Skip to main content
Log in

The Role of 3-Canal Biomechanics in Angular Motion Transduction by the Human Vestibular Labyrinth

  • Published:
Annals of Biomedical Engineering Aims and scope Submit manuscript

Abstract

The present work examines the role of the complex geometry of the human vestibular membranous labyrinth in the process of angular motion transduction by the semicircular canals. A morphologically descriptive mathematical model was constructed to address the biomechanical origins of temporal signal processing and directional coding in determining the inputs to the brain. The geometrical model was developed based on shrinkage-corrected temporal bone sections using a segmentation/data-fitting procedure. Endolymph fluid dynamics within the 3-canal labyrinth was modeled using an asymptotic form of the Navier–Stokes equations and solved to estimate endolymph and cupulae volume displacements. The geometrical model was manipulated to study the role of major morphological features on directional and temporal coding. Anatomical results show that the bony osseous canals provide reasonable estimates of the orientation of the delicate membranous canals—the two differed by only 3.48 ± 1.89°. Biomechanical results show that the maximal response directions are distinct from the anatomical canal planes, but can be closely approximated by fitting a flat plane to the centerline of the canal of interest and weighting each location along the centerline with the inverse of the cross-sectional area squared. Vector cross-products of these maximal response directions, in turn, determine the null planes and prime directions that transmit the direction of angular motion to the brain as three independent directional channels associated with the nerve bundles. Finally, parameter studies indicate that changes in canal cross-sectional area and shape only moderately affect canal temporal and directional coding, while three-canal orientation is critical to directional coding.

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.

Figure 1.
Figure 2.
Figure 3.
Figure 4.
Figure 5.
Figure 6.
Figure 7.
Figure 8.
Figure 9.
Figure 10.
Figure 11.
Figure 12.

Similar content being viewed by others

References

  1. Arnold B., Jager L., Grevers G. (1996) Visualization of inner ear structures by three-dimensional high-resolution magnetic resonance imaging. Am. J. Otol. 17(3): 480–485

    PubMed  CAS  Google Scholar 

  2. Baird R. A., et al. (1988) The vestibular nerve of the chinchilla. II. Relation between afferent response properties and peripheral innervation patterns in the semicircular canals. J. Neurophysiol. 60(1):182–203

    PubMed  CAS  Google Scholar 

  3. Blanks R. H., et al. (1985) Planar relationships of the semicircular canals in rhesus and squirrel monkeys. Brain Res. 340(2): 315–324

    Article  PubMed  CAS  Google Scholar 

  4. Blanks R. H., Curthoys I. S., Markham C. H. (1972) Planar relationships of semicircular canals in the cat. Am. J. Physiol. 223(1): 55–62

    PubMed  CAS  Google Scholar 

  5. Blanks R. H., Curthoys I. S., Markham C. H. (1975) Planar relationships of the semicircular canals in man. Acta Otolaryngol. 80(3–4): 185–196

    PubMed  CAS  Google Scholar 

  6. Boyle R., Highstein S. M. (1990) Resting discharge and response dynamics of horizontal semicircular canal afferent of the toadfish, Opsanus tau. J. Neurosci. 10: 1557–1569

    PubMed  CAS  Google Scholar 

  7. Cohen B. (1988) Representation of three-dimensional space in the vestibular, oculomotor, and visual systems. Concluding remarks. Ann. N.Y. Acad. Sci. 545:239–247

    Article  PubMed  CAS  Google Scholar 

  8. Cohen B., Suzuki J. I., Bender M. B. (1964) Eye Movements from Semicircular Canal Nerve Stimulation in the Cat. Ann. Otol. Rhinol. Laryngol. 73:153–169

    PubMed  CAS  Google Scholar 

  9. Counter S. A., et al. (2003) 3D MRI of the in vivo vestibulo-cochlea labyrinth during Gd-DTPA-BMA uptake. Neuroreport 14(13):1707–1712

    Article  PubMed  CAS  Google Scholar 

  10. Curthoys I. S., et al. (1975) The orientation of the semicircular canals in the guinea pig. Acta Otolaryngol. 80(3–4): 197–205

    PubMed  CAS  Google Scholar 

  11. Curthoys I. S., Oman C. M. (1987) Dimensions of the horizontal semicircular duct, ampulla and utricle in the human. Acta Otolaryngol. 103(3–4): 254–261

    PubMed  CAS  Google Scholar 

  12. Curthoys I. S., Markham C. H., Curthoys E. J. (1977) Semicircular duct and ampulla dimensions in cat, guinea pig and man. J. Morphol. 151(1): 17–34

    Article  PubMed  CAS  Google Scholar 

  13. Damiano E. R. (1997) A bi-phasic model of the cupula and the low-frequency mechanics of the vestibular semicircular canal. Soc. Mech. Eng. Bed. 35: 61–62

    Google Scholar 

  14. Damiano E. R., Rabbitt R. D. (1996) A singular perturbation model for fluid dynamics in the vestibular semicircular canal and ampulla. J. Fluid Mech. 307:333–372

    Article  Google Scholar 

  15. Della Santina, C. C., et al. Orientation of human semicircular canals measured by three-dimensional multiplanar CT reconstruction. J. Assoc. Res. Otolaryngol. 6(3):191–206, 2005.

    PubMed  Google Scholar 

  16. Dickman J. D. (1996) Spatial orientation of semicircular canals and afferent sensitivity vectors in pigeons. Exp. Brain Res. 111(1): 8–20

    Article  PubMed  CAS  Google Scholar 

  17. Estes M. S., Blanks R. H., Markham C. H. (1975) Physiologic characteristics of vestibular first-order canal neurons in the cat. I. Response plane determination and resting discharge characteristics. J. Neurophysiol. 38(5): 1232–1249

    PubMed  CAS  Google Scholar 

  18. Ezure K., Graf W.(1984) A quantitative analysis of the spatial organization of the vestibulo-ocular reflexes in lateral- and frontal-eyed animals–I. Orientation of semicircular canals and extraocular muscles. Neuroscience 12(1): 85–93

    Article  PubMed  CAS  Google Scholar 

  19. Ghanem T. A., Rabbitt R. D., Tresco P. A. (1998) Three-dimensional reconstruction of the membranous vestibular labyrinth in the toadfish, Opsanus tau. Hear Res. 124(1–2): 27–43

    Article  PubMed  CAS  Google Scholar 

  20. Groen J. J. (1957) The semicircular canal system of the organs of equilibrium. Phys. Med. Biol. 1(3):225–242

    Article  PubMed  CAS  Google Scholar 

  21. Haque A., Angelaki D. E., Dickman J. D. (2004) Spatial tuning and dynamics of vestibular semicircular canal afferents in rhesus monkeys. Exp. Brain. Res. 155(1): 81–90

    Article  PubMed  Google Scholar 

  22. Hoffmann K. P. (1988) Responses of single neurons in the pretectum of monkeys to visual stimuli in three-dimensional space. Ann. N.Y. Acad. Sci. 545:180–186

    Article  PubMed  CAS  Google Scholar 

  23. Igarashi M. (1967) Dimensional study of the vestibular apparatus. Laryngoscope 77(10): 1806–1817

    Article  PubMed  CAS  Google Scholar 

  24. Igarashi M., O-Uchi T., Alford B. R. (1981) Volumetric and dimensional measurements of vestibular structures in the squirrel monkey. Acta Otolaryngol. 91(5–6): 437–444

    PubMed  CAS  Google Scholar 

  25. Igarashi M., Ohashi K., Ishii M. (1986) Morphometric comparison of endolymphatic and perilymphatic spaces in human temporal bones. Acta Otolaryngol. 101(3–4): 161–164

    PubMed  CAS  Google Scholar 

  26. Jackler R. K., Dillon W. P. (1988) Computed tomography and magnetic resonance imaging of the inner ear. Otolaryngol. Head Neck Surg. 99(5): 494–504

    PubMed  CAS  Google Scholar 

  27. Koizuka I., et al. (1991) High-resolution magnetic resonance imaging of the human temporal bone. ORL J. Otorhinolaryngol. Relat. Spec. 53(6): 357–361

    PubMed  CAS  Google Scholar 

  28. Mayne R. (1950) The dynamic characteristics of the semicircular canals. J. Comp. Physiol. Psychol. 43: 304–319

    Article  Google Scholar 

  29. McCrea R. A., et al. (1987) Anatomical and physiological characteristics of vestibular neurons mediating the horizontal vestibulo-ocular reflex of the squirrel monkey. J. Comp. Neurol. 264(4): 547–570

    Article  PubMed  CAS  Google Scholar 

  30. Oman C. M., Marcus E. N., Curthoys I. S. (1987) The influence of semicircular canal morphology on endolymph flow dynamics. An anatomically descriptive mathematical model. Acta Otolaryngol. 103(1–2): 1–13

    PubMed  CAS  Google Scholar 

  31. Rabbitt R. D. (1999) Directional coding of three-dimensional movements by the vestibular semicircular canals. Biol. Cybern. 80(6):417–431

    Article  PubMed  CAS  Google Scholar 

  32. Rabbitt R. D., et al. (2004) Hair-cell versus afferent adaptation in the semicircular canals. J. Neruophysiol. 93(1): 424–436

    Article  Google Scholar 

  33. Rabbitt R. D., Boyle R., Highstein S. M. (1994) Sensory transduction of head velocity and acceleration in the toadfish horizontal semicircular canal. J. Neurophysiol. 72(2):1041–1048

    PubMed  CAS  Google Scholar 

  34. Rabbitt R. D., Damiano E. R., Grant J. W. (2003) Biomechanics of the vestibular semicircular canals and otolith organs. In: Highstein S. M., Popper A., Fay R. (eds) The Vestibular System. Springer-Verlag, New York, pp. 153–201

    Google Scholar 

  35. Rabbitt R. D., Damiano E. R., Grant J. W. (2004) Biomechanics of the semicircular canals and otolith organs. In: Highstein S. M., Fay R. R., Popper A. N. (eds) The Vestibular System. Springer-Verlag, New York

    Google Scholar 

  36. Rajguru S. M., Ifediba M. A., Rabbitt R. D. (2004) Three-dimensional biomechanical model of benign paroxysmal positional vertigo. Ann. Biomed. Eng. 32(6):831–846

    Article  PubMed  Google Scholar 

  37. Reisine H., Simpson J. I., Henn V. (1988) A geometric analysis of semicircular canals and induced activity in their peripheral afferents in the rhesus monkey. Ann. N.Y. Acad. Sci. 545:10–20

    Article  PubMed  CAS  Google Scholar 

  38. Robinson D. A. (1982) The use of matrices in analyzing the three-dimensional behavior of the vestibulo-ocular reflex. Biol. Cybern. 46(1):53–66

    Article  PubMed  CAS  Google Scholar 

  39. Schuknecht H. (1974) Pathology of the Ear. Harvard University Press, Cambridge, MA

    Google Scholar 

  40. Spoor F., Zonneveld F. (1998) Comparative review of the human bony labyrinth. Am. J. Phys. Anthropol. Suppl. 27: 211–251

    Article  Google Scholar 

  41. Steinhausen W. (1931) Uber den Nachweis der Bewegung der Cupula in der intakten Bogengangsampulle des Labyrinthes bei der naturlichen rotatorischen und calorischen Reizung. Plugers Arch Gesamte Physiol 228: 322–328

    Article  Google Scholar 

  42. Steinhausen W. (1933) Uber die beobachtungen der cupula in der bognegangsampullen des labyrinthes des libenden hecths. Pflugers Arch. 232: 500–512

    Article  Google Scholar 

  43. Takagi A., Sando I., Takahashi H. (1989) Computer-aided three-dimensional reconstruction and measurement of semicircular canals and their cristae in man. Acta Otolaryngol. 107(5–6):362–365

    PubMed  CAS  Google Scholar 

  44. Thorne M., et al. (1999) Cochlear fluid space dimensions for six species derived from reconstructions of three-dimensional magnetic resonance images. Laryngoscope 109(10):1661–1668

    Article  PubMed  CAS  Google Scholar 

  45. Uchino Y., et al. (1979) Horizontal canal input to cat extraocular motoneurons. Brain Res. 177(2): 231–240

    Article  PubMed  CAS  Google Scholar 

  46. Van Buskirk W. C., Watts R. G., Liu Y. K. (1976) The fluid mechanics of the semicircular canals. J. Fluid Mech. 78:87–98

    Article  Google Scholar 

  47. Van Egmond A. A. J., Groen J. J., Jongkees L. B. W. (1949) The mechanics of the semicircular canals. J. Physiol. Lond. 110:1–17

    Google Scholar 

  48. Voie A. H. (2002) Imaging the intact guinea pig tympanic bulla by orthogonal-plane fluorescence optical sectioning microscopy. Hear Res. 171(1–2): 119–128

    Article  PubMed  Google Scholar 

  49. Wilson V. J., Melvill Jones G. (1979) Mammalian Vestibular Physiology. Plenum Press, New York

    Google Scholar 

Download references

Acknowledgements

The authors would like to thank Dr. George Nager, Department of Otolaryngology—Head and Neck Surgery, Johns Hopkins University for providing the histological sections used in this work and Daniel Heibert for checking the equations and correcting a typographical sign error.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Richard D. Rabbitt.

Appendix

Appendix

The fluid filled labyrinth was modeled using the approach of Damiano and Rabbitt14 adjusted to the three-dimensional geometry of the human labyrinth. The endolymph was modeled as a Newtonian fluid undergoing low Reynolds number, low Stokes number, small displacement laminar flow. Inertial forcing due to acceleration of the head is introduced using a Galilean transformation. A slender body asymptotic expansion14 or, alternatively a control volume approach,34 reduces the Navier–Stokes equations to an ordinary differential equation acting along the curved centerline of each duct segment. Following the notation of Rabbitt31, for each short segment n of the labyrinth (n = HC, PC, AC, CC, UA, or UP), the volume displacement of the endolymph (Q) during head movements are represented by:

$$ m_n \frac{d^2Q_n}{dt^2}+c_n \frac{dQ_n}{dt}+k_n Q_n =P_n (l_n)-P_n (0)+f_n.$$
(1)

Parameters m n , c n and k n are proportional to the equivalent mass, damping and stiffness of the endolymph (or cupula depending on location in the canal, as will be discussed later), respectively, and can be estimated from endolymph density (ρ), viscosity (μ), shear stress (γ), and cross-sectional area function A(s) according to the following expressions:

$$ m_n =\int_0^{l_n} {\frac{\rho }{A(s)}ds}, $$
(2)
$$ c_n =\int_0^{l_n } {\frac{\mu \lambda }{A(s)^2}ds}, $$
(3)
$$ k_n =\int_0^{l_n } {\frac{\gamma \lambda }{A(s)^2}ds}. $$
(4)

The independent variable s n defines a curved coordinate running along the centerline of the canal segment. Integration limits are along the length (l n ) of the duct segment (n). The term λ is a dimensionless number that relates the viscous drag to the flow rate based on the cross-sectional shape and frequency of excitation.14, 30 For low Reynolds number, low Stokes number flow, the velocity distribution in an elliptical cross-section is\({u=\frac{\Updelta P}{2\mu \Updelta \ell }\frac{\left({ab} \right)^2}{\left({a^2+b^2} \right)}\left({1-\frac{x^2}{a^2}-\frac{y^2}{b^2}} \right)}\), where (x,y) are cross-sectional coordinates with origin at the center, (a,b) are the major and minor radii, and \({\Updelta P/\Updelta \ell}\) is the pressure gradient. This results in a viscous drag factor \({\uplambda =8\uppi \frac{2-\upvarepsilon ^2}{2\left({1-\upvarepsilon ^2} \right)^{3/2}}},\) where the elliptical eccentricity \({\varepsilon =\sqrt {1-\left(a/b \right)^2}}\). The parameter λ reduces to 8π for a circular cross-section and is easily extended to higher frequencies (Stokes number >1) by making λ a frequency-dependent complex-valued parameter following Damiano.14

The right hand side of Eq. (1) includes a pressure gradient, \({\Updelta P=P_{n}(l_{n})-P_{n}(0),}\) relating the pressures on either end of the labyrinth segment, and inertial forcing f n , due to angular acceleration of the head. The inertial forcing was calculated as

$$f_n =\int_0^{l_n } {\rho (\ddot {\vec {\Upomega }}\times \vec {R}(s))} \cdot d\vec {s}, $$
(5)

where \({\vec {R}}\) is a vector extending from the stereotactic head-fixed coordinate system origin to the centerline of the segment of interest. Angular acceleration (\({\ddot {\vec {\Upomega }})}\) is presented as a vector resolved in the head-fixed coordinate frame (Fig. A.1). It is calculated from angular acceleration relative to the ground-fixed inertial frame (\({\ddot {\vec {\Upomega}}_{\rm inertial})}\) by applying the time dependent orthonormal rotation matrix, N, relating the head-fixed frame to the ground-fixed frame using:

$$ \ddot {\vec {\Upomega }}=N\ddot {\vec {\Upomega }}_{\rm inertial}. $$
(6)
Figure A.1.
figure 13

Model labyrinth geometry. The human membranous labyrinth has 3 natural bifurcation points (1–3) where the six labyrinthine segments join. The three-dimensional motion of each segment was specified by time-dependent angular acceleration in the ground-fixed intertial frame and resolved into the moving head-fixed system. This introduces a Galilean transformation and an inertial force that appears in the Navier–Stokes equations. Poiseuille flow and slender body approximations were assumed to further reduce the equations to a set of coupled ordinary equations (Damiano and Rabbitt14). Equations for six segments were coupled together at the bifurcations by conservation of mass and pressure continuity.

Equations (1)–(6) were employed to compute the endolymph fluid displacements and resultant pressure gradients in a single canal or other labyrinth component. Comparable equations have been used to describe the dynamic responses of a single the HC14, 30 and a 3-canal labyrinth.31, 35 In the single-canal models the pressure gradient, ΔP, corresponded to transcupular pressure. In the current work, pressure gradients and volume displacements apply to six individual duct segments, the HC, AC, PC, CC, UA, and UP. Equations for the segments were coupled together to form a coupled 3-canal model. This was accomplished by identifying three bifurcation points (Fig. A.1) where the labyrinth segments naturally connect to one another and applying conservation of fluid volume and pressure continuity to create a matrix equation governing whole labyrinth endolymph (e) fluid mechanics:

$$ {\mathbf M}^e\frac{d^2\vec {Q}^e }{dt^2}+{{\mathbf C}^e}^e\frac{d\vec {Q}^e }{dt}+{{\mathbf K}^e}^e\vec {Q}^e =\vec {F}^e-\Delta \vec {P}. $$
(7)

The mass, \({{\mathbf M}^e,}\) damping \({{\mathbf C}^e,}\) and stiffness \({{\mathbf K}^e}\) are matrices are

$$ {\mathbf M}^e=\left[ {\begin{array}{lll} {\mathbf M}_{\rm HC} & m_{\rm UA} & -m_{\rm UP} \\ m_{\rm UA} & M_{\rm AC} & m_{\rm CC} \\ -m_{\rm UP}& m_{\rm CC} & M_{\rm PC} \\ \end{array} } \right], $$
(8)
$$ {\mathbf C}^e=\left[ {{\begin{array}{lll} C_{\rm HC}& c_{\rm UA} & -c_{\rm UP} \\ c_{\rm UA} & C_{\rm AC} & c_{\rm CC} \\ -c_{\rm UP} & c_{\rm CC} & C_{\rm PC} \\ \end{array} }} \right], $$
(9)

and

$$ {\mathbf K}^e=\left[ {{\begin{array}{lll} K_{\rm HC} & k_{\rm UA} & -k_{\rm UP} \\ k_{\rm UA} & K_{\rm AC} & k_{\rm CC} \\ -k_{\rm UP} & k_{\rm CC} & K_{\rm PC} \\ \end{array} }} \right]. $$
(10)

The diagonal elements are the addition of the segments forming a closed loop around the respective canal; e.g., \({M_{\rm HC} =m_{\rm HC} +m_{\rm UP} +m_{\rm UA}}\), \({M_{\rm PC} =m_{\rm PC} +m_{\rm CC} +m_{\rm UP} }\) and, \({M_{\rm AC}=m_{\rm AC} +m_{\rm CC} +m_{\rm UA}}\) . In the present work we have selected the local coordinate systems such that \({m_{\rm HC},m_{\rm UP},m_{\rm UA},m_{\rm PC} ,m_{\rm CC} }\) and m AC are all positive. This simplifies the sign convention and clarifies the sign of each element in the matrix (but differs from previous work31, 36). Each element in the above matrices was calculated from Eqs. (2)–(4). The pressure across the three cupulae is

$$\Updelta \vec {P}=\left[ {{\begin{array}{l} \Updelta P_{\rm HC} \\ \Updelta P_{\rm AC} \\ \Updelta P_{\rm PC} \\ \end{array} }} \right] $$
(11)

and inertial forcing becomes

$$ \vec {F}^e=\left[ {{\begin{array}{l} f_{\rm HC} +f_{\rm UA} +f_{\rm UP} \\ f_{\rm AC} +f_{\rm CC} +f_{\rm UA} \\ f_{\rm PC} +f_{\rm CC} -f_{\rm UP} \\ \end{array} }} \right]. $$
(12)

The vector \({\vec {Q}^e }\) contains the volume displacements of the endolymph at the HC, AC, and PC cupulae:

$$ \vec {Q}^e =\left[ {{\begin{array}{l} Q_{\rm HC}^e \\ Q_{\rm AC}^e \\ Q_{\rm PC}^e \\ \end{array} }} \right] $$
(13)

The preceding equations provide a method of relating endolymph volume displacements to prescribed head angular accelerations. In order to characterize the effect of these fluid displacements on cupula movement, it is necessary to develop a model for the cupula. The cupula has often been modeled as an elastic or viscoelastic material impermeable to endolymph.30, 33, 46 These models neglect the intrinsic porosity of the cupula’s mucopolysaccharide matrix, resulting in a 1:1 relationship between endolymph and cupula displacement. In this model, we represent cupula porosity by assuming that the structure is composed of a fluid and a solid phase. The solid portion responds to pressure gradients according to the Equation

$$ {\mathbf M}^c\frac{d^2\vec {Q}^c }{dt^2}+{\mathbf C}^c\frac{d\vec {Q}^c }{dt}+{\mathbf K}^c\vec {Q}^c =\Delta \vec {P}+\vec {F}^c, $$
(14)

in which \({{\mathbf M}^c,}\) \({{\mathbf C}^c,}\) and \({{\mathbf K}^c}\) are diagonal matrices corresponding to the equivalent mass, stiffness, and viscosity of each cupula. For these calculations, the upper limit of integration is cupula thickness, h c . Inertial forcing, \({\vec {F}^c}\), is calculated from Eq. (5), where ρ refers to the total density of the solid and fluid phase components of the cupula. The pressure gradient, \({\Updelta \vec {P}}\), is the interaction force, where

$$ \Updelta \vec {P}=\Upgamma \left({\frac{d\vec {Q}^e}{dt}-\frac{d\vec {Q}^c}{dt}} \right). $$
(15)

The constant Γ is related to Darcy’s constant (D a ), a term used to describe fluid flow through a porous medium, the cupula solid phase volume fraction (ψ), and the cross-sectional area of the cupula (A c) by

$$ \Upgamma =(D_a \psi A^c)^{-1}. $$
(16)

The equations governing endolymph and cupula responses were combined to produce the following differential equation describing the entire system:

$${\mathbf {M}}\frac{d^2\vec {Q}}{dt^2}+{\mathbf C}\frac{d\vec {Q}}{dt}+{\mathbf K}\vec {Q}=\vec {F}. $$
(17)

Note, the forcing vector \({\vec {F}}\) has units of pressure and the displacement \({\vec {Q}}\) has units of volume. The effective mass, stiffness, and damping matrices are

$$ {\mathbf M}=\left[ {\begin{array}{ll} {\mathbf M}^e & 0 \\ 0 & {\mathbf M}^c \\ \end{array} } \right], $$
(18)
$$ {\mathbf C}=\left[ {{\begin{array}{ll} {\mathbf C}^e+\Upgamma & -\Upgamma \\ -\Upgamma & {\mathbf C}^c+\Upgamma \\ \end{array} }} \right] $$
(19)

and

$$ {\mathbf K}=\left[ {\begin{array}{ll}{\mathbf K}^e & 0 \\ 0 & {\mathbf K}^c \\ \end{array} } \right], $$
(20)

respectively, where the matrix \({\Upgamma=\Upgamma {\mathbf I}}\) .

The inertial forcing vector is condensed to

$$\vec {F}=\left[ {\begin{array}{l} \vec {F}^e \\ \vec {F}^c \\ \end{array} } \right] $$
(21)

and endolymph, and cupula volume displacements are simply

$$ \vec {Q}=\left[ {\begin{array}{l} \vec {Q}^e \\ \vec {Q}^c \\ \end{array} } \right]. $$
(22)

Thus any three-dimensional head movement may be described in terms of cupula volume displacements in the three canals. The physical parameter values used in these equations are listed in Table A.1.

Table A.1. Model parameter values.

The mass, damping, and stiffness matrix elements were computed using the geometry and the physical parameters listed in Table A.1. For labyrinth 1, the effective mass matrix (g/cm4) was

$${\mathbf M}=\left[ {\begin{array}{llllll} 1134 & 18.6 & 0 & 0 & 0 & 0 \\ 18.6 & 1165 & 38.2 & 0 & 0 & 0 \\ 0 & 38.2 & 1517 & 0 & 0 & 0 \\ 0 & 0 & 0 & 2.39 & 0 & 0 \\ 0 & 0 & 0 & 0 & 2.37 & 0 \\ 0 & 0 & 0 & 0 & 0 & 2.98 \\ \end{array} } \right], $$
(23)

the damping matrix (dyn s/cm5) was

$${\mathbf C}=\left[ {\begin{array}{llllll} 2.19e6 & 287 & 0 & -2e6 & 0 & 0\\ 287 & 2.21e6 & 1105 & 0 & -2e6 &0 \\ 0 & 1105 & 2.26e6 & 0 & 0 & -2e6\\ -2e6 & 0 & 0 & 2e6 & 0 & 0 \\ 0 & -2e6 & 0 & 0 & 2e6 & 0 \\ 0 & 0 & -2e6 & 0 & 0 & 2e6 \\ \end{array} } \right], $$
(24)

and the stiffness matrix (dyn/cm5) was

$$ {\mathbf K}=\left[{{\begin{array}{llllll} 0 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 17016 & 0 & 0 \\ 0 & 0 & 0 & 0 & 16715 & 0 \\ 0 & 0 & 0 & 0 & 0 & 26522 \\ \end{array} }} \right]. $$
(25)

The pressure forcing vector (dyn/cm2) was determined from

$$\vec {F}=\left[ {\begin{array}{l} -0.6107 \\ -0.00757 \\ -0.00480 \\ 0.00990 \\ 0.00606 \\ 0.00423 \\ \end{array}} \right] {\ddot {\Upomega}}_{{\rm HC}^{\prime}}+ \left[{\begin{array}{l} -0.000348 \\ 0.6399 \\ 0.00981 \\ 0.00115 \\ 0.00393 \\ -0.00559 \\ \end{array}} \right] {\ddot {\Upomega}}_{{\rm AC}^{\prime}} +\left[{\begin{array}{l} -0.000348 \\ 0.00906 \\ 0.7991 \\ 0.00364 \\ -0.00522 \\ -0.0106 \\ \end{array}} \right] {\ddot {\Upomega}}_{{\rm PC}^{\prime}}, $$
(26)

where \({\ddot {\Upomega}_{\rm HC}^{\prime} }\) , \({\ddot {\Upomega}_{\rm AC}^{\prime}}\) and \({\ddot {\Upomega}_{\rm PC}^{\prime}}\) are the projected components of head angular acceleration (rad/s2) in the prime directions \({{n}_{\rm HC}^{\prime} =[-0.064,0.042,-0.997]}\), \({{n}_{\rm AC}^{\prime} =[0.702,0.699,-0.136]}\) and \({{n}_{\rm PC}^{\prime} =[0.857,-0.557,-0.103]}\), respectively.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ifediba, M.A., Rajguru, S.M., Hullar, T.E. et al. The Role of 3-Canal Biomechanics in Angular Motion Transduction by the Human Vestibular Labyrinth. Ann Biomed Eng 35, 1247–1263 (2007). https://doi.org/10.1007/s10439-007-9277-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10439-007-9277-y

Keywords

Navigation