31
Draft of 1 September 2013 The central pattern generator for rhythmic whisking David Kleinfeld 1 , Martin Deschênes 2 and Jeffrey D. Moore 3 1 Department of Physics and Section on Neurobiology, UCSD, La Jolla, CA, USA ([email protected]). 2 Department of Psychiatry and Neuroscience, Laval University, Quebec City, Canada ([email protected]). 3 Graduate Program in Neuroscience, UCSD, La Jolla, CA, USA ([email protected]). To appear in: Sensorimotor Integration in the Whisker System Editors: Patrik Krieger and Alexander Groh Publisher: Tucker Seven Editorial Associates Abstract: Six sentences Narrative ~2400 words Legends: ~1400 words References: 44 Figures 14 Editorial correspondence: David Kleinfeld Department of Physics 0374 University of California 9500 Gilman Drive La Jolla, CA 92093 Office: 858-822-0342 Mobile: 858-922-4664 Fax: 858-534-7697 Email: [email protected]

Draft of 1 September 2013 -  · Neuronal output from this region is reset at ... functions as a master clock to coordinate orofacial ... in which an epoch of whisking appears

Embed Size (px)

Citation preview

Draft of 1 September 2013

The central pattern generator for rhythmic whisking David Kleinfeld1, Martin Deschênes2 and Jeffrey D. Moore3 1Department of Physics and Section on Neurobiology, UCSD, La Jolla, CA, USA ([email protected]).

2Department of Psychiatry and Neuroscience, Laval University, Quebec City, Canada ([email protected]).

3Graduate Program in Neuroscience, UCSD, La Jolla, CA, USA ([email protected]).

To appear in: Sensorimotor Integration in the Whisker System Editors: Patrik Krieger and Alexander Groh Publisher: Tucker Seven Editorial Associates Abstract: Six sentences Narrative ~2400 words Legends: ~1400 words References: 44 Figures 14 Editorial correspondence: David Kleinfeld Department of Physics 0374 University of California 9500 Gilman Drive La Jolla, CA 92093 Office: 858-822-0342 Mobile: 858-922-4664 Fax: 858-534-7697 Email: [email protected]

Draft of 1 September 2013

2

Abstract

Whisking and sniffing are predominant aspects of exploratory behavior in rodents. We review evidence that these motor rhythms are coordinated by the respiratory patterning circuitry in the ventral medulla. A region in the intermediate reticular zone, dorsomedial to the preBotzinger inspiratory complex, provides rhythmic input to the facial motoneurons that drive protraction of the vibrissae. Neuronal output from this region is reset at each inspiration by direct input from the preBötzinger complex. High frequency breathing, or sniffing, has a one-to-one coordination with whisking while basal respiration is accompanied by intervening whisks that occur between breaths. We conjecture that the preBötzinger complex, which projects to neighboring premotor regions for the control of other orofacial muscles, functions as a master clock to coordinate orofacial behaviors with breathing.

Draft of 1 September 2013

3

Some 50 years ago Welker [1] observed that whisking and sniffing by

rodents appeared to be synchronous. More than a curious observation, it leads to

the suggestion that breathing may be at the root of all rhythmic orofacial

behaviors. We now understand the basis for Welker's observation and, in this

review, summarize the hunt for the neuronal circuitry that controls rapid

exploratory whisking by rats [2]. This is followed by a discussion of the potential

role of sniffing as carrier signal that binds different orofacial sensory inputs.

Before we begin our narrative on whisking, it is worth considering an

abbreviated wiring diagram of the anatomy of the vibrissa sensorimotor system

[3] (figure 1), as such circuit maps can constrain the potential mechanisms that

that generate behavior. The business end of the rodent vibrissa system is the

mystacial pad, which includes the muscles that drive the vibrissae and the

sensory fibers that innervate the follicle. We see that the nervous system already

forms a feedback loop at the level of the brainstem, from the trigeminal ganglion

inputs through interneurons in the trigeminal nuclei and back to the facial

motoneurons that drive the muscles. This disynaptic feedback loop is the

shortest sensorimotor pathway [4], and it is paralleled by loops that involve the

cerebellum, by loops that involved the midbrain, e.g., the superior colliculus, and

by loops that involve the forebrain, the most extensive of which includes vibrissa

primary sensory (vS1) and primary motor cortices (vM1).

What is the nature of whisking? Rodents will whisk in air as they explore a

novel environment and search for objects and conspecifics [5]. The nature of

whisking can change upon contact with a surface, especially as the animal turns

and whisking is no longer symmetric [6, 7]. Here we focus on the case of

rhythmic symmetric whisking by rats [8], in which an epoch of whisking appears

almost periodic (figure 2). This suggests that from a mathematical perspective,

whisking may be conceived as a rhythmic process, with a rapidly evolving phase

and a slowly evolving envelope, i.e., amplitude and midpoint, much like the

signals in AM radio. Formally, we can define the angle of a vibrissa relative to the

face as θ(t), which evolves according to θ(t) = θamplitude(t)•cos [φ(t)] + θmidpoint(t)

with dφ(t)/dt = 2πfwhisk, where fwhisk is the instantaneous whisking frequency [9].

Draft of 1 September 2013

4

This definition raises an interesting question. Does the brain drive whisking

through a combination of signals, each with a different time-scale as implied by

the mathematical decomposition, or is there a single control signal with a broad

dynamic range in time? Past experimental results suggest that nature has

chosen to control whisking through narrow band signals that code on the scale of

the fast rhythm, i.e., roughly 10 Hz, and the slowly evolving envelope, i.e.,

roughly 1 Hz. First, Pietr et al. [10] showed that systemic infusion of an agonist of

endocannabinoid receptors leads to a decrease in the amplitude of whisking,

while infusion of an antagonist these same receptors leads to high-amplitude

whisking without variation in amplitude (figure 3a). Critically, in both cases the

pharmacological interventions had no effect on the frequency of whisking

(figure 3b). Second, Hill et al. [11] showed that the firing patterns of neurons in

vM1 cortex preferentially report the slowly varying amplitude and midpoint of a

whisking bout independent of the rapidly varying phase (figure 4). These data

suggest that the rapidly varying phase and the slowly varying change in the

envelope of whisking are controlled by separate mechanisms.

We now revisit Welker’s [1] observation regarding the synchrony of

whisking and sniffing. Quantitative concurrent measurements of breathing and

whisking in head-restrained and freely moving rats reveal key aspects of their

coordination [2, 12] (insert in figure 5). First, breathing over a wide range of

rates can occur without substantial whisking (figure 5a). To test whether

whisking can also occur without breathing, a puff of ammonia was applied to the

snout to inactivate the central inspiratory drive [13] and temporarily inhibit

respiration. Critically, rats can whisk during such a disruption in breathing

(figure 5b), which implies that the oscillator for breathing and the putative

oscillator for whisking are separately gated. While sniffing is indeed accompanied

by a one-to-one relation with whisking (figure 5a), basal breathing is

accompanied by whisks that are coincident with an inspiration, which we denote

"inspiratory-locked whisks", and also "intervening whisks" that occur between

successive breaths and successively decrease in amplitude (figure 5c).

Therefore there is an incommensurate many-to-one relation between whisking

Draft of 1 September 2013

5

and breathing. These data imply that there are separate, or separable, oscillators

for breathing and whisking, and that the breathing rhythm may reset the whisking

rhythm.

The detailed timing between whisking and breathing is quantified though a

frequency-ordered plot of the correlation of whisking with breathing across a

large data set of whisks (figure 6). Vibrissa protractions are time-locked to the

onset of inspiration across the entire range of breathing frequencies. Basal

respiration cycles are accompanied by multiple whisks per breath, with an

instantaneous whisking frequency of approximately 8 Hz for the intervening

whisks. These data imply a unidirectional connection from the breathing oscillator

[14-16] to a putative CPG for whisking.

Where is the pattern generator for the fast, rhythmic vibrissa motion? One

possibility is that rapid whisking is generated by the previously discussed

disynaptic feedback loop in the brainstem. That is, sensory signals generated by

the motion of the vibrissa could directly drive the facial motoneurons, which

would in turn generate a new sensory signal, and so on. The frequency of

whisking would depend of propagation delays in this circuit. Against this

hypothesis, lesion of the sensory nerve (infraorbial nerve in figure 1) has

minimal effect on whisking [1, 8, 17]; in fact the rhythmic pattern becomes more

stable [8]! Similarly, lesioning of any of a large number of midbrain and forebrain

nuclei only minimally interrupt whisking [1, 18]. This leads to the hypothesis that

an oscillator which drives whisking is located in the brainstem. But where? Given

the close phase relationship between whisking and breathing, the neighborhood

of the ventral respiratory column in the medulla [19] is a clear suspect. A role for

breathing in the control of whisking is further suggested by the shared facial

musculature between the two behaviors [20, 21]. Lastly, the CPGs for chewing,

i.e., the oral portion of the gigantocellular reticular nucleus [22], for airway

control, i.e., the ambiguus nucleus [23], and for licking, i.e., the medial edge of

the parvocellular reticular formation bordering the IRt, are located near each

other [24] and near the preBötzinger complex, which has been demonstrated to

generate the inspiratory breathing rhythm [14, 25]. This collective proximity is

Draft of 1 September 2013

6

consistent with the need to synchronize orofacial behaviors [26]. In particular, the

coordination of whisking with breathing and the resetting of whisking by

inspiration suggests that a brainstem whisking CPG is reset, and possibly driven,

by the preBötzinger complex.

The difference in the basal respiration, at frequencies < 3 Hz, and

whisking patterns provides a signature to discriminate between breathing and

potential whisking neuronal centers [2] (figure 5c). As a control, we first recorded

multiunit spiking activity within the preBötzinger complex and rostral ventral

respiratory group in awake rats, and we observe units whose spiking occurred in

phase with inspiration and with vibrissa protraction during inspiratory-locked

whisks (figure 7a and red dots in figure 7c). Critically, the activity did not track

the intervening whisks. In contrast, we located a subset of units in the

intermediate band of the reticular formation (IRt) whose spiking was tightly

phase-locked to the protraction of both inspiratory-locked and intervening whisks

(figure 7b and blue dots in figure 7c). These units are potential pre-motor

drivers of the intrinsic muscles that serve rhythmic whisking (figure 2a) and are

henceforth referred to as "whisking units". They are located in the ventral part of

the IRt, medial to the ambiguus nucleus and immediately doromedial to the

preBötzinger complex. We denote this new region the vibrissa zone of the IRt

(vIRt) (figure 1).

The hypothesis that whisking units in the vIRt constitute the oscillator for

whisking predicts that activation of this region will lead to prolonged autonomous

rhythmic whisking [2]. Indeed, microinjection of the glutamate receptor agonist

kainic acid in the vicinity of the vIRt is a salutary means to induce prolonged

rhythmic vibrissa movement, near 10 Hz, in the lightly anesthetized rat

(figure 8a). The frequency of whisking decreases over time as the effect of

anesthesia declines, while the frequency of breathing remains a constant basal

rate. This implies that the chemical activation is sufficiently strong to decouple

rhythmic protraction from breathing. The ability to induce whisking in an immobile

animal further provides a means to stably record from units whose firing times

were coherent with rhythmic protraction (figure 8b). We identified neuronal units

Draft of 1 September 2013

7

that spiked in synchrony with protraction, as in the case of the units identified

during intervening whisks in the behaving animal (figure 7c), as well as units that

spiked in anti-phase. Injection of an anterograde tracer in the vIRt led to in

labeling of axon terminals in the ventrolateral part of the facial nucleus, where

motoneurons that innervate the intrinsic muscles are clustered [2].

The above results provide evidence for the sufficiency of neurons in the

vIRt to drive rhythmic protraction. We now consider the necessity of the vIRt for

rhythmic motion and test if a lesion to this zone suppresses whisking [2]. First,

small electrolytic lesions of the vIRt abolish whisking on the side of the lesion,

while whisking persists on the contralateral side (figure 9a). Critically, lesions

within the vIRt that were as small as 200 µm in diameter were sufficient to

severely impair whisking on the ipsilateral side, whereas off-site lesions have

minnimal effects on whisking (figure 9c). Qualitatively similar results were found

with ibotenic acid or Sindbis viral lesions (figure 9c). These data lead to the

conclusion that units in the vIRt play an obligatory role in the generation of

whisking.

The behavioral (figures 5 and 6) and physiological (figures 7 and 8) data

suggest that neurons in the inspiratory CPG reset an oscillatory network of

whisking units in the vIRt that can drive protraction of the vibrissa concurrent with

each inspiration. We assessed this hypothesized connection by tract tracing

methods [2]. Injections of tracer into the preBötzinger complex (figure 10a),

identified by the phase relation of units relative to breathing (figure 10b), led to

dense anterograde labeling of terminals in the vIRt, in the same region where we

observed whisking units and where lesions extinguished ipsilateral whisking

(figure 10c). These results support a direct connection from the preBötzinger

complex to the vIRt.

We next delineated the projections from neurons in the vIRt to facial

motoneurons [2]. Tracer was injected in the lateral aspect of the facial nucleus

(figure 11a), and we observed a number of retrogradely labeled cells in the vIRt

(figure 11b). A detailed map of the location of cells that were retrogradely

Draft of 1 September 2013

8

labeled from this injection reveals the spatial extent of the high-density cluster of

facial projecting vIRt cells (figure 11c). In toto, these and previous [27, 28]

patterns of neuronal labeling in the IRt support a direct connection from the vIRt

to the facial nucleus and substantiate the role of the vIRt as a premotor nucleus.

This zone functions as the premotor pattern generator for rhythmic whisking and

is part of a larger circuit whereby cells in nuclei that are obligatory for inspiration

[25, 29, 30] reset the phase of vIRt units with each breath (figure 12).

We conclude that whisking concurrent with sniffing is effectively driven on

a cycle-by-cycle basis by the inspiratory rhythm generator, while intervening

whisks between successive inspirations result from oscillations of the whisking

units in vIRt [2]. This result bears on the generation of other rhythmic orofacial

behaviors, for which licking is particularly well described. First, tongue

protrusions are coordinated with the respiratory cycle [31]. Second, like the vIRt

facial premotoneurons, a cluster of hypoglossal premotoneurons are

concentrated dorsomedially to the preBötzinger complex within the IRt [24, 32]

and, further, are driven by bursts of spikes that are locked to inspiration [33].

Third, the output of units in the hypoglossal IRt zone locks to rhythmic licking

[34]. Lastly, infusion of an inhibitory agonist into the IRt blocks licking [35]. These

past results are consistent with a model in which preBötzinger units reset the

phase of bursting in a network of hypoglossal premotor neurons in the IRt zone,

in parallel with our circuit for whisking (figure 12).

A final issue concerns the potential binding of touch-based and olfaction-

based sensory inputs. In the vibrissa system, the spike rate of neurons in vS1

cortex (figure 1) is modulated by the phase of the vibrissa in the whisk cycle [36-

40]. In particular, the spike rate of neurons in layers 4 and 5a is most pronounced

when the vibrissae contact an objects at a particular phase [38] (figure 13a,b).

Different neurons have different preferred phases so that all phases in the whisk

cycle, corresponding to contact upon retraction as well as protraction, are

covered. In contrast, the output of the same neurons appears untuned when

contact is plotted as a function of absolute contact angle (figure 13c).

Draft of 1 September 2013

9

Tuning of the neuronal response in terms of phase implies that the rodent

codes touch in a coordinate system that is locked to the CPG for whisking. It is of

interest that neurons in the olfactory bulb tend to spike in phase with breathing,

as opposed to spiking in a manner time-locked to the presentation of an odorant

[41] (figure 14). When rodents are actively exploring, the precise one-to-one

phase locking between whisking and sniffing (figure 5a) could ensure that spikes

induced by both tactile and olfactory stimuli occur with a fixed temporal

relationship to one another, with a delay that corresponds to a particular location

and smell. This implies that sensory inputs from touch, which enter the brain at

the level of the brainstem, and inputs from smell, which enter the brain at the

rostral pole, can in principle be linked by the breathing rhythm via computations

in downstream neurons. Thus, coordination by the output of the preBötzinger

complex can ensure that orofacial behaviors do not confound each other as well

as serve to perceptually bind concurrent tactile and olfactory inputs. Whether the

binding mechanism further involves coherent transient theta-band oscillations in

neocortical and hippocampal circuits remains an open issue [42, 43].

Acknowledgements. Our work was funded by the Canadian Institutes of Health

Research (grant MT-5877), the National Institutes of Health (grants NS058668,

NS066664 and NS047101), and the US-Israeli Binational Science Foundation

(grant 2003222).

Draft of 1 September 2013

10

Reference

1. Welker, W.I., Analysis of sniffing of the albino rat. Behaviour, 1964. 12: 223-244. 2. Moore*, J.D., M. Deschênes*, T. Furuta, D. Huber, M.C. Smear, M. Demers, and

D. Kleinfeld, Hierarchy of orofacial rhythms revealed through whisking and breathing. Nature, 2013. 469: 53-57.

3. Kleinfeld, D. and M. Deschênes, Neuronal basis for object location in the vibrissa scanning sensorimotor system. Neuron, 2011. 72: 455-468.

4. Nguyen, Q.-T. and D. Kleinfeld, Positive feedback in a brainstem tactile sensorimotor loop. Neuron, 2005. 45: 447-457.

5. Deschênes, M., J.D. Moore, and D. Kleinfeld, Sniffing and whisking in rodents. Current Opinion in Neurobiology, 2012. 22: 243-250.

6. Mitchinson, B., C.J. Martin, R.A. Grant, and T.J. Prescott, Feedback control in active sensing: Rat exploratory whisking is modulated by environmental contact. Proceedings of the Royal Society of London: Biological Sciences, 2007. 274: 1035-1041.

7. Towal, R.B. and M.J. Hartmann, Right-left asymmetries in the whisking behavior of rats anticipate movements. Journal of Neuroscience, 2006. 26: 8838-8846.

8. Berg, R.W. and D. Kleinfeld, Rhythmic whisking by rat: Retraction as well as protraction of the vibrissae is under active muscular control. Journal of Neurophysiology, 2003. 89: 104-117.

9. Hill, D.N., R. Bermejo, H.P. Zeigler, and D. Kleinfeld, Biomechanics of the vibrissa motor plant in rat: Rhythmic whisking consists of triphasic neuromuscular activity. Journal of Neuroscience, 2008. 28: 3438-3455.

10. Pietr, M.D., P.M. Knutsen, D.I. Shore, E. Ahissar, and Z. Vogel, Cannabinoids reveal separate controls for whisking amplitude and timing in rats. Journal of Neurophysiology, 2010. 104: 2532-2542.

11. Hill, D.N., J.C. Curtis, J.D. Moore, and D. Kleinfeld, Primary motor cortex reports efferent control of vibrissa position on multiple time scales. Neuron, 2011. 72: 344–356.

12. Ranade, S., B. Hangya, and A. Kepecs, Multiple modes of phase locking between sniffing and whisking during active exploration. Journal of Neuroscience, 2013. 33: 8250-8256.

13. Lawson, E.E., D.W. Richter, M.F. Czyzyk-Krzeska, A. Bischoff, and R.C. Rudesill, Respiratory neuronal activity during apnea and other breathing patterns induced by laryngeal stimulation. Journal of Applied Ohysiology, 1991. 70: 2742-2749.

14. Smith, J.C., H.H. Ellenberger, K. Ballanyi, D.W. Richter, and J.L. Feldman, Pre-Botzinger complex: A brainstem region that may generate respiratory rhythm in mammals. Science, 1991. 254: 726-729.

15. Feldman, J.L., C.A. Del Negro, and P.A. Gray, Understanding the rhythm of breathing: So near, yet so far. Annual Review of Physiology, 2013. 75: 423-452.

Draft of 1 September 2013

11

16. Garcia, A.J., S. Zanella, H. Koch, A. Doi, and J.M. Ramirez, Networks within networks: The neuronal control of breathing. Progress in Brain Research, 2011. 188: 31-50.

17. Gao, P., R. Bermejo, and H.P. Zeigler, Vibrissa deaffentation and rodent whisking patterns: Behavioral evidence for a central pattern generator. Journal of Neuroscience, 2001. 21: 5374-5380.

18. Lovick, T.A., The behavioural repertoire of precollicular decerebrate rats. Journal of Physiology, 1972. 226: 4P-6P.

19. Smith, J.C., A.P.L. Abdala, I.A. Rybak, and J.F.R. Paton, Structural and functional architecture of respiratory networks in the mammalian brainstem. Philosophical Transactions of the Royal Society of London B, 2009. 364: 2577-2587.

20. Sherrey, J.H. and D. Megirian, State dependence of upper airway respiratory motoneurons: Functions of the cricothyroid and nasolabial muscles of the unanesthetized rat. Electroencephalography and Clinical Neurophysiology, 1977. 43: 218-228.

21. Haidarliu, S., D. Golomb, D. Kleinfeld, and E. Ahissar, Dorsorostral snout muscles in the rat subserve coordinated movement for whisking and sniffing. Anatomical Record, 2012: in press.

22. Nakamura, Y. and N. Katakura, Generation of masticatory rhythm in the brainstem. Neuroscience Research, 1995. 23: 1-19.

23. Bieger, D. and D.A. Hopkins, Viscerotopic representation of the upper alimentary tract in the medulla oblongata in the rat: The nucleus ambiguus. Journal of Comparative Neurology, 1987. 262: 546-562.

24. Travers, J.B., L.A. Dinardo, and H. Karimnamazi, Motor and premotor mechanisms of licking. Neuroscience and Biobehavioral Reviews, 1997. 21: 631-647.

25. Tan, W., W.A. Janczewski, P. Yang, X.M. Shao, E.M. Callaway, and J.L. Feldman, Silencing preBötzinger Complex somatostatin-expressing neurons induces persistent apnea in awake rat. Nature Neuroscience, 2008. 11: 538 - 540.

26. Travers, J.B., Oromotor nuclei, in The Rat Nervous System - Second Ediition, G. Paxinos, Editor 1995, Academic Press: San Diego. p. 239-255.

27. Takatoh, J., A. Nelson, X. Zhou, M.M. Bolton, M.D. Ehlers, B.R. Arenkiel, R. Mooney, and F. Wang, New modules are added to vibrissal premotor circuitry with the emergence of exploratory whisking. Neuron, 2013. 77: 346-360.

28. Isokawa-Akesson, M. and B.R. Komisaruk, Difference in projections to the lateral and medial facial nucleus: Anatomically separate pathways for rhythmical vibrissa movement in rats. Experimental Brain Research, 1987. 65: 385-398.

29. Gray, P.A., J.A. Hayes, G.Y. Ling, I. Llona, S. Tupal, M.C. Picardo, S.E. Ross, T. Hirata, J.G. Corbin, J. Eugenín, and C.A. Del Negro, Developmental origin of preBötzinger complex respiratory neurons. Journal of Neuroscience, 2010. 30: 14883-148895.

Draft of 1 September 2013

12

30. Bouvier, J., M. Thoby-Brisson, R. N., V. Dubreuil, J. Ericson, J. Champagnat, A. Pierani, C. A., and G. Fortin, Hindbrain interneurons and axon guidance signaling critical for breathing. Nature Neuroscience, 2010. 13: 1066-1074.

31. Welzl, H. and J. Bures, Lick-synchronized breathing in rats. Physiological Behavior, 1977. 18: 751-753.

32. Koizumi, H., C.G. Wilson, S. Wong, T. Yamanishi, N. Koshiya, and J.C. Smith, Functional imaging, spatial reconstruction, and biophysical analysis of a respiratory motor circuit isolated in vitro. Journal of Neuroscience, 2008. 28: :2353–2365.

33. Ono, T., Y. Ishiwata, N. Inaba, T. Kuroda, and Y. Nakamura, Modulation of the inspiratory-related activity of hypoglossal premotor neurons during ingestion and rejection in the decerebrate cat. Journal of Neurophysiology, 1998. 80: 48-58.

34. Travers, J.B., L.A. DiNardo, and K. H., Medullary reticular formation activity during ingestion and rejection in the awake rat. Experimental Brain Research, 2000. 130: 78-92.

35. Chen, Z., S.P. Travers, and J.B. Travers, Muscimol infusions in the brain stem reticular formation reversibly block ingestion in the awake rat. American Journal of Physiology: Regulatory, Intergrative, and Comparative Physiology, 2001. 280: R1085-1094.

36. de Kock, C.P. and B. Sakmann, Spiking in primary somatosensory cortex during natural whisking in awake head-restrained rats is cell-type specific. Proceedings of the National Academy of Sciences USA, 2009. 106: 16446-16450.

37. Fee, M.S., P.P. Mitra, and D. Kleinfeld, Central versus peripheral determinates of patterned spike activity in rat vibrissa cortex during whisking. Journal of Neurophysiology, 1997. 78: 1144-1149.

38. Curtis, J.C. and D. Kleinfeld, Phase-to-rate transformations encode touch in cortical neurons of a scanning sensorimotor system. Nature Neuroscience, 2009. 12: 492-501.

39. Crochet, S. and C.C.H. Petersen, Correlating membrane potential with behaviour using whole-cell recordings from barrel cortex of awake mice. Nature Neuroscience, 2006. 9: 608-609.

40. Gentet, L.J., M. Avermann, F. Matyas, J.F. Staiger, and C.C.H. Petersen, Membrane potential dynamics of GABAergic neurons in the barrel cortex of behaving mice. Neuron, 2010. 65: 422–435.

41. Shusterman, R., M.C. Smear, A.A. Koulakov, and D. Rinberg, Precise olfactory responses tile the sniff cycle. Nature Neuroscience, 2011. 14: 1039-1044.

42. Berg, R.W., D. Whitmer, and D. Kleinfeld, Exploratory whisking by rat is not phase-locked to the hippocampal theta rhythm. Journal of Neuroscience, 2006. 26: 6518–6522.

43. Lisman, J.E. and O. Jensen, The θ-γ neural code. Neuron, 2013. 77: 1002-1016. 44. Fukuda, Y. and Y. Honda, Differences in respiratory neural activities between

vagal (superior laryngeal), hypoglossal, and phrenic nerves in the anesthetized rat. Japanese Journal of Physiology, 1982. 32: 387-398.

Draft of 1 September 2013

13

Figure Legends Figure 1. The anatomy of the vibrissa somatosensorimotor system. Major

pathways from the vibrissae to the brainstem and up through neocortex are

shown. Abbreviations: PrV, principal trigeminal nucleus; SpVO, SpVI, SpVMu,

and SpVC, spinal nuclei oralis, interpolaris, muralis, and caudalis, respectively;

VPMdm, dorsomedial aspect of the ventral posterior medial nucleus of dorsal

thalamus; Po, medial division of the posterior group nucleus; nRt, nucleus

reticularis; ZIv, ventral aspect of the zona incerta; SC, superior colliculus; and

vIRT, the vibrissa region of the intermediate reticular zone and the central pattern

generator for whisking. Black arrows indicate excitatory projections while red

arrows are inhibitory projections. Adapted from [3].

Figure 2. Decomposition of rhythmic whisking into a varying phase component and slowly varying envelope parameters. (a) Schematic of the

angular parameters and the representation of phase in the whisk cycle. (b) Top

panel shows vibrissa angular position, θ(t). Lower panels show the phase, φ(t),

as calculated from the Hilbert transform, along with the amplitude, θamplitude, and

midpoint, θmidpoint, of the whisking angle calculated from individual whisk cycles.

Broken vertical lines indicate wrapping of phase from π to –π. The red line in the

top panel is the reconstruction calculated from θ = θamplitude•cos[φ] + θmidpoint at

each time point. Adapted from [11].

Figure 3. Effect of a cannabinoid receptor type 1 agonist and an antagonist on whisking kinematics. (a) Typical traces of vibrissa angle executed by an

animal four hours after administration of Δ9-THC (2.5 mg/kg), an agonist (dark

gray), vehicle (black), or SR141716A (2 mg/kg), an antagonist (light gray).

Vertical calibration bar corresponds to 50°. (b) Cumulative probability distribution

functions of protraction amplitudes and whisking durations across all animals.

Adapted from [10].

Draft of 1 September 2013

14

Figure 4. Neurons in vM1 cortex report the amplitude and midpoint of rhythmic whisking. Firing rate profiles for two example units in vM1 cortex as a

function of slowly varying parameters, i.e., amplitude and midpoint, of vibrissa

motion (figure 2). The left and middle columns are profiles of units that show

different relative modulation. Each plot is calculated by dividing the distribution of

the respective signal at spike time by the distribution of that signal over the entire

behavioral session. Green lines are fits from a smoothing algorithm along with

the 95 % confidence band. The right column shows composite data across units

and illustrates that, on average, the rate is unaffected by whisking, consistent

with the presence of units that both increase (green) and decrease (red) their

rate with increasing angle; blue dots correspond to a non-monotonic change.

Adapted from [11].

Figure 5. Simultaneous measurements of vibrissa angular position (blue) and breathing (red). (a) Measurement showing epochs of breathing without

whisking and sniffing with whisking. (b) Measurement showing an epoch of

whisking without breathing. (c) Measurement showing breathing with intervening

whisks between inspirations. These data illustrate phase resetting of whisking by

breathing (figure 6). (insert) Procedure to measure whisking via videography

and breathing via a thermocouple in a head-restrained rat. The additional

chamber is a port for electrode measurements. Adapted from [2].

Figure 6. Reset of whisking by inspiration. Rasters of inspiration onset times

(red) and protraction onset times (blue) relative to the onset of inspiration for

individual breaths are ordered by the duration of the breath; green arrow

accounts for the 30 ms lead of inspiratory drive to facial muscles as opposed to

the measured inspiration [44]. Whisk and inspiration onset times are significantly

correlated during both sniffing and basal respiration. Adapted from [2].

Draft of 1 September 2013

15

Figure 7. Units in the intermediate reticular formation (IRt) that report inspiration versus protraction. (a) Concurrent recordings of breathing (red),

whisking (blue), and multiunit activity (black) in the preBötzinger complex. The

location of the recording site is labeled with Chicago sky blue and is shown in a

sagittal section counterstained with neutral red. LRt denotes the lateral reticular

nucleus, FN the facial nucleus, Amb the ambiguus nucleus, and IO the inferior

olive. (b) Multiunit spike activity in the vibrissa zone of the intermediate reticular

formation. The section is counterstained with neutral red. (c) The recording sites

for all data imposed on a three dimensional reconstruction of the medulla.

Whisking units are located medial to the preBötzinger complex in the IRt.

Adapted from [2].

Figure 8. Injection of kainic acid activates the vIRt, which drives facial motoneurons and induces whisking. (a) Vibrissa motion (blue), breathing

(red), intrinsic (green) and extrinsic (black) electromyogram (EMG). (b) Polar

plots of the coherence between spiking activity and vibrissa motion at the peak

frequency of whisking. Open circles represent multiunit activity and closed circles

represent single units. The green bar represents the coherence of the ∇EMG for

the intrinsic muscle (panel b) with vibrissa motion. (Inserts) Spiking activity of

neuronal units in the vIRt (black) in relation to vibrissa motion (blue). Adapted

from [2].

Figure 9. Lesion of the vIRt impairs ipsilateral whisking. (a) Example of

whisking bout following an electrolytic lesion. (b) Composite histological results

across all lesion sites were mapped onto a three dimensional reconstruction of

the medulla and selected anatomical substructures. The lesion centroids are

denoted with symbols, with circles for electrolytic lesions, triangles for lesion via

transport of Sindbis virus, and squares for chemical lesion by ibotenic acid.

Adapted from [2].

Draft of 1 September 2013

16

Figure 10. Anatomical evidence that preBötzinger units project to the vibrissa zone of the IRt. (a) Recording of a single inspiratory unit in the

preBötzinger complex, together with breathing. (b) Injection of the anterograde

tracer biotinylated dextran amine through the same pipette used to record (panel

a). (c) Labeling of axons and terminals in the vIRt from cells in the preBötzinger

complex. Adapted from [2].

Figure 11. Anatomical evidence that the facial nucleus receives input from the vibrissa zone of the IRt. (a) Injection of retrograde tracer Neurobiotin™

(green) into the facial nucleus (FN). Labeling with α-choline acetyl-transferase

highlights motoneurons in the facial nucleus (red). (b) Retrograde labeling of

neurons in the vIRt (white arrow). Labeling with α-choline acetyl-transferase

highlights neurons in the ambiguus nucleus (red). (c) Compendium of the

locations of cells that were retrogradely labeled from the facial nucleus with

Neurobiotin™, superimposed on a three dimensional reconstruction of the

medulla. Note labeled cells in the vIRt, located between coronal planes spaced

500 µm apart and that span 200 µm along the lateral-medial axis. pFRG denotes

the parafacial respiratory group and PCRt the parvocellular reticular nucleus.

Adapted from [2].

Figure 12. Schematic of the brainstem circuitry that generates whisking in coordination with breathing. The neurotransmitters refer to glycine (Gly),

glutamate (Glu), and γ-aminobutyric acid (Gaba) that were found from in situ

hybridization measurements [2].

Figure 13. Evidence that neurons in vS1 cortex encode contact with an object relative to the phase of the vibrissae in the whisk cycle (a) The

scheme used to measure the spike response of units in vS1 cortex as animals

Draft of 1 September 2013

17

rhythmically whisk first in air then whisk to touch a contact sensor. Vibrissa

position is determined from videography while contact is determined via

displacement of the sensor. A critical aspect of this behavioral task is that touch

is recorded across all phases of the whisk cycle; the case shown here is touch

soon after the onset of retraction (red dot). (b) The left plot shows the peak

values of the touch response as a function of phase in the whisk cycle (left panel in figure 2a). The uncertainty represents the 95 % confidence interval. A smooth

curve through this data defines the phase of maximal touch response, denoted

φtouch (red dot and corresponding dot in panel a). The right plot is the same data

parsed according to the angular position of the vibrissa upon contact. The angle

is measured relative to the midline of the animal’s head (right panel in figure 2a). Unlike the case for phase, there is no significant tuning for angle. Adapted

from [38].

Figure 14. Odor response in olfactory bulb in a representative neuron in the

olfactory bulb of an awake mouse. (a) Schematic of the experiment. A head-fixed

animal was positioned in front of the odor delivery port. It was implanted with

intranasal cannula to derive the breathing cycle from the measured pressure and

a multi-electrode chamber. (b) Raster plots for single unit spikes from a mitral or

tufted neuron in the olfactory bulb in response to an odor stimulus. The data is

displayed as synchronized by odor onset. The blue bands show the respiration

cycles for each trace. (c) The same data as used for panel b after alignment to

the onset of inspiration and temporally warped by breathing, so that it is now in

phase coordinates. The uniformity of the blue bands indicates that the individual

respiration cycles are now aligned. Adapted from [41].

Exaf

fere

nce

Perip

hera

l rea

ffere

nce

Efference

Zona Incerta

ZIv

L3 & L4 L1 & L5a

Sensory (vS1) cortex

Infraorbital branch of trigeminal nerve

Facial motor nucleus

Facial nerve

PrV map

L5b

Trigeminal ganglion

Thalamus

Po

VPMdm

Motor (vM1) cortex

VPMdm Po

Follicles

Mystacial pad(Vibrissa sensorimotor complex)

Follicle muscles

VPMdm map

Trigeminal nuclei

PrVSpVO SpVI

SpVC

nRt

L6

SC

L5b

Brain

stem

vIRt

Efference

copy

Perip

hery

Midb

rain

Endb

rain

Inte

rbra

in

Figure 1. Kleinfeld, Deschenes and Moore

SpVMu

Pad

Pha

se,

φ

Whisking bout

Ang

le, θ

π0

−π

Hilbert transform

Mid

poin

t,θ m

id

Reconstruction

Am

plitu

de,

θ amp

25°

Time1 s

105°

75°45°

a

φ = ±π

φ = 0Maximum protraction

Maximum retractionMidpoint

0o

180o

Vibrissa angle, θ(t) Phase in whisk cycle, φ(t)

b

105°

75°45°

Figure 2. Kleinfeld, Deschenes and Moore

Endocannabinoid agonist

Endocannabinoid antagonist

Figure 3. Kleinfeld, Deschenes and Moore

Control

b

0.25

0.50

0.75

1.00

50 1000Whisking amplitude (deg)

100 2000Whisking period (ms)

a

Time

0

AgonistAntagonist

Control

Cum

ulat

ive

dist

ribut

ion

1 s

50o

Whi

skin

g an

gle

25°15° 35° 25°15°

50° 75° 50°25° 75°0 0

0 0

10 5

10 5

Spi

ke ra

te (H

z)

Average rate (Hz)

0.1

100

10

1.0

0.1 100101.0

0.1

100

10

1.0

0.1 100101.0

Mod

ulat

ion

byam

plitu

de (H

z)M

odul

atio

n by

mid

poin

t (H

z)

Average rate (Hz)

Spi

ke ra

te (H

z)

Midpoint, θmid Midpoint, θmid

Amplitude, θamp Amplitude, θamp

Figure 4. Kleinfeld, Deschenes and Moore

Thermocouple

Camera

Intervening whisks

Breathing w/o whisking

Whisking w/o breathing

20°

0.1°C

Time

Sniffing and whisking

Ammonia

1 s

Figure 5. Kleinfeld, Deschenes and Moore

a

b

c

Figure 6. Kleinfeld, Deschenes and Moore

-0.5 00

0.5 1.0

Basal

Sniff

18000

Bre

ath

num

ber

Time from inspiration onset (s)

Protraction onset

Interveningwhisks

3 Hz

5 Hz

Inspiration onset

Slowestbasal

breathing

Fastestsniffing

Corrected onset

50˚

1 s

Dorsal Dorsal

Ros

tral

Rostral

Late

ral

Late

ral

Inferior olive

Ambiguus

Lateral reticularnucleus

PreBötzinger

Breathing/Whisking

500 µm

•••• •Facial nucleus

Vibrissa IRt

0.5 mV

Sagittal Coronal Horizontal

Intervening whisks

Intervening whisksBreathing/Whisking

Inspiratory/protractionWhisking

a

b

c

Figure 7. Kleinfeld, Deschenes and Moore

25˚

200 µV

1 s

NasalisEMG

IntrinsicEMG

Vibrissaangle

Retraction units in the vIRt

Breathing

�/2

3�/2

00.5

|Coherence|1.00

Phase

(rad

ians)

800 µV

Figure 8. Kleinfeld, Deschenes and Moore

Protraction units in the vIRt200 µV

20º0.1 s

a

b

0.5 s

Ipsilateral side

Contralateral side

Whi

skin

g am

plitu

de

*

*

*

*

*

*o

oo

o

oooooo

o

oo

∆∆

∆∆

IO

Dorsal

Late

ral Severe

deficit

Minimaldeficit

Rostral

Late

ral

IO

Coronal

Ambiguus

1 mm

ooo

oo

o

oo

*

*

*

*

*

*

o

o

oo

o

∆∆

Horizontal

Figure 9. Kleinfeld, Deschenes and Moore

20o

Time

a

b

20o

Inferior olive

Ambiguus

Vibrissaintermediatereticular zone

preBötzinger(injection site)

1 mm 100 µm

Breathing

preBötzinger unit1 s

Figure 10. Kleinfeld, Deschenes and Moore

a

b c

Figure 11. Kleinfeld, Deschenes and Moore

250 µm

Ambiguus

vIRt

InjectionChAT

FN

500 µm

1 mm

IO

Facialnucleus

LRt

vIRt

PCRt

Rostral

Late

ral

Böt

preBötzinger

Inferior olive

Horizontal

a

b

c

Figure 12. Kleinfeld, Deschenes and Moore

FacialReset phase

Intrinsic muscles(vibrissa protraction)

Extrinsic muscles(mystacial pad motion)

Dorsal

Ros

tral

Sensoryfeedback

Inspiration

Modulators

Glut, Gly, Gaba

Midpoint

vIRt

Böt preBötpFRG

Amplitude

130o 160o–π 0 πPhase in whisk cycle Angle in whisk cycle

0

100

200

Spike

rate

at c

onta

ct (H

z)

100o

Single unit50 ms

Touch

Angle120o

160o

Contacta

b

Video

Figure 13. Kleinfeld, Deschenes and Moore

a

b

Figure 14. Kleinfeld, Deschenes and Moore

Ras

ters

alig

ned

byod

or p

rese

ntat

ion

Alig

ned

by o

nset

of in

spira

tion

and

war

ped

to b

reat

hing

Time

Time after onset of inspiration (ms)

Odorized air

Pre

ssur

e(P

a)

−50

0

50

100 ms

Onset ofinspiration

Offset ofinspiration

Phase in breathing cycle

Pressure

transducer

−247 0 247 494

Phase in breathing cycle (ms)−2π 0 2π 4π

c

Amplifier2π0