine 27(8):E215–E220

    Article  Google Scholar 

  • Legnani G, Casalo F, Righettini P, Zappa B (1996a) A homogeneous matrix approach to 3D kinematics and dynamics-II. Applications to chains of rigid bodies and serial manipulators. Mech Mach Theory 31(5):589–605

  • Legnani G, Casolo F, Righettini P, Zappa B (1996b) A homogeneous matrix approach to 3D kinematics and dynamics-I. Theory. Mech Mach Theory 31(5):573–587

  • Li G, Wang S, Passias P, Xia Q, Li G, Wood K (2009) Segmental in vivo vertebral motion during functional human lumbar spine activities. Eur Spine J 18(7):1013–1021

    Article  Google Scholar 

  • Macintosh JE, Bogduk N (1986) The biomechanics of the lumbar multifidus. Clin Biomech 1(4):205–213

    Article  Google Scholar 

  • Macintosh JE, Bogduk N (1987) Volvo award in basic science. The morphology of the lumbar erector spinae. Spine 12(7):658–668

    Article  Google Scholar 

  • Macintosh JE, Bogduk N (1991) The attachments of the lumbar erector spinae. Spine 16(7):783–792

    Article  Google Scholar 

  • Macintosh JE, Valencia F, Bogduk N, Munro R (1986) The morphology of the human lumbar multifidus. Clin Biomech 1(4):196–204

    Article  Google Scholar 

  • McGill SM, Norman RW (1987) Effects of an anatomically detailed erector spinae model on L4/L5 disc compression and shear. J Biomech 20(6):591–600

    Article  Google Scholar 

  • Millecamps M, Tajerian M, Naso L, Sage EH, Stone LS (2012) Lumbar intervertebral disc degeneration associated with axial and radiating low back pain in ageing SPARC-null mice. Pain 153(6):1167–1179

    Article  Google Scholar 

  • Miller JA, Schmatz C, Schultz AB (1988) Lumbar disc degeneration: correlation with age, sex, and spine level in 600 autopsy specimens. Spine 13(2):173–178

    Article  Google Scholar 

  • Monteiro NMB (2009) Analysis of the intervertebral discs adjacent to interbody fusion using a multibody and finite element co-simulation. Ph.D. thesis, Universidade Técnica de Lisboa

  • Mörl F, Siebert T, Schmitt S, Blickhan R, Günther M (2012) Electro-mechanical delay in Hill-type muscle models. J Mech Med Biol 12(5):85–102

    Article  Google Scholar 

  • Nachemson AL, Evans JH (1968) Some mechanical properties of the third human lumbar interlaminar ligament (ligamentum flavum). J Biomech 1(3):211–220

    Article  Google Scholar 

  • NASA (1978) Anthropometric source book. A handbook of anthropometric data (NASA RP-1024) vol 2. National Aeronautics and Space Administration, Texas (NTIS No. N79-13711)

  • Nigg B, Denoth J (1980) Belastung des menschlichen Bewegungsapparates aus der Sicht der Biomechanik. In: Nigg B, Denoth J (eds) Sportplatzbeläge, chapter 3. Juris Verlag, Zürich, pp 31–77

    Google Scholar 

  • Nigg BM, Herzog W (2007) Biomechanics of the musculo-skeletal system. Wiley, New York

    Google Scholar 

  • Otten E (2003) Inverse and forward dynamics: models of multi-body systems. Philos Trans R Soc B Biol Sci 358(1437):1493–1500

    Article  Google Scholar 

  • Pandy MG, Zajac FE, Sim E, Levine WS (1990) An optimal control model for maximum-height human jumping. J Biomech 23(12):1185–1198

    Article  Google Scholar 

  • Panjabi MM, Goel V, Oxland T, Takata K, Duranceau J, Krag, M, Price M (1992) Human lumbar vertebrae. Quantitative three-dimensional anatomy. Spine 17(3):299–306

  • Panjabi MM, Goel V, Takata K (1982) Physiologic strains in the lumbar spinal ligaments an in vitro biomechanical study 1981 Volvo award in biomechanics. Spine 7(3):192–203

    Article  Google Scholar 

  • Panjabi MM, Takata K, Goel V, Federico D, Oxland T, Duranceau J, Krag M (1991) Thoracic human vertebrae. Quantitative three-dimensional anatomy. Spine 16(8):888–901

    Article  Google Scholar 

  • Pearcy M, Portek I (1984) Three-dimensional X-ray analysis of normal movement in the lumbar spine. Spine 9(3):294–297

  • Pintar F, Yoganandan N, Myers T, Elhagediab A, Sances A (1992) Biomechanical properties of human lumbar spine ligaments. J Biomech 25(11):1351–1356

    Article  Google Scholar 

  • Polly DW, Kilkelly FX, McHale KA, Asplund LM, Mulligan M, Chang AS (1996) Measurement of lumbar lordosis. Evaluation of intraobserver, interobserver, and technique variability. Spine 21(13):1530–1535 discussion 1535–6

    Article  Google Scholar 

  • Robertson DJ, Von Forell GA, Alsup J, Bowden AE (2013) Thoracolumbar spinal ligaments exhibit negative and transverse pre-strain. J Mech Behav Biomed Mater 23:44–52

    Article  Google Scholar 

  • Rohlmann A, Nabil Boustani H, Bergmann G, Zander T (2010) Effect of a pedicle-screw-based motion preservation system on lumbar spine biomechanics: a probabilistic finite element study with subsequent sensitivity analysis. J Biomech 43(15):2963–2969

    Article  Google Scholar 

  • Rohlmann A, Zander T, Rao M, Bergmann G (2009) Realistic loading conditions for upper body bending. J Biomech 42(7):884–890

    Article  Google Scholar 

  • Rosatelli AL, Ravichandiran K, Agur AM (2008) Three-dimensional study of the musculotendinous architecture of lumbar multifidus and its functional implications. Clin Anat 21(6):539–546

    Article  Google Scholar 

  • Sarzi-Puttini P, Atzeni F, Fumagalli M, Capsoni F, Carrabba M (2004) Osteoarthritis of the spine. Semin Arthritis Rheum 34(6):38–43

    Article  Google Scholar 

  • Schmidt H, Kettler A, Rohlmann A, Claes L, Wilke H-J (2007) The risk of disc prolapses with complex loading in different degrees of disc degeneration—a finite element analysis. Clin Biomech 22(9):988–998

    Article  Google Scholar 

  • Schmidt H, Shirazi-Adl A, Galbusera F, Wilke H-J (2010) Response analysis of the lumbar spine during regular daily activities—a finite element analysis. J Biomech 43(10):1849–1856

    Article  Google Scholar 

  • Schmitt S (2003) Modellierung und Simulation biomechanischer Prozesse am Beispiel des Skisprungs. Diploma thesis, Universität Stuttgart

  • Schmitt S, Haeufle DFB, Blickhan R, Günther M (2012) Nature as an engineer: one simple concept of a bio-inspired functional artificial muscle. Bioinspir Biomim 7(3):036022

    Article  Google Scholar 

  • Schmitt S, Günther M (2011) Human leg impact: energy dissipation of wobbling masses. Arch Appl Mech 81(7):887–897

    Article  MATH  Google Scholar 

  • Schultz A, Warwick DN, Berkson MH, Nachemson AL (1979) Mechanical-properties of human lumbar spine motion segments Part 1: responses in flexion, extension, lateral bending, and torsion. J Biomech Eng Trans ASME 101(1):46–52

    Article  Google Scholar 

  • Siebert T, Till O, Blickhan R (2014a) Work partitioning of transversally loaded muscle: experimentation and simulation. Comput Methods Biomech Biomed Eng 17(3):217–229

    Article  Google Scholar 

  • Siebert T, Till O, Stutzig N, Günther M, Blickhan R (2014b) Muscle force depends on the amount of transversal muscle loading. J Biomech 47(8):1822–1828

    Article  Google Scholar 

  • Soest AJKv, Haenen WP, Rozendaal LA (2003) Stability of bipedal stance: the contribution of cocontraction and spindle feedback. Biol Cybern 88(4):293–301

  • Stokes IA, Gardner-Morse M (1995) Lumbar spine maximum efforts and muscle recruitment patterns predicted by a model with multijoint muscles and joints with stiffness. J Biomech 28(2):173–186

    Article  Google Scholar 

  • Stokes IA, Gardner-Morse M (1999) Quantitative anatomy of the lumbar musculature. J Biomech 32(3):311–316

    Article  Google Scholar 

  • Van Soest AJ, Bobbert MF (1993) The contribution of muscle properties in the control of explosive movements. Biol Cybern 69(3):195–204

  • Virgin WJ (1951) Experimental investigations into the physical properties of the intervertebral disc. J Bone Joint Surg Br 33–B(4):607–611

    Google Scholar 

  • Waddell G (1987) Volvo award in clinical sciences. A new clinical model for the treatment of low-back pain. Spine 12(7):632–644

    Article  Google Scholar 

  • Whatley BR, Wen X (2012) Intervertebral disc (IVD): structure, degeneration, repair and regeneration. Mater Sci Eng C 32(2):61–77

    Article  Google Scholar 

  • White AA, Panjabi MM (1990) Clinical biomechanics of the spine. Lippincott Williams & Wilkins, Philadelphia

  • Wilke H, Neef P, Caimi M, Hoogland T, Claes L (1999) New in vivo measurements of pressures in the intervertebral disc in daily life. Spine 24(8):755

  • Wilke H, Neef P, Hinz B, Seidel H, Claes L (2001) Intradiscal pressure together with anthropometric data-a data set for the validation of models. Clin Biomech 16(Suppl 1):S111–S126

    Article  Google Scholar 

  • Wilkenfeld AJ, Audu ML, Triolo RJ (2006) Feasibility of functional electrical stimulation for control of seated posture after spinal cord injury: a simulation study. J Rehabil Res Dev 43(2):139–152

    Article  Google Scholar 

  • Wong KWN, Leong JCY, Chan M-K, Luk KDK, Lu WW (2004) The flexion-extension profile of lumbar spine in 100 healthy volunteers. Spine 29(15):1636–1641

    Article  Google Scholar 

  • Wong KWN, Luk KDK, Leong JCY, Wong SF, Wong KKY (2006) Continuous dynamic spinal motion analysis. Spine 31(4):414–419

    Article  Google Scholar 

Download references

Acknowledgments

The authors would like to thank the German Research Foundation (DFG) for financial support of the project within the Cluster of Excellence in Simulation Technology (EXC 310/1) at the University of Stuttgart.

T.R. and S.S. have received funding from the (European Union) Seventh Framework Programme (FP7/2007–2013) under Grant Agreement No. 246994.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to S. Schmitt.

Appendices

Appendix 1: Generated anthropometric data set: vertebral position, orientation, and dimension

An average data set of vertebral positions was calculated from the literature data (Ashton-Miller and Schultz 1997; Belytschko 1978; El-Rich and Shirazi-Adl 2004; Kitazaki and Griffin 1997) to be used by calcman. They determined global orientation and position of vertebrae using skin markers. The vertebral position is defined as the position of the geometric centre of each vertebral body with respect to the global reference frame. S1 was positioned in the centre of the pelvis. Position data from each literature data set was normalised in length to one reference trunk height, i.e. each data set was normalised to the same caudocranial distance (\(z\)-value) of T1-L5, \(h_{\text {T1-L5,Avg}}=458.1\hbox {mm} \approx h_{\text {T1-L5,Belytschko}}\).

Then, a polynomial fit was estimated for each data set (see Fig. 9, \(f_{\text {Ashton}}\), \(f_{\text {Belytschko}}\), \(f_{\text {Rich}}\) and \(f_{\text {Kitazaki}}\)). These four fits for \(x=f(z)\) were averaged by calculating the mean coefficients of each polynomial. The resulting function (sixth-order polynomial) writes

$$\begin{aligned} f_{\text {Avg}}(z)&= 0.015 \,\text {mm} \!+\!0.201 \!\times \! z\!-\!1.12 \times 10^{-3} \,\text {mm}^{-1} \times z^2\nonumber \\&-1.2 \!\times \! 10^{-5} \,\text {mm}^{-2} \!\times \! z^3 \!+\!7.0 \!\times \! 10^{-8} \,\text {mm}^{-3} \times z^4\nonumber \\&-1.5 \times 10^{-10} \,\text {mm}^{-4} \times z^5\nonumber \\&+1.0 \!\times \! 10^{-13} \,\text {mm}^{-5} \times z^6. \end{aligned}$$
(2)

The average position data on the longitudinal axis including standard deviation and the basis data sets are presented in Fig. 9. Calcman calculates the subject-derived height (\(h_{\text {T1-L5}}\)) and, with that, scales the z-position of each vertebra by the ratio \(h_{\text {T1-L5}}/h_{\text {T1-L5,Avg}}\) while leaving the \(x\)-position unscaled (Table 1).

Fig. 9
figure 9

Anthropometric data set functions fitted to the position data of vertebral bodies (centre) in healthy male humans. All polynomial fits \(f_{\text {Ashton}}(z)\) (Ashton-Miller and Schultz 1997), \(f_{\text {Belytschko}}(z)\) (Belytschko 1978), \(f_{\text {Rich}}(z)\) (El-Rich and Shirazi-Adl 2004), and \(f_{\text {Kitazaki}}(z)\) (Kitazaki and Griffin 1997) were scaled to the reference height of \(h_{\text {T1-L5,Avg}}=458.1\) mm from T1 to L5. \(f_{\text {Avg}}(z)\) depicts the mean data set used for scaling in calcman, the thin grey lines capture its standard deviation \(f_{\text {Avg}}(z)+/-\sigma \). Plus sign indicates the position of each vertebral centre L5 to T1. The origin is equal to the position point of L5. The sagittal axis from dorsal to ventral is equal to the \(x\)-axis, while the longitudinal axis from caudal to cranial is equal to the \(z\)-axis

Each vertebral orientation was defined as the normal of the curve (Eq. 2) at its position aligning with the anterior–posterior direction of the vertebra. Each vertebra was modelled as a cylinder (see below), although real vertebrae are slightly wedge-shaped. The calculated orientations were compared with the experimental data from the literature (Ashton-Miller and Schultz 1997; Campbell-Kyureghyan et al. 2005), which are usually presented as upper endplate orientations in the sagittal plane. For this, we fitted graphic primitives from computed tomography (CT) data (Anatomium (TM), 21st Century Solutions Ltd/Gibraltar) that represent T1 to S1, to their respective cylinder primitives and estimated the upper endplate orientation (Table 1). All three data sets vary locally in the range of a few degrees, which is well represented by their standard deviations.

Vertebral dimensions, i.e. vertebral height, depth, and width, were determined for the entire spine, i.e. C3 to S1. The study by Gilad and Nissan (1986) was used as a basis. However, it only includes cervical and lumbar dimensions. Missing data were determined by parameter scaling of data sets from Panjabi et al. (1991, 1992), White and Panjabi (1990). For example, depths for the whole spine included in Panjabi et al. (1992, 1991), White and Panjabi (1990) were scaled to the basis depths in Gilad and Nissan (1986) in order to determine missing vertebral depths in the thoracic spine.

The cylinder’s height was set to the maximum of dorsal and ventral height (Eq. 3 \(_{5}\)), its radius to the maximum of depth and width (Eq. 3 \(_6\)). This allows an easy inclusion of the mass of the spinous and transverse processes. It is assumed that their mass approximately balances the overestimation by using the maximum height and ignoring the wedge shape of each vertebral body. The vertebral mass (Eq. 3 \(_3\)) was calculated on the basis of each cylinder’s volume (Eq. 3 \(_4\)) and the lumbar vertebrae bone density \(\rho _{\text {lumbar}} = 600{-}1{,}000\,\text {kg/m}3\) (Nigg and Herzog 2007). The moment of inertia tensor follows from the cylinder’s parameters (see Eqs. (3 \()_{1,2}\)):

$$\begin{aligned} I_{\text {cyl,xx}} = I_{\text {cyl,yy}}&= 1/12 \times m_{\text {vert,i}} \times h_{\text {vert,i}}^2+ 1/4\nonumber \\&\times \,\, m_{\text {vert,i}} \times r_{\text {vert,i}}^2, \nonumber \\ I_{\text {cyl,zz}}&= 1/2 \times m_{\text {vert,i}} \times r_{\text {vert,i}}^2 \nonumber \\ \text{ with }\qquad m_{\text {vert,i}}&= V_{\text {vert,i}} \times \rho _{\text {lumbar}} \times 3/2, \nonumber \\ V_{\text {vert,i}}&= \pi \times r_{\text {vert,i,width}}^{2} \times h_{\text {vert,i}}, \nonumber \\ h_{\text {vert,i}}&= \max \{h_{\text {vert,i,dorsal}},h_{\text {vert,i,ventral}}\}, \nonumber \\ r_{\text {vert,i}}&= \max \{r_{\text {vert,i,depth}},r_{\text {vert,i,width}}\}. \end{aligned}$$
(3)

Appendix 2: Nonlinear ligament model

Various studies on ligaments confirm a nonlinear stress–strain characteristic for ligaments (Chazal et al. 1985; Pintar et al. 1992). For our lumbar spine model, the spinal ligaments were modelled as nonlinear straight-line elements based on the experimental results by Chazal et al. (1985), Panjabi et al. (1982). The function describing the spinal ligament’s stress–strain characteristics was assumed to be continuously differentiable (\(C^1\) continuity) between an initial nonlinear region (up to A) and a constant stiffness at higher loads (A-B, see Fig. 10). The concave region at even higher forces is not shown here. Hence, the force–length characteristic of a spinal ligament is modelled as

$$\begin{aligned}&F_{\text {el,LIG}}(l_{\text {LIG}})\nonumber \\&\quad = \left\{ \begin{array}{l@{\quad }l} 0 &{} \, l_{\text {LIG}} < l_{\text {LIG,0}} \\ K_{\text {LIG,nl}} \, (l_{\text {LIG}}-l_{\text {LIG,0}})^{\nu _{\text {LIG,nll}}} &{}\, l_{\text {LIG}} < l_{\text {LIG,nll}} \\ \Delta F_{\text {LIG,0}} + K_{\text {LIG,l}} \,(l_{\text {LIG}} - l_{\text {LIG,nll}}) &{} \, l_{\text {LIG}} \ge l_{\text {LIG,nll}} \end{array} \right. \!, \nonumber \\ \end{aligned}$$
(4)

where \(l_{\text {LIG}}\) is its length. Its parameters are

$$\begin{aligned} l_{\text {LIG,nll}}&= (1+\Delta U_{\text {LIG,nll}}) \, l_{\text {LIG,0}}, \nonumber \\ \nu _{\text {LIG,nll}}&= \Delta U_{\text {LIG,nll}} / \Delta U_{\text {LIG,l}}, \nonumber \\ K_{\text {LIG,nl}}&= \Delta F_{\text {LIG,0}}/(\Delta U_{\text {LIG,nll}}\,l_{\text {LIG,0}})^{\nu _{\text {LIG,nll}}}, \nonumber \\ K_{\text {LIG,l}}&= \Delta F_{\text {LIG,0}}/(\Delta U_{\text {LIG,l}}\,l_{\text {LIG,0}}) . \end{aligned}$$
(5)
Fig. 10
figure 10

Force–length relation of the spinal ligament. Four independent parameters were used such as \(l_{\text {LIG,0}}\) (rest length), \(\Delta U_{\text {LIG,nll}}\) (relative stretch at nonlinear–linear transition), \(\Delta F_{\text {LIG,0}}\) (both force at the transition and force increase in the linear part), and \(\Delta U_{\text {LIG,l}}\) (relative additional stretch in the linear part providing a force increase in \(\Delta F_{\text {LIG,0}}\)). The transition between the initial nonlinear and the linear region is located at \(l_{\text {LIG,nll}} = (1+\Delta U_{\text {LIG,nll}})\), the exponent for the initial nonlinear relation is \(\nu _{\text {LIG,nll}} = \Delta U_{\text {LIG,nll}} / \Delta U_{\text {LIG,l}}\). The parameters of relevant and included spinal ligaments are given in Table 5. They were derived from experimental data (Chazal et al. 1985)

They were derived from four independent parameters such as \(l_{\text {LIG,0}}\) (rest length of ligament), \(\Delta U_{\text {LIG,nll}}\) (relative stretch at nonlinear–linear transition), \(\Delta F_{\text {LIG,0}}\) (both force at transition and force increase in linear part), and \(\Delta U_{\text {LIG,l}}\) (relative additional stretch in linear part providing the force increase \(\Delta F_{\text {LIG,0}}\)), cf.  Fig. 10. These parameters were partly extracted from tensile experiments on 43 human spinal ligaments performed by Chazal et al. (1985). For five spinal ligaments, i.e. anterior longitudinal (ALL) and posterior longitudinal ligament (PLL), ligamentum flavum (LF), supra-spinal (SSL), and interspinal (ISL) ligament, Chazal et al. (1985) estimated tension and elongation at transition (point A), at the end of the linear phase (point B) and at the apex of the tension elongation curve. The arithmetic means of the extracted force and elongation data at A and B (\(F_{\text {A}}\), \(l_{\text {A}}\), \(F_{\text {B}}\), \(l_{\text {B}}\)) were used to calculate three input parameters of the ligament characteristic (\(\Delta U_{\text {LIG,nll}}\), \(\Delta U_{\text {LIG,l}}\), \(\Delta F_{\text {LIG,0}}\)). These input parameters were determined separately for each of the five ligaments.

The ligament’s rest length (\(l_{\text {LIG,0}}\)) was extracted from the model’s geometry. It is accepted that spinal ligaments are not in a strain-free state when the spine is in its neutral anatomical position during upright stance. Pre-strain data for the spinal ligaments were taken from the literature (Aspden 1992; Nachemson and Evans 1968; Robertson et al. 2013) and used to determine the ligament’s rest length in the neutral position. The neutral position was approximated by simulating an initial settling of the lumbar spine model (see 2.6, neutral position). In that state, each ligament’s rest length was calculated by the distance between two ligament’s attachment points (\(l_{\text {LIG,SET}}\) [m]) and the respective pre-strain (\(\epsilon _{\text {LIG}}\) [ ]): \(l_{\text {LIG,0}}=l_{\text {LIG,SET}}/(1+\epsilon _{\text {LIG}})\).

As proposed by Panjabi et al. (1982), ligaments with planar attachment areas, like the ALL, LF, and ISL, can be approximated by more than one pair of attachment points, i.e. multiple lines for one ligament. Their characteristic (Eq. 4) is then equally distributed by equally distributing \(\Delta F_{\text {LIG,0}}\) between multiple lines.

A velocity-dependent damper was also included in order to account for the ligament’s nonlinear viscoelastic behaviour, i.e.

$$\begin{aligned}&F_{\text {damp,LIG}}= F_{\text {el,LIG}} \times d_{\text {LIG}} \times v_{\text {LIG}} \;\; \text {with} \nonumber \\&\text {with} \;\; \;\; v_{\text {LIG}}=\frac{{\mathrm {dl}}_{\text {LIG}(t)}}{{\mathrm {dt}}}, \;\; d_{\text {LIG}} = 1.0 \,\text {s}\,\text {m}^{-1} \nonumber \\&\text {and consequently} \;\; \;\; F_{\text {tot,LIG}} = F_{\text {el,LIG}} \times (1 + d_{\text {LIG}} \times v_{\text {LIG}})\mathrm {.}\nonumber \\ \end{aligned}$$
(6)

According to Gerritsen et al. (1995), the damping force depends on the elastic force \(F_{\text {el,LIG}}\) and is scaled by a damping coefficient \(d_{\text {LIG}}\). This avoids discontinuity and numerical stabilities.

Appendix 3: Ligament parameters

See Table 5.

Table 5 Ligament modelling parameters for the nonlinear MB model version, i.e. \(l_{\text {LIG,0}}\) (rest length of ligament), \(\Delta U_{\text {LIG,nll}}\) (relative stretch at nonlinear–linear transition), \(\Delta F_{\text {LIG,0}}\) (both force at transition and force increase in linear part), \(\Delta U_{\text {LIG,l}}\) (relative additional stretch in linear part providing the force increase \(\Delta F_{\text {LIG,0}}\)) and \(\epsilon \) (pre-strain of ligament), for all included spinal ligaments, i.e. anterior longitudinal (ALL) and posterior longitudinal ligament (PLL), ligamentum flavum (LF), supraspinal (SSL), and interspinal (ISL) ligament

Appendix 4: Development of specific IVD model

In MB models, IVDs are either modelled as one force element or one force element is used to capture the mechanical response of the IVD as well as its surrounding ligaments and muscles, i.e. lumped modelling approach for one functional spinal unit. The mechanical response of an IVD (Huynh et al. 2010) or a functional spinal unit (Christophy et al. 2011; Monteiro 2009) is usually modelled using a linear and decoupled bushing element (6-DOF spring-dashpot element) as a connecting force element. Stiffness and damping parameters are taken from experimental measurements (Ashton-Miller and Schultz 1997; Berkson et al. 1979; Eberlein et al. 2004; Gardner-Morse and Stokes 2004; Hirsch and Nachemson 1954; Schultz et al. 1979; Virgin 1951). Experimental data clearly demonstrate the nonlinear nature of these biological tissues. Hence, it is linearised to be used as model parameters for the linear elements in other MB models (Huynh et al. 2010; Monteiro 2009). So far, model parameters differ, as experimental results cover a wider range.

Instead, a recently developed IVD model was used in our MB model and implemented in demoa. It captures the nonlinear and coupled mechanical response of IVDs in the lumbar region from pelvis to vertebra L1 (Karajan et al. 2013). In particular, a FE model of the IVD (Karajan 2012) was used to pre-compute the mechanical behaviour of the IVD in terms of stresses and strains. The FE model of the IVD is based on the theory of porous media. It is capable of capturing the porous microstructure of the IVD. The mechanical response of the IVD due to several applied deformation states was captured at its centre of gravity by homogenisation via an integration of the resulting stresses at the connecting surface of the adjacent vertebra. Thus, each of the homogenised results represents a reaction force and moment due to an applied deformation state (combination of displacement and rotation). A parameterisation using a cubic polynomial turned out to deliver the best fit of the resulting forces and moments gathered at the varying deformation states, i. e. sampling points, which can be related to the displacements and rotations of the MB system defined in demoa. For more detailed information on the representation, the reader is referred to Karajan et al. (2013).

Appendix 5: Hill-type muscles in MB models

In forward dynamics MB models, muscles are often modelled as Hill-type MTCs (Haeufle et al. 2014). This labelling refers to A.V. Hills’s experimental finding (Hill 1938) that the force–velocity relation of a muscles fibre is a hyperbola. Thus, Hill-type models are phenomenologically based macroscopic lumped parameter models. Such a model usually consists of a contractile element (CE), a serial, and a parallel elastic element. The corresponding contraction (Eq. 7 \(_1\), CE length \(l_{\text {CE}}),\) and activation dynamics (Eq. 7 \(_2\), activity \(q\) with \(0\le q\le 1\)) of a muscle are described by first-order differential equations:

$$\begin{aligned}&\dot{l}_{\text {CE}}=f_{l}(l_{\text {CE}},l_{\text {MTC}},q),\nonumber \\&\dot{q}=f_{\text {q}}(q,\hbox {STIM}) \end{aligned}$$
(7)

with \(l_{\text {MTC}}\): length of MTC, \(\hbox {STIM}\) stimulation input. \(\hbox {STIM}\) is calculated by our simple lambda-control model, a bio-inspired motor control model. It represents the electrical stimulation signal (summed transmembrane potential) on the muscles’ surface and can be compared with surface electromyography (sEMG). Günther et al. (2007) included an additional damping component parallel to the serial elastic element. Hence, the right-hand side of Eq. (7)\(_1\) also depends on \(\dot{l}_{\text {MTC}}\). By using this modified Hill-type muscle model, movement control is expected to be more feasible, because eigenoscillations are damped due to the interaction of muscle and segment inertia (Günther et al. 2007). We refer to Günther et al. (2007), Haeufle et al. (2014), Mörl et al. (2012) for a detailed description of the muscle model. It was implemented in demoa and used for all MTCs in the presented full human MB model.

Appendix 6: Muscle parameters

See Table  6.

Table 6 Muscle modelling parameters: maximum isometric force of the contractile element \(F_{\text {max}}\), optimal fibre length \(l_{\text {CE,opt}}\), and serial elastic element \(l_{\text {SEE,0}}\)

Appendix 7: Comparison with the literature data

Kinematic data from the literature cannot be compared directly with results from experimental studies because investigated overall movements differ. Some studies present flexion–extension results in a combined way, others present data, measured with less or more flexion amplitudes. Therefore, relative changes and the order of magnitude were used for comparison. In the following, vertical translational and angular displacements, and IVD forces of the presented lumbar spine model are compared with the literature data. Additionally, the resultant IVD forces of the nonlinear intervertebral disc model included into the spine model are compared with results of three lumbar spine model variations including the classical, linear intervertebral disc model for three different parameter sets. The resultant IVD forces again are compared with the literature data (Fig. 11).

Fig. 11
figure 11

A Vertical displacements in the local sagittal plane (along the local longitudinal axis (\(z\): caudocranial direction)) of the subjacent body and B angular displacements of intervertebral discs \(\varphi _{\text {y}}\) in the local sagittal plane of the subjacent body (its local \(x\)\(z\)-plane) as a function of the negative lumbar lordosis angle. Vertical and rotational displacements are depicted relative to displacement at \(t_{\text {SET}}\), i.e. \(r_{\text {z},t_{\text {SET}}} = 0\) m and \(\Delta \varphi _{\text {y},t_{\text {SET}}} = 0 ^\circ \). Bold lines are used to emphasise translational and rotational displacements in L4/5

Vertical displacements in the IVDs increase from cranial to caudal (top to bottom, see Fig. 12). For a simple flexion movement simulated with our full human body model, the displacement amplitudes increase from 0.34 mm (L1/2), to 0.45 mm (L2/3), to 0.58 mm (L3/4), to 0.62 mm (L4/5). In vitro studies by Hayes et al. (1989) report displacements for a complete flexion–extension movement in the same order: L1/2 (1.9 mm) \(\rightarrow \) L2/3 (2.4 mm) \(\rightarrow \) L3/4 (2.5 mm) \(\rightarrow \) L4/5 (3.0 mm). The same way, Li et al. (2009) report displacements for a flexion–extension movement in the order: L2/3 (\(0.2\pm 0.2\) mm) \(\rightarrow \) L3/4 (\(0.6\pm 0.4\) mm) \(\rightarrow \) L4/5 (\(0.7\pm 0.6\) mm). Additionally, the order of magnitude of all three studies lies in the same range.

Fig. 12
figure 12

Vertical displacements in the lumbar joints, w.r.t local joint coordinate system, increase from cranial to caudal. In vitro studies (Hayes et al. 1989; Li et al. 2009) report displacements in the same order; however, they measured flexion–extension movements of higher amplitudes. Additional lines depict trend of values

More experimental data exist for the comparison of angular displacements in the IVDs. Ibarz et al. (2013), White and Panjabi (1990) present various in vivo measurements of angular joint motion during flexion–extension movement. Interestingly, data presented in the literature do not match, i.e. some present increasing angular motion from cranial to caudal (Hayes et al. 1989; Ibarz et al. 2013; Pearcy and Portek 1984; White and Panjabi 1990) while others present the opposite (Lee et al. 2002; Li et al. 2009; Wong et al. 2004, 2006). The simulation results that were presented in this study showed an increase in angular displacement during flexion in the order L1/2 (\(2.6^{\circ }\)) \(\rightarrow \) L2/3 (\(3.6^{\circ }\)) \(\rightarrow \) L3/4 (\(4.4^{\circ }\)) \(\rightarrow \) L4/5 (\(4.6^{\circ }\)). This caudocranial increase in angular IVD joint motion is in accordance with flexion–extension studies by Hayes et al. (1989), Ibarz et al. (2013), Pearcy and Portek (1984), White and Panjabi (1990): L1/2 (\(5-16^{\circ }\)) \(\rightarrow \) L2/3 (\(7.8-18^{\circ }\)) \(\rightarrow \) L3/4 (\(6-17^{\circ }\)) \(\rightarrow \) L4/5 (\(9-21^{\circ }\)). Again, only relative changes could be presented (see Fig. 13).

Fig. 13
figure 13

Angular displacements in the lumbar joints, w.r.t local joint coordinate system, increase from cranial to caudal. In vitro studies (Hayes et al. 1989; Ibarz et al. 2013; Pearcy and Portek 1984; White and Panjabi 1990) report displacements in the same order; however, they measured flexion–extension movements of higher amplitudes. Note, we did not include studies (Lee et al. 2002; Li et al. 2009; Wong et al. 2004, 2006) with the opposite trend in this figure. Additional lines depict trend of values

Additionally, we compared our angular displacements with the intervertebral flexion–extension (IVFE) presented by Lee et al. (2002), Wong et al. (2004, 2006). It indicates the relationship between net lumbar motion and motion of the individual vertebrae. The IVFE is often used as input data for modelling of the IVD element in inverse dynamic models, cf. slope \(k\) in Christophy et al. (2011). Our presented joint kinematics plotted against the lumbar lordosis angle are equivalent to the IVFE. Our simulation results for the IVFE also show a linear-like pattern, like experiments determining the IVFE characteristics. However, simulated IVFE in all joints of our model increase caudocranial, whereas in experiments IVFE slopes increase with investigated joints moving cranial. This is due to the fact that so far the new IVD model (Karajan et al. 2013) was used for all lumbar IVDs with the same parameters. The new, nonlinear IVD model was developed for the L4/5 intervertebral disc and then applied to all other IVDs. We argue that by applying different scaled material properties and an enhanced IVD element’s geometry, a more realistic IVFE pattern would occur. Additionally, it will furthermore result in a more physiological angular displacement behaviour. An enhanced nonlinear IVD model accounting for the mentioned shortcomings is already available, but was not ready to be included in this study yet (Karajan et al. 2014).

In vivo measurements of lumbar joint dynamics are rare. For a first validation, we used intradiscal pressure measurement data of the IVD L4/5. Measurements were taken with one healthy male test subject in different postures by Wilke et al. (1999, 2001). They measured a disc pressure of 0.43–0.5 MPa during relaxed standing of the test subject (body mass 70 kg, body height 1.74 m). In a \(36^\circ \) flexed standing posture, the disc pressure increased almost linearly to 1.08 MPa. The flexion angle was defined differently, i.e. between the thoracolumbar junction and the sacrum (\(\alpha _{\text {fl,S1--T12}}\)). The disc area, measured using magnetic resonance imaging, was 18 cm\(^2\). Thus, the respective axial joint force was 774–900 N in relaxed standing and 1,944 N in a flexed position. In our model, axial forces increased from \(F_{\text {z,upright}}=942\) N during upright stance to \(F_{\text {z,flex}}=1{,}681\) N at \(t_{\text {sim,end-nl}}\) with \(\alpha _{\text {fl,S1--T12}}=3^\circ \) and \(\alpha _{\text {fl,S1--T12}}=33^\circ \), respectively. Taking into account our slightly different flexion angle (higher at the beginning, smaller at the end), simulated longitudinal forces using the nonlinear IVD model in L4/5 show good agreement with experimental data.

In comparison with in vivo measurements of joint dynamics (Wilke et al. 1999, 2001), our new, nonlinear IVD model included in the spine model shows the best agreement (see 3.1). The classical, linear IVD model variations using either Monteiro parameters or Gardner–Morse parametres result in higher forces while standing and lower forces when reaching a flexed position. Restoring forces during stance in the Huynh model show good agreement with our model results, while forces in flexion are underestimated.

Appendix 8: A simulation of heel impact with the ground

We simulated a double heel contact of the model with the ground after a free fall of about 8 mm, i.e. after about 0.04 s, starting from a completely symmetrical initial position. To insure ball contacts would not interfere with the impact, the ankle joints were fixed in a slightly dorsally flexed orientation as bracket joints, the latter being a method to set the number of degrees of freedom in a joint to zero. Just alike, we fixed the upper body to L1, resulting in a lumped L1-plus upper body. Due to the stiffness \(4\times 10^{5}\) N/m of the linear elastic spring modelling the normal ground reaction force component of each heel contact element, both heel elements were deformed about 2 mm in vertical direction during the impact. The elements’ vertical damping force was modelled in proportion to the product of vertical deformation and velocity, with the corresponding coefficient of proportionality set to \(4\times 10^{5}\) Ns/m\(^2\). Correspondingly, the net vertical ground reaction force reached a peak of about 65 m/s\(^2\) (Fig. 14).

The mass distribution in the model was chosen such that all segment masses represented just the bony parts. This is because it is only the skeleton that is decelerated during this impact period of about 20.0 ms (Denoth et al. 1985; Nigg and Denoth 1980), whereas all soft tissue material like the muscles is coupled viscoelastically to the bones (Denoth 1985; Gruber et al. 1998; Günther et al. 2003; Schmitt and Günther 2011) and decelerated with a time delay. In modelling its mechanical response, the soft tissue material is denoted ‘wobbling mass’ (Denoth 1985). The ratio of wobbling to segment mass was set to zero in the feet, to \(1/2\) in the shanks, and to \(2/3\) in the thighs (Günther et al. 2003; Schmitt and Günther 2011). For the pelvis’ bony part, we assumed 3.0 kg, neglecting an assumed wobbling mass portion of 5.3 kg. The bony masses of the 24 vertebrae were assumed to decrease from about 50 g in L5 to 10 g in C1. In our model, the inertia of all vertebrae from T12 up to C1 is included in the upper body. For the upper body, we applied the ‘effective mass’ concept introduced by Denoth (1986), Denoth et al. (1985): during the 20.0 ms impact period, the effectively decelerated mass may depend on both net bony mass of the body and the initial joint angular configuration. We estimated that one-third of the 36.8 kg act effectively during impact. Altogether, the decelerated body masses summed to 25.0 kg in our impact model.

The resulting distribution of vertical acceleration components measured in the global reference frame is plotted in Fig. 14. There is a shock wave propagating from the feet to the head. The acceleration amplitudes decrease from distal to proximal in the legs, with a slight deviation from this rule: the foot acceleration is a little lower than the shank acceleration. This fact is due to the foot centre of mass being located about 10 cm anterior to the ankle and its upward movement determined by a superposition of heel translation and foot–shank rotation. Above the pelvis, the amplitudes decrease along the lumbar spine from caudal to cranial. Quantitatively, the acceleration maxima are time-shifted according to the following segmental sequence: shank(s) (0.06 ms), thigh(s) (0.02 ms), pelvis (\(-\)0.01 ms) L5 (0.08 ms) L4 (0.26 ms) L3 (0.18 ms), whole body (0.22 ms) L2 (8.73 ms) L1-plus upper body. The L1-plus upper body mass is the highest of all segments and located most distantly from the ground. Its maximum acceleration occurs clearly later than in any other segment including the whole body.

Fig. 14
figure 14

Vertical components of the model’s net ground reaction force and each segment’s centre of mass acceleration during an impact of the upright model on the heels from 8 mm falling height

These numbers can be compared to known experimental data of shock wave propagation from the ankle to the head (Lafortune et al. 1996). For this distance, these authors found propagation times between 5.9 and 4.7 ms, for knee angles between full extension and \(40^{\circ }\) flexion, respectively. Our simulation results demonstrate that the propagation along the chain of the four distal bodies connected by hard joint constraints (foot, shank, thigh, pelvis) is faster than the propagation via viscoelastically deforming soft constraints (like IVDs in our case). Therefore, we would expect that summing up the time shifts of the \(24\) IVDs will dominate the shock wave propagation time from heel to head, because physiologically there are only four additional distal cartilage connections from heel to pelvis: lower and upper ankle joints, knee, and hip. To estimate the propagation velocity depending on IVD properties, we first calculate the mean time shift per IVD for the four IVDs located between pelvis and L2 as \(\frac{-0.01+0.08+0.26+0.40\,\hbox {ms}}{4}=\frac{0.73\,\hbox {ms}}{4}\). For the propagation time along the vertebral column comprising 24 IVDs, we found \(24\times \frac{0.73\,\hbox {ms}}{4}\approx 4.4\,\hbox {ms}\). The deviation from the data by Lafortune et al. (1996) can be easily explained by two delaying contributions missing in our model so far: (1) those by the cartilage in the leg joints and (2) those by acoustic wave propagation in the bones. However, the latter effects should be negligible because the elastic modulus of bones is about 4 orders of magnitude higher than that of IVDs, with approximately the same mass density.

The distance between L5 and C1 is roughly 60 cm. Thus, we can estimate the propagation velocity from our computer experiment to be \(c_{\mathrm{long}}=60\,\hbox {cm}/4.4\,\hbox {ms}=136\,\hbox {m}/\hbox {s}\). Based on theory, we may predict this number based on elastic modulus \(E=\frac{\sigma }{\varepsilon }=\frac{\Delta \,F}{\Delta \,l}\times \frac{A_{0}}{l_{0}}\) with the stress \(\sigma =\frac{\Delta \,F}{A_{0}}\), the strain \(\varepsilon =\frac{\Delta \,l}{l_{0}}\), and the stiffness \(K=\frac{\Delta \,F}{\Delta \,l}\). The area of a vertebra’s upper endplate is roughly 1 cm\(^{2}\) and the range of IVD thicknesses is about 1 cm\(\ldots \)0.5 cm from L5 to C1. In general, wave propagation velocity \(c_{\mathrm{long}}\) in solids depends also on the Poisson ratio \(\mu =-\frac{\Delta \,A}{A_{0}}/\frac{\Delta \,l}{l_{0}}\). For longitudinal waves, theory predicts \(c_{\mathrm{long}}=\sqrt{\frac{E\times (1-\mu )}{\rho \times (1-\mu -\mu ^{2})}}\), where \(\rho \) symbolises the mass density. With a lumbar IVD stiffness estimation of \(5\times 10^{5}\) N/m (compare Table 2), lumbar thickness of 1 cm, and a small Poisson ratio \(\mu \ll 0.5\), a value of \(c_{\mathrm{long}}=73\) m/s can be predicted from theory. This is nearby the theoretical minimum value for \(\mu =0\). This theoretical \(c_{\mathrm{long}}\) increases for \(\mu \) approaching the pole \(0.5\). For \(\mu =0.45\), \(c_{\mathrm{long}}\) would be about twice the minimum value.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rupp, T.K., Ehlers, W., Karajan, N. et al. A forward dynamics simulation of human lumbar spine flexion predicting the load sharing of intervertebral discs, ligaments, and muscles. Biomech Model Mechanobiol 14, 1081–1105 (2015). https://doi.org/10.1007/s10237-015-0656-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10237-015-0656-2

Keywords

Navigation