Our Sense of Direction - Progress, Controversies and Challenges Kathleen E Cullen1 & Jeffrey S Taube2 (2017)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

F O C U S O N S PAT I A L C O G N I T I O N 

perspective

Our sense of direction: progress, controversies and


challenges
Kathleen E Cullen1 & Jeffrey S Taube2
In this Perspective, we evaluate current progress in moved9–12, and subjects can keep track of changes in heading during
understanding how the brain encodes our sense of direction, active motion in the absence of vision5.
within the context of parallel work focused on how early In everyday life, the vestibular system plays a vital role in a wide
vestibular pathways encode self-motion. In particular, we range of functions from reflexes to the highest levels of voluntary
discuss how these systems work together and provide evidence behavior. Notably, as we move through our environment, the ves-
that they involve common mechanisms. We first consider the tibular system explicitly detects and encodes our self-motion in six
classic view of the head direction cell and results of recent dimensions: the semicircular canals and otolith end organs of the ves-
experiments in rodents and primates indicating that inputs tibular system detect three dimensions of rotational head velocity and
to these neurons encode multimodal information during self- three dimensions of translational head acceleration, respectively. This
motion, such as proprioceptive and motor efference copy self-motion information is in turn essential for generating the funda-
signals, including gaze-related information. We also consider mental reflexes that guarantee stable gaze and posture during our eve-
the paradox that, while the head-direction network is generally ryday activities, namely the vestibulo-ocular, vestibulocervical, and
assumed to generate a fixed representation of perceived vestibulospinal reflexes13. Importantly, vestibular information is also
directional heading, this computation would need to be sent to higher centers via ascending pathways. In particular, two main
dynamically updated when the relationship between voluntary thalamocortical pathways—one anterior and the other posterior—
motor command and its sensory consequences changes. transmit vestibular information to the cortex (Fig. 1; reviewed in
Such situations include navigation in virtual reality and head- refs. 14,15). Of the two, the anterior pathway has been more exten-
restricted conditions, since the natural relationship between sively studied. It comprises direct projections from the anterior dorsal
visual and extravisual cues is altered. thalamus to the retrosplenial cortex and multisynaptic projections
to the entorhinal cortex (via presubiculum and parasubiculum), and
it is thought to encode our static sense of direction during naviga-
Sense of direction versus a self-motion-detection system tion and spatial memory tasks16. In contrast, the posterior pathway
Our ability to keep track of where we are going relative to where we originates in the vestibular nuclei and projects to the ventral posterior
have been requires knowledge of our ongoing location and orienta- thalamus, and then on to the somatosensory cortex and the parieto-
tion. In everyday life, this ability, or sense of direction, depends on insular vestibular cortex17. This pathway has been less extensively
the integration of both visual and nonvisual cues, including vestibular studied despite its likely contribution to the accurate coordination of
and proprioceptive information. The visual system provides retinal perception and detection of self-motion (reviewed in ref. 18).
image-motion (optic flow) cues, which have long been known to This anatomical division then raises the question: are these two
provide a powerful sense of motion through an otherwise stationary ascending pathways independent or do they involve common mech-
environment1. In addition, extravisual cues make important contribu- anisms? Below, we first consider the prevailing view that during
tions to navigation and are thought to contribute to path integration navigation, vestibular information transmitted to the anterior tha-
or dead reckoning (Box 1) in the absence of visual cues (for review lamocortical pathway is integrated to provide our sense of direction
see refs. 2,3). Notably, information sent to the brain by the vestibular by providing a critical input to the head-direction cell network. In
nerve allows the estimation of head displacement (both linear and this context, neurons within the anterior thalamic nuclei are tuned
angular) in the absence of visual cues4–8. Further, during locomotion, to specific static head directions. We then consider the implications
proprioceptive and/or motor-related signals also contribute signifi- of recent work establishing that early vestibular pathways integrate
cantly to the estimation of self-motion. For example, estimates of dis- multimodal information to preferentially encode vestibular exaffer-
tance traveled and self-velocity are more accurate when self-motion ence (i.e., motion that is not self-generated). The implications of these
is generated by active locomotion than when subjects are passively recent studies are then considered in relation to both the ascending
sense of direction and self-motion detection pathways.
1Department of Biomedical Engineering, The Johns Hopkins University, Baltimore,
Maryland, USA. 2Department of Psychological & Brain Sciences, Dartmouth College,
A quick review of the classic head direction system
Hanover, New Hampshire, USA. Correspondence should be addressed to Head direction (HD) cells, originally discovered in the rat dorsal pre-
K.E.C. (kathleen.cullen@jhu.edu) or J.S.T. (jeffrey.taube@dartmouth.edu). subiculum (postsubiculum) by James Ranck in 1984 (see foreword19),
Received 2 June; accepted 14 September; published online 26 October 2017; discharge in relation to the animal’s directional heading in the hori-
doi:10.1038/nn.4658 zontal plane, independent of the animal’s location and behavior20,21.

nature neuroscience  VOLUME 20 | NUMBER 11 | NOVEMBER 2017 1465


perspective

Box 1  Terms
Allothetic cues: external sensory cues that can be used as landmarks; these cues are used in an episodic (as opposed to continuous) manner.
Forward model: a process that predicts the expected sensory feedback of a motor command, allowing rapid error detection when actual and predicted feedback do
not match.
Idiothetic cues: internal body cues used to detect how the body and/or head is moving through space; to maintain spatial accuracy these cues must be used in a
continuous (as opposed to episodic) manner; for example, vestibular, proprioceptive, motor efference copy.
Optic flow: the flow of visual stimuli across the retina; this information can be used to determine how objects or the head is moving through the environment.
Path integration (dead reckoning): the process whereby an organism updates its spatial orientation through the use of self-motion cues as it moves
through the environment.
Spatial updating: the process whereby an organism updates its spatial orientation after movement or a period of time has elapsed.

In this regard their firing is analogous to a compass, which always systems, along with information about visual landmarks, work inter-
points in a particular direction (north) no matter where the compass actively and in synch with one another, but situations can arise in
is located, although for HD cells their activity does not appear to be which there is a conflict amongst these different types of sensory and
sensitive to the earth’s geomagnetic field22 but rather to landmark cues motor cues, an issue that we will return to below.
and the earth’s gravity axis21,23. Each HD cell is tuned to a different
direction, and a population of such cells uniformly represents the 360° The generation of a HD signal: vestibular pathways and ring
range. Their firing is thought to encode a continuous representation attractor networks
of the animal’s perceived directional heading. Peak firing rates vary The HD signal can be considered to be composed of two components:
across different cells and can range from a few spikes per s to more (i) the directional signal itself, which is theoretically based on the inte-
than 100 spikes per s. Figure 2a depicts three typical HD cells from gration of self-movement cues, and (ii) the anchoring of this signal
three different brains areas. Note the variations in peak firing rates, to a source, which subsequently defines the reference frame for the
as well as tuning widths, across the three cells, findings that are not signal. To understand how and where the HD signal is constructed,
well understood in terms of their functional roles. The variations researchers have recorded from one brain area following the lesioning
depicted are not necessarily characteristic of a particular brain area or inactivation of a second brain area. Using rats as the experimental
but in general can be observed across all areas where HD cells have
been identified. HD cells are found in many brain areas throughout
the limbic system, and studies have identified several brainstem sites Vestibular cortical Limbic/entorhinal
in rats that are responsible for generating the signal. Self-movement areas (PIVC)
Thalamus
cortex
cues are integrated in these brainstem areas, and the processed HD
signal is propagated rostrally through a network that includes the
anterodorsal thalamus (ADN), dorsal presubiculum, and entorhinal VPL ADN
cortex (reviewed in ref. 19). Visual landmark information is integrated
into this circuit via direct projections from visual cortex to the dorsal
presubiculum (and to a lesser extent, the retrosplenial cortex), which wa
y
th HDC
then exerts top-down control with projections to the lateral mammil- rp
a
network
io
lary nuclei (reviewed in ref. 24). HD cells have been mostly studied in er
st
Po
rats, and to a lesser extent in mice, but they have also been identified y
NPH/ wa
in monkeys25 and indirectly in humans26. SGN ath
rp
Early studies demonstrated the importance of the vestibular system r io
te Thalamocortical
An
for the HD signal, as bilateral neurotoxic lesions of the vestibular laby- VN pathways
rinth abolished the HD signal in the anterior thalamus27. Yet other
studies have shown how self-motion cues that are not vestibular in
origin can influence HD cell activity. For example, while HD cells
retain directional tuning during passive motion, in which only ves- Navigation
self-motion
tibular cues are present28, there can be differences in how HD cells perception
fire between active and passive locomotion29–31. Moreover, when HD
Debbie Maizels/Springer Nature

cells are monitored as rats move from a familiar environment to a


Gaze
novel one (in which they are unfamiliar with the visual landmark Afferents
VN stabilization
cues), the cells’ preferred firing directions remain stable between the (VOR)
Vestibular
two environments when animals are allowed to actively locomote to labyrinth
the novel environment32,33 but not when they are passively moved
Postural
(via a wheeled cart) to the novel environment33,34. Thus, despite the stabilization
presence of a normal vestibular signal when the animals are pas- (VSR)

sively transported, HD cells require information from other types of


movement signals (proprioception, motor efference copy) to main- Figure 1  Overview highlighting functions and projections of the
tain a stable orientation between a familiar and a novel environment. two ascending vestibular pathways: (i) the anterior vestibulothalamic
pathway through the nucleus prepositus and supragenual nucleus (NPH/SGN)
Further, other studies that have manipulated optic flow cues (Box 1)
to the HD network (blue; see Fig. 2 for details) and (ii) the posterior
have shown that visual motion cues can influence HD cell activity35,36, vestibulothalamic pathway through the ventral posterior lateral nucleus
as well as make hippocampal place cells that are nondirectional in a (red). PIVC, parieto-insular vestibular cortex; VN, vestibular nuclei; VPL,
natural environment more directional37. Normally, these movement ventral posterior lateral thalamic nucleus.

1466 VOLUME 20 | NUMBER 11 | NOVEMBER 2017  nature neuroscience


perspective

a 16 Presubiculum b Visual Other


cortex cortical areas
Firing rate (spike per s) 12

8 Entorhinal
Retrosplenial Dorsal cortex
4 cortex presubiculum
Granular Dysgranular

0
0 60 120 180 240 300 360 Hippocampus
Anterodorsal
Head direction (degrees) thalamic
nucleus
Anterodorsal thalamus
60

50 Lateral
Firing rate (spike/s)

mammillary
40 nucleus
30 HD signal
generation
20
Dorsal
10 tegmental Interpeduncular
nucleus nucleus
0
0 60 120 180 240 300 360 ?
Head direction (degrees)
Nucleus Supragenual
120 Lateral mammillary prepositus nucleus

Debbie Maizels/Springer Nature


100
Firing rate (spike/s)

Inhibitory
80 Excitatory
Medial PGRNd
60 HD cells
vestibular
AHV cells
nucleus
40 Place cells
20 Grid cells

0 Vestibular Proprioception
0 60 120 180 240 300 360
labyrinth Motor efference
Head direction (degrees)

Figure 2  Representative plots for HD cells and HD circuit. (a) Three typical HD cells are shown across three different brain areas. Peak firing rates in
each cell’s preferred firing direction can range from low (top) to medium (middle) to high (bottom) across different cells. Cells with low and high peak
firing rates can be observed in all brain areas. (b) Circuit diagram showing principal connections of areas containing HD, angular head velocity, place,
and grid cells. The bracket shows the site of the postulated ring attractor network that generates the HD signal. Red dashed line shows point in network
where lesions to brain areas below this level induce burst firing patterns in recorded ADN neurons; lesions above this level lead to the loss of the HD
signal in ADN without the presence of burstiness among the recorded neurons.

model, these studies have highlighted a circuit starting in the ves- brainstem nuclei, which all contain neurons whose firing also cor-
tibular nuclei and projecting through a series of brainstem nuclei, relates to angular head velocity40–42.
including the nucleus prepositus, supragenual nucleus, and dorsal The most parsimonious computational models that have been pro-
tegmental nucleus, and then projecting to the lateral mammillary posed to account for HD cell firing use continuous ring attractor
nucleus and ADN thalamus19,38. A large percentage of the neurons networks, which receive input from cells sensitive to angular head
within the ADN can be classified as ‘classic’ HD cells, where the signal velocity (reviewed in ref. 43). These networks contain interconnected
is quite robust with a high signal-to-noise ratio. From the ADN, the neurons, such that HD cells with similar preferred firing directions
HD signal is projected rostrally to a number of areas, each of which have recurrent excitatory connections with one another, and HD cells
contains HD cells, including the presubiculum and parasubiculum, with opposing preferred firing directions inhibit one another. Once
retrosplenial, entorhinal and medial precentral cortices, and striatum activity is initiated, the network can sustain activity without outside
(Fig. 2b). There is good evidence that inputs from early vestibular excitation. The local area of activity (referred to as the activity hill)
pathways are critical for the HD signal, as various types of vestibular is moved around the ring to different directional headings following
manipulations that interfere with a normal vestibular signal lead to inputs from self-motion-based (idiothetic; for example, angular head
the loss of the HD signal. For example, lesions, inactivation, or occlu- velocity cells) or allothetic-based (landmark) sources (Box 1).
sions of the semicircular canals, as well as experiments in horizontal Based on studies that have investigated the anatomical substrate
canal-deficient mice, all lead to the loss of head-direction-specific of the HD signal, researchers have hypothesized that a ring attrac-
activity when recording from the ADN or postsubiculum (reviewed in tor network may operate across the reciprocal connections between
refs. 16,39). Given that the vestibular system conveys information the dorsal tegmental nuclei and the lateral mammillary nuclei
about how head orientation changes with head movements, the ves- (Fig. 2b)39,44,45; cf. ref. 46. Consistent with this view are findings from
tibular system is an ideal candidate to provide information for updat- lesion and recording studies, which show that where in the HD cir-
ing the animal’s resultant head orientation following a head turn. cuit a lesion was placed determined the type of response observed in
Specifically, the prevailing view is that information about the animals’ downstream recording areas (Fig. 2b). Theoretically, a lesion placed
angular head velocity, which is encoded by neurons within the ves- within the attractor network, or any area efferent to it, should lead
tibular nuclei, is conveyed from here to various cell types within the to disrupted firing in neurons further downstream to the lesion.

nature neuroscience  VOLUME 20 | NUMBER 11 | NOVEMBER 2017 1467


perspective

In contrast, a lesion placed afferent to the site of the attractor network However, this comparable activity is not the case at the next stage of
will leave the network’s internal connections intact, but the external vestibular processing in the vestibular nuclei. Instead, the responses of
inputs will no longer be able to drive or reset the activity hill; under both categories of vestibular nuclei neurons—VO neurons and VOR
these conditions, the network will be ‘untethered’ and the activity neurons—are modulated in a behaviorally dependent manner.
hill will drift around the attractor ring. Precisely such responses were
found when lesions were made to brain areas associated with the Preferential encoding of passive versus active self-motion in the
ascending vestibulothalamic pathway (i.e., nucleus prepositus, supra- vestibular nuclei
genual nucleus; Fig. 2b)16,39. Experiments in head-unrestrained monkeys have established that
the sensory responses of VO neurons are suppressed during actively
The vestibular system integrates multimodal information and generated movements (Fig. 3b)58–61. Specifically, while VO neurons
motor signals robustly respond to passive head movements their responses are
Prior studies have demonstrated that multimodal information, including markedly (~70%) attenuated during active head movement. This is
proprioceptive, optic flow, and motor efference copy signals, can the case for both rotational and translational motion, as well as across
influence HD cell activity, yet to date the question of how HD cells different species of monkeys (reviewed in ref. 62) and in mice48. More
actually integrate vestibular signals with other self-motion cues recent experiments controlling the association between intended and
remains open. Notably, as detailed above, the type of information actual head movement have established that a vestibular cancellation
encoded by vestibular afferents is ideally suited to act as a spatial signal is exclusively generated when the activation of neck proprio-
updater for the HD network, and the prevailing view is that neu- ceptors matches the motor-generated expectation63,64.
rons within the vestibular nuclei relay this information to the HD Because VO neurons mediate vestibulospinal reflex pathways, the
cell network. However, the vestibular nuclei do much more than just finding that their responses are suppressed during active motion has
receive and encode a vestibular signal (i.e., inputs from the vestibular important implications for voluntary motor control versus balance.
sensory organs). Indeed, these nuclei integrate multimodal informa- Specifically, while the vestibulospinal reflex is vital for stabilizing pos-
tion originating from a number of different sources including head, ture in response to unexpected movements, the stabilizing commands
body, and eye (gaze) motor commands, neck proprioceptive informa- produced by an intact vestibulospinal reflex would be counterproduc-
tion, and visual inputs in both rodents47,48 and primates (reviewed in tive during voluntary movements. Accordingly, turning off vestibu-
refs. 49,50). Thus, a remaining challenge is understanding how these lospinal reflexes is functionally advantageous. Further, because these
different sources of information contribute to the HD signal genera- central vestibular neurons send ascending projections to the ventral
tion, as well as the implications of this integration for understanding posterior thalamus, it follows that the self-motion signals relayed to
the nature of the HD signal itself. the higher-level cortical areas responsible for perception are attenu-
Vestibular nuclei neurons can be grouped into two main catego- ated during active-motion vestibular information. Accordingly, these
ries on the basis of differences in their sensitivity to eye motion findings are problematic for the classic view of how the HD signal is
and passive head motion, as well as connectivity. The first category, generated, resulting in the following conundrum: if the main ascend-
vestibular-only (VO, alternatively termed non-eye-movement) neu- ing vestibular pathway from the vestibular nuclei does not robustly
rons, contribute to posture and self-motion perception. Specifically, encode self-motion during active movements because the signal is
these neurons receive direct inputs from the vestibular nerve and dramatically suppressed, how is the HD signal computed?65,66
in turn drive vestibular spinal reflexes via their direct and indirect Finally, the finding that the vestibular pathway does not encode
projections to the spinal cord (reviewed in ref. 50). In addition, VO self-motion during active movements raises the question of whether
neurons are a source of vestibular input to thalamocortical path- HD cell responses differ between active versus passive head move-
ways51–53 and are reciprocally interconnected with the nodulus and ments. Some studies have shown that the HD signal is attenuated
uvula of the vestibular cerebellum54. The second category of vestibu- during passive head turns, particularly when the animal is tightly
lar nuclei neurons are vestibulo-ocular reflex (VOR) neurons. These restrained in a towel29–31, but other studies using head-fixed animals
neurons also receive direct inputs from the vestibular nerve and have shown that HD cells fire similarly for both active and passive
project to extraocular motor neurons, thereby driving compensatory motion—at least for cells in the ADN thalamus28. Thus, we return to
VOR eye movements to ensure stabile gaze during self-motion. The the initial quandary: if vestibular activity at the level of the vestibular
majority of VOR neurons are the position-vestibular-pause (PVP) nuclei is largely suppressed during active head movements, how does
neurons, a distinct group of neurons deriving their name from the the vestibular system drive HD cell discharge?
signals they carry during passive head rotations and eye movements.
A second subclass of VOR pathway neurons, floccular target neu- Conundrum: building HD signal during active head movements
rons (FTN), receives input from the flocculus of the cerebellum as One possible answer to the conundrum raised above is that other
well as from the vestibular nerve. FTNs similarly encode eye- as self-motion cues carried by VO neurons during active movements are
well as head-related signals, and their responses complement those integrated by the HD network to compute an estimate of direction.
of PVP neurons by ensuring the VOR remains calibrated during While this is not the case in rhesus monkeys60,61, recordings from
daily activities. the vestibular nuclei in mice have shown that, in contrast to monkey
During everyday activities, such as gaze behavior and locomotion, neurons, VO neurons in rodents encode a static neck-position signal
the vestibular system is activated as a result of our own actively gener- in both active and passive conditions48. Notably, this propriocep-
ated motor commands (Fig. 3a). Single-unit recording experiments tive-derived and/or neck motor signal appears to remain robust (i.e.,
in monkeys have established that the afferent activity in the eighth unchanged) for active as well as passive movements. Nevertheless, it
cranial nerve is the same for both actively self-generated movement remains the case that, in both mice and monkeys, the signals coded
and externally applied movement, during both rotations and trans- by VO neurons significantly differ between active and passive condi-
lations55–57. Further, these studies have shown that activity in the tions, with neurons displaying less modulation during active move-
eighth nerve is also independent of the animal’s current gaze behavior. ments compared to during passive ones. This finding then raises the

1468 VOLUME 20 | NUMBER 11 | NOVEMBER 2017  nature neuroscience


perspective

related possibility that the VO neuron signal is decoded by the HD a Cerebellum

network in a context-dependent manner. Specifically, while the head- Internal


motion-related modulation of VO neurons is markedly suppressed model
Estimate of
during active head motion in both species, it is not completely can- sensory feedback
Actual sensory feedback
celled. Thus, theoretically, this attenuated input could be up-weighted during self-motion Match
in the active condition and/or combined with other available inputs
Cancellation
at the level of the HD network (for example, proprioceptive-derived signal
signals and/or motor efference). This possibility raises the questions Vestibular input
Selective
coding of
of whether it is the case and if so, what circuits underlie the required • passively–applied
Vestibular
sensors
Vestibular
nuclei
passive
• actively–generated vestibular
up-weighting and/or integration? stimulation

Eye-movement information and the HD network? b Normal conditions c Learning


As reviewed above, the vestibular nuclei contain a second category (match) (initial mismatch)
Passive Active Early Middle Late
of neurons in addition to VO neurons, namely VOR (i.e., PVP and (1–5) (26–30) (46–50)
FTN) neurons. These neurons encode eye- as well as head-motion

100° per s 50 spikes per s


Head
information and project to oculomotor centers to control VOR eye velocity
movements. Unlike VO neurons, VOR neurons do not preferentially
encode passive head motion, thus raising the question of whether 100 ms

inputs from these neurons might be combined by the HD network to Firing


compute an estimate of HD. However, these neurons integrate eye- rate

movement commands with vestibular information such that they


encode head motion in a manner that depends fundamentally on Time
the animal’s current gaze strategy. Specifically, the responses of PVP
neurons are suppressed whenever animals redirect their gaze (i.e.,
during saccades and smooth pursuit) by an inhibitory gaze-command
Internal
signal. Similarly, FTN neurons carry gaze-related as well as vestibular Initial model
information. Thus, while these neurons could theoretically provide mismatch updated

head-movement-related inputs to the HD network, they would also


transmit gaze-related inputs. d
Indeed, VOR neurons project to the nucleus prepositus, a structure
implicated in the circuit connecting the vestibular nuclei to HD net-

Debbie Maizels/Springer Nature


work (Fig. 2). The nucleus prepositus is also known as the oculomotor
integrator, due to its central role in shaping eye movements (for review,
see ref. 67). Accordingly, understanding the role played by the nucleus
prepositus, an area in which lesions have been shown to disrupt the HD
signal68, is complicated by the fact that this structure also contributes to
controlling gaze. Moreover, this observation raises the possibility that
the ascending HD signal from the vestibular nuclei to the ADN thala-
mus does not exclusively transmit angular head-direction information
derived from the vestibular labyrinth but also encodes eye-position Figure 3  Learning new relationships between motor commands and
information. In sum, none of the three classes of output neurons in the sensory feedback during self-motion.(a,b) A cancellation signal is sent
monkey vestibular nuclei (VO, PVP, and FTNs), neurons that would be to the vestibular nuclei when sensory feedback matches the expected
sensory consequence of motor command (a), resulting in the suppression
the potential relays to the HD network, actually encode purely vestibular
of responses to self-generated vestibular stimulation at the first central
information during voluntary behavior (active head movements); rather stage of processing in the vestibular nuclei (b; compare blue and green
they are either attenuated during voluntary self-motion or encode some traces). (c) When the normal relationship between the motor command
type of eye-position signal along with any head-movement signal. and resultant movement is altered for active head movements, there is
In monkeys, the vast majority of nucleus prepositus neurons an initial mismatch (red arrow) between expected and actual sensory
respond only to eye movements, with a subset (~20%) responding to feedback. As a result, these vestibular neurons robustly respond to self-
both eye movements and vestibular stimulation69–71. Thus, the pri- generated vestibular stimulation (dashed red trace shows predicted
response based on sensitivity to passive vestibular stimulation). Then, after
mate nucleus prepositus predominately outputs eye-in-head position a learning phase (gray box), the brain updates its model of the expected
and velocity information, even during voluntary head motion (for sensory consequence of the motor command (green arrow). (d) In head-
example, see ref. 69). Given that the nucleus prepositus targets regions fixed or head-restrained VR-based protocols there is effectively a mismatch
projecting multisynaptically to the presubiculum 19, a key question between multisensory feedback (both vestibular and proprioceptive) and
arises: to what extent is the HD cell an indicator of the animal’s head optic flow. Dynamic updating of the brain’s model of the expected sensory
direction, as opposed to the direction of its gaze? Unfortunately, to consequence of the motor command requires tracking the comparison of
the predictive and actual sensory feedback signals.
date, no studies in rodents have recorded HD cells while simultane-
ously tracking eye position, so it is currently unknown whether rodent
HD cells respond to changes in gaze direction. We must therefore
turn our attention to primate studies that have recorded from hip- In a series of studies in monkeys, Rolls and colleagues mostly iden-
pocampal and parahippocampal areas while monitoring eye position tified spatial-view cells that responded when a certain part of the
and gaze direction. environment was in the animal’s field of view, as opposed to location

nature neuroscience  VOLUME 20 | NUMBER 11 | NOVEMBER 2017 1469


perspective

or HD72,73. The majority of these cells were localized to the hippocam- expected and actual sensory feedback. This constraint is particularly
pus rather than areas now known to contain HD cells. More recently, relevant to studies of navigation based on experimental designs in
recordings from the entorhinal cortex in head-fixed monkeys viewing which this relationship is mismatched relative to what occurs in eve-
images on a video monitor showed that many cells fired based on the ryday life. For example, in virtual reality (VR) the lack of normal
direction of the monkey’s saccadic eye movements74. It is notable, multisensory feedback (both vestibular and proprioceptive) during
however, that this signal appears to be based on an egocentric refer- VR effectively results in a mismatch condition between the system’s
ence frame, rather than the allocentric one seen for rodent HD cells. internal state and resultant optic flow (Fig. 3d). Specifically, the infor-
Nonetheless, together the above studies demonstrate the importance mation about the subject’s movement based on visual information
of eye position- and eye movement-related information as important would conflict with vestibular and proprioceptive cues. Furthermore,
determinants for cell firing—at least for monkeys. In contrast, other there would also be a mismatch between the expected and actual
studies have recorded cells in monkeys that resemble HD cells that sensory vestibular and proprioceptive feedback experienced in studies
use an allocentric framework. Robertson et al.25 originally identified using head-fixed or head-restrained protocols. Thus, in this context,
five such cells in parahippocampal areas (four in presubiculum and knowing one’s perceived directional heading in the real world versus
one in entorhinal cortex), and more recently, others have identified in an immersive world (i.e., in a video game), requires that the brain
HD cells in the anterior thalamus75. It is noteworthy, however, that learns to switch back and forth between states in which the relation-
the proportion of HD cells in these studies was small (~5%) in com- ships between these cues are different. Further experiments will be
parison to rodent studies, which found 50% and 25% of cells in the required to understand the circuits and mechanisms that underlie
ADN and postsubiculum, respectively20,30. Furthermore, while the this ability.
HD signals recorded in these primate studies were independent of
eye position, whether rodent HD cells encode true HD and not gaze- Implications for navigation in immersive environments
direction responses remains an open question. The considerations discussed above raise the fundamental question of
how HD cells respond in mismatched VR and head-restrained condi-
Rapid updating of a forward model underlies the suppression of tions versus during actual navigation in the real world. Specifically,
vestibular signals during active movement are the pathways that encode our sense of direction updated in such
The suppression of vestibular input resulting from active motion is conditions to reflect a modified forward model that no longer expects
common across species including mice48, squirrel monkeys76, and the same sensory feedback from extravisual sources (i.e., vestibular
macaque monkeys (rhesus60 and fasicularis77). As mentioned above, and proprioceptive), as would be experienced under natural condi-
recent studies in rhesus monkeys that focused on understanding the tions? And if so, what is the time course of this updating and what
mechanism underlying this cancellation suggest that the brain gener- are its implications for understanding how the brain encodes our
ates a sensory expectation based on its motor output, which is then sense of direction? Can this updating be observed in HD cells? If so,
compared with the actual sensory input (reviewed in refs. 49,50). in contrast with common wisdom, this outcome could imply that the
Specifically, vestibular responses are suppressed at the level of the network generating the HD cell does not generate a constant signal
vestibular nucleus when there is a match between the expected conse- representing the subject’s perceived directional heading in the real
quences of self-generated movement and resultant sensory feedback63 world but rather a perception of their directional heading in the real-
(Fig. 3a). Moreover, this mechanism appears to be cerebellar-depend- ity they are immersed in, whether it is a VR environment or during
ent, and it allows selective encoding of passive self-motion at the sleep and dreaming.
earliest central stage of vestibular processing78,79. How spatial cells behave under VR conditions is an important issue
A recent experiment emphasized the importance of prior experi- because most spatial studies in humans employ functional imaging
ence in determining what constitutes a match between the expected techniques on subjects who are performing a spatial task while lying
sensory consequences of self-generated movement and the actual supine with their head fixed and viewing a video monitor—meaning
sensory feedback during motion 78. As monkeys made voluntary that there is no active vestibular signal, motor-efference feedback, or
head movements, the relationship between the motor command proprioceptive feedback. These issues have been reviewed in more
produced to generate the movement and the actual resultant move- detail80,81, but the bottom line is that navigation in the virtual world
ment was altered to produce a mismatch. Specifically, constant is quite different than navigation in the real world, particularly in
resistive torque was applied that initially reduced the monkey’s terms of the sensory and motor systems that are activated, which has
head motion by half. In this condition, vestibular nuclei neurons important implications for interpreting underlying mechanisms.
first responded as if head movement was externally generated, What we do know, based on numerous behavioral studies, is that
but the cancellation of active vestibular input re-emerged over the accurately performing a spatial task using a video monitor versus
same time course as the ongoing behavioral learning (~40–50 head navigating successfully are two different things. For instance, rodents
movements; Fig. 3c). Indeed, this result provides clear evidence for can accurately perform a spatial task, even when their HD cell system
the rapid updating of a forward model (Box 1) that enables the ves- is not functional82, suggesting that different spatial tasks employ dif-
tibular system to learn to expect unexpected head motion inputs. ferent computations that most likely require multiple brain areas and
Accordingly, what constitutes a match between (i) the expected neural systems. Establishing how and where these systems interact
sensory consequences of self-generated movement and (ii) the and are integrated is critical for understanding the neural mechanisms
actual sensory feedback is continually updated as a function of underlying accurate navigation in both the real world and in VR con-
recent experience. ditions. The extent to which accurate performance can be maintained
In this context, it is noteworthy that the HD network is generally in either desktop or VR spatial tasks in the face of disruption to the
assumed to generate a stable representation of the animal’s current or HD system, and by default the grid cell system, which is dependent
anticipated head direction. Yet the expected correspondence between on the HD system83, is not known.
the sensory and motor cues that are inputs into this network require In this context, it is interesting to consider how HD cells might
updating whenever there is a change in the relationship between the fire when a subject is immersed in a video game in which they adopt

1470 VOLUME 20 | NUMBER 11 | NOVEMBER 2017  nature neuroscience


perspective

the spatial perspective of the avatar but still maintain an awareness refs. 88,89). HD cells in bats, a species that routinely moves in a three-
of their orientation with the surrounding environment. Under these dimensional world, display three-dimensional tuning and maintain
circumstances, would HD signals adopt the reference frame related directional tuning when inverted90. In contrast, while rat HD cells
to the video game or to the surrounding real world? Would HD cells display directional tuning in the vertical plane, their tuning is lost
across all brain areas adopt the same reference frame simultaneously when the rat is inverted82,91. One hypothesis proposed to explain
and dynamically switch between representations, or would the HD these results is that HD cells define their reference frame based on
cell population split into different representations? Recent recordings the animal’s plane of locomotion such that the vertical plane is rep-
from the rat retrosplenial cortex have found different types of HD resented as an extension of the floor89. An alternative hypothesis,
cells, some of which are tied to the global environment and some of referred to as the mosaic hypothesis, is that the animal uses two rules
which are tied to the local environment84. Remarkably, a third cell for updating their directional heading: (i) a rule for yaw rotations
type, located only in the dysgranular retrosplenial cortex, encoded around the dorsal–ventral axis and (ii) a rule for rotations of the ani-
both local and global reference frames concurrently, thus having two mal’s dorsal–ventral axis around the gravity-defined vertical axis92.
preferred firing directions in the same environment and potentially Notably, neither of these proposals can account for the absence of
providing a basis for how we simultaneously maintain multiple spatial directional firing when the animal is inverted.
representations for different environments. Recently, Shinder and Taube23 recorded HD cells from head-fixed
Similarly to immersion in a video game, one might consider how rats as they were passively positioned in different orientations in all
HD cells respond when imagining a situation that involves navigation: three rotational planes. Responses in large part could be explained
for example, when sitting at home and thinking about how you might in relation to the animal’s position relative to the earth’s gravitational
navigate between two places that are distant from your home. Again, axis. Using a reference frame that is at least partially defined by gravity
would HD signals adopt the reference frame related to the imagined would account for the finding that disruption of the otolith organs,
locations or to the surrounding real world? Acquiring answers to which are sensitive to gravity, impairs HD cell activity93. Indeed,
this question will be important to fully understand the brain mecha- linking the reference frame to gravity is consistent with recent
nisms underlying spatial orientation and navigation. A few studies findings in monkeys that reported neurons in the anterior thalamus
have begun to address this issue. For instance, Peyrache et al.46 moni- tuned to pitch-and-roll orientation relative to gravity94. Finally,
tored multiple HD cells simultaneously during sleep and found that vestibular nuclei neurons show strong attenuation to otolith-related
their firing retained the temporal correlation structure between each input during voluntary changes in head position 95. Thus, taken
other. Further, network activity that occurred during the awake state together, while the vestibular system, at the level of both the semi-
was replayed during sleep at a 10× higher rate, suggesting that HD circular canals and the otolith organs, appears deeply involved in the
cell activity is not always tied to real world events. Whether HD cell HD signal, future experiments will be required to understand how
activity across other brain areas will contain these same properties is vestibular and extravestibular sensory signals, as well as motor inputs,
an interesting question, as one could imagine that certain brain areas are integrated to build an estimation of three-dimensional orientation
may only be associated with real-world events, while other brain areas during active navigation.
are more involved in imaginative states.
It is important to consider these findings in light of what’s known Outlook and conclusions
about hippocampal place cell activity in VR, where results have been In sum, to date there has been much progress in understanding the
less consistent. Some labs have reported robust place cell firing in VR neural mechanisms underlying our sense of direction, as well as how
compared to the real-world85, while others have reported either an early vestibular pathways encode self-motion. Disruption of the
absence of such firing or a lower percentage of cells showing loca- peripheral vestibular system impairs the directional firing patterns
tion-specific firing86,87. This discrepancy is not due to species differ- of HD cells27, as well as theta rhythm in the hippocampus (reviewed
ences, because all the studies used rats. Currently, the most plausible in ref. 96) and medial entorhinal cortex97. Further the grid-cell signal
reason(s) is that substantial vestibular, proprioceptive, and motor in the medial entorhinal cortex is dependent on both an intact HD
information is needed in order to observe robust place-like activity. signal83 and a theta signal98,99 and is believed to be intimately involved
In the studies involving less location-specific activity, the rats were in navigational processes. Taken together, it follows that the vestibular
either head- or body-fixed, which limited the amount of vestibular system is also critically involved in the generation of the grid-cell sig-
and proprioceptive stimuli. In contrast, in the study that observed nal, although this point has not yet been tested directly. Interestingly,
robust place cell firing85, which also reported the presence of HD cells the distinction between the effects of active versus passive motion
and grid cells, the rats were in a harness, and though they were con- on vestibular neurons is recapitulated for grid cells, in the sense that
fined in place (because their limb movements were directed to move active movement is required for a normal grid cell signal100.
a track ball), they were relatively free to rotate their head and body. Neurons at the first central stage of vestibular processing encode
Thus, although there was an absence in activating the linear compo- proprioceptive, gaze, and head motor-efference copy, as well as ves-
nent of the vestibular system (otoliths), the angular component of the tibular signals—all cues that affect both HD and grid cell firing.
vestibular system (canals) and proprioceptors in the neck and body Moreover, this multimodal information is transmitted in a behavio-
would have been activated normally. In sum, if both vestibular and rally dependent manner that is rapidly updated when the relationship
proprioceptive stimuli are needed in order to activate normal spatial between voluntary motion and the actual sensory input to the brain
firing, then monitoring subjects in VR conditions where these stimuli changes. Thus, while the ascending vestibular pathway through the
are limited may not reflect an accurate representation of the underly- anterior thalamus is likely important for navigation, as it conveys the
ing neural processes that occur during real world navigation. HD signal rostrally, there are open questions regarding how this path-
way utilizes vestibular signals versus other sensory cues and motor
HD cells in three dimensions and the vestibular system signals during navigation. In contrast, the ascending vestibular path-
Over the past decade, several studies in rodents have examined way through the ventral posterior thalamus is likely more involved
how HD cells respond outside the horizontal plane (reviewed in with one’s perceived sense of self-motion (whether such motion is

nature neuroscience  VOLUME 20 | NUMBER 11 | NOVEMBER 2017 1471


perspective

relative to a visual scene on a video monitor or relative to the exter- 26. Shine, J.P., Valdés-Herrera, J.P., Hegarty, M. & Wolbers, T. The human retrosplenial
cortex and thalamus code head direction in a global reference frame. J. Neurosci.
nal world). Where and how these two ascending vestibular pathways 36, 6371–6381 (2016).
are merged and integrated in the cortex remains an open question, 27. Stackman, R.W. & Taube, J.S. Firing properties of head direction cells in the rat
and further appreciation of this integration will ultimately be key to anterior thalamic nucleus: dependence on vestibular input. J. Neurosci. 17,
4349–4358 (1997).
understanding the neural mechanisms underlying our sense of direc- 28. Shinder, M.E. & Taube, J.S. Active and passive movement are encoded equally
tion and accurate navigation in both the three-dimensional real world by head direction cells in the anterodorsal thalamus. J. Neurophysiol. 106,
and in VR conditions. 788–800 (2011).
29. Knierim, J.J., Kudrimoti, H.S. & McNaughton, B.L. Place cells, head direction
cells, and the learning of landmark stability. J. Neurosci. 15, 1648–1659
COMPETING FINANCIAL INTERESTS (1995).
The authors declare no competing financial interests. 30. Taube, J.S. Head direction cells recorded in the anterior thalamic nuclei of freely
moving rats. J. Neurosci. 15, 70–86 (1995).
Reprints and permissions information is available online at http://www.nature.com/ 31. Zugaro, M.B., Tabuchi, E., Fouquier, C., Berthoz, A. & Wiener, S.I. Active
reprints/index.html. Publisher’s note: Springer Nature remains neutral with regard to locomotion increases peak firing rates of anterodorsal thalamic head direction
jurisdictional claims in published maps and institutional affiliations. cells. J. Neurophysiol. 86, 692–702 (2001).
32. Taube, J.S. & Burton, H.L. Head direction cell activity monitored in a novel
1. Gibson, J.J. The perception of visual surfaces. Am. J. Psychol. 63, 367–384 environment and during a cue conflict situation. J. Neurophysiol. 74, 1953–1971
(1950). (1995).
2. Loomis, J.M., Blascovich, J.J. & Beall, A.C. Immersive virtual environment 33. Yoder, R.M. et al. Both visual and idiothetic cues contribute to head direction
technology as a basic research tool in psychology. Behav. Res. Methods Instrum. cell stability during navigation along complex routes. J. Neurophysiol. 105,
Comput. 31, 557–564 (1999). 2989–3001 (2011).
3. McNaughton, B.L., Battaglia, F.P., Jensen, O., Moser, E.I. & Moser, M.B. Path 34. Stackman, R.W., Golob, E.J., Bassett, J.P. & Taube, J.S. Passive transport disrupts
integration and the neural basis of the ‘cognitive map’. Nat. Rev. Neurosci. 7, directional path integration by rat head direction cells. J. Neurophysiol. 90,
663–678 (2006). 2862–2874 (2003).
4. Berthoz, A., Israël, I., Georges-François, P., Grasso, R. & Tsuzuku, T. Spatial 35. Arleo, A. et al. Optic flow stimuli update anterodorsal thalamus head direction
memory of body linear displacement: what is being stored? Science 269, 95–98 neuronal activity in rats. J. Neurosci. 33, 16790–16795 (2013).
(1995). 36. Blair, H.T. & Sharp, P.E. Visual and vestibular influences on head-direction cells
5. Glasauer, S., Amorim, M.A., Vitte, E. & Berthoz, A. Goal-directed linear locomotion in the anterior thalamus of the rat. Behav. Neurosci. 110, 643–660 (1996).
in normal and labyrinthine-defective subjects. Exp. Brain Res. 98, 323–335 37. Acharya, L., Aghajan, Z.M., Vuong, C., Moore, J.J. & Mehta, M.R. Causal influence
(1994). of visual cues on hippocampal directional selectivity. Cell 164, 197–207
6. Grasso, R., Ivanenko, Y. & Lacquaniti, F. Time course of gaze influences on postural (2016).
responses to neck proprioceptive and galvanic vestibular stimulation in humans. 38. Biazoli, C.E. Jr., Goto, M., Campos, A.M. & Canteras, N.S. The supragenual
Neurosci. Lett. 273, 121–124 (1999). nucleus: a putative relay station for ascending vestibular signs to head direction
7. Israël, I., Grasso, R., Georges-Francois, P., Tsuzuku, T. & Berthoz, A. Spatial cells. Brain Res. 1094, 138–148 (2006).
memory and path integration studied by self-driven passive linear displacement. 39. Clark, B.J. & Taube, J.S. Vestibular and attractor network basis of the head
I. Basic properties. J. Neurophysiol. 77, 3180–3192 (1997). direction cell signal in subcortical circuits. Front. Neural Circuits https://dx.doi.
8. Ivanenko, Y.P. & Grasso, R. Integration of somatosensory and vestibular inputs in org/10.3389/fncir.2012.00007 (2012).
perceiving the direction of passive whole-body motion. Brain Res. 5, 323–327 40. Bassett, J.P. & Taube, J.S. Neural correlates for angular head velocity in the rat
(1997). dorsal tegmental nucleus. J. Neurosci. 21, 5740–5751 (2001).
9. Becker, W., Nasios, G., Raab, S. & Jürgens, R. Fusion of vestibular and 41. Dumont, J.R., Winter, S.S., Farnes, K.B. & Taube, J.S. Neural correlates within
podokinesthetic information during self-turning towards instructed targets. Exp. nucleus prepositus and paragigantocellularis during active and passive movement.
Brain Res. 144, 458–474 (2002). Soc. Neurosci. Poster 631.26 (2015).
10. Frissen, I., Campos, J.L., Souman, J.L. & Ernst, M.O. Integration of vestibular and 42. Sharp, P.E., Tinkelman, A. & Cho, J. Angular velocity and head direction signals
proprioceptive signals for spatial updating. Exp. Brain Res. 212, 163–176 (2011). recorded from the dorsal tegmental nucleus of Gudden in the rat: implications
11. Jürgens, R. & Becker, W. Perception of angular displacement without landmarks: for path integration in the head direction cell circuit. Behav. Neurosci. 115,
evidence for Bayesian fusion of vestibular, optokinetic, podokinesthetic, and 571–588 (2001).
cognitive information. Exp. Brain Res. 174, 528–543 (2006). 43. Knierim, J.J. & Zhang, K. Attractor dynamics of spatially correlated neural activity
12. Telford, L., Howard, I.P. & Ohmi, M. Heading judgments during active and passive in the limbic system. Annu. Rev. Neurosci. 35, 267–285 (2012).
self-motion. Exp. Brain Res. 104, 502–510 (1995). 44. Bassett, J.P., Tullman, M.L. & Taube, J.S. Lesions of the tegmentomammillary
13. Goldberg, J.M. et al. The Vestibular System: a Sixth Sense (Oxford University circuit in the head direction system disrupt the head direction signal in the
Press, 2012). anterior thalamus. J. Neurosci. 27, 7564–7577 (2007).
14. Hitier, M., Besnard, S. & Smith, P.F. Vestibular pathways involved in cognition. 45. Sharp, P.E., Blair, H.T. & Cho, J. The anatomical and computational basis of the
Front. Integr. Nuerosci. 8, 59 (2014). rat head-direction cell signal. Trends Neurosci. 24, 289–294 (2001).
15. Shinder, M.E. & Taube, J.S. Differentiating ascending vestibular pathways to the 46. Peyrache, A., Lacroix, M.M., Petersen, P.C. & Buzsáki, G. Internally organized
cortex involved in spatial cognition. J. Vestib. Res. 20, 3–23 (2010). mechanisms of the head direction sense. Nat. Neurosci. 18, 569–575 (2015).
16. Yoder, R.M. & Taube, J.S. The vestibular contribution to the head direction signal 47. Beraneck, M. & Cullen, K.E. Activity of vestibular nuclei neurons during vestibular
and navigation. Front. Integr. Neurosci. https://dx.doi.org/10.3389/ and optokinetic stimulation in the alert mouse. J. Neurophysiol. 98, 1549–1565
fnint.2014.00032 (2014). (2007).
17. Guldin, W.O. & Grüsser, O.J. Is there a vestibular cortex? Trends Neurosci. 21, 48. Medrea, I. & Cullen, K.E. Multisensory integration in early vestibular processing
254–259 (1998). in mice: the encoding of passive vs. active motion. J. Neurophysiol. 110, 2704–
18. Clark, B.J. & Harvey, R.E. Do the anterior and lateral thalamic nuclei make distinct 2717 (2013).
contributions to spatial representation and memory? Neurobiol. Learn. Mem. 133, 49. Cullen, K.E. The neural encoding of self-motion. Curr. Opin. Neurobiol. 21,
69–78 (2016). 587–595 (2011).
19. Taube, J.S. The head direction signal: origins and sensory-motor integration. Annu. 50. Cullen, K.E. The vestibular system: multimodal integration and encoding of self-
Rev. Neurosci. 30, 181–207 (2007). motion for motor control. Trends Neurosci. 35, 185–196 (2012).
20. Taube, J.S., Muller, R.U. & Ranck, J.B. Jr. Head-direction cells recorded from 51. Grüsser, O.J., Pause, M. & Schreiter, U. Vestibular neurones in the parieto-insular
the postsubiculum in freely moving rats. I. Description and quantitative analysis. cortex of monkeys (Macaca fascicularis): visual and neck receptor responses.
J. Neurosci. 10, 420–435 (1990). J. Physiol. (Lond.) 430, 559–583 (1990).
21. Taube, J.S., Muller, R.U. & Ranck, J.B. Jr. Head-direction cells recorded from 52. Lang, W., Büttner-Ennever, J.A. & Büttner, U. Vestibular projections to the monkey
the postsubiculum in freely moving rats. II. Effects of environmental manipulations. thalamus: an autoradiographic study. Brain Res. 177, 3–17 (1979).
J. Neurosci. 10, 436–447 (1990). 53. Marlinski, V. & McCrea, R.A. Self-motion signals in vestibular nuclei neurons
22. Tryon, V.L. et al. Magnetic field polarity fails to influence the directional projecting to the thalamus in the alert squirrel monkey. J. Neurophysiol. 101,
signal carried by the head direction cell network and the behavior of rats in a 1730–1741 (2009).
task requiring magnetic field orientation. Behav. Neurosci. 126, 835–844 54. Cullen, K.E. & Minor, L.B. Semicircular canal afferents similarly encode active
(2012). and passive head-on-body rotations: implications for the role of vestibular
23. Shinder, M.E. & Taube, J.S. Responses of head direction cells in the anterodorsal efference. J. Neurosci. 22, RC226 (2002).
thalamus during inversion. Soc. Neurosci. Poster 895.1 (2010). 55. Reisine, H. & Raphan, T. Unit activity in the vestibular nuclei of monkeys during
24. Yoder, R.M., Clark, B.J. & Taube, J.S. Origins of landmark encoding in the brain. off-vertical axis rotation. Ann. NY Acad. Sci. 656, 954–956 (1992).
Trends Neurosci. 34, 561–571 (2011). 56. Jamali, M., Sadeghi, S.G. & Cullen, K.E. Response of vestibular nerve afferents
25. Robertson, R.G., Rolls, E.T., Georges-François, P. & Panzeri, S. Head direction innervating utricle and saccule during passive and active translations.
cells in the primate pre-subiculum. Hippocampus 9, 206–219 (1999). J. Neurophysiol. 101, 141–149 (2009).

1472 VOLUME 20 | NUMBER 11 | NOVEMBER 2017  nature neuroscience


perspective

57. Sadeghi, S.G., Minor, L.B. & Cullen, K.E. Response of vestibular-nerve afferents 78. Brooks, J.X., Carriot, J. & Cullen, K.E. Learning to expect the unexpected: rapid
to active and passive rotations under normal conditions and after unilateral updating in primate cerebellum during voluntary self-motion. Nat. Neurosci. 18,
labyrinthectomy. J. Neurophysiol. 97, 1503–1514 (2007). 1310–1317 (2015).
58. Carriot, J., Brooks, J.X. & Cullen, K.E. Multimodal integration of self-motion cues 79. Brooks, J.X. & Cullen, K.E. The primate cerebellum selectively encodes unexpected
in the vestibular system: active versus passive translations. J. Neurosci. 33, self-motion. Curr. Biol. 23, 947–955 (2013).
19555–19566 (2013). 80. Donato, F. & Moser, E.I. A world away from reality. Nature 533, 325 (2016).
59. Carriot, J., Jamali, M., Brooks, J.X. & Cullen, K.E. Integration of canal and otolith 81. Taube, J.S., Valerio, S. & Yoder, R.M. Is navigation in virtual reality with fMRI
inputs by central vestibular neurons is subadditive for both active and really navigation? J. Cogn. Neurosci. 25, 1008–1019 (2013).
passive self-motion: implication for perception. J. Neurosci. 35, 3555–3565 82. Gibson, B., Butler, W.N. & Taube, J.S. The head-direction signal is critical for
(2015). navigation requiring a cognitive map but not for learning a spatial habit. Curr.
60. Roy, J.E. & Cullen, K.E. Selective processing of vestibular reafference during Biol. 23, 1536–1540 (2013).
self-generated head motion. J. Neurosci. 21, 2131–2142 (2001). 83. Winter, S.S., Clark, B.J. & Taube, J.S. Spatial navigation. Disruption of the head
61. Roy, J.E. & Cullen, K.E. Dissociating self-generated from passively applied head direction cell network impairs the parahippocampal grid cell signal. Science 347,
motion: neural mechanisms in the vestibular nuclei. J. Neurosci. 24, 2102–2111 870–874 (2015).
(2004). 84. Jacob, P.-Y. An independent, landmark-dominated head-direction signal in
62. Straka, H., Zwergal, A. & Cullen, K.E. Vestibular animal models: contributions to dysgranular retrosplenial cortex. Nat. Neurosci. 20, 173–175 (2017).
understanding physiology and disease. J. Neurol. 263 (Suppl. 1), 10–23 (2016). 85. Aronov, D. & Tank, D.W. Engagement of neural circuits underlying 2D spatial
63. Brooks, J.X. & Cullen, K.E. Early vestibular processing does not discriminate navigation in a rodent virtual reality system. Neuron 84, 442–456 (2014).
active from passive self-motion if there is a discrepancy between predicted and 86. Aghajan, Z.M. et al. Impaired spatial selectivity and intact phase precession in
actual proprioceptive feedback. J. Neurophysiol. 111, 2465–2478 (2014). two-dimensional virtual reality. Nat. Neurosci. 18, 121–128 (2015).
64. Cullen, K.E. & Brooks, J.X. Neural correlates of sensory prediction errors in 87. Chen, G., King, J.A., Burgess, N. & O’Keefe, J. How vision and movement combine
monkeys: evidence for internal models of voluntary self-motion in the cerebellum. in the hippocampal place code. Proc. Natl. Acad. Sci. USA 110, 378–383
Cerebellum 14, 31–34 (2015). (2013).
65. Cullen, K.E. The neural encoding of self-generated and externally applied 88. Jeffery, K.J., Wilson, J.J., Casali, G. & Hayman, R.M. Neural encoding of large-
movement: implications for the perception of self-motion and spatial memory. scale three-dimensional space—properties and constraints. Front. Psychol. 6, 297
Front. Integr. Neurosci. 7, 108 (2014). (2015).
66. Shinder, M.E. & Taube, J.S. Resolving the active versus passive conundrum for 89. Taube, J.S. Head direction cell firing properties and behavioural performance in
head direction cells. Neuroscience 270, 123–138 (2014). 3-D space. J. Physiol. (Lond.) 589, 835–841 (2011).
67. McCrea, R.A. & Horn, A.K.E. Nucleus prepositus. Prog. Brain Res. 151, 205–230 90. Finkelstein, A. et al. Three-dimensional head-direction coding in the bat brain.
(2006). Nature 517, 159–164 (2015).
68. Butler, W.N., Smith, K.S., van der Meer, M.A.A. & Taube, J.S. The head-direction 91. Calton, J.L. & Taube, J.S. Degradation of head direction cell activity during
signal plays a functional role as a neural compass during navigation. Curr. Biol. inverted locomotion. J. Neurosci. 25, 2420–2428 (2005).
27, 1259–1267 (2017). 92. Wilson, J.J., Page, H. & Jeffery, K.J. A proposed rule for updating of the head
69. Dale, A. & Cullen, K.E. The nucleus prepositus predominantly outputs eye direction cell reference frame following rotations in three dimensions. Preprint at
movement-related information during passive and active self-motion. bioRxiv https://doi.org/10.1101/043711 (2016).
J. Neurophysiol. 109, 1900–1911 (2013). 93. Yoder, R.M. & Taube, J.S. Head direction cell activity in mice: robust
70. Dale, A. & Cullen, K.E. Local population synchrony and the encoding of eye directional signal depends on intact otolith organs. J. Neurosci. 29, 1061–1076
position in the primate neural integrator. J. Neurosci. 35, 4287–4295 (2015). (2009).
71. McFarland, J.L. & Fuchs, A.F. Discharge patterns in nucleus prepositus hypoglossi 94. Laurens, J., Kim, B., Dickman, J.D. & Angelaki, D.E. Gravity orientation tuning
and adjacent medial vestibular nucleus during horizontal eye movement in in macaque anterior thalamus. Nat. Neurosci. 19, 1566–1568 (2016).
behaving macaques. J. Neurophysiol. 68, 319–332 (1992). 95. Mackrous, I., Carriot, J., Jamali, M., Brooks, J. & Cullen, K. Selective encoding
72. Georges-François, P., Rolls, E.T. & Robertson, R.G. Spatial view cells in the of unexpected head tilt by the central neurons takes into account the cerebellar
primate hippocampus: allocentric view not head direction or eye position or place. computation output. Soc. Neurosci. Poster 718.04 http://www.abstractsonline.
Cereb. Cortex 9, 197–212 (1999). com/pp8/#!/4071/presentation/15239 (2016).
73. Rolls, E.T., Robertson, R.G. & Georges-François, P. Spatial view cells in the 96. Aitken, P., Zheng, Y. & Smith, P.F. The regulation of hippocampal theta rhythm
primate hippocampus. Eur. J. Neurosci. 9, 1789–1794 (1997). by the vestibular system. J. Neurophysiol. (in the press).
74. Killian, N.J., Potter, S.M. & Buffalo, E.A. Saccade direction encoding in the 97. Jacob, P.-Y., Poucet, B., Liberge, M., Save, E. & Sargolini, F. Vestibular control
primate entorhinal cortex during visual exploration. Proc. Natl. Acad. Sci. USA of entorhinal cortex activity in spatial navigation. Front. Integrative Neurosci. 8,
112, 15743–15748 (2015). 38 (2014).
75. Kim, B., Dickman, J.D., Taube, J.S. & Angelaki, D.E. Head direction tuned cells 98. Brandon, M.P. et al. Reduction of theta rhythm dissociates grid cell spatial
in the macaque anterior thalamus. Soc. Neurosci. Poster 444.11 (2015). periodicity from directional tuning. Science 332, 595–599 (2011).
76. Gdowski, G.T. & McCrea, R.A. Neck proprioceptive inputs to primate vestibular 99. Koenig, J., Linder, A.N., Leutgeb, J.K. & Leutgeb, S. The spatial periodicity of
nucleus neurons. Exp. Brain Res. 135, 511–526 (2000). grid cells is not sustained during reduced theta oscillations. Science 332,
77. Sadeghi, S.G., Mitchell, D.E. & Cullen, K.E. Different neural strategies for 592–595 (2011).
multimodal integration: comparison of two macaque monkey species. Exp. Brain 100. Winter, S.S., Mehlman, M.L., Clark, B.J. & Taube, J.S. Passive transport disrupts
Res. 195, 45–57 (2009). grid signals in the parahippocampal cortex. Curr. Biol. 25, 2493–2502 (2015).

nature neuroscience  VOLUME 20 | NUMBER 11 | NOVEMBER 2017 1473

You might also like