Skip to main content
Log in

Striola magica. A functional explanation of otolith geometry

  • Published:
Journal of Computational Neuroscience Aims and scope Submit manuscript

Abstract

Otolith end organs of vertebrates sense linear accelerations of the head and gravitation. The hair cells on their epithelia are responsible for transduction. In mammals, the striola, parallel to the line where hair cells reverse their polarization, is a narrow region centered on a curve with curvature and torsion. It has been shown that the striolar region is functionally different from the rest, being involved in a phasic vestibular pathway. We propose a mathematical and computational model that explains the necessity of this amazing geometry for the striola to be able to carry out its function. Our hypothesis, related to the biophysics of the hair cells and to the physiology of their afferent neurons, is that striolar afferents collect information from several type I hair cells to detect the jerk in a large domain of acceleration directions. This predicts a mean number of two calyces for afferent neurons, as measured in rodents. The domain of acceleration directions sensed by our striolar model is compatible with the experimental results obtained on monkeys considering all afferents. Therefore, the main result of our study is that phasic and tonic vestibular afferents cover the same geometrical fields, but at different dynamical and frequency domains.

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
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Berthoz, A., Allain, R., Bennequin, D., David, R., Janvier, P. (2011). Sortir de l’eau. Le passage de la vie aquatique la vie terrestre., chap. Sortir de l’eau et systeme vestibulaire. P. Corvol et J.-L. Elghozi (dir.), Odile Jacob.

  • Boyle, R., Goldberg, J.M., Highstein, S.M. (1992). Inputs from regularly and irregularly discharging vestibular nerve afferents to secondary neurons in squirrel monkey vestibular nuclei. III. Correlation with vestibulospinal and vestibuloocular output pathways. Journal of Neurophysiology, 68(2), 471–484. http://jn.physiology.org/content/68/2/471.abstract.

    PubMed  CAS  Google Scholar 

  • Brichta, A.M., & Goldberg, J.M. (1998). The Papilla Neglecta of Turtles: A Detector of Head Rotations with Unique Sensory Coding Properties. Journal of Neuroscience, 18(11), 4314–4324. http://www.jneurosci.org/cgi/content/abstract/18/11/4314.

    PubMed  CAS  Google Scholar 

  • Bronstein, A.M., & Gresty, M.A. (1988). Short latency compensatory eye movement responses to transient linear head acceleration: a specific function of the otolith-ocular reflex. Experimental Brain Research, 71(2), 406–410.

    Article  CAS  Google Scholar 

  • Chang, J., Popper, A., Saidel, W. (1992). Heterogeneity of sensory hair cells in a fish ear. Journal of Comparative Neurology, 4(324), 621–640.

    Article  Google Scholar 

  • Coxeter, H. (1969). Introduction to Geometry. New York: Wiley.

    Google Scholar 

  • Curthoys, I.S., Betts, G.A., Burgess, A.M., MacDougall, H.G., Cartwright, A.D., Halmagyi, G.M. (1999). The planes of the utricular and saccular maculae of the guinea pig. Annals of the New York Academy of Sciences, 871(1), 27–34. doi:10.1111/j.1749-6632.1999.tb09173.x.

    Article  PubMed  CAS  Google Scholar 

  • Curthoys, I.S., Black, R.A., Goldberg, J.M., Fernandez, C. (1995). New representations of otolithic primary afferent spatial tuning–a re-processing of the Fernandez & Goldberg (1976) data. Acta oto-laryngologica Supplement, 520 Pt 2, 427–429. http://view.ncbi.nlm.nih.gov/pubmed/8749180.

    Article  CAS  Google Scholar 

  • Darboux, G. (1887). Leçons sur la Theorie generale des Surfaces et les applications geometriques du Calcul Infinitesimal (Vol. 1). Gauthier-Villars.

  • Deliagina, T.G., Beloozerova, I.N., Zelenin, P.V., Orlovsky, G.N. (2008). Spinal and supraspinal postural networks. Brain Research Reviews, 57(1), 212–221. doi:10.1016/j.brainresrev.2007.06.0172007.06.017.

    Article  PubMed  CAS  Google Scholar 

  • Desai, S.S., Zeh, C., Lysakowski, A. (2005). Comparative morphology of rodent vestibular periphery. I. Saccular and utricular maculae. Journal of Neurophysiology, 93(1), 251–266. doi:10.1152/jn.00746.2003.

    Article  PubMed  Google Scholar 

  • Dickman, J., Angelaki, D., Correia, M. (1991). Response properties of gerbil otolith afferents to small-angle pitch and roll tilts. Brain Research, 556, 303–310.

    Article  PubMed  CAS  Google Scholar 

  • Eatock, R., & Songer, J. (2011). Vestibular hair cells and afferents: two channels for head motion signals. Annual Review of Neuroscience, 34, 501–534.

    Article  PubMed  CAS  Google Scholar 

  • Eugene, D., Idoux, E., Beraneck, M., Moore, L., Vidal, P.P. (2011). Intrinsic membrane properties of central vestibular neurons in rodents. Experimental Brain Research, 1–14. doi:10.1007/s00221-011-2569-3.

  • Fernandez, C., & Goldberg, J.M. (1976a). Physiology of peripheral neurons innervating otolith organs of the squirrel monkey. I. Response to static tilts and to long-duration centrifugal force. Journal of Neurophysiology, 39(5), 970–984. http://view.ncbi.nlm.nih.gov/pubmed/824412.

    PubMed  CAS  Google Scholar 

  • Fernandez, C., & Goldberg, J.M. (1976b). Physiology of peripheral neurons innervating otolith organs of the squirrel monkey. II. Directional selectivity and force-response relations. Journal of Neurophysiology, 39(5), 985–995. http://view.ncbi.nlm.nih.gov/pubmed/824413.

    PubMed  CAS  Google Scholar 

  • Fernandez, C., & Goldberg J.M. (1976c). Physiology of peripheral neurons innervating otolith organs of the squirrel monkey. III. Response dynamics. Journal of Neurophysiology, 39(5), 996–1008. http://view.ncbi.nlm.nih.gov/pubmed/824414.

    PubMed  CAS  Google Scholar 

  • Fernandez, C., Goldberg, J.M., Baird, R.A. (1990). The vestibular nerve of the chinchilla. III. Peripheral innervation patterns in the utricular macula. Journal of Neurophysiology, 63(4), 767–780. http://jn.physiology.org/content/63/4/767.abstract.

    PubMed  CAS  Google Scholar 

  • Georgopoulos, A., Schwartz, A., Kettner, R. (1986). Neuronal population coding of movement direction. Science, 233, 1416–1419.

    Article  PubMed  CAS  Google Scholar 

  • Goldberg, J.M. (2000). Afferent diversity and the organization of central vestibular pathways. Experimental Brain Research, 130(3), 277–297. doi:10.1007/s002210050033.

    Article  CAS  Google Scholar 

  • Goldberg, J.M., Desmadryl, G., Baird, R.A., Fernandez, C. (1990). The vestibular nerve of the chinchilla. IV. Discharge properties of utricular afferents. Journal of Neurophysiology, 63(4), 781–790. http://jn.physiology.org/content/63/4/781.abstract.

    PubMed  CAS  Google Scholar 

  • Greenhill, A. (1892). The applications of elliptic functions. New York: MacMillan and Co.

    Google Scholar 

  • Griffith, P., & Harris, J. (1978). Principles of algebraic geometry. New York: Wiley.

    Google Scholar 

  • Hess, B.J. (1992). Three-dimensional head angular velocity detection from otolith afferent signals. Biological Cybernetics, 67(4), 323–333. http://view.ncbi.nlm.nih.gov/pubmed/1515511.

    Article  PubMed  CAS  Google Scholar 

  • Highstein, S., Rabbitt, R., Holstein, G., Boyle, R. (2005). Determinants of spatial and temporal coding by semicircular canal afferents. Journal of Neurophysiology, 93(5), 2359–2370.

    Article  PubMed  Google Scholar 

  • Jaeger, R. (2002). Modeling the relation between head orientations and otolith responses in humans. Hearing Research, 173(1–2), 29–42. doi:10.1016/S0378-5955(02)00485-9.

    Article  PubMed  CAS  Google Scholar 

  • Jaeger, R., & Haslwanter, T. (2004). Otolith responses to dynamical stimuli: results of a numerical investigation. Biological Cybernetics, 90(3), 165–175. doi:10.1007/s00422-003-0456-0.

    Article  PubMed  CAS  Google Scholar 

  • Jaeger, R., Kondrachuk, A.V., Haslwanter, T. (2008). The distribution of otolith polarization vectors in mammals: comparison between model predictions and single cell recordings. Hearing research, 239(1–2), 12–19. doi:10.1016/j.heares.2008.01.004.

    Article  PubMed  CAS  Google Scholar 

  • Jones, T., Jones, S., Colbert, S. (1998). The adequate stimulus for avian short latency vestibular responses to linear translation. Journal of Vestibular Research, 3(8), 253–72.

    Article  Google Scholar 

  • Lanford, P.J., Platt, C., Popper, A.N. (2000). Structure and function in the saccule of the goldfish (Carassius auratus): a model of diversity in the non-amniote ear. Hearing Research, 143(1–2), 1–13. http://view.ncbi.nlm.nih.gov/pubmed/10771179.

    Article  PubMed  CAS  Google Scholar 

  • Lange, M., & Jones, T. (1990). Short latency electrophysiological responses to pulsed linear acceleration in the mammal. Association for Research in Otolaryngology, 343–390.

  • Leonard, R. (2002). Molecular probes of the vestibular nerve I. Peripheral termination patterns of calretinin, calbindin and peripherin containing fibers. Brain Research, 928(1–2), 8–17. doi:10.1016/S0006-8993(01)03268-1.

    Article  PubMed  CAS  Google Scholar 

  • Li, A., Xue, J., Peterson, E. (2007). Architecture of the mouse utricle: macular organization and hair bundle heights. Journal of Neurophysiology, 99(2), 718–733.

    Article  PubMed  Google Scholar 

  • Li, A., Xue, J., Peterson, E.H. (2008). Architecture of the mouse utricle: macular organization and hair bundle heights. Journal of Neurophysiology, 99(2), 718–733. doi:10.1152/jn.00831.2007.

    Article  PubMed  CAS  Google Scholar 

  • Lin, V., Golub, J., Nguyen, T., Hume, C., Oesterle, E., Stone, J. (2011). Inhibition of notch activity promotes nonmitotic regeneration of hair cells in the adult mouse utricles. The Journal of Neuroscience, 31(43), 15329–15339.

    Article  PubMed  CAS  Google Scholar 

  • Lindeman, H. (1969). Studies on the morphology of the sensory regions of the vestibular apparatus. Erg Anat Entw Gesch, 42, 1–113.

    CAS  Google Scholar 

  • Lindeman, H.H. (1973). Anatomy of the otolith organs. Advances in Oto-rhino-laryngology, 20, 405–433. http://view.ncbi.nlm.nih.gov/pubmed/4267996.

    PubMed  CAS  Google Scholar 

  • Lorincz, E.N., & Hess, B.J.M. (2008). Dynamic effects on the subjective visual vertical after roll rotation. Journal of Neurophysiology, 100(2), 657–669. doi:10.1152/jn.01248.2007.

    Article  PubMed  Google Scholar 

  • Lu, Z. (1998). Morphological polarizations of sensory hair cells in the three otolithic organs of a teleost fish: fluorescent imaging of ciliary bundles. Hearing Research, 126(1–2), 47–57. doi:10.1016/S0378-5955(98)00149-X.

    Article  PubMed  CAS  Google Scholar 

  • Lyford-Pike, S., Vogelheim, C., Chu, E., Santina, C.D., Carey, J. (2007). Gentamicin is primarily localized in vestibulat type I hair cells after intratympanic administration. Journal of the Association for Research in Otolaryngology (JARO), 8, 497–508.

    Article  Google Scholar 

  • Minor, L.B., & Goldberg, J.M. (1991). Vestibular-nerve inputs to the vestibulo-ocular reflex: a functional- ablation study in the squirrel monkey. The Journal of Neuroscience, 11(6), 1636–1648. http://www.jneurosci.org/content/11/6/1636.abstract.

    PubMed  CAS  Google Scholar 

  • Moravec, W., & Peterson, E. (2004). Differences between stereocilia numbers on type i and type ii vestibular hair cells. Journal of Neurophysiology, 92, 3153–3160.

    Article  PubMed  CAS  Google Scholar 

  • Naganuma, H., Tokumasu, K., Okamoto, M., Hashimoto, S., Yamashina S. (2001). Three-dimensional analysis of morphological aspects of the human saccular macula. The Annals of Otology, Rhinology, and Laryngology, 110(11), 1017–1024. http://view.ncbi.nlm.nih.gov/pubmed/11713911.

    PubMed  CAS  Google Scholar 

  • Naganuma, H., Tokumasu K., Okamoto, M., Hashimoto, S., Yamashina, S. (2003). Three-dimensional analysis of morphological aspects of the human utricular macula. The Annals of Otology, Rhinology, and Laryngology, 112(5), 419–424. http://view.ncbi.nlm.nih.gov/pubmed/12784980.

    PubMed  Google Scholar 

  • Nam, J., Cotton, J., Grant, W. (2007a). A virtual hair cell, I: evaluation of mechanoelectric transduction parameters. Biophysical Journal, 92(6), 1918–1928.

    Article  PubMed  CAS  Google Scholar 

  • Nam, J., Cotton, J., Grant, W. (2007b). A virtual hair cell, II: evaluation of mechanoelectric transduction parameters. Biophysical Journal, 92(6), 1929–1937.

    Article  PubMed  CAS  Google Scholar 

  • Peterson, E. (1988). Are there parallel channels in the vestibular nerve? New Physiology Science, 13, 194–201.

    Google Scholar 

  • Platt, C., Jrgensen, J.M., Popper, A.N. (2004). The inner ear of the lungfish Protopterus. Journal of Comparative Neurology, 471(3), 277–288. doi:10.1002/cne.20038.

    Article  PubMed  Google Scholar 

  • Rabbitt, R. (1999). Directional coding of three-dimensional movements by the vestibular semicircular canals. Biological Cybernetics, 80(6), 417–431.

    Article  PubMed  CAS  Google Scholar 

  • Ross, M. (1988). Morphological evidence for parallel processing of information in rat macula. Acta Otolaryngol, 106(3–4), 213–218.

    Article  PubMed  CAS  Google Scholar 

  • Ross M. (1997). Morphological evidence for local microcircuits in rat vestibular maculae. Journal of Comparative Neurology, 379(3), 333–346. doi:10.1002/(SICI)1096-9861(19970317)379:3%3C333::AID-CNE2%3E3.0.CO;2-4.

    Article  PubMed  CAS  Google Scholar 

  • Ross, M. (2001). Complex vestibular macular anatomical relationships need a synthetic approach. Acta Otolaryngol Supply, 545(*), 25–28.

    Article  CAS  Google Scholar 

  • Ross, M. (2003). The evolution of concepts of vestibular peripheral information processing: toward the dynamic, adaptive, parallel processing macular model. Acta Oto-laryngologica, 123(7), 784–794. http://view.ncbi.nlm.nih.gov/pubmed/14575392.

    PubMed  Google Scholar 

  • Ross, M., Dayhoff, J., Mugler, D. (1990). Toward modeling a dynamic biological neural network. Mathematical and Computer Modelling, 13(7), 97–105.

    Article  PubMed  CAS  Google Scholar 

  • Ross, M., Linton S., Parnas, B. (2000). Simulation studies of vestibular macular afferent-discharge patterns using a new, quasi-3-d finite volume method. Journal of Computational Neuroscience, 8(1), 5–18.

    Article  PubMed  CAS  Google Scholar 

  • Rowe, H., & Peterson, E. (2004). Quantitative analysis of stereociliari arrays on vestibular hair cells. Hearing Research, 190, 10–24.

    Article  PubMed  Google Scholar 

  • Rowe, M., & Peterson, E. (2006). Autocorrelation analysis of hair bundle structure in the utricle. Journal of Neurophysiology, 96(5), 2653–2669.

    Article  PubMed  CAS  Google Scholar 

  • Rumelhart, D.E., Hintont, G.E., Williams R.J. (1986). Learning representations by back-propagating errors. Nature, 323(6088), 533–536. London.

    Article  Google Scholar 

  • Sato, H., Sando, I., Takahashi, H. (1992). Computer-aided three-dimensional measurement of the human vestibular apparatus. Otolaryngology–Head and Neck Surgery: Official Journal of American Academy of Otolaryngology-Head and Neck Surgery, 107(3), 405–409. http://view.ncbi.nlm.nih.gov/pubmed/1408226.

    CAS  Google Scholar 

  • Si, X., Angelaki, D., Dickman, J. (1997). Response properties of pigeon otolith afferents to linear acceleration. Experimental Brain Research, 117, 242–250.

    Article  CAS  Google Scholar 

  • Si, X., Zakir, M.M., Dickman, J.D. (2003). Afferent innervation of the utricular macula in Pigeons. Journal of Neurophysiology, 89(3), 1660–1677. doi:10.1152/jn.00690.2002.

    Article  PubMed  Google Scholar 

  • Simmons, D.D., Tong, B., Schrader, A.D., Hornak, A.J. (2010). Oncomodulin identifies different hair cell types in the mammalian inner ear. Journal of Comparative Neurology, 518(18), 3785–3802. doi:10.1002/cne.22424.

    Article  PubMed  CAS  Google Scholar 

  • Smith, C., & Goldberg, J. (1986). A stochastic afterhyperpolarization model of repetitive activity in vestibular afferents. Biological Cybernetics, 54(1), 41–51.

    Article  PubMed  CAS  Google Scholar 

  • Spoedlin, H. (1966). The vestibular system and its deseases In R.J. Wolfson (Ed.), The ultrastructure of the vestibular sense organ. Philadelphia: University of Pennsylvania Press.

    Google Scholar 

  • Spoon, C., Moravec, W., Rowe, M., Grant, J., Peterson, E. (2011). Steady-state stiffness of utricular hair cells depends on macular location and hair bundle structure. Journal of Neurophysiology, 106(6), 2950–2963.

    Article  PubMed  Google Scholar 

  • Straka, H., Beraneck, M., Rohregger, M., Moore, L.E., Vidal, P.P., Vibert, N. (2004). Second-order vestibular neurons form separate populations with different membrane and discharge properties. Journal of Neurophysiology, 92(2), 845–861. doi:10.1152/jn.00107.2004.

    Article  PubMed  CAS  Google Scholar 

  • Takagi, A., & Sando, I. (1988). Computer-aided three-dimensional reconstruction and measurement of the vestibular end-organs. Otolaryngology–Head and Neck Surgery: Official Journal of American Academy of Otolaryngology-Head and Neck Surgery, 98(3), 195–202. http://view.ncbi.nlm.nih.gov/pubmed/2895448.

    CAS  Google Scholar 

  • Tomko, D., Peterka, R., Schor, R., O’Leary, D. (1981). Response dynamics of horizontal canal afferents in barbiturate-anesthetized cats. Journal of Neurophysiology, 45(3), 376–396.

    PubMed  CAS  Google Scholar 

  • Tribukait, A., & Rosenhall, U. (2001). Directional sensitivity of the human macula utriculi based on morphological characteristics. Audiology & Neuro-otology, 6(2), 98–107. http://view.ncbi.nlm.nih.gov/pubmed/11385183.

    Article  CAS  Google Scholar 

  • Tribukait, A., Rosenhall, U., Osterdahl, B. (2005). Morphological characteristics of the human macula sacculi. Audiology & Neuro-otology, 10(2), 90–96. doi:10.1159/000083364.

    Article  Google Scholar 

  • Vilares, I., & Kording, K. (2011). Bayesian models: the structure of the world, uncertainty, behavior, and the brain. Annals of New-York Adademy of Sciences, 1224, 22–39.

    Article  Google Scholar 

  • Watanuki, K., Kawamoto, K., Katagiri, S. (1971). Structure of the otolithic layers on the maculae sacculi and utriculi in the guinea pig. Equilibrium Research Supplementary, 2, 41–48.

    Article  Google Scholar 

  • Xue, J., & Peterson, E. (2006). Hair bundle heights on the utricle: differences between macular locations and hair cell types. Journal of Neurophysiology, 95, 176–186.

    Google Scholar 

  • Zakir, M., Huss, D., Dickman, J.D. (2003). Afferent innervation patterns of the saccule in pigeons. Journal of Neurophysiology, 89(1), 534–550. doi:10.1152/jn.00817.2001.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgments

This work was supported in part by the European Projects CLONS (FP7-ITC-2007.8.0. Project 225929). We are very grateful to Dr. Rudi Jaeger and Professor Emeritus Gerald Jay Goldberg to give us the permission of using their data and reproducing their images. D.B. thanks Ruth-Anne Eatock for very interesting discussions. The authors thank warmly the two anonymous referees which helped them to improve their work and its presentation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mariella Dimiccoli.

Additional information

Action Editor: Alain Destexhe

Mathematical appendix

Mathematical appendix

This appendix provides the mathematical definitions, theorems and proofs, underlying our Striola model. In particular it describes the virtual surface generated by a twisted curve in the Euclidian space and the optimal shapes of curves allowing an uniform detection of acceleration directions.

The suggested information processing of the striolar region on the otoliths macula lays on the correspondence between the pairs of points of a space curve and the directions orthogonal to the tangents at the points. The shape of the striola is dictated by the range and the regularity of this correspondence, thus we will conduct a study of its properties. In Section A.1 we show that the regularity depends on the existence of curvature and torsion. This gives an elementary introduction in the affine framework to the more sophisticated investigations of Sections A.2 and A.3.

Notations: we consider a smooth curve S in the three dimensional Euclidian space, and we study the correspondence between the pairs of points \((P_{1},P_{2})\) of S and the directions D that are orthogonal to S in \(P_{1}\) and \(P_{2}\). We are particularly interested in the case where this correspondence defines a mapping \(D=F(P_{1},P_{2})\) for pairs of points \(P_{1},P_{2}\) that are sufficiently closed in S. We will consider several natural measures of area on the set of pairs of points on S and on the set of directions in the space. The regularity and uniformity of the mapping F is encoded by its jacobian determinant J, then we will study the link between J and the geometry of S. From the Euclidian point of view it is natural to retain on the set of points pairs, the measure associated to the canonical coordinates \((\sigma _{1},\sigma _{2})\), where \(\sigma _{1}=s_{1}+s_{2}\) and \(\sigma _{2}=s_{1}s_{2}\) are the elementary symmetric functions of the curvilinear abscissas \(s_{1},s_{2}\) of the respective points \(P_{1},P_{2}\) on S, and on the set of directions D the measure given by solid angle; this is the framework of Section A.2. In Section A.3 we restrict our attention to the case of a spherical curve S, for application to our model of the striola curve. However ,we begin in Section A.1 with the Cartesian point of view, where computations are more elementary. In this case a natural parametrization for points on S comes from the orthogonal projection on one axis and natural coordinates for directions in the space come from the intersection with a fixed plane. (Figs. 13 and 14)

Fig. 13
figure 13

Representation of the preferred direction encoded by each couples of cells. It is here computed for the curve \(x(t)=t/\sqrt (2), y=t^{2}, z = \sqrt (2)/3 \times t^{3}\), for \(t \in [-10;10]\), with 25 cells distributed equally along t. Panels show the projection of this surface in the \((x,y)\)\((y,z)\) and \((x,z)\) planes, the color palette represents the value of missing component (z, x and y respectively)

Fig. 14
figure 14

Representation of the virtual surface \(\Sigma \), defined by Eq. (3436), whose normals define the perceived directions. It is here computed for the curve \(x(t)=t/\sqrt (2), y=t^{2}, z = \sqrt (2)/3 \times t^{3}\), for \(t \in [-10;10]\). Panels show the projection of this surface in the \((x,y)\)\((z,y)\) and \((x,z)\) planes, as well as in an arbitrary plane, the color palette represents the same z value in each of the panels

1.1 A.1 Cartesian optimal striola

Let us consider a space curve S described in parametric form by cartesian coordinates

$$ x=f(t),\quad y=g(t),\quad z=h(t). $$
(32)

If a vector \(\overrightarrow {A}\) has coordinates \(u,v,w\), its scalar product with the tangent \(\overrightarrow {T}\) of S is given by

$$ \overrightarrow{A}.\overrightarrow{T}=f'(t)u+g'(t)v+h'(t)w. $$
(33)

The orthogonality equation \(\overrightarrow {A}.\overrightarrow {T}=0\) gives

$$ f'(t)u+g'(t)v+h'(t)w=0. $$
(34)

We suppose that \(\overrightarrow {T}\) is nowhere 0, and that there exist a conical domain of \(\overrightarrow {A}\) where the Eq. (34) has two solutions \(t_{1}(u,v,w)\)\(t_{2}(u,v,w)\) (eventually equals). In this case, for each \(\overrightarrow {A}\) in the domain, there exist two parameters \(t_{1},t_{2}\) and a non-zero real number c such that

$$ c\overrightarrow{A}=\overrightarrow{T}(t_{1})\times\overrightarrow{T}(t_{2}), $$
(35)

which translates in coordinates as

$$cu=g'(t_{1})h'(t_{2})-g'(t_{2})h'(t_{1}),$$
(36)
$$cv=h'(t_{1})f'(t_{2})-h'(t_{2})f'(t_{1}),$$
(37)
$$cw=f'(t_{1})g'(t_{2})- f'(t_{2})g'(t_{1}).$$
(38)

A classical theorem asserts that any smooth function \(\varphi \) (resp. polynomial function) of two real variables \(t_{1},t_{2}\) (defined in a domain of the plane which is symmetric with respect to the diagonal \(t_{1}=t_{2}\)) which satisfies \(\varphi (t_{1},t_{2})=\varphi (t_{2},t_{1})\) for any pair \((t_{1},t_{2})\), is equal to a smooth (resp. polynomial) function \(\Phi (\sigma _{1},\sigma _{2})\) where \(\sigma _{1}=t_{1}+t_{2}\) and \(\sigma _{2}=t_{1}t_{2}\) are the elementary symmetric functions of \(t_{1}\) and \(t_{2}\).In addition we can suppose that one of the coordinates \(u,v,w\) can be arbitrarily chosen, say for instance \(w=1\). Then we get a transformation F from a domain of \((\sigma _{1},\sigma _{2})\) to a domain of \((u,v)\). Let us denote \(J(\sigma _{1},\sigma _{2})\) the jacobian determinant of the transformation F.

Theorem 1

Let us suppose that J is regular and non-vanishing in the neighborhood of a point (0,0) corresponding to a pair of coincident points, then it exists an Euclidian affine change of coordinates \(x,y,z\) , where we have \(f(0)=0\)\(f'(0)\neq 0\)\(g(0)=g'(0)=0\) but \(g''(0)\neq 0\) , and \(h(0)=h'(0)=h''(0)=0\) but \(h^{(3)}\neq 0\). In other terms, after an Euclidian affine change of frame, the valuations of the Taylor series of g and h are the same as the ones of the twisted cubic.

Proof

We begin by fixing the x axis in such a way that \(f(0)=0\)\(f'(0)\neq 0\). We change the parametrization t to have \(f(t)=at\) in a neighborhood of \(t=0\). It is easy to show this has no effect on the conclusions of the theorem. We suppose \(w=1\). In the considered domain of \(u,v\) there exist two solutions \(t_{1}(u,v)\)\(t_{2}(u,v)\) (eventually equals) of the orthogonality equation, that is:

$$ au+g'(t_{1})v+h'(t_{1})=0,\quad\quad au+g'(t_{2})v+h'(t_{2})=0. $$
(39)

By solving this linear system we obtain \(u,v\) from \(t_{1},t_{2}\):

$$u=\frac{h'(t_{2})g'(t_{1})-h'(t_{1})g'(t_{2})}{a(g'(t_{2})-g'(t_{1}))},$$
(40)
$$v=\frac{h'(t_{1})-h'(t_{2})}{g'(t_{2})-g'(t_{1})}.$$
(41)

On another side, by the implicit function theorem we have

$$(g''(t_{1})v+h''(t_{1}))\frac{\partial t_{1}}{\partial u}=-a,$$
(42)
$$(g''(t_{1})v+h''(t_{1}))\frac{\partial t_{1}}{\partial v}=-g'(t_{1}),$$
(43)

and similar equations for \(t_{2}\). Thus the jacobian determinant of the transformation from \(t_{1},t_{2}\) to \(u,v\) is equal to

$$\begin{array}{rll} J'(u,v;t_{1},t_{2})&=&a^{-1}(g'(t_{2})-g'(t_{1}))^{-1}(g''(t_{1})v+h''(t_{1}))\\ &&\quad \times (g''(t_{2})v+h''(t_{2})) \end{array} $$
(44)

We remark that where \(t_{1}=t_{2}\), the function \(J'\) has a priori a simple zero \((t_{1}-t_{2})\) along the diagonal subset. However the jacobian of the transformation which maps \((t_{1},t_{2})\) to \((\sigma _{1},\sigma _{2})\) is equal to \(t_{1}-t_{2}\), so the jacobian of the mapping F in the coordinates \(u,v\) and \(\sigma _{1},\sigma _{2}\) is given by the regular symmetric function

$$\begin{array}{rll} J(\sigma_{1},\sigma_{2})&=&[a(t_{1}-t_{2})(g'(t_{2})-g'(t_{1}))]^{-1}\\ &&\quad \times (g''(t_{1})v+h''(t_{1}))(g''(t_{2})v+h''(t_{2}))\\ \end{array} $$
(45)

that is in general finite and non-zero along the diagonal.

Let us suppose that J is a smooth function without zero in an open set of \((t_{1},t_{2})\) containing (0,0). By substitution of the values of u and v we obtain the following functional equation for g and h:

$$\begin{array}{lll} &&[g''(t_{1})(h'(t_{1})-h'(t_{2}))+h''(t_{1})(g'(t_{2})-g'(t_{1}))]\\ &&[g''(t_{2})(h'(t_{1})-h'(t_{2}))+h''(t_{2})(g'(t_{2})-g'(t_{1}))]\\&&\quad=J(t_{1},t_{2})a(t_{1}-t_{2})(g'(t_{2})-g'(t_{1}))^{3},\end{array}$$
(46)

Let us suppose that \(t_{1}\) and \(t_{2}\) both approach t, in such a manner that \(t=\alpha t_{1}+\beta t_{2}\) with \(\alpha ,\beta \) bounded and \(\alpha +\beta =1\). Up to the order 4 in \(t_{2}-t_{1}\) (or \(t_{1}-t\) or \(t_{2}-t\) because it is the same order) we have

$$\begin{array}{lll}&&[((t_{1}-t)(t_{1}-t_{2})+\frac{1}{2}(t_{2}-t_{1})(t_{1}+t_{2}-2t))h''(t)g^{(3)}(t))\\&&\;\;+((t_{1}-t)(t_{2}-t_{1})+\frac{1}{2}(t_{1}-t_{2})(t_{1}+t_{2}-2t))h^{(3)}(t)g''(t))]\\ &&[((t_{2}-t)(t_{1}-t_{2})+\frac{1}{2}(t_{2}-t_{1})(t_{1}+t_{2}-2t))h''(t)g^{(3)}(t))\\&&\;\;+((t_{2}-t)(t_{2}-t_{1})+\frac{1}{2}(t_{1}-t_{2})(t_{1}+t_{2}-2t))h^{(3)}(t)g''(t))]\\&&\qquad=-a J(t,t)(t_{1}-t_{2})^{4}g''(t)^{3},\end{array}$$
(47)

We also have \(t_{1}-t=\beta (t_{1}-t_{2}), t_{2}-t=\alpha (t_{2}-t_{1}), t_{1}+t_{2}-2t=(2\beta -1)(t_{1}-t_{2})\), then, at the limit, we get

$$\begin{array}{lll}&&\left[\frac{1}{2}h''(t)g^{(3)}(t))-\frac{1}{2}h^{(3)}(t)g''(t))\right]\\&& \left[-\frac{1}{2}h''(t)g^{(3)}(t))+\frac{1}{2}h^{(3)}(t)g''(t))\right]\\&&\qquad\qquad\quad=-a J(t,t)g''(t)^{3}.\end{array}$$
(48)

We would have found the same formula by fixing \(t_{1}=t\) and looking at \(t_{2}\) tending to t.

Let us write \(G=g''\) and \(H=h''\), we obtain

$$ (G'H-GH')^{2}=4aJG^{3}, $$
(49)

From this we see that G cannot vanish, meaning that g is necessarily convex or concave. By changing the sign of the coordinate y in the ambient space, we can suppose that g is convex, then G is positive and J positive.

By using a translation and a rotation in the plane \(x,y\) we can suppose that \(g(0)=g'(0)=0\), but \(g''(0)=G(0)> 0\), which gives a development

$$ g(t)=\frac{1}{2}bt^{2}+\frac{1}{3}b_{1}t^{3}+... $$
(50)

When g is known as well as the initial condition \(H(0)=h''(0)\), the linear differential equation

$$ H'=(G'/G)H+2(aGJ)^{1/2}, $$
(51)

determines H as

$$ H(t)=G(t)\frac{H(0)}{G(0)}+2G(t)\int_{0}^{t}\sqrt{\frac{aJ(s)}{G(s)}}ds. $$
(52)

Then, knowing the values of \(h(0),h'(0)\) in addition to \(h''(0)\), we obtain h by two other quadratures.

With a vertical translation we can assume \(h(0)=0\), and with a rotation in the plane \(x,z\) we can assume \(h'(0)=0\). By rotating the plane \(y,z\) we can then make \(h''(0)=0\). Now we can apply the formula (52) and the Theorem 1 follows.

When \(f,g,h\) satisfy the conclusions of the Theorem 1, the axis \(x,y,z\) are respectively aligned with the tangent, the normal and the binormal axis of the curve S at the origin point 0. The curvature \(\kappa (0)\) and the torsion \(\tau (0)\) of S in 0 are given by

$$ \kappa(0)=\frac{g''(0)}{f'(0)^{2}},\quad \tau(0)=-\frac{h^{(3)}(0)}{f'(0)g''(0)}. $$
(53)

Let us verify by direct computation that every polynomials \(g,h\) of respective degree 2 and 3 satisfy the formula (46) for all points \((t_{1},t_{2})\) with J a non-zero constant:We can write

$$\begin{array}{rll}f(t)&=&at,\; g(t)\\ &=&\frac{b}{2}t^{2}+b_{1}t+b_{0},\quad h(t)=\frac{c}{3}t^{3}+c_{2}t^{2}+c_{1}t+c_{0},\end{array} $$
(54)

then we have

$$g'(t_{2})-g'(t_{1})=b(t_{2}-t_{1}),$$
(55)
$$h'(t_{1})-h'(t_{2})=c(t_{1}-t_{2})(t_{2}+t_{1})+2c_{2}(t_{1}-t_{2}),$$
(56)
$$g''(t)=b,$$
(57)
$$h''(t)=2(ct+c_{2}),$$
(58)

then

$$g''(t_{1})(h'(t_{1}) - h'(t_{2})) + h''(t_{1})(g'{\kern-1pt}(t_{2})-g'(t_{1})) = -bc(t_{2}-t_{1})^{2},$$
(59)
$$g''(t_{2})(h'(t_{1})-h'(t_{2})) + h''(t_{2})(g'(t_{2})-g'(t_{1})) = bc(t_{2}-t_{1})^{2},$$
(60)
$$(t_{2}-t_{1})(g'(t_{2})-g'(t_{1}))^{3} = b^{3}(t_{2}-t_{1})^{4},$$
(61)

and Eq. (46) is satisfied with

$$ \frac{b^{2}c^{2}}{b^{3}}=aJ, $$
(62)

that is

$$ J=\frac{c^{2}}{ab}. $$
(63)

Theorem 2

Suppose that \(f,g,h\) satisfy the conclusions of the Theorem 1, and that f is a linear function of t; then J is constant in the neighborhood of the diagonal if and only if g is a polynomial of degree 2 and h is a polynomial of degree 3.

Proof

Let us fix the constant J and the parameter \(t_{2}\); the two Eqs. (46) and (49) constitute a system of two non autonomous ordinary differential equations in two unknown functions \(g',h'\) of order one in \(g'\) and order two in \(h'\), so by the Cauchy-Lipschitz theorem, the general solution depends on three independent real parameters when the initial time is chosen, thus the general solution of the system when initial time is let free depends on four real constants. But we already know solutions of the system depending on four real parameters, they are polynomials in t of degree 1 for \(g'\) and of degree 2 for \(h'\). CQFD.

Thus, in the case of Theorem 1, we obtain a semi-cubic parabola

$$ x=at,\quad y=\frac{1}{2}a^{2}\kappa(0)t^{2},\quad z=-\frac{1}{6}a^{3}\kappa(0)\tau(0)t^{3}. $$
(64)

1.2 A.2 Euclidian optimal striola

The natural measure on directions in a 3D Euclidian space is the solid angle \(d\sigma \), and the natural parametrization of a curve is by its arc-length s. Therefore it is natural to look at the correspondence between pairs of points and directions in these coordinates. We will first establish a formula for the Jacobian determinant in these coordinates. As before we will constat it has a simple zero. As in the Cartesian study, the symmetric coordinates operate a regularization. But the new ingredient here is the surface obtained by translating the curve by itself, that was considered by Sophus Lie, for application to minimal surfaces. The points on this auxiliary surface form another symmetric parametrization of pairs of points, giving another natural regularization.

Let us introduce a surface \(\Sigma \) associated to the curve S, it is formed by adding the pair of points of the curve:

$$x(s_{1},s_{2})=f(s_{1})+f(s_{2}),$$
(65)
$$y(s_{1},s_{2})=g(s_{1})+g(s_{2}),$$
(66)
$$z(s_{1},s_{2})=h(s_{1})+h(s_{2}).$$
(67)

Let us denote by \(P_{1},P_{2}\) the points of abscissas \(s_{1},s_{2}\) respectively on the curve S. By fixing \(s_{2}\) (resp. \(s_{1}\)) we see that the tangent vector \(\overrightarrow {T_{1}}\) (resp. \(\overrightarrow {T_{2}}\))to S in \(P_{1}\) (resp. \(P_{2}\)) is tangent to the surface \(\Sigma \). Then the cross product \(\overrightarrow {T_{1}}\times \overrightarrow {T_{2}}\) is normal to \(\Sigma \). Thus for generic points, when \(\overrightarrow {T_{1}}\) and \(\overrightarrow {T_{2}}\) are not collinear, the unit normal vector \(\overrightarrow {n}\) to \(\Sigma \) at the point corresponding to \((s_{1},s_{2})\) is parallel to the intersection of the normal planes at S in \(P_{1}\) and \(P_{2}\). This is the way \(\Sigma \) represents the pair of points \((P_{1},P_{2})\) on S, by developing their associated direction D in the space.

This surface \(\Sigma \) is a particular case or a “Translation Surface”, in the sense of S. Lie, H. Poincaré, G. Darboux (Leçons sur la Théorie gé né rale des Surfaces et les applications gé omé triques du Calcul Infinité simal; Part I, L.III, ch. VI, VII, VIII, IX; (Darboux 1887)). The general translation surfaces are the surfaces described by the sum of points of two given space curves. In our case, the curve is translated by itself.

Note that the surface which is defined as the locus of middle points of the segments that join the pairs of points of the curve S, is the surface homothetic to \(\Sigma \) in the ratio \(1/2\), then it has the same normals and could be used in the place of \(\Sigma \) in what follows.

According to Lie, by using imaginary parameters, any minimal surface is described as a translation surface starting from a complex curve with tangent vectors of zero lengths.

Caution: the surface \(\Sigma \) does not represent the macula, it represents the directions of acceleration detected by the striola by its normal vectors. It is a kind of virtual macula associated to the striola. The parametrization by \(s_{1},s_{2}\) is singular, because two pairs \((s_{1},s_{2})\) and \((s_{2},s_{1})\) represent the same point, but the surface \(\Sigma \) itself is in general regular. Let us see what happens in the example of the surface generated by the twisted cubic; in this case we have:

$$x(t_{1},t_{2})=t_{1}+t_{2},$$
(68)
$$y(t_{1},t_{2})=t_{1}^{2}+t_{2}^{2},$$
(69)
$$z(t_{1},t_{2})=t_{1}^{3}+t_{2}^{3}.$$
(70)

We then have

$$ z-xy=t_{1}t_{2}^{2}+t_{2}t_{1}^{2}=x(t_{1}t_{2})=\frac{1}{2}x(x^{2}-y), $$
(71)

so \(\Sigma \) is included in the smooth graph of a polynomial function:

$$ z=\frac{1}{2}x(y+x^{2}). $$
(72)

And it is easy to show that every point \((x,y,z)\) in this graph such that \(y\geq x^{2}/2\) belongs to \(\Sigma \). Again the parametrization problem comes from the symmetry between \(t_{1}\) and \(t_{2}\), but the regular prolongation comes from the fact that every symmetric polynomial of \(t_{1},t_{2}\), like \(t_{1}^{3}+t_{2}^{3}\), is a function of \(x=t_{1}+t_{2}\) and \(y=t_{1}^{2}+t_{2}^{2}\). (This theorem needs rational coefficients, integers are not sufficient are they were for \(t_{1}+t_{2}\) and \(t_{1}t_{2}\), but this has no consequence in the present context.) As the example of twisted cubic shows, the set of directions that are detected by real parameters \(t_{1},t_{2}\) is the set of normals of a regular surface with boundary.

For the general twisted cubic

$$ x=at,\quad y=\frac{b}{2}t^{2},\quad z=\frac{c}{3}t^{3} $$
(73)

we find the region \(y\geq bx^{2}/4a^{2}\) on the gather surface

$$ z=\frac{c}{6a^{3}b}(bx^{3}+2a^{2}xy). $$
(74)

In what follows, the numbers \(s_{1},s_{2}\) will design arc-lengths on the curve S. We denote by \(\overrightarrow {T_{1}}\) resp. \(\overrightarrow {T_{2}}\) the unit tangent to the curve S at \(s_{1}\) resp. \(s_{2}\). Let us suppose they are linearly independent and denote by \(\overrightarrow {n}\) the unit normal to the plane they generate, such that \((\overrightarrow {T_{1}},\overrightarrow {T_{2}},\overrightarrow {n})\) constitute a right handed oriented reference frame, which corresponds to the vector product formula: \(|\sin (arcos(\overrightarrow {T_{1}}.\overrightarrow {T_{2}}))|\overrightarrow {n}=\overrightarrow {T_{1}}\times \overrightarrow {T_{2}}\).

For any surface in \(\mathbb {R}^{3}\) the Gauss mapping is the map to the unit sphere, which sends a point to the unit normal of the surface in this point, thus in our case the Gauss mapping \(\gamma \) sends the point of \(\Sigma \) corresponding to \(s_{1},s_{2}\) onto the vector \(\overrightarrow {n}\). This map is specially interesting for us because \(\overrightarrow {n}\) represents the preferred direction of acceleration of an afferent cell associated to the parameters \((s_{1},s_{2})\).

By a theorem of Gauss, the determinant of the jacobian of \(\gamma \) from orthonormal coordinates on \(\Sigma \) to solid angle is the Gauss curvature function K of the surface \(\Sigma \).

The Riemannian metric on \(\Sigma \) has coefficients \(g_{ij}\) defined by the scalar products \(\overrightarrow {T_{i}}.\overrightarrow {T_{j}}\), cf. Coxeter, Introduction to Geometry (Wiley, 1969). In our case \(g_{11}=g_{22}=1\). The superficial area measure is given in these coordinates by the formula

$$ dA=\sqrt{g}ds_{1}ds_{2}, $$
(75)

where we noted

$$ g=g_{11}g_{22}-g_{12}^{2}=1-(\overrightarrow{T_{1}}.\overrightarrow{T_{2}})^{2} $$
(76)

Observe that the two families of curves respectively defined by fixing \(s_{1}+ s_{2}\) or \(s_{1}-s_{2}\) equal to constants, are orthogonal one to each other.

We denote by \(g^{ij}\) the coefficients of the inverse matrix of \((g_{ij})\); they are

$$ g^{11}=g^{22}=g^{-1},\quad g^{12}=g^{21}=-g^{-1}g_{12}. $$
(77)

The first Christoffel symbols (describing the intrinsic parallel transport on the surface) are defined as

$$ \Gamma_{ij;k}=\frac{1}{2}(\partial_{i}g_{jk}+\partial_{j}g_{ik}-\partial_{k}g_{ij}). $$
(78)

In our case the only non-zero symbols are

$$ \Gamma_{11;2}=\partial_{1}(g_{12}),\quad\Gamma_{22;1}=\partial_{2}(g_{12}). $$
(79)

The second (and more usual) Christoffel symbols are defined as

$$ \Gamma^{k}_{ij}=\sum g^{kl}\Gamma_{ij;k}. $$
(80)

Thus in our case they are non-zero only if \(i=j\). All the non-zero symbols are given by:

$$\Gamma_{11}^{1}=\frac{-g_{12}\partial_{1}(g_{12})}{g},\quad\Gamma_{11}^{2}=\frac{\partial_{1}(g_{12})}{g},$$
(81)
$$\Gamma_{22}^{1}=\frac{\partial_{2}(g_{12})}{g},\quad\Gamma_{22}^{2}=\frac{-g_{12}\partial_{2}(g_{12})}{g}, $$
(82)

It is useful, in our case, to relate these intrinsic quantities with the parametric representation; for that we introduce the second derivatives vectors

$$ \overrightarrow{T_{ij}}=\frac{\partial}{\partial s_{j}}\overrightarrow{T_{i}}; $$
(83)

Then the Christoffel symbols are given by scalar products as

$$ \Gamma_{ij;k}=\overrightarrow{T_{k}}.\overrightarrow{T_{ij}} $$
(84)

We can also define

$$ b_{ij}=\overrightarrow{n}.\overrightarrow{T_{ij}}. $$
(85)

They are the coefficients of the second fundamental form of \(\Sigma \). In our present case only \(b_{11}\) and \(b_{22}\) could be non-zero. The corresponding Weingarten coefficients are

$$ b^{i}_{j}=\sum g^{ik}b_{kj}, $$
(86)

thus, in our case

$$ b^{1}_{1}=g^{-1}b_{11},\quad b^{2}_{2}=g^{-1}b_{22},\quad b^{1}_{2}=b^{2}_{1}=-g^{-1}g_{12}b_{12}. $$
(87)

Then we obtain the “equations of Gauss”:

$$ \overrightarrow{T_{ij}}=\sum \Gamma^{k}_{ij}\overrightarrow{T_{k}}+b_{ij}\overrightarrow{n}. $$
(88)

If we denote by b the discriminant of the second fundamental form

$$ b=b_{11}b_{22}-b_{12}^{2}, $$
(89)

the Gauss curvature is defined by

$$ K=\frac{b}{g}. $$
(90)

The Theorema egregium of Gauss says that K can be expressed by the derivatives of the second symbols. A nice formula was found by Liouville (cf. Coxeter p.367):

$$ \sqrt{g}K=\frac{\partial}{\partial s_{2}}\left(\frac{\sqrt{g}}{g_{11}}\Gamma_{11}^{2}\right) -\frac{\partial}{\partial s_{1}}\left(\frac{\sqrt{g}}{g_{11}}\Gamma_{12}^{2}\right) $$
(91)

Note, in our case we have \(\Gamma _{12}^{2}=0\), and

$$ \Gamma_{11}^{2}=\frac{\overrightarrow{T_{11}}.\overrightarrow{T_{2}}}{g}. $$
(92)

(This is because \(\overrightarrow {T_{11}}\) is orthogonal to \(\overrightarrow {T_{1}}\).)

The Gauss map has the curvature as its jacobian determinant:

$$ \frac{d\sigma}{\sqrt{g}ds_{1}ds_{2}}=K, $$
(93)

Thus the jacobian determinant from \(ds_{1}ds_{2}\) to \(d\sigma \) is given by

$$ J'=\sqrt{g}K=\frac{\partial}{\partial s_{2}}\frac{\overrightarrow{T_{11}}.\overrightarrow{T_{2}}}{\sqrt{g}} $$
(94)

Which can be rewritten as

$$ J'=\frac{(\overrightarrow{T_{1}}.\overrightarrow{T_{2}})(\overrightarrow{T_{11}}.\overrightarrow{T_{2}}) (\overrightarrow{T_{22}}.\overrightarrow{T_{1}}) +g(\overrightarrow{T_{11}}.\overrightarrow{T_{22}})}{g^{3/2}} $$
(95)

From Section A.1 we know that \(J'\) has a simple zero along the diagonal (of the first order like \((s_{1}-s_{2})\) multiplied by a constant). This invitees to make a re-normalization of \(J'\). A natural choice is to divide by \(s_{1}-s_{2}\) because this is the jacobian of the transformation from the coordinates \((s_{1},s_{2})\) to the elementary symmetric function \(\sigma _{1}=s_{1}+s_{2}, \sigma _{2}=s_{1}s_{2}\). But also \(\sqrt {g}\) has a first order zero on the diagonal \(s_{1}=s_{2}\), then an alternative natural choice to re-normalize \(J'\) is to take the curvature K itself. This corresponds to the representation of the set of pairs of points on S by the surface \(\Sigma \) itself. With this second choice, the optimality problem becomes: to find the curves S such that the associated double surface of translation \(\Sigma \) has the less possible changing Gaussian curvature. We will compute now this curvature K, and the density \(\sqrt {g}\).

The jacobian of the mapping from affine coordinates \((u,v,1)\) on directions to the spherical angle measure is given by

$$ J(u,v;\sigma)=\frac{1}{\sqrt{1+u^{2}+v^{2}}}, $$
(96)

where \(u,v\) are given in function of \(s_{1},s_{2}\) by the formulas

$$u=\frac{g_{1}h_{2}-h_{1}g_{2}}{f_{1}g_{2}-g_{1}f_{2}}$$
(97)
$$v=\frac{h_{1}f_{2}-f_{1}h_{2}}{f_{1}g_{2}-g_{1}f_{2}}$$
(98)

For computation another formula is helpful, coming from Weingarten (see Coxeter 1969 p.354):

$$ \sqrt{g}K=\left[\overrightarrow{n};\partial_{1}\overrightarrow{n};\partial_{2}\overrightarrow{n}\right], $$
(99)

where the bracket denotes the determinant of vectors, or triple product. In fact this Weingarten formula is the origin of the recalled Liouville formula. In our present situation the formula can be rewritten by using vector products as follows:

$$ g^{2}K=\left[\overrightarrow{T_{1}}\times \overrightarrow{T_{2}}; \overrightarrow{T_{11}} \times \overrightarrow{T_{2}}; \overrightarrow{T_{1}}\times \overrightarrow{T_{22}}\right] $$
(100)

Theorem 3

Let us denote by \(\kappa ,\tau \) the curvature and torsion of S, and \(\dot {\kappa },\dot {\tau }\) their derivatives with respect to the arc-length s on S; then along the double curve S on \(\Sigma \) we have the limit formula

$$ K=-\frac{\tau^{2}}{4}. $$
(101)

At the first order in \(s_{1}-s,s_{2}-s\), we have

$$ K(s_{1},s_{2})=-\frac{\tau^{2}}{4}\big[1+(s_{1}+s_{2}-2s)(\dot{\tau}/\tau)\big], $$
(102)

and

$$ g(s_{1},s_{2})=\kappa^{2}(s_{1}-s_{2})^{2}\big[1+(s_{1}+s_{2}-2s)(\dot{\kappa}/\kappa)\big]. $$
(103)

Proof

We use Taylor formula for evaluating \(\overrightarrow {t_{1}},\overrightarrow {t_{2}}\), and recurrently all terms in the Eq. (100). (Note that we verified this result by using Eq. (90). The proof is a straightforward but tedious computation. As a corollary we obtain the amazing result that for twisted curves, when \(\tau \) has no zero, the Gaussian curvature of the double surface is everywhere finite and strictly negative.

Theorem 3 is a precision of the Theorem 1, because it shows directly that a curve giving good vector detection in space must be twisted, but it uses the fact that curvature and torsion are well defined, which is one of the results of Theorem 1.

If we choose the surface \(\Sigma \) as the natural parametrization of pairs of points \((P_{1},P_{2})\) on S, the absolute value of the Jacobian determinant from pairs of points to directions in space is approximatively given by

$$ |K|\approx \frac{1}{4}\tau^{2}(1+\frac{d(\ln (\tau))}{ds}(\sigma_{1}-2s)) $$
(104)

for near points, thus, at the first order of approximation, the optimal curves for uniform detection of directions must have a constant torsion \(\tau \). But, if we choose the elementary symmetric functions \((\sigma _{1},\sigma _{2})\) as the natural parametrization of pairs of points \((P_{1},P_{2})\) on S, the absolute value of the Jacobian determinant for near points is given by

$$ |J|= \sqrt{g}|K|/|s_{1}-s_{2}|\approx\frac{1}{4}\kappa\tau^{2}(1+\frac{1}{2}\frac{d(\ln (\kappa\tau^{2}))}{ds}(\sigma_{1}-2s)). $$
(105)

Thus the optimal curves correspond to constant value of the quantity \(\tau \sqrt {\kappa }\).

At higher orders in the formula of K it appears polynomials in the two variables \(s_{1}-s_{2},s_{1}+s_{2}-2s\) whose coefficients are polynomials in the higher derivatives of \(\kappa ,\tau \) with respect to s. At the order two for instance we obtain two coefficients of the following form

$$\begin{array}{lll}&&P(\kappa,\tau,\dot{\kappa},\dot{\tau},\ddot{\kappa},\ddot{\tau}) \\ &&\quad\quad=\kappa^{2}\tau^{2}\big[A\ddot{\kappa}/\kappa+B\ddot{\tau}/\tau+C(\dot{\kappa}/\kappa)^{2}+D(\dot{\tau}/\tau)^{2} \\&&\qquad\qquad\qquad +E(\dot{\kappa}\dot{\tau}/\kappa\tau)+F(\kappa^{2}+\tau^{2})\big]\end{array} $$
(106)

In particular, the only possibility that the order two computation of K gives a development \(-\frac {\tau ^{2}}{4}\big [1+\alpha (s_{1}-s_{2})^{2}\big ]\) with a constant \(\alpha \), is that \(\kappa \) and \(\tau \) are constants. That is S must be a circular helix.

The standard circular helix in the vertical cylinder of radius a is defined as

$$ x=a \cos t,\quad y=a\sin t,\quad z=ct. $$
(107)

One has

$$ \kappa=\frac{a}{a^{2}+c^{2}},\quad \tau=\frac{c}{a^{2}+c^{2}}, $$
(108)
$$ a=\frac{\kappa}{\kappa^{2}+\tau^{2}}. $$
(109)

(cf. Coxeter pp.323-325). The principal normal \(\overrightarrow {p}=\dot {\overrightarrow {T}}/\kappa \) is always horizontal and its binormal \(\textbf {b}\), as its tangent \(\overrightarrow {T}\), makes a constant angle with the vertical. This binormal vector \(\overrightarrow {b}\) is important for us because it is the limit for coincident points \(P_{1},P_{2}\) of the normal vector \(\overrightarrow {n}\) to the surface \(\Sigma \).

In the case of helix we have the following formula for the curvature of \(\Sigma \):

$$ K=-\frac{\tau^{2}}{4}\left[1+a\frac{\kappa}{2}(\cos(u_{1}-u_{2})-1)\right]^{-2}. $$
(110)

In perception, some acceleration directions \(\overrightarrow {A}\) could be more natural than others. But, on directions in space, the 2D spherical Von-Mises distributions have maximum entropy among the laws when mean and variances are given; then it would be interesting to look at curves giving approximatively such laws for the preferred directions.

There is also a notion of generalized helix (cf. Coxeter p.325). They are the curves where the ration \(\tau /\kappa \) has a constant value. They are characterized by the fact that their tangent vectors make a constant angle with a fixed vector. The standard twisted cubic is not a circular helix, but a linear transform of it is a generalized helix, that is \(\tau /\kappa \) is a constant, and it has near its vertex a torsion and a curvature which are stationary. In fact let us consider the following special twisted cubic:

$$ x=\frac{t}{\sqrt{2}},\quad y=t^{2},\quad z=\frac{\sqrt{2}}{3}t^{3}; $$
(111)

then its curvature and torsion are given by

$$ \kappa=\tau=\frac{1}{(1+t^{2})^{2}} $$
(112)

(cf. Coxeter, exercise 17.8.2.) Note that a usual name for the twisted cubic (or rational normal curve) is rational helix. □

1.3 A.3 Spherical optimal striola

Our computational models of striola (cf. Methods) belong to a spherical surface M representing the macula. The normal vector \(\overrightarrow {N}\) along the curve S which represents the polarization of hair cells disposed along S is the normal to the tangent vector \(\overrightarrow {T}\) which is also tangent to the sphere M. The discussion in Section A.2 invites to consider as optimal the spherical curves with constant regularized jacobian determinant. We saw two natural choices; the first one, with algebraic coordinates as \((\sigma _{1}=s_{1}+s_{2}, \sigma _{2}=s_{1}s_{2})\) or any regular functions of these coordinates, gives a constant product \(\kappa \tau ^{2}\); the second one with geometric parametrization by \(\Sigma \), gives more simply a constant torsion \(\tau \).

We suppose that M is part of the sphere in the \(3D\) Euclidian space, centered in 0 of radius R. The variable point on the curve S is written now \(X(s)\), where s denotes a curvilinear abscissa on S. By definition, we have

$$ \dot{X}=\overrightarrow{T},\quad \dot{\overrightarrow{T}}=\kappa \overrightarrow{p},\quad \dot{\overrightarrow{p}}=-\kappa \overrightarrow{T}+\tau \overrightarrow{b}. $$
(113)

Let us denote \(C=1/\kappa \)\(D=1/\tau \) the radius of curvature and the radius of torsion respectively. We have

$$ \overrightarrow{p}=C\ddot{X},\quad \dot{\overrightarrow{p}}=C\dddot{X}+\dot{C}\ddot{X}, $$
(114)

thus

$$ \overrightarrow{b}=\frac{1}{\tau}\dot{\overrightarrow{p}}+\frac{\kappa}{\tau}\overrightarrow{T} =DC\dddot{X}+D\dot{C}\ddot{X}+DC^{-1}\dot{X} $$
(115)

That gives the following formulas for the tangent, normal and binormal:

$$ \overrightarrow{T}=\dot{X},\quad \overrightarrow{p}=C\ddot{X},\quad \overrightarrow{b}=DC\dddot{X}+D\dot{C}\ddot{X}+DC^{-1}\dot{X} $$
(116)

As \(X.X\) is constant equal to \(R^{2}\), we have \(X.\overrightarrow {T}=0\), and by derivation of this identity we get \(\overrightarrow {T}.\overrightarrow {T}+\kappa X.\overrightarrow {p}=0\). But \(\overrightarrow {T}\) has a norm equal to 1, thus we get \(X.\overrightarrow {p}=-C\), or equivalently \(X.\ddot {X}=-1\). By derivation again, we get \(\dot {X}.\ddot {X}+X.\dddot {X}=0\), then \(X.\dddot {X}=0\) because \(\overrightarrow {T}\) and \(\overrightarrow {p}\) are orthogonal. Then the scalar product \(X.\overrightarrow {b}\) is equal to \(D\dot {C}X.\ddot {X}\) that is \(-D\dot {C}\). Thus

$$ X=-C\overrightarrow{p}-D\dot{C}\overrightarrow{b}, $$
(117)

and by equating the norm of X to R, this gives the well known intrinsic equation for spherical curves:

$$ C^{2}+D^{2}\dot{C}^{2}=R^{2}. $$
(118)

When D is constant, this equation has for only solutions

$$ C(s)=R\cos\left(\frac{s-s_{0}}{D}\right). $$
(119)

When \(D\sqrt {C}\) is constant, equals to A, this equation has for solutions

$$ C(s)=R.cnl^{2}\left(\frac{R^{1/2}(s-s_{0})}{2A}\right), $$
(120)

where cnl denotes the lemniscate elliptic function coslemn, that inverses the lemniscate elliptic integral of Bernoulli and Fagnano:

$$ I(u)=\int_{u}^{1}\frac{dt}{\sqrt{1-t^{4}}} $$
(121)

With the Jacobi-Guderman notations we have \(cnl(x)=cn(x\sqrt {2},1/\sqrt {2})\). (A nice reference for elliptic functions is Greenhill (1892))

The associated functions \(\overrightarrow {T},\overrightarrow {p},\overrightarrow {b},X\) can easily be deduced from the solutions of the following two by two linear differential system:

$$ \frac{dY(s)}{ds}=\frac{i}{2}\begin{pmatrix} \kappa (s)& \tau(s)\\ -\tau(s)&-\kappa (s) \end{pmatrix} Y(s) $$
(122)

In fact, if \(Y(s)\) is a resolvent matrix of this system, and if we introduce the imaginary Pauli matrices

$$ e_{1}{} ={}i\sigma_{1}{}={} \begin{pmatrix} 0& i\\ i&0 \end{pmatrix}\; e_{2}{}=i\sigma_{2}{}={} \begin{pmatrix} 0& 1\\ -1&0 \end{pmatrix}\; e_{3}{}=i\sigma_{3}{}={}\begin{pmatrix} i& 0\\ 0&-i \end{pmatrix} $$
(123)

satisfying

$$ e_{1}e_{2}-e_{2}e_{1}=-2e_{3},\; e_{2}e_{3}-e_{3}e_{2}=-2e_{1},\; e_{3}e_{1}-e_{1}e_{3}=-2e_{2}, $$
(124)

then the matrices

$$ \frac{1}{2}Ye_{1}Y^{-1}, \quad \frac{1}{2}Ye_{2}Y^{-1}, \quad\frac{1}{2}Ye_{3}Y^{-1}, $$
(125)

decomposed in the basis \(e_{1},e_{2},e_{3}\) give respectively the vectorial functions \(\overrightarrow {T}(s),\overrightarrow {p}(s),\overrightarrow {b}(s)\). Then the curve \(X(s)\) can be deduced by one quadrature.

However there is no known explicit formula for \(Y(s)\) when \(\kappa \) and \(\tau \) correspond to the particular curvature function \(C(s)\) we found before. We can only compute their Taylor polynomials.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Dimiccoli, M., Girard, B., Berthoz, A. et al. Striola magica. A functional explanation of otolith geometry. J Comput Neurosci 35, 125–154 (2013). https://doi.org/10.1007/s10827-013-0444-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-013-0444-x

Keywords

Navigation