ArticlePDF Available

Abstract and Figures

The sympathetic nervous system plays a critical role in regulating many autonomic functions, including cardiac rhythm. The postganglionic neurons in the sympathetic chain ganglia are essential components that relay sympathetic signals to target tissues and disruption of their activity leads to poor health outcomes. Despite this importance, the neurocircuitry within sympathetic ganglia is poorly understood. Canonically, postganglionic sympathetic neurons are thought to simply be activated by monosynaptic inputs from preganglionic cholinergic neurons of the intermediolateral cell columns of the spinal cord. Early electrophysiological studies of sympathetic ganglia where the peripheral nerve trunks were electrically stimulated identified excitatory cholinergic synaptic events in addition to retrograde action potentials, leading some to speculate that excitatory collateral projections are present. However, this seemed unlikely since sympathetic postganglionic neurons were known to synthesize and release norepinephrine and expression of dual neurochemical phenotypes had not been well recognized. In vitro studies clearly established the capacity of cultured sympathetic neurons to express and release acetylcholine and norepinephrine throughout development and even in pathophysiological conditions. Given this insight, we believe that the canonical view of ganglionic transmission needs to be reevaluated and may provide a mechanistic understanding of autonomic imbalance in disease. Further studies likely will require genetic models manipulating neurochemical phenotypes within sympathetic ganglia to resolve the function of cholinergic collateral projections between postganglionic neurons. In this perspective article, we will discuss the evidence for collateral projections in sympathetic ganglia, determine if current laboratory techniques could address these questions, and discuss potential obstacles and caveats.
This content is subject to copyright.
PERSPECTIVE
published: 10 February 2022
doi: 10.3389/fcvm.2022.842656
Frontiers in Cardiovascular Medicine | www.frontiersin.org 1February 2022 | Volume 9 | Article 842656
Edited by:
Tania Zaglia,
University of Padova, Italy
Reviewed by:
Nazareno Paolocci,
Johns Hopkins University,
United States
Matthew W. Kay,
George Washington University,
United States
Michael Rubart,
Indiana University Bloomington,
United States
*Correspondence:
Beth A. Habecker
habecker@ohsu.edu
Specialty section:
This article was submitted to
Hypertension,
a section of the journal
Frontiers in Cardiovascular Medicine
Received: 24 December 2021
Accepted: 19 January 2022
Published: 10 February 2022
Citation:
Clyburn C, Andresen MC, Ingram SL
and Habecker BA (2022) Untangling
Peripheral Sympathetic Neurocircuits.
Front. Cardiovasc. Med. 9:842656.
doi: 10.3389/fcvm.2022.842656
Untangling Peripheral Sympathetic
Neurocircuits
Courtney Clyburn 1, Michael C. Andresen 1, Susan L. Ingram 2and Beth A. Habecker 1
*
1Department of Chemical Physiology and Biochemistry, Oregon Health and Science University, Portland, OR, United States,
2Department of Neurological Surgery, Oregon Health and Science University, Portland, OR, United States
The sympathetic nervous system plays a critical role in regulating many autonomic
functions, including cardiac rhythm. The postganglionic neurons in the sympathetic
chain ganglia are essential components that relay sympathetic signals to target tissues
and disruption of their activity leads to poor health outcomes. Despite this importance,
the neurocircuitry within sympathetic ganglia is poorly understood. Canonically,
postganglionic sympathetic neurons are thought to simply be activated by monosynaptic
inputs from preganglionic cholinergic neurons of the intermediolateral cell columns of the
spinal cord. Early electrophysiological studies of sympathetic ganglia where the peripheral
nerve trunks were electrically stimulated identified excitatory cholinergic synaptic events
in addition to retrograde action potentials, leading some to speculate that excitatory
collateral projections are present. However, this seemed unlikely since sympathetic
postganglionic neurons were known to synthesize and release norepinephrine and
expression of dual neurochemical phenotypes had not been well recognized. In vitro
studies clearly established the capacity of cultured sympathetic neurons to express
and release acetylcholine and norepinephrine throughout development and even in
pathophysiological conditions. Given this insight, we believe that the canonical view
of ganglionic transmission needs to be reevaluated and may provide a mechanistic
understanding of autonomic imbalance in disease. Further studies likely will require
genetic models manipulating neurochemical phenotypes within sympathetic ganglia to
resolve the function of cholinergic collateral projections between postganglionic neurons.
In this perspective article, we will discuss the evidence for collateral projections in
sympathetic ganglia, determine if current laboratory techniques could address these
questions, and discuss potential obstacles and caveats.
Keywords: sympathetic ganglia, neurocircuits, synaptic inputs, co-transmission, collaterals
INTRODUCTION
Technical advances have long driven new insights into the mechanistic basis of neurophysiology.
From the identification of the action potential in the nineteenth century (1) to the work of
Joseph Erlanger and Herbert Gasser who revolutionized neurophysiological research by developing
sensitive oscilloscopes that allowed for the visualization and analysis of nerve impulses previously
below the threshold of detection (2,3). With these technological advancements, researchers set
out to decode the complex autonomic signals that regulate visceral functions, including cardiac
Clyburn et al. Neurocircuitry of the Sympathetic Ganglia
activity (46). Several important hypotheses arose from this
work, but were untested due to experimental limitations that
existed at the time. These questions include whether or
not collateral projections are present between postganglionic
neurons in sympathetic ganglia and what neurotransmitters
may be involved in sympathetic signaling pathways. Since these
first studies, significant advancements in neurophysiological
equipment and scientific methods have allowed for substantial
progress in understanding the neurophysiology of peripheral
sympathetic activity (711), but the neurocircuitry within
sympathetic ganglia remains poorly understood. The purpose
of this article is to review the primary data that support
these untested hypotheses, determine if current technology and
experimental methods can be used to answer these important
questions, and discuss any potential obstacles that may arise.
Canonical Neurocircuitry of Cervical
Sympathetic Ganglia
Several well-written review articles are available that thoroughly
summarize our current understanding of the sympathetic
neurocircuitry that regulates visceral functions and cardiac
activity (1216). To briefly summarize, sympathetic signals
originate in the hypothalamus and brainstem and travel to the
preganglionic neurons in the intermediolateral cell column of
the spinal cord. These preganglionic neurons send cholinergic
projections to postganglionic neurons in the cervical sympathetic
chain ganglia. These neurons in the sympathetic ganglia then
send noradrenergic projections through the postganglionic nerve
trunk to innervate the target tissue. This general organization
is conserved across species. However, there is significant
biological variability in the anatomy and physiology of the
cervical sympathetic ganglia which clouds our understanding of
peripheral sympathetic neurocircuits. The cervical sympathetic
chain includes the superior, middle, and inferior cervical
ganglia and exhibits significant anatomical variability between
species, between individuals, and even between the left and
right ganglia. In approximately 80% of humans, for example,
the inferior cervical ganglion is fused with the first thoracic
ganglion to form the cervicothoracic (commonly known as
stellate) ganglion. Additionally, the middle cervical ganglion
is completely absent in 20% of the population (1720). In
considering the variability in the sympathetic nervous system
in humans and animal models as well as the murkiness
that surrounds the functional organization of the cervical
sympathetic ganglia, we will consider data from all cervical
sympathetic ganglia interchangeably as we discuss peripheral
sympathetic neurocircuitry.
Evidence for Collateral Projections in
Sympathetic Ganglia
In addition to the ambiguity surrounding the anatomy of
the sympathetic ganglia, the functional organization of the
sympathetic neurocircuitry within these ganglia remains poorly
defined. Current literature usually describes and illustrates the
neurocircuitry within cervical sympathetic ganglia as simple
monosynaptic connections from cholinergic preganglionic
neurons to the noradrenergic postganglionic neurons that
innervate visceral targets and the heart (21,22). However,
Erulkar and Woodward (23) made intracellular recordings
of postganglionic sympathetic neurons from rabbit superior
cervical ganglia (SCG) in situ which suggest this neurocircuitry
may be more complicated. In these early experiments, they found
that stimulation of the external carotid nerve produced single,
short latency spikes followed by long-lasting hyperpolarizations
(presumably action potentials) in some neurons. In these
experiments, electrical stimulation of the postganglionic nerve
trunk evoked antidromic, or retrograde, action potentials that
traveled up the axon to the cell body in the SCG. Surprisingly,
the majority of neurons exhibited an early spike followed by a
long-lasting depolarization. This depolarization suggested that
stimulation of the postganglionic nerve trunk elicited excitatory
synaptic activity within the SCG. Erulkar and Woodward
proposed several hypotheses that would account for this
phenomenon, but state that the simplest explanation would be
that recurrent excitatory collateral fibers from postganglionic
axons project to neighboring cells within the SCG (Figure 1).
Upon stimulation of the preganglionic nerve trunk, Erulkar and
Woodward also observed that a significant proportion of SCG
neurons exhibited multi-spike responses which were uncovered
at increasing stimulation intensities. While they acknowledge
these results may be caused by different populations of
preganglionic fibers with different thresholds and conduction
velocities that converge onto the same cell, this explanation
seems highly unlikely. Instead, they note that the multi-spike
responses are consistent with the collateral hypothesis and that
the late-arriving impulses may have traveled through another
indirect pathway.
Excitatory synaptic events in postganglionic sympathetic
neurons following electrical stimulation of the postganglionic
nerve trunk have also been observed in the cat, rat and guinea
pig (2426). In 1970, Perri et al. extended the observations by
Erulkar and Woodward by investigating the neurotransmitter
that may be involved in these synaptic events. Curare, a
nicotinic acetylcholine (ACh) receptor antagonist, applied to
the ganglionic preparation abolished these synaptic events,
indicating they were likely cholinergic in nature (25). If
these excitatory events originated from postganglionic collateral
projections, this would mean that these sympathetic neurons
express ACh and norepinephrine (NE). Although we now know
that postganglionic sympathetic neurons and other neurons
can express a dual neurochemical phenotype (2731), at the
time this was thought to be unlikely. Thus, researchers sought
alternative explanations. These included the possibility that
errant preganglionic fibers may be running up the postganglionic
nerve trunk and forming excitatory synaptic connections with
postganglionic neurons (Figure 1) (25). A second possibility
raised was that sensory afferent neurons were the source of
synaptic excitation. Sympathetic ganglia have long been linked
to sensory afferents of the cardiac region of the thoracic viscera
(6) but on the bases of such anatomical findings broadly
dismissed as only passing through the ganglia on their way
to the spinal cord. Thus, sketches commonly depict that a
portion of these afferent paths include sympathetic ganglia
Frontiers in Cardiovascular Medicine | www.frontiersin.org 2February 2022 | Volume 9 | Article 842656
Clyburn et al. Neurocircuitry of the Sympathetic Ganglia
FIGURE 1 | Schematic diagram illustrating the errant preganglionic fiber and collateral hypotheses. In the preganglionic hypothesis (left), mis-routed cholinergic (black)
preganglionic fibers travel back up the efferent nerves through which sympathetic axons travel to the target tissue. In the collateral hypothesis (right) postganglionic
sympathetic neurons send noradrenergic projections (red) to the target tissue and cholinergic projections (black) to other sympathetic neurons in the sympathetic
chain ganglia.
(e.g., cervical and stellate) (32). However, functional studies
related to cardiovascular regulation, for example, indicate that
activation of cardiac sensory afferents traveling via the stellate
ganglia act to increase blood pressure by augmenting sympathetic
motor neuron activity directed to the systemic vasculature and
such responses are abolished by severing ganglion connections
to the spinal cord (33). Dye tracing and electrical activity
recordings indicate that such afferent neurons arise from the
dorsal root ganglia and traverse peripheral sympathetic ganglia
to their visceral targets (34). Although it seemed unlikely at
the time, our current understanding of dual neurochemical
phenotypes in sympathetic neurons has persuaded us that
further studies are needed to investigate the potential presence
of cholinergic collateral projections between postganglionic
neurons in sympathetic ganglia.
Evidence for Dual Neurochemical
Phenotypes in Sympathetic Neurons
In 1935, Henry Dale put forth the “one neuron, one
neurotransmitter” hypothesis which was later termed as “Dale’s
Principle” (35,36) and stated that one neuron was only
capable of releasing one neurotransmitter at any one time
[reviewed by: (37)]. By the mid-1970s, evidence disputing
Dale’s Principle began to emerge when Jan and colleagues
made intracellular recordings of sympathetic frog neurons and
observed slow synaptic potentials which were mediated by the
peptide, luteinizing hormone-releasing hormone, in addition to
the canonical cholinergic transmission (38,39). Since these initial
studies, co-transmission of several neuromodulators, including
ATP and neuropeptide Y (NPY) (4044), as well as the co-
transmission of fast primary neurotransmitters have been well
established throughout the central and peripheral nervous
system (4547).
Some of the earliest evidence supporting co-expression
of primary neurotransmitters comes from cardiac myocyte—
sympathetic neuron co-cultures (48). In 1976, Furshpan
et al. used electrophysiological recordings of sympathetic SCG
principal neurons cultured with cardiac myocytes from newborn
rats and described a population of neurons that secreted
ACh and NE (4851). This work sparked a series of studies
to determine if sympathetic neurons produced ACh in vivo
and to identify the differentiation factors that determine
the fates of sympathetic neurons (52). Several “cholinergic
differentiation factors” are now known to induce ACh expression
in noradrenergic sympathetic neurons including leukemia
inhibitor factor (LIF), ciliary neurotrophic factor (CNTF),
cardiotrophin-1 (CT-1), neurotrophin-3 (NT-3) and glial cell
line-derived neurotrophic factor (GDNF) (5361). While the
cholinergic transdifferentiation of postganglionic sympathetic
neurons in vivo was first thought to be confined to developmental
periods, more recent studies have shown that inflammatory
cytokines can induce cholinergic transdifferentiation in cardiac
Frontiers in Cardiovascular Medicine | www.frontiersin.org 3February 2022 | Volume 9 | Article 842656
Clyburn et al. Neurocircuitry of the Sympathetic Ganglia
disease (30,62). Early studies have revealed key aspects
of the neurochemical phenotype of sympathetic neurons:
dual and malleable neurochemical capacity. These studies of
cultured neurons indicated that these postganglionic neurons
could form synapses between neurons and could display
adrenergic/cholinergic dual function (50). In culture, the
neurotrophin brain-derived neurotrophic factor (BDNF) rapidly
shifted sympathetic neuron release of norepinephrine to
ACh indicating the dual neurochemical phenotype of these
postganglionic neurons (31). Subsequent work also in cultured
neurons supported both segregation as well as plasticity of
functional release sites of NE and ACh (63). In light of this
evidence, it appears that the cholinergic hypothesis originally
proposed by Erulkar and Woodward in 1968 was dismissed
prematurely before fully testing the idea. Gaining a clear insight
into the neurocircuitry within sympathetic ganglia can provide
better understanding of the mechanisms underlying autonomic
imbalance in disease and may highlight novel therapeutic targets
for cardiac patients.
Challenges and Potential Caveats
The dual expression of ACh and NE in sympathetic neurons
provides support for the collateral hypothesis, but there are
still many questions that remain. When dual neurochemical
phenotypes were first described, these neurons were generally
classified into two mechanistic subtypes, co-release and co-
transmission (37,64). Co-release describes the packaging of
multiple neurotransmitters into the same synaptic vesicle
so they there are released together and co-transmission
originally described different neurotransmitters packaged into
different vesicles within the same presynaptic terminal (37,64).
Dale’s Principle was updated in 1986 to state that neurons
release the same group of neurotransmitters/peptides from
all presynaptic terminals (64,65). However, the collateral
hypothesis challenges this view because the primary data
indicates that ACh is released from the collateral terminals.
The excitatory synaptic events that are observed following
stimulation of the postganglionic nerve trunk are completely
blocked by the application of nicotinic acetylcholine antagonists
(24,25). In the collateral model, this could suggest the
biased expression of ACh and NE at different synapses within
the same neuron. Based on the updated version of Dale’s
hypothesis, we would expect that both ACh and NE would
be released from collateral synapses. As adrenergic receptors
are G-protein coupled receptors, synaptic release of NE cannot
be detected with the electrophysiological techniques used in
these previous studies (25). Future experiments could utilize
new biosensor technology to determine if NE is also released
from collateral projections (6668). If NE is not released
from collateral projections, there are several mechanisms
that have been described in the CNS that could explain
biased release of different vesicular pools. These include
differing release probabilities and altered coupling to presynaptic
Ca2+channels (69,70), “kiss-and-run” release mechanisms,
and even neurotransmitter segregation to separate axons
(64,7174). Studies in cultured sympathetic neurons have
observed segregation of the vesicular NE and ACh transporters
(VMAT2 and VAChT, respectively) in distinct varicosities,
suggesting release of each transmitter is independent and
spatially segregated (63). Considering how target-derived factors
and cytokines can induce cholinergic transdifferentiation in
postganglionic sympathetic neurons, it is conceivable that factors
in the microenvironment are involved in determining different
neurochemical identities of synaptic terminals in the ganglia and
target tissues.
Putative afferent fibers could have an important physiological
role in allowing sensory feedback to regulate the activity of
postganglionic sympathetic neurons in the event that inputs
from the central nervous system are disrupted. This sensory
feedback loop would also allow for the fast adaptation of
sympathetic neuron activity to rapidly changing stimuli (24).
However, the synaptic events that are observed following
stimulation of the postganglionic nerve trunk are mediated
by acetylcholine and, therefore, unlikely to be mediated by
sensory fibers, which utilize glutamatergic signaling. The
physiological consequence of putative cholinergic collateral
projections is less obvious. Collateral projections between
postganglionic sympathetic neurons may also have a significant
physiological role in maintaining autonomic control of the
heart and other tissues. Excitatory collateral projections
between postganglionic sympathetic neurons would increase
the probability that these neurons fire synchronously and may
amplify preganglionic signals from the spinal cord. Given the
importance of neural timing for cardiopulmonary integration,
the synchronization of sympathetic inputs could potentially
play a critical role in maintaining appropriate cardiac activity.
Furthermore, disruption of this synchronization following
injury or disease could exacerbate autonomic imbalance.
Early studies hypothesized that single spikes in recordings
of cardiac sympathetic nerve activity represent synchronized
activity of multiple postganglionic neurons (75), but the
potential role of collateral synapses in this synchronization is
purely speculative.
It was not possible in the past to distinguish preganglionic
cholinergic transmission from putative post-ganglionic collateral
projections, but transgenic mouse models and genetic tools
provide the means to address this issue now. Targeted deletion
of ACh production (chat gene) or release (slc18A3 gene)
from neurons expressing tyrosine hydroxylase (th gene),
dopamine beta hydroxylase (dbh gene), or norepinephrine
transporter (slc6a2 gene) would remove cholinergic transmission
selectively from post-ganglionic neurons within the ganglion.
Likewise, targeting expression of channelrhodopsin-2 or
other stimulatory optogenetic tools to noradrenergic neurons
would allow selective stimulation of those cells, coupled
with recordings of resulting synaptic activity. Combining
these approaches with the use of genetically encoded calcium
sensors would allow testing the functional impact of putative
cholinergic collateral transmission in situ. Mouse lines to
facilitate these studies, which directly address the issue of
potential cholinergic collaterals, are available from public
repositories. If cholinergic collaterals are an important aspect
of transmission by post-ganglionic sympathetic neurons then
we would expect them to be present and detectable throughout
Frontiers in Cardiovascular Medicine | www.frontiersin.org 4February 2022 | Volume 9 | Article 842656
Clyburn et al. Neurocircuitry of the Sympathetic Ganglia
the sympathetic chain, and lead to changes in transmission to
target tissues.
DISCUSSION
While some of the first studies investigating sympathetic
nerve impulses proposed that collateral projections may be
present in the sympathetic ganglia, this model was dismissed
because of our incomplete understanding of dual neurochemical
phenotypes. In hindsight, given our understanding of co-
transmission of ACh and NE in postganglionic sympathetic
neurons, it is clear that studies should be carried out to
clarify the neurocircuitry within the sympathetic ganglia.
The postganglionic neurons within sympathetic ganglia are
critical components of the neurocircuitry responsible for
relaying autonomic signals from the central nervous system
to several visceral targets, including the heart. The activity
of postganglionic sympathetic neurons is disrupted in many
pathologies and sympathetic hyperactivity elevates risk for
cardiac arrhythmias and sudden cardiac death and contributes
to the development of heart failure. A detailed understanding
sympathetic neurocircuitry is needed to understand how
peripheral sympathetic activity is disrupted in pathophysiological
conditions and to develop novel therapeutic strategies.
DATA AVAILABILITY STATEMENT
The original contributions presented in the study are included
in the article, further inquiries can be directed to the
corresponding author.
AUTHOR CONTRIBUTIONS
CC, MA, SI, and BH contributed to drafting the work and
revising it critically for important intellectual content. All authors
approved the final version of the manuscript and agree to be
accountable for all aspects of the work. All persons designated
as authors qualify for authorship, and all those who qualify for
authorship are listed.
FUNDING
This work was supported by the National Institutes of Health
Grant HL146833.
ACKNOWLEDGMENTS
The figure in this article was created with BioRender.com.
REFERENCES
1. Schuetze SM. The discovery of the action potential. Trends Neurosci. (1983)
6:164–8. doi: 10.1016/0166-2236(83)90078-4
2. Erlanger J, Bishop G, Gasser H. Experimental analysis of the simple action
potential wave in nerve by the cathode ray oscillograph. Am J Physiol Legacy
Content. (1926) 78:537–73. doi: 10.1152/ajplegacy.1926.78.3.537
3. Erlanger J, Gasser H. The action potential in fibers of slow conduction
in spinal roots and somatic nerves. Am J Physiol. (1930) 92:43–
82. doi: 10.1152/ajplegacy.1930.92.1.43
4. Adrian ED, Bronk DW, Phillips G. Discharges in mammalian sympathetic
nerves. J Physiol. (1932) 74:115–33. doi: 10.1113/jphysiol.1932.sp002
832
5. Langley JN. On axon-reflexes in the pre-ganglionic fibres of the sympathetic
system. J Physiol. (1900) 25:364–98. doi: 10.1113/jphysiol.1900.sp000803
6. Ranson S, Billingsley P. The superior cervical ganglion and the cervical
portion of the sympathetic trunk. J Comp Neurol. (1918) 29:313–
358. doi: 10.1002/cne.900290403
7. Briant LJ, Paton JF, Pickering AE, Champneys AR. Modelling the vascular
response to sympathetic postganglionic nerve activity. J Theor Biol. (2015)
371:102–16. doi: 10.1016/j.jtbi.2015.01.037
8. Carter JR. Microneurography and sympathetic nerve activity: a decade-
by-decade journey across 50 years. J Neurophysiol. (2019) 121:1183–
94. doi: 10.1152/jn.00570.2018
9. Greaney JL, Kenney WL. Measuring and quantifying skin sympathetic
nervous system activity in humans. J Neurophysiol. (2017) 118:2181–
93. doi: 10.1152/jn.00283.2017
10. Hart EC, Head GA, Carter JR, Wallin BG, May CN, Hamza SM, et al.
Recording sympathetic nerve activity in conscious humans and other
mammals: guidelines and the road to standardization. Am J Physiol Heart Circ
Physiol. (2017) 312:H1031–51. doi: 10.1152/ajpheart.00703.2016
11. Macefield VG. Sympathetic microneurography. Handb Clin Neurol. (2013)
117:353–64. doi: 10.1016/B978-0-444-53491-0.00028-6
12. Dampney RA. Central neural control of the cardiovascular
system: current perspectives. Adv Physiol Educ. (2016) 40:283–
96. doi: 10.1152/advan.00027.2016
13. Fukuda K, Kanazawa H, Aizawa Y, Ardell JL, Shivkumar K. Cardiac
innervation and sudden cardiac death. Circ Res. (2015) 116:2005–
19. doi: 10.1161/CIRCRESAHA.116.304679
14. Habecker BA, Anderson ME, Birren SJ, Fukuda K, Herring N, Hoover DB,
et al. Molecular and cellular neurocardiology: development, and cellular
and molecular adaptations to heart disease. J Physiol. (2016) 594:3853–
75. doi: 10.1113/JP271840
15. Rajendran PS, Challis RC, Fowlkes CC, Hanna P, Tompkins JD, Jordan
MC, et al. Identification of peripheral neural circuits that regulate heart
rate using optogenetic and viral vector strategies. Nat Commun. (2019)
10:1944. doi: 10.1038/s41467-019-09770-1
16. Scalco A, Moro N, Mongillo M, Zaglia T. Neurohumoral cardiac
regulation: optogenetics gets into the groove. Front Physiol. (2021)
12:726895. doi: 10.3389/fphys.2021.726895
17. Fazliogullari Z, Kilic C, Karabulut AK, Yazar F. A morphometric analysis
of the superior cervical ganglion and its surrounding structures.
Surg Radiol Anat. (2016) 38:299–302. doi: 10.1007/s00276-015-
1551-3
18. Narouze S. Ultrasound-guided stellate ganglion block: safety and efficacy. Curr
Pain Headache Rep. (2014) 18:424. doi: 10.1007/s11916-014-0424-5
19. Pather N, Partab P, Singh B, Satyapal KS. The sympathetic
contributions to the cardiac plexus. Surg Radiol Anat. (2003)
25:210–5. doi: 10.1007/s00276-003-0113-2
20. Yin Z, Yin J, Cai J, Sui T, Cao X. Neuroanatomy and clinical analysis of the
cervical sympathetic trunk and longus colli. J Biomed Res. (2015) 29:501–7.
doi: 10.7555/JBR.29.20150047
21. Jänig W. Organization of the sympathetic ner vous system:
peripheral and central aspects. Neuroimmune Biol. (2008)
7:55–85. doi: 10.1016/S1567-7443(07)00204-9
22. Vegh AMD, Duim SN, Smits AM, Poelmann RE, Ten Harkel ADJ, DeRuiter
MC, et al. Part and parcel of the cardiac autonomic nerve system: unravelling
its cellular building blocks during development. J Cardiovasc Dev Dis. (2016)
3:28. doi: 10.3390/jcdd3030028
23. Erulkar SD, Woodward JK. Intracellular recording from mammalian
superior cervical ganglion in situ. J Physiol. (1968) 199:189–
203. doi: 10.1113/jphysiol.1968.sp008648
Frontiers in Cardiovascular Medicine | www.frontiersin.org 5February 2022 | Volume 9 | Article 842656
Clyburn et al. Neurocircuitry of the Sympathetic Ganglia
24. Bosnjak ZJ, Kampine JP. Cardiac sympathetic afferent cell bodies are located
in the peripheral nervous system of the cat. Circ Res. (1989) 64:554–
62. doi: 10.1161/01.RES.64.3.554
25. Perri V, Sacchi O, Casella C. Synaptically mediated potentials elicited by
the stimulation of post-ganglionic trunks in the guinea-pig superior cervical
ganglion. Pflugers Arch. (1970) 314:55–67. doi: 10.1007/BF00587046
26. Wallis D, Watson AH, Mo N. Cardiac neurones of
autonomic ganglia. Microsc Res Tech. (1996) 35:69–
79. doi: 10.1002/(SICI)1097-0029(19960901)35:1andlt;69::AID-
JEMT6andgt;3.0.CO;2-N
27. Bosnjak ZJ, Kampine JP. Intracellular recordings from the stellate ganglion of
the cat. J Physiol. (1982) 324:273–83. doi: 10.1113/jphysiol.1982.sp014112
28. Bosnjak ZJ, Kampine JP. Electrophysiological and morphological
characterization of neurons in stellate ganglion of cats. Am J Physiol.
(1985) 248:R288–292. doi: 10.1152/ajpregu.1985.248.3.R288
29. Hoard JL, Hoover DB, Mabe AM, Blakely RD, Feng N, Paolocci N.
Cholinergic neurons of mouse intrinsic cardiac ganglia contain noradrenergic
enzymes, norepinephrine transporters, and the neurotrophin receptors
tropomyosin-related kinase A and p75. Neuroscience. (2008) 156:129–
42. doi: 10.1016/j.neuroscience.2008.06.063
30. Kanazawa H, Ieda M, Kimura K, Arai T, Kawaguchi-Manabe H, Matsuhashi
T, et al. Heart failure causes cholinergic transdifferentiation of cardiac
sympathetic nerves via gp130-signaling cytokines in rodents. J Clin Invest.
(2010) 120:408–21. doi: 10.1172/JCI39778
31. Yang B, Slonimsky JD, Birren SJ. A rapid switch in sympathetic
neurotransmitter release properties mediated by the p75 receptor. Nat
Neurosci. (2002) 5:539–45. doi: 10.1038/nn0602-853
32. Coleridge JC, Kidd C, Sharp JA. The distribution, connexions and histology
of baroreceptors in the pulmonary artery, with some observations on the
sensory innervation of the ductus arteriosus. J Physiol. (1961) 156:591–
602. doi: 10.1113/jphysiol.1961.sp006695
33. Peterson DF, Brown AM. Pressor reflexes produced by stimulation of afferent
fibers in the cardiac sympathetic nerves of the cat. Circ Res. (1971) 28:605–
10. doi: 10.1161/01.RES.28.6.605
34. Oldfield BJ, McLachlan EM. Localization of sensory neurons traversing
the stellate ganglion of the cat. J Comp Neurol. (1978) 182:915–
22. doi: 10.1002/cne.901820509
35. Dale H. Pharmacology and nerve-endings (walter ernest dixon memorial
lecture): (section of therapeutics and pharmacology). Proc R Soc Med. (1935)
28:319–32. doi: 10.1177/003591573502800330
36. Eccles JC, Fatt P, Koketsu K. Cholinergic and inhibitory synapses in a pathway
from motor-axon collaterals to motoneurones. J Physiol. (1954) 126:524–
62. doi: 10.1113/jphysiol.1954.sp005226
37. Vaaga CE, Borisovska M, Westbrook GL. Dual-transmitter neurons:
functional implications of co-release and co-transmission. Curr Opin
Neurobiol. (2014) 29:25–32. doi: 10.1016/j.conb.2014.04.010
38. Jan LY, Jan YN. Peptidergic transmission in sympathetic ganglia of the frog. J
Physiol. (1982) 327:219–46. doi: 10.1113/jphysiol.1982.sp014228
39. Jan YN, Jan LY, Kuffler SW. A peptide as a possible transmitter in
sympathetic ganglia of the frog. Proc Natl Acad Sci USA. (1979) 76:1501–
5. doi: 10.1073/pnas.76.3.1501
40. Alston EN, Parrish DC, Hasan W, Tharp K, Pahlmeyer L, Habecker BA.
Cardiac ischemia-reperfusion regulates sympathetic neuropeptide expression
through gp130-dependent and independent mechanisms. Neuropeptides.
(2011) 45:33–42. doi: 10.1016/j.npep.2010.10.002
41. Browning KN, Lees GM. Inhibitory effects of NPY on ganglionic transmission
in myenteric neurones of the guinea-pig descending colon. Neurogastroenterol
Motil. (2000) 12:33–41. doi: 10.1046/j.1365-2982.2000.00178.x
42. Burnstock G. Physiology and pathophysiology of
purinergic neurotransmission. Physiol Rev. (2007) 87:659–
797. doi: 10.1152/physrev.00043.2006
43. Pernow J, Schwieler J, Kahan T, Hjemdahl P, Oberle J, Wallin
BG, et al. Influence of sympathetic discharge pattern on
norepinephrine and neuropeptide Y release. Am J Physiol. (1989)
257:H866–872. doi: 10.1152/ajpheart.1989.257.3.H866
44. Verhaeghe RH, Vanhoutte PM, Shepherd JT. Inhibition of sympathetic
neurotransmission in canine blood vessels by adenosine and adenine
nucleotides. Circ Res. (1977) 40:208–15. doi: 10.1161/01.RES.40.2.208
45. Johnson MD. Synaptic glutamate release by postnatal rat
serotonergic neurons in microculture. Neuron. (1994) 12:433–
42. doi: 10.1016/0896-6273(94)90283-6
46. Jonas P, Bischofberger J, Sandkuhler J. Corelease of two fast
neurotransmitters at a central synapse. Science. (1998) 281:419–
24. doi: 10.1126/science.281.5375.419
47. Walker MC, Ruiz A, Kullmann DM. Monosynaptic GABAergic
signaling from dentate to CA3 with a pharmacological and
physiological profile typical of mossy fiber synapses. Neuron. (2001)
29:703–15. doi: 10.1016/S0896-6273(01)00245-8
48. Furshpan EJ, MacLeish PR, O’Lague PH, Potter DD. Chemical
transmission between rat sympathetic neurons and cardiac myocytes
developing in microcultures: evidence for cholinergic, adrenergic,
and dual-function neurons. Proc Natl Acad Sci USA. (1976)
73:4225–9. doi: 10.1073/pnas.73.11.4225
49. Furshpan EJ, Landis SC, Matsumoto SG, Potter DD. Synaptic
functions in rat sympathetic neurons in microcultures. I
Secretion of norepinephrine and acetylcholine. J Neurosci. (1986)
6:1061–79. doi: 10.1523/JNEUROSCI.06-04-01061.1986
50. Potter DD, Landis SC, Matsumoto SG, Furshpan EJ. Synaptic
functions in rat sympathetic neurons in microcultures. II
Adrenergic/cholinergic dual status and plasticity. J Neurosci. (1986)
6:1080–98. doi: 10.1523/JNEUROSCI.06-04-01080.1986
51. Reichardt LF, Patterson PH. Neurotransmitter synthesis and uptake by
isolated sympathetic neurones in microcultures. Nature. (1977) 270:147–
51. doi: 10.1038/270147a0
52. Schotzinger RJ, Landis SC. Cholinergic phenotype developed by
noradrenergic sympathetic neurons after innervation of a novel cholinergic
target in vivo. Nature. (1988) 335:637–9. doi: 10.1038/335637a0
53. Apostolova G, Dorn R, Ka S, Hallbook F, Lundeberg J, Liser
K, et al. Neurotransmitter phenotype-specific expression changes
in developing sympathetic neurons. Mol Cell Neurosci. (2007)
35:397–408. doi: 10.1016/j.mcn.2007.03.014
54. Brodski C, Schaubmar A, Dechant G. Opposing functions of GDNF and NGF
in the development of cholinergic and noradrenergic sympathetic neurons.
Mol Cell Neurosci. (2002) 19:528–38. doi: 10.1006/mcne.2001.1093
55. Brodski C, Schnurch H, Dechant G. Neurotrophin-3 promotes the cholinergic
differentiation of sympathetic neurons. Proc Natl Acad Sci USA. (2000)
97:9683–8. doi: 10.1073/pnas.160080697
56. Habecker BA, Pennica D, Landis SC. Cardiotrophin-1 is not the
sweat gland-derived differentiation factor. Neuroreport. (1995)
7:41–4. doi: 10.1097/00001756-199512000-00009
57. Hiltunen PH, Airaksinen MS. Sympathetic cholinergic target innervation
requires GDNF family receptor GFR alpha 2. Mol Cell Neurosci. (2004)
26:450–7. doi: 10.1016/j.mcn.2004.04.003
58. Rao MS, Patterson PH, Landis SC. Multiple cholinergic differentiation factors
are present in footpad extracts: comparison with known cholinergic factors.
Development. (1992) 116:731–44. doi: 10.1242/dev.116.3.731
59. Rao MS, Sun Y, Escary JL, Perreau J, Tresser S, Patterson PH, et al. Leukemia
inhibitory factor mediates an injury response but not a target-directed
developmental transmitter switch in sympathetic neurons. Neuron. (1993)
11:1175–85. doi: 10.1016/0896-6273(93)90229-K
60. Stanke M, Duong CV, Pape M, Geissen M, Burbach G, Deller T, et al. Target-
dependent specification of the neurotransmitter phenotype: cholinergic
differentiation of sympathetic neurons is mediated in vivo by gp 130 signaling.
Development. (2006) 133:141–50. doi: 10.1242/dev.02189
61. Yamamori T, Fukada K, Aebersold R, Korsching S, Fann MJ, Patterson
PH. The cholinergic neuronal differentiation factor from heart cells
is identical to leukemia inhibitory factor. Science. (1989) 246:1412–
6. doi: 10.1126/science.2512641
62. Olivas A, Gardner RT, Wang L, Ripplinger CM, Woodward WR, Habecker
BA. Myocardial infarction causes transient cholinergic transdifferentiation
of cardiac sympathetic nerves via gp130. J Neurosci. (2016) 36:479–
88. doi: 10.1523/JNEUROSCI.3556-15.2016
63. Vega A, Luther JA, Birren SJ, Morales MA. Segregation of the classical
transmitters norepinephrine and acetylcholine and the neuropeptide Y in
sympathetic neurons: modulation by ciliary neurotrophic factor or prolonged
growth in culture. Dev Neurobiol. (2010) 70:913–28. doi: 10.1002/dneu.20834
Frontiers in Cardiovascular Medicine | www.frontiersin.org 6February 2022 | Volume 9 | Article 842656
Clyburn et al. Neurocircuitry of the Sympathetic Ganglia
64. Cifuentes F, Morales MA. Functional implications of
neurotransmitter segregation. Front Neural Circuits. (2021)
15:738516. doi: 10.3389/fncir.2021.738516
65. Eccles JC. Chemical transmission and Dale’s principle. Prog Brain Res. (1986)
68:3–13. doi: 10.1016/S0079-6123(08)60227-7
66. Condon AF, Robinson BG, Asad N, Dore TM, Tian L, Williams JT. The
residence of synaptically released dopamine on D2 autoreceptors. Cell Rep.
(2021) 36:109465. doi: 10.1016/j.celrep.2021.109465
67. Tedjo W, Nejad JE, Feeny R, Yang L, Henry CS, Tobet S, et al. Electrochemical
biosensor system using a CMOS microelectrode array provides high
spatially and temporally resolved images. Biosens Bioelectron. (2018) 114:78–
88. doi: 10.1016/j.bios.2018.04.009
68. Tjahjono N, Jin Y, Hsu A, Roukes M, Tian L. Letting the little light of mind
shine: Advances and future directions in neurochemical detection. Neurosci
Res Nov. (2021) 2021:30. doi: 10.1016/j.neures.2021.11.012
69. Silm K, Yang J, Marcott PF, Asensio CS, Eriksen J, Guthrie DA.
Synaptic vesicle recycling pathway determines neurotransmitter
content and release properties. Neuron. (2019). 102:786–800
e785. doi: 10.1016/j.neuron.2019.03.031
70. Takacs VT, Cserep C, Schlingloff D, Posfai B, Szonyi A, Sos KE, et al. Co-
transmission of acetylcholine and GABA regulates hippocampal states. Nat
Commun. (2018) 9:2848. doi: 10.1038/s41467-018-05136-1
71. Fortin GM, Ducrot C, Giguere N, Kouwenhoven WM, Bourque MJ, Pacelli
C, et al. Segregation of dopamine and glutamate release sites in dopamine
neuron axons: regulation by striatal target cells. FASEB J. (2019) 33:400–
17. doi: 10.1096/fj.201800713RR
72. Granger AJ, Wang W, Robertson K, El-Rifai M, Zanello AF, Bistrong
K, et al. Cortical ChAT(+) neurons co-transmit acetylcholine and
GABA in a target- and brain-region-specific manner. Elife. (2020)
9:57749. doi: 10.7554/eLife.57749.sa2
73. Liang K, Wei L, Chen L. Exocytosis, endocytosis, and
their coupling in excitable cells. Front Mol Neurosci. (2017)
10:109. doi: 10.3389/fnmol.2017.00109
74. von Twickel A, Kowatschew D, Salturk M, Schauer M, Robertson B, Korsching
S, et al. Individual dopaminergic neurons of lamprey SNc/VTA project to both
the striatum and optic tectum but restrict co-release of glutamate to striatum
only. Curr Biol. (2019) 29:677–685 e676. doi: 10.1016/j.cub.2019.01.004
75. Ninomiya I, Malpas SC, Matsukawa K, Shindo T, Akiyama T. The amplitude
of synchronized cardiac sympathetic nerve activity reflects the number of
activated pre- and postganglionic fibers in anesthetized cats. J Auton Nerv Syst.
(1993) 45:139–47. doi: 10.1016/0165-1838(93)90125-E
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.
Copyright © 2022 Clyburn, Andresen, Ingram and Habecker. 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.
Frontiers in Cardiovascular Medicine | www.frontiersin.org 7February 2022 | Volume 9 | Article 842656
... Sympathetic ganglia were traditionally described as simple relays between cholinergic preganglionic neurons and noradrenergic postganglionic neurons that innervate peripheral tissues (Jänig, 2008;Vegh et al., 2016). Discrepancies from this simple organization suggest that additional steps may exist in sympathetic ganglionic neurocircuitry (Clyburn et al., 2022). Almost 50 years ago, intracellular recordings of postganglionic sympathetic neurons from rabbit superior cervical ganglia (SCG) indicated that stimulation of the postganglionic nerve trunk unexpectedly activated late arriving, long-lasting depolarizations following retrograde activation of most neurons (Erulkar & Woodward, 1968). ...
... The remaining eEPSCs most probably represent incomplete deletion of the floxed ChAT allele in some cells because Cre induction did not occur in 100% of sympathetic neurons (Fig. 1). Our results clearly eliminate alternative explanations for retrograde eEPSCs such as postganglionic axons originating from other ganglia or sensory fibers traveling up an otherwise efferent nerve track to the sympathetic ganglia (Clyburn et al., 2022;Perri et al., 1970). Functional studies have suggested that activation of cardiac sensory afferents traveling via the SG increase blood pressure by influencing efferent motoneuron activity and that these effects are lost after severing connections to the spinal cord (Peterson & Brown, 1971). ...
Article
Full-text available
The sympathetic nervous system vitally regulates autonomic functions, including cardiac activity. Postganglionic neurons of the sympathetic chain ganglia relay signals from the central nervous system to autonomic peripheral targets. Disrupting this flow of information often dysregulates organ function and leads to poor health outcomes. Despite the importance of these sympathetic neurons, fundamental aspects of the neurocircuitry within peripheral ganglia remain poorly understood. Conventionally, simple monosynaptic cholinergic pathways from preganglionic neurons are thought to activate postganglionic sympathetic neurons. However, early studies suggested more complex neurocircuits may be present within sympathetic ganglia. The present study recorded synaptic responses in sympathetic stellate ganglia neurons following electrical activation of the pre‐ and postganglionic nerve trunks and used genetic strategies to assess the presence of collateral projections between postganglionic neurons of the stellate ganglia. Orthograde activation of the preganglionic nerve trunk, T‐2, uncovered high jitter synaptic latencies consistent with polysynaptic connections. Pharmacological inhibition of nicotinic acetylcholine receptors with hexamethonium blocked all synaptic events. To confirm that high jitter, polysynaptic events were due to the presence of cholinergic collaterals from postganglionic neurons within the stellate ganglion, we knocked out choline acetyltransferase in adult noradrenergic neurons. This genetic knockout eliminated orthograde high jitter synaptic events and EPSCs evoked by retrograde activation. These findings suggest that cholinergic collateral projections arise from noradrenergic neurons within sympathetic ganglia. Identifying the contributions of collateral excitation to normal physiology and pathophysiology is an important area of future study and may offer novel therapeutic targets for the treatment of autonomic imbalance. image Key points Electrical stimulation of a preganglionic nerve trunk evoked fast synaptic transmission in stellate ganglion neurons with low and high jitter latencies. Retrograde stimulation of a postganglionic nerve trunk evoked direct, all‐or‐none action currents and delayed nicotinic EPSCs indistinguishable from orthogradely‐evoked EPSCs in stellate neurons. Nicotinic acetylcholine receptor blockade prevented all spontaneous and evoked synaptic activity. Knockout of acetylcholine production in noradrenergic neurons eliminated all retrogradely‐evoked EPSCs but did not change retrograde action currents, indicating that noradrenergic neurons have cholinergic collaterals connecting neurons within the stellate ganglion.
... Sympathetic ganglia were traditionally described as simple relays between cholinergic preganglionic neurons and noradrenergic postganglionic neurons that innervate peripheral tissues (Jänig, 2008;Vegh et al., 2016). Discrepancies from this simple organization suggest additional steps may exist in sympathetic ganglionic neurocircuitry (reviewed by (Clyburn et al., 2022). Nearly 50 years ago, intracellular recordings of postganglionic sympathetic neurons from rabbit superior cervical ganglia (SCG) indicated that stimulation of the postganglionic nerve trunk unexpectedly activated late arriving, longlasting depolarizations following retrograde activation of most neurons (Erulkar & Woodward, 1968). ...
... The remaining eEPSCs most likely represent incomplete deletion of the floxed ChAT allele in some cells, as Cre induction did not occur in 100% of sympathetic neurons (Fig. 1). Our results clearly eliminate alternative explanations for retrograde eEPSCs such as postganglionic axons originating from other ganglia or sensory fibers traveling up an otherwise efferent nerve track to the sympathetic ganglia (Perri et al., 1970;Clyburn et al., 2022). Functional studies have suggested that activation of cardiac sensory afferents traveling via the SG increase blood pressure by influencing efferent motoneuron activity and that these effects are lost after severing connections to the spinal cord (Peterson & Brown, 1971). ...
Article
Postganglionic neurons of the stellate ganglia (SG) are an integral component of sympathetic regulation of the heart. SG neurons receive preganglionic cholinergic inputs via T‐2 and send noradrenergic projections to the heart via the inferior cardiac nerve (iCN) to influence cardiac activity. SG neurons are thought to receive simple monosynaptic inputs from preganglionic neurons, but early studies have hinted that cholinergic collateral projections may be present in the SG. In culture, SG neurons form noradrenergic and cholinergic synapses between one another, suggesting that SG neurons express dual neurochemical phenotypes. This study tests the hypothesis that cholinergic collateral projections are present between sympathetic neurons of the intact SG in mice. Whole‐cell patch clamp recordings were made of intact SG neurons from male (N=8) and female (N=6) ChAT TH+CreERT2/lox mice with and without tamoxifen treatment (N=7 ChAT KO mice and N=7 control mice, respectively). Tamoxifen treatment in ChAT TH+CreERT2/lox mice induced the deletion of choline acetyltransferase (ChAT) in tyrosine hydroxylase positive (i.e. noradrenergic) cells. This strategy blocks putative acetylcholine synthesis in SG neurons while leaving preganglionic cholinergic synapses intact. To search for collaterals, the preganglionic nerve trunk, T‐2, was electrically stimulated to evoke excitatory postsynaptic currents (eEPSCs). The amplitude and jitter (i.e. variability in latency) of eEPSCs were examined. Additionally, the postganglionic nerve trunk, the iCN, was electrically stimulated to evoke retrograde action currents (rAC) and the presence of evoked synaptic currents was assessed. The amplitude and frequency of spontaneous EPSCs (sEPSCs) were also examined. ChAT KO significantly reduced the frequency of sEPSCs compared to control mice (1.21±0.559events/s vs. 0.13±0.048events/s, P<0.05 from Student’s unpaired t‐test, N=5‐6 cells) while the amplitude was unaffected (74.4±26.10pA vs. 39.4±10.46pA, P>0.05 from Student’s unpaired t‐test, N=5‐6 cells). This suggests that ChAT KO reduces the presynaptic activity of postganglionic noradrenergic neurons within the SG. In control mice, stimulation of T‐2 evoked high jitter eEPSCs, which was significantly reduced in ChAT KO mice (327.1±92.14μs vs. 39.2±15.69μs, P<0.05 from Student’s unpaired t‐test, N=5‐6 cells) Additionally, the proportion of neurons that exhibited evoked synaptic currents following stimulation of the iCN was significantly reduced in ChAT KO mice (3/4 cells vs. 1/9 cells, P<0.05 from χ² test). Chemical transmission at additional synapses accounts for the high jitter events observed in disynaptic pathways. High jitter responses following T‐2 stimulation and evoked synaptic responses after iCN stimulation in control neurons are both consistent with the collateral hypothesis. The loss of high jitter responses and evoked synaptic currents in ChAT KO mice also suggests these collateral projections are cholinergic. This suggests that collateral synaptic activity contributes to the increased frequency of sEPSCs in control mice. Understanding SG neurocircuitry is critical in deciphering changes in sympathetic activity in many pathophysiological conditions such as myocardial infarction and heart failure and may identify novel therapeutic targets for the treatment of autonomic imbalance.
... Sympathetic ganglia have been canonically understood as a simple monosynaptic serial connection between cholinergic preganglionic neurons and noradrenergic postganglionic neurons that innervate peripheral organs, vasculature and the heart. However, as reviewed by Clyburn et al. (2022), this connection may not be so simple, as noradrenergic sympathetic postganglionic neurons may host cholinergic synapses; these would function to relay sympathetic postganglionic activity in parallel postganglionic circuits in response to activation of a single preganglionic fibre. To expose this connectome, in an article recently published in The Journal of Physiology, Clyburn's group set out to experimentally determine whether cholinergic collaterals arise from noradrenergic neurons within the sympathetic stellate ganglia in mice (Clyburn et al., 2023). ...
Article
Altered autonomic input to the heart plays a major role in atrial fibrillation (AF). Autonomic neurons termed ganglionated plexi (GP) are clustered on the heart surface to provide the last point of neural control of cardiac function. To date the properties of GP neurons in humans are unknown. Here we have addressed this knowledge gap in human GP neuron structure and physiology in patients with and without AF. Human right atrial GP neurons embedded in epicardial adipose tissue were excised during open heart surgery performed on both non‐AF and AF patients and then characterised physiologically by whole cell patch clamp techniques. Structural analysis was also performed after fixation at both the single cell and at the entire GP levels via three‐dimensional confocal imaging. Human GP neurons were found to exhibit unique properties and structural complexity with branched neurite outgrowth. Significant differences in excitability were revealed between AF and non‐AF GP neurons as measured by lower current to induce action potential firing, a reduced occurrence of low action potential firing rates, decreased accommodation and increased synaptic density. Visualisation of entire GPs showed almost all neurons are cholinergic with a small proportion of noradrenergic and dual phenotype neurons. Phenotypic distribution differences occurred with AF including decreased cholinergic and dual phenotype neurons, and increased noradrenergic neurons. These data show both functional and structural differences occur between GP neurons from patients with and without AF, highlighting that cellular plasticity occurs in neural input to the heart that could alter autonomic influence on atrial function. image Key points The autonomic nervous system plays a critical role in regulating heart rhythm and the initiation of AF; however, the structural and functional properties of human autonomic neurons in the autonomic ganglionated plexi (GP) remain unknown. Here we perform the first whole cell patch clamp electrophysiological and large tissue confocal imaging analysis of these neurons from patients with and without AF. Our data show human GP neurons are functionally and structurally complex. Measurements of action potential kinetics show higher excitability in GP neurons from AF patients as measured by lower current to induce action potential firing, reduced low firing action potential rates, and decreased action potential accommodation. Confocal imaging shows increased synaptic density and noradrenergic phenotypes in patients with AF. Both functional and structural differences occur in GP neurons from patients with AF that could alter autonomic influence on atrial rhythm.
Article
Full-text available
Cardiac function is tightly regulated by the autonomic nervous system (ANS). Activation of the sympathetic nervous system increases cardiac output by increasing heart rate and stroke volume, while parasympathetic nerve stimulation instantly slows heart rate. Importantly, imbalance in autonomic control of the heart has been implicated in the development of arrhythmias and heart failure. Understanding of the mechanisms and effects of autonomic stimulation is a major challenge because synapses in different regions of the heart result in multiple changes to heart function. For example, nerve synapses on the sinoatrial node (SAN) impact pacemaking, while synapses on contractile cells alter contraction and arrhythmia vulnerability. Here, we present a multiscale neurocardiac modelling and simulator tool that predicts the effect of efferent stimulation of the sympathetic and parasympathetic branches of the ANS on the cardiac SAN and ventricular myocardium. The model includes a layered representation of the ANS and reproduces firing properties measured experimentally. Model parameters are derived from experiments and atomistic simulations. The model is a first prototype of a digital twin that is applied to make predictions across all system scales, from subcellular signalling to pacemaker frequency to tissue level responses. We predict conditions under which autonomic imbalance induces proarrhythmia and can be modified to prevent or inhibit arrhythmia. In summary, the multiscale model constitutes a predictive digital twin framework to test and guide high‐throughput prediction of novel neuromodulatory therapy. image Key points A multi‐layered model representation of the autonomic nervous system that includes sympathetic and parasympathetic branches, each with sparse random intralayer connectivity, synaptic dynamics and conductance based integrate‐and‐fire neurons generates firing patterns in close agreement with experiment. A key feature of the neurocardiac computational model is the connection between the autonomic nervous system and both pacemaker and contractile cells, where modification to pacemaker frequency drives initiation of electrical signals in the contractile cells. We utilized atomic‐scale molecular dynamics simulations to predict the association and dissociation rates of noradrenaline with the β‐adrenergic receptor. Multiscale predictions demonstrate how autonomic imbalance may increase proclivity to arrhythmias or be used to terminate arrhythmias. The model serves as a first step towards a digital twin for predicting neuromodulation to prevent or reduce disease.
Article
Purpose: Cardiac autonomic dysfunction is one of the main pillars of cardiovascular pathophysiology. The purpose of this review is to provide an overview of the current state of the art on the pathological remodeling that occurs within the autonomic nervous system with cardiac injury and available neuromodulatory therapies for autonomic dysfunction in heart failure. Methods: Data from peer-reviewed publications on autonomic function in health and after cardiac injury are reviewed. The role of and evidence behind various neuromodulatory therapies both in preclinical investigation and in-use in clinical practice are summarized. Results: A harmonic interplay between the heart and the autonomic nervous system exists at multiple levels of the neuraxis. This interplay becomes disrupted in the setting of cardiovascular disease, resulting in pathological changes at multiple levels, from subcellular cardiac signaling of neurotransmitters to extra-cardiac, extra-thoracic remodeling. The subsequent detrimental cycle of sympathovagal imbalance, characterized by sympathoexcitation and parasympathetic withdrawal, predisposes to ventricular arrhythmias, progression of heart failure, and cardiac mortality. Knowledge on the etiology and pathophysiology of this condition has increased exponentially over the past few decades, resulting in a number of different neuromodulatory approaches. However, significant knowledge gaps in both sympathetic and parasympathetic interactions and causal factors that mediate progressive sympathoexcitation and parasympathetic dysfunction remain. Conclusions: Although our understanding of autonomic imbalance in cardiovascular diseases has significantly increased, specific, pivotal mediators of this imbalance and the recognition and implementation of available autonomic parameters and neuromodulatory therapies are still lagging.
Article
Full-text available
Here, we present and discuss the characteristics and properties of neurotransmitter segregation, a subtype of neurotransmitter cotransmission. We review early evidence of segregation and discuss its properties, such as plasticity, while placing special emphasis on its probable functional implications, either in the central nervous system (CNS) or the autonomic nervous system. Neurotransmitter segregation is a process by which neurons separately route transmitters to independent and distant or to neighboring neuronal processes; it is a plastic phenomenon that changes according to synaptic transmission requirements and is regulated by target-derived signals. Distant neurotransmitter segregation in the CNS has been shown to be related to an autocrine/paracrine function of some neurotransmitters. In retinal amacrine cells, segregation of acetylcholine (ACh) and GABA, and glycine and glutamate to neighboring terminals has been related to the regulation of the firing rate of direction-selective ganglion cells. In the rat superior cervical ganglion, segregation of ACh and GABA to neighboring varicosities shows a heterogeneous regional distribution, which is correlated to a similar regional distribution in transmission strength. We propose that greater segregation of ACh and GABA produces less GABAergic inhibition, strengthening ganglionic transmission. Segregation of ACh and GABA varies in different physiopathological conditions; specifically, segregation increases in acute sympathetic hyperactivity that occurs in cold stress, does not vary in chronic hyperactivity that occurs in hypertension, and rises in early ages of normotensive and hypertensive rats. Given this, we propose that variations in the extent of transmitter segregation may contribute to the alteration of neural activity that occurs in some physiopathological conditions and with age.
Article
Full-text available
The cardiac autonomic nervous system (ANS) is the main modulator of heart function, adapting contraction force, and rate to the continuous variations of intrinsic and extrinsic environmental conditions. While the parasympathetic branch dominates during rest-and-digest sympathetic neuron (SN) activation ensures the rapid, efficient, and repeatable increase of heart performance, e.g., during the “fight-or-flight response.” Although the key role of the nervous system in cardiac homeostasis was evident to the eyes of physiologists and cardiologists, the degree of cardiac innervation, and the complexity of its circuits has remained underestimated for too long. In addition, the mechanisms allowing elevated efficiency and precision of neurogenic control of heart function have somehow lingered in the dark. This can be ascribed to the absence of methods adequate to study complex cardiac electric circuits in the unceasingly moving heart. An increasing number of studies adds to the scenario the evidence of an intracardiac neuron system, which, together with the autonomic components, define a little brain inside the heart, in fervent dialogue with the central nervous system (CNS). The advent of optogenetics, allowing control the activity of excitable cells with cell specificity, spatial selectivity, and temporal resolution, has allowed to shed light on basic neuro-cardiology. This review describes how optogenetics, which has extensively been used to interrogate the circuits of the CNS, has been applied to untangle the knots of heart innervation, unveiling the cellular mechanisms of neurogenic control of heart function, in physiology and pathology, as well as those participating to brain–heart communication, back and forth. We discuss existing literature, providing a comprehensive view of the advancement in the understanding of the mechanisms of neurogenic heart control. In addition, we weigh the limits and potential of optogenetics in basic and applied research in neuro-cardiology.
Article
Full-text available
Neuromodulation mediated by synaptically released endogenous transmitters acting in G-protein-coupled receptors (GPCRs) is slow primarily because of multistep downstream signaling. What is less well understood is the spatial and temporal kinetics of transmitter and receptor interaction. The present work uses the combination of the dopamine sensor, dLight, to detect the spatial release and diffusion of dopamine and a caged form of a D2-dopamine receptor antagonist, CyHQ-sulpiride, to rapidly block the D2 autoreceptors. Photoactivation of the CyHQ-sulpiride blocks receptors in milliseconds such that the time course of dopamine/receptor interaction is mapped onto the downstream signaling. The results show that highly localized release, but not dopamine diffusion, defines the time course of the functional interaction between dopamine and D2 autoreceptors, which determines downstream inhibition.
Article
Full-text available
The mouse cerebral cortex contains neurons that express choline acetyltransferase (ChAT) and are a potential local source of acetylcholine. However, the neurotransmitters released by cortical ChAT⁺ neurons and their synaptic connectivity are unknown. We show that the nearly all cortical ChAT⁺ neurons in mice are specialized VIP⁺ interneurons that release GABA strongly onto other inhibitory interneurons and acetylcholine sparsely onto layer 1 interneurons and other VIP⁺/ChAT⁺ interneurons. This differential transmission of ACh and GABA based on the postsynaptic target neuron is reflected in VIP⁺/ChAT⁺ interneuron pre-synaptic terminals, as quantitative molecular analysis shows that only a subset of these are specialized to release acetylcholine. In addition, we identify a separate, sparse population of non-VIP ChAT⁺ neurons in the medial prefrontal cortex with a distinct developmental origin that robustly release acetylcholine in layer 1. These results demonstrate both cortex-region heterogeneity in cortical ChAT⁺ interneurons and target-specific co-release of acetylcholine and GABA.
Article
Full-text available
The mouse cerebral cortex contains neurons that express choline acetyltransferase (ChAT) and are a potential local source of acetylcholine. However, the neurotransmitters released by cortical ChAT+ neurons and their synaptic connectivity are unknown. We show that the nearly all cortical ChAT+ neurons in mice are specialized VIP+ interneurons that release GABA strongly onto other inhibitory interneurons and acetylcholine sparsely onto layer 1 interneurons and other VIP+/ChAT+ interneurons. This differential transmission of ACh and GABA based on the postsynaptic target neuron is reflected in VIP+/ChAT+ interneuron pre-synaptic terminals, as quantitative molecular analysis shows that only a subset of these are specialized to release acetylcholine. In addition, we identify a separate, sparse population of non-VIP ChAT+ neurons in the medial prefrontal cortex with a distinct developmental origin that robustly release acetylcholine in layer 1. These results demonstrate both cortex-region heterogeneity in cortical ChAT+ interneurons and target-specific co-release of acetylcholine and GABA.
Article
Full-text available
Heart rate is under the precise control of the autonomic nervous system. However, the wiring of peripheral neural circuits that regulate heart rate is poorly understood. Here, we developed a clearing-imaging-analysis pipeline to visualize innervation of intact hearts in 3D and employed a multi-technique approach to map parasympathetic and sympathetic neural circuits that control heart rate in mice. We anatomically and functionally identify cholinergic neurons and noradrenergic neurons in an intrinsic cardiac ganglion and the stellate ganglia, respectively, that project to the sinoatrial node. We also report that the heart rate response to optogenetic versus electrical stimulation of the vagus nerve displays different temporal characteristics and that vagal afferents enhance parasympathetic and reduce sympathetic tone to the heart via central mechanisms. Our findings provide new insights into neural regulation of heart rate, and our methodology to study cardiac circuits can be readily used to interrogate neural control of other visceral organs.
Article
Synaptic transmission via neurochemical release is the fundamental process that integrates and relays encoded information in the brain to regulate physiological function, cognition, and emotion. To unravel the biochemical, biophysical, and computational mechanisms of signal processing, one needs to precisely measure the neurochemical release dynamics with molecular and cell-type specificity and high resolution. Here we reviewed the development of analytical, electrochemical, and fluorescence imaging approaches to detect neurotransmitter and neuromodulator release. We discussed the advantages and practicality in implementation of each technology for ease-of-use, flexibility for multimodal studies, and challenges for future optimization. We hope this review will provide a versatile guide for tool engineering and applications for recording neurochemical release.
Article
In contrast to temporal coding by synaptically acting neurotransmitters such as glutamate, neuromodulators such as monoamines signal changes in firing rate. The two modes of signaling have been thought to reflect differences in release by different cells. We now find that midbrain dopamine neurons release glutamate and dopamine with different properties that reflect storage in different synaptic vesicles. The vesicles differ in release probability, coupling to presynaptic Ca ²⁺ channels and frequency dependence. Although previous work has attributed variation in these properties to differences in location or cytoskeletal association of synaptic vesicles, the release of different transmitters shows that intrinsic differences in vesicle identity drive different modes of release. Indeed, dopamine but not glutamate vesicles depend on the adaptor protein AP-3, revealing an unrecognized linkage between the pathway of synaptic vesicle recycling and the properties of exocytosis. Storage of the two transmitters in different vesicles enables the transmission of distinct signals. In this work on glutamate corelease by dopamine neurons, Silm et al. show that an individual neuron expresses two classes of synaptic vesicle that form through distinct mechanisms and transmit distinct information due to differences in frequency dependence.
Article
Dopaminergic neurons in the substantia nigra (SNc) innervate both striatum and the superior colliculus in mammals, as well as its homolog the optic tectum in lampreys, belonging to the oldest group of living vertebrates [1–3]. In the lamprey, we have previously shown that the same neuron sends axonal branches to both striatum and the optic tectum [3]. Here, we show that most neurons in the lamprey SNc and ventral tegmental area (VTA) (also referred to as the nucleus of the posterior tuberculum) express not only tyrosine hydroxylase (TH), in lamprey a marker of dopaminergic neurons [4], but also the vesicular glutamate transporter (vGluT), suggesting that glutamate is a co-transmitter. Remarkably, the axonal branches that project to striatum elicit both dopaminergic and glutamatergic synaptic effects on striatal neurons, whereas the axonal projections to the optic tectum only evoke dopaminergic effects. Thus, axonal branches from the same neuron can use two transmitters in one branch and only one in the other. Previous studies suggest that, along an individual dopaminergic axon, there can be microdomains of either TH or vGluT [5–8]. In addition, the present results demonstrate that entire axonal branches to one target structure can differ from that of branches to another target, both originating from the same dopamine neuron. This implies that a given dopamine neuron can exert different effects on two different target structures. The combined release of dopamine and glutamate may be appropriate in striatum, whereas the effects exerted on the tectal motor center may be better served with a selective dopaminergic modulation. von Twickel et al. show that most dopaminergic neurons in the lamprey SNc/VTA co-express glutamate and individual co-expressing neurons project to both the striatum and optic tectum. Remarkably, glutamate co-release is restricted to striatum, indicating differential neurotransmitter release depending on target area.
Article
The technique of microneurography has advanced the field of neuroscience for the past 50 years. While there have been a number of reviews on microneurography, this paper takes an objective approach to exploring the impact of microneurography studies. Briefly, Web of Science (Thomson Reuters) was used to identify the highest citation articles over the past 50 years, and key findings are presented in a decade-by-decade highlight. This includes the establishment of microneurography in the 1960s, the acceleration of the technique by Gunnar Wallin in the 1970s, the international collaborations of the 1980s and 1990s, and finally the highest impact studies from 2000 to present. This journey through 50 years of microneurographic research related to peripheral sympathetic nerve activity includes a historical context for several of the laboratory interventions commonly used today (e.g., cold pressor test, mental stress, lower body negative pressure, isometric handgrip, etc.) and how these interventions and experimental approaches have advanced our knowledge of cardiovascular, cardiometabolic, and other human diseases and conditions.