Synaptic integration across first-order tactile neurons can discriminate edge orientations with high acuity and speed

Our ability to manipulate objects relies on tactile inputs signaled by first-order tactile neurons that innervate the glabrous skin of the hand. A fundamental feature of these first-order tactile neurons is that their distal axon branches in the skin and innervates many mechanoreceptive end organs, yielding spatially-complex receptive fields with several highly sensitive zones. Recent work indicates that this peripheral arrangement constitutes a neural mechanism enabling first-order tactile neurons to signal high-level geometric features of touched objects, such as the orientation of an edge moving across the skin. Here we show how second-order tactile neurons could integrate these complex peripheral signals to compute edge-orientation. We first derive spiking models of human first-order tactile neurons that fit and predict responses to moving edges with high accuracy. Importantly, our models suggest that first-order tactile neurons innervate mechanoreceptors in the fingertips following a random sampling scheme. We then use the model neurons as the basis for a simulation of the peripheral neuronal population and its synaptic integration by second-order tactile neurons (i.e. in the spinal cord and brainstem). Our model networks indicate that computations done by second-order neurons could underlie the human ability to process edge orientations with high acuity and speed. In particular, our models suggest that synaptic integration of AMPA inputs within short timescales are critical for discriminating fine orientations, whereas NMDA-like synapses refine discrimination and maintain robustness over longer timescales. Taken together, our results provide new insight into the computations occurring in the earliest stages of the tactile processing pathway and how they may be critical for supporting hand function.

[1]  Anthony R. McIntosh,et al.  Multiregional integration in the brain during resting-state fMRI activity , 2017, PLoS Comput. Biol..

[2]  H. Sompolinsky,et al.  Compressed sensing, sparsity, and dimensionality in neuronal information processing and data analysis. , 2012, Annual review of neuroscience.

[3]  J. Randall Flanagan,et al.  Coding and use of tactile signals from the fingertips in object manipulation tasks , 2009, Nature Reviews Neuroscience.

[4]  A. Beitz,et al.  NMDA R1 mRNA distribution in motor and thalamic-projecting sensory neurons in the rat spinal cord and brain stem , 1995, Neuroscience Letters.

[5]  Gregory J. Gerling,et al.  Validating a Population Model of Tactile Mechanotransduction of Slowly Adapting Type I Afferents at Levels of Skin Mechanics, Single-Unit Response and Psychophysics , 2014, IEEE Transactions on Haptics.

[6]  Y. Fukami Interaction of impulse activities originating from individual Golgi tendon organs innervated by branches of a single axon. , 1980, The Journal of physiology.

[7]  Henry Markram,et al.  A Novel Multiple Objective Optimization Framework for Constraining Conductance-Based Neuron Models by Experimental Data , 2007, Front. Neurosci..

[8]  Shaul Hestrin,et al.  Activation and desensitization of glutamate-activated channels mediating fast excitatory synaptic currents in the visual cortex , 1992, Neuron.

[9]  Kenneth O. Johnson,et al.  A continuum mechanical model of mechanoreceptive afferent responses to indented spatial patterns. , 2006, Journal of neurophysiology.

[10]  Vincent Hayward,et al.  Spatio-temporal skin strain distributions evoke low variability spike responses in cuneate neurons , 2014, Journal of The Royal Society Interface.

[11]  Henry Markram,et al.  Preserving axosomatic spiking features despite diverse dendritic morphology. , 2013, Journal of neurophysiology.

[12]  R. Johansson,et al.  Properties of cutaneous mechanoreceptors in the human hand related to touch sensation. , 1984, Human neurobiology.

[13]  Interaction of activity in frog skin touch afferent units. , 1981, Journal of neurophysiology.

[14]  K W Horch,et al.  Impulse generation in type I cutaneous mechanoreceptors. , 1974, Journal of neurophysiology.

[15]  L. Dahlin,et al.  Consequences and adaptation in daily life – patients’ experiences three decades after a nerve injury sustained in adolescence , 2013, BMC Musculoskeletal Disorders.

[16]  Kenneth O. Stanley,et al.  Deep Neuroevolution: Genetic Algorithms Are a Competitive Alternative for Training Deep Neural Networks for Reinforcement Learning , 2017, ArXiv.

[17]  J. DiCarlo,et al.  Structure of Receptive Fields in Area 3b of Primary Somatosensory Cortex in the Alert Monkey , 1998, The Journal of Neuroscience.

[18]  Roland S Johansson,et al.  Fast and accurate edge orientation processing during object manipulation , 2017, bioRxiv.

[19]  Benoit P. Delhaye,et al.  Simulating tactile signals from the whole hand with millisecond precision , 2017, Proceedings of the National Academy of Sciences.

[20]  L. Abbott,et al.  Random Convergence of Olfactory Inputs in the Drosophila Mushroom Body , 2013, Nature.

[21]  E. G. Jones,et al.  Cortical and subcortical contributions to activity-dependent plasticity in primate somatosensory cortex. , 2000, Annual review of neuroscience.

[22]  R. Johansson,et al.  First spikes in ensembles of human tactile afferents code complex spatial fingertip events , 2004, Nature Neuroscience.

[23]  Wolfgang Maass,et al.  To Spike or Not to Spike: That Is the Question , 2015, Proc. IEEE.

[24]  Henry Markram,et al.  Models of Neocortical Layer 5b Pyramidal Cells Capturing a Wide Range of Dendritic and Perisomatic Active Properties , 2011, PLoS Comput. Biol..

[25]  Yoshichika Baba,et al.  Computation identifies structural features that govern neuronal firing properties in slowly adapting touch receptors , 2014, eLife.

[26]  G. J. Gerling,et al.  Predicting SA-I mechanoreceptor spike times with a skin-neuron model. , 2009, Mathematical biosciences.

[27]  R. Purple,et al.  Afferent fibers with multiple encoding sites. , 1974, Brain research.

[28]  J Andrew Pruszynski,et al.  Neural network models of the tactile system develop first-order units with spatially complex receptive fields , 2017, bioRxiv.

[29]  J. Phillips,et al.  Responses of human mechanoreceptive afferents to embossed dot arrays scanned across fingerpad skin , 1992, The Journal of neuroscience : the official journal of the Society for Neuroscience.

[30]  Maria Nolano,et al.  Quantification of myelinated endings and mechanoreceptors in human digital skin , 2003, Annals of neurology.

[31]  V. Hayward,et al.  Segregation of Tactile Input Features in Neurons of the Cuneate Nucleus , 2014, Neuron.

[32]  David Cai,et al.  Sparsity and Compressed Coding in Sensory Systems , 2014, PLoS Comput. Biol..

[33]  Tatyana O. Sharpee,et al.  A Robust Feedforward Model of the Olfactory System , 2016, PLoS Comput. Biol..

[34]  J. A. Pruszynski,et al.  Edge-orientation processing in first-order tactile neurons , 2014, Nature Neuroscience.

[35]  Nicole L. Carlson,et al.  Sparse Codes for Speech Predict Spectrotemporal Receptive Fields in the Inferior Colliculus , 2012, PLoS Comput. Biol..

[36]  Aneesha K. Suresh,et al.  Edge orientation signals in tactile afferents of macaques. , 2016, Journal of neurophysiology.

[37]  Paul A. Rhodes,et al.  The Properties and Implications of NMDA Spikes in Neocortical Pyramidal Cells , 2006, The Journal of Neuroscience.

[38]  D. Tapper,et al.  Integration of impulse activity in a peripheral sensory unit. , 1966, Experimental neurology.

[39]  D. L. Donoho,et al.  Compressed sensing , 2006, IEEE Trans. Inf. Theory.

[40]  C. Rivadulla,et al.  Intracuneate mechanisms underlying primary afferent cutaneous processing in anaesthetized cats , 2004, The European journal of neuroscience.