Skip to main content

On the Duration of Human Movement: From Self-paced to Slow/Fast Reaches up to Fitts’s Law

  • Chapter
  • First Online:

Part of the book series: Springer Tracts in Advanced Robotics ((STAR,volume 117))

Abstract

In this chapter, we present a mathematical theory of human movement vigor. At the core of the theory is the concept of the cost of time. According to it, natural movement cannot be too slow because the passage of time entails a cost which makes slow moves undesirable. Within this framework, an inverse methodology is available to reliably and robustly characterize how the brain penalizes time from experimental motion data. Yet, a general theory of human movement pace should not only account for the self-selected speed but should also include situations where slow or fast speed instructions are given by an experimenter or required by a task. In particular, the limit case of a “maximal speed” instruction is linked to Fitts’s law, i.e. the speed/accuracy trade-off. This chapter first summarizes the cost of time theory and the procedure used for its accurate identification. Then, the case of slow/fast movements is investigated but changing the duration of goal-directed movements can be done in various ways in this framework. Here we show that only one strategy seems plausible to account for both slow/fast and self-paced reaching movements. By relying upon a free-time optimal control formulation of the motor planning problem, this chapter provides a comprehensive treatment of the linear-quadratic case for single degree of freedom arm movements but the principles are easily extendable to multijoint and/or artificial systems.

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

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD   169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Notes

  1. 1.

    Note that we did not use the standard way to define the value function: for a movement duration equal to t, this is usually \(\tilde{V}_{\mathbf {x}^{f}}(w,\mathbf {x}^{0}(w))=\inf {\int _{w}^{t}{\displaystyle l\big (\mathbf {x_{u}}(s),\mathbf {u}(s)\big )ds}}.\) Here we set \(V_{\mathbf {x}^{f}}(t-w,\mathbf {x}^{0}(w))=\tilde{V}_{\mathbf {x}^{f}}(w,\mathbf {x}^{0}(w))\), hence \(\frac{\partial V_{\mathbf {x}^{f}}}{\partial t}=-\frac{\partial \tilde{V}_{\mathbf {x}^{f}}}{\partial t}\).

  2. 2.

    We assume here that there are no abnormal extremals (an hypothesis which is satisfied in particular by controllable linear systems). As a consequence, it is not necessary to put a Lagrange multiplier in front of l in \(\mathscr {H}_{0}\).

References

  1. L.B. Bagesteiro, R.L. Sainburg, Handedness: dominant arm advantages in control of limb dynamics. J. Neurophysiol. 88(5), 2408–2421 (2002). doi:10.1152/jn.00901.2001

  2. D. Bennequin, R. Fuchs, A. Berthoz, T. Flash, Movement timing and invariance arise from several geometries. PLoS Comput. Biol. 5(7), e1000,426 (2009). doi:10.1371/journal.pcbi.1000426

  3. A. Berardelli, J.C. Rothwell, P.D. Thompson, M. Hallett, Pathophysiology of bradykinesia in parkinson’s disease. Brain 124(Pt 11), 2131–2146 (2001)

    Article  Google Scholar 

  4. B. Berret, F. Jean, Why don’t we move slower? the value of time in the neural control of action. J. Neurosci. 36(4), 1056–1070 (2016). doi:10.1523/JNEUROSCI.1921-15.2016

  5. B. Berret, C. Darlot, F. Jean, T. Pozzo, C. Papaxanthis, J.P. Gauthier, The inactivation principle: mathematical solutions minimizing the absolute work and biological implications for the planning of arm movements. PLoS Comput. Biol. 4(10), e1000,194 (2008). doi:10.1371/journal.pcbi.1000194

  6. B. Berret, J.P. Gauthier, C. Papaxanthis, How humans control arm movements. Proc. Steklov Inst. Math. 261, 44–58 (2008)

    Google Scholar 

  7. B. Berret, E. Chiovetto, F. Nori, T. Pozzo T, Evidence for composite cost functions in arm movement planning: an inverse optimal control approach. PLoS Comput. Biol. 7(10), e1002,183 (2011). doi:10.1371/journal.pcbi.1002183

  8. J.M.M. Brown, W. Gilleard, Transition from slow to ballistic movement: development of triphasic electromyogram patterns. Eur. J. Appl. Physiol. Occup. Physiol. 63(5), 381–386 (1991). doi:10.1007/BF00364466

  9. S.H. Brown, H. Hefter, M. Mertens, H.J. Freund, Disturbances in human arm movement trajectory due to mild cerebellar dysfunction. J. Neurol. Neurosurg. Psychiatry 53(4), 306–313 (1990)

    Article  Google Scholar 

  10. S. Card, T. Moran, A. Newell, The Psychology of Human-computer Interaction (L. Erlbaum Associates, Hillsdale, 1983)

    Google Scholar 

  11. M.M. Churchland, G. Santhanam, K.V. Shenoy, Preparatory activity in premotor and motor cortex reflects the speed of the upcoming reach. J. Neurophysiol. 96(6), 3130–3146 (2006). doi:10.1152/jn.00307.2006

  12. E.M. Connelly, A control model: an alternative interpretation of fitts’ law. Proc. Hum. Factors Ergon. Soc. Annu. Meet. 28(7), 625–628 (1984). doi:10.1177/154193128402800722

    Article  Google Scholar 

  13. M. Desmurget, S. Grafton, Forward modeling allows feedback control for fast reaching movements. Trends Cogn. Sci. 4(11), 423–431 (2000)

    Article  Google Scholar 

  14. J.A. Doeringer, N. Hogan, Intermittency in preplanned elbow movements persists in the absence of visual feedback. J. Neurophysiol. 80(4), 1787–1799 (1998)

    Google Scholar 

  15. D. Elliott, W.F. Helsen, R. Chua, A century later: Woodworth’s (1899) two-component model of goal-directed aiming. Psychol. Bull. 127(3), 342–357 (2001)

    Article  Google Scholar 

  16. A. Ferrante, G. Marro, L. Ntogramatzidis, A parametrization of the solutions of the finite-horizon lq problem with general cost and boundary conditions. Automatica 41, 1359–1366 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  17. P.M. Fitts, The information capacity of the human motor system in controlling the amplitude of movement. J. Exp. Psychol. 47(6), 381–391 (1954)

    Article  Google Scholar 

  18. T. Flash, N. Hogan, The coordination of arm movements: an experimentally confirmed mathematical model. J. Neurosci. 5(7), 1688–1703 (1985)

    Google Scholar 

  19. J.P. Gauthier, B. Berret, F. Jean, A biomechanical inactivation principle. Proc. Steklov Inst. Math. 268, 93–116 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  20. J. Gaveau, B. Berret, L. Demougeot, L. Fadiga, T. Pozzo, C. Papaxanthis, Energy-related optimal control accounts for gravitational load: comparing shoulder, elbow, and wrist rotations. J. Neurophysiol. 111(1), 4–16 (2014). doi:10.1152/jn.01029.2012

  21. R. Gentili, V. Cahouet, C. Papaxanthis, Motor planning of arm movements is direction-dependent in the gravity field. Neuroscience 145(1), 20–32 (2007). doi:10.1016/j.neuroscience.2006.11.035

  22. M. Hallett, C.D. Marsden, Ballistic flexion movements of the human thumb. J. Physiol. 294, 33–50 (1979)

    Article  Google Scholar 

  23. C.M. Harris, D.M. Wolpert, Signal-dependent noise determines motor planning. Nature 394(6695), 780–784 (1998). doi:10.1038/29528

  24. C.M. Harris, D.M. Wolpert, The main sequence of saccades optimizes speed-accuracy trade-off. Biol. Cybern. 95(1), 21–29 (2006). doi:10.1007/s00422-006-0064-x

  25. B. Hoff, A model of duration in normal and perturbed reaching movement. Biol. Cybern. 71, 481–488 (1994)

    Google Scholar 

  26. N. Hogan, An organizing principle for a class of voluntary movements. J. Neurosci. 4(11), 2745–2754 (1984)

    Google Scholar 

  27. C. Isenberg, B. Conrad, Kinematic properties of slow arm movements in parkinson’s disease. J. Neurol. 241(5), 323–330 (1994)

    Article  Google Scholar 

  28. M.T. Johnson, J.D. Coltz, T.J. Ebner, Encoding of target direction and speed during visual instruction and arm tracking in dorsal premotor and primary motor cortical neurons. Eur. J. Neurosci. 11(12), 4433–4445 (1999)

    Article  Google Scholar 

  29. H.J. Kappen, Optimal control theory and the linear bellman equation, in Bayesian Time Series Models, ed. by D. Barber, A.T. Cemgil, S. Chiappa (Cambridge University Press, Cambridge, 2011), pp. 363–387. http://dx.doi.org/10.1017/CBO9780511984679.018. Cambridge Books Online

  30. D.E. Kirk, Optimal Control Theory: An Introduction (Prentice-Hall, New Jersey, 1970)

    Google Scholar 

  31. E.B. Lee, L. Markus, Foundations of Optimal Control Theory (Wiley, New York, 1967)

    MATH  Google Scholar 

  32. C. MacKenzie, T. Iberall, The Grasping Hand, Advances in Psychology (North-Holland, London, 1994)

    Google Scholar 

  33. C.L. MacKenzie, R.G. Marteniuk, C. Dugas, D. Liske, B. Eickmeier, Three-dimensional movement trajectories in fitts’ task: implications for control. Q. J. Exp. Psychol. Sect. A 39(4), 629–647 (1987). doi:10.1080/14640748708401806

  34. I.S. MacKenzie, Fitts’ law as a research and design tool in human-computer interaction. Hum.-Comput. Interact. 7(1), 91–139 (1992)

    Article  Google Scholar 

  35. N. Mansard, O. Stasse, P. Evrard, A. Kheddar, A versatile generalized inverted kinematics implementation for collaborative working humanoid robots: the Stack of Tasks, in ICAR’09: International Conference on Advanced Robotics (Munich, Germany, 2009), pp 1–6. http://hal-lirmm.ccsd.cnrs.fr/lirmm-00796736

  36. P. Mazzoni, A. Hristova, J.W. Krakauer, Why don’t we move faster? Parkinson’s disease, movement vigor, and implicit motivation. J. Neurosci. 27(27), 7105–7116 (2007). doi:10.1523/JNEUROSCI.0264-07.2007

  37. O. Missenard, L. Fernandez, Moving faster while preserving accuracy. Neuroscience 197, 233–241 (2011). doi:10.1016/j.neuroscience.2011.09.020

  38. W.L. Nelson, Physical principles for economies of skilled movements. Biol. Cybern. 46(2), 135–147 (1983)

    Article  MATH  Google Scholar 

  39. F. Nori, R. Frezza, Linear optimal control problems and quadratic cost functions estimation, in 12th Mediterranean Conference on Control and Automation, MED’04 (Kusadasi, Aydin, Turkey, 2004)

    Google Scholar 

  40. U. Pattacini, F. Nori, L. Natale, G. Metta, G. Sandini, An experimental evaluation of a novel minimum-jerk cartesian controller for humanoid robots, in 2010 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS) (2010), pp 1668–1674. doi:10.1109/IROS.2010.5650851

  41. R. Plamondon, A.M. Alimi, Speed/accuracy trade-offs in target-directed movements. Behav. Brain Sci. 20, 279–303 (1997)

    Google Scholar 

  42. L.S. Pontryagin, V.G. Boltyanskii, R.V. Gamkrelidze, E.F. Mishchenko, The Mathematical Theory of Optimal Processes (Pergamon Press, New York, 1964)

    Google Scholar 

  43. N. Qian, Y. Jiang, Z.P. Jiang, P. Mazzoni, Movement duration, fitts’s law, and an infinite-horizon optimal feedback control model for biological motor systems. Neural. Comput. 25(3), 697–724 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  44. R. Shadmehr, Control of movements and temporal discounting of reward. Curr. Opin. Neurobiol. 20(6), 726–730 (2010). doi:10.1016/j.conb.2010.08.017

  45. R. Shadmehr, J.W. Krakauer, A computational neuroanatomy for motor control. Exp. Brain Res. 185(3), 359–381 (2008). doi:10.1007/s00221-008-1280-5

  46. R. Shadmehr, S. Mussa-Ivaldi, Biological Learning and Control (MIT Press, Cambridge, 2012)

    Google Scholar 

  47. R. Shadmehr, J.J. Orban de Xivry, M. Xu-Wilson, T.Y. Shih, Temporal discounting of reward and the cost of time in motor control. J. Neurosci. 30(31), 10,507–10,516 (2010). doi:10.1523/JNEUROSCI.1343-10.2010

  48. R. Stengel, Optimal Control and Estimation, Dover books on advanced mathematics (Dover Publications, Mineola, 1986)

    MATH  Google Scholar 

  49. H. Tanaka, J.W. Krakauer, N. Qian, An optimization principle for determining movement duration. J. Neurophysiol. 95(6), 3875–3886 (2006). doi:10.1152/jn.00751.2005

  50. E. Todorov, Optimality principles in sensorimotor control. Nat. Neurosci. 7(9), 907–915 (2004). doi:10.1038/nn1309

  51. E. Todorov, in Optimal control theory, Bayesian Brain: Probabilistic Approaches to Neural Coding, ed. by K. Doya (2006), pp. 269–298

    Google Scholar 

  52. E. Todorov, M.I. Jordan, Optimal feedback control as a theory of motor coordination. Nat. Neurosci. 5(11), 1226–1235 (2002). doi:10.1038/nn963

  53. J. Trommershäuser, L.T. Maloney, M.S. Landy, Decision making, movement planning and statistical decision theory. Trends Cogn. Sci. 12(8), 291–297 (2008). doi:10.1016/j.tics.2008.04.010

  54. R.S. Turner, M. Desmurget, Basal ganglia contributions to motor control: a vigorous tutor. Curr. Opin. Neurobiol. 20(6), 704–716 (2010). doi:10.1016/j.conb.2010.08.022

  55. Y. Uno, M. Kawato, R. Suzuki, Formation and control of optimal trajectory in human multijoint arm movement minimum torque-change model. Biol. Cybern. 61(2), 89–101 (1989)

    Article  Google Scholar 

  56. R.P.R.D. van der Wel, D. Sternad, D.A. Rosenbaum, Moving the arm at different rates: slow movements are avoided. J. Mot. Behav. 42(1), 29–36 (2010). doi:10.1080/00222890903267116

  57. D. Winter, Biomechanics and Motor Control of Human Movement (Wiley, New York, 1990)

    Google Scholar 

  58. S.J. Young, J. Pratt, T. Chau, Target-directed movements at a comfortable pace: movement duration and fitts’s law. J. Mot. Behav. 41(4), 339–346 (2009). doi:10.3200/JMBR.41.4.339-346

Download references

Acknowledgements

This work is supported by a public grant overseen by the French National Research Agency (ANR) as part of the “Investissement d’ Avenir” program, through the “iCODE Institute project” funded by the IDEX Paris-Saclay, ANR-11-IDEX-0003-02.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Frédéric Jean .

Editor information

Editors and Affiliations

Appendix: Technical Details

Appendix: Technical Details

1.1 Asymptotic Study

We describe in this asymptotic studies the behavior of the solutions of the linear quadratic model introduced in Sect. 2.2 when some parameters of the problem go to zero or infinity.

1.1.1 Asymptotic Study for Small/Large Time and Fixed Cost

Let us study the behavior of the optimal solutions when the final time \(\tau \) varies, the quadratic cost \(l(\mathbf {x},u)=u^{2}+\mathbf {x}^{T}Q\mathbf {x}+2\mathbf {x}^{T}Su\) and the terminal conditions \(\mathbf {x}^{0}=(\theta ^{0},0,\dots ,0)\), \(\mathbf {x}^{f}=0\) being fixed. For every \(\tau >0\) we denote by \(\mathbf {x}_{\tau }(t)= \left( \theta _{\tau }(t),\dots , \theta _{\tau }^{(n-1)}(t)\right) ,\) \(t\in [0,\tau ],\) the solution of the free-time OC problem in fixed time \(\tau \) whose expression is given by Eq. (6). Consider first the case of large times, that is the case where \(\tau \rightarrow \infty \). Remind that in this case \(e^{\tau A_{+}}\) and \(e^{-\tau A_{-}}\) tend to zero.

Lemma 4

When \(\tau \rightarrow \infty \), there holds

$$ \mathbf {x}_{\tau }(t)=e^{tA_{+}}\mathbf {x}^{0}+O(\Vert e^{\tau A_{+}}\mathbf {x}^{0}\Vert ). $$

As a consequence, there exists constants \(c,\alpha >0\), and \(\varepsilon \in (0,1)\), such that

$$ \theta _{\tau }(t)=c\theta ^{0}e^{-\alpha t}+O(e^{-\alpha \tau }),\qquad \hbox {for any }\,t\in [\varepsilon \tau ,\tau ]. $$

We thus recover a somewhat intuitive result: the solution of a LQ problem in fixed time converges to the solution of the same LQ problem with infinite horizon when the time goes to infinity (see Remark 2).

Proof

We deduce directly from the conditions of Eq. (7) the values of \(\mathbf {p}_{1}=\mathbf {p}_{1}(\tau )\) and \(\mathbf {p}_{2}=\mathbf {p}_{2}(\tau )\) in function of \(\mathbf {x}^{0}\). By putting these values into Eq. (6), we obtain

$$ \mathbf {x}_{\tau }(t)=e^{tA_{+}}\left( I-e^{-\tau A_{-}}e^{\tau A_{+}}\right) ^{-1}\mathbf {x^{0}}-e^{(t-\tau )A_{-}}e^{\tau A_{+}}\mathbf {x^{0}}, $$

which is of the form \(e^{tA_{+}}\mathbf {x}^{0}+O(\Vert e^{\tau A_{+}}\mathbf {x}^{0}\Vert )\) since \(A_{+}\) is stable and \(A_{-}\) is anti-stable. Now, \(e^{tA_{+}}\mathbf {x}^{0}\) is a function of t which can be written as a sum of decreasing exponential terms. Denoting by \(e^{-\alpha t}\) the less decreasing term in this sum, it appears that all other exponential terms in \(e^{tA_{+}}\mathbf {x}^{0}\) are negligible in front of \(e^{-\alpha \tau }\) for \(t/\tau \) not too small and we obtain the formula for \(\theta _{\tau }\) (note that in general \(\alpha =\min \left\{ -\mathfrak {R}(\lambda )\ :\ \lambda \ eigenvalue\ of\ A_{+}\right\} \)).   \(\square \)

Consider now the case of small times, i.e. the case where \(\tau \rightarrow 0\). In that case we can prove the following result.

Lemma 5

Let p(s) be the polynomial function of degree \(2n-1\) defined by \(\left( p(0), p'(0),\dots ,p^{(n-1)}(0)\right) =\mathbf {x}^{0}\) and \(\left( p(1),p'(1),\dots ,p^{(n-1)}(1)\right) =\mathbf {x}^{f}\). Then

$$ \theta _{\tau }(t)=p\left( \frac{t}{\tau }\right) +O(\tau ). $$

As a consequence, \(\theta _{\tau }(t)\approx p(\frac{t}{\tau })\) for small times \(\tau \): a change of the final time induces approximately a temporal rescaling of the solutions.

Remark 6

Note that since the terminal conditions are equilibriums, the polynomial \(p(\cdot )\) satisfies \(\dot{p}(t)=\dot{p}(1-t)\), which implies that the velocity profiles of \(\theta _{\tau }\) have an almost symmetric shape for small times \(\tau \). Indeed, the polynomial function \(\widetilde{p}(t)=\theta ^{0}-p(1-t)\) satisfies the same conditions at \(t=0\) and \(t=1\) as p(t), which implies by unicity of the solution that \(\widetilde{p}(t)=p(t)\), and so the conclusion.

Proof

Let us start with a preliminary remark on the optimal solution \(\theta _{\tau }\). On one hand, it follows from Eq. (6) that \(\theta _{\tau }(t)\) is an analytic function (i.e. it is equal to its Taylor series) which depends linearly on the vectors \(\mathbf {p}_{1}=\mathbf {p}_{1}(\tau )\) and \(\mathbf {p}_{2}=\mathbf {p}_{2}(\tau )\). Hence, all derivatives of \(\theta _{\tau }\) at 0 depend linearly on the pair \((\mathbf {p}_{1},\mathbf {p}_{2})\). On the other hand, due to the particular properties of the matrices \(A_{-},A_{+}\) (see [16, Lemma 1]), there is a one-to-one correspondence between \((\mathbf {p}_{1},\mathbf {p}_{2})\) and the 2n first derivatives of \(\theta _{\tau }\) at 0, i.e. by \(\theta _{\tau }^{(k)}(0),\ 0\le k\le 2n-1\). As a consequence, all derivatives of \(\theta _{\tau }\) at 0 depend linearly on the 2n first ones: for every integer k there exists a constant \(C_{k}\) such that, for any \(\tau ,\) \(|\theta _{\tau }^{(k)}(0)|\le C_{k}\varTheta _{\tau },\) where

$$ \varTheta _{\tau }=\max \left\{ |\theta _{\tau }^{(k)}(0)|,\ 0\le k\le 2n-1\right\} . $$

Set \(\phi _{\tau }(t)=\theta _{\tau }(t)-p(\frac{t}{\tau })\). We have to prove that \(\phi _{\tau }(t)=O(\tau )\). The above remark and the fact that \(\left( p(0),\dots ,p^{(n-1)}(0)\right) =\left( \theta _{\tau }(0),\dots ,\theta _{\tau }^{(n-1)}(0)\right) =(\theta ^{0},0,\dots ,0)\) imply that the Taylor expansion of \(\phi _{\tau }\) has the form,

$$\begin{aligned} \phi _{\tau }(t)=\sum _{k=n}^{2n-1}\frac{t^{k}}{k!}\left( \theta _{\tau }^{(k)}(0)-\frac{p^{(k)}(0)}{\tau ^{k}}\right) +\varTheta _{\tau }O(t^{2n}), \end{aligned}$$
(11)

where all \(O(\cdot )\) are uniform with respect to \(\tau .\) By definition of \(p(\cdot )\) we have also \(\phi _{\tau }^{(j)}(\tau )=0\) for \(j=0,\dots ,n-1\), and from Eq. (11) we get

$$ \sum _{k=n}^{2n-1}\frac{1}{(k-j)!}\left( \tau ^{k}\theta _{\tau }^{(k)}(0)-p^{(k)}(0)\right) =\varTheta _{\tau }O(\tau ^{2n}),\qquad j=0,\dots ,n-1. $$

It follows that, for \(k=n,\dots ,2n-1\) there holds \(\tau ^{k}\theta _{\tau }^{(k)}(0)-p^{(k)}(0)=\varTheta _{\tau }O(\tau ^{2n})\), and thus from the definition of \(\varTheta _{\tau }\) we obtain that \(\varTheta _{\tau }\tau ^{2n}=O(\tau )\). This and Eq. (11) give \(\phi _{\tau }(t)=O(\tau )\), which proves the lemma.   \(\square \)

1.1.2 Asymptotic Study for Fixed Time

Let us try to understand now how the optimal solutions behave when some coefficients in the cost function are modified. We fix an initial state \(\mathbf {x}^{0}=(\theta ^{0},0,\dots ,0)\), a final one \(\mathbf {x}^{f}=0\), and an infinitesimal CoT g(t). We consider a family of costs \(l_{r}(\mathbf {x},u)\) depending on a parameter r of the form

$$ l_{r}(\mathbf {x},u)=u^{2}+r\theta ^{2}+\mathbf {x}^{T}Q_{0}\mathbf {x}+2\mathbf {x}^{T}Su, $$

that is, with a matrix \(Q(r)=Q_{0}+re_{1}e_{1}^{T}\) (\(e_{1}=(1,0,\dots ,0)\) denotes the first vector of the canonical basis of \(\mathbb {R}^{n}\)). We want to study the behavior when r tends to \(\infty \) of the optimal solutions of the following free-time OC problem: minimize the cost

$$ C_{r}(u,t_{u})=\int _{0}^{t_{u}}\left( g(t)+{\displaystyle l_{r}\big (\mathbf {x}_{u}(t),u(t)\big )}\right) \,dt, $$

among all inputs \(u(\cdot )\) and all times \(t_{u}\) such that \(\mathbf {x}_{u}(0)=\mathbf {x}^{0}\) and \(\mathbf {x}_{u}(t_{u})=\mathbf {x}^{f}\). As we have seen previously, the time \(\tau =\tau (r)\) is determined by Eq. (9) and the optimal solutions are the one of the OC problem \(\min \int _{0}^{\tau }l_{r}(\mathbf {x},u)\) in fixed time \(\tau \).

Lemma 7

For every \(r>0\) we denote by \(\mathbf {x}_{r}(t)= \left( \theta _{r}(t),\dots ,\theta _{r}{}^{(n-1)}(t)\right) ,\) \(t\in [0,\tau (r)],\) the solution of the free-time OC problem associated with \(C_{r}\). Assume that the infinitesimal cost of time \(g(\cdot )\) is a bounded function. Then there exists constants \(c,\alpha >0\), and \(\varepsilon \in (0,1)\), such that, when \(r\rightarrow \infty ,\) we have \(r^{1/2n}\tau (r)\rightarrow \infty \) and

$$ \theta _{r}(t)=c\theta ^{0}e^{-\alpha r^{1/2n}t}+O\left( e^{-\alpha r^{1/2n}\tau (r)}\right) ,\qquad \hbox {for any }\,t\in [\varepsilon \tau (r),\tau (r)]. $$

Note that the boundedness assumption on g is very natural and seems to be verified experimentally since we obtain functions g(t) that are decreasing for large t.

Proof

To simplify the study, we give only the proof in the case where the matrices \(Q_{0}\) and S are zero, and the dynamics (Eq. (5)) is of the form \(\theta ^{(n)}=u\). The proof of the complete result can be obtained by showing that this case actually gives the highest order terms with respect to r. With the preceding hypothesis, \(\theta _{r}(t)\) is the solution of the OC problem in fixed time \(\tau =\tau (r)\) associated with the infinitesimal cost \(u^{2}+r\theta ^{2}\), or equivalently with \(\frac{1}{r}u^{2}+\theta ^{2}\). Set \(\widetilde{\theta _{r}}(t)=\theta _{r}(tr^{-1/2n})\). Then \(\widetilde{\theta _{r}}(t)\) is the solution of the OC problem in fixed time \(r^{1/2n}\tau \) associated with the infinitesimal cost \(u^{2}+\theta ^{2}\). In the latter problem, nothing depends on r except the duration \(r^{1/2n}\tau \). It results from the analysis of Sect. 2.2.2 that there exists a universal function of time \(\varphi (\cdot )\) such that \(\widetilde{u_{r}}(r^{1/2n}\tau )=\theta ^{0}\varphi (r^{1/2n}\tau )\). Since we have \(u_{r}(t)=r^{1/2}\widetilde{u_{r}}(r^{1/2n}t)\), we obtain

$$ u_{r}(\tau )=r^{1/2}\theta ^{0}\varphi (r^{1/2n}\tau ). $$

Now remember (see Eq. (8)) that the time \(\tau \) must satisfy \(g(\tau )=\left( u_{r}(\tau )\right) ^{2}\), which gives \(g(\tau )=r\,\left( \theta ^{0}\varphi (r^{1/2n}\tau )\right) ^{2}\). Assume by contradiction that the quantity \(r^{1/2n}\tau (r)\) is bounded as \(r\rightarrow \infty \). Then \(\varphi (r^{1/2n}\tau )\) is bounded away from zero (\(\varphi \) is positive and continuous on \((0,+\infty )\), and converges to \(+\infty \) as \(t\rightarrow 0\), see Fig. 2), and therefore \(g(\tau (r))\rightarrow \infty \) as \(r\rightarrow \infty \), which contradicts the boundedness of g. Thus we get \(r^{1/2n}\tau (r)\rightarrow \infty \).

Since \(\widetilde{\theta _{r}}(t)\) is the solution of an OC problem in fixed time with a very large time \(r^{1/2n}\tau (r)\), it results from Lemma 4 that \(\widetilde{\theta _{r}}(t)=c\theta ^{0}e^{-\alpha t}+O\left( e^{-\alpha r^{1/2n}\tau (r)}\right) \) for t larger than \(\varepsilon r^{1/2n}\tau (r)\) for some \(\varepsilon \in (0,1)\). The conclusion follows from \(\theta _{r}(t)=\widetilde{\theta _{r}}(tr^{1/2n})\).    \(\square \)

1.2 Model for Arm Reaching Movements

Single degree-of-freedom (dof) limb. For a 1-dof arm moving in the horizontal plane, the basic model used throughout the study was already described in numerous other studies (e.g. [4, 20, 21, 26, 49]) and is as follows:

$$\begin{aligned} {\left\{ \begin{array}{ll} I\ddot{\theta } &{} =\tau -b\dot{\theta }\\ \dot{\tau } &{} =u \end{array}\right. } \end{aligned}$$
(12)

where is \(\theta \) the shoulder joint angle, \(\tau \) is the muscle torque, b is the friction coefficient (\(b=0.87\) here), I is the moment of inertia of the arm with respect to the shoulder joint (value estimated based upon Winter’s table for each participant; [57]) and u is the single control variable.

For the trajectory cost we typically considered canonical quadratic costs of the form \(l(\mathbf {x},u)=u^{2}+\mathbf {x}^{T}Q\mathbf {x}+2\mathbf {x}^{T}Su\), where \(\mathbf {x}=(\theta ,\dot{\theta },\ddot{\theta })\in \mathbb {R}^{3}\) denotes the system state. The two most famous examples are the minimum torque change corresponding to \(l(\mathbf {x},u)=u^{2}\) [55] and the minimum jerk corresponding to \(l(\mathbf {x},u)=\dddot{\theta }^{2}\) [18]. Other costs, possibly composite, may account for such planar movements in fixed time but such an investigation is out of the scope of the present chapter (but see [4,5,6,7, 19] for studies related to the trajectory cost identification).

Rights and permissions

Reprints and permissions

Copyright information

© 2017 Springer International Publishing AG

About this chapter

Cite this chapter

Jean, F., Berret, B. (2017). On the Duration of Human Movement: From Self-paced to Slow/Fast Reaches up to Fitts’s Law. In: Laumond, JP., Mansard, N., Lasserre, JB. (eds) Geometric and Numerical Foundations of Movements . Springer Tracts in Advanced Robotics, vol 117. Springer, Cham. https://doi.org/10.1007/978-3-319-51547-2_3

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-51547-2_3

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-51546-5

  • Online ISBN: 978-3-319-51547-2

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics