ArticlePDF AvailableLiterature Review

Variations on the theme: focus on cerebellum and emotional processing

Authors:

Abstract and Figures

The cerebellum operates exploiting a complex modular organization and a unified computational algorithm adapted to different behavioral contexts. Recent observations suggest that the cerebellum is involved not just in motor but also in emotional and cognitive processing. It is therefore critical to identify the specific regional connectivity and microcircuit properties of the emotional cerebellum. Recent studies are highlighting the differential regional localization of genes, molecules, and synaptic mechanisms and microcircuit wiring. However, the impact of these regional differences is not fully understood and will require experimental investigation and computational modeling. This review focuses on the cellular and circuit underpinnings of the cerebellar role in emotion. And since emotion involves an integration of cognitive, somatomotor, and autonomic activity, we elaborate on the tradeoff between segregation and distribution of these three main functions in the cerebellum.
Content may be subject to copyright.
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 1
TYPE Review
PUBLISHED 10 May 2023
DOI 10.3389/fnsys.2023.1185752
OPEN ACCESS
EDITED BY
Conor J. Houghton,
University of Bristol, United Kingdom
REVIEWED BY
Yongsoo Kim,
Penn State Health Milton S. Hershey Medical
Center, United States
Anders Rasmussen,
Lund University, Sweden
*CORRESPONDENCE
Egidio D’Angelo
dangelo@unipv.it
These authors share first authorship
RECEIVED 13 March 2023
ACCEPTED 18 April 2023
PUBLISHED 10 May 2023
CITATION
Ciapponi C, Li Y, Osorio Becerra DA,
Rodarie D, Casellato C, Mapelli L and
D’Angelo E (2023) Variations on the theme:
focus on cerebellum and emotional
processing.
Front. Syst. Neurosci. 17:1185752.
doi: 10.3389/fnsys.2023.1185752
COPYRIGHT
© 2023 Ciapponi, Li, Osorio Becerra, Rodarie,
Casellato, Mapelli and D’Angelo. This is an
open-access article distributed under the terms
of the Creative Commons Attribution License
(CC BY). The use, distribution or reproduction
in other forums is permitted, provided the
original author(s) and the copyright owner(s)
are credited and that the original publication in
this journal is cited, in accordance with
accepted academic practice. No use,
distribution or reproduction is permitted which
does not comply with these terms.
Variations on the theme: focus on
cerebellum and emotional
processing
Camilla Ciapponi1, Yuhe Li1, Dianela A. Osorio Becerra1,
Dimitri Rodarie1,2 , Claudia Casellato1, Lisa Mapelli1and
Egidio D’Angelo1,3*
1Department of Brain and Behavioral Sciences, University of Pavia, Pavia, Italy, 2Centro Ricerche Enrico
Fermi, Rome, Italy, 3Brain Connectivity Center, IRCCS Mondino Foundation, Pavia, Italy
The cerebellum operates exploiting a complex modular organization and a
unified computational algorithm adapted to different behavioral contexts. Recent
observations suggest that the cerebellum is involved not just in motor but
also in emotional and cognitive processing. It is therefore critical to identify
the specific regional connectivity and microcircuit properties of the emotional
cerebellum. Recent studies are highlighting the differential regional localization
of genes, molecules, and synaptic mechanisms and microcircuit wiring. However,
the impact of these regional differences is not fully understood and will require
experimental investigation and computational modeling. This review focuses on
the cellular and circuit underpinnings of the cerebellar role in emotion. And
since emotion involves an integration of cognitive, somatomotor, and autonomic
activity, we elaborate on the tradeoff between segregation and distribution of
these three main functions in the cerebellum.
KEYWORDS
cerebellum, emotion, Purkinje cell, microcircuit functioning, modular organization
Emotion and the cerebellum
The term emotion is used to designate a collection of neurophysiological responses
triggered from parts of the brain to the body and to other parts of the brain, elicited by
stimuli from the external world. A collection of such responses results in an emotional state,
defined by changes within the body. The term feeling is used to describe the complex mental
state that results from the emotional state and reflects the ability to subjectively experience
the mental states resulting from the experience of emotions (Damasio, 1998). Several studies
are striving to understand the neurophysiological underpinnings of emotional processing,
intended as the ensemble of mechanisms that allow emotion to be generated, learned, and
managed.
There is consensus that emotion in mammals involves complex brain networks
centered on structures classically referred to as the limbic system (Papez, 1937;
Maclean, 1990;Nakano, 1998). These include a core composed by the cingulate cortex,
Frontiers in Systems Neuroscience 01 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 2
Ciapponi et al. 10.3389/fnsys.2023.1185752
amygdala, hypothalamus, and hippocampus and have been defined
quite circularly by their involvement in emotion (Damasio, 1998).
It has recently been noted that a discrete system of components
may not be sufficient to describe emotion (Roxo et al., 2011;
Barrett and Satpute, 2019;Steffen et al., 2022) and indeed the
classical networks of the limbic system have been extended to
include the periaqueductal gray (PAG), striatum, prefrontal cortex
and parietal cortex, and cerebellum (Apps and Strata, 2015;
Adamaszek et al., 2017;Rolls, 2019). The cerebellum is typically
recognized for its role in movement coordination and motor
learning, but increasing evidence suggests it may also be involved
in higher-order functions, including emotional and cognitive
processing (Schmahmann, 2004). Over the past three decades,
insights into the role of the cerebellum in emotion have increased
to include perception, recognition, forwarding, encoding, and
learning of emotional information. These processes are thought
to substantially contribute to the generation of experiences and
the regulation of emotional states in relation to motor, cognitive,
and social behaviors, with implications for pain, speech, and mood
disorders.
In humans, neuroimaging indicates cerebellar activation in
fear learning paradigms (Lange et al., 2015;Ernst et al., 2019).
Structural and functional abnormalities also appear to be correlated
with impaired mood regulation and anxiety disorders, generating
what is known as the cerebellar cognitive-affective syndrome
(here, “affective” corresponds to “emotional”) (Hoche et al., 2018).
While investigations in humans are providing a privileged point
of view on emotion, investigations in experimental animals are
revealing most of what we know about the underlying neuronal and
circuit mechanisms. Therefore, we will consider the two aspects in
turn.
This review takes a physiological perspective evaluating the
cellular and circuit underpinnings of the cerebellar role in
emotion. And since emotion involves an integration of cognitive,
somatomotor, and autonomic activity (Itô, 1984;Ito, 1993,
2005, 2008;Schmahmann, 2004;Schmahmann and Caplan, 2006;
D‘Angelo and Casali, 2013;Apps et al., 2018), we will elaborate on
the possible distinction, convergence, and overlap of these three
main functions in the cerebellum.
How could the cerebellum support
emotional processing?
This chapter provides a succinct summary of cerebellar
structure, function, and dynamics, to understand how they
contribute to processing different aspects of behavior, and in
particular emotion.
Abbreviations: AA, ascending axon; BC, basket cell; BLA, basolateral
amygdala; CeA, central amygdala; CF, climbing fiber; CS, conditioned
stimulus; DCN, deep cerebellar nuclei; FC, fear conditioning; fMRI,
functional magnetic resonance imaging; GoC, Golgi cell; GrC, granule cell;
IO, inferior olive; LA, lateral nucleus of the amygdala; MF, Mossy fiber;
MLI, molecular layer interneuron; MRI, magnetic resonance imaging; PAG,
periaqueductal Gray; PC, Purkinje cell; PF, parallel fiber; SC, Stellate cell;
UBC, unipolar brush cell; US, unconditioned stimulus; vlPAG, ventrolateral
periaqueductal Gray; VTA, ventral tegmental area; Z-, zebrin II negative; Z+,
zebrin II positive.
Structure
The cerebellum contains a well-defined set of cells and
fibers and is organized in parasagittal modules and transverse
zones, that are further subdivided based on biochemical mapping
(Figure 1). The intersection of these sagittal and transverse maps
generates smaller units called cortical microzones which, once
connected to the deep cerebellar nuclei (DCN) and inferior
olive (IO) neurons (see abbreviation list), form the olivo-cortico-
nuclear microcomplexes (Apps and Hawkes, 2009;Apps et al.,
2018).
The concept of microcomplex is rooted in the specific
organization of the climbing fibers (CFs) emitted by the IO and
in their projections to Purkinje cells (PCs) and DCN cells (Apps
and Hawkes, 2009;Figure 2). Indeed, CFs first contact DCN
cells and then branch on the sagittal plane contacting several
PCs (about 7 in rodents). These PCs in turn project back to
the same DCN cells receiving a common IO input and then
these DCN cells, through inhibitory interneurons, project to the
same IO neurons that generated the CF, forming a closed loop
circuit. Mossy fiber (MFs) are the other major inputs to these
circuits. Opposite to CFs, MFs spread on the transverse plane
(some branches can cross the midline reaching the contralateral
cerebellum) and do not respect the borders of microcomplexes.
And parallel fibers (PFs), the axon of granule cells (GrCs), bifurcate
in the molecular layer traveling transversally and connecting
regions several millimeters away. In addition, intrinsic biochemical
properties of PCs generate maps, often referred to as stripes
(Apps and Hawkes, 2009), oriented along the longitudinal axis,
increasing the possibility for MFs to intercept regions with
different PC properties. Recently, Zebrin II (Aldolase C) positive
and negative stripes have been identified (Figure 1). The zebrin
stripes largely overlap with longitudinal cerebellar modules to
generate microzones (Apps and Hawkes, 2009) and PCs within
a microzone show synchronous firing when activated by the IO
loops through the CFs (Witter and De Zeeuw, 2015). Functionally,
zebrin stripes show upbound and downbound properties depending
on CF connectivity, PC discharge regulation and PF-PC long-term
depression (see below).
Thus, we have an intricate mesoscopic organization that
overlaps with a broad mapping of anatomical subdivisions to bodily
and mental functions. This brings about two main consequences.
First, specific behavioral functions impinge on complex anatomic-
functional maps made of many microcomplexes and patches
rather than on simply maps residing in uniform cerebellar
regions and differentiated by their input-output connectivity.
Secondly, mapping of bodily and mental functions involves
multiple cerebellar regions, mostly due to the distribution of MF
terminal branches and PFs. These considerations are challenging
the historical view that the cerebellum is made of multiple
parallel non-communicating processing units, leaving space to
what we may call “multimodal fusion”. As a consequence, the
focus moves from the definition of cerebellar functions on
a pure anatomical basis to that of the cerebellar algorithm
that operates with different inputs and outputs (D‘Angelo
and Casali, 2013). This algorithm, in turn, might be tuned
depending on the local properties of the microcomplexes that are
involved.
Frontiers in Systems Neuroscience 02 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 3
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 1
The rodent cerebellum includes the midline vermis and the lateral hemispheres. The left side shows the anatomical subdivisions and the striped
pattern deriving from Aldoc (zebrin II) expression levels in PC (increasing expression levels from light to dark brown). In the vermis, subdivisions in
the anterior-posterior orientation are delimited by circumvolutions and fissures and are further segmented into 10 lobules: the anterior cerebellum
(lobules I V), the central cerebellum (lobules VI VII), the posterior cerebellum (VIII-IXb), nodular cerebellum (lobule IXc-X), flocculus (FI) and
paraflocculus (PFI). Lobules VI-VIII expand into large hemispheric lobules including crus I and crus II. The right side shows the cerebellar
olivo-cortico-nuclear modules defined by IO-CF-PC-CN closed-loop connectivity, identifying seven parallel longitudinal zones (A, B, C1, C2, C3,
D1, D2). C, caudal; Cop, copula pyramidis; Cr I, crus I of ansiform lobule; Cr II, crus II of ansiform lobule; PFl, paraflocculus; Fl, flocculus; L, lateral; Lt,
left; Par, paramedian lobule; R, rostral; Rt, right; Sim, simple lobule. Modified from Zhou et al. (2014) and Voogd (2014).
Function
Independent from the nature of inputs, being them motor,
sensory, cognitive, or emotional, the cerebellar algorithm learns to
predict the precise timing of correlated events (D‘Angelo and Casali,
2013).This algorithm establishes a causal relationship between
afferent signals in sequence and can apply to different contexts,
e.g., correlating elementary components of movement in motor
coordination, sensory stimuli in eye-blink classical conditioning
or in fear conditioning, or abstract items in logical or emotional
operations. In the realm of voluntary functions, the cerebellum
operates in a similar way by comparing intention (conveyed
through descending cortico-cerebellar pathways) with execution
(conveyed by afferent sensory fibers) to predict possible errors.
This de facto anticipates and corrects execution errors, generating
the precise predictive motor control typical of vertebrates.
These observations, combined with insightful theories, have been
extended to explain fluid reasoning and mental coordination and
are arguably important also for emotional control.
Dynamics and mechanisms
On the mechanistic side, the main issue is how the cerebellar
units process input signals through their internal algorithm. In
other words, how do cerebellar neurons and synapses implement
such algorithm. This issue has been addressed by several
physiological and computational works, revealing a complex set
of neuronal and synaptic properties that can generate the input-
output transformation and plastic changes required to process
the input signals (for review see De Schepper et al., 2022). The
circuit operates a complex spatio-temporal transformation of the
MF input and uses the CF input to drive plasticity. On top of this,
the circuit governs intrinsic oscillatory dynamics that allow the time
correlation of the processes involved (D’Angelo, 2011;D’Angelo
et al., 2011).
A unified framework for motor,
cognitive, and emotional processing?
The intricate neuronal circuitry of the cerebellum is thought to
encode internal models that reproduce the dynamic properties of
body parts allowing the brain to precisely control the movement
without the need for sensory feedback. It is thought that the
cerebellum might also encode internal models that reproduce the
essential properties of mental representations elaborated in the
cerebral cortex. This hypothesis suggests a possible mechanism by
which intuition and implicit thought might function and explains
Frontiers in Systems Neuroscience 03 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 4
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 2
Three layers compose the cerebellar cortex: the molecular layer, Purkinje cell layer, and the granular layer. The principal inputs to the cerebellar
cortex are conveyed by mossy fibers (MF) originating from different brain stem and spinal cord nuclei and climbing fibers (CF) originating from the
inferior olive. MFs contact GrCs and GoCs in the granular layer. GrC axon ascends to the molecular layer and bifurcates originating the parallel fibers
(PF). GrC axons contact Purkinje cells (PCs) and molecular layer interneurons (MLI). PCs provide the sole output of the cerebellar cortex and feature
a wide, flat, highly branching dendritic tree and a single long axon projecting to the deep cerebellar nuclei (DCN), from where cerebellar efferent
projections are sent to other brain areas. The MLI are Stellate and basket cells (SC, BC) located in the molecular layer. The candelabrum cells (CC)
(Lainé and Axelrad, 1994;Osorno and Regehr, 2022) are present in the PC layer, while unipolar brush cells (UBC) and lugaro cells (LC) are reported in
the granular layer.
aspects of cognitive and emotional reasoning as well as symptoms
exhibited by psychiatric patients (Ito, 2008).
Recently it was argued that a similar circuit structure in
all cerebellar areas may carry out various operations using a
common computational scheme (D‘Angelo and Casali, 2013).
The authors attributed different roles of the cerebellum to
the specific connectivity of the cerebellar modules with motor,
cognitive, and emotional functions at least partially segregated
into different cerebro-cerebellar loops. This led to formulate
ameta-levels hypothesis: from cellular/molecular to network
mechanisms leading to generation of computational primitives,
thence to high-level cognitive/emotional processing, and finally
to the sphere of mental function and dysfunction. It has been
proposed that the intimate interplay between timing and learning
(reminiscent of the “timing and learning machine” capabilities long
attributed to the cerebellum) reverberates from cellular to circuit
mechanisms. Subsequently, integration within large-scale brain
loops could generate the disparate cognitive/emotional and mental
functions in which the cerebellum is implicated. The cerebellum
should therefore operate as a general-purpose co-processor, whose
effects depend on the specific brain centers to which individual
modules are connected. Abnormal functioning in these loops
could eventually contribute to the pathogenesis of major brain
pathologies including not just ataxia but also dyslexia, autism,
schizophrenia, and depression. These functions and dysfunctions
are deeply involving cognitive and emotional processing.
A system view
In engineering, the cerebellum has been assimilated to a double
forward and inverse controller, capable of predicting system states
based on contextual information and previous memory (Ito, 1993,
2008).The fact that our understanding of cellular phenomena
is sufficient to explain the predictive capabilities of the system
is demonstrated by the ability of neuro-robots, embedding a
canonical cellular representation of the cerebellar circuit dynamics
and mechanisms, to reproduce a wide set of sensory-motor control
tasks (Casellato et al., 2015;Antonietti et al., 2016, 2018, 2022). This
same system approach may be applied to cognitive and emotional
processing, provided the controller, to which the cerebellar circuit is
connected, is appropriately designed and implemented. Eventually,
a precise representation of the neuroanatomical principles based on
cerebellar connectivity and subdivisions, may help understanding
Frontiers in Systems Neuroscience 04 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 5
Ciapponi et al. 10.3389/fnsys.2023.1185752
the computational principles and mechanisms underlying the
cerebellar involvement in emotion (D’Angelo et al., 2016;D’Angelo
and Jirsa, 2022).
Cellular and circuit variants in the
rodent cerebellum
Since the cerebellar computational algorithm is likely to be
tuned depending on the local properties of microcomplexes,
it is important to consider how circuit properties vary across
cerebellar regions. This section is based on rodent studies, in which
accurate investigations have addressed the molecular and cellular
mechanisms and the potential sources of heterogeneity underlying
functional specialization.
Anatomical and physiological studies have shown that the
midline cerebellar vermis is particularly involved in emotional
processing and that lesions and other interventions on this
cerebellar compartment affect emotion-related behaviors (reviewed
in Apps et al., 2018). The cerebellar vermis receives olivo-
cerebellar (CF) input from the caudal medial accessory olive and
emits a cortico-nuclear output to the medial (fastigial) cerebellar
nucleus which, in turn, has widespread projections to the cingulate
cortex, prefrontal cortex, parietal cortex, amygdala, hypothalamus,
hippocampus, periaqueductal gray (PAG), and striatum (Voogd
and Ruigrok, 2004;Blatt et al., 2013;Rolls, 2019) (see also next
chapter).
Different cerebellar vermal lobules are associated with different
aspects of emotion (see Figure 1). From rostral to caudal: lobules
IV–VI with fear learning/memory and affective state; lobule VI–
VII with orientation of gaze; lobule VIII with fear-induced freezing
behavior; and lobule IX and X with cardiorespiratory control (Apps
et al., 2018). Different parts of the vermis could therefore regulate
and integrate the cognitive, motor, and autonomic aspects of fear-
related behavior (Clausi et al., 2017).
Here, neuron and microcircuit variants are considered in
turn and a summary of properties that may be more relevant to
emotional processing is reported in the section on “Conclusion”.
The cerebellar microcircuit and its
variants
The cerebellar circuit includes several cell types connected
according to the general circuit scheme shown in Figures 2,3A.
A recent modeling work summarized the general knowledge on
the cerebellar microcircuit generating a reference ground-truth that
binds together structure, function, and dynamics (De Schepper
et al., 2022). In the granular layer, each glomerulus hosts 50
excitatory and 50 inhibitory synapses on as many GrC dendrites,
plus 2 excitatory synapse on basolateral dendrites of as many
Golgi cells (GoCs). Each one of the 4 GrC dendrites receives
an excitatory and (in most cases) an inhibitory input from as
many different MFs and GoCs, respectively. Each GoC receives
320 ascending axon (AA) synapses on basolateral dendrites
and 910 PF synapses on apical dendrites, and there are 7-
8 electrical synapses plus 160 GABAergic (using the Gamma-
aminobutyric acid neurotransmitter) synapses per GoC-GoC pair.
In the molecular layer, 25% of AAs contact the distal dendrites
of the overlaying PCs (each AA forming 2.4 synapses on average),
while PFs form 1 synapse per PC dendritic intersection. In
summary, each PC receives 12% of the whole GrC inputs from
AAs. Among the molecular layer interneurons (MLI), we consider
the Stellate cells (SCs) and basket cells (BCs) populations. There
are 25 SC-PC and BC-PC synapses altogether. Moreover, there
are 17˙
600 PF-MLI-PC synapses (2˙
600 PF-SC-PC and 15˙
000
PF-BC-PC synapses).
Beyond the canonical description of the cerebellar network
reported above (Figure 3A), there is growing evidence that the
cerebellar cortex is not homogenous in structure, physiology, or
gene expression (Nishiyama and Linden, 2004;Ito, 2006;Sillitoe
et al., 2009;Fujita et al., 2012;Martinez et al., 2013;Nedelescu and
Abdelhack, 2013;Chen et al., 2022) and these variations of regional
features (Figure 3) may influence information processing. Several
cell clusters have recently been reported based on transcriptomic
profile: there are 9 PCs clusters overlapping with zebrin stripes,
3 GrC clusters showing antero-posterior segregation (anterior-
central zone, central-posterior zone, nodular zone), 2 GoC clusters,
and two clusters of MLI including both SCs and BCs (Kozareva
et al., 2021).
Purkinje cells
Two large groups of PCs were identified based on the
expression of Aldoc (aldolase C, the enzyme responsible for the
zebrin II pattern) (Figure 3). The Aldoc positive and negative PCs
define the “zebrin stripes”, consisting of longitudinal alternating
bands of positive (Z +) and negative (Z-) PCs (see Figure 1). This
distribution is typical of the anterior and posterior cerebellar zones
(comprising lobules I - V, and VIII - dorsal IX, respectively), while
the central and nodular zones (lobules VI - VII and ventral IX -
X) are uniformly characterized by Z + PCs. From the functional
point of view, Z- PCs are characterized by higher spontaneous firing
compared to Z + PCs (Xiao et al., 2014;Zhou et al., 2014). More
recently, Z- PCs were shown to undergo simple spike suppression
(downbound) (Wu et al., 2019) and Z + PCs to undergo simple spike
facilitation (upbound) (De Zeeuw, 2021).
In addition to heterogeneity between stripes, PCs also display
functional variations along the transverse axis. PCs in the anterior
lobes have more pronounced adaption of tonic firing compared
to those in the nodular zone (Kim et al., 2013), where PCs show
lower firing rates and reduced excitability (Witter and De Zeeuw,
2015). These differences likely make PCs in the nodular lobe more
suitable to integrate vestibular inputs, which are slower compared
to the faster and stronger inputs arriving in the anterior zone. An
anteroposterior gradient has been also reported for the regularity of
firing (increasing toward the nodular zones) (Witter and De Zeeuw,
2015). The difference in PCs firing along the transverse axis may
simply match the different distribution pattern of zebrin, with more
zebrin negative PCs in the anterior cerebellum, and more positive
PCs in the posterior and nodular cerebellum.
PCs regional diversity also impacts the microcircuit
connectivity. In addition to the general scheme, in which
granule cells activate PCs and DCN cells in sequence, recent
evidence showed that PCs axon collaterals in the granular layer
form functionally active contacts onto GrCs (Guo et al., 2016) and
Unipolar Brushed Cells (UBCs) (Guo et al., 2021) that are less
represented in the anterior cerebellum (Figures 3B, C). Moreover,
Frontiers in Systems Neuroscience 05 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 6
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 3
Cerebellar microcircuit variants. (A) Canonical circuit of the cerebellum. The primary input of mossy fibers (MF) project to deep cerebellar nuclei
(DCN), granule cells (GrC), and Golgi cells (GoC). GoC inhibits GrC in a feedforward and feedback loop. Besides GoC, GrC axons activates Purkinje
cell (PC) and molecular layer interneurons (MLI). PCs provide the sole output of the cerebellar cortex and inhibit the cerebellar nuclei. (B) A first
variant of the canonical circuit is represented by PCs axon collaterals projecting to GrC in the granular layer. (C) A second variant is represented by
unipolar brush cells (UBC), that are contacted by MF and excite GrC. UBC receive inhibition through PC axon collaterals. (D) A third variant shows
the nucleo-cortical projections: DCN axon collaterals contact GrCs with excitatory connections and inhibit GoCs, in the granular layer.
these collaterals show more complex morphology and more
numerous contacts in the granular layer starting from lobule VI
toward lobule X (Guo et al., 2016). This connection is characterized
by a low probability of GABA release, suggesting that a robust
activation of PCs is needed to activate this additional source of
feedback inhibition to GrCs. These results show that, in lobules
VI-X, PCs contribute to both phasic and tonic inhibition in the
granular layer, therefore affecting the spatiotemporal dynamics of
GrCs firing (D‘Angelo et al., 2013;Mapelli et al., 2014).
Unipolar brush cells
The canonical cerebellar microcircuit also differs based on the
regional distribution of UBCs (Figure 3C). UBCs are excitatory
neurons involved in the MF-GrC pathway. These neurons are
located in the granular layer, with highest density in the medial
cerebellum, particularly lobules I, IX, X, and the lowest density
in the lobules IV-VI in mammals (Takács et al., 1999;Mugnaini
et al., 2011). UBCs receive a robust synaptic input by a single
MF and contact several GrCs, and are therefore called “intrinsic
mossy fibers” (Nunzi and Mugnaini, 2000). Their response
patterns amplify the vestibular inputs, significantly modifying the
transmission pathway from MFs to PCs (Nunzi and Mugnaini,
2000;Locatelli et al., 2013;Balmer and Trussell, 2019).
Granule cells
GrCs functions and connectivity are heterogeneously
distributed in different subregions of the cerebellum. First, GrCs
receive excitatory inputs from UBCs, where these neurons are
enriched (mainly in the nodular zone). Second, PCs axon collaterals
contribute to GrC inhibition with a regional gradient increasing
toward posterior and nodular zones. It has also been reported
an antero-posterior gradient for GrCs efficiency in responding
to burst vs. tonic inputs, based on the differential expression of
the CaV3-Kv4 complex (Heath et al., 2014;Witter and De Zeeuw,
2015). This finding probably matches the differences in the inputs
(with slower vestibular inputs arriving at the posterior and nodular
zones). Moreover, GrCs receive excitation from nucleo-cortical
projections in several lobules, with a sagittal arrangement (Houck
and Person, 2014), which generates an internal amplification loop
(Gao et al., 2016).
Golgi cells (and molecular layer interneurons)
Cerebellar inhibitory interneurons also show some specific
spatial arrangement, in particular concerning GoCs. GoC
dendrites are topographically organized within striped boundaries
determined by the zebrin II expression patterns in PCs. There
is no evidence in literature that different GoC subtypes are
present in positive and negative stripes, though two clusters have
been identified based on the transcriptomic profile (Kozareva
et al., 2021). Moreover, a recent study highlighted the functional
heterogeneity of GoCs depending on the expression of the human
glycine neuronal transporter type 2 (GlyT2), affecting the inhibitory
control over GrCs (Dumontier et al., 2023). Additionally, GoCs
receive inhibition from nucleo-cortical projections in several
lobules (Ankri et al., 2015), decreasing GoC inhibition over GrCs.
MLIs are involved in complex inhibitory loops that shape PC
responses (Brown et al., 2019;Prestori et al., 2019). To the best of
our knowledge, a differential spatial distribution of MLIs subtypes
in different zones or lobules has not been reported, yet.
Deep cerebellar nuclei cells
The cellular heterogeneity in the DCN is less studied, compared
to the cerebellar cortex. Nevertheless, DCN can be subdivided
based on the connectivity pattern in caudo-ventral and rostro-
dorsal axes, based on innervation from zebrin II positive and
negative PC, respectively [for a comprehensive review on this
subject see Kebschull et al. (2023)]. Different functional nucleo-
cortical projections have also been identified in several lobules,
arranged in a sagittal organization mainly matching the territories
of PCs sending inputs to those DCN regions (Houck and Person,
2014). Therefore, nucleo-cortical projections (Figure 3D) match
the cortico-nuclear ones, originating closed loops, but the amount
of reciprocal connections can differ in different regions (Houck
and Person, 2014). Excitatory and inhibitory projections from DCN
affect GrCs and GoCs activity, respectively (Ankri et al., 2015;
Gao et al., 2016). Anatomical tracing studies in rats reported that
certain regions of the cerebellar cortex seem to lack nucleo-cortical
projections, such as the lateral vermis (longitudinal zone B, see
Figure 1), the lateral paravermis (C3 zone), and some patches of
the paravermis C1 zone and the hemisphere (D zone) (Buisseret-
Delmas and Angaut, 1988, 1989). As a result, each lobule shows
Frontiers in Systems Neuroscience 06 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 7
Ciapponi et al. 10.3389/fnsys.2023.1185752
different patterns of connectivity. For example, in cats, Lobule V
has been reported to lack nucleo-cortical projections in the C1 zone
(medial paravermis) (Trott et al., 1998a) and be enriched in them
in the C2 zone (lateral paravermis) (Trott et al., 1998b). Inside
zones, the distribution can be uneven, too. For example, the C2
zone of the paraflocculus lacks nucleo-cortical projections (Trott
et al., 1998b). Further anatomical and functional characterization of
these projections is needed to clarify the specific impact on regional
processing.
Structural variants
In addition to differences in local microcircuit connectivity,
there are also differences in the density and distribution of the
different cell types as well as in the thickness of cerebellar regions.
Cell types and density
Geographical differences in PC pertain to neuronal locations,
packing density and diameter, as well as axon and dendritic
arbor morphology. In rats, there are fewer PCs at the base of
each cerebellar folium than at the apex (Braitenberg and Atwood,
1958;Eccles, 1967;Armstrong and Schild, 1970), and the packing
density of PCs in the anterior lobe is greater than that of the
posterior lobe (Armstrong and Schild, 1970). For example, Lobule
X is known to have a greater PCs density than other lobules,
such as lobules II and VI (Keller et al., 2018). There are also
significant regional differences in PC size, with larger PC diameter
and organelle volume in phylogenetically older cerebellar regions
(such as the vermis) (Parma, 1969;Müller and Heinsen, 1984).
Consistent changes in PC axons diameter between white matter
compartments and dendritic arbor morphology at the base of a
folium compared to the apex have been described (Voogd, 2011;
Nedelescu and Abdelhack, 2013;Nedelescu et al., 2018). The
biophysical parameters and energy consumption of individual PC
in various cortical regions could be drastically altered by these
variations. Similarly, GrCs exhibit regional variations in packing
density, higher in the apex than in the base of a folium, and
bigger cell size in the vermis than in the hemispheres (Friede, 1955;
Cerminara et al., 2015). The packing density of GoCs in a variety of
mammals, including humans, is lower in the hemispheres (except
for the flocculus) than in the vermis, where the densities are highest
in the flocculus and lobules IX (Lange, 1974;Palay and Chan-Palay,
1982). Moreover, GoCs in the hemisphere are smaller than those in
the vermis (Geurts et al., 2001, 2003).
Cortical thickness
On average, the granular layer is 73% thicker than the molecular
layer (Zheng et al., 2022), whereas the Purkinje cell layer is
single cell thick. The difference in layer thickness constrains the
area occupied by neurons on the parasagittal surface, e.g., size
and width of PCs dendritic trees. Clearly, synapse development
and maturation as well as plasticity impact the branching and
elongation of dendrites (Takeo et al., 2021). Difference in layer
thickness will also impact the number of neurons, as in the granular
layer.
Granular layer thickness changes from the anterior to posterior
cerebellum, being much larger at lobules VI-VIII than lobules I-V
and smaller in lobules IX -X than in the flocculus (Zheng et al.,
2022). In the depth of the sulci, all three layers are thicker than at
the crowns of the gyri. Globally, the lateral-posterior-inferior area
has a thicker granular layer than the medial-superior region.
Unique PCs dendritic tree shapes in the sulcus result in
more arbor field overlap between nearby PCs than at the apex
(Nedelescu and Abdelhack, 2013) and cause quantitative variations
in dendritic branching patterns between lobules V and IX. Two
physically different PC subtypes have been identified. Type 1
PCs adhered to the conventional concept of a PC morphology,
with a single primary dendrite and a rather uniform branching
density across the molecular layer’s thickness. On the other hand,
Type 2 PCs have two major dendrites and little branching in the
bottom portion. PCs in the anterior lobe (lobule V) of adult mice
exhibited a preponderance of type 1 dendritic arbors, while the
posterior lobe (lobule IX) contained a bigger proportion of type 2
isoforms. Overlapping in adjacent PC arbors varies across lobules,
with more overlap in lobule V than in lobule IX, and arbors
in lobule V using surrounding space more efficiently than those
in lobule IX. Those distinctions in neurons may throw out the
potential question concerning the afferents system differentiation.
For example, lobules IV and V are the key area receiving significant
MF projections from the lateral reticular nucleus, while lobules VI-
VIII and IX get a less dense distribution of this input (Wu et al.,
1999).
Atlas data for microcircuit and structural
variations
The cerebellum holds remarkable variations in cell
morphology, density, volume, and cellular properties. These
variations are further enriched considering microcircuit
connectivity of different cerebellar regions, that go by lobules,
stripes, modules and eventually identify specific microzones and
microcomplexes. Using atlases introduce a new method to combine
evidence collected through the different studies. For example, the
extensive datasets from the Allen Brain Institute (ISH Data.,
2023: Allen Brain Atlas: Mouse Brain.) and Blue Brain Cell Atlas
(Rodarie et al., 2022) provides exhaustive information on cell types
and densities that might be used for detailed model reconstructions
(Figure 4). This information must now be crossed with that on
cerebellar regional connectivity and behavioral localization of
functions addressing those regions and pathways that are critical
for emotional processing.
Cerebellar circuits in rodent emotional
behavior
While the cerebellum has canonically been implicated in motor
control and motor learning, which it exerts through a set of well-
studied circuits and pathways (Figure 5A), neuroimaging and
clinical data recently identified the cerebellum as a key region
among the emotion and cognitive relevant structures (Strick et al.,
2009;Apps and Strata, 2015;Strata, 2015;Adamaszek et al.,
2017). Anatomical studies in rodents have highlighted indirect
connections from the DCN, the sole output of the cerebellum,
Frontiers in Systems Neuroscience 07 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 8
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 4
Neuron density and thickness in the mouse cerebellum based on the Blue Brain Cell Atlas. (A) Estimations of the thickness of the molecular layer in
mm for each subregion of the cerebellar cortex, using the algorithm proposed by Rodarie et al. (2022). (B) Estimations of the density of excitatory
and inhibitory neurons in the molecular and granular layer for each region of the cerebellar cortex. The data are based on the publicity available
Blue Brain Cell Atlas (2018) website.
to various brain structures involved in emotional (Figure 5B)
and cognitive processing (Figure 5C). For example, monosynaptic
projections from the DCN to the ventral tegmental area (VTA),
a structure included in the dopaminergic reward system, have
been described (Carta et al., 2019). Cerebellar output to the VTA
mainly originates from the dentate nucleus and, to a lesser extent,
from the interpositus nucleus (Baek et al., 2022). The optogenetic
activation of the cerebellum-VTA pathway is rewarding and
can condition mice behavior during place preference and social
interaction tests (Carta et al., 2019;Baek et al., 2022). Moreover, the
cerebellum and hypothalamus are highly interconnected through
direct and indirect routes. The reciprocal cerebellar-hypothalamic
pathway has been suggested to provide circuits through which the
cerebellum can influence the autonomic hypothalamic processes
activated during emotional states (Dietrichs and Haines, 1989;
Gritti, 2013). In fact, the cerebellum, via its connection with
the hypothalamus, is involved in controlling vasomotor reflexes
and blood pressure, carotid sinus reflexes, and the somatic
and autonomic manifestations of shame rage, pupil nictitating
membrane, respiration, gastrointestinal functions (Zhu et al., 2011).
Although a strong functional connectivity between the cerebellum
and a key affective center, the amygdala, have been observed
(Supple et al., 1987;Tovote et al., 2015;LeDoux, 2000), the
anatomical substrate for this connectivity is still unclear. In fact, no
direct connections between the cerebellum and the amygdala exist.
Nevertheless, a putative di-synaptic pathway between the DCN
and the basolateral amygdala (BLA) through the thalamus (mainly
medial dorsal nucleus, centro-median, and parafascicular) has been
described (Gornati et al., 2018). Therefore, the cerebellar output
might influence the BLA, known to process affect-relevant salience
and valence information (Jung et al., 2022).
Studying cerebellar role in fear using
animal models
Rodent behavioral studies have been extensively used to
investigate neurobiological aspects of emotions (see Box
“Emotional tasks in rodents”). However, finding valid and
objective measures of animal emotions can be challenging. The
most widely used emotional behavioral tests in rodents aim to
trigger a fear response (Fi. In fact, fear tests elicit a clear behavioral
readout and are highly reproducible. In fear conditioning, rodents
show a robust conditional response during fear memory retrieval
which is manifested in long periods of freezing behavior. Fear
conditioning is an associative learning task. For conditioning to
occur, pathways transmitting the conditioned stimulus (CS) and
unconditioned stimulus (US) have to converge in the brain. It is
widely believed that the amygdala is the site of CS-US convergence
(LeDoux et al., 1988;Koutsikou et al., 2014). Auditory (LeDoux,
1990;LeDoux et al., 1990;Mascagni et al., 1993;Romanski
and LeDoux, 1993) and other sensory inputs representing the
information about the CS terminate mainly in the lateral nucleus
of the amygdala (LA). LA, in turns, project to the central amygdala
(CeA) which controls the expression of fear response through
its projections to the brainstem. Fear conditioning to auditory
CS is thought to be mediated by these pathways. In contextual
fear conditioning, CS is represented by both the auditory and
the context cue. Information on the contextual stimulus involves
communication between the hippocampus and the basal and
accessory basal nuclei of the amygdala (LeDoux, 1992, n.d.).
These nuclei project to the CeA which controls the expression
of the response. Somatosensory information concerning the US,
including the nociceptive stimuli generated by the footshock
(commonly used in fear conditioning tests), are transmitted from
cortical and thalamic areas to the LA (Turner and Zimmer, 1984;
McDonald, 1998), mirroring the CS pathway. Moreover, CeA
receives nociceptive inputs from the parabrachial area (Bernard
and Besson, 1990) and directly from the spinal cord (Burstein
and Potrebic, 1993). CeA ultimately projects to different brain
areas to drive the expression of fear response. Particularly, CeA
modulates the activity of the bed nucleus of the stria terminalis
on the release of the pituitary-adrenal stress hormone (LeDoux
et al., 1988;Herman et al., 2005;Pruessner et al., 2010), and the
lateral hypothalamus on the control of blood pressure. In addition,
CeA activates the ventrolateral periaqueductal gray (PAG) and
elicit freezing response. The PAG has long been implicated
as the organizer of behavioral components of the response to
Frontiers in Systems Neuroscience 08 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 9
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 5
Efferent pathways from the cerebellum. Left: Schematic diagrams of historically studied efferent pathways related to motor functions (A), and of
direct or indirect efferent pathways that are considered to be associated with cognitive (B) and emotional (C) functions. Middle: target brain regions
from each DCN. Right: summary of broad functions of the pathways illustrated on the left. ALM, anterior lateral motor cortex; SMC, sensorimotor
cortex; SC, superior colliculus; Th, thalamus; RN, red nucleus; VN, vestibula nucleus; IO, inferior olive; RF, reticular formation; SpC spinal cord;
mPFC, medial prefrontal cortex; VTA, ventral tegmental area; PPN, peduncolopontine tegmental nucleus; RTN, reticulotegmental nuclei; DRN,
dorsal Raphe nucleus; HY, hypothalamus; PB, parabrachial nucleus; CeA, central amygdala; BLA, basolateral amygdala; PAG, periacqueductal gray.
Modified from Kang et al. (2021).
threat. It exerts direct control of freezing via glutamatergic
projections to premotor neurons targets in the magnocellular
nucleus of the medulla. Freezing behavior can be elicited from
the ventrolateral periaqueductal gray (vlPAG) in decerebrate
animals, therefore descending projections are necessary and
sufficient to elicit such behavioral response (Keay and Bandler,
2001). The cerebellum may also participate in the freezing
response. Indeed, anatomical evidence of projections from the
PAG to the cerebellar cortex have been reported (Dietrichs, 1983),
as well as direct projections from the cerebellar nuclei to the
vlPAG.
Based on early clinical and anatomical evidence, initial works
were conducted on the vermis. Using inactivation procedures,
the interpositus nucleus appeared to be involved in memory
formation of the freezing response to acoustic CS, whereas the
vermis appeared to be involved in the memorization of the freezing
response both to acoustic CS and in contextual tests (Sacchetti
et al., 2002). Furthermore, fear conditioning elicited long-term
potentiation at PF-PC synapses, in particular at cerebellar lobules
V and VI but not in IX and X. Interestingly, fear conditioning did
not modify CF synapses onto PC (Sacchetti et al., 2004).
Moreover, in vivo field potential mapping techniques in
anesthetized rats revealed potent physiological connections
between vlPAG and cerebellar vermal lobule VIII (Koutsikou
et al., 2014). In fact, electrical stimulation of vlPAG on one
side of the brain evoked field potentials on the cerebellar
cortical surface. Responses were largest in vermis lobule VIII,
predominantly localized laterally in the VIII, both sides. This
response is thought to be mediated by CFs activation since
barbituric anesthesia, used in the study, likely weakened MFs
activity. When lobule VIII was inactivated right after the animal
conditioning to CS-US pairings, fear memory retrieval showed a
Frontiers in Systems Neuroscience 09 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 10
Ciapponi et al. 10.3389/fnsys.2023.1185752
reduction in duration of freezing behavior to the CS tone only.
It has been hypothesized that the inactivation of lobule VIII
during the consolidation of the fear memory resulted in the
inability to retrieve the associative memory. However, a reduction
in freezing duration also occurred when CS-only memory
retrieval test was replaced with an innate fear test (predator
odor).
Cerebellum in autonomic response to
fear
Response to fear also includes body reactions as heart
rate and respiratory rate changes, vasomotor activity/blood flow
modification, and consequently sweating, trembling muscles,
increased body temperature (Nisimaru, 2004). The main pathway
known to mediate these effects travels from the amygdala to the
hypothalamus which in turns activates the pituitary gland to secrete
hormones in the bloodstream and the autonomic nervous system.
In order to acknowledge the complex pattern of body reactions
to fear, Signoret-Genest and colleagues recently introduced the
concept of cardio-behavioral defensive state (Signoret-Genest et al.,
2023). From this perspective, freezing is not considered an isolated,
independent response to fear, rather the behavioral component
of a complex response pattern to fear. Such complex response
comprises rapid microstates associated with specific behavior and
heart rate dynamics, both affected by long-lasting macrostates
reflecting context-dependent threat levels. If the cerebellum has a
role in freezing, which is part of such cardio-behavioral response,
then it might have a role in the other bodily responses to
fear. Indeed, the cerebellum has been proven able to influence
autonomic functions. For example, the electrical stimulation of the
vermis lobule IX causes changes in blood pressure and heart rate,
together with inhibition of the baroreceptor reflex (Bradley et al.,
1991).
On this matter, the direct and indirect projections from the
cerebellum to the hypothalamus, spinal visceral nuclei, dorsal
motor vagal nucleus, nucleus tracti solitari, parabrachial nucleus,
raphe nucleus, ambiguous nucleus, are expected to take part (Zhu
et al., 2011).
Furthermore, recent works have highlighted a strong functional
connectivity between the cerebellum and a region of the ventro-
lateral brainstem known as the pre-Botzinger complex (Lu et al.,
2013). The pre-Botzinger complex is considered the central pattern
generator of the respiratory rhythm (Smith et al., 1991) and
possibly the site of integration and coordination of different
orofacial behaviors in mice, such as whisking, licking and
breathing (Lu et al., 2013;Romano et al., 2020). Moreover,
pre-Botzinger complex inhibitory neurons modulate sympathetic
vasomotor neuron activity, generating heart rate and blood
pressure oscillations in phase with respiration (Menuet et al., 2020).
Although there are no direct projections from the cerebellum
to the pre-Botzinger complex (Teune et al., 2000), cerebellar
nuclei project to the gigantocellular reticular formation and to
the parabrachial complex, which are both downstream of the pre-
Botzinger complex. In turns, these areas project to motor neurons
in the spinal cord to drive orofacial movements (Dobbins and
Feldman, 1994).
Anatomical and functional mapping
in humans
Thanks to recent developments in neuroimaging techniques,
direct studies on the brain of living humans, with or without
pathology, have become possible using minimally invasive tools.
These techniques can be used to study the structure of the brain,
e.g., Magnetic Resonance Imaging (MRI) and Computed Axial
Tomography, or its function, e.g., Functional Magnetic Resonance
Imaging (fMRI), Single Photon Emission Computed Tomography,
Positron Emission Tomography, Magnetoencephalography,
Electroencephalography and Long Latency Evoked Potentials
(Aine, 1995).
Functional neuroimaging studies provide a dynamic view of
brain activity. These studies expose the subject to a given stimulus
or context, like a cognitive task, to observe its behavior and record
the underlying brain activity. Among all, fMRI (Ogawa et al.,
1992) has become the most widely used functional neuroimaging
technique. fMRI not only allowed to quantify task-dependent
activations of brain regions by detecting changes in blood flow, but
it also boosted the study of brain networks and connectome. In
particular, resting-state fMRI allows the identification of functional
connectivity networks between specific areas (Fox and Raichle,
2007;Van Dijk et al., 2010). fMRI is also applicable to non-human
species (Gorges et al., 2017) and its use is rapidly extending to
rodent research [see for example (Pan et al., 2015, 2018;Grandjean
et al., 2020;Pradier et al., 2021)]. Nonetheless, fMRI has limitations
and its results are sometimes misinterpreted (Maestú et al., 2003;
Logothetis, 2008;Elliott et al., 2020).
Since emotion involves the integration of cognitive,
somatomotor, and autonomic functions, we will exploit the
privileged point of view offered by human investigations to address
the concept of segregation and distribution of these same functions
across regions of the cerebellum.
Neuroimaging studies in human
cerebellum
A growing literature on functional neuroimaging (Stoodley
and Schmahmann, 2009;Guell and Schmahmann, 2020) evaluates
structure-function relationships in motor and non-motor areas
of the human cerebellum, together with clinical findings. These
studies highlight the involvement of different circuits depending
on the task being performed. The imaging studies in humans
were used to reconstruct a functional topographic map of the
cerebellum. The results of these studies can be interpreted
according to Larsell’s nomenclature (Schmahmann et al., 1999),
that divides the cerebellum along the anterior-posterior axis
in lobules denoted by roman numerals from I to X (optional
labels, H and v, are used to denote the hemispheres and
the vermal region). In addition, the two largest hemispheric
lobules are separated in two parts, namely lobule VII (into
VIIA comprising crus I and crus II, and VIIB), and lobule VIII
(into VIIIA and VIIIB), according to the intrabiventer fissures
(Figure 7A). It should be noted that, in human studies, the gross
subdivisions of the cerebellum are anterior and posterior, this
Frontiers in Systems Neuroscience 10 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 11
Ciapponi et al. 10.3389/fnsys.2023.1185752
BOX 1
Emotional tasks in rodents
Fear conditioning (FC) is a type of associative learning, in which rodents learn to associate a neutral stimulus (conditioned stimulus, CS; more often a tone but
sometimes a visual stimulus) with an aversive stimulus (unconditioned stimulus, US; usually a mild electrical foot shock), and show a consistent behavioral response
(conditioned response) (Figure 6A). The conditioned response is the freezing behavior, defined as the complete absence of movements except those related to breathing
for at least 2 seconds. The FC protocol consists first of an acquisition phase, often performed on day 1, where the animal is presented with a set of CS-US pairings and.
learns to fear both the tone and the training context. FC is learned rapidly, and after one conditioning session a very stable long-lasting behavioral change is produced
(consolidation). The next day, the amount of freezing during the presentation of CS alone (the tone) is used as a behavioral readout of memory retrieval. Depending on
the goal of the study, the retrieval phase can be performed in the same training environment as the conditioning, with or without the auditory cues (context FC, for
hippocampal-dependent learning), or in a different environment, to rule out the contribution of the context in fear memory retrieval (amygdala-dependent learning).
Moreover, consolidation and stability of the fear memory can be assessed by performing the retrieval phase at some distance from the acquisition (delayed FC). Finally,
fear memory can undergo extinction when CS is presented alone for enough times (extinction phase).
The looming test (Figure 6B) is a non-associative fear test that triggers rodents innate behavioral response to approaching aerial predators. The innate fear response
does not require any previously acquired conditioning. A rodent is placed into an arena with a display monitor covering most of the ceiling, and an opaque/dark nest in
a corner of the arena representing a hiding spot. The looming stimulus consists of a black disc appearing on the display monitor above the animal and expanding,
mimicking the shadow of an approaching predator. This setting reliably and consistently triggers one of two defensive behaviors: escape toward the hiding spot or
freezing for a prolonged period. The sensory modality involved in triggering the behavioral response is vision. In particular, a subpopulation of neurons in the superior
colliculus seems to be necessary and sufficient to orchestrate these dimorphic defensive behaviors via two distinct pathways, eventually targeting a different nucleus in
the amygdala (Wilensky et al., 2006;Tovote et al., 2015;Shang et al., 2018). Interestingly, pairing a neutral tone (CS) with a looming stimulus (US) fails to drive the
associative learning of the defensive response, in contrast with the robust learned response of the classical FC (Heinemans and Moita, 2022). However, comparison
between innate and learned fear studies suggest considerable overlap between circuit for innate and learned freezing and escape response, thus leaving the topic open for
further investigation.
In the predator odor test, rodents are exposed to a predator odor, a natural threat, and show an increase in fear behavior, particularly freezing and avoidance. Olfaction is
the most preserved sense throughout mammalian species and is the most prominent sensory modality in rodents. Moreover, early neuroanatomists and scientists
recognized a strong link between olfactory processing and emotions, including the olfactory areas in the emotional circuit (Papez, 1937). Indeed, later studies confirmed
the role of the connections between olfactory structures and specific nuclei in the amygdala in predator odor-induced fear behavior, including freezing. Nevertheless, the
predator odor test has been progressively disregarded, possibly due to the issues in standardizing the odor stimulus, the high variability in the behavioral response, and
the low reproducibility of the test.
Anxiety tests
Elevated plus maze, Open field test, Dark-light box test, Hole board test, Social novelty test, Novelty exploration, Marble burying test.
An array of behavioral tests has been developed to investigate anxiety-like behaviors in rodents, so called because they resemble anxiety behaviors in humans (although
in rodents they might represent something else entirely). In humans, anxiety and fear produce similar behavioral responses, including but not limited to increasing
vigilance, freezing and/or hypoactivity, elevated heart rate, and suppressed food consumption. As opposed to fear, anxiety is elicited by aversive stimuli or a sense of
threat which are diffuse, unpredictable, and of long duration. In rodents, anxiety-like behavior is assessed by measuring the response to a novel and potentially
threatening environment or object. Anxiety tests depend on locomotion and take advantage of the innate aversion rodents show for bright light and open areas. This is
the case for approach-avoidant paradigms such as Elevated plus maze (Figure 6C), Open field test (Figure 6D), Dark-light box test (Figure 6E), and Hole board test
(Figure 6F). These tests assess the natural exploratory behavior of the rodents, which is supposed to be reduced or suppressed when the animals feel a threat. Other
widely used tests rely on the curiosity towards new objects (Figure 6G) or social novelty (Figure 6H), which is altered by stress and anxiety. More creative and less
used tests, such as the Marble burying test (Figure 6I), exploit rodents’ tendency to dig the bedding of their cages and hide objects when stressed.
Often, these tests are inconclusive and non-reproducible because of the high inter-animal and inter-session variability of the results. These inconsistencies are attributed
to the individual genetic variations (as genetic background and strains), testing environment (handling, testing rooms, equipment), protocols, and the rearing
environment of the rodents (housing environment).
latter includes superior and inferior parts and the flocculo-nodular
lobe.
In general, in functional studies, the cerebellum is usually
divided into two representations, motor and non-motor (Stoodley
and Schmahmann, 2009;Buckner et al., 2011;Stoodley et al.,
2012;Guell et al., 2018b, 2019;Guell and Schmahmann, 2020).
Specifically, areas involved in motor representations are the
anterior lobe and parts of lobule VI and VIII, while the areas
involved in non-motor representations are lobules VI, crus I, crus
II, VIIB and IX with extensions to lobule X. Some authors, such
as King et al. (2019), promotes a functional parcellation to re-
define lobar boundaries as it better matches the experimental data.
Also, Boillat et al., 2020 highlighted that the activation patterns
observed in functional studies do not strictly adhere to lobar
boundaries, and suggested that the cerebellar cortex is a continuous
sheet.
Although extensive fMRI studies have been performed, a
composite unified picture of the cerebellar functional topography
is still lacking. Figure 7B represents the activation of the different
cerebellar lobules in motor, cognitive, emotional, and social tasks.
The most significant observation coming from our data metanalysis
is that, besides a preference for specific modalities, there is a
considerable overlap of functions in several cerebellar areas.
Motor cerebellum
As commonly used in literature, we distinguish motor tasks into
simple motor tasks,eye movements tasks, and complex motor tasks.
Studies on simple motor tasks (Riecker et al., 2005;Stoodley
and Schmahmann, 2009;Guell et al., 2018a;Boillat et al., 2020)
report activation of the anterior lobe and lobule VIII, confirming
the electrophysiological studies of body maps (Snider and Eldred,
1951). More specifically:
1. Right/left leg and foot movements activate ipsilateral lobules II,
III, and IV with extensions to lobule I,
2. Orofacial movements activate medial regions of lobule VI,
3. Right/left hand movements activate ipsilateral lobules V, VIII,
4. Tongue movement activates ipsilateral lobule VI,
5. Articulation activates medial hemispheric lobules IV to VI and
VIII.
Studies on eye movements (Dieterich et al., 2000;Dimitrova
et al., 2002;Alvarez et al., 2010;King et al., 2019;Boillat et al., 2020)
report activation of the anterior lobe and lobule VIII, including
vermal areas and lobule VII. For example: optokinetic reflex, smooth
pursuit, vergence and saccades activate vermal lobules VI and VII
Frontiers in Systems Neuroscience 11 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 12
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 6
Schematic representation of emotional tasks in rodents, as detailed in the Box. Fear conditioning (A). Looming test (B). Elevated plus maze (C). Open
field test (D). Dark-light box test (E). Hole board test (F). Novel object test (G). Social novelty test (H). Marble burying test (I).
and extend into vermis lobules IV-V and adjacent medial regions
of crus I and crus II.
Studies on complex motor tasks (Debas et al., 2010;Ma et al.,
2010;Schlerf et al., 2010;Stoodley et al., 2012;Hardwick et al.,
2013;Ashida et al., 2019;King et al., 2019;Tzvi et al., 2020) report
activation of the anterior lobe, lobule VIII as well as lobule VII.
More specifically:
1. Motor learning activates bilateral anterior/medial cerebellum
extending into lobule VI,
2. Motor planning activates lobules VI, crus I and crus II,
3. Motor sequencing activates anterior regions extending into
lobule VI, VIIIA/B and crus II,
4. Motor adaptation activates anterior lobules IV-V, VIII
bilaterally,
5. Visuomotor adaptation activates lobules IV-V, VI, VIII, crus II,
6. Finger tapping activates ipsilateral anterior lobe (lobules IV-V)
and lobules VI, VIII,
7. Left and right button press activates ipsilateral anterior lobe
and lobule VIII,
8. Skilled movement activates lobules VI, crus I,
9. Execution (reported in both motor and cognitive studies)
activates lobes V, VI, VIII,
Clinical studies (Urban et al., 2003;Kumral et al., 2007;
Schmahmann et al., 2009;Stoodley and Schmahmann, 2009;
Maderwald et al., 2012;Bultmann et al., 2014;Stoodley et al., 2016)
support the previously reported experimental findings on motor
tasks. Affections of these lobules or regions have been additionally
found to be related to neuromotor disorders. For example, patients
with poor finger tapping have lesions in the ipsilateral anterior
cerebellum, ataxia scores correlate with damage of the anterior lobe
(lobules II to V extending into lobule VI), and lesions responsible
for cerebellar dysarthria include lobules I, II, III.
Cognitive cerebellum
fMRI analyses in cognitive tasks (Fink et al., 2000;Stoodley and
Schmahmann, 2009;Stoodley et al., 2012;Balsters et al., 2014;E
et al., 2014;Guell et al., 2018a;Ashida et al., 2019;Casiraghi et al.,
2019;King et al., 2019) report vermal activation of lobules VI and
VII as well as hemispheric activation of these lobules with lobule
VIIIA and extensions to lobule V. More specifically:
1. Action/execution tasks (reported in both motor and cognitive
studies) activate lobules V, VI, VIII and regions of crus I
2. Action/observation tasks activate areas of lobule VI, crus I and
crus II,
3. Working memory activates hemispheric lobules VI, crus I, and
VIIB to VIIIA,
4. Visuospatial working memory and attention activate
hemispheric lobules VI, crus I,
5. Linguistic and reading tasks activate right lobules VI, VII, and
VIIIA,
6. Spatial (mental rotation) activate left-lateralized lobules VI,
VII to VIIIA, crus I and midline VIIA,
7. Spatial (navigation/orientation/mutual rotation) activate right
crus I, bilateral lobules VI, and VII.
Clinic evidence (Schmahmann and Sherman, 1998;Mariën
et al., 2014;Stoodley et al., 2016;Hoche et al., 2018) proves
Frontiers in Systems Neuroscience 12 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 13
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 7
Contribution of the human cerebellum to motor, cognitive, emotional, social and complex task. (A) Classical nomenclature of the division of the
human cerebellum, from anterior lobe (I to V) to posterior lobe (Superior posterior: lobules VI to VIIB and Inferior posterior: lobules VIIIA to IX)
separated by intrabiventer fissures (dark black line), and flocculo-nodular lobe. (B) Association of the activation of the different cerebellar lobes
(in colors) to different functional tasks, based on the consulted bibliography. The tasks were divided into motor, cognitive, emotional, and social.
Complex experiences cannot be strictly classified into motor, cognitive, emotional, or social tasks. (C) Lobules involvement in functional tasks of
panel (B). The height of each bar corresponds to the number of tasks (out of the 17 listed) related to the activation of each lobule.
Frontiers in Systems Neuroscience 13 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 14
Ciapponi et al. 10.3389/fnsys.2023.1185752
FIGURE 8
Venn’s diagram based on the consulted bibliography of three functional domains (motor, cognitive, emotional/social). In this diagram, the emotional
and social tasks are joined within the same functional domain, and complex experiences are excluded. The Venn’s diagram shows the lobules
involved in motor, cognitive, and emotional/social tasks. Overlapping parts contain the lobules that are common to two or all regions. Notice that
lobules involved in the cognitive function are in common with both the motor and emotional/social function.
that cognitive impairments correlate with posterior lobe lesions
(lobules VI and VII, including crus I, crus II) in abstract
reasoning, working memory, planning, executive functioning,
spatial cognition, language, and cognitive regulation. Damage to
the left hemisphere is more related to spatial difficulties, while the
right cerebellar hemisphere to language problems.
Emotional cerebellum
Most studies (Kilts et al., 2003;Wildgruber et al., 2005;Stoodley
and Schmahmann, 2009;Stoodley et al., 2010, 2012;Moulton et al.,
2011;Baumann and Mattingley, 2012;Schraa-Tam et al., 2012;
Adamaszek et al., 2014, 2017;Nummenmaa et al., 2014;Lange et al.,
2015;An et al., 2018;Ernst et al., 2019;King et al., 2019;Thomasson
et al., 2019) agree that the regions involved in emotional processing
are in the posterior lobes, including the vermal area.
Vermal and paravermal regions are involved in fundamental or
primary emotions, namely:
1. Anger: vermal lobule IX and paravermal lobules VI, crus I and
crus II,
2. Sadness: vermal lobules, crus I, crus II and paravermal lobule
VI,
3. Disgust; vermal lobules V, VIIIA to IX and paravermal lobule
VI,
4. Fear: vermal lobule VIIB and paravermal lobes VI, crus I and
crus II,
5. Happiness: paravermal lobe VIIIB.
In the domain of Emotional recognition:
1. Seeing emotional vs neutral images activates posterior lobes,
including lobules VI, VII
2. Seeing pleasant images activates the right lobules VI and crus
I
3. Seeing unpleasant images activates bilateral lobules VI to
VIIIB, vermis of lobule IX, and left crus I
4. Processing angry facial expression activates the right lobule V
5. Identifying emotional intonations activates midline lobule VII
and lateral lobules VI and crus I
6. Listening to other’s spoken emotional narratives induces
activation in lobules VI and IX, right lobules VIIB, VIII, and
IX
7. Processing of emotion-laden visual and auditory art activates
crus I and crus II is reported.
Emotion regulation activates vermal and hemispheric lobules
VI, crus I, VIIB, VIIIA/B and IX. Emotional memory activates crus
II and lobules VI, IX, X.
Finally, Fear learning activates the culmen, right and left
hemispheric lobule IV-V and left lobule VI, and right and left lobule
hemispheric lobule IX.
Social cerebellum
The study of the role of cerebellum in social behaviors (Van
Overwalle et al., 2014, 2020;Leggio and Olivito, 2018) has recently
gained attention. More precisely, on mentalizing task which
evaluate the ability to understand the mental state of oneself or
others that underlies overt behavior:
1. social mentalizing activates crus I, crus II, and lobule IX,
2. person mentalizing activates crus I, lobules IV and VI,
3. event mentalizing activates crus I,
4. abstract mentalizing activates crus I, lobules VI and IX.
Mirroring is the behavior in which one person
subconsciously imitates the gesture, speech pattern,
or attitude of another). This task activates crus I,
lobules VIIB and VI.
Clinical data (Levisohn et al., 2000;Schmahmann
et al., 2007;Tavano et al., 2007;Adamaszek et al., 2014)
report the relationship between damage to the posterior
vermis with extensions in the medial cerebellar regions
Frontiers in Systems Neuroscience 14 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 15
Ciapponi et al. 10.3389/fnsys.2023.1185752
with social, behavioral, affective deficits and emotional
disorders in patients with affect dulling, anxiety, depression,
impulsivity, irritability, aggression and autism spectrum
disorder.
Complex experiences involving multiple
cerebellar areas
fMRI studies indicate that the cerebellum responds to other
complex experiences such as aversive and painful stimuli (Becerra
et al., 1999;Ploghaus et al., 1999;Dimitrova et al., 2003;Singer
et al., 2004;Jackson et al., 2005;Moulton et al., 2011;Borsook et al.,
2018), interoceptive stimuli (Parsons et al., 2000) as well as sexual
experiences (Holstege and Huynh, 2011;Wise et al., 2017).
Aversive stimuli (noxious heat, electricity, unpleasant images)
activate lobule VI, crus I, VIIB, while painful stimulation evokes
bilateral activation of lobules III, IV and VIII as well as the vermal
and hemispheric parts of lobule VI and crus I. Therefore, pain
experiences activate multiple areas: the anterior regions in relation
to the motor component as reflex, the posterior regions in relation
to the anticipation of pain or unpleasant situation, and the posterior
vermis for the painful experience itself including fear or startle
reactions. Pain empathy activates bilaterally lobule VI and crus I.
Interoceptive stimuli such as thirst activates vermal lobe III,
whereas during maximal thirst lobes X, VI and crus I are activated,
and the left lobe VI is strongly activated after drinking to satiety.
Man ejaculation evokes activation in lobule VIII, right vermal
lobule V, left vermal lobule VII, left lobule VI, and bilateral crus
I, moreover women orgasm elicited activity left lobule V and right
lobule III, while during the post-orgasm recovery activates vermal
and hemispheric lobule VI and lobule VII.
It would be wrong to limit an area of the cerebellum strictly to
either motor, cognitive, emotional or social tasks. However, lobules
task-dependent predominance can be delineated (Figures 7,8).
Lobule VI as an integrating region
Figure 7B shows how the lobule VI is involved in all the
reported tasks of our collected literature studies. This lobule
might represent an integrating or modulating node between motor,
cognitive, emotional, and social areas. This may be due to the
connection of this area with multiple cortical and brain stem areas.
It would be interesting to see whether a multimodal input implies
cells with larger dendrites to integrate more synaptic information.
Extra cerebellar connectivity
Studies on resting-state functional connectivity showed that
most of the cerebellar regions partake in 5 intrinsically-connected
circuits or networks (Habas et al., 2009;Brissenden et al., 2018):
1. the sensorimotor network (lobule V-VI and VIII),
2. the default-mode network (lobules VIIa and IX),
3. the right and left central executive network (lobule VIIa),
4. the dorsal attentional network (lobules VIIB and VIIIA)
5. the salience network (lobules VI and VII).
Furthermore, at least 2 functional macroscale circuits involving
the cerebellum have been identified (Leggio and Olivito, 2018):
asensorimotor zone (anterior cerebellum) functionally correlated
with premotor, motor, somatosensory, visual, and auditory cortical
regions, and a supramodal zone (posterior cerebellum) functionally
correlated with dorsolateral prefrontal and inferior posterior-
parietal regions.
Resting-state fMRI studies report connections of the
cerebellum with cerebral cortex (Krienen and Buckner, 2009;
O’Reilly et al., 2010), striatum (Stoodley and Schmahmann,
2009), basal ganglia (Bostan et al., 2010, 2013), antero-medial
and mediodorsal thalamus, and habenula (Milardi et al., 2016;
Habas and Manto, 2018). Moreover, tracts have been identified
between the medial part of posterior lobule of the cerebellum with
insula (Cauda et al., 2011;Baur et al., 2013), frontal operculum,
anterior cingulate cortex, medial prefrontal cortex, amygdala, and
hippocampus, as well as between the lateral part of the cerebellum
with the hypothalamus and anterior cingulate cortex (Sang et al.,
2012;Bostan et al., 2013;Murty et al., 2014;Habas, 2018).
Limitations of in vivo neuroimaging
studies of the cerebellum
Neuroimaging studies are affected by limited spatial and
temporal resolution, although this has recently been increased by
technological improvements such as fMRI, electroencephalography
(Huster et al., 2012;Jorge et al., 2014;Croce et al., 2016) and
advanced analysis methods (Spencer and Goodfellow, 2022). The
fine folded pattern of the layers of the cerebellar cortex makes it
even more complicated to define cerebellar connectivity in detail
in humans. Indeed, the surface of the cerebellar cortex is much
more tightly folded than that of the cerebral cortex. Thus, although
the visible surface of the cerebellum is several times smaller than
that of the neocortex, once unfolded it equals 78% of the total
surface area of the neocortex (Sereno et al., 2020). This, together
with the extended connectivity to associative cortical areas (Palesi
et al., 2015), suggests a prominent role for the cerebellum in the
evolution of distinctive human behaviors including cognition and
emotion.
Conclusions and outstanding
questions
Segregation, distribution, and
specialization
The evidence reviewed here reveals, as a fundamental aspect of
emotional processing in the cerebellum, the interplay of segregation
and distribution of functions in multiple cerebellar regions. The
corollary is the concept of specialization, that could adapt the
general circuit mechanisms of the cerebellum to the specific needs
of emotional control.
Segregation. Numerous studies (Stoodley et al., 2010, 2021;
Solstrand Dahlberg et al., 2020) have shown that lobules of the
Frontiers in Systems Neuroscience 15 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 16
Ciapponi et al. 10.3389/fnsys.2023.1185752
central cerebellum, in particular lobules VI, VII, and VIII, show
most of the fMRI-related activity during emotional tasks. Likewise,
the data summarized in Figure 7B indicate that several cerebellar
regions are engaged in multiple functions, as opposed to being
specific for a single function. It is plausible to ask whether and to
what extent the functional specialization of lobules VI, VII, and
VIII entails molecular, cellular, or local network specializations as
compared to other modules. The search for circuit specializations
in the cerebellum may start from central lobules, especially lobule
VI that presents the most extensive implications for emotional
processing (Figure 7C).
Distribution. Since the emotional response entails
sensorimotor, cognitive, and autonomic aspects, numerous
cerebellar lobules that generically operate motor or autonomic
control are indeed also contributing to emotion. As a corollary,
some lobules should be convergence sites for emotional, cognitive
and motor responses, as we have pointed out for lobule VI.
Specialization. The processing role of the cerebellum at system
level notwithstanding, the cerebellar network repeats itself in
the general motive but not in the details, suggesting that some
properties may differentiate microcircuit computation. Focusing
on central lobules, the following pattern emerges:
1. Lobules VI and VII are almost entirely Z + (or upbound),
PCs have lower firing rate and show PF-PC long-term
potentiation. Lobule VIII has alternating Z + /Z- stripes.
2. Lobules VI to X show an increasing amount of PC axonal
collaterals contacting GrCs, UBCs, and GoCs.
3. Lobules VI to VIII have almost no UBCs.
4. Lobule VI to VIII have higher cortical thickness than other
lobules, and lobule VI has a higher PC density than most of
the other lobules.
5. There is an antero-posterior gradient in GrC and PC firing
adaptation.
It is possible that these different properties subtend the higher
need for multimodal integration of lobules VI-VIII bringing about
larger input-output fiber sets, different management of recurrent
loops and local processing of signals on specific frequency bands,
differential processing of burst and tonic firing depending on
the nature of the input, different needs for plasticity generation
and memory storage. To what extent these properties impact on
microcircuit function and spatiotemporal dynamics is currently
unknown and its understanding will require accurate physiological
recordings and model simulations.
Implications for behavioral processing
Since the cerebellum is involved in learning the association of
time-correlated events and it is likely to do so in all its modules
(D‘Angelo and Casali, 2013), a core hypothesis is that the cerebellum
orchestrates the time domain for the suppression/initiation of fear
responses as much as it does in other conditioning tasks, e.g., in
eye-blink classical conditioning. In other words, the cerebellum
should perform in emotional control the same operation it
performs in motor control and coordinate the many components
of the emotional response including motor and visceral reactions.
Lobules VI to VIII are involved and studies in rodents support
their functional connection to the amygdala, hypothalamus, and
periaqueductal gray. Regions IX and X play a major role in
orchestrating the autonomic response. In humans, the pattern of
explorable emotions is larger than in rodents, but evidence still
supports a specialization of lobules VI to VIII. In humans, the
high resolution of fMRI techniques during tasks has reinforced
the concept of a remarkable distribution and overlapping of
areas involved in emotional control with those controlling motor,
cognitive, and social behaviors. In addition to coordinate the
components of the emotional response, the cerebellum may
likewise be involved in comparing the cortical commands with their
execution to predict possible errors and correct them. Again, the
cerebellum may play the same coordinating role in emotion as it
plays in sensorimotor control.
Author contributions
CCia, YL, and DO: writing original draft preparation
and visualization. ED’A, LM, CCas, and DR: writing
review and editing. ED’A, LM, and CCas: supervision. ED’A:
conceptualization. All authors contributed to the article and
approved the submitted version.
Funding
This work has received funding from the European Union’s
Horizon 2020 Research and Innovation Program under the Marie
Skłodowska-Curie grant agreement No. 956414. “Cerebellum and
emotional networks” and under the Flagship program No. 945539
“Human Brain Project, SGA3”. This work was supported by
#nextgenerationEU (NGEU) and funded by the Ministry of
University and Research (MUR), National Recovery and Resilience
Plan (NRRP), project MNESYS (PE0000006) A Multiscale
integrated approach to the study of the nervous system in health
and disease (DN. 1553 11.10.2022). DR has received funding from
the Centro Ricerche Enrico Fermi.
Conflict of interest
The authors declare that the research was conducted in the
absence of any commercial or financial relationships that could be
construed as a potential conflict of interest.
Publisher’s note
All claims expressed in this article are solely those of the
authors and do not necessarily represent those of their affiliated
organizations, or those of the publisher, the editors and the
reviewers. Any product that may be evaluated in this article, or
claim that may be made by its manufacturer, is not guaranteed or
endorsed by the publisher.
Frontiers in Systems Neuroscience 16 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 17
Ciapponi et al. 10.3389/fnsys.2023.1185752
References
Adamaszek, M., D’Agata, F., Ferrucci, R., Habas, C., Keulen, S., Kirkby, K. C.,
et al. (2017). Consensus paper: cerebellum and emotion. Cerebellum 16, 552–576.
doi: 10.1007/s12311-016- 0815-8
Adamaszek, M., D’Agata, F., Kirkby, K. C., Trenner, M. U., Sehm, B., Steele, C. J.,
et al. (2014). Impairment of emotional facial expression and prosody discrimination
due to ischemic cerebellar lesions. Cerebellum 13, 338–345. doi: 10.1007/s12311-013-
0537-0
Aine, C. J. (1995). A conceptual overview and critique of functional neuroimaging
techniques in humans: I. MRI/FMRI and PET. Crit. Rev. Neurobiol. 9, 229–309.
Alvarez, T. L., Alkan, Y., Gohel, S., Douglas Ward, B., and Biswal, B. B. (2010).
Functional anatomy of predictive vergence and saccade eye movements in humans: a
functional MRI investigation. Vis.Res. 50, 2163–2175. doi: 10.1016/j.visres.2010.08.018
An, S., Han, X., Wu, B., Shi, Z., Marks, M., Wang, S., et al. (2018). Neural activation
in response to the two sides of emotion. Neurosci. Lett. 684, 140–144. doi: 10.1016/j.
neulet.2018.07.011
Ankri, L., Husson, Z., Pietrajtis, K., Proville, R., Léna, C., Yarom, Y., et al. (2015).
A novel inhibitory nucleo-cortical circuit controls cerebellar Golgi cell activity. Elife
4:e06262. doi: 10.7554/eLife.06262
Antonietti, A., Casellato, C., D’Angelo, E., and Pedrocchi, A. (2016). “Model-driven
analysis of eyeblink classical conditioning reveals the underlying structure of cerebellar
plasticity and neuronal activity, in Proceedings of the IEEE Transactions on Neural
Networks and Learning Systems, (Piscataway, NJ), doi: 10.1109/TNNLS.2016.2598190
Antonietti, A., Geminiani, A., Negri, E., D’Angelo, E., Casellato, C., and Pedrocchi,
A. (2022). Brain-inspired spiking neural network controller for a neurorobotic whisker
system. Front. Neurorobot. 16:817948. doi: 10.3389/fnbot.2022.817948
Antonietti, A., Monaco, J., D’Angelo, E., Pedrocchi, A., and Casellato, C.
(2018). Dynamic redistribution of plasticity in a cerebellar spiking neural network
reproducing an associative learning task perturbed by TMS. Int. J. Neural Syst.
28:1850020. doi: 10.1142/S012906571850020X
Apps, R., and Hawkes, R. (2009). Cerebellar cortical organization: a one-map
hypothesis. Nat. Rev. Neurosci. 10, 670–681. doi: 10.1038/nrn2698
Apps, R., Hawkes, R., Aoki, S., Bengtsson, F., Brown, A. M., Chen, G., et al.
(2018). Cerebellar modules and their role as operational cerebellar processing units.
Cerebellum 17, 654–682. doi: 10.1007/s12311-018- 0952-3
Apps, R., and Strata, P. (2015). Neuronal circuits for fear and anxiety the missing
link. Nat. Rev. Neurosci. 16, 642–642. doi: 10.1038/nrn4028
Armstrong, D. M., and Schild, R. F. (1970). A quantitative study of the purkinje cells
in the cerebellum of the albino rat. J. Comp. Neurol. 139, 449–456. doi: 10.1002/cne.
901390405
Ashida, R., Cerminara, N. L., Edwards, R. J., Apps, R., and Brooks, J. C. W. (2019).
Sensorimotor, language, and working memory representation within the human
cerebellum. Hum. Brain Mapp. 40, 4732–4747. doi: 10.1002/hbm.24733
Baek, S. J., Park, J. S., Kim, J., Yamamoto, Y., and Tanaka-Yamamoto, K.
(2022). VTA-projecting cerebellar neurons mediate stress-dependent depression-like
behaviors. eLife 11:e72981. doi: 10.7554/eLife.72981
Balmer, T. S., and Trussell, L. O. (2019). Selective targeting of unipolar brush cell
subtypes by cerebellar mossy fibers. eLife 8:e44964. doi: 10.7554/eLife.44964
Balsters, J. H., Laird, A. R., Fox, P. T., and Eickhoff, S. B. (2014). Bridging the gap
between functional and anatomical features of cortico-cerebellar circuits using meta-
analytic connectivity modeling. Hum. Brain Mapp. 35, 3152–3169. doi: 10.1002/hbm.
22392
Barrett, L. F., and Satpute, A. B. (2019). Historical pitfalls and new directions in the
neuroscience of emotion. Neurosci. Lett. 693, 9–18. doi: 10.1016/j.neulet.2017.07.045
Baumann, O., and Mattingley, J. B. (2012). Functional topography of primary
emotion processing in the human cerebellum. NeuroImage 61, 805–811. doi: 10.1016/
j.neuroimage.2012.03.044
Baur, V., Hänggi, J., Langer, N., and Jäncke, L. (2013). Resting-state functional and
structural connectivity within an insula-amygdala route specifically index state and
trait anxiety. Biol. Psychiatry 73, 85–92. doi: 10.1016/j.biopsych.2012.06.003
Becerra, L. R., Breiter, H. C., Stojanovic, M., Fishman, S., Edwards, A., Comite,
A. R., et al. (1999). Human brain activation under controlled thermal stimulation
and habituation to noxious heat: an fMRI study. Magn. Resonan. Med. 41,
1044–1057.
Bernard, J. F., and Besson, J. M. (1990). The spino(trigemino)pontoamygdaloid
pathway: electrophysiological evidence for an involvement in pain processes.
J. Neurophysiol. 63, 473–490. doi: 10.1152/jn.1990.63.3.473
Blatt, G. J., Oblak, A. L., and Schmahmann, J. D. (2013). “Cerebellar connections
with limbic circuits: anatomy and functional implications, in Handbook of the
Cerebellum and Cerebellar Disorders, eds M. Manto, J. D. Schmahmann, F. Rossi, D. L.
Gruol, and N. Koibuchi (Dordrecht: Springer Netherlands), doi: 10.1007/978-94- 007-
1333-8_22
Blue Brain Cell Atlas (2018). Blue Brain Portal. Available online at: https://portal.
bluebrain.epfl.ch/resources/models/cell-atlas/ (accessed March 9, 2023).
Boillat, Y., Bazin, P. L., and van der Zwaag, W. (2020). Whole-body somatotopic
maps in the cerebellum revealed with 7T fMRI. NeuroImage 211:116624. doi: 10.1016/
j.neuroimage.2020.116624
Borsook, D., Youssef, A. M., Simons, L., Elman, I., and Eccleston, C. (2018). When
pain gets stuck: the evolution of pain chronification and treatment resistance. Pain 159,
2421–2436. doi: 10.1097/j.pain.0000000000001401
Bostan, A., Dum, R., and Strick, P. (2010). The basal ganglia communicate with
the cerebellum. Proc. Natl. Acad. Sci. U. S. A. 107, 8452–8456. doi: 10.1073/pnas.
1000496107
Bostan, A. C., Dum, R. P., and Strick, P. L. (2013). Cerebellar networks with the
cerebral cortex and basal ganglia. Trends Cognit. Sci. 17, 241–254. doi: 10.1016/j.tics.
2013.03.003
Bradley, D. J., Ghelarducci, B., and Spyer, K. M. (1991). The role of the posterior
cerebellar vermis in cardiovascular control. Neurosci. Res. 12, 45–56. doi: 10.1016/
0168-0102(91)90099- K
Braitenberg, V., and Atwood, R. P. (1958). Morphological observations on the
cerebellar cortex. J. Comp. Neurol. 109, 1–33. doi: 10.1002/cne.901090102
Brissenden, J. A., Tobyne, S. M., Osher, D. E., Levin, E. J., Halko, M. A., and Somers,
D. C. (2018). Topographic cortico-cerebellar networks revealed by visual attention and
working memory. Curr. Biol. 28, 3364–3372. doi: 10.1016/j.cub.2018.08.059
Brown, A. M., Arancillo, M., Lin, T., Catt, D. R., Zhou, J., Lackey, E. P., et al. (2019).
Molecular layer interneurons shape the spike activity of cerebellar Purkinje cells. Sci.
Rep. 9:1742. doi: 10.1038/s41598-018- 38264-1
Buckner, R. L., Krienen, F. M., Castellanos, A., Diaz, J. C., and Thomas Yeo, B. T.
(2011). The organization of the human cerebellum estimated by intrinsic functional
connectivity. J. Neurophysiol. 106, 2322–2345. doi: 10.1152/jn.00339.2011
Buisseret-Delmas, C., and Angaut, P. (1988). The cerebellar nucleocortical
projections in the rat. A retrograde labelling study using horseradish peroxidase
combined to a lectin. Neurosci. Lett. 84, 255–260. doi: 10.1016/0304-3940(88)90516- 2
Buisseret-Delmas, C., and Angaut, P. (1989). Anatomical mapping of the cerebellar
nucleocortical projections in the rat: a retrograde labeling study. J. Comp. Neurol. 288,
297–310. doi: 10.1002/cne.902880208
Bultmann, U., Pierscianek, D., Gizewski, E. R., Schoch, B., Fritsche, N., Timmann,
D., et al. (2014). Functional recovery and rehabilitation of postural impairment and
gait ataxia in patients with acute cerebellar stroke. Gait Post. 39, 563–569. doi: 10.1016/
j.gaitpost.2013.09.011
Burstein, R., and Potrebic, S. (1993). Retrograde labeling of neurons in the spinal
cord that project directly to the amygdala or the orbital cortex in the rat. J. Comp.
Neurol. 335, 469–485. doi: 10.1002/cne.903350402
Carta, I., Chen, C. H., Schott, A. L., Dorizan, S., and Khodakhah, K.
(2019). Cerebellar modulation of the reward circuitry and social behavior. Science
363:eaav0581. doi: 10.1126/science.aav0581
Casellato, C., Antonietti, A., Garrido, J. A., Ferrigno, G., D’Angelo, E., and
Pedrocchi, A. (2015). Distributed cerebellar plasticity implements generalized
multiple-scale memory components in real-robot sensorimotor tasks. Front. Comput.
Neurosci. 9:24. doi: 10.3389/fncom.2015.00024
Casiraghi, L., Alahmadi, A. A. S., Monteverdi, A., Palesi, F., Castellazzi, G., Savini,
G., et al. (2019). I see your effort: force-Related BOLD effects in an extended action
execution-observation network involving the cerebellum. Cereb. Cortex 29, 1351–
1368. doi: 10.1093/cercor/bhy322
Cauda, F., D’Agata, F., Sacco, K., Duca, S., Geminiani, G., and Vercelli, A. (2011).
Functional connectivity of the insula in the resting brain. NeuroImage 55, 8–23. doi:
10.1016/j.neuroimage.2010.11.049
Cerminara, N. L., Lang, E. J., Sillitoe, R. V., and Apps, R. (2015). Redefining the
cerebellar cortex as an assembly of non-uniform Purkinje cell microcircuits. Nat. Rev.
Neurosci. 16, 79–93. doi: 10.1038/nrn3886
Chen, X., Du, Y., Broussard, G. J., Kislin, M., Yuede, C. M., Zhang, S., et al. (2022).
Transcriptomic mapping uncovers Purkinje neuron plasticity driving learning. Nature
605, 722–727. doi: 10.1038/s41586-022- 04711-3
Clausi, S., Iacobacci, C., Lupo, M., Olivito, G., Molinari, M., and Leggio, M. (2017).
The role of the cerebellum in unconscious and conscious processing of emotions: a
review. Appl. Sci. 7:521. doi: 10.3390/app7050521
Croce, P., Basti, A., Marzetti, L., Zappasodi, F., and Gratta, C. D. (2016). EEG-fMRI
Bayesian framework for neural activity estimation: a simulation study. J. Neural Eng.
13:066017. doi: 10.1088/1741-2560/13/6/066017
Damasio, A. R. (1998). Emotion in the perspective of an integrated nervous system.
Brain Res. Brain Res Rev. 26, 83–86.
D’Angelo, E. (2011). Neural circuits of the cerebellum: hypothesis for function.
J. Integr. Neurosci. 10, 317–352. doi: 10.1142/S0219635211002762
Frontiers in Systems Neuroscience 17 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 18
Ciapponi et al. 10.3389/fnsys.2023.1185752
D’Angelo, E., Antonietti, A., Casali, S., Casellato, C., Garrido, J. A., Luque, N. R.,
et al. (2016). Modeling the cerebellar microcircuit: new strategies for a long-standing
issue. Front. Cell. Neurosci. 10:176. doi: 10.3389/fncel.2016.00176
D’Angelo, E., and Jirsa, V. (2022). The quest for multiscale brain modeling. Trends
Neurosci. 45, 777–790. doi: 10.1016/j.tins.2022.06.007
D’Angelo, E., Mazzarello, P., Prestori, F., Mapelli, J., Solinas, S., Lombardo, P., et al.
(2011). The cerebellar network: from structure to function and dynamics. Brain Res.
Rev. 66, 5–15. doi: 10.1016/j.brainresrev.2010.10.002
De Schepper, R., Geminiani, A., Masoli, S., Rizza, M. F., Antonietti, A., Casellato, C.,
et al. (2022). Model simulations unveil the structure-function-dynamics relationship
of the cerebellar cortical microcircuit. Commun. Biol. 5, 1–19. doi: 10.1038/s42003-
022-04213- y
De Zeeuw, C. I. (2021). Bidirectional learning in upbound and downbound
microzones of the cerebellum. Nat. Rev. Neurosci. 22, 92–110. doi: 10.1038/s41583-
020-00392- x
Debas, K., Carrier, J., Orban, P., Barakat, M., Lungu, O., Vandewalle, G., et al.
(2010). Brain plasticity related to the consolidation of motor sequence learning and
motor adaptation. Proc. Natl. Acad. Sci. U. S. A. 107, 17839–17844. doi: 10.1073/pnas.
1013176107
Dieterich, M., Bucher,S. F., Seelos, K. C., and Brandt, T. (2000). Cerebellar activation
during optokinetic stimulation and saccades. Neurology 54, 148–155. doi: 10.1212/wnl.
54.1.148
Dietrichs, E. (1983). Cerebellar cortical afferents from the periaqueductal grey in the
cat. Neurosci. Lett. 41, 21–26. doi: 10.1016/0304-3940(83)90217- 3
Dietrichs, E., and Haines, D. E. (1989). Interconnections between hypothalamus and
cerebellum. Anat. Embryol. 179, 207–220. doi: 10.1007/BF00326585
Dimitrova, A., Kolb, F. P., Elles, H. G., Maschke, M., Forsting, M., Diener, H. C., et al.
(2003). Cerebellar responses evoked by nociceptive leg withdrawal reflex as revealed by
event-related fMRI. J. Neurophysiol. 90, 1877–1886. doi: 10.1152/jn.00053.2003
Dimitrova, A., Weber, J., Maschke, M., Elles, H. G., Kolb, F. P., Forsting, M., et al.
(2002). Eyeblink-related areas in human cerebellum as shown by fMRI. Hum. Brain
Mapp. 17, 100–115. doi: 10.1002/hbm.10056
Dobbins, E. G., and Feldman, J. L. (1994). Brainstem network controlling
descending drive to phrenic motoneurons in rat. J. Comp. Neurol. 347, 64–86. doi:
10.1002/cne.903470106
Dumontier, D., Mailhes-Hamon, C., Supplisson, S., and Dieudonné, S. (2023).
Neurotransmitter content heterogeneity within an interneuron class shapes inhibitory
transmission at a central synapse. Front. Cell Neurosci. 16:1060189. doi: 10.3389/fncel.
2022.1060189
D‘Angelo, E., and Casali, S. (2013). Seeking a unified framework for cerebellar
function and dysfunction: from circuit operations to cognition. Front. Neural Circuits
6:116. doi: 10.3389/fncir.2012.00116
D‘Angelo, E., Solinas, S., Mapelli, J., Gandolfi, D., Mapelli, L., and Prestori, F. (2013).
The cerebellar Golgi cell and spatiotemporal organization of granular layer activity.
Front. Neural Circ 7:93. doi: 10.3389/fncir.2013.00093
E, K. H., Chen, S. H. A., Ho, M. H. R., and Desmond, J. E. (2014). A meta-analysis of
cerebellar contributions to higher cognition from PET and fMRI studies. Hum. Brain
Mapp. 35, 593–615. doi: 10.1002/hbm.22194
Eccles, J. C. (1967). Circuits in the cerebellar control of movement. Proc. Natl. Acad.
Sci. U. S. A. 58, 336–343. doi: 10.1073/pnas.58.1.336
Elliott, M. L., Knodt, A. R., Ireland, D., Morris, M. L., Poulton, R., Ramrakha,
S., et al. (2020). What is the test-retest reliability of common task-functional MRI
measures? New empirical evidence and a meta-analysis. Psychol. Sci. 31, 792–806.
doi: 10.1177/0956797620916786
Ernst, T. M., Brol, A. E., Gratz, M., Ritter, C., Bingel, U., Schlamann, M., et al. (2019).
The cerebellum is involved in processing of predictions and prediction errors in a fear
conditioning paradigm. eLife 8:e46831. doi: 10.7554/eLife.46831
Fink, G. R., Marshall, J. C., Shah, N. J., Weiss, P. H., Halligan, P. W., Grosse-
Ruyken, M., et al. (2000). Line bisection judgments implicate right parietal cortex and
cerebellum as assessed by fMRI. Neurology 54, 1324–1331. doi: 10.1212/WNL.54.6.
1324
Fox, M. D., and Raichle, M. E. (2007). Spontaneous fluctuations in brain activity
observed with functional magnetic resonance imaging. Nat. Rev. Neurosci. 8, 700–711.
doi: 10.1038/nrn2201
Friede, R. (1955). [Quantitative displacement of the layers within the
convolutions of the cerebellar cortex and its biologic importance]. Acta Anat. 25,
65–72.
Fujita, H., Morita, N., Furuichi, T., and Sugihara, I. (2012). Clustered
fine compartmentalization of the mouse embryonic cerebellar cortex and its
rearrangement into the postnatal striped configuration. J. Neurosci. 32, 15688–15703.
doi: 10.1523/JNEUROSCI.1710-12.2012
Gao, Z., Proietti-Onori, M., Lin, Z., ten Brinke, M. M., Boele, H.-J., Potters, J.-W.,
et al. (2016). Excitatory cerebellar nucleocortical circuit provides internal amplification
during associative conditioning. Neuron 3, 645–657. doi: 10.1016/j.neuron.2016.01.
008
Geurts, F. J., De Schutter, E., and Dieudonné, S. (2003). Unraveling the cerebellar
cortex: cytology and cellular physiology of large-sized interneurons in the granular
layer. Cerebellum 2, 290–299. doi: 10.1080/14734220310011948
Geurts, F. J., Timmermans, J., Shigemoto, R., and De Schutter, E. (2001).
Morphological and neurochemical differentiation of large granular layer interneurons
in the adult rat cerebellum. Neuroscience 104, 499–512. doi: 10.1016/s0306-4522(01)
00058-6
Gorges, M., Roselli, F., Müller, H. P., Ludolph, A. C., Rasche, V., and Kassubek,
J. (2017). Functional connectivity mapping in the animal model: principles and
applications of resting-state fMRI. Front. Neurol. 8:200. doi: 10.3389/fneur.2017.00200
Gornati, S. V., Schäfer, C. B., Eelkman Rooda, O. H. J., Nigg, A. L., De Zeeuw, C. I.,
and Hoebeek, F. E. (2018). Differentiating cerebellar impact on thalamic nuclei. Cell
Rep. 23, 2690–2704. doi: 10.1016/j.celrep.2018.04.098
Grandjean, J., Canella, C., Anckaerts, C., Ayrancı, G., Bougacha, S., Bienert, T., et al.
(2020). Common functional networks in the mouse brain revealed by multi-centre
resting-state fMRI analysis. NeuroImage 205:116278. doi: 10.1016/j.neuroimage.2019.
116278
Gritti, I. (2013). The cerebellum, the hypothalamus and behavior. Nat. Sci. 5,
832–834. doi: 10.4236/ns.2013.57100
Guell, X., Gabrieli, J. D. E., and Schmahmann, J. D. (2018a). Triple representation
of language, working memory, social and emotion processing in the cerebellum:
convergent evidence from task and seed-based resting-state fMRI analyses in a single
large cohort. NeuroImage 172, 437–449. doi: 10.1016/j.neuroimage.2018.01.082
Guell, X., Schmahmann, J. D., Gabrieli, J. D. E., and Ghosh, S. S. (2018b). Functional
gradients of the cerebellum. eLife 7:e36652. doi: 10.7554/eLife.36652
Guell, X., Goncalves, M., Kaczmarzyk, J. R., Gabrieli, J. D. E., Schmahmann, J. D.,
and Ghosh, S. S. (2019). Littlebrain: a gradient-based tool for the topographical
interpretation of cerebellar neuroimaging findings. PLoS One 14:e0210028. doi: 10.
1371/journal.pone.0210028
Guell, X., and Schmahmann, J. (2020). Cerebellar functional anatomy: a didactic
summary based on human fMRI evidence. Cerebellum 19, 1–5. doi: 10.1007/s12311-
019-01083- 9
Guo, C., Rudolph, S., Neuwirth, M. E., and Regehr, W. G. (2021). Purkinje cell
outputs selectively inhibit a subset of unipolar brush cells in the input layer of the
cerebellar cortex. eLife 10:e68802. doi: 10.7554/eLife.68802
Guo, C., Witter,L., Rudolph, S., Elliott, H. L., Ennis, K. A., and Regehr, W. G. (2016).
Purkinje cells directly inhibit granule cells in specialized regions of the cerebellar
cortex. Neuron 91, 1330–1341. doi: 10.1016/j.neuron.2016.08.011
Habas, C. (2018). Research note: a resting-state, cerebello-amygdaloid
intrinsically connected network. Cereb. Ataxias 5:4. doi:10.1186/s40673-01
8-0083-0
Habas, C., Kamdar, N., Nguyen, D., Prater, K., Beckmann, C. F., Menon, V., et al.
(2009). Distinct cerebellar contributions to intrinsic connectivity networks. J. Neurosci.
29, 8586–8594. doi: 10.1523/JNEUROSCI.1868-09.2009
Habas, C., and Manto, M. (2018). Probing the neuroanatomy of the cerebellum
using tractography. Handb. Clin. Neurol 154, 235–249. doi: 10.1016/B978-0-444-
63956-1.00014- X
Hardwick, R. M., Rottschy, C., Miall, R. C., and Eickhoff, S. B. (2013). A quantitative
meta-analysis and review of motor learning in the human brain. NeuroImage 67,
283–297. doi: 10.1016/j.neuroimage.2012.11.020
Heath, N. C., Rizwan, A. P., Engbers, J. D. T., Anderson, D., Zamponi, G. W., and
Turner, R. W. (2014). The expression pattern of a Cav3–Kv4 complex differentially
regulates spike output in cerebellar granule cells. J. Neurosci. 34, 8800–8812. doi:
10.1523/JNEUROSCI.0981-14.2014
Heinemans, M., and Moita, M. A. (2022). Looming stimuli reliably drive innate, but
not learned, defensive responses in rats. bioRxiv[Preprint]. doi: 10.1101/2022.02.07.
479432
Herman, J. P., Ostrander, M. M., Mueller, N. K., and Figueiredo, H. (2005). Limbic
system mechanisms of stress regulation: hypothalamo-pituitary-adrenocortical axis.
Prog. Neuro Psychopharmacol. Biol. Psychiatry 29, 1201–1213. doi: 10.1016/j.pnpbp.
2005.08.006
Hoche, F., Guell, X., Vangel, M. G., Sherman, J. C., and Schmahmann, J. D. (2018).
The cerebellar cognitive affective/Schmahmann syndrome scale. Brain 141, 248–270.
doi: 10.1093/brain/awx317
Holstege, G., and Huynh, H. K. (2011). Brain circuits for mating behavior in cats
and brain activations and de-activations during sexual stimulation and ejaculation and
orgasm in humans. Hormones Behav. 59, 702–707. doi: 10.1016/j.yhbeh.2011.02.008
Houck, B. D., and Person, A. L. (2014). Cerebellar loops: a review of the
nucleocortical pathway. Cerebellum 13, 378–385. doi: 10.1007/s12311-013- 0543-2
Huster, R. J., Debener, S., Eichele, T., and Herrmann, C. S. (2012). Methods for
simultaneous EEG-fMRI: an introductory review. J. Neurosci. 32, 6053–6060. doi:
10.1523/JNEUROSCI.0447-12.2012
ISH Data. (2023). Allen Brain Atlas: Mouse Brain. Washington, DC: Allen Institute
for Brain Science
Itô, M. (1984). The Cerebellum and Neural Control. Billings: Raven Press.
Frontiers in Systems Neuroscience 18 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 19
Ciapponi et al. 10.3389/fnsys.2023.1185752
Ito, M. (1993). Movement and thought: identical control mechanisms by the
cerebellum. Trends Neurosci. 16, 448–450. doi: 10.1016/0166-2236(93)90073- U
Ito, M. (2005). Bases and implications of learning in the cerebellum–adaptive
control and internal model mechanism. Prog. Brain Res. 148, 95–109. doi: 10.1016/
S0079-6123(04)48009- 1
Ito, M. (2006). Cerebellar circuitry as a neuronal machine. Prog. Neurobiol. 78,
272–303. doi: 10.1016/j.pneurobio.2006.02.006
Ito, M. (2008). Control of mental activities by internal models in the cerebellum.
Nat. Rev. Neurosci. 9, 304–313. doi: 10.1038/nrn2332
Jackson, P. L., Meltzoff, A. N., and Decety, J. (2005). How do we perceive the pain
of others? A window into the neural processes involved in empathy. NeuroImage 24,
771–779. doi: 10.1016/j.neuroimage.2004.09.006
Jorge, J., van der Zwaag, W., and Figueiredo, P. (2014). EEG–fMRI integration
for the study of human brain function. NeuroImage 102, 24–34. doi: 10.1016/j.
neuroimage.2013.05.114
Jung, S. J., Vlasov, K., D’Ambra, A. F., Parigi, A., Baya, M., Frez, E. P., et al. (2022).
Novel cerebello-amygdala connections provide missing link between cerebellum and
limbic system. Front. Syst. Neurosci. 16:879634. doi: 10.3389/fnsys.2022.879634
Kang, S., Jun, S., Baek, S. J., Park, H., Yamamoto, Y., Tanaka-Yamamoto K., et al.
(2021). Recent advances in the understanding of specific efferent pathways emerging
from the cerebellum. Front. Neuroanat. 5:759948. doi: 10.3389/fnana.2021.759948
Keay, K. A., and Bandler, R. (2001). Parallel circuits mediating distinct emotional
coping reactions to different types of stress. Neurosci. Biobehav. Rev. 25, 669–678.
doi: 10.1016/S0149-7634(01)00049- 5
Kebschull, J. M., Casoni, F., Consalez, G. G., Goldowitz, D., Hawkes, R., Ruigrok,
T. J. H., et al. (2023). Cerebellum lecture: the cerebellar nuclei—core of the cerebellum.
Cerebellum [Online ahead of print]. doi: 10.1007/s12311-022- 01506-0
Keller, D., Erö, C., and Markram, H. (2018). Cell densities in the mouse brain: a
systematic review. Front. Neuroanat. 12:83. doi: 10.3389/fnana.2018.00083
Kilts, C. D., Egan, G., Gideon, D. A., Ely, T. D., and Hoffman, J. M. (2003).
Dissociable neural pathways are involved in the recognition of emotion in static and
dynamic facial expressions. NeuroImage 18, 156–168. doi: 10.1006/nimg.2002.1323
Kim, C.-H., Shin, J. J., Kim, J., and Kim, S. J. (2013). Reduced spike frequency
adaptation in Purkinje cells of the vestibulocerebellum. Neurosci. Lett. 535, 45–50.
doi: 10.1016/j.neulet.2012.12.045
King, M., Hernandez-Castillo, C. R., Poldrack, R. A., Ivry, R. B., and Diedrichsen, J.
(2019). Functional boundaries in the human cerebellum revealed by a multi-domain
task battery. Nat. Neurosci. 22, 1371–1378. doi: 10.1038/s41593-019-0436- x
Koutsikou, S., Crook, J. J., Earl, E. V., Leith, J. L., Watson, T. C., Lumb, B. M.,
et al. (2014). Neural substrates underlying fear-evoked freezing: the periaqueductal
grey–cerebellar link. J. Physiol. 592, 2197–2213. doi: 10.1113/jphysiol.2013.268714
Kozareva, V., Martin, C., Osorno, T., Rudolph, S., Guo, C., Vanderburg, C., et al.
(2021). A transcriptomic atlas of mouse cerebellar cortex comprehensively defines cell
types. Nature 598, 214–219. doi: 10.1038/s41586-021-03220- z
Krienen, F. M., and Buckner, R. L. (2009). Segregated fronto-cerebellar circuits
revealed by intrinsic functional connectivity. Cereb. Cortex 19, 2485–2497. doi: 10.
1093/cercor/bhp135
Kumral, E., Çelebisoy, M., Çelebisoy, N., Canbaz, D. H., and Çalli, C. (2007).
Dysarthria due to supratentorial and infratentorial ischemic stroke: a diffusion-
weighted imaging study. Cerebrovasc. Dis. 23, 331–338. doi: 10.1159/000099131
Lainé, J., and Axelrad, H. (1994). The candelabrum cell: a new interneuron in the
cerebellar cortex. J. Comp. Neurol. 339, 159–173. doi: 10.1002/cne.903390202
Lange, I., Kasanova, Z., Goossens, L., Leibold, N., De Zeeuw, C. I., van Amelsvoort,
T., et al. (2015). The anatomy of fear learning in the cerebellum: a systematic meta-
analysis. Neurosci. Biobehav. Rev. 59, 83–91. doi: 10.1016/j.neubiorev.2015.09.019
Lange, W. (1974). Regional differences in the distribution of golgi cells in the
cerebellar cortex of man and some other mammals. Cell Tissue Res. 153, 219–226.
doi: 10.1007/BF00226610
LeDoux, J., Cicchetti, P., Xagoraris, A., and Romanski, L. (1990). The lateral
amygdaloid nucleus: sensory interface of the amygdala in fear conditioning.
J. Neurosci. 10:1062. doi: 10.1523/JNEUROSCI.10-04- 01062.1990
LeDoux, J., Iwata, J., Cicchetti, P., and Reis, D. (1988). Different projections of
the central amygdaloid nucleus mediate autonomic and behavioral correlates of
conditioned fear. J. Neurosci. 8:2517. doi: 10.1523/JNEUROSCI.08-07- 02517.1988
LeDoux, J. E. (1990). “Fear pathways in the brain: implications for a theory of the
emotional brain, in Fear and Defence, eds P. F. Brain, S. Parmigiani, R. J. Blanchard,
and D. Mainardi (Amsterdam: Harwood Academic Publishers), 163–177.
LeDoux, J. E. (1992). Brain mechanisms of emotion and emotional learning. Curr.
Opin. Neurobiol. 2, 191–197. doi: 10.1016/0959-4388(92)90011- 9
LeDoux, J. E. (2000). Emotion circuits in the brain. Annu Rev Neurosci. 23, 155–184
Leggio, M., and Olivito, G. (2018). Topography of the cerebellum in relation to social
brain regions and emotions. Handb. Clin. Neurol. 154, 71–84. doi: 10.1016/B978-0-
444-63956- 1.00005-9
Levisohn, L., Cronin-Golomb, A., and Schmahmann, J. D. (2000).
Neuropsychological consequences of cerebellar tumour resection in children.
Cerebellar cognitive affective syndrome in a paediatric population. Brain 123,
1041–1050. doi: 10.1093/brain/123.5.1041
Locatelli, F., Bottà, L., Prestori, F., Masetto, S., and D’Angelo, E. (2013). Late-onset
bursts evoked by mossy fibre bundle stimulation in unipolar brush cells: evidence for
the involvement of H- and TRP-currents. J. Physiol. 4, 899–918. doi: 10.1113/jphysiol.
2012.242180
Logothetis, N. K. (2008). What we can do and what we cannot do with fMRI. Nature
453, 869–878. doi: 10.1038/nature06976
Lu, L., Cao, Y., Tokita, K., Heck, D., and Boughter, J. (2013). Medial cerebellar
nuclear projections and activity patterns link cerebellar output to orofacial and
respiratory behavior. Front. Neural Circuits 7:56. doi: 10.3389/fncir.2013.00056
Ma, L., Wang, B., Narayana, S., Hazeltine, E., Chen, X., Robin, D. A., et al.
(2010). Changes in regional activity are accompanied with changes in inter-regional
connectivity during 4 weeks motor learning. Brain Res. 1318, 64–76. doi: 10.1016/j.
brainres.2009.12.073
Maclean (1990). The Triune Brain in Evolution: Role in Paleocerebral Functions.
Berlin: Springer Science+Business Media.
Maderwald, S., Thürling, M., Küper, M., Theysohn, N., Müller, O., Beck, A., et al.
(2012). Direct visualization of cerebellar nuclei in patients with focal cerebellar lesions
and its application for lesion-symptom mapping. NeuroImage 63, 1421–1431. doi:
10.1016/j.neuroimage.2012.07.063
Maestú, F., Quesney-Molina, F., Ortiz-Alonso, T., Campo, P., Fernández-Lucas, A.,
and Amo, C. (2003). Cognition and neural networks, a new perspective based on
functional neuroimaging. Rev. Neurol. 37, 962–966. doi: 10.33588/rn.3710.2003069
Mapelli, L., Solinas, S., and D‘Angelo, E. (2014). Integration and regulation of
glomerular inhibition in the cerebellar granular layer circuit. Front. Cell. Neurosci.
8:55. doi: 10.3389/fncel.2014.00055
Mariën, P., Ackermann, H., Adamaszek, M., Barwood, C. H. S., Beaton, A.,
Desmond, J., et al. (2014). Consensus paper: language and the cerebellum: an ongoing
enigma. Cerebellum 13, 386–410. doi: 10.1007/s12311-013- 0540-5
Martinez, S., Andreu, A., Mecklenburg, N., and Echevarria, D. (2013). Cellular and
molecular basis of cerebellar development. Front. Neuroanat. 7:18. doi: 10.3389/fnana.
2013.00018
Mascagni, F., McDonald, A. J., and Coleman, J. R. (1993). Corticoamygdaloid
and corticocortical projections of the rat temporal cortex: a Phaseolus vulgaris
leucoagglutinin study. Neuroscience 57, 697–715. doi: 10.1016/0306-4522(93)90016- 9
McDonald, A. J. (1998). Cortical pathways to the mammalian amygdala. Progr.
Neurobiol. 55, 257–332. doi: 10.1016/S0301-0082(98)00003- 3
Menuet, C., Connelly, A. A., Bassi, J. K., Melo, M. R., Le, S., Kamar, J., et al. (2020).
PreBötzinger complex neurons drive respiratory modulation of blood pressure and
heart rate. Elife 9:e57288. doi: 10.7554/eLife.57288
Milardi, D., Arrigo, A., Anastasi, G., Cacciola, A., Marino, S., Mormina, E.,
et al. (2016). Extensive direct subcortical cerebellum-basal ganglia connections in
human brain as revealed by constrained spherical deconvolution tractography. Front.
Neuroanat. 10:29. doi: 10.3389/fnana.2016.00029
Moulton, E. A., Elman, I., Pendse, G., Schmahmann, J., Becerra, L., and Borsook,
D. (2011). Aversion-related circuitry in the cerebellum: responses to noxious heat
and unpleasant images. J. Neurosci. 31, 3795–3804. doi: 10.1523/JNEUROSCI.6709-
10.2011
Mugnaini, E., Sekerkova, G., and Martina, M. (2011). The unipolar brush cell: a
remarkable neuron finally receiving the deserved attention. Brain Res. Rev. 66:220.
doi: 10.1016/j.brainresrev.2010.10.001
Müller, U., and Heinsen, H. (1984). Regional differences in the ultrastructure
of purkinje cells of the rat. Cell Tissue Res. 235, 91–98. doi: 10.1007/BF0021
3728
Murty, V. P., Shermohammed, M., Smith, D. V., Carter, R. M. K., Huettel, S. A.,
and Adcock, R. A. (2014). Resting state networks distinguish human ventral tegmental
area from substantia nigra. NeuroImage 100, 580–589. doi:10.1016/j.neuroimage.201
4.06.047
Nakano, I. (1998). The limbic system: an outline and brief history of its concept.
Neuropathology 18, 211–214. doi: 10.1111/j.1440-1789.1998.tb00102.x
Nedelescu, H., and Abdelhack, M. (2013). Comparative morphology of dendritic
arbors in populations of Purkinje cells in mouse sulcus and apex. Neural Plast.
2013:948587. doi: 10.1155/2013/948587
Nedelescu, H., Abdelhack, M., and Pritchard, A. T. (2018). Regional differences in
Purkinje cell morphology in the cerebellar vermis of male mice. J. Neurosci. Res. 96,
1476–1489. doi: 10.1002/jnr.24206
Nishiyama, H., and Linden, D. J. (2004). Differential maturation of climbing
fiber innervation in cerebellar vermis. J. Neurosci. 24, 3926–3932. doi: 10.1523/
JNEUROSCI.5610-03.2004
Nisimaru, N. (2004). Cardiovascular modules in the cerebellum. Jpn. J. Physiol. 54,
431–448. doi: 10.2170/jjphysiol.54.431
Frontiers in Systems Neuroscience 19 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 20
Ciapponi et al. 10.3389/fnsys.2023.1185752
Nummenmaa, L., Glerean, E., Hari, R., and Hietanen, J. K. (2014). Bodily maps
of emotions. Proc. Natl. Acad. Sci. U. S. A. 111, 646–651. doi:10.1073/pnas.13216
64111
Nunzi, M. G., and Mugnaini, E. (2000). Unipolar brush cell axons form a large
system of intrinsic mossy fibers in the postnatal vestibulocerebellum. J. Comp. Neurol.
422, 55–65.
Ogawa, S., Tank, D. W., Menon, R., Ellermann, J. M., Kim, S. G., Merkle, H.,
et al. (1992). Intrinsic signal changes accompanying sensory stimulation: functional
brain mapping with magnetic resonance imaging. Proc. Natl. Acad. Sci. U. S. A. 89,
5951–5955. doi: 10.1073/pnas.89.13.5951
O’Reilly, J. X., Beckmann, C. F., Tomassini, V., Ramnani, N., and Johansen-Berg, H.
(2010). Distinct and overlapping functional zones in the cerebellum defined by resting
state functional connectivity. Cereb. Cortex 20, 953–965. doi: 10.1093/cercor/bhp157
Osorno, T., and Regehr, W. G. (2022). Candelabrum cells are ubiquitous cerebellar
cortex interneurons with specialized circuit properties. Nat. Neurosci. 25, 702–713.
doi: 10.1038/s41593-022- 01057-x
Palay, S. L., and Chan-Palay, V. (1982). The Cerebellum: New Vistas. Berlin: Springer-
Verlag.
Palesi, F., Tournier, J. D., Calamante, F., Muhlert, N., Castellazzi, G., Chard,
D., et al. (2015). Contralateral cerebello-thalamo-cortical pathways with prominent
involvement of associative areas in humans in vivo. Brain Struct. Funct. 220, 3369–
3384. doi: 10.1007/s00429-014- 0861-2
Pan, W. J., Billings, J., Nezafati, M., Abbas, A., and Keilholz, S. (2018). Resting state
fMRI in rodents. Curr. Protocols Neurosci. 83:e45. doi: 10.1002/cpns.45
Pan, W. J., Billings, J. C. W., Grooms, J. K., Shakil, S., and Keilholz, S. D. (2015).
Considerations for resting state functional MRI and functional connectivity studies in
rodents. Front. Neurosci. 9:269. doi: 10.3389/fnins.2015.00269
Papez J. W. (1937). A proposed mechanism of emotion. J. Neuropsychiatry Clin.
Neurosci. 7, 103–112
Parma, A. (1969). [The size of cell corpuscles of Purkinje cells in the paleocerebellum
and neocerebellum of some mammals as compared to man]. Acta Anat. Suppl. 56,
337–346.
Parsons, L. M., Denton, D., Egan, G., McKinley, M., Shade, R., Lancaster,
J., et al. (2000). Neuroimaging evidence implicating cerebellum in support of
sensory/cognitive processes associated with thirst. Proc. Natl. Acad. Sci. U. S. A. 97,
2332–2336. doi: 10.1073/pnas.040555497
Ploghaus, A., Tracey, I., Gati, J. S., Clare, S., Menon, R. S., Matthews, P. M., et al.
(1999). Dissociating pain from its anticipation in the human brain. Science 284,
1979–1981. doi: 10.1126/science.284.5422.1979
Pradier, B., Wachsmuth, L., Nagelmann, N., Segelcke, D., Kreitz, S., Hess, A., et al.
(2021). Combined resting state-fMRI and calcium recordings show stable brain states
for task-induced fMRI in mice under combined ISO/MED anesthesia. NeuroImage
245:118626. doi: 10.1016/j.neuroimage.2021.118626
Prestori, F., Mapelli, L., and D’Angelo, E. (2019). Diverse neuron properties and
complex network dynamics in the cerebellar cortical inhibitory circuit. Front. Mol.
Neurosci. 12:267. doi: 10.3389/fnmol.2019.00267
Pruessner, J. C., Dedovic, K., Pruessner, M., Lord, C., Buss, C., Collins, L., et al.
(2010). Stress regulation in the central nervous system: evidence from structural and
functional neuroimaging studies in human populations - 2008 Curt Richter Award
Winner. Psychoneuroendocrinology 35, 179–191. doi: 10.1016/j.psyneuen.2009.02.016
Riecker, A., Mathiak, K., Wildgruber, D., Erb, M., Hertrich, I., Grodd, W., et al.
(2005). fMRI reveals two distinct cerebral networks subserving speech motor control.
Neurology 64:89. doi: 10.1212/01.WNL.0000152156.90779.89
Rodarie, D., Verasztó, C., Roussel, Y., Reimann, M., Keller, D., Ramaswamy, S.,
et al. (2022). A method to estimate the cellular composition of the mouse brain from
heterogeneous datasets. PLoS Comput. Biol. 18:e1010739. doi: 10.1371/journal.pcbi.
1010739
Rolls, E. T. (2019). The cingulate cortex and limbic systems for action, emotion, and
memory. Handb. Clin. Neurol. 166, 23–37. doi: 10.1016/B978-0-444- 64196-0.00002-9
Romano, V., Reddington, A. L., Cazzanelli, S., Mazza, R., Ma, Y., Strydis, C., et al.
(2020). Functional convergence of autonomic and sensorimotor processing in the
lateral cerebellum. Cell Rep. 32:107867. doi: 10.1016/j.celrep.2020.107867
Romanski, L. M., and LeDoux, J. E. (1993). Information cascade from primary
auditory cortex to the amygdala: corticocortical and corticoamygdaloid projections of
temporal cortex in the rat. Cereb. Cortex 3, 515–532. doi: 10.1093/cercor/3.6.515
Roxo, M. R., Franceschini, P. R., Zubaran, C., Kleber, F. D., and Sander, J. W. (2011).
The limbic system conception and its historical evolution. Sci. World J. 11, 2427–2440.
doi: 10.1100/2011/157150
Sacchetti, B., Baldi, E., Lorenzini, C. A., and Bucherelli, C. (2002). Cerebellar role
in fear-conditioning consolidation. Proc. Natl. Acad. Sci. U. S. A. 99, 8406–8411.
doi: 10.1073/pnas.112660399
Sacchetti, B., Scelfo, B., Tempia, F., and Strata, P. (2004). Long-term synaptic
changes induced in the cerebellar cortex by fear conditioning. Neuron 42, 973–982.
doi: 10.1016/j.neuron.2004.05.012
Sang, L., Qin, W., Liu, Y., Han, W., Zhang, Y., Jiang, T., et al. (2012). Resting-state
functional connectivity of the vermal and hemispheric subregions of the cerebellum
with both the cerebral cortical networks and subcortical structures. NeuroImage 61,
1213–1225. doi: 10.1016/j.neuroimage.2012.04.011
Schlerf, J. E., Verstynen, T. D., Ivry, R. B., and Spencer, R. M. C. (2010). Evidence
of a novel somatopic map in the human neocerebellum during complex actions.
J. Neurophysiol. 103, 3330–3336. doi: 10.1152/jn.01117.2009
Schmahmann, J. D. (2004). Disorders of the cerebellum: ataxia, dysmetria of
thought, and the cerebellar cognitive affective syndrome. J. Neuropsychiatry Clin.
Neurosci. 16, 367–378. doi: 10.1176/jnp.16.3.367
Schmahmann, J. D., and Caplan, D. (2006). Cognition, emotion and the cerebellum.
Brain 129, 290–292. doi: 10.1093/brain/awh729
Schmahmann, J. D., Doyon, J., McDonald, D., Holmes, C., Lavoie, K., Hurwitz, A.,
et al. (1999). Three-dimensional MRI atlas of the human cerebellum in proportional
stereotaxic space. NeuroImage 10, 233–260. doi: 10.1006/nimg.1999.0459
Schmahmann, J. D., Gardner, R., MacMore, J., and Vangel, M. G. (2009).
Development of a brief ataxia rating scale (BARS) based on a modified form of the
ICARS. Move. Disord. 24, 1820–1828. doi: 10.1002/mds.22681
Schmahmann, J. D., and Sherman, J. C. (1998). The cerebellar cognitive affective
syndrome. Brain 121, 561–579. doi: 10.1093/brain/121.4.561
Schmahmann, J. D., Weilburg, J. B., and Sherman, J. C. (2007). The neuropsychiatry
of the cerebellum - insights from the clinic. Cerebellum 6, 254–267. doi: 10.1080/
14734220701490995
Schraa-Tam, C. K. L., Rietdijk, W. J. R., Verbeke, W. J. M. I., Dietvorst, R. C.,
Van Den Berg, W. E., Bagozzi, R. P., et al. (2012). fMRI activities in the emotional
cerebellum: a preference for negative stimuli and goal-directed behavior. Cerebellum
11, 233–245. doi: 10.1007/s12311-011- 0301-2
Sereno, M. I., Diedrichsen, J. R. N., Tachrount, M., Testa-Silva, G., D Arceuil, H.,
and De Zeeuw, C. (2020). The human cerebellum has almost 80% of the surface area
of the neocortex. Proc. Natl. Acad. Sci. U. S. A. 117, 19538–19543. doi: 10.1073/pnas.
2002896117
Shang, C., Chen, Z., Liu, A., Li, Y., Zhang, J., Qu, B., et al. (2018). Divergent midbrain
circuits orchestrate escape and freezing responses to looming stimuli in mice. Nat.
Commun. 9:1232. doi: 10.1038/s41467-018- 03580-7
Signoret-Genest, J., Schukraft, N. L., Reis, S., Segebarth, D., Deisseroth, K., and
Tovote, P. (2023). Integrated cardio-behavioral responses to threat define defensive
states. Nat. Neurosci. 26, 447–457. doi: 10.1038/s41593-022-01252- w
Sillitoe, R. V., Gopal, N., and Joyner, A. L. (2009). Embryonic origins of ZebrinII
parasagittal stripes and establishment of topographic Purkinje cell projections.
Neuroscience 162, 574–588. doi: 10.1016/j.neuroscience.2008.12.025
Singer, T., Dolan, R. J., and Frith, C. D. (2004). Empathy for pain involves the
affective but not sensory components of pain. Science 303, 1157–1162.
Smith, J. C., Ellenberger, H. H., Ballanyi, K., Richter, D. W., and Feldman, J. L.
(1991). Pre-Bötzinger complex: a brainstem region that may generate respiratory
rhythm in mammals. Science 254, 726–729. doi: 10.1126/science.1683005
Snider, R., and Eldred, E. (1951). Electro-anatomical studies on cerebro-cerebellar
connections in the cat. J. Comp. Neurol. 95, 1–16. doi: 10.1002/cne.900950102
Solstrand Dahlberg, L., Lungu, O., and Doyon, J. (2020). Cerebellar contribution
to motor and non-motor functions in parkinson’s disease: a meta-analysis of fMRI
findings. Front. Neurol. 11:127. doi: 10.3389/fneur.2020.00127
Spencer, A. P. C., and Goodfellow, M. (2022). Using deep clustering to improve
fMRI dynamic functional connectivity analysis. NeuroImage 257:119288. doi: 10.1016/
j.neuroimage.2022.119288
Steffen, P. R., Hedges, D., and Matheson, R. (2022). The brain is adaptive not triune:
how the brain responds to threat, challenge, and change. Front. Psychiatry 13:802606.
doi: 10.3389/fpsyt.2022.802606
Stoodley, C. J., Desmond, J. E., Guell, X., and Schmahmann, J. D. (2021). “Functional
topography of the human cerebellum revealed by functional neuroimaging studies,
in Handbook of the Cerebellum and Cerebellar Disorders, eds M. Manto, D. Gruol, J.
Schmahmann, N. Koibuchi, and R. Sillitoe (Berlin: Springer). doi: 10.1007/978-3- 319-
97911-3_30- 2
Stoodley, C. J., MacMore, J. P., Makris, N., Sherman, J. C., and Schmahmann, J. D.
(2016). Location of lesion determines motor vs. cognitive consequences in patients
with cerebellar stroke. NeuroImage 12, 765–775. doi: 10.1016/j.nicl.2016.10.013
Stoodley, C. J., and Schmahmann, J. D. (2009). Functional topography in the human
cerebellum: a meta-analysis of neuroimaging studies. NeuroImage 44, 489–501. doi:
10.1016/j.neuroimage.2008.08.039
Stoodley, C. J., Valera,E. M., and Schmahmann, J. D. (2010). An fMRI study of intra-
individual functional topography in the human cerebellum. Behav. Neurol. 23, 65–79.
doi: 10.1155/2010/840942
Stoodley, C. J., Valera,E. M., and S chmahmann, J. D. (2012). Functional topography
of the cerebellum for motor and cognitive tasks: an fMRI study. NeuroImage 59,
1560–1570. doi: 10.1016/j.neuroimage.2011.08.065
Frontiers in Systems Neuroscience 20 frontiersin.org
fnsys-17-1185752 May 3, 2023 Time: 13:42 # 21
Ciapponi et al. 10.3389/fnsys.2023.1185752
Strata, P. (2015). The emotional cerebellum. Cerebellum 14, 570–577. doi: 10.1007/
s12311-015- 0649-9
Strick, P. L., Dum, R. P., and Fiez, J. A. (2009). Cerebellum and nonmotor
function. Annu. Rev. Neurosci. 32, 413–434. doi:10.1146/annurev.neuro.31.060407.1
25606
Supple, W. F., Leaton, R. N., and Fanselow, M. S. (1987). Effects of cerebellar vermal
lesions on species-specific fear responses, neophobia, and taste-aversion learning in
rats. Physiol. Behav. 39, 579–586. doi: 10.1016/0031-9384(87)90156-9
Takács, J., Markova, L., Borostyánköi, Z., Görcs, T. J., and Hámori, J. (1999).
Metabotrop glutamate receptor type 1a expressing unipolar brush cells in the
cerebellar cortex of different species: a comparative quantitative study. J. Neurosci. Res.
55, 733–748.
Takeo, Y. H., Shuster, S. A., Jiang, L., Hu, M. C., Luginbuhl, D. J., Rülicke, T.,
et al. (2021). GluD2- and Cbln1-mediated competitive interactions shape the dendritic
arbors of cerebellar Purkinje cells. Neuron 109, 629–644. doi: 10.1016/j.neuron.2020.
11.028
Tavano, A., Grasso, R., Gagliardi, C., Triulzi, F., Bresolin, N., Fabbro, F., et al. (2007).
Disorders of cognitive and affective development in cerebellar malformations. Brain
130, 2646–2660. doi: 10.1093/brain/awm201
Teune, T. M., van der Burg, J., van der Moer, J., Voogd, J., and Ruigrok,
T. J. H. (2000). Topography of cerebellar nuclear projections to the brain
stem in the rat. Prog. Brain Res. 124, 141–172. doi:10.1016/S0079-6123(00)24
014-4
Thomasson, M., Saj, A., Benis, D., Grandjean, D., Assal, F., and Péron, J.
(2019). Cerebellar contribution to vocal emotion decoding: insights from stroke and
neuroimaging. Neuropsychologia 132:107141. doi: 10.1016/j.neuropsychologia.2019.
107141
Tovote, P., Fadok, J. P., and Lüthi, A. (2015). Neuronal circuits for fear and anxiety.
Nat. Rev. Neurosci. 16, 317–331. doi: 10.1038/nrn3945
Trott, J. R., Apps, R., and Armstrong, D. M. (1998a). Zonal organization of
cortico-nuclear and nucleo-cortical projections of the paramedian lobule of the cat
cerebellum. 1. The C1 zone. Exp. Brain Res. 118, 298–315. doi:10.1007/s0022100
50285
Trott, J. R., Apps, R., and Armstrong, D. M. (1998b). Zonal organization of
cortico-nuclear and nucleo-cortical projections of the paramedian lobule of the cat
cerebellum. 2. The C2 zone. Exp. Brain Res. 118, 316–330. doi:10.1007/s0022100
50286
Turner, B. H., and Zimmer, J. (1984). The architecture and some of the
interconnections of the rat’s amygdala and lateral periallocortex. J. Comp. Neurol. 227,
540–557. doi: 10.1002/cne.902270406
Tzvi, E., Koeth, F., Karabanov, A. N., Siebner, H. R., and Krämer, U. M.
(2020). Cerebellar Premotor cortex interactions underlying visuomotor
adaptation. NeuroImage 220, 117142. doi: 10.1016/j.neuroimage.2020.
117142
Urban, P. P., Marx, J., Hunsche, S., Gawehn, J., Vucurevic, G., Wicht,
S., et al. (2003). Cerebellar speech representation: lesion topography in
dysarthria as derived from cerebellar ischemia and functional magnetic
resonance imaging. Arch. Neurol. 60, 965–972. doi: 10.1001/archneur.60.
7.965
Van Dijk, K. R. A., Hedden, T., Venkataraman, A., Evans, K. C., Lazar, S. W.,
and Buckner, R. L. (2010). Intrinsic functional connectivity as a tool for human
connectomics: theory, properties, and optimization. J. Neurophysiol. 103, 297–321.
doi: 10.1152/jn.00783.2009
Van Overwalle, F., Baetens, K., Mariën, P., and Vandekerckhove, M. (2014). Social
cognition and the cerebellum: a meta-analysis of over 350 fMRI studies. NeuroImage
86, 554–572. doi: 10.1016/j.neuroimage.2013.09.033
Van Overwalle, F., Manto, M., Cattaneo, Z., Clausi, S., Ferrari, C., Gabrieli, J. D. E.,
et al. (2020). Consensus paper: cerebellum and social cognition. Cerebellum 19,
833–868. doi: 10.1007/s12311-020- 01155-1
Voogd, J. (2011). Cerebellar zones: a personal history. Cerebellum 10, 334–350.
doi: 10.1007/s12311-010- 0221-6
Voogd, J. (2014). What we do not know about cerebellar systems neuroscience.
Front. Syst. Neurosci. 8:227. doi: 10.3389/fnsys.2014.00227
Voogd, J., and Ruigrok, T. J. H. (2004). The organization of the corticonuclear and
olivocerebellar climbing fiber projections to the rat cerebellar vermis: the congruence
of projection zones and the zebrin pattern. J. Neurocytol. 33, 5–21. doi: 10.1023/B:
NEUR.0000029645.72074.2b
Wildgruber,D., Riecker, A., Hertrich, I., Erb, M., Grodd, W., Ethofer,T., et al. (2005).
Identification of emotional intonation evaluated by fMRI. NeuroImage 24, 1233–1241.
doi: 10.1016/j.neuroimage.2004.10.034
Wilensky, A. E., Schafe, G. E., Kristensen, M. P., and LeDoux, J. E. (2006).
Rethinking the fear circuit: the central nucleus of the amygdala is required for the
acquisition, consolidation, and expression of Pavlovian fear conditioning. J. Neurosci.
26, 12387–12396. doi: 10.1523/JNEUROSCI.4316-06.2006
Wise, N. J., Frangos, E., and Komisaruk, B. R. (2017). Brain activity unique to orgasm
in women: an fMRI analysis. J. Sex. Med. 14, 1380–1391. doi: 10.1016/j.jsxm.2017.08.
014
Witter, L., and De Zeeuw, C. I. (2015). Regional functionality of the cerebellum.
Curr. Opin. Neurobiol. 33, 150–155. doi: 10.1016/j.conb.2015.03.017
Wu, B., Blot, F. G., Wong, A. B., Osório, C., Adolfs, Y., Pasterkamp, R. J.,
et al. (2019). TRPC3 is a major contributor to functional heterogeneity of cerebellar
Purkinje cells. eLife 8:e45590. doi: 10.7554/eLife.45590
Wu, H. S., Sugihara, I., and Shinoda, Y. (1999). Projection patterns of single mossy
fibers originating from the lateral reticular nucleus in the rat cerebellar cortex and
nuclei. J. Comp. Neurol. 411, 97–118.
Xiao, J., Cerminara, N. L., Kotsurovskyy, Y., Aoki, H., Burroughs, A., Wise, A. K.,
et al. (2014). Systematic regional variations in Purkinje cell spiking patterns. PLoS One
9:e105633. doi: 10.1371/journal.pone.0105633
Zheng, J., Yang, Q., Makris, N., Huang, K., Liang, J., Ye, C., et al. (2022). Three-
dimensional digital reconstruction of the cerebellar cortex: lobule thickness, surface
area measurements, and layer architecture. Cerebellum 22, 249–260. doi: 10.1007/
s12311-022- 01390-8
Zhou, H., Lin, Z., Voges, K., Ju, C., Gao, Z., Bosman, L. W., et al. (2014). Cerebellar
modules operate at different frequencies. eLife 3:e02536. doi: 10.7554/eLife.02536
Zhu, L., Sacco, T., Strata, P., and Sacchetti, B. (2011). Basolateral amygdala
inactivation impairs learning-induced long-term potentiation in the cerebellar cortex.
PLoS One 6:e16673. doi: 10.1371/journal.pone.0016673
Frontiers in Systems Neuroscience 21 frontiersin.org
... (1) Cognitive domain: The multiscale analysis of circuit operations supports the cerebellar involvement in the main cognitive domains of CIAS, attention, learning, decision-making (D'Angelo and Casali, 2012), akin to the involvement of cerebellum in cognitive, emotional, and behavioral control (Ciapponi et al., 2023;Supplementary Figure S1). (2) Processing speed: The cerebellum contributes substantially to mechanisms of CIAS, like processing speed, by allowing mental processing to move from controlled to automatic mode (Wong et al., 2021). ...
... On the other, the hypothesis of the involvement of the cerebellum in SZ is gaining credit. The cerebellum is involved in multiple aspects of cognitive processing (Schmahmann, 2016;D'Angelo, 2018;D'Angelo, 2019;Schmahmann, 2019;Jacobi et al., 2021;Ciapponi et al., 2023;Nguyen et al., 2023) and is connected functionally and anatomically to brain regions that are core domains of CIAS, such as PFC, basal ganglia, and VTA. Cerebellar alterations can either be primary (genetic and epi-genetic) or secondary (compensatory) in origin and emerge on different scales (Figures 3-6): ...
... The main one is whether the universal cerebellar transform (Ito, 2008;D'Angelo and Casali, 2012) is altered and how, in turn, this impacts cognitive performance in SZ. Related to this is the differentiation of activity and neuromodulation among specific cerebellar regions (Ciapponi et al., 2023). This is particularly pertinent to the posterior lobules, which hold a pivotal role in cognitive processing. ...
Article
Full-text available
Schizophrenia (SZ) is a complex neuropsychiatric disorder associated with severe cognitive dysfunction. Although research has mainly focused on forebrain abnormalities, emerging results support the involvement of the cerebellum in SZ physiopathology, particularly in Cognitive Impairment Associated with SZ (CIAS). Besides its role in motor learning and control, the cerebellum is implicated in cognition and emotion. Recent research suggests that structural and functional changes in the cerebellum are linked to deficits in various cognitive domains including attention, working memory, and decision-making. Moreover, cerebellar dysfunction is related to altered cerebellar circuit activities and connectivity with brain regions associated with cognitive processing. This review delves into the role of the cerebellum in CIAS. We initially consider the major forebrain alterations in CIAS, addressing impairments in neurotransmitter systems, synaptic plasticity, and connectivity. We then focus on recent findings showing that several mechanisms are also altered in the cerebellum and that cerebellar communication with the forebrain is impaired. This evidence implicates the cerebellum as a key component of circuits underpinning CIAS physiopathology. Further studies addressing cerebellar involvement in SZ and CIAS are warranted and might open new perspectives toward understanding the physiopathology and effective treatment of these disorders.
... Based on these and other works, there is now an increased recognition for cerebellar involvement in various higher-order nonmotor functions (for reviews, see Adamaszek et al. 2017, Clausi et al. 2022, Heck et al. 2023, Strick et al. 2009, Van Overwalle et al. 2020). This growing understanding is supported by neuroanatomical data revealing substantial connections between the cerebellum and cerebral cortex association regions, particularly with the prefrontal cortex (PFC) and limbic system, shedding light on the observed cognitive deficits in individuals with cerebellar damage (for reviews, see Bodranghien et al. 2016, Ciapponi et al. 2023, Koziol et al. 2014. ...
Article
Full-text available
The cerebellum has a well-established role in controlling motor functions, including coordination, posture, and the learning of skilled movements. The mechanisms for how it carries out motor behavior remain under intense investigation. Interestingly though, in recent years the mechanisms of cerebellar function have faced additional scrutiny since nonmotor behaviors may also be controlled by the cerebellum. With such complexity arising, there is now a pressing need to better understand how cerebellar structure, function, and behavior intersect to influence behaviors that are dynamically called upon as an animal experiences its environment. Here, we discuss recent experimental work that frames possible neural mechanisms for how the cerebellum shapes disparate behaviors and why its dysfunction is catastrophic in hereditary and acquired conditions—both motor and nonmotor. For these reasons, the cerebellum might be the ideal therapeutic target.
... Nuclear plasticity should become critical in more complex associative learning protocols involving memory consolidation and saving [35,36,101,102]. Generalizing the current model to simulate more complex behaviours would require extending spatiotemporal representations to multiple scales, e.g., by introducing dendritic processing, synaptic dynamics, additional plasticity rules, and DCN and PC wiring loops [14,45,87,[103][104][105][106][107], as well as consolidation of the synaptic plasticity by sealing of the perineuronal nets [108]. ...
Article
Full-text available
According to the motor learning theory by Albus and Ito, synaptic depression at the parallel fibre to Purkinje cells synapse ( pf -PC) is the main substrate responsible for learning sensorimotor contingencies under climbing fibre control. However, recent experimental evidence challenges this relatively monopolistic view of cerebellar learning. Bidirectional plasticity appears crucial for learning, in which different microzones can undergo opposite changes of synaptic strength (e.g. downbound microzones–more likely depression, upbound microzones—more likely potentiation), and multiple forms of plasticity have been identified, distributed over different cerebellar circuit synapses. Here, we have simulated classical eyeblink conditioning (CEBC) using an advanced spiking cerebellar model embedding downbound and upbound modules that are subject to multiple plasticity rules. Simulations indicate that synaptic plasticity regulates the cascade of precise spiking patterns spreading throughout the cerebellar cortex and cerebellar nuclei. CEBC was supported by plasticity at the pf -PC synapses as well as at the synapses of the molecular layer interneurons (MLIs), but only the combined switch-off of both sites of plasticity compromised learning significantly. By differentially engaging climbing fibre information and related forms of synaptic plasticity, both microzones contributed to generate a well-timed conditioned response, but it was the downbound module that played the major role in this process. The outcomes of our simulations closely align with the behavioural and electrophysiological phenotypes of mutant mice suffering from cell-specific mutations that affect processing of their PC and/or MLI synapses. Our data highlight that a synergy of bidirectional plasticity rules distributed across the cerebellum can facilitate finetuning of adaptive associative behaviours at a high spatiotemporal resolution.
... Patients with cerebellar disease exhibit flattening or blunting of affect, irritability, agitation, and emotional lability (Schmahmann, 2004). Both human and animal studies have demonstrated the cerebellum's involvement in modulating emotional signals and behavior through extensive cerebello-cortical and -subcortical circuits (Stoodley and Schmahmann, 2009;Strick et al., 2009;Buckner et al., 2011;Adamaszek et al., 2017;Ciapponi et al., 2023) (Fig. 1). For example, the cerebellum has been implicated in what is perhaps the best studied type of emotional learning, fear conditioning. ...
Article
Full-text available
The cerebellum, traditionally associated with motor coordination and balance, also plays a crucial role in various aspects of higher-order function and dysfunction. Emerging research has shed light on the cerebellum's broader contributions to cognitive, emotional, and reward processes. The cerebellum's influence on autonomic function further highlights its significance in regulating motivational and emotional states. Perturbations in cerebellar development and function have been implicated in various neurodevelopmental disorders, including autism spectrum disorder and attention deficit hyperactivity disorder. An increasing appreciation for neuropsychiatric symptoms that arise from cerebellar dysfunction underscores the importance of elucidating the circuit mechanisms that underlie complex interactions between the cerebellum and other brain regions for a comprehensive understanding of complex behavior. By briefly discussing new advances in mapping cerebellar function in affective, cognitive, autonomic, and social processing and reviewing the role of the cerebellum in neuropathology beyond the motor domain, this Mini-Symposium review aims to provide a broad perspective of cerebellar intersections with the limbic brain in health and disease.
Preprint
Full-text available
The motor learning theory anticipates that cerebro-cerebellar loops perform sensorimotor prediction thereby regulating motor control. This operation has been identified during action execution (AE) and observation (AO) but the causal interaction between the cerebellum and cerebral cortex remained unclear. Here we used Dynamic Causal Modelling (DCM) to study functional MRI (fMRI) data obtained during a squeeze ball task in either the AE or AO conditions. In both cases, active regions included bilateral primary visual cortex (V1), left primary motor cortex (M1), left supplementary motor and premotor cortex (SMAPMC), left cingulate cortex (CC), left superior parietal lobule (SPL), and right cerebellum (CRBL). AE and AO networks showed the same fixed effective connectivity, with pathways between V1, CRBL, SMAPMC and CC wired in a closed loop. However, the cerebellar communication towards the cerebral cortex switched from excitatory in AE to inhibitory in AO. Moreover, in AE only, signal modulation was non-linear from SMAPMC to CRBL and within the CRBL self-connection, supporting the role of the CRBL in elaborating motor plans received from SMAPMC. Thus, the need for motor planning and the presence of a sensorimotor feedback in AE discriminate the modality of forward control operated by the CRBL on SMAPMC. While the underlying circuit mechanisms remain to be determined, these results reveal that the CRBL differentially controls the excitatory/inhibitory dynamics of inter-regional effective connectivity depending on its functional engagement, opening new prospective for the design of artificial sensorimotor controllers and for the investigation of neurological diseases.
Article
Full-text available
Cerebellar networks have traditionally been linked to sensorimotor control. However, a large body of evidence suggests that cerebellar functions extend to non-motor realms, such as fear-based emotional processing and that these functions are supported by interactions with a wide range of brain structures. Research related to the cerebellar contributions to emotional processing has focussed primarily on the use of well-constrained conditioning paradigms in both human and non-human subjects. From these studies, cerebellar circuits appear to be critically involved in both conditioned and unconditioned responses to threatening stimuli in addition to encoding and storage of fear memory. It has been hypothesised that the computational mechanism underlying this contribution may involve internal models, where errors between actual and expected outcomes are computed within the circuitry of the cerebellum. From a clinical perspective, cerebellar abnormalities have been consistently linked to neurodevelopmental disorders, including autism. Importantly, atypical adaptive behaviour and heightened anxiety are also common amongst autistic individuals. In this review, we provide an overview of the current anatomical, physiological and theoretical understanding of cerebellar contributions to fear-based emotional processing to foster further insights into the neural circuitry underlying emotional dysregulation observed in people with autism.
Article
Full-text available
Nascent cerebellar neuropsychiatry is rewriting complex human relations. In daily practice this sheds light on subsets of therapy-resistant patients, who feel hampered by a lack of skills in predictively presensing the trajectories to where especially their interpersonal appropriations might end up. Humans affected by "dysmetric" social phobia often lead minimal lives, strongly dislike exposition, suffer fatiguability also from immune dysfunctions, and anhedonia. In social dysmetria especially on the cerebellar cortex`s both lateralmost Crus-II (Van Overwalle) seem damaged, the left Crus-I may add agentic failure (Guell). The cerebellum, and the basal ganglia together discipline the cortex through parallel closed circuits partly interconnected through the thalamus. "Patho-trajectories" of the cere-bellum in daily care emerge through also neurological ataxiology and MRI-imaging. A uniquely pontine axonal diffusive damage to an executive-control-loop as core of "p-factor", the prime broad risk factor, points to an immune-arterial hotspot at the root of cerebellar arterial supply, as supported by the pediatric mostly inflammatory somatic "d-factor". While localizations refer back to his "phrenology", F. G. Gall's contention, that skulls, plausibly via mycobacteria, were relevant, are rehabilitated by calvario-menigeal vessels, which added the skull to the new brain logistics. Mast cells are uniquely positioned to cause superficial and deep cortical pathologies, and, enclosed in its subtentorial posterior fossa, also the cerebellum is exposed to various, often intracellular, smoldering originally dental or ORL-infections exemplifying non-neural psychiatric etiologies.
Conference Paper
Full-text available
Nascent cerebellar neuropsychiatry is rewriting complex human relations. In daily practice this sheds light on subsets of therapy-resistant patients, who feel hampered by a lack of skills in predictively presensing the trajectories to where especially their interpersonal appropriations might end up. Humans affected by "dysmetric" social phobia often lead minimal lives, strongly dislike exposition, suffer fatiguability also from immune dysfunctions, and anhedonia. In social dysmetria especially on the cerebellar cortex`s both lateralmost Crus-II (Van Overwalle) seem damaged, the left Crus-I may add agentic failure (Guell). The cerebellum, and the basal ganglia together discipline the cortex through parallel closed circuits partly interconnected through the thalamus. "Patho-trajectories" of the cere-bellum in daily care emerge through also neurological ataxiology and MRI-imaging. A uniquely pontine axonal diffusive damage to an executive-control-loop as core of "p-factor", the prime broad risk factor, points to an immune-arterial hotspot at the root of cerebellar arterial supply, as supported by the pediatric mostly inflammatory somatic "d-factor". While localizations refer back to his "phrenology", F. G. Gall's contention, that skulls, plausibly via mycobacteria, were relevant, are rehabilitated by calvario-menigeal vessels, which added the skull to the new brain logistics. Mast cells are uniquely positioned to cause superficial and deep cortical pathologies, and, enclosed in its subtentorial posterior fossa, also the cerebellum is exposed to various, often intracellular, smoldering originally dental or ORL-infections exemplifying non-neural psychiatric etiologies.
Preprint
Nascent Harvardian Italo-Belgian cerebellar neuropsychiatry is rewriting complex human relations. In daily practice this sheds light on subsets of therapy-resistant patients, who feel hampered by a lack of skills in predictably pre-sensing the trajectories to where especially their interpersonal appropriations might end up. Humans affected by "dysmetric" social phobia often lead minimal lives, strongly dislike exposition, suffer fatigability also from immune dysfunctions, and anhedonia. In social dysmetria especially on the cerebellar cortex`s both lateral-most Crus-II (Van Overwalle) seem damaged, the left Crus-I may add agentic failure (Guell). The cerebellum, and the basal ganglia together discipline the cortex through parallel closed circuits partly interconnected through the thalamus. "Patho-trajectories" of the cere-bellum in daily care emerge through also neurological ataxiology and MRI-imaging. A uniquely pontine axonal diffusive damage to an executive-control-loop as core of "p-factor", the prime broad risk factor, points to an immune-arterial hotspot at the root of cerebellar arterial supply, as supported by the pediatric mostly inflammatory somatic "d-factor". While localizations refer back to his "phrenology", F. G. Gall's contention, that skulls, plausibly via mycobacteria, were relevant, are rehabilitated by calvario-meningeal vessels, which added the skull to the new brain logistics. Mast cells are uniquely positioned to cause superficial and deep cortical pathologies, and, enclosed in its subtentorial posterior fossa, also the cerebellum is exposed to various, often intracellular, smoldering originally dental or ORL-infections exemplifying non-neural psychiatric etiologies. - In memoriam Riccardo Delle Chiaie
Article
Full-text available
The cerebellum is a key player in many brain functions and a major topic of neuroscience research. However, the cerebellar nuclei (CN), the main output structures of the cerebellum, are often overlooked. This neglect is because research on the cerebellum typically focuses on the cortex and tends to treat the CN as relatively simple output nuclei conveying an inverted signal from the cerebellar cortex to the rest of the brain. In this review, by adopting a nucleocentric perspective we aim to rectify this impression. First, we describe CN anatomy and modularity and comprehensively integrate CN architecture with its highly organized but complex afferent and efferent connectivity. This is followed by a novel classification of the specific neuronal classes the CN comprise and speculate on the implications of CN structure and physiology for our understanding of adult cerebellar function. Based on this thorough review of the adult literature we provide a comprehensive overview of CN embryonic development and, by comparing cerebellar structures in various chordate clades, propose an interpretation of CN evolution. Despite their critical importance in cerebellar function, from a clinical perspective intriguingly few, if any, neurological disorders appear to primarily affect the CN. To highlight this curious anomaly, and encourage future nucleocentric interpretations, we build on our review to provide a brief overview of the various syndromes in which the CN are currently implicated. Finally, we summarize the specific perspectives that a nucleocentric view of the cerebellum brings, move major outstanding issues in CN biology to the limelight, and provide a roadmap to the key questions that need to be answered in order to create a comprehensive integrated model of CN structure, function, development, and evolution.
Article
Full-text available
Fear and anxiety are brain states that evolved to mediate defensive responses to threats. The defense reaction includes multiple interacting behavioral, autonomic and endocrine adjustments, but their integrative nature is poorly understood. In particular, although threat has been associated with various cardiac changes, there is no clear consensus regarding the relevance of these changes for the integrated defense reaction. Here we identify rapid microstates that are associated with specific behaviors and heart rate dynamics, which are affected by long-lasting macrostates and reflect context-dependent threat levels. In addition, we demonstrate that one of the most commonly used defensive behavioral responses—freezing as measured by immobility—is part of an integrated cardio-behavioral microstate mediated by Chx10⁺ neurons in the periaqueductal gray. Our framework for systematic integration of cardiac and behavioral readouts presents the basis for a better understanding of complex neural defensive states and their associated systemic functions.
Article
Full-text available
Neurotransmitter content is deemed the most basic defining criterion for neuronal classes, contrasting with the intercellular heterogeneity of many other molecular and functional features. Here we show, in the adult mouse brain, that neurotransmitter content variegation within a neuronal class is a component of its functional heterogeneity. Golgi cells (GoCs), the well-defined class of cerebellar interneurons inhibiting granule cells (GrCs), contain cytosolic glycine, accumulated by the neuronal transporter GlyT2, and GABA in various proportions. By performing acute manipulations of cytosolic GABA and glycine supply, we find that competition of glycine with GABA reduces the charge of IPSC evoked in GrCs and, more specifically, the amplitude of a slow component of the IPSC decay. We then pair GrCs recordings with optogenetic stimulations of single GoCs, which preserve the intracellular transmitter mixed content. We show that the strength and decay kinetics of GrCs IPSCs, which are entirely mediated by GABA A receptors, are negatively correlated to the presynaptic expression of GlyT2 by GoCs. We isolate a slow spillover component of GrCs inhibition that is also affected by the expression of GlyT2, leading to a 56% decrease in relative charge. Our results support the hypothesis that presynaptic loading of glycine negatively impacts the GABAergic transmission in mixed interneurons, most likely through a competition for vesicular filling. We discuss how the heterogeneity of neurotransmitter supply within mixed interneurons like the GoC class may provide a presynaptic mechanism to tune the gain of microcircuits such as the granular layer, thereby expanding the realm of their possible dynamic behaviors.
Article
Full-text available
The mouse brain contains a rich diversity of inhibitory neuron types that have been characterized by their patterns of gene expression. However, it is still unclear how these cell types are distributed across the mouse brain. We developed a computational method to estimate the densities of different inhibitory neuron types across the mouse brain. Our method allows the unbiased integration of diverse and disparate datasets into one framework to predict inhibitory neuron densities for uncharted brain regions. We constrained our estimates based on previously computed brain-wide neuron densities, gene expression data from in situ hybridization image stacks together with a wide range of values reported in the literature. Using constrained optimization, we derived coherent estimates of cell densities for the different inhibitory neuron types. We estimate that 20.3% of all neurons in the mouse brain are inhibitory. Among all inhibitory neurons, 18% predominantly express parvalbumin (PV), 16% express somatostatin (SST), 3% express vasoactive intestinal peptide (VIP), and the remainder 63% belong to the residual GABAergic population. We find that our density estimations improve as more literature values are integrated. Our pipeline is extensible, allowing new cell types or data to be integrated as they become available. The data, algorithms, software, and results of our pipeline are publicly available and update the Blue Brain Cell Atlas. This work therefore leverages the research community to collectively converge on the numbers of each cell type in each brain region.
Article
Full-text available
The cerebellar network is renowned for its regular architecture that has inspired foundational computational theories. However, the relationship between circuit structure, function and dynamics remains elusive. To tackle the issue, we developed an advanced computational modeling framework that allows us to reconstruct and simulate the structure and function of the mouse cerebellar cortex using morphologically realistic multi-compartmental neuron models. The cerebellar connectome is generated through appropriate connection rules, unifying a collection of scattered experimental data into a coherent construct and providing a new model-based ground-truth about circuit organization. Naturalistic background and sensory-burst stimulation are used for functional validation against recordings in vivo, monitoring the impact of cellular mechanisms on signal propagation, inhibitory control, and long-term synaptic plasticity. Our simulations show how mossy fibers entrain the local neuronal microcircuit, boosting the formation of columns of activity travelling from the granular to the molecular layer providing a new resource for the investigation of local microcircuit computation and of the neural correlates of behavior. A computational modelling framework is developed, and used to reconstruct and simulate the entire cerebellar cortex, showing how mossy fibers entrain the neuronal microcircuit. This sheds light on local computation and neural correlates of behaviors
Article
Full-text available
It is common for animals to use self-generated movements to actively sense the surrounding environment. For instance, rodents rhythmically move their whiskers to explore the space close to their body. The mouse whisker system has become a standard model for studying active sensing and sensorimotor integration through feedback loops. In this work, we developed a bioinspired spiking neural network model of the sensorimotor peripheral whisker system, modeling trigeminal ganglion, trigeminal nuclei, facial nuclei, and central pattern generator neuronal populations. This network was embedded in a virtual mouse robot, exploiting the Human Brain Project's Neurorobotics Platform, a simulation platform offering a virtual environment to develop and test robots driven by brain-inspired controllers. Eventually, the peripheral whisker system was adequately connected to an adaptive cerebellar network controller. The whole system was able to drive active whisking with learning capability, matching neural correlates of behavior experimentally recorded in mice.
Article
Full-text available
To understand how the cerebellar cortex transforms mossy fiber (MF) inputs into Purkinje cell (PC) outputs, it is vital to delineate the elements of this circuit. Candelabrum cells (CCs) are enigmatic interneurons of the cerebellar cortex that have been identified based on their morphology, but their electrophysiological properties, synaptic connections and function remain unknown. Here, we clarify these properties using electrophysiology, single-nucleus RNA sequencing, in situ hybridization and serial electron microscopy in mice. We find that CCs are the most abundant PC layer interneuron. They are GABAergic, molecularly distinct and present in all cerebellar lobules. Their high resistance renders CC firing highly sensitive to synaptic inputs. CCs are excited by MFs and granule cells and are strongly inhibited by PCs. CCs in turn primarily inhibit molecular layer interneurons, which leads to PC disinhibition. Thus, inputs, outputs and local signals converge onto CCs to allow them to assume a unique role in controlling cerebellar output.
Article
Full-text available
The cerebellum is emerging as a powerful regulator of cognitive and affective processing and memory in both humans and animals and has been implicated in affective disorders. How the cerebellum supports affective function remains poorly understood. The short-latency (just a few milliseconds) functional connections that were identified between the cerebellum and amygdala—a structure crucial for the processing of emotion and valence—more than four decades ago raise the exciting, yet untested, possibility that a cerebellum-amygdala pathway communicates information important for emotion. The major hurdle in rigorously testing this possibility is the lack of knowledge about the anatomy and functional connectivity of this pathway. Our initial anatomical tracing studies in mice excluded the existence of a direct monosynaptic connection between the cerebellum and amygdala. Using transneuronal tracing techniques, we have identified a novel disynaptic circuit between the cerebellar output nuclei and the basolateral amygdala. This circuit recruits the understudied intralaminar thalamus as a node. Using ex vivo optophysiology and super-resolution microscopy, we provide the first evidence for the functionality of the pathway, thus offering a missing mechanistic link between the cerebellum and amygdala. This discovery provides a connectivity blueprint between the cerebellum and a key structure of the limbic system. As such, it is the requisite first step toward obtaining new knowledge about cerebellar function in emotion, thus fundamentally advancing understanding of the neurobiology of emotion, which is perturbed in mental and autism spectrum disorders.
Article
Full-text available
Cellular diversification is critical for specialized functions of the brain including learning and memory1. Single-cell RNA sequencing facilitates transcriptomic profiling of distinct major types of neuron2–4, but the divergence of transcriptomic profiles within a neuronal population and their link to function remain poorly understood. Here we isolate nuclei tagged5 in specific cell types followed by single-nucleus RNA sequencing to profile Purkinje neurons and map their responses to motor activity and learning. We find that two major subpopulations of Purkinje neurons, identified by expression of the genes Aldoc and Plcb4, bear distinct transcriptomic features. Plcb4+, but not Aldoc+, Purkinje neurons exhibit robust plasticity of gene expression in mice subjected to sensorimotor and learning experience. In vivo calcium imaging and optogenetic perturbation reveal that Plcb4+ Purkinje neurons have a crucial role in associative learning. Integrating single-nucleus RNA sequencing datasets with weighted gene co-expression network analysis uncovers a learning gene module that includes components of FGFR2 signalling in Plcb4+ Purkinje neurons. Knockout of Fgfr2 in Plcb4+ Purkinje neurons in mice using CRISPR disrupts motor learning. Our findings define how diversification of Purkinje neurons is linked to their responses in motor learning and provide a foundation for understanding their differential vulnerability to neurological disorders. Subpopulations of Purkinje neurons display distinct transcriptomic responses and functions in associative learning.
Article
Addressing the multiscale organization of the brain, which is fundamental to the dynamic repertoire of the organ, remains challenging. In principle, it should be possible to model neurons and synapses in detail and then connect them into large neuronal assemblies to explain the relationship between microscopic phenomena, large-scale brain functions, and behavior. It is more difficult to infer neuronal functions from ensemble measurements such as those currently obtained with brain activity recordings. In this article we consider theories and strategies for combining bottom-up models, generated from principles of neuronal biophysics, with top-down models based on ensemble representations of network activity and on functional principles. These integrative approaches are hoped to provide effective multiscale simulations in virtual brains and neurorobots, and pave the way to future applications in medicine and information technologies.