Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Touch and tactile neuropathic pain sensitivity are set by corticospinal projections

Abstract

Current models of somatosensory perception emphasize transmission from primary sensory neurons to the spinal cord and on to the brain1,2,3,4. Mental influence on perception is largely assumed to occur locally within the brain. Here we investigate whether sensory inflow through the spinal cord undergoes direct top-down control by the cortex. Although the corticospinal tract (CST) is traditionally viewed as a primary motor pathway5, a subset of corticospinal neurons (CSNs) originating in the primary and secondary somatosensory cortex directly innervate the spinal dorsal horn via CST axons. Either reduction in somatosensory CSN activity or transection of the CST in mice selectively impairs behavioural responses to light touch without altering responses to noxious stimuli. Moreover, such CSN manipulation greatly attenuates tactile allodynia in a model of peripheral neuropathic pain. Tactile stimulation activates somatosensory CSNs, and their corticospinal projections facilitate light-touch-evoked activity of cholecystokinin interneurons in the deep dorsal horn. This touch-driven feed-forward spinal–cortical–spinal sensitization loop is important for the recruitment of spinal nociceptive neurons under tactile allodynia. These results reveal direct cortical modulation of normal and pathological tactile sensory processing in the spinal cord and open up opportunities for new treatments for neuropathic pain.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Adult CST ablation impairs light touch but not nociceptive behavioural responses.
Fig. 2: Hindlimb somatosensory CSN ablation causes specific loss of light touch response and mechanical allodynia.
Fig. 3: Light-touch-elicited activation of somatosensory CSNs and spinal dorsal horn neurons.
Fig. 4: Lumbar CCK+ neurons receive convergent Aβ fibre and CST inputs and are required for mechanical allodynia.

Data availability.

The datasets generated during and/or analysed during the current study are available from the corresponding author on reasonable request.

References

  1. Owens, D. M. & Lumpkin, E. A. Diversification and specialization of touch receptors in skin. Cold Spring Harb. Perspect. Med. 4, a013656 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Abraira, V. E. & Ginty, D. D. The sensory neurons of touch. Neuron 79, 618–639 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Basbaum, A. I., Bautista, D. M., Scherrer, G. & Julius, D. Cellular and molecular mechanisms of pain. Cell 139, 267–284 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Costigan, M., Scholz, J. & Woolf, C. J. Neuropathic pain: a maladaptive response of the nervous system to damage. Annu. Rev. Neurosci. 32, 1–32 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Lemon, R. N. Descending pathways in motor control. Annu. Rev. Neurosci. 31, 195–218 (2008).

    Article  CAS  PubMed  Google Scholar 

  6. Liu, K. et al. PTEN deletion enhances the regenerative ability of adult corticospinal neurons. Nat. Neurosci. 13, 1075–1081 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Jin, D. et al. Restoration of skilled locomotion by sprouting corticospinal axons induced by co-deletion of PTEN and SOCS3. Nat. Commun. 6, 8074 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Boada, M. D. & Woodbury, C. J. Myelinated skin sensory neurons project extensively throughout adult mouse substantia gelatinosa. J. Neurosci. 28, 2006–2014 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Boada, M. D. & Woodbury, C. J. Physiological properties of mouse skin sensory neurons recorded intracellularly in vivo: temperature effects on somal membrane properties. J. Neurophysiol. 98, 668–680 (2007).

    Article  PubMed  Google Scholar 

  10. Lemon, R. N. & Griffiths, J. Comparing the function of the corticospinal system in different species: organizational differences for motor specialization? Muscle Nerve 32, 261–279 (2005).

    Article  PubMed  Google Scholar 

  11. Abraira, V. E. et al. The cellular and synaptic architecture of the mechanosensory dorsal horn. Cell 168, 295–310.e19 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wang, X. et al. Deconstruction of corticospinal circuits for goal-directed motor skills. Cell 171, 440–455.e14 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kinoshita, M. et al. Genetic dissection of the circuit for hand dexterity in primates. Nature 487, 235–238 (2012).

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Liu, Y. et al. A sensitized IGF1 treatment restores corticospinal axon-dependent functions. Neuron 95, 817–833.e4 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. O’Leary, D. D. Development of connectional diversity and specificity in the mammalian brain by the pruning of collateral projections. Curr. Opin. Neurobiol. 2, 70–77 (1992).

    Article  PubMed  Google Scholar 

  16. O’Leary, D. D. & Koester, S. E. Development of projection neuron types, axon pathways, and patterned connections of the mammalian cortex. Neuron 10, 991–1006 (1993).

    Article  PubMed  Google Scholar 

  17. Decosterd, I. & Woolf, C. J. Spared nerve injury: an animal model of persistent peripheral neuropathic pain. Pain 87, 149–158 (2000).

    Article  CAS  PubMed  Google Scholar 

  18. Atasoy, D., Betley, J. N., Su, H. H. & Sternson, S. M. Deconstruction of a neural circuit for hunger. Nature 488, 172–177 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  19. Li, X. et al. Skin suturing and cortical surface viral infusion improves imaging of neuronal ensemble activity with head-mounted miniature microscopes. J. Neurosci. Methods 291, 238–248 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Todd, A. J. Neuronal circuitry for pain processing in the dorsal horn. Nat. Rev. Neurosci. 11, 823–836 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Cheng, L. et al. Identification of spinal circuits involved in touch-evoked dynamic mechanical pain. Nat. Neurosci. 20, 804–814 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ji, R. R., Baba, H., Brenner, G. J. & Woolf, C. J. Nociceptive-specific activation of ERK in spinal neurons contributes to pain hypersensitivity. Nat. Neurosci. 2, 1114–1119 (1999).

    Article  CAS  PubMed  Google Scholar 

  23. Torsney, C. & MacDermott, A. B. Disinhibition opens the gate to pathological pain signaling in superficial neurokinin 1 receptor-expressing neurons in rat spinal cord. J. Neurosci. 26, 1833–1843 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Duan, B. et al. Identification of spinal circuits transmitting and gating mechanical pain. Cell 159, 1417–1432 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Bourane, S. et al. Identification of a spinal circuit for light touch and fine motor control. Cell 160, 503–515 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Taniguchi, H. et al. A resource of Cre driver lines for genetic targeting of GABAergic neurons in cerebral cortex. Neuron 71, 995–1013 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Sun, F. et al. Sustained axon regeneration induced by co-deletion of PTEN and SOCS3. Nature 480, 372–375 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Arlotta, P. et al. Neuronal subtype-specific genes that control corticospinal motor neuron development in vivo. Neuron 45, 207–221 (2005).

    Article  CAS  PubMed  Google Scholar 

  29. Metz, G. A., Dietz, V., Schwab, M. E. & van de Meent, H. The effects of unilateral pyramidal tract section on hindlimb motor performance in the rat. Behav. Brain Res. 96, 37–46 (1998).

    Article  CAS  PubMed  Google Scholar 

  30. Muir, G. D. & Whishaw, I. Q. Complete locomotor recovery following corticospinal tract lesions: measurement of ground reaction forces during overground locomotion in rats. Behav. Brain Res. 103, 45–53 (1999).

    Article  CAS  PubMed  Google Scholar 

  31. Mastwal, S. et al. Phasic dopamine neuron activity elicits unique mesofrontal plasticity in adolescence. J. Neurosci. 34, 9484–9496 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Cao, V. Y. et al. In vivo two-photon imaging of experience-dependent molecular changes in cortical neurons. J. Vis. Exp. 71, 50148 (2013).

    Google Scholar 

  33. Cao, V. Y. et al. Motor learning consolidates Arc-expressing neuronal ensembles in secondary motor cortex. Neuron 86, 1385–1392 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Ziv, Y. et al. Long-term dynamics of CA1 hippocampal place codes. Nat. Neurosci. 16, 264–266 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Mukamel, E. A., Nimmerjahn, A. & Schnitzer, M. J. Automated analysis of cellular signals from large-scale calcium imaging data. Neuron 63, 747–760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Hyvärinen, A. & Oja, E. Independent component analysis: algorithms and applications. Neural Netw. 13, 411–430 (2000).

    Article  PubMed  Google Scholar 

  37. Chen, T. W. et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hill, D. N., Varga, Z., Jia, H., Sakmann, B. & Konnerth, A. Multibranch activity in basal and tuft dendrites during firing of layer 5 cortical neurons in vivo. Proc. Natl Acad. Sci. USA 110, 13618–13623 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  39. Peters, A. J., Lee, J., Hedrick, N. G., O’Neil, K. & Komiyama, T. Reorganization of corticospinal output during motor learning. Nat. Neurosci. 20, 1133–1141 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Cichon, J., Blanck, T. J. J., Gan, W. B. & Yang, G. Activation of cortical somatostatin interneurons prevents the development of neuropathic pain. Nat. Neurosci. 20, 1122–1132 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Peters, A. J., Chen, S. X. & Komiyama, T. Emergence of reproducible spatiotemporal activity during motor learning. Nature 510, 263–267 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  42. Holloway, B. B. et al. Monosynaptic glutamatergic activation of locus coeruleus and other lower brainstem noradrenergic neurons by the C1 cells in mice. J. Neurosci. 33, 18792–18805 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank T. Huang, Y. Zhang and Q. Ma for advice and D. Ginty, S. Hegarty, Q. Ma, F. Wang and P. Williams for critical reading. This study was supported by grants from the Craig Neilsen Foundation (Y.L. and X.W.), Paralyzed Veterans of America Foundation (Y.L.), Dr. Miriam and Sheldon G. Adelson Medical Research Foundation and NINDS (C.J.W. and Z.H.) and NIMH intramural research program ZIA MH002897 (K.H.W. and X.L.). IDDRC and viral cores supported by the grants NIH P30 HD018655 and P30EY012196 were used for this study.

Reviewer information

Nature thanks R. Ji and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

Y.L., A.L., X.L., Z.Z., K.H.W., C.J.W. and Z.H. conceived the experiments. Y.L., A.L., X.L., Z.Z., M.C., X.W., C.F., C.A., J.Z., Z.G., B.C., X.D., J.-Y.Z. and Y.Z. performed the experiments. Y.L., A.L., X.L., C.C., K.H.W., C.J.W. and Z.H. prepared the manuscript with input from all authors. K.H.W., C.J.W. and Z.H. co-supervised the project.

Corresponding authors

Correspondence to Kuan Hong Wang, Clifford J. Woolf or Zhigang He.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Effects of pyramidotomy on tactile behaviour and gross locomotion in mice.

a, Correlation between CST ablation and tactile behaviours in mice with pyramidotomy. For individual animals that received pyramidotomy, tag number, image and quantification of L3 spinal cord sections stained with anti-PKCγ antibodies showing remaining CST axons, percentage of withdrawal response to low-threshold von Frey filament (0.16g) and light brush, and sense time to the tape are shown. b, c, Performance on ground walking (b, P = 0.14, hindlimb weight support; P = 0.81, hindlimb retraction; P = 0.41, hindlimb protraction; P = 0.81, inter-limb coordination; P = 0.20, fore–hindlimb coordination), and rotarod test (c, P = 0.87) in mice with sham surgery (n = 8) or pyramidotomy (n = 12). n.s., no statistical significance. Two-sided student’s t-test. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 2 CST axon termination in the lumbar spinal cord.

a, Representative transverse spinal section (L3) from an Emx1-tdTomato (red) reporter line. b, Sections were co-stained with IB4 (green), a lamina IIi marker, and anti-PKCγ, a laminae IIi/III marker, in the spinal dorsal horn. Scale bar, 500 μm. For a and b, three and four mice, respectively, showed similar results.

Extended Data Fig. 3 Efficiency of somatosensory CSN ablation by P14 intraspinal injection.

a, Left, schematic of regional CSN ablation by P14 lumbar (T13–L6) intraspinal injection. Right, representative image (n = 8 animals with similar results) of the cortex with GFP+ areas covering hindlimb S1/S2. b, Representative images (n = 6 animals with similar results) of cortical sections showing hindlimb CSNs retrogradely labelled by lumbar (T13–L6) intraspinal injection of HiRet-GFP at P14. Scale bar, 100 μm. c, To assess ablation efficiency, at the end point, retrograde-targeting rAAV-mCherry was injected into the lumbar spinal cord in some animals. Representative images of cortical sections showing retrogradely labelled mCherry+ CSNs (left) within the GFP+ cortical areas (right) (S1/S2) in control or AAV-FLEX-DTR injected animals with quantification (normalized to those in controls as 100). **P < 0.01 (P < 0.0001), two-sided Student’s t-test. n = 5 and 5 for control and AAV-FLEX-DTR injected mice, respectively. Scale bar, 100 μm. d, Representative images of transverse lumbar spinal cord sections showing residual CST axons labelled by GFP (from mice co-injected with AAV-GFP to S1/S2 and AAV-FLEX-DTR) in control or S1/S2 CSN ablated animals with quantification. **P < 0.01 (P < 0.0001), two-sided Student’s t-test. n =  7 or 8 for control or AAV-FLEX-DTR-injected mice, respectively. Scale bar, 500 μm. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 4 Mechanical allodynia induced by SNI or CFA injection is compromised in mice with pyramidotomy, but cold allodynia and mechanical hyperalgesia induced by SNI are not.

a, Schematic drawing of experimental paradigm. be, Measurement of punctate (b) and dynamic (c) mechanical allodynia, cold allodynia (d) and mechanical hyperalgesia (e) after SNI in mice that underwent sham surgery (n = 8) or pyramidotomy (n = 9) 1–21 days after SNI. b, P < 0.0001, P < 0.0001, P = 0.0012, P = 0.0004 and P = 0.041; c, P < 0.0001, P < 0.0001, P = 0.0004, P = 0.001 and P = 0.0045; d, P = 0.26, P = 0.33, P = 0.29, P > 0.99 and P > 0.99; e, P > 0.99, P = 0.15, P = 0.56, P > 0.99 and P = 0.74 for 1, 3, 7, 14 and 21 d post SNI, respectively. f, g, Measurement of punctate (f) and dynamic (g) mechanical allodynia in mice that underwent sham surgery (n = 6) or pyramidotomy (n = 6) 1–7 days after hindpaw CFA injection. f, P = 0.01, P = 0.01, P = 0.01, P < 0.0001 and P < 0.0001; g, P < 0.0001 for 1, 2, 3, 5 and 7d post CFA injection, respectively. Two-way repeated measures ANOVA followed by Bonferroni correction. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 5 Calcium imaging of CSN activity in intact and SNI mice.

a, Schematic of experimental procedures. b, Confocal fluorescence images of coronal brain sections showing specific expression of GCaMP6s in CSNs. Left, a 10× image showing labelled CSN soma and dendrites; right, a 25× image showing the magnified view of apical dendritic trunks. Dotted line, expected focal plane of head-mounted microscope. Scale bars, 100 μm. c, Procedures for identifying the active events of CSN dendrites. Left, example of dendrites identified from a calcium movie by ICA analysis. The brightest spot in a dendritic tree, corresponding to the trunk (red circle), is used as a region of interest for temporal signal analysis. Top trace, temporal signal of the dendrite. Bottom trace, magnified calcium events. Horizontal bars indicate rising phases of calcium events, which are associated with neuronal activation and used in subsequent analysis. Scale bar, 100 µm. d, Example calcium movie frames showing dendritic activities of hindlimb S1 CSNs upon different sensory stimuli in intact mice. In these examples, brush stimuli activated CSNs, whereas von Frey (0.04 g) and laser heat stimuli did not. Calcium signals are expressed as ∆F/F0 (F0 is the time-average fluorescence of the whole movie). Scale bar, 200 µm. For b, d, the experiments were repeated independently four times with similar results. e, Pie charts showing the proportions of neurons that responded to brushes, von Frey filaments or both brush and von Frey stimulation before and after SNI. Few neurons responded to both stimuli before SNI, but this overlapping proportion increased after SNI.

Extended Data Fig. 6 Mechanical allodynia induced by spinal disinhibition (treatment with bicuculline and strychnine) is compromised in mice with pyramidotomy.

a, Drawing of experimental paradigm. b, c, Measurement of punctate (b) and dynamic (c) mechanical allodynia after intrathecal injection of bicuculline and strychnine in mice that had undergone sham surgery (n = 6) or pyramidotomy (n = 7). **P < 0.01 (P < 0.0001 for b and c at 10, 30, and 90 min post drug), two-way repeated measures ANOVA followed by Bonferroni correction. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 7 Neuronal activity in cortical and subcortical areas upon light touch after SNI.

ac, Drawings of c-Fos immunostaining in intact mice (a), mice with SNI only (b) and mice with SNI after light brush stimulation (c) from control mice and mice with pyramidotomy. mPFC, medial prefrontal cortex; ACC, anterior cingulate cortex; S1HL, hindlimb primary somatosensory cortex; S2, secondary somatosensory cortex; Pir, piriform cortex; PV, periventricular nucleus of the thalamus; VM, ventromedial nucleus of the hypothalamus; Amyg, amygdala. d, Quantification of c-Fos+ cells in multiple cortical areas in intact mice (with CSN, n = 3; with pyramidotomy, n = 3), mice with SNI only (with CSN, n = 3; with pyramidotomy, n = 3), and mice with SNI after light brush stimulation (with CSN, n = 4; with pyramidotomy, n = 3). **P < 0.01; n.s., no statistical significance. PFC SNI only with or without Py: P > 0.99; PFC SNI + brush with or without Py: P < 0.0001; ACC SNI only with or without Py: P > 0.99; ACC SNI + brush with or without Py: P = 0.0002; M1 SNI only with or without Py: P > 0.99; M1 SNI + brush with or without Py: P > 0.99; S1 SNI only with or without Py: P > 0.99; S1 SNI + brush with or without Py: P < 0.0001; insula SNI only with or without Py: P > 0.99; insula SNI + brush with or without Py: P < 0.0001; S2 SNI + brush with or without Py: P = 0.0075; Pir all conditions (ANOVA): P = 0.82. One-way ANOVA followed by Bonferroni correction. e, Representative images of multiple cortical areas stained with c-Fos (red) and GFP (green, HiRet-GFP injection) in mice with SNI after light brush stimulation. Arrowheads mark the co-localization of c-Fos+ and GFP+ CSNs. Scale bars, 20 μm. f, Quantification of c-Fos+ and GFP+ CSN co-localization in multiple cortical areas in animals with SNI after light brush stimulation without (n = 4) or with pyramidotomy (Py, n = 3). *P < 0.05; **P < 0.01, P = 0.03, 0.04, 0.0009, and 0.02 for mPFC, ACC, S1, and S2, respectively. Two-sided Student’s t-test. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 8 Ablation of lumbar CCK+ interneurons reduces tactile sensitivity and dorsal horn neuronal activation after SNI, but not nociceptive response or gross locomotion.

ad, Measurements of sensitivity to laser heat (a, P = 0.76), acetone (b, P = 0.86), von Frey filaments (c, P = 0.03 for 0.016 g) and brush (d, P = 0.002) stimuli in control mice (n = 8) or mice in which CCK–tdTomato interneurons had been ablated (n = 7). a, b, d, Two-sided Student’s t-test; c, two-way repeated measures ANOVA followed by Bonferroni correction. e, f, Performance on open field (e, P = 0.54) and ground walking (f, P = 0.68, 0.72, and 0.50 for hindlimb weight support, protraction, and retraction, respectively) in control mice (n = 8) or mice in which CCK–tdTomato interneurons had been ablated (n = 7). n.s., no statistical significance, two-sided Student’s t-test. gi, Representative images of c-Fos (green) activity (g) and quantification of CCK+/c-Fos+ cells (h) and c-Fos+ neurons in different laminae of the dorsal horn of the spinal cord (L3–4) of CCK–tdTomato (red) mice after SNI and brush stimulation in mice that had undergone sham surgery (n = 4), pyramidotomy (n = 5) or lumbar CCK–tdTomato ablation (n = 3) (i). Scale bars, 500 μm (g, right, middle, and left columns) and 50 μm (g, zoom in panels). **P < 0.01, two-sided Student’s t-test. h, P < 0.0001; i, P = 0.0045 and 0.0028 for laminae I–II and laminae III–V, respectively. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 9 Characterization of Aβ fibre and CST inputs onto CCK–tdTomato neurons.

a, Schematic of the stimulation and whole-cell patch recording set-up for tdTomato-labelled CCK+ interneurons. CST axons labelled by AAV-ChR2-YFP were stimulated with a 473-nm laser. A single dorsal root (L4–L6) was stimulated with a glass suction electrode. b, c, Representative consecutive traces (n = 3) of Aβ fibre (left) and opto-CST (right) stimulation-evoked responses (b) and summarization (c) of whole cell patch-clamp recordings on CCK–tdTomato interneurons. Three recording conditions were used: first, to detect evoked EPSCs (eEPSCs), we held the membrane potential at −70 mV, which is the equilibrium potential of Cl and therefore minimizes the flow of eIPSCs. Second, by holding the membrane potential at 0 mV, we examined the polysynaptic, inhibitory inputs (eIPSCs) onto CCK–tdTomato interneurons. Third, we used current clamp mode to examine whether the stimulation drove action potential firing at the resting membrane potential. Type 1: CCK–tdTomato neurons receive only excitatory inputs, and few of them generated AP outputs when Aβ or CST inputs were stimulated. Type 2: CCK–tdTomato neurons receive both excitatory inputs and feed-forward inhibitory inputs, with no AP output. Type 3: CCK–tdTomato neurons receive predominant feed-forward inhibition, with no AP output. Type 4: CCK–tdTomato neurons show no response at either voltage or current clamp recording. d, Left, representative recording of an Aβ (25 μA) dorsal root evoked EPSC at −70 mV. Latency and jitter properties (magnified in inset) with quantifications (n = 8 neurons) are consistent with monosynaptic sensory connectivity. Right, opto-stimulation evoked EPSCs (averaged traces) at −70 mV in the same cell as shown on the left. The evoked EPSC was blocked by AMPA and NMDA antagonists (NBQX (5 μM)/CPP (20 μM)). In addition, opto-stimulation evoked EPSCs were eliminated by TTX (0.5 μM), and reinstated by 4-AP (2 mM), indicating a monosynaptic connection between the CST and CCK–tdTomato interneurons. Bar graph, quantification of eEPSC amplitudes after administration of drugs as shown. **P < 0.0001; one-way ANOVA followed by Bonferroni correction. n = 8 neurons. Data shown as mean ± s.e.m.

Source data

Extended Data Fig. 10 Reinforcement of tactile allodynia in mice with SNI by optogenetic stimulation of somatosensory CSNs.

a, Drawing of experimental paradigm. b, c, Measurement of punctate (b, von Frey filaments) and dynamic (c, brush) mechanical allodynia upon opto-stimulation in control mice (n = 6) and mice expressing ChR2-YFP in CSNs (n = 6) after SNI. n.s., no statistical significance; *P < 0.05. b, c, P = 0.18, 0.08 and 0.41, 0.02 for PLAP and Cre, without or with laser, respectively; two-sided Student’s t-test. d, Representative images and quantification of pERK (red) and NK1R (green) immunostaining in the superficial dorsal horn (laminae I–II) of the spinal cord (L3–4) in control mice (n = 3) or mice expressing ChR2-YFP in hindlimb CSNs (n = 3) with SNI and brush stimulation coupled with opto-stimulation. Scale bar, 100 μm. *P < 0.05; P = 0.02 and 0.01 for pERK and pERK/NK1R ratio, respectively; two-sided Student’s t-test. Six sections crossing the lumbar spinal cord (L3–4) were quantified for individual animals. Data shown as mean ± s.e.m.

Source data

Supplementary information

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, Y., Latremoliere, A., Li, X. et al. Touch and tactile neuropathic pain sensitivity are set by corticospinal projections. Nature 561, 547–550 (2018). https://doi.org/10.1038/s41586-018-0515-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-018-0515-2

Keywords

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing