ArticlePDF AvailableLiterature Review

Abstract

The locus coeruleus (LC) is a small brainstem structure located in the lower pons and is the main source of noradrenaline (NA) in the brain. Via its phasic and tonic firing, it modulates cognition and autonomic functions and is involved in the brain's immune response. The extent of degeneration to the LC in healthy ageing remains unclear, however, noradrenergic dysfunction may contribute to the pathogenesis of Alzheimer's (AD) and Parkinson's disease (PD). Despite their differences in progression at later disease stages, the early involvement of the LC may lead to comparable behavioural symptoms such as preclinical sleep problems and neuropsychiatric symptoms as a result of AD and PD pathology. In this review, we draw attention to the mechanisms that underlie LC degeneration in ageing, AD and PD. We aim to motivate future research to investigate how early degeneration of the noradrenergic system may play a pivotal role in the pathogenesis of AD and PD which may also be relevant to other neurodegenerative diseases.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
Available online 11 July 2023
0149-7634/© 2023 Published by Elsevier Ltd.
Noradrenergic neuromodulation in ageing and disease
F. Krohn
a
,
b
,
1
,
2
, E. Lancini
a
,
b
,
*
,
1
,
2
, M. Ludwig
b
,
d
,
2
, M. Leiman
a
,
b
,
2
, G. Guruprasath
a
,
b
,
2
,
L. Haag
b
,
2
, J. Panczyszyn
b
,
2
, E. Düzel
a
,
b
,
c
,
d
,
2
,
3
, D. H¨
ammerer
a
,
b
,
c
,
d
,
e
,
4
, M. Betts
a
,
b
,
d
,
2
a
German Center for Neurodegenerative Diseases (DZNE), Otto-von-Guericke University Magdeburg, Magdeburg, Germany
b
Institute of Cognitive Neurology and Dementia Research (IKND), Otto-von-Guericke University Magdeburg, Magdeburg, Germany
c
Institute of Cognitive Neuroscience, University College London, London UK-WC1E 6BT, UK
d
CBBS Center for Behavioral Brain Sciences, University of Magdeburg, Magdeburg, Germany
e
Department of Psychology, University of Innsbruck, A-6020 Innsbruck, Austria
ARTICLE INFO
Keywords:
Noradrenaline
Noradrenergic neuromodulation
Locus coeruleus
ABSTRACT
The locus coeruleus (LC) is a small brainstem structure located in the lower pons and is the main source of
noradrenaline (NA) in the brain. Via its phasic and tonic ring, it modulates cognition and autonomic functions
and is involved in the brains immune response. The extent of degeneration to the LC in healthy ageing remains
unclear, however, noradrenergic dysfunction may contribute to the pathogenesis of Alzheimers (AD) and Par-
kinsons disease (PD). Despite their differences in progression at later disease stages, the early involvement of the
LC may lead to comparable behavioural symptoms such as preclinical sleep problems and neuropsychiatric
symptoms as a result of AD and PD pathology. In this review, we draw attention to the mechanisms that underlie
LC degeneration in ageing, AD and PD. We aim to motivate future research to investigate how early degeneration
of the noradrenergic system may play a pivotal role in the pathogenesis of AD and PD which may also be relevant
to other neurodegenerative diseases.
1. Introduction
The locus coeruleus (LC), Latin for blue spot, is a brainstem nucleus
rst described in 1786 by F´
elix Vicq-dAzyr (Vicq-dAzyr, 1786). Despite
its small size, the LC projects to and receives input from widespread
brain regions (Liebe et al., 2022; Szabadi, 2013) and is thus involved in
numerous functions related to cognition such as memory formation
(Amaral and Foss, 1975; Gibbs et al., 2010; Hansen, 2017; Kety, 1972;
Zornetzer and Gold, 1976), attention, sensory processing (Bouret and
Sara, 2002; Lecas, 2004), novelty (Vankov et al., 1995; Yamasaki and
Takeuchi, 2017) and emotional memory (H¨
ammerer et al., 2018). It is
involved in autonomic functions such as blood pressure (Sved and
Felsten, 1987), immune function (Lehnert et al., 1998; Rassnick et al.,
1994) and the sleep-wake cycle (for a review see Osorio-Forero et al.,
2022). Furthermore, it is involved in the ght or ight response by
modulating heart rate, blood pressure, salivation and pupil dilation
(Ross and Van Bockstaele, 2021; Samuels and Szabadi, 2008). Alter-
ations to the noradrenergic system occur as a result of degeneration, that
therefore may impair these functions. Indeed, in humans, the LC has
recently become increasingly relevant in healthy aging because several
cognitive functions supported by the noradrenergic system, such as
verbal intelligence (Clewett et al., 2016), response inhibition (Liu et al.,
2020; Tomassini et al., 2022), memory (Calarco et al., 2022; Dahl et al.,
2019; Langley et al., 2022; Liu et al., 2020), emotional memory
(H¨
ammerer et al., 2018; Sterpenich et al., 2006), attention and pro-
cessing speed (Calarco et al., 2022) decline in older age (Harada et al.,
2013).
In addition to ageing, degeneration and dysfunction of the LC
noradrenergic system (LC-NA) also occurs during the rst stages of
Alzheimer´s (AD) and Parkinson´s (PD) disease and may contribute to the
* Correspondence to: Institute of Cognitive Neurology and Dementia Research Otto-von-Guericke University IKND (Haus 64), Leipziger Str. 44, 39120 Magdeburg,
Germany.
E-mail address: elisa.lancini@dzne.de (E. Lancini).
1
Authors contributed equally to this work
2
Leipziger Str. 44/Haus 64, 39120 Magdeburg (Germany)
3
Leipziger Str. 44/Haus 64, 39120 Magdeburg (Germany), Institute of Cognitive Neuroscience, University College London, London, UK-WC1E 6BT, UK
4
Leipziger Str. 44/Haus 64, 39120 Magdeburg (Germany)
,
University College London, London, UK-WC1E 6BT, UK 3, Universit¨
atsstrasse 57, 6020 Innsbruck,
Austria
Contents lists available at ScienceDirect
Neuroscience and Biobehavioral Reviews
journal homepage: www.elsevier.com/locate/neubiorev
https://doi.org/10.1016/j.neubiorev.2023.105311
Received 29 April 2023; Received in revised form 29 June 2023; Accepted 7 July 2023
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
2
spread of pathology (see Section 4.). There is signicant loss in LC vol-
ume and MRI contrast in both AD and PD (Betts et al., 2019a; Jacobs
et al., 2021a; Madelung et al., 2022; Theolas et al., 2017; Zarow et al.,
2003). The LC-NA system is mainly linked to AD pathology through its
metabolites, whose levels increased and interact with tau and
amyloid-beta (Aß) pathology: a recent review and meta-analysis found
elevated cerebrospinal uid (CSF) MHPG levels in AD compared to
healthy controls, while noradrenaline (NA) levels were unchanged
(Lancini et al., 2023). Moreover, as the noradrenergic metabolite 3,
4-Dihydroxyphenyl Glycolaldehyde (DOPEGAL) interacts with tau, NA
has been suggested to indirectly accelerate the progression of AD disease
(Kang et al., 2022). In PD, the LC also shows structural degeneration
early in the disease, which might be linked to non-motor symptoms that
occur prior to substantia nigra (SN) degeneration such as anxiety (Lapiz
et al., 2001), depression (Remy et al., 2005) and REM sleep disturbances
(Ehrminger et al., 2016; García-Lorenzo et al., 2013).
Finally, given the link between LC integrity and symptoms of
neurodegenerative diseases and ageing, the LC presents as a natural
target for therapeutic interventions to ameliorate those symptoms. And
indeed, different therapeutic interventions, that increase noradrenergic
levels, have successfully been used to improve cognition in individuals
with PD and in individuals with mild cognitive impairment (MCI) and
AD.
In this review, we summarise the involvement of the LC-NA system in
healthy aging and in neurodegenerative disease. The aim of this review
is to summarise the latest research ndings and encourage further
research on the involvement of the noradrenergic system in both healthy
aging and in the aetiology of AD and PD.
2. The noradrenergic system
The LC is the main source of NA in the brain. The LCs azure look
stems from high levels of neuromelanin (NM), a byproduct of NA syn-
thesis and NA metabolism that binds metal ions (Zucca et al., 2006) Due
to its exposed location next to the 4th ventricle, the LC is more vulner-
able to inammatory molecules and toxins (for a review see Evans et al.,
2022; Matchett et al., 2021). Moreover, LC axons are partially myelin-
ated therefore less cost-efcient, resulting in higher energy consump-
tion, which in turn leads to higher levels of highly reactive oxygen
species (ROS) (Lushchak et al., 2021). Thus, the LC is more vulnerable to
the effects of ROS (for a review see Evans et al., 2022).
The LC receives afferent connections from the neocortex (Cedarbaum
and Aghajanian, 1978; Luppi et al., 1995), prefrontal cortex (PFC)
(Arnsten and Goldman-Rakic, 1984; Sesack et al., 1989), amygdala
(Cedarbaum and Aghajanian, 1978; Charney et al., 1998), and ventral
tegmental area (VTA) (Deutch et al., 1986) among other structures.
Efferent projections have been grouped into three major noradren-
ergic pathways: 1) the cortical (or ascendant) pathway (Szabadi,
2013), which includes projections to the ventral tegmental area (VTA)
(Alvarado et al., 2023; Mejias-Aponte, 2016), SN (Mejias-Aponte, 2016;
Rommelfanger and Weinshenker, 2007), amygdala (McCall et al., 2017;
Uematsu et al., 2017), hippocampus (Haring and Davis, 1985; Jones and
Moore, 1977; Takeuchi et al., 2016), hypothalamus (Giorgi et al., 2021;
Schwarz and Luo, 2015), thalamus (Beas et al., 2018; Rodenkirch et al.,
2019), basal forebrain (Espa˜
na and Berridge, 2006), PFC and sensory
cortices (McBurney-Lin et al., 2019; Schwarz and Luo, 2015; Szabadi,
2013), 2) the spinal pathway (or descendent) which includes brain-
stem nuclei such as the nervus vagi (Nosaka et al., 1982), sympathetic
premotor nuclei (Head et al., 1998; Tavares et al., 1996) and the
oculomotor nucleus (Carpenter et al., 1992) as well as various spinal
nuclei via 3) cerebellar pathways (Fu et al., 2011), where it connects to
both cerebellar cortex and nuclei (Dietrichs, 1988) and potentiates
Purkinje cell spiking (Moises et al., 1981), for a comprehensive review
on LC connections see Szabadi and colleagues (Szabadi, 2013).
Via the cortical pathways, the LC-NA system has been shown to be
involved in various cognitive processes such as memory (Clewett et al.,
2016; Takeuchi et al., 2016), attention (Dahl et al., 2019; Unsworth and
Robison, 2017) and sleep (Van Egroo et al., 2022). Via these projections,
the LC plays a crucial role in modulating the dopaminergic system.
Notably, noradrenergic terminals in cortical regions and the hippo-
campus have been found to co-release dopamine (DA), indicating a
functional overlap (Devoto et al., 2020, 2008, 2005; Kempadoo et al.,
2016; Pozzi et al., 1994; Smith and Greene, 2012; Takeuchi et al., 2016).
Moreover, the LC projects to the ventral tegmental area (VTA), a key hub
of the dopaminergic pathway, and modulates dopamine release in the
nucleus accumbens (nACC) and prefrontal cortex (PFC) (Sara, 2009).
Additionally, the LC has direct projections to the PFC, which, in turn,
sends inhibitory relay signals to the VTA. Reciprocal connections exist
between the VTA, PFC, and LC, forming a complex network that allows
for bidirectional modulation (Sara, 2009). NA modulation is also
implicated in psychiatric disorders such as PTSD (Debiec and LeDoux,
2006), depression (Brunello et al., 2002; Kobayashi et al., 2022; Remy
et al., 2005), ADHD (del Campo et al., 2011), schizophrenia (Meisenzahl
et al., 2007; Toda and Abi-Dargham, 2007), and substance use disorders
(Downs and McElligott, 2022), where alterations in noradrenaline-
dopamine signalling play a role.
The descendent pathways have been linked to sympathetic functions
such as blood pressure (Anselmo-Franci et al., 1999) and muscle tonus
(Kiyashchenko et al., 2001) while the LC cerebellar pathways remain
understudied up to date.
Signalling between LC and the projecting areas occur via NA. NA
synthesis starts with the amino acid tyrosine, which is converted into
dihydroxyphenylalanine (DOPA) and then into DA by the enzymes
tyrosine hydroxylase (TH), and DOPA decarboxylase. After its synthesis,
DA is transported into noradrenergic neurons, which contain the
enzyme dopamine β-hydroxylase (DBH), that converts DA to NA
(Molinoff and Axelrod, 1971). The synthesised NA is then stored in
presynaptic vesicles and released into the synaptic cleft where it binds to
the G-coupled adrenergic receptors (Ars) type β,
α
1, and
α
2 with
respectively intermediate, higher, and lowest afnity (Molinoff, 1984).
The effect of NA depends on the type of receptors expressed on the
postsynaptic neurons. Binding to
α
1 or β adrenergic receptors produces
stimulatory effects while binding to
α
2 adrenergic receptors produces
inhibitory effects on cell signalling (Schwarz and Luo, 2015; Szabadi,
2013).
After exerting its modulatory effect, NA is removed from the synaptic
cleft by either reuptake via NA transporters or by degradation by the
enzymes monoamine oxidase (MAO) and catechol-O-methyltransferase
(COMT) and aldehyde reductase (AR) into several metabolites.
NA is metabolised into normetanephrine (NMN) and 3-methoxy 4
hydroxyphenylglycol aldehyde (MOPEGAL) by COMT and MAO, or into
DOPEGAL, 3,4-dihydroxyphenylglycol (DHPG) and subsequently 3-
Methoxy-4-hydroxyphenylglycol (MHPG) by MAO, AR and COMT.
MHPG, considered as the major metabolite of NA in the brain (Schan-
berg et al., 1968), is further converted to MOPEGAL via alcohol dehy-
drogenase (ADH), and into vanillylmandelic acid through aldehyde
dehydrogenase (Eisenhofer et al., 2004; Kamal and Lappin, 2022;
Molinoff and Axelrod, 1971).
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
3
2.1. LC ring modalities
The LC has two ring modes: tonic (persistent) and phasic (short
bursts). Animal studies suggest that electric LC stimulation in rats leads
to higher NA release when the neurons are stimulated by phasic rather
than tonic stimulation (e.g., Florin-Lechner et al., 1996). By shifting the
balance between these two ring modes (Berridge and Waterhouse,
2003), the LC modulates arousal levels (Aston-Jones and Cohen, 2005;
Chen and Sara, 2007, see Section 2) and supports memory encoding
(Klukowski and Harley, 1994; Mather et al., 2016; Sara, 2015, 2009, see
Section 2). Thalamic alpha wave strength, which has been linked to
cortical inhibition and desynchronization as well as selective attention
(Jensen and Mazaheri, 2010), have also been linked to shifts between
phasic and tonic ring (McCormick, 1989).
Thus, the LC might facilitate information processing overall
(Rodenkirch et al., 2019) and in particular the processing of
high-priority stimuli over less-prioritised stimuli (Mather et al., 2016).
Although differences in LC ring have not yet been systematically
analysed in healthy aging and disease, a disturbance, such as increased
noradrenergic levels during the early stages of AD (Palmer et al., 1987)
and a decline in LC connectivity in healthy aging (Langley et al., 2022),
may impair the ability of the LC to modulate downstream targets and
functions such as corticothalamic waves and vice versa: As reviewed by
Chalermpalanupp and colleagues (Chalermpalanupap et al., 2017), re-
sults from a tau pathology rat mouse model suggested that increasing
noradrenergic levels later in the disease might alleviate tau-mediated
pathology.
2.2. Role in neuroinammation and neurovascular system
The locus coeruleus-noradrenergic (LC-NA) system plays a major
role in modulating neuroinammation (Evans et al., 2022) and the
neurovascular system (Giorgi et al., 2020).
NA released by LC neurons play a role in the inammatory responses
through adrenergic receptors (Feinstein et al., 2002) as it can regulate
the expression of proinammatory cytokines (Flierl et al., 2009; Li et al.,
2015), but it can also increase the peripheral inammatory response
(Grisanti et al., 2010; Kavelaars et al., 1997; Spengler et al., 1990). NA
also regulates the activity of T cells and microglia (Kahn et al., 1985;
ONeill and Harkin, 2018; Tanaka et al., 2002).
LC-lesioned animal models show neuronal damage, increased
microglia and inammatory response (Bharani et al., 2017; Song et al.,
2019a, 2019b) and the anti-inammatory effects of LC-NA system play
also a role in neurodegenerative diseases: direct stimulation of the
LC-NA system via vagus nerve stimulation (VNS) reduced both inam-
matory activation and a-synuclein accumulation in LC-lesioned rats
(Farrand et al., 2017) and In the presence of amyloid pathology, lower
NA levels in the LC target areas might impair the microglia in those
areas, therefore precluding amyloid to be cleared or surrounded by it
(Heneka et al., 2010). Therefore, NA can be an endogenous
anti-inammatory agent (Feinstein et al., 2002) and a loss in LC neurons
and consequent alterations of NA signalling could further exacerbate the
inammatory component of AD (Kelly et al., 2019) and PD (Qin et al.,
2007). This modulation helps maintain a balanced immune environment
in the brain. Moreover, it inuences the cerebral blood ow (CBF) via
the modulation of the neurovascular unit (NVU), a group of cells that are
involved in the mechanism of coupling between energy demand and
cerebral blood ow (Iadecola, 2017): in the neurovascular system,
LC-NA induces vasoconstriction, resulting in a global reduction of ce-
rebral blood ow (CBF) and redistribution of blood primarily to acti-
vated brain regions (Bekar et al., 2012).
Furthermore, the LC-NA plays a signicant role in maintaining the
homeostasis of the blood-brain barrier (BBB) (Erd˝
o et al., 2017) as it
controls the expression of tight junctions (TJs) (Kalinin et al., 2006) and
consequently the permeability of the BBB to water and solutes (Thomsen
et al., 2017).
3. Cognitive functions
Recently, several reviews on noradrenergic involvement in cognitive
function have been published, most notably a review by Gina Poe and
colleagues providing a comprehensive overview of LC embryonal
development, anatomy and function (Poe et al., 2020). Here we provide
a brief overview of recent reviews on LC function and cognition The LC
is tightly linked to sleep architecture: Decreases in LC activity (Foote
et al., 1980) and noradrenergic levels (Kalen et al., 1989) are linked to
the transition from wakefulness to sleep where the LC is virtually silent
during REM sleep (Foote et al., 1980). Similarly, increases in LC activity
precede unprovoked rousing from sleep (Aston-Jones and Bloom, 1981).
Increasing noradrenergic levels pharmacologically shortens REM sleep
and prolongs wakefulness across species (De Sarro et al., 1987; Spiegel
and DeVos, 1980). Interestingly, LC activity during NREM sleep has also
been linked to memory consolidation and sleep spindle activity (Kjaerby
et al., 2022; Osorio-Forero et al., 2021). Several afferents to and effer-
ents from sleep promiting regions enable this tight involvement in sleep
(Lew et al., 2021; Samuels and Szabadi, 2008; Saper and Fuller, 2017).
For comprehensive reviews on the topic see Van Egroo et al. (2022) and
Osorio-Forero et al. (2021).
Through its links to the entire neocortex (Szabadi, 2013), the LC has
been tightly linked to arousal and attention. As described by the GANE
model, the LC is capable of increasing the neuronal activity required for
the processing of currently prioritized stimuli while suppressing
neuronal activity of other brain regions (Mather et al., 2016). The
complex activity of the LC acts to coordinate adaptive neural dynamics
as recently reviewed by (Wainstein et al., 2022). Another theory on the
LCs involvement in attentional modulation proposes that the LC mod-
ulates the magnitude of the cortical alpha wave, which is a measure for
disengagement, through a thalamic gating mechanism. This allows for
optimal sensory processing (Dahl et al., 2022). Although it has been
shown in vitro that the LC is capable of inactivating about 90% of rat
midbrain cholinergic neurons (Williams and Reiner, 1993), which also
are essential to maintaining attention (Knudsen, 2011), the behavioural
consequence of such inactivation is currently unknown. In humans,
higher connectivity between the LC and nucleus basalis of Meinert has
been linked to more dynamic global shifts in cortical functional con-
nectivity states (Taylor et al., 2022). For a concise review on the
noradrenergic-cholinergic interaction see Slater et al. (2022).
Through its dopaminergic (Kempadoo et al., 2016; Takeuchi et al.,
2016) and noradrenergic (Guo and Li, 2007; OMalley et al., 1998)
connections to the hippocampus, the LC is shown to be crucial for
memory consolidation (Duszkiewicz et al., 2019; Pintus et al., 2018;
Takeuchi et al., 2016; Titulaer et al., 2021) and for novelty (Takeuchi
et al., 2016). It has been hypothesized that while the LC is important for
the processing and consolidation of intense, rst-time novel experiences,
the VTA is involved in semantic novelty of objects of a known category
(Duszkiewicz et al., 2019). It has also suggested that the dopaminergic
and noradrenergic system are important for rule-shifting during cogni-
tive tasks (Pajkossy et al., 2018). For a comprehensive overview over
noradrenergic projections see Szabadi (2013).
4. Noradrenergic dysfunction in healthy ageing
From an early age on, the LC accumulates tau (Harley et al., 2021)
Mounting evidence has shown that high LC integrity correlates well
with cognitive performance in older age (Bachman et al., 2021; Dahl
et al., 2020; Parent et al., 2022), which is not surprising giving its vast
cortical connections (Szabadi, 2013), its crucial role in cognition (Poe
et al., 2020), and it its protective role against neuroinammation (Giorgi
et al., 2020; McNamee et al., 2010).
4.1. Age-related differences in functional and structural LC connectivity
The functional and structural connectivity of the NA system in aging
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
4
and AD has recently gained more interest as a means for understanding
the role of the LC-NA system in aging and age-related cognitive decline.
Liebe and colleagues (Liebe et al., 2022) investigated both the
structural and functional connectivity of the LC using 7 T MRI and found
that brain areas with higher structural connection to the LC also showed
increased resting state functional connectivity to the LC. These include
the thalamus, ventral diencephalon, basal ganglia, motor cortex, cere-
bellum, amygdala, nucleus accumbens, and temporal brain areas
including the EC, presubiculum, and hippocampus (Liebe et al., 2022).
However, this link diminishes with age. Interestingly, connectivity also
was linked to subjective anxiety and alertness (Liebe et al., 2022).
High-resolution diffusion tensor imaging-based tractography (DTI)
showed that compared to younger adults, older adults showed reduced
integrity of the central tegmental tract (CTT), which branches into the
thalamus and rostral LC (Langley et al., 2022). Reduced integrity of the
CTT also correlated with verbal memory delayed recall scores, measured
by Rey- Auditory Verbal Learning Test (RAVLT). Porat and colleagues
(Porat et al., 2022) also found decreased FA in the ascending norad-
renergic bundle in older adults compared to younger adults (Porat et al.,
2022). It should be emphasised that both studies found an increase in LC
FA in older compared with younger adults (Langley et al., 2022; Porat
et al., 2022). These results suggest that the integrity of the LC and its
projections measured using DTI, are independently affected by ageing.
Cross-sectional differences in high-resolution resting state functional
connectivity (rfMRI) in a cohort of 1974 years of age followed a
nonlinear correlation between left LC and most areas of frontal, tem-
poral, parietal regions and the nucleus basalis of Meyner (MBN) (Jacobs
et al., 2018): they showed an increase in connectivity until 25 years of
age followed by a continuous decrease until 60 years of age and a sub-
sequent increase at old ages (Jacobs et al., 2018). In contrast to this,
Song and colleagues (Song et al., 2021) showed an inverted U-shape
functional connectivity between the LC and visual, sensory and auditory
cortices and a U-shape correlation to frontal areas. The differences may
be explicable by a 10x larger sample size used in the Song study.
Moreover, it was shown that cross-sectionally, older adults showed
stronger functional connectivity between LC and visual processing re-
gions (Lee et al., 2020). Additionally, cross-sectionally, in older adults,
there is a decrease in LC-frontoparietal functional connectivity (Lee
et al., 2018) and a decrease in functional connectivity between the LC
and areas linked to the salience network (Lee et al., 2020) while the
functional activity between the salience and the frontoparietal network
increases (Lee et al., 2020). However, no differences in
LC-parahippocampal functional connectivity were found (Lee et al.,
2018). These results suggest that in older adults attentional selectivity,
which relies on frontoparietal regions is reduced, while stimuli
perception is intact.
It remains to be determined whether these differences underlie dif-
ferences in LC MRI signal, driven in part by age-related differences in LC
NM (Keren et al., 2015), and not accounting for additional physiological
measures, such as the extent of tau pathology or differential imaging
techniques and parameters across studies (Liu et al., 2017).
4.2. The role of tau and amyloid on LC function
Tau is a small soluble protein with a exible structure that helps with
microtubule stabilisation (Avila et al., 2016). Accumulation of misfolded
tau tangles, one hallmark feature of AD pathology, may occur in the LC
already at an early stage in life.
As reviewed by Harley and colleagues (Harley et al., 2021), even at
adolescence, most brains have pre-tangles, a precursor to tau tangles, in
the LC, that might spread to other regions such as the transentorhinal
(TEC) and/or entorhinal cortices (EC) (Harley et al., 2021). These
AT8-positive, misfolded pre-tangle tau seeds spreading from the LC
might be the origin of tau pathology in AD (Braak et al., 2011; Stratmann
et al., 2016).
This has also been shown in animals, where tau pathology in AD mice
models spread from LC to the forebrain and other regions affected by tau
pathology early in human AD thus potentially recapitulating tau human
pathology spread (Kang et al., 2019). However, another group has
theorised that tau pathology originates in the EC and then spreads to the
LC (Kaufman et al., 2018).
Memory symptoms typically appear when tau tangles reach the
hippocampus, temporal insular and association cortex at Braak stage III-
IV in AD (Therriault et al., 2022), which corresponds to tau spread at
very old age in healthy ageing.
Aß plaques are the other hallmark feature of AD. At nanomolar
concentration, Aß might regulate cellular activation (Turner et al.,
2003). However, when accumulated, Aß plaques hinder normal cell
functioning and cause hyperactivity (for review see Hector and Brouil-
lette, 2021). In the absence of Aβ, episodic-memory performance is best
explained by tau burden in the areas dened by Braak staging as I-III, the
TER/EC, suggesting a link between tau and cognitive decline (Maass
et al., 2018).
In a group composed of healthy older adults, patients with MCI and
AD patients, Jacobs and colleagues (Jacobs et al., 2021a) showed a
negative correlation between LC positron emission tomography (PET)
signal intensity, a measure of LC integrity and tau deposition in the EC as
well as the medial and lateral temporal regions and medial
parietal-frontal regions. For healthy participants, this correlation
remained only in the EC. In a subset with elevated Aβ levels, LC integrity
was associated with greater tau pathology outside the temporal lobe,
extra medial-temporal lobe, lateral temporal and medial-lateral parietal
and frontal regions. The autopsy data conrmed these results as LC tau
tangle density correlated with Braak stages and in the context of
elevated Aβ burden, also with steeper retrospective memory decline,
supporting the role of LC in tau pathology progression (Jacobs et al.,
2021a).
In the context of Aβ burden, Ciampa and colleagues (Ciampa et al.,
2022) showed that higher LC catecholamine synthesis capacity,
measured with PET in older adults, was related to lower tau in the
temporal lobe.
Moreover, adjusting for Aß status, catecholamine synthesis in the LC
but not in the raphe, midbrain, and striatum, was associated with lower
rates of tau accumulation over time and with better-than-expected
memory performance adjusting for individuals tau burden (Ciampa
et al., 2022).
Low LC catecholamine synthesis is also related to vulnerability to
affective dysregulation, measured as high neuroticism and depression,
and tau PET burden in the amygdala (Parent et al., 2022). Interestingly,
low conscientiousness and high neuroticism are indirectly related to
increased tau burden in the amygdala, via their association with low LC
catecholamine synthesis capacity (Parent et al., 2022). Therefore LC
vulnerability could be involved in affective dysregulation and
neuroticism.
These results are strong indicators for the LC and the brainstem as
among the regions of earliest tau accumulation, if not the earliest,
showing a prevalent involvement of tau in LC dysregulation.
4.3. LC integrity is linked to cognitive performance in older age
As explained in Section 1, the LC contains high levels of NM, which
permits the visualisation of the LC using NM-sensitive MRI (for a review
see Betts et al., 2019b). Normalised NM-sensitive MRI contrast values
inside the LC have become a popular method to assess the integrityof
the LC in vivo (Betts et al., 2019b). It has been shown that LC MRI
contrast (consistent with an increase in neuromelanin) increases up until
about 60 years of age and then decreases thereafter (Liu et al., 2019;
Zecca et al., 2004). Several studies have also now shown, LC MRI
contrast is related to cognitive function. In older participants, a positive
association has been found between LC MRI contrast and episodic
memory performance (Dahl et al., 2022; H¨
ammerer et al., 2018), as well
as a delayed memory score (RAVLT) (Dahl et al., 2019). Overall LC
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
5
integrity has been linked to a cognition score composed by education,
occupational attainment and verbal IQ, which was particularly pro-
nounced in the rostral subsection, and reading intelligence score (Cle-
wett et al., 2016). In particular, the rostral subsection has also been
linked to delayed memory performance (Dahl et al., 2019) and response
inhibition (Tomassini et al., 2022) as well as attention, delayed memory
performance and processing speed (Calarco et al., 2022). In addition, LC
integrity has also been associated with cortical thickness (Bachman
et al., 2021).
In older age, signicant variability in LC integrity has been observed
(Betts et al., 2019a) and age-related differences in LC have not been
consistently observed in cognitively intact older adults in the range of
6080 (Betts et al., 2019b, 2017; Giorgi et al., 2021).
It remains to be determined whether these differences in the MRI
signal of the LCs are due in part to age-related differences in LC neu-
romelanin (Keren et al., 2015; Zucca et al., 2006), although additional
physiological measures such as the extent of tau pathology or different
imaging techniques and parameters are not considered in the various
studies (Liu et al., 2017).
4.4. The effect on neuroinammation and neurovascular system
Ageing and age-related disorders are characterized by elevated pro-
inammatory markers (Franceschi et al., 2000), which contribute to the
progression of neuropathology. Systemic inammation further exacer-
bates this process by triggering the activation of microglia, which
inltrate the brain through the blood-brain barrier (Bettcher and
Kramer, 2014; Heneka et al., 2015; Perry, 2010; Perry et al., 2007). The
presence of pro-inammatory cytokines in older adults has been asso-
ciated with cognitive impairment, highlighting the detrimental effects of
inammation on brain function (Cunningham et al., 2009; Cunningham
and Hennessy, 2015; Franceschi and Campisi, 2014; Holmes, 2013; Qin
et al., 2007; Simen et al., 2011).
Moreover, the degeneration of noradrenergic neurons in the LC and
the natural process compromise the integrity of the blood-brain barrier
(BBB) (Kalinin et al., 2006; Luissint et al., 2012). This compromised BBB
amplies the effects of systemic inammation, allowing inammatory
substances to inltrate the brain more easily. In aged rats with LC-NA
lesions, the inammatory response is intensied, and there are notable
alterations in brain-derived neurotrophic factor (BDNF) levels, showing
the role of the LC/NA system in modulating neuroinammation (Bhar-
ani et al., 2017).
Additionally, microglial dysfunction, combined with the compro-
mised BBB (Hussain et al., 2021) and reduced tissue perfusion in aging,
plays a signicant role in impairing synaptic plasticity (Erd˝
o et al.,
2017), that is an important mechanism for memory formation which
becomes impaired in aging and neurodegenerative diseases (Blau et al.,
2012).
5. The noradrenergic system and neurodegenerative diseases
5.1. Alzheimers disease
AD is a neurodegenerative disease which constitutes the vast ma-
jority of the cases of dementia in the older population (Evans et al.,
2022) with a prevalence in Europe of around 1%, an incidence that is
expected to triple by 2050 (Prince et al., 2015). AD can be considered a
proteinopathy as it is dened by the presence of amyloid plaques (Braak
and Braak, 1991) and tau aggregates called neurobrillary tangles
(NFTs) (Goedert, 1993).
NA plays a complex role in the pathogenesis and progression of AD,
encompassing both protective and contributory effects. NA exerts its
protective effect by destabilising Aβ protobril, impeding their assembly
into insoluble plaques (Zou et al., 2019). Moreover, it might have a
protective effect against oxidative stress caused by amyloid, by stimu-
lating CAMP production through b-adrenergic receptors, resulting in the
activation of the nerve growth factor (NGF) or brain-derived neuro-
trophic factor (BDNF) (Counts and Mufson, 2010) and regulates
inammation through microglia modulation (Mori et al., 2002; Sugama
and Kakinuma, 2021).
However, it might also be implicated in the further progression of AD
pathology as compensatory mechanism for LC neuronal loss during
aging and AD, like changes in the receptors and increased excitability of
remaining NA neurons might be implicated in the aetiology or further
progression of Aß pathology.
NA signalling is implicated in amyloid toxic effects through
α
2-
adrenergic receptors. Amyloid activated the glycogen synthase kinase
(GSK), enzyme responsible for tau phosphorylation, by binding on an
allosteric site of
α
2-adrenergic receptor, therefore redirecting NA sig-
nalling. The activation of GSK can be facilitated by concurrent binding
of amyloid on the allosteric site and NA on the main site of the receptor.
In fact, when this happens, amyloid can exert its effect at a concentration
as low as 1% of what is typically required for amyloid to exert the effect
alone.
The increase of NA turnover and thus the production of the toxic NA
metabolite 3,4-dihydroxyphenylglycolaldehyde (DOPEGAL) (Wein-
shenker, 2018), might play an important role in tau LC-derived toxicity
(Burke et al., 1999). DOPEGAL is the product of NA metabolism by the
enzyme monoamine oxidase A (MAO-A) (see Section 1. for an overview
on NA biosynthesis). It can be considered toxic as, in contrast to NE, it
promotes tau aggregation and propagation (Kang et al., 2022, 2019) and
induces tau-mediated neuronal cell apoptosis (Burke et al., 1998). Mice
lacking DBH, the protein that converts dopamine to NA, and therefore
the ability to produce NA, showed reduced tau pathology. Models in
which the asparagine endopeptidase (AEP), which is necessary for tau
cleavage, was knocked out, showed attenuated LC neuronal degenera-
tion in the presence of DOPEGAL (Kang et al., 2020). A further link
between AD and tau, mediated by DOPEGAL has been shown in mice
and AD brain tissue. Apoe4 allele induced DOPEGAL tau-mediated
toxicity, while Apoe3 prevents it by binding tau and blocking the
cleavage into misfolded tau (Kang et al., 2021). Tau neurotoxicity has
also been shown to lead to hippocampal atrophy and cognitive decits
(Kang et al., 2021). Together these studies show that NA, but specically
its metabolite DOPEGAL, have a toxic neuronal effect, which is mediated
by tau (for a review Kang et al., 2020). DOPEGAL is then processed to
MHPG, the major NA metabolites in the brain (Burke et al., 1999), which
has been also associated with compensatory mechanisms of LC surviving
neurons and AD progression (Hoogendijk et al., 1995; Nakamura and
Sakaguchi, 1990). Higher NA turnover (MHPG:NA ratio) inversely
correlated with LC neuronal cells (Hoogendijk et al., 1995). Higher CSF
levels of MHPG were also associated with greater tau and amyloid
concentration (Riphagen et al., 2021) in the CSF and with lower cortical
thickness in mild stages of AD (van Hooren et al., 2021). Finally, the
presence of MHPG facilitates tau spreading in AD mice model (Koppel
et al., 2019) suggesting a parallel with the DOPEGAL action of neuronal
apoptosis and pathology enhancement in the presence of tau pathology.
In the presence of amyloid pathology, lower NA levels in the LC
target areas might impair the microglia in those areas, therefore pre-
cluding amyloid to be cleared or surrounded by it (Heneka et al., 2010).
Therefore, NA can be an endogenous anti-inammatory agent (Feinstein
et al., 2002) and a loss in LC neurons and consequent alterations of NA
signalling could further exacerbate the inammatory component of AD
pathology (Kelly et al., 2019).
5.2. The LC-NA system in the context of AD
The loss of LC neurons in AD occurs at early stages (Betts et al.,
2019a, 2019b; Braak and Braak, 1991), is most pronounced in the rostral
and middle LC portion (Beardmore et al., 2021; Betts et al., 2019a;
Theolas et al., 2017; Tomlinson et al., 1981) and appears to correlate
with disease duration (Zarow et al., 2003) and disease severity (Cassidy
et al., 2022; Olivieri et al., 2019). In fact, a post-mortem analysis showed
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
6
a decrease in LC volume by circa 8% for every Braak stage (Theolas
et al., 2017). Reduced LC volume during AD is also supported by in vivo
studies showing reduced LC MRI contrast in AD compared to healthy
controls (Betts et al., 2019a; Cassidy et al., 2022; Li et al., 2022; Olivieri
et al., 2019).
Animal studies suggest that the NA system might interact with Aβ
pathology. Catecholamines, including NA showed a dose-dependent
inhibitory effect on the aggregation of amyloid brils potentially by
destabilising amyloid protobrils in vitro (Huong et al., 2010). In rats,
the LC might protect against amyloid pathology as lesioning of the LC
led to an increase in amyloid and inammation compared to
non-lesioned controls (Kelly et al., 2019). Conversely, decreased NA
levels might impair amyloid clearance thus further propagating amyloid
accumulation (Dobarro et al., 2013; Ni et al., 2006).
As LC integrity and NA production are affected in AD, these results
support the idea that the loss of LC/NA-mediated protection in AD may
enhance the risk of amyloid aggregation in the brain, as the anti-amyloid
aggregatory and inammatory effect of NA might be impaired due to its
reduced levels.
The NA system might also link tau and amyloid pathology in AD:
Wang and colleagues demonstrated in vitro and in an AD mouse model
that amyloid binds hyperactive alpha 2 receptors on an allosteric site
activating a tau phosphorylation cascade (Wang, 2020).
Tau pathology is related to LC integrity and functioning: lower LC
neuron count is associated with higher plasma tau levels (Murray et al.,
2022) and LC integrity is shown to be decreased in tau-positive partic-
ipants (Cassidy et al., 2022) and inversely correlated to tau in the EC in
healthy participants (Jacobs et al., 2021a). As tau induces neuronal
hyperexcitability (Crimins et al., 2012; Rudy et al., 2015), this might
also be hypothesised to be the case for LC neurons (Weinshenker, 2018),
facilitating the spread of tau through its axonal pathways, as tau is
capable of trans-synaptic propagation (Kang et al., 2022, 2019). How-
ever, Liu and colleagues (Liu et al., 2021) did not observe an increase in
LC hyperactivity in AD compared to healthy controls with respect to
uorodeoxyglucose (FDG) PET, a tracer used to image metabolic activity
(Liu et al., 2021).
Increased activity of the remaining LC neurons has also been sug-
gested by studies reporting increased NA levels in AD, hypothesising this
increased activation as the cause of increased NA production and
therefore NA levels (Raskind, 1984; Raskind et al., 1997, 1999), how-
ever, this has not been seen consistently in the literature. In fact, a recent
meta-analysis focusing on NA and MHPG data in AD found a trend to-
wards lower levels of NA in CSF markers and signicantly increased
MHPG CSF markers in participants with AD dementia compared to
healthy controls (Lancini et al., 2023).
The discrepancies between studies on NA levels in participants with
AD may have several causes, such as the use of different diagnosis
criteria, methods of severity calculation, as well as small sample sizes
that may limit statistical power. Furthermore, although the meta-
analysis did not detect any differences between studies caused by con-
founding variables, inter- and intra-laboratory differences may have
played a role: differences in sample storage, handling and equipment
could have inuenced the results, despite the use of similar or identical
protocols.
As discussed by Lancini and colleagues (Lancini et al., 2023), in the
presence of amyloid and tau, NA metabolite MHPG seems to be a more
sensitive measure of AD symptoms characterisation, therefore altered
metabolism, more than production might play a substantial role in AD.
Higher CSF levels of MHPG, are associated with higher levels of p-tau
(Jacobs et al., 2021b; Riphagen et al., 2021) and higher amyloid-β 42
under high levels of inammation in CSF (Riphagen et al., 2021) and
lower cortical thickness in mild stages of AD (van Hooren et al., 2021) in
humans. Additionally, a structural equation model (SEM) analysis
showed a strong link between neuropsychiatric symptoms and MHPG
and p-tau, suggesting that noradrenergic dysfunction is coupled to AD
pathology and clinical symptoms (Jacobs et al., 2021b). Also, the
increase in the noradrenergic metabolite DOPEGAL has been shown to
promote tau aggregation and propagation in mice models (Kang et al.,
2022, 2019) and to induce tau-mediated neuronal cell apoptosis in AD
postmortem human brain tissue (Burke et al., 1999, 1998, 1997).
5.3. LC neurodegeneration and AD symptoms
In line with accumulating evidence of NA being intricately linked
with AD pathology, links between an LC-NA system and cognitive and
clinical symptoms of AD dementia are increasingly reported. General
cognition as measured by the Mini-Mental State Examination (MMSE)
and Montreal Cognitive Assessment (MoCA) and memory performance
has been shown to be positively correlated to LC integrity (Li et al.,
2022) and CSF noradrenergic levels (Elrod et al., 1997). Also, CSF MHPG
is correlated to memory decits in participants with Subjective cognitive
decline (SCD), MCI and AD (Jacobs et al., 2021b).
LC neuronal loss results in the dysregulation of circadian rhythms
early in the disease process (Van Egroo et al., 2019) and in fact sleep
impairment is another symptom that occurs early in AD and can also
precede cognitive symptoms (Ehrenberg et al., 2018; for a review see
Van Egroo et al., 2022). Critically, as studies in animal models showed
an intact sleep architecture is required to clear neurotoxic waste prod-
ucts from the brain, such as Aß (Xie et al., 2013), and that sleep loss
affected tau (Barth´
elemy et al., 2020), sleep disruption might accelerate
the propagation of AD pathology starting a vicious cycle (Tekieh et al.,
2022).
5.4. The effect on neuroinammation and neurovascular system
In AD, degeneration to the LC-NA may lead to the breakdown of the
blood-brain barrier (BBB) and microglia activation, impairing amyloid
clearance, exacerbating amyloid accumulation, promoting neuro-
inammation, and causing neuronal death. Decreased levels of NA
disrupt the regulation of pro-inammatory cytokines, leading to further
LC cell loss creating a negative feedback loop (I et al., 2014).
In-vivo, neuroinammation can be assessed with PET tracers that
bind mitochondrial translocator protein (TSPO), that is overexpressed in
cells with activated microglia (Werry et al., 2019). In AD participants,
TSPO tracer binding was found to be increased (Hamelin et al., 2016;
Kreisl et al., 2016), indicating inammation and microglial activation
(Werry et al., 2019) and to be correlated with amyloid pathology (Fan
et al., 2015) and severity of disease (Kreisl et al., 2013). Controls and
MCI did not show elevated TSPO showing that inammation occurs after
the conversion to MCI into AD (Kreisl et al., 2013). This nding suggests
that elevated TSPO expression may reect a more pronounced LC-NA
dysregulation and could therefore serve as possible marker of disease
progression in longitudinal studies, despite to this date no study corre-
lated TSPO measure with LC markers (Giorgi et al., 2020). In line with
the PET study from Fan and colleagues (Fan et al., 2015), postmortem
studies have also demonstrated the presence of activated microglia
surrounding amyloid plaques in AD brains (Taipa et al., 2018).
NA also plays a crucial role in clearing Aβ by microglia and sup-
pressing Aβ-induced cytokine production, where a reduction in NA
levels may contribute to insufcient suppression of pro-inammatory
mediators, contributing to AD progression (Heneka et al., 2010; Kali-
nin et al., 2007; Kong et al., 2010). Experiments in mice have demon-
strated that NA β-adrenergic stimulation protects cortical and LC cells
from Aβ-induced cell loss (Evans et al., 2020). Moreover, Both 5xFAD
mice (Kalinin et al., 2012) and P301S tau mice (Chalermpalanupap
et al., 2018) show microglial and astrocyte activation in the LC
compared to wild-type mice. In mice, NA maintains the clearance of Aβ
via microglia and suppresses Aβ-induced cytokine production as
β-adrenergic stimulation by NA protects cortical and LC cells from
Aβ-induced cell loss (Evans et al., 2020). These studies collectively
suggest that a reduction in NA levels may contribute to the progression
of AD and Aβ accumulation, making NA a potential target for
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
7
pharmacological treatment.
5.5. Parkinsons disease
PD - similar to AD - is one of the most common neurodegenerative
diseases worldwide, with an incidence rate for women and men of
respectively 37.55 and 61.21 per 100,000 person-years (Hirsch et al.,
2016). The main hallmark of the disease is a signicant reduction of
dopaminergic cells in the substantia nigra pars compacta (SNpc) and the
presence of
α
-synuclein aggregates, called Lewy Bodies (Dickson et al.,
2008). Recently, the role of NA in the development and progression of
idiopathic PD has drawn increased attention (Espay et al., 2014;
Paredes-Rodriguez et al., 2020).
Cellular accumulation of misfolded alpha-synuclein is one of the
hallmark features of PD (Atik et al., 2016). Spectroscopic and calori-
metric experiments suggest that NA stabilizes alpha-synuclein in inter-
mediate states into stable, cytotoxic species (Singh and Bhat, 2019) thus
possibly contributing to the generation of alpha-synuclein. Some the-
ories propose that alpha-synuclein misfolds in the gut and then spreads
to the brain (Berg et al., 2021; Braak et al., 2003; Engelender and
Isacson, 2017). A recent mouse model paper suggests a protective
noradrenergic mechanism against this spread: acute chemogenic
NA-depletion in PD Mouse models was linked to enteric inammation
and increased enteric alpha-synuclein levels 2 weeks later and Sub-
stantia Nigra neuronal degeneration as well as constipation and motor
decits 35 months later (Song et al., 2023). It has been suggested that
already in healthy elderly, higher LC neuronal density might be pro-
tective against deleterious effects of brainstem Lewy body numbers on
cognition (Wilson et al., 2013). Another PD mouse model suggests that
the noradrenergic system might also be protective against dopaminergic
neuronal depletion despite increasing alpha-synuclein aggregation and
microgliosis (Jovanovic et al., 2022).
In humans, the LC might be benecial against PD pathology even
before symptom onset: already in healthy subjects, higher LC neuronal
density diminishes the association between brainstem Lewy body
numbers (Wilson et al., 2013).
5.6. The LC-NA system in the context of PD
Comparative studies of subcortical nuclei found a high loss of LC
neurons in idiopathic PD, which is greater than the neuronal loss
observed in the nucleus basalis of Meynert and SNpc (Huynh et al.,
2021; Zarow et al., 2003). LC volume is found to be reduced (Hwang
et al., 2022), together with NA and MHPG CSF levels (Lancini et al.,
2023), and NM-sensitive MRI indicates subregion-specic degeneration
of the caudal and middle sub-portion of the LC (Doppler et al., 2021;
Madelung et al., 2022; Ye et al., 2022). Other studies showed an overall
decrease in LC MRI contrast (Schwarz et al., 2017; Wang et al., 2018). A
meta-analysis on white matter degeneration in PD found no change (Wei
et al., 2021; Zhang and Burock, 2020) or even improved white matter
integrity (Taylor et al., 2018) in LC tracts compared to healthy volun-
teers suggesting they may be unaffected in PD.
Several hypotheses such as the Braak model, and recently also the
brain-gut hypothesis (Borghammer and Van Den Berge, 2019; Horsager
et al., 2020) integrate the LC into early- to mid-stages of PD. However,
due to the heterogeneity of the disease (Farrow et al., 2022), a parsi-
monious model describing the whole breadth of variability is still
lacking.
As recently reviewed by Lancini and colleagues (Lancini et al., 2023),
NA transporter density, measured with PET NA transporter (NET) tracer,
is decreased in PD suggesting a functional decline to the noradrenergic
system including LC terminals (Doppler et al., 2021) besides structural
degeneration. Additionally, LC- specic
α
-synuclein expressing mouse
models showed degeneration of noradrenergic neurons (Henrich et al.,
2018), changes in the abundance and integrity of the immune system
(Butkovich et al., 2018; Henrich et al., 2018) and Parkinsonian
symptoms (Henrich et al., 2018).
Hyperactivation of remaining noradrenergic neurons in LC occurs in
rats after lesioning the nigrostriatal pathway (Wang et al., 2009), an
effect that was also in line with in vitro results and in a mouse model of
PD (Matschke et al., 2022). Interestingly, in mice stimulation of the
noradrenergic system protects against neuronal depletion but does not
prevent alpha-synuclein aggregation and immune system degeneration,
suggesting an immune system-independent mechanism (Jovanovic
et al., 2022). However, early LC degeneration is also linked to non-motor
symptoms such as anxiety and depression, that are typical in early stages
of the disease (Bremner et al., 1996; Remy et al., 2005). In LC-lesioned
animal models of PD, neuronal damage leads to increased microglia and
greater inammatory response to alpha-synuclein (Bharani et al., 2017;
Song et al., 2019b). Conversely, indirect stimulation of the LC-NA sys-
tem via invasive vagus nerve stimulation (iVNS) in rats, led to reduced
inammation as well as a-synuclein accumulation and increased loco-
motion in LC-lesioned rats (Farrand et al., 2020, 2017). Considering the
role that chronic inammation plays in the pathogenesis and further
progression of PD (Qin et al., 2007), these results indicate that LC-NA
system dysregulation may be involved in this process. To our knowl-
edge, in humans, there have been no correlations between noradren-
ergic levels in PD and inammation markers, which would be an
important step for future studies.
5.7. LC degeneration correlates with PD symptoms
NM-sensitive MRI correlates with a number of different PD symp-
toms. Rostral LC integrity has been linked to motor symptoms ipsilat-
erally (Ye et al., 2022) and lower LC contrast has also been linked to
self-evaluation of motor accuracy (Hezemans et al., 2022). Several
memory tests such as MoCA and Berlin model of intelligence structure
(BIS) scores (Ye et al., 2022), as well as short-term verbal and numerical
memory performance (Prasuhn et al., 2021), have been shown to be
associated with higher LC MRI contrast. LC integrity in PD is also related
to cognition: scores on episodic and short-term memory as well as lexical
uency (Prasuhn et al., 2021), response inhibition (Ye et al., 2021) and
trail-making test (TMT) results (Li et al., 2019) all correlate with LC
integrity. Additionally, left caudal and whole LC integrity have been
linked to orthostatic blood pressure (Madelung et al., 2022) and sleep
disruption (Doppler et al., 2021): sleep cyclic alternative patterns (CAP)
are diminished in PD and the rate of CAP was shown to be correlated
with noradrenergic PET binding only in patients but not in healthy
control participants while the duration of CAP subphases is altered in PD
patients as well. Finally, increased NA transporter binding and
decreased DA transporter binding in PD were linked to increased anxiety
(Carey et al., 2021).
5.8. The effect on neuroinammation and neurovascular system
In PD, LC-NA degeneration contributes to increased neuro-
inammation through reactive microglia and oxidative stress (Giorgi
et al., 2020). Animal studies have demonstrated that lesioning of the LC
leads to heightened activation of microglial cells (af Bjerk´
en et al., 2019)
and inammation in SN dopaminergic cells and when combined with
systemic inammation, LC degeneration exacerbates degeneration in
the nigrostriatal pathway, hippocampus, and motor cortex in PD models
(Bharani et al., 2017; Herrera et al., 2000; Song et al., 2019b, 2019a).
The presence of inammation in dopaminergic areas in PD has been
shown in vivo, as PET studies using tracers of microglia-activation
showed increased binding in basal ganglia, substantia nigra, and
fronto-temporal cortex in individuals with PD (Edison et al., 2013;
Gerhard et al., 2006; Iannaccone et al., 2013) and individuals diagnosed
with PD and dementia (PDD) displayed a signicantly more extensive
microglia activation pattern compared to individuals with PD (Edison
et al., 2013). This result provides further evidence for the involvement of
neuroinammatory responses in exacerbating the disease despite to this
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
8
date no study correlated this measure with LC markers (Giorgi et al.,
2020).
In PD rat models, indirect stimulation of LC-NA activity through
vagus nerve stimulation has shown potential in reducing neuro-
inammation and protecting dopaminergic cells (Farrand et al., 2020),
whilst administration of NA in dopaminergic neurons reduces the pro-
duction of ROS and rate of neurodegeneration (Butkovich et al., 2018).
Animal studies using models of PD have provided further support for
the role of LC degeneration in neuroinammation and oxidative stress
(see Giorgi et al., 2020 for a review). The reverse inuence of inam-
mation has also been investigated: lipopolysaccharide-induced inam-
mation in DSP-4-induced LC lesion mice exacerbated degeneration to
the nigrostriatal pathway, hippocampus, and motor cortex and
increased microglial activation (Bharani et al., 2017; Song et al., 2019a,
2019b).
6. Methods to investigate the LC-NA system
The currently most established methods to investigate the structure
and function of the LC-NA system include structural and functional
magnetic resonance (sMRI and fMRI), PET, electroencephalography
(EEG) (for a review see Nieuwenhuis et al., 2011, 2005), and pupill-
ometry (for a review see Viglione et al., 2023).
Sasaki and colleagues demonstrated the possibility of measuring LC
cell density using a neuromelanin-sensitive MRI protocol (Sasaki et al.,
2006). In that study, brain images from healthy people and people with
PD were acquired with a T1-weighted Turbo Spin Echo (TSE) protocol, a
neuromelanin-sensitive sequence. In healthy adults, the contrast in-
tensity in the LC resulted in high signal intensity in areas that were
consistent with the anatomical location of the two NM-rich regions LC
and SN. The anatomical precision of this technique has been further
conrmed in postmortem tissue demonstrating colocalization between
MRI signal and NM-containing neurons (Cassidy et al., 2019; Keren
et al., 2015). Different neuromelanin-sensitive MRI sequences have
since then been used as a non-invasive ‘integrity measure for simulta-
neous assessment of the LC (for a review see Betts et al., 2019b; Galgani
et al., 2020; Liu et al., 2017; Sulzer et al., 2018) and SN (for a review see
Sulzer et al., 2018). However, to date, the mechanisms underlying MRI
contrast in the LC remain unclear (see Betts et al., 2019b for a
comprehensive review). LC tracts have been studied in vivo (Liebe et al.,
2022) using magnetic resonance tractography (Jbabdi and
Johansen-Berg, 2011), a method that investigates brain white matter
tracts and through which biologically accurate measures of bre con-
nectivity can be obtained (Smith et al., 2015). DTI can be signicantly
affected by several factors, namely low-resolution MRI, crossing-bres,
noise, and distortions. These inuences introduce complexities and po-
tential limitations, as they may impede the accurate delineation of the
LC. Furthermore, in the context of neurodegenerative disorders such as
AD and PD, the challenges associated with DTI LC measures become
more pronounced, due to the potential reduction in LCs structural
integrity, which adds an additional layer of complexity to the assessment
process (Zhang et al., 2020; Zhang and Burock, 2020).
Changes in brain activation during tasks are approximated using
functional fMRI, a sequence that measures the changes in blood-oxygen
levels in the brain that occur in response to neuronal activity (Buxton
et al., 2004). These images are registered on structural MRI images to
allow a correct identication of the area that was active during the task,
which is especially important in the case of LC activity (Turker et al.,
2021). However, investigating LC activity with fMRI presents challenges
due to its size, location, and inter-individual variability.
The LC shows signicant inter-individual variability, particularly in
the context of aging (Liu et al., 2019), making it challenging to establish
a consistent region of interest (ROI) across individuals (Turker et al.,
2021). Moreover, limited resolution of standard MRI techniques can
lead to partial volume effects, where LC voxels contain non-LC tissue
(Cassidy et al., 2022). The proximity of the LC to large sources of
physiological noise, such as the CSF and the 4th ventricle, means that
spatial smoothing during preprocessing may introduce additional
physiological noise (Turker et al., 2021). Taken together, variations in
voxel size and smoothing kernel across studies contribute to
between-study variability, as evident from the heterogeneity of these
measures between studies (Liu et al., 2017). Recently, Yi and colleagues
proposed an optimised spatial transformation pipeline for the LC, along
with a method to quantify the precision of spatial transformations, that
could be of help to increase future comparability across studies (Yi et al.,
2023). Finally, the BOLD signal of the LC can be inuenced by physio-
logical noise, potentially attenuating the fMRI response (Liu et al.,
2017).
PET imaging uses radiotracers that specically bind to sites and
molecules of interest, thus allowing for an assessment of their spatial
distribution in the brain. Of particular interest for the LC-NA system is
the MeNER tracer (Schou et al., 2003), that was successfully used to
quantify NA transporters levels in healthy and PD (García-Lorenzo et al.,
2013; Sommerauer et al., 2018). Both AD and PD are multisystemic
diseases, therefore it is informative to also use PET tracers that allow
visualisation of brain pathology, neuroinammation, cholinergic and
monoamine neurotransmitter systems, synaptic density as well as
metabolism (for a review on AD see Bao et al., 2021; Guti´
errez et al.,
2022, for a review on PD see Prange et al., 2022). PET tau measures of
LC are currently not feasible as PET tau tracer binds to neuromelanin
rather than specically binding to tau protein leading (off-target bind-
ing) (Jacobs et al., 2021a; Lee et al., 2018; Marqui´
e et al., 2015).
The ring activity of the LC-NA system can be detected during wake
using EEG, as the event-related wave P3b, the parietal subcomponent of
the P3 reects phasic activity of the LC-NA system (Nieuwenhuis et al.,
2005; Polich, 2007). LC-NA modulation of sleep (Van Egroo et al., 2022)
can be assessed by concurrent polysomnography and PET Mener in
aging and disease, as sleep instability measured as cyclic alternating
pattern (CAP) had been shown to correlate with NA transporter density
in brainstem in PD (Doppler et al., 2021).
Pupillometry is also used as an indirect measure for LC activation
(Joshi et al., 2016; Szabadi, 2013) as the uctuations in the ring rate of
LC neurons closely parallel changes in pupil diameter (Aston-Jones and
Cohen, 2005; Gilzenrat et al., 2010; Joshi et al., 2016).
Levels of NA and its metabolites in blood (Katunina et al., 2023;
Teunissen et al., 2022) and CSF (David and Malhotra, 2022; Lotankar
et al., 2017) have also been used and are under constant development.
Despite being the closest measure to the central nervous system (CNS),
CSF biomarkers still require specic protocols and a deeper under-
standing of their dynamics in order to be considered reliable measure-
ments of central NA (Lancini et al., 2023).
7. The locus coeruleus as a therapeutic target
7.1. Drug interventions targeting the noradrenergic system
Numerous noradrenergic drugs have been tested to improve cogni-
tion in neurodegenerative diseases:
Methylphenidate is a noradrenergic and dopaminergic reuptake in-
hibitor that is mainly used to treat Attention decit hyperactivity dis-
order (ADHD). In a randomised trial, methylphenidate improved apathy
but not neuropsychiatric symptoms in participants with AD, compared
to placebo (Mintzer et al., 2021) while in participants with MCI, it
improved global cognitive scores and memory after 3 days of daily
intake (Press et al., 2021). In PD mouse models, methylphenidate
improved sleep and daily functioning (Oakes et al., 2021). A recent re-
view showed that methylphenidate improved cognition in a wide vari-
ety of conditions associated with elderly: it improved depressive
symptoms, accelerated post-stroke recovery and improved outcome
measures after stroke (Swartzwelder and Galanos, 2016).
Atomoxetine, a presynaptic NET inhibitor also used in ADHD treat-
ment, did not improve cognition in a 5-week trial with participants with
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
9
MCI but was associated with an increase in noradrenergic plasma and
CSF levels and a reduction in CSF tau and pTau 181 (Levey et al., 2022).
It is also associated with altered protein levels of pathophysiology
related to synaptic and metabolism, glial immunity and
inammation-related proteins as well as an increase in the
Brain-Derived Neurotrophic Factor (BDNF) (Levey et al., 2022). Finally,
atomoxetine as a treatment for MCI has been shown to be associated
with an increase in fMRI functional connectivity between the insula and
the hippocampus (Levey et al., 2022). In PD, atomoxetine plasma con-
centration correlated with uency in an animal category and letter
uency test (Borchert et al., 2019) and improved inhibition in a
stop-signal task (Borchert et al., 2019; OCallaghan et al., 2021).
Clonidine, an alpha-2 receptor agonist commonly used against hy-
pertonia, has been shown in PD to improve movement initiation in a stop
signal task and to modulate the activity of the dorsomedial prefrontal
and anterior cingulate cortices, which are linked to the inhibitory
network (Criaud et al., 2022).
A meta-analysis of 19 randomised controlled trials that performed
interventions in AD patients using noradrenergic drugs found a positive
effect on the MMSE and apathy but no improvement on attention (David
et al., 2022).
As noradrenergic drugs have shown benecial effects in improving
LC-related affected functions in both healthy participants and partici-
pants with diseases, they are a promising tool for potential early or even
later-stage intervention.
7.2. Targeting the noradrenergic system via the vagus nerve
Transcutaneous vagus nerve stimulation (taVNS) offers the possi-
bility of electrical non-invasive stimulation of the cymba conchae of the
external ear, which has nerve bre connections to the vagus nerve
(Peuker and Filler, 2002) and is thought to stimulate the LC-NE system
via the nucleus tractus solitarius (NTS) (Butt et al., 2020; Ruffoli et al.,
2011). IVNS has already been shown to increase the ring rate of LC
neurons (Hulsey et al., 2017) to stimulus-specic cortical plasticity in
the motor cortex (Hulsey et al., 2019) and to a reduction in inamma-
tory markers (for a review see Ludwig et al., 2021). Additionally, iVNS
led to an increase in NA release in brain regions pathologically affected
early in AD and PD, such as the hippocampus, basolateral amygdala and
cortical target areas (Hassert et al., 2004; Hulsey et al., 2019; Manta
et al., 2013). This is consistent with the results of electrical tonic stim-
ulation of the LC, which resulted in higher NA release and increased
activation of the PFC (Florin-Lechner et al., 1996).
As taVNS is less invasive than iVNS, it is a more promising therapy
avenue: results combining taVNS and fMRI have demonstrated the ef-
cacy of taVNS to effectively increase activity in the NTS and LC as well
as in LC-NE projection areas such as the amygdala and hippocampus
(Sclocco et al., 2019; Yakunina et al., 2017). Meanwhile, taVNS is also
being investigated as a potential therapy to improve cognition in early
AD (for a review see Vargas-Caballero et al., 2022) as well as motor and
nonmotor symptoms in PD (Cakmak et al., 2017; Zaehle and Krauel,
2021). Further taVNS studies should take into account interindividual
differences in the integrity of the stimulated LC-NE system to accurately
validate the benet of taVNS for an individual (Ludwig et al., 2021).
7.3. Targeting the noradrenergic system via environmental enrichment
Environmental enrichment is the concept of providing a cognitively
or physically stimulating and challenging environment for humans and
animals, which improves cognition and overall brain health: it has been
shown that mice in an enriched environment display increased cortical
thickness in sensory areas (Engineer et al., 2004) and perform better in
cognitive tasks (Yuan et al., 2012). The number of years of education
(L¨
ovd´
en et al., 2020), a large social circle (Smith et al., 2018) and a
cognitively demanding occupation (Stebbins et al., 2022), all considered
to be an enriched environment, be benecial for cognitive performance
potentially from early childhood onwards (Schoentgen et al., 2020).
The LC seems to be an important link between environmental
enrichment and cognitive reserve: in mice, exposure to different odours
caused neuronal growth in murine olfactory bulbs (Veyrac et al., 2009).
However, this was blocked by the administration of ß-antagonists during
the environmental enrichment period (Veyrac et al., 2009). As moderate
LC activity is crucial for attention (Aston-Jones and Cohen, 2005) and as
LC activation also has been linked to anti-inammatory function
(McNamee et al., 2010) and BDNF secretion (Ivy et al., 2003), frequent
LC activation would conceptually be benecial for overall brain health
(for reviews on the importance of NA in healthy aging see Mather, 2021;
Mather and Harley, 2016). Results of the effect of cognitive training on
delaying dementia have been conicting (Sharp and Gatz, 2011; Wilson
et al., 2013) and more research is required in that eld.
An increasing number of studies have investigating whether physical
exercise can be benecial in aging and neurodegeneration.
In humans, it has been shown across age groups that exercise training
across a variety of tasks such as strength training (A˘
gg¨
on et al., 2020),
very light exercise (Kuwamizu et al., 2022), cycling and heavy running
(Strobel et al., 1997) can increase blood plasma noradrenergic levels
(A˘
gg¨
on et al., 2020; Strobel et al., 1997) and dilate pupil size, that is a
proxy of LC activity (Kuwamizu et al., 2022). The increase in norad-
renergic levels during physical exercise correlates with better cognitive
performance: during light cycling for 1 min, the increased level of the
noradrenergic metabolite MHPG of healthy young males from the rest to
exercise condition correlates with their faster response time in a mildly
mentally challenging 4-choice reaction time task (McMorris et al.,
2008). However, this effect could have also been explained by differ-
ences in exercise intensity (McMorris, 2003).
Recently, it has been shown in animal studies that exercise and
cognitive training through environmental enrichment can also increase
levels of NA (Naka et al., 2002) improve cognition (Bindra et al., 2021)
and may in part be mediated by ß-adrenergic receptors (Ebrahimi et al.,
2010). In Male Wistar rat model, it was shown that a pharmacological
decrease in hippocampal noradrenergic levels with timolol, β-adrenergic
receptor antagonist, decreased long-term object recognition memory
persistence while intrahippocampal NE injection had the opposite effect
(da Silva de Vargas et al., 2017). The same group furthermore showed
that in this rat model, object recognition memory consolidation was
accompanied by an increase in NE and BDNF (Mello-Carpes et al., 2016).
They furthermore showed that the increase in BDNF was eliminated if
the increase in NE was eliminated suggesting a causal relationship be-
tween both increases. Additionally, In Male Sprague-Dawley rats, the
benecial effects of exercise on cognition were shown to be partially
dependent on NA as giving the noradrenergic blocker propranolol
immediately after training abolished the increase in hippocampal BDNF
mRNA levels (Ivy et al., 2003).
Physical activity decreases the risk for dementia, although the type of
exercise and optimal frequency is still to be dened (for a review see
Iso-Markku et al., 2022; L´
opez-Ortiz et al., 2023). In mice, the benecial
effect of environmental enrichment on protection against AD has been
linked to the activation of ß-adrenergic receptors (Li et al., 2013).
8. Future outlook
It has been shown that LC degeneration is strongly linked to age and
pathology-related cognitive and physical decline, and may even precede
it. This might be expressed by a decline in LC-related functions such as
decreased muscle tonus, decreased working memory, sleeping problems
and hypertension.
As in healthy older adults, direct links between LC activity and task
performance are lacking, future human studies in both healthy young
and older adults should link differences in CSF NA levels or LC fMRI
activity tasks to differences in task performance to have a more direct
link between LC and other outcome measures such as task performance.
More emphasis should be put on LC tract integrity as it might drive
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
10
the link between LC integrity and cognition. However, this is more
challenging methodologically: the LC is a small structure, DTI requires
more advanced systems than structural MRI and LC white matter tracts
can be difcult to differentiate from surrounding tracts. Additionally, LC
tractography is not as well established as LC integrity measurements.
Therefore, future research is needed to rst establish standardised
methods for measuring LC tract integrity and to improve the accuracy
and reliability of this outcome measure.
The hypothesis that degeneration of the LC might accelerate the
progression of AD dementia, which in turn might further impair the LC,
is often complicated by the heterogeneity of data in cross-sectional de-
signs. However, longitudinal studies such as the DELCODE study (Jessen
et al., 2018), which collect structural measures and CSF data from the
same participants every year, or the longitudinal Harvard Aging Brain
Study (HABS) (Dagley et al., 2017), will allow matching changes in in-
dividual differences in LC integrity with individual changes in task
performance. Also studies that include longitudinal cognitive measures
and post-mortem data like the Religious Orders Study (ROS) and Aging
Project (MAP) (Bennett et al., 2018), allow to compare measures ob-
tained in vivo.
The LC-NA system can be targeted through exercise, pharmacologi-
cally and potentially via the vagus nerve in humans and has shown
promise as a target for healthy aging, AD and PD. However, large-scale
trials are lacking. Also, to maximise the effect, a combination of these
methods should be attempted.
Given the reciprocal relationship between degenerated LC and
inammation, more research should focus on the effects and combina-
tion of noradrenergic drugs with anti-inammatory agents. This bidi-
rectional interaction highlights the importance of targeting
inammation as a potential strategy to mitigate LC degeneration and
subsequently delay the aging process.
A viable yet often overlooked target for ADis MHPG, which is closely
related to tau, amyloid, inammation, cortical thickness (Jacobs et al.,
2021b; Riphagen et al., 2021; van Hooren et al., 2021) and, through its
link with p-tau, to neuropsychiatric symptoms (Jacobs et al., 2021b).
9. Conclusion
Recent developments in AD research have shown that the LC-NA
system might be exacerbating pathology in a vicious cycle with NA
metabolites promoting tau aggregation and tau promoting LC hyperac-
tivity causing higher levels of NE metabolites. Additionally, a decreased
protection from AD pathology due to decreased immune protection
caused by LC degeneration might further exacerbate the disease (Box
1a). However, NA also seems to be protective against amyloid aggre-
gation indicating a multifaceted LC involvement in AD.
In PD (Box 1b), the LC-NA system has been found to be involved in
early-to-mid stages, as it is linked to early non-motor symptoms such as
anxiety and depression, and its degeneration may accelerate PD
progression.
LC degeneration leads to neuroinammation in aging and neurode-
generative through shared increased microglia mechanisms but also
disease-specic mechanisms. In AD, LC degeneration contributes to
increased microglial reactivity and disruption of the BBB), thereby
accelerating the accumulation of Aß and tau proteins. On the other hand,
in PD, LC degeneration primarily leads to increased microglial reactivity
and oxidative damage. These mechanisms highlight the multifaceted
role of LC degeneration in promoting neuroinammation and the
pathological processes associated with AD and PD.
Finally, multiple noradrenergic drugs, vagus nerve stimulation and
physical exercise have all been shown to be promising therapeutic tar-
gets (Box 2) by potentially boosting noradrenergic function in aging and
neurodegenerative disease thus alleviating some of their debilitating
symptoms.
Box 1
Involvement of the LC-NA system in (A) AD and (B) PD.
.
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
11
Funding sources
F.K is supported by the German Federal Ministry of Education and
Research (BMBF, funding code 01ED2102B) under the aegis of the EU
Joint Programme Neurodegenerative Disease Research (JPND).
E.L, J.P and L.H are supported by the Deutsche For-
schungsgemeinschaft (DFG, German Research Foundation)
362321501/ Research Training Group (RTG) 2413 SynAGE.
M.Lu is supported by the federal state of Saxony-Anhalt and the
European Regional Development Fund (ERDF) in the Center for
Behavioral Brain Sciences (CBBS, ZS/2016/04/78113).
M.Le and G.G are supported by the Deutsche For-
schungsgemeinschaft (DFG, German Research Foundation) Project-ID
425899994 Sonderforschungsbereiche 1436 (SFB 1436).
E.D has received nancial support for his institution by Deutsche
Forschungsgemeinschaft (DFG, German Research Foundation) Project-
ID 425899994 Sonderforschungsbereiche 1436 (SFB 1436), Human
Brain Project, Specic Grant Agreement 3 (SGA3), Deutsche For-
schungsgemeinschaft (DFG, German Research Foundation) Sonder-
forschungsbereiche 1315 (SFB 1315).
M.J.B is supported by the Deutsche Forschungsgemeinschaft (DFG,
German Research Foundation) Project-ID 425899994 Sonderfor-
schungsbereiche 1436 (SFB 1436), Center for Behavioral Brain Sciences
(CBBS) NeuroNetzwerk 17, and by the German Federal Ministry of Ed-
ucation and Research (BMBF, funding code 01ED2102B) under the aegis
of the EU Joint Programme Neurodegenerative Disease Research
(JPND).
D.H is supported by Sonderforschungsbereich 1315, Project B06,
Box 2
LC-NA as a therapeutic target.
.
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
12
Sonderforschungsbereich 1436, Project A08, ARUK SRF2018B-004,
CBBS Neural Network (CBBS, ZS/2016/04/78113), and NIH
R01MH126971.
Competing interests
E.D has received payments for his role and works as consultant for
Roche, Biogen, RoxHealth and expert testimony for UCL Consultancy,
served at scientic advisory boards for EdoN Initiative and Ebsen Alz-
heimers Center (no payment) and Roche (personal nancial support),
and is a co-founder of the digital health start-up Neotiv.
F.K, E.L, M.Lu, M.Le, G.G, L.H, J.P, M.J.B and D.H have nothing to
disclose.
Data Availability
No data was used for the research described in the article.
References
af Bjerk´
en, S., Stenmark Persson, R., Barkander, A., Karalija, N., Pelegrina-Hidalgo, N.,
Gerhardt, G.A., Virel, A., Str¨
omberg, I., 2019. Noradrenaline is crucial for the
substantia nigra dopaminergic cell maintenance. Neurochem. Int. 131, 104551
https://doi.org/10.1016/j.neuint.2019.104551.
A˘
gg¨
on, E., Agirbas¸ , ¨
O., Alp, H.H., Uçan, I., Gürsoy, R., Hackney, A.C., 2020. Effect of
dynamic and static strength training on hormonal activity in elite boxers. BJHPA 12,
110. https://doi.org/10.29359/BJHPA.12.3.01.
Alvarado, J.S., Hateld, J., Mooney, R., 2023. Divergent projections from locus coeruleus
to the corticobasal ganglia system and ventral tegmental area of the adult male zebra
nch. J. Comp. Neurol. 531, 921934. https://doi.org/10.1002/cne.25474.
Amaral, D.G., Foss, J.A., 1975. Locus coeruleus lesions and learning. Science 188,
377378. https://doi.org/10.1126/science.1118734.
Anselmo-Franci, J.A., Rocha, M.J.A., Peres-Polon, V.L., Moreira, E.R., Antunes-
Rodrigues, J., Franci, C.R., 1999. Role of the locus coeruleus on blood pressure
response and atrial natriuretic peptide secretion following extracellular volume
expansion. Brain Res. Bull. 50, 173177. https://doi.org/10.1016/S0361-9230(99)
00183-5.
Arnsten, A.F.T., Goldman-Rakic, P.S., 1984. Selective prefrontal cortical projections to
the region of the locus coeruleus and raphe nuclei in the rhesus monkey. Brain Res.
306, 918. https://doi.org/10.1016/0006-8993(84)90351-2.
Aston-Jones, G., Bloom, F.E., 1981. Activity of norepinephrine-containing locus
coeruleus neurons in behaving rats anticipates uctuations in the sleep-waking
cycle. J. Neurosci. 1, 876886. https://doi.org/10.1523/JNEUROSCI.01-08-
00876.1981.
Aston-Jones, G., Cohen, J.D., 2005. An integrative theory of locus coeruleus-
norepinephrine function: adaptive gain and optimal performance. Annu. Rev.
Neurosci. 28, 403450. https://doi.org/10.1146/annurev.neuro.28.061604.135709.
Atik, A., Stewart, T., Zhang, J., 2016. Alpha-synuclein as a biomarker for parkinsons
disease: alpha-synuclein as a biomarker for PD. Brain Pathol. 26, 410418. https://
doi.org/10.1111/bpa.12370.
Avila, J., Jim´
enez, J.S., Sayas, C.L., Bol´
os, M., Zabala, J.C., Rivas, G., Hern´
andez, F.,
2016. Tau structures. Front. Aging Neurosci. 8, 262. https://doi.org/10.3389/
fnagi.2016.00262.
Bachman, S.L., Dahl, M.J., Werkle-Bergner, M., Düzel, S., Forlim, C.G., Lindenberger, U.,
Kühn, S., Mather, M., 2021. Locus coeruleus MRI contrast is associated with cortical
thickness in older adults. Neurobiol. Aging 100, 7282. https://doi.org/10.1016/j.
neurobiolaging.2020.12.019.
Bao, W., Xie, F., Zuo, C., Guan, Y., Huang, Y.H., 2021. PET neuroimaging of Alzheimers
disease: radiotracers and their utility in clinical research. Front. Aging Neurosci. 13,
624330 https://doi.org/10.3389/fnagi.2021.624330.
Barth´
elemy, N.R., Liu, H., Lu, W., Kotzbauer, P.T., Bateman, R.J., Lucey, B.P., 2020.
Sleep deprivation affects tau phosphorylation in human cerebrospinal uid. Ann.
Neurol. 87, 700709. https://doi.org/10.1002/ana.25702.
Beardmore, R., Hou, R., Darekar, A., Holmes, C., Boche, D., 2021. The locus coeruleus in
aging and Alzheimers disease: a postmortem and brain imaging review. JAD 83,
522. https://doi.org/10.3233/JAD-210191.
Beas, B.S., Wright, B.J., Skirzewski, M., Leng, Y., Hyun, J.H., Koita, O., Ringelberg, N.,
Kwon, H.-B., Buonanno, A., Penzo, M.A., 2018. The locus coeruleus drives
disinhibition in the midline thalamus via a dopaminergic mechanism. Nat. Neurosci.
21, 963973. https://doi.org/10.1038/s41593-018-0167-4.
Bekar, L.K., Wei, H.S., Nedergaard, M., 2012. The locus coeruleus-norepinephrine
network optimizes coupling of cerebral blood volume with oxygen demand. J. Cereb.
Blood Flow Metab. 32, 21352145. https://doi.org/10.1038/jcbfm.2012.115.
Bennett, D.A., Buchman, A.S., Boyle, P.A., Barnes, L.L., Wilson, R.S., Schneider, J.A.,
2018. Religious orders study and rush memory and aging project. J. Alzheimers Dis.
64, S161S189. https://doi.org/10.3233/JAD-179939.
Berg, D., Borghammer, P., Fereshtehnejad, S.-M., Heinzel, S., Horsager, J., Schaeffer, E.,
Postuma, R.B., 2021. Prodromal Parkinson disease subtypes key to understanding
heterogeneity. Nat. Rev. Neurol. 17, 349361. https://doi.org/10.1038/s41582-
021-00486-9.
Berridge, C.W., Waterhouse, B.D., 2003. The locus coeruleusnoradrenergic system:
modulation of behavioral state and state-dependent cognitive processes. Brain Res.
Rev. 42, 3384. https://doi.org/10.1016/S0165-0173(03)00143-7.
Bettcher, B.M., Kramer, J.H., 2014. Longitudinal inammation, cognitive decline, and
Alzheimers disease: a mini-review. Clin. Pharmacol. Ther. 96, 464469. https://doi.
org/10.1038/clpt.2014.147.
Betts, M.J., Cardenas-Blanco, A., Kanowski, M., Jessen, F., Düzel, E., 2017. In vivo MRI
assessment of the human locus coeruleus along its rostrocaudal extent in young and
older adults. NeuroImage 163, 150159. https://doi.org/10.1016/j.
neuroimage.2017.09.042.
Betts, M.J., Cardenas-Blanco, A., Kanowski, M., Spottke, A., Teipel, S.J., Kilimann, I.,
Jessen, F., Düzel, E., 2019a. Locus coeruleus MRI contrast is reduced in Alzheimers
disease dementia and correlates with CSF Aβ levels. Alzheimers Dement.: Diagn.
Assess. Dis. Monit. 11, 281285. https://doi.org/10.1016/j.dadm.2019.02.001.
Betts, M.J., Kirilina, E., Otaduy, M.C.G., Ivanov, D., Acosta-Cabronero, J., Callaghan, M.
F., Lambert, C., Cardenas-Blanco, A., Pine, K., Passamonti, L., Loane, C., Keuken, M.
C., Trujillo, P., Lüsebrink, F., Mattern, H., Liu, K.Y., Priovoulos, N., Fliessbach, K.,
Dahl, M.J., Maaß, A., Madelung, C.F., Meder, D., Ehrenberg, A.J., Speck, O.,
Weiskopf, N., Dolan, R., Inglis, B., Tosun, D., Morawski, M., Zucca, F.A., Siebner, H.
R., Mather, M., Uludag, K., Heinsen, H., Poser, B.A., Howard, R., Zecca, L., Rowe, J.
B., Grinberg, L.T., Jacobs, H.I.L., Düzel, E., H¨
ammerer, D., 2019b. Locus coeruleus
imaging as a biomarker for noradrenergic dysfunction in neurodegenerative
diseases. Brain 142, 25582571. https://doi.org/10.1093/brain/awz193.
Bharani, Krishna L., Derex, R., Granholm, A.-C., Ledreux, A., 2017. A noradrenergic
lesion aggravates the effects of systemic inammation on the hippocampus of aged
rats. PLoS One 12, e0189821. https://doi.org/10.1371/journal.pone.0189821.
Bindra, S., LaManna, J.C., Xu, S.C., 2021. Environmental enrichment improved cognitive
performance in mice under normoxia and hypoxia. Advances in Experimental
Medicine and Biology. Springer International Publishing, Cham. https://doi.org/
10.1007/978-3-030-48238-1.
Blau, C.W., Cowley, T.R., OSullivan, J., Grehan, B., Browne, T.C., Kelly, L., Birch, A.,
Murphy, N., Kelly, A.M., Kerskens, C.M., Lynch, M.A., 2012. The age-related decit
in LTP is associated with changes in perfusion and blood-brain barrier permeability.
Neurobiol. Aging. https://doi.org/10.1016/j.neurobiolaging.2011.09.035.
Borchert, R.J., Rittman, T., Rae, C.L., Passamonti, L., Jones, S.P., Vatansever, D., V´
azquez
Rodríguez, P., Ye, Z., Nombela, C., Hughes, L.E., Robbins, T.W., Rowe, J.B., 2019.
Atomoxetine and citalopram alter brain network organization in Parkinsons disease.
Brain Commun. 1, fcz013 https://doi.org/10.1093/braincomms/fcz013.
Borghammer, P., Van Den Berge, N., 2019. Brain-rst versus gut-rst Parkinsons
disease: a hypothesis. JPD 9, S281S295. https://doi.org/10.3233/JPD-191721.
Bouret, S., Sara, S.J., 2002. Locus coeruleus activation modulates ring rate and
temporal organization of odour-induced single-cell responses in rat piriform cortex.
Eur. J. Neurosci. 16, 23712382. https://doi.org/10.1046/j.1460-9568.2002.02413.
x.
Braak, H., Braak, E., 1991. Neuropathological stageing of Alzheimer-related changes.
Acta Neuropathol. 82, 239259. https://doi.org/10.1007/BF00308809.
Braak, H., Tredici, K.D., Rüb, U., de Vos, R.A.I., Jansen Steur, E.N.H., Braak, E., 2003.
Stageing of brain pathology related to sporadic Parkinsons disease. Neurobiol.
Ageing 24, 197211. https://doi.org/10.1016/S0197-4580(02)00065-9.
Braak, H., Thal, D.R., Ghebremedhin, E., Tredici, K.D., 2011. Stages of the pathologic
process in alzheimer disease: age categories from 1 to 100 years. J. Neuropathol.
Exp. Neurol. 70, 10.
Bremner, D., Krystal, J.H., Southwick, M., Charney, D.S., 1996. Noradrenergic
mechanisms in stress and anxiety: I. preclinical studies 11.
Brunello, N., Mendlewicz, J., Kasper, S., Leonard, B., Montgomery, S., Nelson, J.C.,
Paykel, E., Versiani, M., Racagni, G., 2002. The role of noradrenaline and selective
noradrenaline reuptake inhibition in depression. Eur. Neuropsychopharmacol. 12,
461475. https://doi.org/10.1016/S0924-977X(02)00057-3.
Burke, W.J., Schmitt, C.A., Miller, C., Li, S.W., 1997. Norepinephrine transmitter
metabolite induces apoptosis in differentiated rat pheochromocytoma cells. Brain
Res. 760, 290293. https://doi.org/10.1016/S0006-8993(97)00447-2.
Burke, W.J., Kristal, B.S., Yu, B.P., Li, S.W., Lin, T.-S., 1998. Norepinephrine transmitter
metabolite generates free radicals and activates mitochondrial permeability
transition: a mechanism for DOPEGAL-induced apoptosis. Brain Res. 787, 328332.
https://doi.org/10.1016/S0006-8993(97)01488-1.
Burke, W.J., Li, S.W., Schmitt, C.A., Xia, P., Chung, H.D., Gillespie, K.N., 1999.
Accumulation of 3,4-dihydroxyphenylglycolaldehyde, the neurotoxic monoamine
oxidase A metabolite of norepinephrine, in locus ceruleus cell bodies in Alzheimers
disease: mechanism of neuron death. Brain Res. 816, 633637. https://doi.org/
10.1016/S0006-8993(98)01211-6.
Butkovich, L.M., Houser, M.C., Tansey, M.G., 2018.
α
-synuclein and noradrenergic
modulation of immune cells in Parkinsons disease pathogenesis. Front Neurosci. 12,
626. https://doi.org/10.3389/fnins.2018.00626.
Butt, M.F., Albusoda, A., Farmer, A.D., Aziz, Q., 2020. The anatomical basis for
transcutaneous auricular vagus nerve stimulation. J. Anat. 236, 588611. https://
doi.org/10.1111/joa.13122.
Buxton, R.B., Uluda˘
g, K., Dubowitz, D.J., Liu, T.T., 2004. Modeling the hemodynamic
response to brain activation. NeuroImage 23, S220S233. https://doi.org/10.1016/
j.neuroimage.2004.07.013.
Cakmak, Y.O., Apaydin, H., Kiziltan, G., Gündüz, A., Ozsoy, B., Olcer, S., Urey, H.,
Cakmak, O.O., Ozdemir, Y.G., Ertan, S., 2017. Rapid alleviation of Parkinsons
disease symptoms via electrostimulation of intrinsic auricular muscle zones. Front.
Hum. Neurosci. 11.
Calarco, N., Cassidy, C.M., Selby, B., Hawco, C., Voineskos, A.N., Diniz, B.S., Nikolova, Y.
S., 2022. Associations between locus coeruleus integrity and diagnosis, age, and
cognitive performance in older adults with and without late-life depression: An
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
13
exploratory study. NeuroImage: Clin. 36, 103182 https://doi.org/10.1016/j.
nicl.2022.103182.
Carey, G., G¨
ormezo˘
glu, M., de Jong, J.J.A., Hofman, P.A.M., Backes, W.H., Dujardin, K.,
Leentjens, A.F.G., 2021. Neuroimageing of anxiety in Parkinsons disease: a
systematic review. Mov. Disord. 36, 327339. https://doi.org/10.1002/mds.28404.
Carpenter, M.B., Periera, A.B., Guha, N., 1992. Immunocytochemistry of oculomotor
afferents in the squirrel monkey (Saimiri sciureus). J. Hirnforsch. 33, 151167.
Cassidy, C.M., Zucca, F.A., Girgis, R.R., Baker, S.C., Weinstein, J.J., Sharp, M.E.,
Bellei, C., Valmadre, A., Vanegas, N., Kegeles, L.S., Brucato, G., Kang, U.J.,
Sulzer, D., Zecca, L., Abi-Dargham, A., Horga, G., 2019. Neuromelanin-sensitive MRI
as a noninvasive proxy measure of dopamine function in the human brain. Proc.
Natl. Acad. Sci. USA 116, 51085117. https://doi.org/10.1073/pnas.1807983116.
Cassidy, C.M., Therriault, J., Pascoal, T.A., Cheung, V., Savard, M., Tuominen, L.,
Chamoun, M., McCall, A., Celebi, S., Lussier, F., Massarweh, G., Soucy, J.-P.,
Weinshenker, D., Tardif, C., Ismail, Z., Gauthier, S., Rosa-Neto, P., 2022. Association
of locus coeruleus integrity with Braak stage and neuropsychiatric symptom severity
in Alzheimers disease. Neuropsychopharmacol 47, 11281136. https://doi.org/
10.1038/s41386-022-01293-6.
Cedarbaum, J.M., Aghajanian, G.K., 1978. Afferent projections to the rat locus coeruleus
as determined by a retrograde tracing technique. J. Comp. Neurol. 178, 115.
https://doi.org/10.1002/cne.901780102.
Chalermpalanupap, T., Weinshenker, D., Rorabaugh, J.M., 2017. Down but not out: the
consequences of pretangle tau in the locus coeruleus. Neural Plast. 2017, 19.
https://doi.org/10.1155/2017/7829507.
Chalermpalanupap, T., Schroeder, J.P., Rorabaugh, J.M., Liles, L.C., Lah, J.J., Levey, A.I.,
Weinshenker, D., 2018. Locus coeruleus ablation exacerbates cognitive decits,
neuropathology, and lethality in p301s tau transgenic mice. J. Neurosci. 38, 7492.
https://doi.org/10.1523/JNEUROSCI.1483-17.2017.
Charney, D.S., Grillon, C., Bremner, J.D., 1998. Review: the neurobiological basis of
anxiety and fear: circuits, mechanisms, and neurochemical interactions (Part I).
Neuroscientist 4, 3544. https://doi.org/10.1177/107385849800400111.
Chen, F.-J., Sara, S.J., 2007. Locus coeruleus activation by foot shock or electrical
stimulation inhibits amygdala neurons. Neuroscience 144, 472481. https://doi.
org/10.1016/j.neuroscience.2006.09.037.
Ciampa, C.J., Parent, J.H., Harrison, T.M., Fain, R.M., Betts, M.J., Maass, A., Winer, J.R.,
Baker, S.L., Janabi, M., Furman, D.J., DEsposito, M., Jagust, W.J., Berry, A.S., 2022.
Associations among locus coeruleus catecholamines, tau pathology, and memory in
ageing. Neuropsychopharmacology 47, 11061113. https://doi.org/10.1038/
s41386-022-01269-6.
Clewett, D.V., Lee, T.-H., Greening, S., Ponzio, A., Margalit, E., Mather, M., 2016.
Neuromelanin marks the spot: identifying a locus coeruleus biomarker of cognitive
reserve in healthy ageing. Neurobiol. Ageing 37, 117126. https://doi.org/10.1016/
j.neurobiolageing.2015.09.019.
Counts, S.E., Mufson, E.J., 2010. Noradrenaline activation of neurotrophic pathways
protects against neuronal amyloid toxicity. J. Neurochem. 113, 649660. https://
doi.org/10.1111/j.1471-4159.2010.06622.x.
Criaud, M., Laurencin, C., Poisson, A., Metereau, E., Redout´
e, J., Thobois, S.,
Boulinguez, P., Ballanger, B., 2022. Noradrenaline and movement initiation
disorders in parkinsons disease: a pharmacological functional MRI study with
clonidine. Cells 11, 2640. https://doi.org/10.3390/cells11172640.
Crimins, J.L., Rocher, A.B., Luebke, J.I., 2012. Electrophysiological changes precede
morphological changes to frontal cortical pyramidal neurons in the rTg4510 mouse
model of progressive tauopathy. Acta Neuropathol. 124, 777795. https://doi.org/
10.1007/s00401-012-1038-9.
Cunningham, C., Hennessy, E., 2015. Co-morbidity and systemic inammation as drivers
of cognitive decline: new experimental models adopting a broader paradigm in
dementia research. Alzheimers Res. Ther. 7, 33. https://doi.org/10.1186/s13195-
015-0117-2.
Cunningham, C., Campion, S., Lunnon, K., Murray, C.L., Woods, J.F.C., Deacon, R.M.J.,
Rawlins, J.N.P., Perry, V.H., 2009. Systemic inammation induces acute behavioral
and cognitive changes and accelerates neurodegenerative disease. Biol. Psychiatry
65, 304312. https://doi.org/10.1016/j.biopsych.2008.07.024.
da Silva de Vargas, L., Neves, B.-H.S., das, Roehrs, R., Izquierdo, I., Mello-Carpes, P.,
2017. One-single physical exercise session after object recognition learning promotes
memory persistence through hippocampal noradrenergic mechanisms. Behav. Brain
Res. 329, 120126. https://doi.org/10.1016/j.bbr.2017.04.050.
Dagley, A., LaPoint, M., Huijbers, W., Hedden, T., McLaren, D.G., Chatwal, J.P., Papp, K.
V., Amariglio, R.E., Blacker, D., Rentz, D.M., Johnson, K.A., Sperling, R.A.,
Schultz, A.P., 2017. Harvard ageing brain study: dataset and accessibility.
Neuroimage 144, 255258. https://doi.org/10.1016/j.neuroimage.2015.03.069.
Dahl, M.J., Mather, M., Düzel, S., Bodammer, N.C., Lindenberger, U., Kühn, S., Werkle-
Bergner, M., 2019. Rostral locus coeruleus integrity is associated with better memory
performance in older adults. Nat. Hum. Behav. 3, 12031214. https://doi.org/
10.1038/s41562-019-0715-2.
Dahl, M.J., Mather, M., Sander, M.C., Werkle-Bergner, M., 2020. Noradrenergic
responsiveness supports selective attention across the adult lifespan. J. Neurosci. 40,
43724390. https://doi.org/10.1523/JNEUROSCI.0398-19.2020.
Dahl, M.J., Bachman, S.L., Dutt, S., Düzel, S., Bodammer, N.C., Lindenberger, U.,
Kühn, S., Werkle-Bergner, M., Mather, M., 2022. The integrity of dopaminergic and
noradrenergic brain regions is associated with different aspects of late-life memory
performance. Neuroscience. https://doi.org/10.1101/2022.10.12.511748.
David, M., Malhotra, P.A., 2022. New approaches for the quantication and targeting of
noradrenergic dysfunction in Alzheimers disease. Ann. Clin. Transl. Neurol. 9,
582596. https://doi.org/10.1002/acn3.51539.
David, M.C.B., Del Giovane, M., Liu, K.Y., Gostick, B., Rowe, J.B., Oboh, I., Howard, R.,
Malhotra, P.A., 2022. Cognitive and neuropsychiatric effects of noradrenergic
treatment in Alzheimers disease: systematic review and meta-analysis. J. Neurol.
Neurosurg. Psychiatry 93, 10801090. https://doi.org/10.1136/jnnp-2022-329136.
De Sarro, G. b, Ascioti, C., Froio, F., Libri, V., Nistic`
o, G., 1987. Evidence that locus
coeruleus is the site where clonidine and drugs acting at
α
1- and
α
2-adrenoceptors
affect sleep and arousal mechanisms. Br. J. Pharmacol. 90, 675685. https://doi.
org/10.1111/j.1476-5381.1987.tb11220.x.
Debiec, J., LeDoux, J.E., 2006. Noradrenergic signaling in the amygdala contributes to
the reconsolidation of fear memory: treatment implications for PTSD. Ann. N. Y.
Acad. Sci. 1071, 521524. https://doi.org/10.1196/annals.1364.056.
del Campo, N., Chamberlain, S.R., Sahakian, B.J., Robbins, T.W., 2011. The roles of
dopamine and noradrenaline in the pathophysiology and treatment of attention-
decit/hyperactivity disorder. Biol. Psychiatry Prefrontal Cortical Circuits Regul.
Atten. Behav. Emot. 69, e145e157. https://doi.org/10.1016/j.
biopsych.2011.02.036.
Deutch, A.Y., Goldstein, M., Roth, R.H., 1986. Activation of the locus coeruleus induced
by selective stimulation of the ventral tegmental area. Brain Res. 363, 307314.
https://doi.org/10.1016/0006-8993(86)91016-4.
Devoto, P., Flore, G., Saba, P., F`
a, M., Gessa, G.L., 2005. Stimulation of the locus
coeruleus elicits noradrenaline and dopamine release in the medial prefrontal and
parietal cortex. J. Neurochem 92, 368374. https://doi.org/10.1111/j.1471-
4159.2004.02866.x.
Devoto, P., Flore, G., Saba, P., Castelli, M.P., Piras, A.P., Luesu, W., Viaggi, M.C.,
Ennas, M.G., Gessa, G.L., 2008. 6-Hydroxydopamine lesion in the ventral tegmental
area fails to reduce extracellular dopamine in the cerebral cortex. J. Neurosci. Res.
86, 16471658. https://doi.org/10.1002/jnr.21611.
Devoto, P., Sagheddu, C., Santoni, M., Flore, G., Saba, P., Pistis, M., Gessa, G.L., 2020.
Noradrenergic source of dopamine assessed by microdialysis in the medial prefrontal
cortex. Front. Pharmacol. 11, 588160 https://doi.org/10.3389/fphar.2020.588160.
Dickson, D.W., Fujishiro, H., DelleDonne, A., Menke, J., Ahmed, Z., Klos, K.J., Josephs, K.
A., Frigerio, R., Burnett, M., Parisi, J.E., Ahlskog, J.E., 2008. Evidence that incidental
Lewy body disease is pre-symptomatic Parkinsons disease. Acta Neuropathol. 115,
437444. https://doi.org/10.1007/s00401-008-0345-7.
Dietrichs, E., 1988. Cerebellar cortical and nuclear afferents from the feline locus
coeruleus complex. Neuroscience 27, 7791. https://doi.org/10.1016/0306-4522
(88)90220-5.
Dobarro, M., Orejana, L., Aguirre, N., Ramírez, M.J., 2013. Propranolol restores
cognitive decits and improves amyloid and Tau pathologies in a senescence-
accelerated mouse model. Neuropharmacol. Cogn. Enhanc.: Mol. Mech. Minds 64,
137144. https://doi.org/10.1016/j.neuropharm.2012.06.047.
Doppler, C.E.J., Kinnerup, M.B., Brune, C., Farrher, E., Betts, M., Fedorova, T.D.,
Schaldemose, J.L., Knudsen, K., Ismail, R., Seger, A.D., Hansen, A.K., Stær, K.,
Fink, G.R., Brooks, D.J., Nahimi, A., Borghammer, P., Sommerauer, M., 2021.
Regional locus coeruleus degeneration is uncoupled from noradrenergic terminal
loss in Parkinsons disease. Brain 144, 27322744. https://doi.org/10.1093/brain/
awab236.
Downs, A.M., McElligott, Z.A., 2022. Noradrenergic circuits and signaling in substance
use disorders. Neuropharmacology 208, 108997. https://doi.org/10.1016/j.
neuropharm.2022.108997.
Duszkiewicz, A.J., McNamara, C.G., Takeuchi, T., Genzel, L., 2019. Novelty and
dopaminergic modulation of memory persistence: a tale of two systems. Trends
Neurosci. 42, 102114. https://doi.org/10.1016/j.tins.2018.10.002.
Ebrahimi, S., Rashidy-Pour, A., Vafaei, A.A., Akhavan, M.M., 2010. Central β-adrenergic
receptors play an important role in the enhancing effect of voluntary exercise on
learning and memory in rat. Behav. Brain Res. 208, 189193. https://doi.org/
10.1016/j.bbr.2009.11.032.
Edison, P., Ahmed, I., Fan, Z., Hinz, R., Gelosa, G., Ray Chaudhuri, K., Walker, Z.,
Turkheimer, F.E., Brooks, D.J., 2013. Microglia, amyloid, and glucose metabolism in
parkinsons disease with and without dementia. Neuropsychopharmacol 38,
938949. https://doi.org/10.1038/npp.2012.255.
Ehrenberg, A.J., Suemoto, C.K., França Resende, E., de, P., Petersen, C., Leite, R.E.P.,
Rodriguez, R.D., Ferretti-Rebustini, R.E., de, L., You, M., Oh, J., Nitrini, R.,
Pasqualucci, C.A., Jacob-Filho, W., Kramer, J.H., Gatchel, J.R., Grinberg, L.T., 2018.
Neuropathologic correlates of psychiatric symptoms in Alzheimers disease.
J. Alzheimers Dis. 66, 115126. https://doi.org/10.3233/JAD-180688.
Ehrminger, M., Latimier, A., Pyatigorskaya, N., Garcia-Lorenzo, D., Leu-Semenescu, S.,
Vidailhet, M., Lehericy, S., Arnulf, I., 2016. The coeruleus/subcoeruleus complex in
idiopathic rapid eye movement sleep behaviour disorder. Brain 139, 11801188.
https://doi.org/10.1093/brain/aww006.
Eisenhofer, G., Kopin, I.J., Goldstein, D.S., 2004. Catecholamine metabolism: a
contemporary view with implications for physiology and medicine. Pharm. Rev. 56,
331349. https://doi.org/10.1124/pr.56.3.1.
Elrod, R., Peskind, E.R., DiGiacomo, L., Brodkin, K.I., Veith, R.C., Raskind, M.A., 1997.
Effects of Alzheimers disease severity on cerebrospinal uid norepinephrine
concentration. Am. J. Psychiatry 154, 2530. https://doi.org/10.1176/ajp.154.1.25.
Engelender, S., Isacson, O., 2017. The threshold theory for Parkinsons disease. Trends
Neurosci. 40, 414. https://doi.org/10.1016/j.tins.2016.10.008.
Engineer, N.D., Percaccio, C.R., Pandya, P.K., Moucha, R., Rathbun, D.L., Kilgard, M.P.,
2004. Environmental enrichment improves response strength, threshold, selectivity,
and latency of auditory cortex neurons. J. Neurophysiol. 92, 7382. https://doi.org/
10.1152/jn.00059.2004.
Erd˝
o, F., Denes, L., de Lange, E., 2017. Age-associated physiological and pathological
changes at the blood-brain barrier: a review. J. Cereb. Blood Flow Metab. 37, 424.
https://doi.org/10.1177/0271678X16679420.
Espa˜
na, R.A., Berridge, C.W., 2006. Organization of noradrenergic efferents to arousal-
related basal forebrain structures. J. Comp. Neurol. 496, 668683. https://doi.org/
10.1002/cne.20946.
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
14
Espay, A.J., LeWitt, P.A., Kaufmann, H., 2014. Norepinephrine deciency in Parkinsons
disease: the case for noradrenergic enhancement: norepinephrine deciency in PD.
Mov. Disord. 29, 17101719. https://doi.org/10.1002/mds.26048.
Evans, A.K., Ardestani, P.M., Yi, B., Park, H.H., Lam, R.K., Shamloo, M., 2020. Beta-
adrenergic receptor antagonism is proinammatory and exacerbates
neuroinammation in a mouse model of Alzheimers Disease. Neurobiol. Dis. 146,
105089 https://doi.org/10.1016/j.nbd.2020.105089.
Evans, A.K., Defensor, E., Shamloo, M., 2022. Selective vulnerability of the locus
coeruleus noradrenergic system and its role in modulation of neuroinammation,
cognition, and neurodegeneration. Front. Pharmacol. 13, 1030609. https://doi.org/
10.3389/fphar.2022.1030609.
Fan, Z., Okello, A.A., Brooks, D.J., Edison, P., 2015. Longitudinal inuence of microglial
activation and amyloid on neuronal function in Alzheimers disease. Brain 138,
36853698. https://doi.org/10.1093/brain/awv288.
Farrand, A.Q., Helke, K.L., Gregory, R.A., Gooz, M., Hinson, V.K., Boger, H.A., 2017.
Vagus nerve stimulation improves locomotion and neuronal populations in a model
of Parkinsons disease. Brain Stimul. 10, 10451054. https://doi.org/10.1016/j.
brs.2017.08.008.
Farrand, A.Q., Verner, R.S., McGuire, R.M., Helke, K.L., Hinson, V.K., Boger, H.A., 2020.
Differential effects of vagus nerve stimulation paradigms guide clinical development
for Parkinsons disease. Brain Stimul. 13, 13231332. https://doi.org/10.1016/j.
brs.2020.06.078.
Farrow, S.L., Cooper, A.A., OSullivan, J.M., 2022. Redening the hypotheses driving
Parkinsons diseases research. npj Park. Dis. 8, 45. https://doi.org/10.1038/s41531-
022-00307-w.
Feinstein, D.L., Heneka, M.T., Gavrilyuk, V., Dello Russo, C., Weinberg, G., Galea, E.,
2002. Noradrenergic regulation of inammatory gene expression in brain.
Neurochem. Int. 41, 357365. https://doi.org/10.1016/s0197-0186(02)00049-9.
Flierl, M.A., Rittirsch, D., Nadeau, B.A., Sarma, J.V., Day, D.E., Lentsch, A.B., Huber-
Lang, M.S., Ward, P.A., 2009. Upregulation of phagocyte-derived catecholamines
augments the acute inammatory response. PLoS One 4, e4414. https://doi.org/
10.1371/journal.pone.0004414.
Florin-Lechner, S.M., Druhan, J.P., Aston-Jones, G., Valentino, R.J., 1996. Enhanced
norepinephrine release in prefrontal cortex with burst stimulation of the locus
coeruleus. Brain Res. 742, 8997. https://doi.org/10.1016/S0006-8993(96)00967-
5.
Foote, S.L., Aston-Jones, G., Bloom, F.E., 1980. Impulse activity of locus coeruleus
neurons in awake rats and monkeys is a function of sensory stimulation and arousal.
PNAS 77, 30333037.
Franceschi, C., Campisi, J., 2014. Chronic inammation (Inammageing) and its
potential contribution to age-associated diseases. J. Gerontol.: Ser. A 69, S4S9.
https://doi.org/10.1093/gerona/glu057.
Franceschi, C., Bonaf`
e, M., Valensin, S., Olivieri, F., De Luca, M., Ottaviani, E., De
Benedictis, G., 2000. Inamm-ageing: an evolutionary perspective on
immunosenescence. Ann. N. Y. Acad. Sci. 908, 244254. https://doi.org/10.1111/
j.1749-6632.2000.tb06651.x.
Fu, Y., Tvrdik, P., Makki, N., Paxinos, G., Watson, C., 2011. Precerebellar cell groups in
the hindbrain of the mouse dened by retrograde tracing and correlated with
cumulative Wnt1-Cre genetic labeling. Cerebellum 10, 570584. https://doi.org/
10.1007/s12311-011-0266-1.
Galgani, A., Lombardo, F., Della Latta, D., Martini, N., Bonuccelli, U., Fornai, F.,
Giorgi, F.S., 2020. Locus coeruleus magnetic resonance imageing in neurological
diseases. Curr. Neurol. Neurosci. Rep. 21, 2. https://doi.org/10.1007/s11910-020-
01087-7.
García-Lorenzo, D., Longo-Dos Santos, C., Ewenczyk, C., Leu-Semenescu, S., Gallea, C.,
Quattrocchi, G., Pita Lobo, P., Poupon, C., Benali, H., Arnulf, I., Vidailhet, M.,
Lehericy, S., 2013. The coeruleus/subcoeruleus complex in rapid eye movement
sleep behaviour disorders in Parkinsons disease. Brain 136, 21202129. https://doi.
org/10.1093/brain/awt152.
Gerhard, A., Pavese, N., Hotton, G., Turkheimer, F., Es, M., Hammers, A., Eggert, K.,
Oertel, W., Banati, R.B., Brooks, D.J., 2006. In vivo imageing of microglial activation
with [11C](R)-PK11195 PET in idiopathic Parkinsons disease. Neurobiol. Dis. 21,
404412. https://doi.org/10.1016/j.nbd.2005.08.002.
Gibbs, M.E., Hutchinson, D.S., Summers, R.J., 2010. Noradrenaline release in the locus
coeruleus modulates memory formation and consolidation; roles for
α
- and
β-adrenergic receptors. Neuroscience 170, 12091222. https://doi.org/10.1016/j.
neuroscience.2010.07.052.
Gilzenrat, M.S., Nieuwenhuis, S., Jepma, M., Cohen, J.D., 2010. Pupil diameter tracks
changes in control state predicted by the adaptive gain theory of locus coeruleus
function. Cogn. Affect. Behav. Neurosci. 10, 252269. https://doi.org/10.3758/
CABN.10.2.252.
Giorgi, F.S., Biagioni, F., Galgani, A., Pavese, N., Lazzeri, G., Fornai, F., 2020. Locus
coeruleus modulates neuroinammation in Parkinsonism and dementia. IJMS 21,
8630. https://doi.org/10.3390/ijms21228630.
Giorgi, F.S., Galgani, A., Puglisi-Allegra, S., Busceti, C.L., Fornai, F., 2021. The
connections of locus coeruleus with hypothalamus: potential involvement in
Alzheimers disease. J. Neural Transm. 128, 589613. https://doi.org/10.1007/
s00702-021-02338-8.
Goedert, M., 1993. TauproteinandtheneurolbrillarypathologyofAIzheimers disease 16,
6.
Grisanti, L.A., Evanson, J., Marchus, E., Jorissen, H., Woster, A.P., DeKrey, W., Sauter, E.
R., Combs, C.K., Porter, J.E., 2010. Pro-inammatory responses in human monocytes
are β1-adrenergic receptor subtype dependent. Mol. Immunol. 47, 12441254.
https://doi.org/10.1016/j.molimm.2009.12.013.
Guo, N.-N., Li, B.-M., 2007. Cellular and subcellular distributions of β1- and β2-
Adrenoceptors in the CA1 and CA3 regions of the rat hippocampus. Neuroscience
146, 298305. https://doi.org/10.1016/j.neuroscience.2007.01.013.
Guti´
errez, I.L., Dello Russo, C., Novellino, F., Caso, J.R., García-Bueno, B., Leza, J.C.,
Madrigal, J.L.M., 2022. Noradrenaline in Alzheimers disease: a new potential
therapeutic target. Int. J. Mol. Sci. 23, 6143. https://doi.org/10.3390/
ijms23116143.
Hamelin, L., Lagarde, J., Doroth´
ee, G., Leroy, C., Labit, M., Comley, R.A., De Souza, L.C.,
Corne, H., Dauphinot, L., Bertoux, M., Dubois, B., Gervais, P., Colliot, O., Potier, M.
C., Bottlaender, M., Sarazin, M., the Clinical IMABio3 team, 2016. Early and
protective microglial activation in Alzheimers disease: a prospective study using
18
F-DPA-714 PET imageing. Brain 139, 12521264. https://doi.org/10.1093/brain/
aww017.
H¨
ammerer, D., Callaghan, M.F., Hopkins, A., Kosciessa, J., Betts, M., Cardenas-
Blanco, A., Kanowski, M., Weiskopf, N., Dayan, P., Dolan, R.J., Düzel, E., 2018.
Locus coeruleus integrity in old age is selectively related to memories linked with
salient negative events. Proc. Natl. Acad. Sci. USA 115, 22282233. https://doi.org/
10.1073/pnas.1712268115.
Hansen, N., 2017. The longevity of hippocampus-dependent memory is orchestrated by
the locus coeruleus-noradrenergic system. Neural Plast. 2017, 19. https://doi.org/
10.1155/2017/2727602.
Harada, C.N., Natelson Love, M.C., Triebel, K.L., 2013. Normal cognitive ageing. Clin.
Geriatr. Med. 29, 737752. https://doi.org/10.1016/j.cger.2013.07.002.
Haring, J.H., Davis, J.N., 1985. Differential distribution of locus coeruleus projections to
the hippocampal formation: anatomical and biochemical evidence. Brain Res. 325,
366369. https://doi.org/10.1016/0006-8993(85)90342-7.
Harley, C.W., Walling, S.G., Yuan, Q., Martin, G.M., 2021. The ‘a, b, cs of pretangle tau
and their relation to ageing and the risk of Alzheimers disease. Semin. Cell Dev.
Biol. 116, 125134. https://doi.org/10.1016/j.semcdb.2020.12.010.
Hassert, D.L., Miyashita, T., Williams, C.L., 2004. The effects of peripheral vagal nerve
stimulation at a memory-modulating intensity on norepinephrine output in the
basolateral amygdala. Behav. Neurosci. 118, 7988. https://doi.org/10.1037/0735-
7044.118.1.79.
Head, G.A., Chan, C.K., Burke, S.L., 1998. Relationship between imidazoline and alpha2-
adrenoceptors involved in the sympatho-inhibitory actions of centrally acting
antihypertensive agents. J. Auton. Nerv. Syst. 72, 163169. https://doi.org/
10.1016/s0165-1838(98)00101-5.
Hector, A., Brouillette, J., 2021. Hyperactivity induced by soluble amyloid-β oligomers in
the early stages of Alzheimers disease. Front. Mol. Neurosci. 13, 600084 https://doi.
org/10.3389/fnmol.2020.600084.
Heneka, M.T., Nadrigny, F., Regen, T., Martinez-Hernandez, A., Dumitrescu-Ozimek, L.,
Terwel, D., Jardanhazi-Kurutz, D., Walter, J., Kirchhoff, F., Hanisch, U.-K.,
Kummer, M.P., 2010. Locus ceruleus controls Alzheimers disease pathology by
modulating microglial functions through norepinephrine. Proc. Natl. Acad. Sci. USA
107, 60586063. https://doi.org/10.1073/pnas.0909586107.
Heneka, M.T., Carson, M.J., Khoury, J.E., Landreth, G.E., Brosseron, F., Feinstein, D.L.,
Jacobs, A.H., Wyss-Coray, T., Vitorica, J., Ransohoff, R.M., Herrup, K., Frautschy, S.
A., Finsen, B., Brown, G.C., Verkhratsky, A., Yamanaka, K., Koistinaho, J., Latz, E.,
Halle, A., Petzold, G.C., Town, T., Morgan, D., Shinohara, M.L., Perry, V.H.,
Holmes, C., Bazan, N.G., Brooks, D.J., Hunot, S., Joseph, B., Deigendesch, N.,
Garaschuk, O., Boddeke, E., Dinarello, C.A., Breitner, J.C., Cole, G.M., Golenbock, D.
T., Kummer, M.P., 2015. Neuroinammation in Alzheimers disease, 388600405
Lancet Neurol. 14. https://doi.org/10.1016/S1474-4422(15)70016-5.
Henrich, M.T., Geibl, F.F., Lee, B., Chiu, W.-H., Koprich, J.B., Brotchie, J.M.,
Timmermann, L., Decher, N., Matschke, L.A., Oertel, W.H., 2018. A53T-
α
-synuclein
overexpression in murine locus coeruleus induces Parkinsons disease-like pathology
in neurons and glia. Acta Neuropathol. Commun. 6, 39. https://doi.org/10.1186/
s40478-018-0541-1.
Herrera, A.J., Casta˜
no, A., Venero, J.L., Cano, J., Machado, A., 2000. The single
intranigral injection of LPS as a new model for studying the selective effects of
inammatory reactions on dopaminergic system. Neurobiol. Dis. 7, 429447.
https://doi.org/10.1006/nbdi.2000.0289.
Hezemans, F.H., Wolpe, N., OCallaghan, C., Ye, R., Rua, C., Simon Jones, P., Murley, A.
G., Holland, N., Regenthal, R., Tsvetanov, K.A., Barker, R.A., Williams-Gray, C.H.,
Robbins, T.W., Passamonti, L., Rowe, J.B., 2022. Noradrenergic decits contribute to
apathy in Parkinsons disease through the precision of expected outcomes. PLoS
Comput. Biol. 18, 130. https://doi.org/10.1371/journal.pcbi.1010079.
Hirsch, L., Jette, N., Frolkis, A., Steeves, T., Pringsheim, T., 2016. The incidence of
Parkinsons disease: a systematic review and meta-analysis. Neuroepidemiology 46,
292300. https://doi.org/10.1159/000445751.
Holmes, C., 2013. Review: Systemic inammation and Alzheimers disease. Neuropathol.
Appl. Neurobiol. 39, 5168. https://doi.org/10.1111/j.1365-2990.2012.01307.x.
Hoogendijk, W.J.G., Pool, C.W., Troost, D., van Zwieten, E., Swaab, D.F., 1995. Image
analyser-assisted morphometry of the locus coeruleus in Alzheimers disease,
Parkinsons disease and amyotrophic lateral sclerosis. Brain 118, 131143. https://
doi.org/10.1093/brain/118.1.131.
Horsager, J., Andersen, K.B., Knudsen, K., Skjærbæk, C., Fedorova, T.D., Okkels, N.,
Schaeffer, E., Bonkat, S.K., Geday, J., Otto, M., Sommerauer, M., Danielsen, E.H.,
Bech, E., Kraft, J., Munk, O.L., Hansen, S.D., Pavese, N., G¨
oder, R., Brooks, D.J.,
Berg, D., Borghammer, P., 2020. Brain-rst versus body-rst Parkinsons disease: a
multimodal imageing case-control study. Brain 143, 30773088. https://doi.org/
10.1093/brain/awaa238.
Hulsey, D.R., Riley, J.R., Loerwald, K.W., Rennaker, R.L., Kilgard, M.P., Hays, S.A., 2017.
Parametric characterization of neural activity in the locus coeruleus in response to
vagus nerve stimulation. Exp. Neurol. 289, 2130. https://doi.org/10.1016/j.
expneurol.2016.12.005.
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
15
Hulsey, D.R., Shedd, C.M., Sarker, S.F., Kilgard, M.P., Hays, S.A., 2019. Norepinephrine
and serotonin are required for vagus nerve stimulation directed cortical plasticity.
Exp. Neurol. 320, 112975 https://doi.org/10.1016/j.expneurol.2019.112975.
Huong, V.T., Shimanouchi, T., Shimauchi, N., Yagi, H., Umakoshi, H., Goto, Y., Kuboi, R.,
2010. Catechol derivatives inhibit the bril formation of amyloid-beta peptides.
J. Biosci. Bioeng. 109, 629634. https://doi.org/10.1016/j.jbiosc.2009.11.010.
Hussain, B., Fang, C., Chang, J., 2021. Bloodbrain barrier breakdown: an emerging
biomarker of cognitive impairment in normal ageing and dementia. Front. Neurosci.
15.
Huynh, B., Fu, Y., Kirik, D., Shine, J.M., Halliday, G.M., 2021. Comparison of locus
coeruleus pathology with nigral and forebrain pathology in Parkinsons disease.
Mov. Disord. 36, 20852093. https://doi.org/10.1002/mds.28615.
Hwang, K.S., Langley, J., Tripathi, R., Hu, X.P., Huddleston, D.E., 2022. Reproducible in
vivo detection of locus coeruleus pathology in Parkinsons Disease. medRxiv,
2022.02.23.22271356.
I, B., Rm, K., Hm, B., Sc, H., S, R., Gl, W., 2014. Age and duration of inammatory
environment differentially affect the neuroimmune response and catecholaminergic
neurons in the midbrain and brainstem. Neurobiol. Ageing 35. https://doi.org/
10.1016/j.neurobiolageing.2013.11.006.
Iadecola, C., 2017. The Neurovascular Unit Coming of Age: A Journey through
Neurovascular Coupling in Health and Disease. Neuron 96, 1742. https://doi.org/
10.1016/j.neuron.2017.07.030.
Iannaccone, S., Cerami, C., Alessio, M., Garibotto, V., Panzacchi, A., Olivieri, S.,
Gelsomino, G., Moresco, R.M., Perani, D., 2013. In vivo microglia activation in very
early dementia with Lewy bodies, comparison with Parkinsons disease. Park. Relat.
Disord. 19, 4752. https://doi.org/10.1016/j.parkreldis.2012.07.002.
Iso-Markku, P., Kujala, U.M., Knittle, K., Polet, J., Vuoksimaa, E., Waller, K., 2022.
Physical activity as a protective factor for dementia and Alzheimers disease:
systematic review, meta-analysis and quality assessment of cohort and casecontrol
studies. Br. J. Sport. Med. 56, 701709. https://doi.org/10.1136/bjsports-2021-
104981.
Ivy, A.S., Rodriguez, F.G., Garcia, C., Chen, M.J., Russo-Neustadt, A.A., 2003.
Noradrenergic and serotonergic blockade inhibits BDNF mRNA activation following
exercise and antidepressant. Pharmacol. Biochem. Behav. 75, 8188. https://doi.
org/10.1016/S0091-3057(03)00044-3.
Jacobs, H.I.L., Müller-Ehrenberg, L., Priovoulos, N., Roebroeck, A., 2018. Curvilinear
locus coeruleus functional connectivity trajectories over the adult lifespan: a 7T MRI
study. Neurobiol. Ageing 69, 167176. https://doi.org/10.1016/j.
neurobiolageing.2018.05.021.
Jacobs, H.I.L., Becker, J.A., Kwong, K., Engels-Domínguez, N., Prokopiou, P.C., Papp, K.
V., Properzi, M., Hampton, O.L., dOleire Uquillas, F., Sanchez, J.S., Rentz, D.M., El
Fakhri, G., Normandin, M.D., Price, J.C., Bennett, D.A., Sperling, R.A., Johnson, K.
A., 2021a. In vivo and neuropathology data support locus coeruleus integrity as
indicator of Alzheimers disease pathology and cognitive decline. Sci. Transl. Med.
13, eabj2511 https://doi.org/10.1126/scitranslmed.abj2511.
Jacobs, H.I.L., Riphagen, J.M., Ramakers, I.H.G.B., Verhey, F.R.J., 2021b. Alzheimers
disease pathology: pathways between central norepinephrine activity, memory, and
neuropsychiatric symptoms. Mol. Psychiatry 26, 897906. https://doi.org/10.1038/
s41380-019-0437-x.
Jbabdi, S., Johansen-Berg, H., 2011. Tractography: where do we go from here? Brain
Connect 1, 169183. https://doi.org/10.1089/brain.2011.0033.
Jensen, O., Mazaheri, A., 2010. Shaping functional architecture by oscillatory alpha
activity: gating by inhibition. Front. Hum. Neurosci. 4. https://doi.org/10.3389/
fnhum.2010.00186.
Jessen, F., Spottke, A., Boecker, H., Brosseron, F., Buerger, K., Catak, C., Fliessbach, K.,
Franke, C., Fuentes, M., Heneka, M.T., Janowitz, D., Kilimann, I., Laske, C.,
Menne, F., Nestor, P., Peters, O., Priller, J., Pross, V., Ramirez, A., Schneider, A.,
Speck, O., Spruth, E.J., Teipel, S., Vukovich, R., Westerteicher, C., Wiltfang, J.,
Wolfsgruber, S., Wagner, M., Düzel, E., 2018. Design and rst baseline data of the
DZNE multicenter observational study on predementia Alzheimers disease
(DELCODE). Alz. Res. Ther. 10, 15. https://doi.org/10.1186/s13195-017-0314-2.
Jones, B.E., Moore, R.Y., 1977. Ascending projections of the locus coeruleus in the rat. II.
Autoradiographic study. Brain Res. 127, 2353. https://doi.org/10.1016/0006-8993
(77)90378-X.
Joshi, S., Li, Y., Kalwani, R.M., Gold, J.I., 2016. Relationships between pupil diameter
and neuronal activity in the locus coeruleus, colliculi, and cingulate cortex. Neuron
89, 221234. https://doi.org/10.1016/j.neuron.2015.11.028.
Jovanovic, P., Wang, Y., Vit, J.P., Novinbakht, E., Morones, N., Hogg, E., Tagliati, M.,
Riera, C.E., 2022. Sustained chemogenetic activation of locus coeruleus
norepinephrine neurons promotes dopaminergic neuron survival in synucleinopathy.
PLoS ONe 17, 121. https://doi.org/10.1371/journal.pone.0263074.
Kahn, M.M., Sansoni, P., Engleman, E.G., Melmon, K.L., 1985. Pharmacologic effects of
autacoids on subsets of T cells. Regulation of expression/function of histamine-2
receptors by a subset of suppressor cells. J. Clin. Investig. 75, 15781583.
Kalen, P., Rosegren, E., Lindvall, O., Bjorklund, A., 1989. Hippocampal noradrenaline
and serotonin release over 24 hours as measured by the dialysis technique in freely
moving rats: correlation to behavioural activity state, effect of handling and tail-
pinch. Eur. J. Neurosci. 1, 181188. https://doi.org/10.1111/j.1460-9568.1989.
tb00786.x.
Kalinin, S., Feinstein, D.L., Xu, H.-L., Huesa, G., Pelligrino, D.A., Galea, E., 2006.
Degeneration of noradrenergic bres from the locus coeruleus causes tight-junction
disorganisation in the rat brain. Eur. J. Neurosci. 24, 33933400. https://doi.org/
10.1111/j.1460-9568.2006.05223.x.
Kalinin, S., Gavrilyuk, V., Polak, P.E., Vasser, R., Zhao, J., Heneka, M.T., Feinstein, D.L.,
2007. Noradrenaline deciency in brain increases beta-amyloid plaque burden in an
animal model of Alzheimers disease. Neurobiol. Ageing 28, 12061214. https://doi.
org/10.1016/j.neurobiolageing.2006.06.003.
Kamal, S., Lappin, S.L., 2022. Biochemistry, Catecholamine Degradation, in: StatPearls.
StatPearls Publishing, Treasure Island (FL).
Kang, S.S., Liu, X., Ahn, E.H., Xiang, J., Manfredsson, F.P., Yang, X., Luo, H.R., Liles, L.C.,
Weinshenker, D., Ye, K., 2019. Norepinephrine metabolite DOPEGAL activates AEP
and pathological Tau aggregation in locus coeruleus. J. Clin. Investig. 130, 422437.
https://doi.org/10.1172/JCI130513.
Kang, S.S., Liu, X., Ahn, E.H., Xiang, J., Manfredsson, F.P., Yang, X., Luo, H.R., Liles, L.C.,
Weinshenker, D., Ye, K., 2020. Norepinephrine metabolite DOPEGAL activates AEP
and pathological Tau aggregation in locus coeruleus. J. Clin. Investig. 130, 422437.
https://doi.org/10.1172/JCI130513.
Kang, S.S., Ahn, E.H., Liu, X., Bryson, M., Miller, G.W., Weinshenker, D., Ye, K., 2021.
ApoE4 inhibition of VMAT2 in the locus coeruleus exacerbates Tau pathology in
Alzheimers disease. Acta Neuropathol. 142, 139158. https://doi.org/10.1007/
s00401-021-02315-1.
Kang, S.S., Meng, L., Zhang, X., Wu, Z., Mancieri, A., Xie, B., Liu, X., Weinshenker, D.,
Peng, J., Zhang, Z., Ye, K., 2022. Tau modication by the norepinephrine metabolite
DOPEGAL stimulates its pathology and propagation. Nat. Struct. Mol. Biol. 29,
292305. https://doi.org/10.1038/s41594-022-00745-3.
Katunina, E.A., Blokhin, V., Nodel, M.R., Pavlova, E.N., Kalinkin, A.L., Kucheryanu, V.G.,
Alekperova, L., Selikhova, M.V., Martynov, M.Yu, Ugrumov, M.V., 2023. Searching
for biomarkers in the blood of patients at risk of developing Parkinsons disease at
the prodromal stage. IJMS 24, 1842. https://doi.org/10.3390/ijms24031842.
Kaufman, S.K., Del Tredici, K., Thomas, T.L., Braak, H., Diamond, M.I., 2018. Tau
seeding activity begins in the transentorhinal/entorhinal regions and anticipates
phospho-tau pathology in Alzheimers disease and PART. Acta Neuropathol. 136,
5767. https://doi.org/10.1007/s00401-018-1855-6.
Kavelaars, A., van de Pol, M., Zijlstra, J., Heijnen, C.J., 1997. β2-Adrenergic activation
enhances interleukin-8 production by human monocytes. J. Neuroimmunol. 77,
211216.
Kelly, S.C., McKay, E.C., Beck, J.S., Collier, T.J., Dorrance, A.M., Counts, S.E., 2019.
Locus coeruleus degeneration induces forebrain vascular pathology in a transgenic
rat model of Alzheimers disease. JAD 70, 371388. https://doi.org/10.3233/JAD-
190090.
Kempadoo, K.A., Mosharov, E.V., Choi, S.J., Sulzer, D., Kandel, E.R., 2016. Dopamine
release from the locus coeruleus to the dorsal hippocampus promotes spatial
learning and memory. Proc. Natl. Acad. Sci. U.S.A. 113, 1483514840. https://doi.
org/10.1073/pnas.1616515114.
Keren, N.I., Taheri, S., Vazey, E.M., Morgan, P.S., Granholm, A.-C.E., Aston-Jones, G.S.,
Eckert, M.A., 2015. Histologic validation of locus coeruleus MRI contrast in post-
mortem tissue. NeuroImage 113, 235245. https://doi.org/10.1016/j.
neuroimage.2015.03.020.
Kety, S.S., 1972. The possible role of the adrenergic systems of the cortex in learning. Res
Publ. Assoc. Res Nerv. Ment. Dis. 50, 376389.
Kiyashchenko, L.I., Mileykovskiy, B.Y., Lai, Y.-Y., Siegel, J.M., 2001. Increased and
decreased muscle tone with orexin (hypocretin) microinjections in the locus
coeruleus and pontine inhibitory area. J. Neurophysiol. 85, 20082016. https://doi.
org/10.1152/jn.2001.85.5.2008.
Kjaerby, C., Andersen, M., Hauglund, N., Untiet, V., Dall, C., Sigurdsson, B., Ding, F.,
Feng, J., Li, Y., Weikop, P., Hirase, H., Nedergaard, M., 2022. Memory-enhancing
properties of sleep depend on the oscillatory amplitude of norepinephrine. Nat.
Neurosci. 25, 10591070. https://doi.org/10.1038/s41593-022-01102-9.
Klukowski, G., Harley, C.W., 1994. Locus coeruleus activation induces perforant path-
evoked population spike potentiation in the dentate gyrus of awake rat. Exp. Brain
Res. 102, 165170. https://doi.org/10.1007/BF00232449.
Knudsen, E.I., 2011. Control from below: the role of a midbrain network in spatial
attention: control from below. Eur. J. Neurosci. 33, 19611972. https://doi.org/
10.1111/j.1460-9568.2011.07696.x.
Kobayashi, K., Shikano, K., Kuroiwa, M., Horikawa, M., Ito, W., Nishi, A., Segi-
Nishida, E., Suzuki, H., 2022. Noradrenaline activation of hippocampal dopamine
D1 receptors promotes antidepressant effects. Proc. Natl. Acad. Sci. USA 119,
e2117903119. https://doi.org/10.1073/pnas.2117903119.
Kong, Y., Ruan, L., Qian, L., Liu, X., Le, Y., 2010. Norepinephrine promotes microglia to
uptake and degrade amyloid β peptide through upregulation of mouse formyl
peptide receptor 2 and induction of insulin-degrading enzyme. J. Neurosci. 30,
1184811857. https://doi.org/10.1523/JNEUROSCI.2985-10.2010.
Koppel, J., Jimenez, H., Adrien, L., Chang, H., Malhotra, E., Davies, P, A.K., 2019.
Increased tau phosphorylation follows impeded dopamine clearance in a P301L and
novel P301L/COMT-deleted (DM) tau mouse model. J. Neurochem 148, 127135.
https://doi.org/10.1111/jnc.14593.
Kreisl, W.C., Lyoo, C.H., McGwier, M., Snow, J., Jenko, K.J., Kimura, N., Corona, W.,
Morse, C.L., Zoghbi, S.S., Pike, V.W., McMahon, F.J., Turner, R.S., Innis, R.B.,
Biomarkers Consortium PET Radioligand Project Team, 2013. In vivo radioligand
binding to translocator protein correlates with severity of Alzheimers disease. Brain
136, 22282238. https://doi.org/10.1093/brain/awt145.
Kreisl, W.C., Lyoo, C.H., Liow, J.-S., Wei, M., Snow, J., Page, E., Jenko, K.J., Morse, C.L.,
Zoghbi, S.S., Pike, V.W., Turner, R.S., Innis, R.B., 2016. 11C-PBR28 binding to
translocator protein increases with progression of Alzheimers disease. Neurobiol.
Ageing 44, 5361. https://doi.org/10.1016/j.neurobiolageing.2016.04.011.
Kuwamizu, R., Yamazaki, Y., Aoike, N., Ochi, G., Suwabe, K., Soya, H., 2022. Pupil-
linked arousal with very light exercise: pattern of pupil dilation during graded
exercise. J. Physiol. Sci. 72, 23. https://doi.org/10.1186/s12576-022-00849-x.
Lancini, E., Haag, L., Bartl, F., Rühling, M., Ashton, N.J., Zetterberg, H., Düzel, E.,
H¨
ammerer, D., Betts, M.J., 2023. CSF and PET biomarkers for noradrenergic
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
16
dysfunction in neurodegenerative diseases: a systematic review and meta-analysis.
Brain Commun. https://doi.org/10.1093/braincomms/fcad085.
Langley, J., Hussain, S., Huddleston, D.E., Bennett, I.J., Hu, X.P., 2022. Impact of locus
coeruleus and its projections on memory and ageing. Brain Connect. 12, 223233.
https://doi.org/10.1089/brain.2020.0947.
Lapiz, M.S., Mateo, Y., Durkin, S., Parker, T., Marsden, C.A., 2001. Effects of central
noradrenaline depletion by the selective neurotoxin DSP-4 on the behaviour of the
isolated rat in the elevated plus maze and water maze. Psychopharmacology 155,
251259. https://doi.org/10.1007/s002130100702.
Lecas, J.C., 2004. Locus coeruleus activation shortens synaptic drive while decreasing
spike latency and jitter in sensorimotor cortex. Implications for neuronal integration.
Eur. J. Neurosci. 19, 25192530. https://doi.org/10.1111/j.0953-
816X.2004.03341.x.
Lee, T.-H., Greening, S.G., Ueno, T., Clewett, D., Ponzio, A., Sakaki, M., Mather, M.,
2018. Arousal increases neural gain via the locus coeruleusnoradrenaline system in
younger adults but not in older adults. Nat. Hum. Behav. 2, 356366. https://doi.
org/10.1038/s41562-018-0344-1.
Lee, T.-H., Kim, S.H., Katz, B., Mather, M., 2020. The decline in intrinsic connectivity
between the salience network and locus coeruleus in older adults: implications for
distractibility. Front. Ageing Neurosci. 12, 2. https://doi.org/10.3389/
fnagi.2020.00002.
Lehnert, H., Schulz, C., Dieterich, K., 1998. Physiological and neurochemical aspects of
corticotropin-releasing factor actions in the brain: the role of the locus coeruleus.
Neurochem Res 23, 10391052. https://doi.org/10.1023/A:1020751817723.
Levey, A.I., Qiu, D., Zhao, L., Hu, W.T., Duong, D.M., Higginbotham, L., Dammer, E.B.,
Seyfried, N.T., Wingo, T.S., Hales, C.M., G´
amez Tansey, M., Goldstein, D.S.,
Abrol, A., Calhoun, V.D., Goldstein, F.C., Hajjar, I., Fagan, A.M., Galasko, D.,
Edland, S.D., Hanfelt, J., Lah, J.J., Weinshenker, D., 2022. A phase II study
repurposing atomoxetine for neuroprotection in mild cognitive impairment. Brain
145, 19241938. https://doi.org/10.1093/brain/awab452.
Lew, C.H., Petersen, C., Neylan, T.C., Grinberg, L.T., 2021. Tau-driven degeneration of
sleep- and wake-regulating neurons in Alzheimers disease. Sleep. Med. Rev. 60,
101541 https://doi.org/10.1016/j.smrv.2021.101541.
Li, M., Yao, W., Li, S., Xi, J., 2015. Norepinephrine induces the expression of interleukin-
6 via β-adrenoreceptor-NAD(P)H oxidase system -NF-κB dependent signal pathway
in U937 macrophages. Biochem. Biophys. Res. Commun. 460, 10291034. https://
doi.org/10.1016/j.bbrc.2015.02.172.
Li, M., Liu, S., Zhu, H., Guo, Z., Zhi, Y., Liu, R., Jiang, Z., Liang, X., Hu, H., Zhu, J., 2022.
Decreased locus coeruleus signal associated with Alzheimers disease based on
neuromelanin-sensitive magnetic resonance imageing technique. Front Neurosci. 16,
1014485. https://doi.org/10.3389/fnins.2022.1014485.
Li, S., Jin, M., Zhang, D., Yang, T., Koeglsperger, T., Fu, H., Selkoe, D.J., 2013.
Environmental novelty activates β2-adrenergic signaling to prevent the impairment
of hippocampal LTP by Aβ oligomers. Neuron 77, 929941. https://doi.org/
10.1016/j.neuron.2012.12.040.
Li, Y., Wang, C., Wang, J., Zhou, Y., Ye, F., Zhang, Y., Cheng, X., Huang, Z., Liu, K.,
Fei, G., Zhong, C., Zeng, M., Jin, L., 2019. Mild cognitive impairment in de novo
Parkinsons disease: a neuromelanin MRI study in locus coeruleus. Mov. Disord. 34,
884892. https://doi.org/10.1002/mds.27682.
Liebe, T., Kaufmann, J., H¨
ammerer, D., Betts, M., Walter, M., 2022. In vivo tractography
of human locus coeruleusrelation to 7T resting state fMRI, psychological measures
and single subject validity. Mol. Psychiatry 27, 49844993. https://doi.org/
10.1038/s41380-022-01761-x.
Liu, K.Y., Marijatta, F., H¨
ammerer, D., Acosta-Cabronero, J., Düzel, E., Howard, R.J.,
2017. Magnetic resonance imageing of the human locus coeruleus: a systematic
review. Neurosci. Biobehav. Rev. 83, 325355. https://doi.org/10.1016/j.
neubiorev.2017.10.023.
Liu, K.Y., Acosta-Cabronero, J., Cardenas-Blanco, A., Loane, C., Berry, A.J., Betts, M.J.,
Kievit, R.A., Henson, R.N., Düzel, E., Howard, R., H¨
ammerer, D., 2019. In vivo
visualization of age-related differences in the locus coeruleus. Neurobiol. Ageing 74,
101111. https://doi.org/10.1016/j.neurobiolageing.2018.10.014.
Liu, K.Y., Kievit, R.A., Tsvetanov, K.A., Betts, M.J., Düzel, E., Rowe, J.B., Cam-CAN,
Tyler, L.K., Brayne, C., Bullmore, E.T., Calder, A.C., Cusack, R., Dalgleish, T.,
Duncan, J., Henson, R.N., Matthews, F.E., Marslen-Wilson, W.D., Rowe, J.B.,
Shafto, M.A., Campbell, K., Cheung, T., Davis, S., Geerligs, L., Kievit, R.,
McCarrey, A., Mustafa, A., Price, D., Samu, D., Taylor, J.R., Treder, M., Tsvetanov, K.
A., Belle, J., van, Williams, N., Bates, L., Emery, T., Erzinçlioglu, S., Gadie, A.,
Gerbase, S., Georgieva, S., Hanley, C., Parkin, B., Troy, D., Auer, T., Correia, M.,
Gao, L., Green, E., Henriques, R., Allen, J., Amery, G., Amunts, L., Barcroft, A.,
Castle, A., Dias, C., Dowrick, J., Fair, M., Fisher, H., Goulding, A., Grewal, A.,
Hale, G., Hilton, A., Johnson, F., Johnston, P., Kavanagh-Williamson, T.,
Kwasniewska, M., McMinn, A., Norman, K., Penrose, J., Roby, F., Rowland, D.,
Sargeant, J., Squire, M., Stevens, B., Stoddart, A., Stone, C., Thompson, T., Yazlik, O.,
Barnes, D., Dixon, M., Hillman, J., Mitchell, J., Villis, L., Howard, R., H¨
ammerer, D.,
2020. Noradrenergic-dependent functions are associated with age-related locus
coeruleus signal intensity differences. Nat. Commun. https://doi.org/10.1038/
s41467-020-15410-w.
Liu, K.Y., Acosta-Cabronero, J., Hong, Y.T., Yi, Y.-J., H¨
ammerer, D., Howard, R., 2021.
FDG-PET assessment of the locus coeruleus in Alzheimers disease. Neuroimage: Rep.
1, 100002 https://doi.org/10.1016/j.ynirp.2020.100002.
L´
opez-Ortiz, S., Lista, S., Valenzuela, P.L., Pinto-Fraga, J., Carmona, R., Caraci, F.,
Caruso, G., Toschi, N., Emanuele, E., Gabelle, A., Nistic`
o, R., Garaci, F., Lucia, A.,
Santos-Lozano, A., 2023. Effects of physical activity and exercise interventions on
Alzheimers disease: an umbrella review of existing meta-analyses. J. Neurol. 270,
711725. https://doi.org/10.1007/s00415-022-11454-8.
Lotankar, S., Prabhavalkar, K.S., Bhatt, L.K., 2017. Biomarkers for Parkinsons disease:
recent advancement. Neurosci. Bull. 33, 585597. https://doi.org/10.1007/s12264-
017-0183-5.
L¨
ovd´
en, M., Fratiglioni, L., Glymour, M.M., Lindenberger, U., Tucker-Drob, E.M., 2020.
Education and cognitive functioning across the life span. Psychol. Sci. Public Interest
21, 641. https://doi.org/10.1177/1529100620920576.
Ludwig, M., Wienke, C., Betts, M.J., Zaehle, T., H¨
ammerer, D., 2021. Current challenges
in reliably targeting the noradrenergic locus coeruleus using transcutaneous
auricular vagus nerve stimulation (taVNS). Auton. Neurosci. 236, 102900 https://
doi.org/10.1016/j.autneu.2021.102900.
Luissint, A.-C., Artus, C., Glacial, F., Ganeshamoorthy, K., Couraud, P.-O., 2012. Tight
junctions at the blood brain barrier: physiological architecture and disease-
associated dysregulation. Fluids Barriers CNS 9, 23. https://doi.org/10.1186/2045-
8118-9-23.
Luppi, P.-H., Aston-Jones, G., Akaoka, H., Chouvet, G., Jouvet, M., 1995. Afferent
projections to the rat locus coeruleus demonstrated by retrograde and anterograde
tracing with cholera-toxin B subunit and Phaseolus vulgaris leucoagglutinin.
Neuroscience 65, 119160. https://doi.org/10.1016/0306-4522(94)00481-J.
Lushchak, V.I., Duszenko, M., Gospodaryov, D.V., Garaschuk, O., 2021. Oxidative stress
and energy metabolism in the brain: midlife as a turning point. Antioxidants 10,
1715. https://doi.org/10.3390/antiox10111715.
Maass, A., Lockhart, S.N., Harrison, T.M., Bell, R.K., Mellinger, T., Swinnerton, K.,
Baker, S.L., Rabinovici, G.D., Jagust, W.J., 2018. Entorhinal tau pathology, episodic
memory decline, and neurodegeneration in ageing. J. Neurosci. 38, 530543.
https://doi.org/10.1523/JNEUROSCI.2028-17.2017.
Madelung, C.F., Meder, D., Fuglsang, S.A., Marques, M.M., Boer, V.O., Madsen, K.H.,
Petersen, E.T., Hejl, A., Løkkegaard, A., Siebner, H.R., 2022. Locus coeruleus shows a
spatial pattern of structural disintegration in Parkinsons disease. Mov. Disord. 37,
479489. https://doi.org/10.1002/mds.28945.
Manta, S., El Mansari, M., Debonnel, G., Blier, P., 2013. Electrophysiological and
neurochemical effects of long-term vagus nerve stimulation on the rat
monoaminergic systems. Int. J. Neuropsychopharmacol. 16, 459470. https://doi.
org/10.1017/S1461145712000387.
Marqui´
e, M., Normandin, M.D., Vanderburg, C.R., Costantino, I.M., Bien, E.A., Rycyna, L.
G., Klunk, W.E., Mathis, C.A., Ikonomovic, M.D., Debnath, M.L., Vasdev, N.,
Dickerson, B.C., Gomperts, S.N., Growdon, J.H., Johnson, K.A., Frosch, M.P.,
Hyman, B.T., G´
omez-Isla, T., 2015. Validating novel tau positron emission
tomography tracer [F-18]-AV-1451 (T807) on postmortem brain tissue: Validation of
PET Tracer. Ann. Neurol. 78, 787800. https://doi.org/10.1002/ana.24517.
Matchett, B.J., Grinberg, L.T., Theolas, P., Murray, M.E., 2021. The mechanistic link
between selective vulnerability of the locus coeruleus and neurodegeneration in
Alzheimers disease. Acta Neuropathol. 141, 631650. https://doi.org/10.1007/
s00401-020-02248-1.
Mather, M., 2021. Noradrenaline in the ageing brain: promoting cognitive reserve or
accelerating Alzheimers disease? Semin. Cell Dev. Biol. 116, 108124. https://doi.
org/10.1016/j.semcdb.2021.05.013.
Mather, M., Harley, C.W., 2016. The locus coeruleus: essential for maintaining cognitive
function and the ageing brain. Trends Cogn. Sci. 20, 214226. https://doi.org/
10.1016/j.tics.2016.01.001.
Mather, M., Clewett, D., Sakaki, M., Harley, C.W., 2016. Norepinephrine ignites local
hotspots of neuronal excitation: How arousal amplies selectivity in perception and
memory. Behav. Brain Sci. 39, e200 https://doi.org/10.1017/S0140525X15000667.
Matschke, L.A., Komadowski, M.A., St¨
ohr, A., Lee, B., Henrich, M.T., Griesbach, M.,
Rinn´
e, S., Geibl, F.F., Chiu, W.H., Koprich, J.B., Brotchie, J.M., Kiper, A.K., Dolga, A.
M., Oertel, W.H., Decher, N., 2022. Enhanced ring of locus coeruleus neurons and
SK channel dysfunction are conserved in distinct models of prodromal Parkinsons
disease. Sci. Rep. 12, 114. https://doi.org/10.1038/s41598-022-06832-1.
McBurney-Lin, J., Lu, J., Zuo, Y., Yang, H., 2019. Locus coeruleus-norepinephrine
modulation of sensory processing and perception: a focused review. Neurosci.
Biobehav. Rev. 105, 190199. https://doi.org/10.1016/j.neubiorev.2019.06.009.
McCall, J.G., Siuda, E.R., Bhatti, D.L., Lawson, L.A., McElligott, Z.A., Stuber, G.D.,
Bruchas, M.R., 2017. Locus coeruleus to basolateral amygdala noradrenergic
projections promote anxiety-like behavior. eLife 6, e18247. https://doi.org/
10.7554/eLife.18247.
McCormick, A., 1989. CholinergRandnoradrenergRmodulationof thalamocortical
processing 12.
McMorris, T., 2003. Incremental exercise, plasma concentrations of catecholamines,
reaction time, and motor time during performance of a noncompatible choice
response time task. PMS 97, 590. https://doi.org/10.2466/PMS.97.6.590-604.
McMorris, T., Collard, K., Corbett, J., Dicks, M., Swain, J.P., 2008. A test of the
catecholamines hypothesis for an acute exercisecognition interaction. Pharmacol.
Biochem. Behav. 89, 106115. https://doi.org/10.1016/j.pbb.2007.11.007.
McNamee, E.N., Ryan, K.M., Grifn, ´
E.W., Gonz´
alez-Reyes, R.E., Ryan, K.J., Harkin, A.,
Connor, T.J., 2010. Noradrenaline acting at central β-adrenoceptors induces
interleukin-10 and suppressor of cytokine signaling-3 expression in rat brain:
Implications for neurodegeneration. Brain Behav. Immun. 24, 660671. https://doi.
org/10.1016/j.bbi.2010.02.005.
Meisenzahl, E.M., Schmitt, G.J., Scheuerecker, J., M¨
oller, H.-J., 2007. The role of
dopamine for the pathophysiology of schizophrenia. Int. Rev. Psychiatry 19,
337345. https://doi.org/10.1080/09540260701502468.
Mejias-Aponte, C.A., 2016. Specicity and impact of adrenergic projections to the
midbrain dopamine system. Brain Res. 1641, 258273. https://doi.org/10.1016/j.
brainres.2016.01.036.
Mello-Carpes, P.B., de Vargas, da Silva, Gayer, L., Roehrs, M.C., Izquierdo, I, R., 2016.
Hippocampal noradrenergic activation is necessary for object recognition memory
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
17
consolidation and can promote BDNF increase and memory persistence. Neurobiol.
Learn Mem. 127, 8492. https://doi.org/10.1016/j.nlm.2015.11.014.
Mintzer, J., Lanctˆ
ot, K.L., Scherer, R.W., Rosenberg, P.B., Herrmann, N., van Dyck, C.H.,
Padala, P.R., Brawman-Mintzer, O., Porsteinsson, A.P., Lerner, A.J., Craft, S.,
Levey, A.I., Burke, W., Perin, J., Shade, D., Kominek, V., Michalak, H., Ni, G.,
Good, S., Mecca, A., Salem-Spencer, S., Keltz, M., Morales, E., Clark, E.D.,
Williams, A., Kindy, A., Freeman, R., Jamil, N., Schultz, M., Sami, S., Padala, K.P.,
Parkes, C., Lah, J., Vaughn, P., Hales, C., Rapoport, M., Gallagher, D., Li, A.,
Sandra, B., Vieira, D., Ruthirakuhan, M., Babani, P., ADMET 2 Research Group,
2021. Effect of methylphenidate on apathy in patients with Alzheimer disease: the
ADMET 2 randomized clinical trial. JAMA Neurol. 1324. https://doi.org/10.1001/
jamaneurol.2021.3356.
Moises, H.C., Waterhouse, B.D., Woodward, D.J., 1981. Locus coeruleus stimulation
potentiates Purkinje cell responses to afferent input: the climbing ber system. Brain
Res. 222, 4364. https://doi.org/10.1016/0006-8993(81)90939-2.
Molinoff, P.B., 1984.
α
- and β-adrenergic receptor subtypes. Drugs 28, 115. https://doi.
org/10.2165/00003495-198400282-00002.
Molinoff, P.B., Axelrod, J., 1971. Biochemistry of catecholamines. Annu. Rev. Biochem.
40, 465500. https://doi.org/10.1146/annurev.bi.40.070171.002341.
Mori, K., Ozaki, E., Zhang, B., Yang, L., Yokoyama, A., Takeda, I., Maeda, N.,
Sakanaka, M., Tanaka, J., 2002. Effects of norepinephrine on rat cultured microglial
cells that express
α
1,
α
2, β1 and β2 adrenergic receptors. Neuropharmacology 43,
10261034. https://doi.org/10.1016/S0028-3908(02)00211-3.
Murray, M.E., Moloney, C.M., Kouri, N., Syrjanen, J.A., Matchett, B.J., Rothberg, D.M.,
Tranovich, J.F., Sirmans, T.N.H., Wiste, H.J., Boon, B.D.C., Nguyen, A.T.,
Reichard, R.R., Dickson, D.W., Lowe, V.J., Dage, J.L., Petersen, R.C., Jack, C.R.,
Knopman, D.S., Vemuri, P., Graff-Radford, J., Mielke, M.M., 2022. Global
neuropathologic severity of Alzheimers disease and locus coeruleus vulnerability
inuences plasma phosphorylated tau levels. Mol. Neurodegener. 17, 85. https://
doi.org/10.1186/s13024-022-00578-0.
Naka, F., Shiga, T., Yaguchi, M., Okado, N., 2002. An enriched environment increases
noradrenaline concentration in the mouse brain. Brain Res. 924, 124126. https://
doi.org/10.1016/S0006-8993(01)03257-7.
Nakamura, S., Sakaguchi, T., 1990. Development and plasticity of the locus coeruleus: a
review of recent physiological and pharmacological experimentation. Prog.
Neurobiol. 34, 505526. https://doi.org/10.1016/0301-0082(90)90018-c.
Ni, Y., Zhao, X., Bao, G., Zou, L., Teng, L., Wang, Z., Song, M., Xiong, J., Bai, Y., Pei, G.,
2006. Activation of β2-adrenergic receptor stimulates γ-secretase activity and
accelerates amyloid plaque formation. Nat. Med 12, 13901396. https://doi.org/
10.1038/nm1485.
Nieuwenhuis, S., Aston-Jones, G., Cohen, J.D., 2005. Decision making, the P3, and the
locus coeruleus-norepinephrine system. Psychol. Bull. 131, 510532. https://doi.
org/10.1037/0033-2909.131.4.510.
Nieuwenhuis, S., De Geus, E.J., Aston-Jones, G., 2011. The anatomical and functional
relationship between the P3 and autonomic components of the orienting response.
Psychophysiology 48, 162175. https://doi.org/10.1111/j.1469-8986.2010.01057.
x.
Nosaka, S., Yasunaga, K., Tamai, S., 1982. Vagal cardiac preganglionic neurons:
distribution, cell types, and reex discharges. Am. J. Physiol. 243, R92R98. https://
doi.org/10.1152/ajpregu.1982.243.1.R92.
OCallaghan, C., Hezemans, F.H., Ye, R., Rua, C., Jones, P.S., Murley, A.G., Holland, N.,
Regenthal, R., Tsvetanov, K.A., Wolpe, N., Barker, R.A., Williams-Gray, C.H.,
Robbins, T.W., Passamonti, L., Rowe, J.B., 2021. Locus coeruleus integrity and the
effect of atomoxetine on response inhibition in Parkinsons disease. Brain 144,
25132526. https://doi.org/10.1093/brain/awab142.
OMalley, S.S., Chen, T.B., Francis, B.E., Gibson, R.E., Burns, H.D., DiSalvo, J., Bayne, M.
L., Wetzel, J.M., Nagarathnam, D., Marzabadi, M., Gluchowski, C., Chang, R.S.L.,
1998. Characterization of specic binding of w125IxL-762,459, a selective a1A-
adrenoceptor radioligand to rat and human tissues.
ONeill, E., Harkin, A., 2018. Targeting the noradrenergic system for anti-inammatory
and neuroprotective effects: implications for Parkinsons disease. Neural Regen. Res.
13, 13321337. https://doi.org/10.4103/1673-5374.235219.
Oakes, H.V., McWethy, D., Ketchem, S., Tran, L., Phillips, K., Oakley, L., Smeyne, R.J.,
Pond, B.B., 2021. Effect of chronic methylphenidate treatment in a female
experimental model of Parkinsonism. Neurotox. Res. 39, 667676. https://doi.org/
10.1007/s12640-021-00347-9.
Olivieri, P., Lagarde, J., Lehericy, S., Valabr`
egue, R., Michel, A., Mac´
e, P., Caill´
e, F.,
Gervais, P., Bottlaender, M., Sarazin, M., 2019. Early alteration of the locus
coeruleus in phenotypic variants of Alzheimers disease. Ann. Clin. Transl. Neurol. 6,
13451351. https://doi.org/10.1002/acn3.50818.
Osorio-Forero, A., Cardis, R., Vantomme, G., Guillaume-Gentil, A., Katsioudi, G.,
Devenoges, C., Fernandez, L.M.J., Lüthi, A., 2021. Noradrenergic circuit control of
non-REM sleep substates. e7 Curr. Biol. 31, 50095023. https://doi.org/10.1016/j.
cub.2021.09.041.
Osorio-Forero, A., Cherrad, N., Banterle, L., Fernandez, L.M.J., Lüthi, A., 2022. When the
locus coeruleus speaks up in sleep: recent insights, emerging perspectives. IJMS 23,
5028. https://doi.org/10.3390/ijms23095028.
Pajkossy, P., Sz˝
oll˝
osi, ´
A., Demeter, G., Racsm´
any, M., 2018. Physiological measures of
dopaminergic and noradrenergic activity during attentional set shifting and reversal.
Front. Psychol. 9.
Palmer, A.M., Francis, P.T., Bowen, D.M., Benton, J.S., Neary, D., Mann, D.M.A.,
Snowden, J.S., 1987. Catecholaminergic neurones assessed ante-mortem in
Alzheimers disease. Brain Res. 414, 365375. https://doi.org/10.1016/0006-8993
(87)90018-7.
Paredes-Rodriguez, E., Vegas-Suarez, S., Morera-Herreras, T., De Deurwaerdere, P.,
Miguelez, C., 2020. The noradrenergic system in Parkinsons disease. Front.
Pharmacol. 11, 435. https://doi.org/10.3389/fphar.2020.00435.
Parent, J.H., Ciampa, C.J., Harrison, T.M., Adams, J.N., Zhuang, K., Betts, M.J.,
Maass, A., Winer, J.R., Jagust, W.J., Berry, A.S., 2022. Locus coeruleus
catecholamines link neuroticism and vulnerability to tau pathology in ageing.
NeuroImage 263, 119658. https://doi.org/10.1016/j.neuroimage.2022.119658.
Perry, V.H., 2010. Contribution of systemic inammation to chronic neurodegeneration.
Acta Neuropathol. 120, 277286. https://doi.org/10.1007/s00401-010-0722-x.
Perry, V.H., Cunningham, C., Holmes, C., 2007. Systemic infections and inammation
affect chronic neurodegeneration. Nat. Rev. Immunol. 7, 161167. https://doi.org/
10.1038/nri2015.
Peuker, E.T., Filler, T.J., 2002. The nerve supply of the human auricle [WWW
Document]. URL https://onlinelibrary.wiley.com/doi/abs/10.1002/ca.1089
(accessed 4.29.23).
Pintus, R., Riggi, M., Cannarozzo, C., Valeri, A., de Leo, G., Romano, M., Gulino, R.,
Leanza, G., 2018. Essential role of hippocampal noradrenaline in the regulation of
spatial working memory and TDP-43 tissue pathology. J. Comp. Neurol. 526,
11311147. https://doi.org/10.1002/cne.24397.
Poe, G.R., Foote, S., Eschenko, O., Johansen, J.P., Bouret, S., Aston-Jones, G., Harley, C.
W., Manahan-Vaughan, D., Weinshenker, D., Valentino, R., Berridge, C.,
Chandler, D.J., Waterhouse, B., Sara, S.J., 2020. Locus coeruleus: a new look at the
blue spot. Nat. Rev. Neurosci. 21, 644659. https://doi.org/10.1038/s41583-020-
0360-9.
Polich, J., 2007. Updating P300: an integrative theory of P3a and P3b. Clin.
Neurophysiol. 118, 21282148. https://doi.org/10.1016/j.clinph.2007.04.019.
Porat, S., Sibilia, F., Yoon, J., Shi, Y., Dahl, M.J., Werkle-Bergner, M., Düzel, S.,
Bodammer, N., Lindenberger, U., Kühn, S., Mather, M., 2022. Age differences in
diffusivity in the locus coeruleus and its ascending noradrenergic tract. NeuroImage
251, 119022. https://doi.org/10.1016/j.neuroimage.2022.119022.
Pozzi, L., Invernizzi, R., Cervo, L., Vallebuona, F., Samanin, R., 1994. Evidence that
extracellular concentrations of dopamine are regulated by noradrenergic neurons in
the frontal cortex of rats. J. Neurochem. 63, 195200. https://doi.org/10.1046/
j.1471-4159.1994.63010195.x.
Prange, S., Klinger, H., Laurencin, C., Danaila, T., Thobois, S., 2022. Depression in
patients with Parkinsons disease: current understanding of its neurobiology and
implications for treatment. Drugs Ageing 39, 417439. https://doi.org/10.1007/
s40266-022-00942-1.
Prasuhn, J., Prasuhn, M., Fellbrich, A., Strautz, R., Lemmer, F., Dreischmeier, S.,
Kasten, M., Münte, T.F., Hanssen, H., Heldmann, M., Brüggemann, N., 2021.
Association of locus coeruleus and substantia nigra pathology with cognitive and
motor functions in patients with Parkinson disease. Neurology 97, e1007e1016.
https://doi.org/10.1212/WNL.0000000000012444.
Press, Y., Punchik, B., Kagan, E., Berzak, A., Freud, T., Dwolatzky, T., 2021.
Methylphenidate for mild cognitive impairment: an exploratory 3-day, randomized,
double-blind, placebo-controlled trial. Front. Med. 8, 594228 https://doi.org/
10.3389/fmed.2021.594228.
Prince, M.J., Wimo, A., Guerchet, M.M., Ali, G.C., Wu, Y.-T., Prina, M., 2015. World
Alzheimer Report 2015 - The Global Impact of Dementia: An analysis of prevalence,
incidence, cost and trends.
Qin, L., Wu, X., Block, M.L., Liu, Y., Breese, G.R., Hong, J.-S., Knapp, D.J., Crews, F.T.,
2007. Systemic LPS causes chronic neuroinammation and progressive
neurodegeneration. Glia 55, 453462. https://doi.org/10.1002/glia.20467.
Raskind, M.A., 1984. Norepinephrine and MHPG levels in CSF and plasma in Alzheimers
disease. Arch. Gen. Psychiatry 41, 343. https://doi.org/10.1001/
archpsyc.1984.01790150033006.
Raskind, M.A., Sadowsky, C.H., Sigmund, W.R., Beitler, P.J., Auster, S.B., 1997. Effect of
tacrine on language, praxis, and noncognitive behavioral problems in Alzheimer
disease. Arch. Neurol. 54, 836840. https://doi.org/10.1001/
archneur.1997.00550190026010.
Raskind, M.H., Goldberg, R.J., Higgins, E.L., Herman, K.L., 1999. Patterns of change and
predictors of success in individuals with learning disabilities: results from a twenty-
year longitudinal study. Learn. Disabil. Res. Pract. 14, 3549. https://doi.org/
10.1207/sldrp1401_4.
Rassnick, S., Sved, A., Rabin, B., 1994. Locus coeruleus stimulation by corticotropin-
releasing hormone suppresses in vitro cellular immune responses. J. Neurosci.: Off.
J. Soc. Neurosci. 14, 60336040. https://doi.org/10.1523/JNEUROSCI.14-10-
06033.1994.
Remy, P., Doder, M., Lees, A., Turjanski, N., Brooks, D., 2005. Depression in Parkinsons
disease: loss of dopamine and noradrenaline innervation in the limbic system. Brain
128, 13141322. https://doi.org/10.1093/brain/awh445.
Riphagen, J.M., van Egroo, M., Jacobs, H.I.L., 2021. Elevated norepinephrine
metabolism gauges Alzheimers disease-related pathology and memory decline.
J. Alzheimers Dis. 80, 521526. https://doi.org/10.3233/JAD-201411.
Rodenkirch, C., Liu, Y., Schriver, B.J., Wang, Q., 2019. Locus coeruleus activation
enhances thalamic feature selectivity via norepinephrine regulation of intrathalamic
circuit dynamics. Nat. Neurosci. 22, 120133. https://doi.org/10.1038/s41593-018-
0283-1.
Rommelfanger, K.S., Weinshenker, D., 2007. Norepinephrine: the redheaded stepchild of
Parkinsons disease. Biochem. Pharmacol. 74, 177190. https://doi.org/10.1016/j.
bcp.2007.01.036.
Ross, J.A., Van Bockstaele, E.J., 2021. The locus coeruleus- norepinephrine system in
stress and arousal: unraveling historical, current, and future perspectives. Front.
Psychiatry 11, 123. https://doi.org/10.3389/fpsyt.2020.601519.
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
18
Rudy, C.C., Hunsberger, H.C., Weitzner, D.S., Reed, M.N., 2015. The role of the tripartite
glutamatergic synapse in the pathophysiology of Alzheimers disease. Ageing Dis. 6,
131148. https://doi.org/10.14336/AD.2014.0423.
Ruffoli, R., Giorgi, F.S., Pizzanelli, C., Murri, L., Paparelli, A., Fornai, F., 2011. The
chemical neuroanatomy of vagus nerve stimulation. J. Chem. Neuroanat. Recent
Adv. Pept. Neurotransm. 42, 288296. https://doi.org/10.1016/j.
jchemneu.2010.12.002.
Samuels, E.R., Szabadi, E., 2008. Functional neuroanatomy of the noradrenergic locus
coeruleus: its roles in the regulation of arousal and autonomic function part II:
physiological and pharmacological manipulations and pathological alterations of
locus coeruleus activity in humans. Curr. Neuropharmacol. 6, 254285. https://doi.
org/10.2174/157015908785777193.
Saper, C.B., Fuller, P.M., 2017. Wakesleep circuitry: an overview. Curr. Opin.
Neurobiol. 44, 186192. https://doi.org/10.1016/j.conb.2017.03.021.
Sara, S., 2015. Locus Coeruleus in time with the making of memories. Curr. Opin.
Neurobiol. 35, 8794. https://doi.org/10.1016/j.conb.2015.07.004.
Sara, S.J., 2009. The locus coeruleus and noradrenergic modulation of cognition. Nat.
Rev. Neurosci. 10, 211223. https://doi.org/10.1038/nrn2573.
Sasaki, M., Shibata, E., Tohyama, K., Takahashi, J., Otsuka, K., Tsuchiya, K.,
Takahashi, S., Ehara, S., Terayama, Y., Sakai, A., 2006. Neuromelanin magnetic
resonance imageing of locus ceruleus and substantia nigra in Parkinsons disease.
NeuroReport 17, 12151218. https://doi.org/10.1097/01.wnr.0000227984.84927.
a7.
Schanberg, S.M., Schildkraut, J.J., Breese, G.R., Kopin, I.J., 1968. Metabolism of
normetanephrine-H3 in rat brain–identication of conjugated 3-methoxy-4-hydro-
phenylglycol as the major metabolite. Biochem Pharm. 17, 247254. https://doi.
org/10.1016/0006-2952(68)90330-4.
Schoentgen, B., Gagliardi, G., D´
efontaines, B., 2020. Environmental and cognitive
enrichment in childhood as protective factors in the adult and ageing brain. Front.
Psychol. 11, 1814. https://doi.org/10.3389/fpsyg.2020.01814.
Schou, M., Halldin, C., S´
ov´
ag´
o, J., Pike, V.W., Guly´
as, B., Mozley, P.D., Johnson, D.P.,
Hall, H., Innis, R.B., Farde, L., 2003. Specic in vivo binding to the norepinephrine
transporter demonstrated with the PET radioligand, (S,S)-[11C]MeNER. Nucl. Med.
Biol. 30, 707714. https://doi.org/10.1016/S0969-8051(03)00079-9.
Schwarz, L.A., Luo, L., 2015. Organization of the locus coeruleus-norepinephrine system.
Curr. Biol. 25, R1051R1056. https://doi.org/10.1016/j.cub.2015.09.039.
Schwarz, S.T., Xing, Y., Tomar, P., Bajaj, N., Auer, D.P., 2017. In vivo assessment of
brainstem depigmentation in parkinson disease: potential as a severity marker for
multicenter studies. Radiology 283, 789798. https://doi.org/10.1148/
radiol.2016160662.
Sclocco, R., Garcia, R.G., Kettner, N.W., Isenburg, K., Fisher, H.P., Hubbard, C.S., Ay, I.,
Polimeni, J.R., Goldstein, J., Makris, N., Toschi, N., Barbieri, R., Napadow, V., 2019.
The inuence of respiration on brainstem and cardiovagal response to auricular
vagus nerve stimulation: a multimodal ultrahigh-eld (7T) fMRI study. Brain Stimul.
12, 911921. https://doi.org/10.1016/j.brs.2019.02.003.
Sesack, S.R., Deutch, A.Y., Roth, R.H., Bunney, B.S., 1989. Topographical organization of
the efferent projections of the medial prefrontal cortex in the rat: an anterograde
tract-tracing study withPhaseolus vulgaris leucoagglutinin. J. Comp. Neurol. 290,
213242. https://doi.org/10.1002/cne.902900205.
Sharp, E.S., Gatz, M., 2011. Relationship between education and dementia: an updated
systematic review. Alzheimer Dis. Assoc. Disord. 25, 289304. https://doi.org/
10.1097/WAD.0b013e318211c83c.
Simen, A.A., Bordner, K.A., Martin, M.P., Moy, L.A., Barry, L.C., 2011. Cognitive
dysfunction with ageing and the role of inammation. Ther. Adv. Chronic Dis. 2,
175195. https://doi.org/10.1177/2040622311399145.
Singh, P., Bhat, R., 2019. Binding of noradrenaline to native and intermediate states
during the brillation of
α
-synuclein leads to the formation of stable and structured
cytotoxic species. ACS Chem. Neurosci. 10, 27412755. https://doi.org/10.1021/
acschemneuro.8b00650.
Slater, C., Liu, Y., Weiss, E., Yu, K., Wang, Q., 2022. The neuromodulatory role of the
noradrenergic and cholinergic systems and their interplay in cognitive functions: a
focused review. Brain Sci. 12, 890. https://doi.org/10.3390/brainsci12070890.
Smith, B.M., Yao, X., Chen, K.S., Kirby, E.D., 2018. A larger social network enhances
novel object location memory and reduces hippocampal microgliosis in aged mice.
Front. Ageing Neurosci. 10, 142. https://doi.org/10.3389/fnagi.2018.00142.
Smith, C.C., Greene, R.W., 2012. CNS dopamine transmission mediated by noradrenergic
innervation. J. Neurosci. 32, 60726080. https://doi.org/10.1523/
JNEUROSCI.6486-11.2012.
Smith, R.E., Tournier, J.-D., Calamante, F., Connelly, A., 2015. SIFT2: enabling dense
quantitative assessment of brain white matter connectivity using streamlines
tractography. NeuroImage 119, 338351. https://doi.org/10.1016/j.
neuroimage.2015.06.092.
Sommerauer, M., Fedorova, T.D., Hansen, A.K., Knudsen, K., Otto, M., Jeppesen, J.,
Frederiksen, Y., Blicher, J.U., Geday, J., Nahimi, A., Damholdt, M.F., Brooks, D.J.,
Borghammer, P., 2018. Evaluation of the noradrenergic system in Parkinsons
disease: an 11C-MeNER PET and neuromelanin MRI study. Brain 141, 496504.
https://doi.org/10.1093/brain/awx348.
Song, I., Neal, J., Lee, T.H., 2021. Age-related intrinsic functional connectivity changes of
locus coeruleus from childhood to older adults. Brain Sci. 11. https://doi.org/
10.3390/brainsci11111485.
Song, S., Jiang, L., Oyarzabal, E.A., Wilson, B., Li, Z., Shih, Y.-Y.I., Wang, Q., Hong, J.-S.,
2019a. Loss of brain norepinephrine elicits neuroinammation-mediated oxidative
injury and selective caudo-rostral neurodegeneration. Mol. Neurobiol. 56,
26532669. https://doi.org/10.1007/s12035-018-1235-1.
Song, S., Wang, Q., Jiang, L., Oyarzabal, E., Riddick, N.V., Wilson, B., Moy, S.S., Shih, Y.
Y.I., Hong, J.S., 2019b. Noradrenergic dysfunction accelerates LPS-elicited
inammation-related ascending sequential neurodegeneration and decits in non-
motor/motor functions. Brain Behav. Immun. 81, 374387. https://doi.org/
10.1016/j.bbi.2019.06.034.
Song, S., Tu, D., Meng, C., Liu, J., Wilson, B., Wang, Q., Shih, Y.-Y.I., Gao, H.-M.,
Hong, J.-S., 2023. Dysfunction of the noradrenergic system drives inammation,
α
-synucleinopathy, and neuronal loss in mouse colon. Front. Immunol. 14, 1083513.
https://doi.org/10.3389/mmu.2023.1083513.
Spengler, R.N., Allen, R.M., Remick, D.G., Strieter, R.M., Kunkel, S.L., 1990. Stimulation
of alpha-adrenergic receptor augments the production of macrophage-derived tumor
necrosis factor. J. Immunol. 145, 14301434.
Spiegel, R., DeVos, J., 1980. Central effects of guanfacine and clonidine during
wakefulness and sleep in healthy subjects. Br. J. Clin. Pharmacol. 10, 165S168S.
https://doi.org/10.1111/j.1365-2125.1980.tb04925.x.
Stebbins, R.C., Yang, Y.C., Reason, M., Aiello, A.E., Belsky, D.W., Harris, K.M.,
Plassman, B.L., 2022. Occupational cognitive stimulation, socioeconomic status, and
cognitive functioning in young adulthood. SSM - Popul. Health 17, 101024. https://
doi.org/10.1016/j.ssmph.2022.101024.
Sterpenich, V., DArgembeau, A., Desseilles, M., Balteau, E., Albouy, G., Vandewalle, G.,
Degueldre, C., Luxen, A., Collette, F., Maquet, P., 2006. The locus ceruleus is
involved in the successful retrieval of emotional memories in humans. J. Neurosci.
26, 74167423. https://doi.org/10.1523/JNEUROSCI.1001-06.2006.
Stratmann, K., Heinsen, H., Korf, H.-W., Del Turco, D., Ghebremedhin, E., Seidel, K.,
Bouzrou, M., Grinberg, L.T., Bohl, J., Wharton, S.B., den Dunnen, W., Rüb, U., 2016.
Precortical phase of Alzheimers disease (AD)-related tau cytoskeletal pathology:
precortical phases of AD. Brain Pathol. 26, 371386. https://doi.org/10.1111/
bpa.12289.
Strobel, G., Freidmann, B., Siebold, R., B¨
artsch, P., 1997. Effect of severe exercise on
plasma catecholamines. Medicine & Science in Sports & Exercise.
Sugama, S., Kakinuma, Y., 2021. Noradrenaline as a key neurotransmitter in modulating
microglial activation in stress response. Neurochem. Int. 143, 104943 https://doi.
org/10.1016/j.neuint.2020.104943.
Sulzer, D., Cassidy, C., Horga, G., Kang, U.J., Fahn, S., Casella, L., Pezzoli, G., Langley, J.,
Hu, X.P., Zucca, F.A., Isaias, I.U., Zecca, L., 2018. Neuromelanin detection by
magnetic resonance imageing (MRI) and its promise as a biomarker for Parkinsons
disease. NPJ Park. Dis. 4, 11. https://doi.org/10.1038/s41531-018-0047-3.
Sved, A.F., Felsten, G., 1987. Stimulation of the locus coeruleus decreases arterial
pressure. Brain Res. 414, 119132. https://doi.org/10.1016/0006-8993(87)91332-
1.
Swartzwelder, R.A., Galanos, A.N., 2016. Methylphenidate Use in the Elderly Population:
What Do We Know Now?
Szabadi, E., 2013. Functional neuroanatomy of the central noradrenergic system.
J. Psychopharmacol. 27, 659693. https://doi.org/10.1177/0269881113490326.
Taipa, R., Ferreira, V., Brochado, P., Robinson, A., Reis, I., Marques, F., Mann, D.M.,
Melo-Pires, M., Sousa, N., 2018. Inammatory pathology markers (activated
microglia and reactive astrocytes) in early and late onset Alzheimer disease: a post
mortem study. Neuropathol. Appl. Neurobiol. 44, 298313. https://doi.org/
10.1111/nan.12445.
Takeuchi, T., Duszkiewicz, A.J., Sonneborn, A., Spooner, P.A., Yamasaki, M.,
Watanabe, M., Smith, C.C., Fern´
andez, G., Deisseroth, K., Greene, R.W., Morris, R.G.
M., 2016. Locus coeruleus and dopaminergic consolidation of everyday memory.
Nature 537, 357362. https://doi.org/10.1038/nature19325.
Tanaka, K.F., Kashima, H., Suzuki, H., Ono, K., Sawada, M., 2002. Existence of functional
β1and β2adrenergic receptors on microglia. J. Neurosci. Res. 70, 232237.
Tavares, A., Handy, D.E., Bogdanova, N.N., Rosene, D.L., Gavras, H., 1996. Localization
of
α
2A- and
α
2B-Adrenergic Receptor Subtypes in Brain. Hypertension 27, 449455.
https://doi.org/10.1161/01.HYP.27.3.449.
Taylor, K.I., Sambataro, F., Boess, F., Bertolino, A., Dukart, J., 2018. Progressive decline
in gray and white matter integrity in de novo Parkinsons disease: an analysis of
longitudinal parkinson progression markers initiative diffusion tensor imageing data.
Front. Ageing Neurosci. 10, 318. https://doi.org/10.3389/fnagi.2018.00318.
Taylor, N.L., DSouza, A., Munn, B.R., Lv, J., Zaborszky, L., Müller, E.J., Wainstein, G.,
Calamante, F., Shine, J.M., 2022. Structural connections between the noradrenergic
and cholinergic system shape the dynamics of functional brain networks.
NeuroImage 260, 119455. https://doi.org/10.1016/j.neuroimage.2022.119455.
Tekieh, T., Robinson, P.A., Postnova, S., 2022. Cortical waste clearance in normal and
restricted sleep with potential runaway tau buildup in Alzheimers disease. Sci. Rep.
12, 13740. https://doi.org/10.1038/s41598-022-15109-6.
Teunissen, C.E., Verberk, I.M.W., Thijssen, E.H., Vermunt, L., Hansson, O.,
Zetterberg, H., van der Flier, W.M., Mielke, M.M., del Campo, M., 2022. Blood-based
biomarkers for Alzheimers disease: towards clinical implementation. Lancet Neurol.
21, 6677. https://doi.org/10.1016/S1474-4422(21)00361-6.
Theolas, P., Ehrenberg, A.J., Dunlop, S., Di Lorenzo Alho, A.T., Nguy, A., Leite, R.E.P.,
Rodriguez, R.D., Mejia, M.B., Suemoto, C.K., Ferretti-Rebustini, R.E.D.L.,
Polichiso, L., Nascimento, C.F., Seeley, W.W., Nitrini, R., Pasqualucci, C.A., Jacob
Filho, W., Rueb, U., Neuhaus, J., Heinsen, H., Grinberg, L.T., 2017. Locus coeruleus
volume and cell population changes during Alzheimers disease progression: A
stereological study in human postmortem brains with potential implication for early-
stage biomarker discovery. Alzheimers Dement. 13, 236246. https://doi.org/
10.1016/j.jalz.2016.06.2362.
Therriault, J., Pascoal, T.A., Lussier, F.Z., Tissot, C., Chamoun, M., Bezgin, G., Servaes, S.,
Benedet, A.L., Ashton, N.J., Karikari, T.K., Lantero-Rodriguez, J., Kunach, P.,
Wang, Y.-T., Fernandez-Arias, J., Massarweh, G., Vitali, P., Soucy, J.-P., Saha-
Chaudhuri, P., Blennow, K., Zetterberg, H., Gauthier, S., Rosa-Neto, P., 2022.
Biomarker modeling of Alzheimers disease using PET-based Braak stageing. Nat.
Ageing 2, 526535. https://doi.org/10.1038/s43587-022-00204-0.
F. Krohn et al.
Neuroscience and Biobehavioral Reviews 152 (2023) 105311
19
Thomsen, M.S., Routhe, L.J., Moos, T., 2017. The vascular basement membrane in the
healthy and pathological brain. J. Cereb. Blood Flow Metab. 37, 33003317. https://
doi.org/10.1177/0271678X17722436.
Titulaer, J., Bj¨
orkholm, C., Feltmann, K., Malml¨
of, T., Mishra, D., Bengtsson Gonzales, C.,
Schilstr¨
om, B., Konradsson-Geuken, Å., 2021. The Importance Of Ventral
Hippocampal Dopamine And Norepinephrine In Recognition Memory. Front. Behav.
Neurosci. 15, 16. https://doi.org/10.3389/fnbeh.2021.667244.
Toda, M., Abi-Dargham, A., 2007. Dopamine hypothesis of schizophrenia: making sense
of it all. Curr. Psychiatry Rep. 9, 329336. https://doi.org/10.1007/s11920-007-
0041-7.
Tomassini, A., Hezemans, F.H., Ye, R., Tsvetanov, K.A., Wolpe, N., Rowe, J.B., 2022.
Prefrontal Cortical Connectivity Mediates Locus Coeruleus Noradrenergic Regulation
Of Inhibitory Control In Older Adults. J. Neurosci. 42, 34843493. https://doi.org/
10.1523/JNEUROSCI.1361-21.2022.
Tomlinson, B.E., Irving, D., Blessed, G., 1981. Cell loss in the locus coeruleus in senile
dementia of Alzheimer type. J. Neurol. Sci. 49, 419428. https://doi.org/10.1016/
0022-510X(81)90031-9.
Turker, H.B., Riley, E., Luh, W.M., Colcombe, S.J., Swallow, K.M., 2021. Estimates of
locus coeruleus function with functional magnetic resonance imageing are
inuenced by localization approaches and the use of multi-echo data. NeuroImage
236, 118047. https://doi.org/10.1016/j.neuroimage.2021.118047.
Turner, P.R., OConnor, K., Tate, W.P., Abraham, W.C., 2003. Roles of amyloid precursor
protein and its fragments in regulating neural activity, plasticity and memory. Prog.
Neurobiol. 70, 132. https://doi.org/10.1016/S0301-0082(03)00089-3.
Uematsu, A., Tan, B.Z., Ycu, E.A., Cuevas, J.S., Koivumaa, J., Junyent, F., Kremer, E.J.,
Witten, I.B., Deisseroth, K., Johansen, J.P., 2017. Modular organization of the
brainstem noradrenaline system coordinates opposing learning states. Nat. Neurosci.
20, 16021611. https://doi.org/10.1038/nn.4642.
Unsworth, N., Robison, M.K., 2017. A locus coeruleus-norepinephrine account of
individual differences in working memory capacity and attention control. Psychon.
Bull. Rev. 24, 12821311. https://doi.org/10.3758/s13423-016-1220-5.
Van Egroo, M., Narbutas, J., Chylinski, D., Villar Gonz´
alez, P., Maquet, P., Salmon, E.,
Bastin, C., Collette, F., Vandewalle, G., 2019. Sleepwake regulation and the
hallmarks of the pathogenesis of Alzheimers disease. Sleep 42, zsz017. https://doi.
org/10.1093/sleep/zsz017.
Van Egroo, M., Koshmanova, E., Vandewalle, G., Jacobs, H.I.L., 2022. Importance of the
locus coeruleus-norepinephrine system in sleep-wake regulation: Implications for
ageing and Alzheimers disease, 101592101592 Sleep. Med. Rev. 62. https://doi.
org/10.1016/j.smrv.2022.101592.
van Hooren, R.W.E., Verhey, F.R.J., Ramakers, I.H.G.B., Jansen, W.J., Jacobs, H.I.L.,
2021. Elevated norepinephrine metabolism is linked to cortical thickness in the
context of Alzheimers disease pathology. Neurobiol. Ageing 102, 1722. https://
doi.org/10.1016/j.neurobiolageing.2021.01.024.
Vankov, A., Herv´
e-Minvielle, A., Sara, S.J., 1995. Response to novelty and its rapid
habituation in locus coeruleus neurons of the freely exploring rat. Eur. J. Neurosci. 7,
11801187. https://doi.org/10.1111/j.1460-9568.1995.tb01108.x.
Vargas-Caballero, M., Warming, H., Walker, R., Holmes, C., Cruickshank, G., Patel, B.,
2022. Vagus nerve stimulation as a potential therapy in early Alzheimers disease: a
review. Front. Hum. Neurosci. 16, 866434 https://doi.org/10.3389/
fnhum.2022.866434.
Veyrac, A., Sacquet, J., Nguyen, V., Marien, M., Jourdan, F., Didier, A., 2009. Novelty
determines the effects of olfactory enrichment on memory and neurogenesis through
noradrenergic mechanisms. Neuropsychopharmacol 34, 786795. https://doi.org/
10.1038/npp.2008.191.
Vicq-dAzyr, F., 1786. Traite dAnatomie de Cerveau., Trait´
e danatomie et de
physiologie.
Viglione, A., Mazziotti, R., Pizzorusso, T., 2023. From pupil to the brain: new insights for
studying cortical plasticity through pupillometry. Front Neural Circuits 17, 1151847.
https://doi.org/10.3389/fncir.2023.1151847.
Wainstein, G., Müller, E.J., Taylor, N., Munn, B., Shine, J.M., 2022. The role of the locus
coeruleus in shaping adaptive cortical melodies. Trends Cogn. Sci. 26, 527538.
https://doi.org/10.1016/j.tics.2022.03.006.
Wang, J., Li, Y., Huang, Z., Wan, W., Zhang, Y., Wang, C., Cheng, X., Ye, F., Liu, K.,
Fei, G., Zeng, M., Jin, L., 2018. Neuromelanin-sensitive magnetic resonance
imageing features of the substantia nigra and locus coeruleus in de novo Parkinsons
disease and its phenotypes. Eur. J. Neurol. 25, 949e73. https://doi.org/10.1111/
ene.13628.
Wang, Q., 2020. Amyloid β redirects norepinephrine signaling to activate the pathogenic
GSK3β/tau cascade: Interventions targeting the noradrenergic system in Alzheimers
and neurodegenerative disease. AlzheimersDement. 16. https://doi.org/10.1002/
alz.044769.
Wang, T., Zhang, Q.-J., Liu, J., Wu, Z.-H., Wang, S., 2009. Firing activity of locus
coeruleus noradrenergic neurons increases in a rodent model of Parkinsonism.
Neurosci. Bull. 25, 1520. https://doi.org/10.1007/s12264-009-1023-z.
Wei, X., Luo, C., Li, Q., Hu, N., Xiao, Y., Liu, N., Lui, S., Gong, Q., 2021. White matter
abnormalities in patients with Parkinsons disease: a meta-analysis of diffusion
tensor imageing using tract-based spatial statistics. Front. Ageing Neurosci. 12,
610962 https://doi.org/10.3389/fnagi.2020.610962.
Weinshenker, D., 2018. Long road to ruin: noradrenergic dysfunction in
neurodegenerative disease. Trends Neurosci. 41, 211223. https://doi.org/10.1016/
j.tins.2018.01.010.
Werry, E.L., Bright, F.M., Piguet, O., Ittner, L.M., Halliday, G.M., Hodges, J.R.,
Kiernan, M.C., Loy, C.T., Kril, J.J., Kassiou, M., 2019. Recent developments in TSPO
PET imageing as a biomarker of neuroinammation in neurodegenerative disorders.
Int. J. Mol. Sci. 20, 3161. https://doi.org/10.3390/ijms20133161.
Williams, J., Reiner, P., 1993. Noradrenaline hyperpolarizes identied rat mesopontine
cholinergic neurons in vitro. J. Neurosci. 13, 38783883. https://doi.org/10.1523/
JNEUROSCI.13-09-03878.1993.
Wilson, R.S., Nag, S., Boyle, P.A., Hizel, L.P., Yu, L., Buchman, A.S., Schneider, J.A.,
Bennett, D.A., 2013. Neural reserve, neuronal density in the locus ceruleus, and
cognitive decline. Neurology 80, 12021208. https://doi.org/10.1212/
WNL.0b013e3182897103.
Xie, L., Kang, H., Xu, Q., Chen, M.J., Liao, Y., Thiyagarajan, M., ODonnell, J.,
Christensen, D.J., Nicholson, C., Iliff, J.J., Takano, T., Deane, R., Nedergaard, M.,
2013. Sleep drives metabolite clearance from the adult brain. Science 342, 373377.
https://doi.org/10.1126/science.1241224.
Yakunina, N., Kim, S.S., Nam, E.-C., 2017. Optimization of transcutaneous vagus nerve
stimulation using functional MRI. Neuromodulation: Technol. Neural Interface 20,
290300. https://doi.org/10.1111/ner.12541.
Yamasaki, M., Takeuchi, T., 2017. Locus coeruleus and dopamine-dependent memory
consolidation. Neural Plast. 2017, 115. https://doi.org/10.1155/2017/8602690.
Ye, R., Hezemans, F.H., OCallaghan, C., Tsvetanov, K.A., Rua, C., Jones, P.S.,
Holland, N., Malpetti, M., Murley, A.G., Barker, R.A., Williams-Gray, C.H.,
Robbins, T.W., Passamonti, L., Rowe, J.B., 2021. Reduced locus coeruleus integrity
linked to response inhibition decits in parkinsonian disorders. medRxiv 44,
2021.10.14.21264996.
Ye, R., OCallaghan, C., Rua, C., Hezemans, F.H., Holland, N., Malpetti, M., Jones, P.S.,
Barker, R.A., Williams-Gray, C.H., Robbins, T.W., Passamonti, L., Rowe, J., 2022.
Locus coeruleus integrity from 7 T MRI relates to apathy and cognition in
Parkinsonian disorders. Mov. Disord. 37, 16631672. https://doi.org/10.1002/
mds.29072.
Yi, Y.-J., Lüsebrink, F., Ludwig, M., Maaß, A., Ziegler, G., Yakupov, R., Kreißl, M.C.,
Betts, M., Speck, O., Düzel, E., H¨
ammerer, D., 2023. It is the locus coeruleus! Oris
it?: a proposition for analyses and reporting standards for structural and functional
magnetic resonance imageing of the noradrenergic locus coeruleus. Neurobiol.
Ageing 129, 137148. https://doi.org/10.1016/j.neurobiolageing.2023.04.007.
Yuan, Z.Y., Wang, M.W., Yan, B.Y., Gu, P., Jiang, X.M., Yang, X.F., Cui, D.S., 2012. An
enriched environment improves cognitive performance in mice from the senescence-
accelerated prone mouse 8 strain: Role of upregulated neurotrophic factor
expression in the hippocampus. Neural Regen. Res. 7, 17971804. https://doi.org/
10.3969/j.issn.1673-5374.
Zaehle, T., Krauel, K., 2021. Chapter 7: Transcutaneous vagus nerve stimulation in
patients with attention-decit/hyperactivity disorder: A viable option? In:
Kadosh, R.C., Zaehle, T., Krauel, K. (Eds.), Progress in Brain Research, Non-Invasive
Brain Stimulation (NIBS) in Neurodevelopmental Disorders. Elsevier, pp. 171190.
https://doi.org/10.1016/bs.pbr.2021.03.001.
Zarow, C., Lyness, S.A., Mortimer, J.A., Chui, H.C., 2003. Neuronal Loss Is Greater In The
Locus Coeruleus Than Nucleus Basalis And Substantia Nigra in Alzheimer and
Parkinson Diseases. Arch. Neurol. 60, 337. https://doi.org/10.1001/
archneur.60.3.337.
Zecca, L., Stroppolo, A., Gatti, A., Tampellini, D., Toscani, M., Gallorini, M., Giaveri, G.,
Arosio, P., Santambrogio, P., Fariello, R.G., Karatekin, E., Kleinman, M.H., Turro, N.,
Hornykiewicz, O., Zucca, F.A., 2004. The role of iron and copper molecules in the
neuronal vulnerability of locus coeruleus and substantia nigra during ageing. Proc.
Natl. Acad. Sci. USA 101, 98439848. https://doi.org/10.1073/pnas.0403495101.
Zhang, Y., Burock, M.A., 2020. Diffusion tensor imageing in Parkinsons disease and
parkinsonian syndrome: a systematic review. Front. Neurol. 11, 531993 https://doi.
org/10.3389/fneur.2020.531993.
Zhang, Y., Vakhtin, A.A., Jennings, J.S., Massaband, P., Wintermark, M., Craig, P.L.,
Ashford, J.W., Clark, J.D., Furst, A.J., 2020. Diffusion tensor tractography of
brainstem bers and its application in pain. PLoS ONE 15, e0213952. https://doi.
org/10.1371/journal.pone.0213952.
Zornetzer, S.F., Gold, M.S., 1976. The locus coeruleus: its possible role in memory
consolidation. Physiol. Behav. 16, 331336. https://doi.org/10.1016/0031-9384
(76)90140-2.
Zou, Y., Qian, Z., Chen, Y., Qian, H., Wei, G., Zhang, Q., 2019. Norepinephrine inhibits
Alzheimers amyloid-β peptide aggregation and destabilizes amyloid-β protobrils: a
molecular dynamics simulation study. ACS Chem. Neurosci. 10, 15851594. https://
doi.org/10.1021/acschemneuro.8b00537.
Zucca, F.A., Bellei, C., Giannelli, S., Terreni, M.R., Gallorini, M., Rizzio, E., Pezzoli, G.,
Albertini, A., Zecca, L., 2006. Neuromelanin and iron in human locus coeruleus and
substantia nigra during ageing: consequences for neuronal vulnerability. J. Neural
Transm. 113, 757767. https://doi.org/10.1007/s00702-006-0453-2.
F. Krohn et al.
... On the other hand, the substantial reduction of NE-positive neurons within the LC of SMNΔ7 mice is in line with the high vulnerability of this brain nucleus to neuronal degeneration under a neuroinflammatory environment typically occurring in SMA pathology [25][26][27]61 . Accordingly, loss of noradrenergic LC neurons is described in animal models and patients with other neurodegenerative disorders, including multiple sclerosis, Alzheimer's disease, and Parkinson's disease [62][63][64] . ...
... It is well-established that NE is a key regulator of neuroimmune responses, exerting both anti-inflammatory and neuroprotective effects through the activation of noradrenergic receptors located on peripheral immune cells, microglia, and astrocytes 62 . Accordingly, perturbation of NE signaling contributes to the pathophysiology of different inflammatory neurodegenerative disorders [62][63][64] . Based on these findings, we cannot exclude that the anti-inflammatory response elicited by Nusinersen treatment previously reported in the CSF of SMA1 patients 26 could rely on the upregulation of central NE levels identified in this study on the same patients' cohort. ...
Article
Full-text available
Beyond motor neuron degeneration, homozygous mutations in the survival motor neuron 1 (SMN1) gene cause multiorgan and metabolic defects in patients with spinal muscular atrophy (SMA). However, the precise biochemical features of these alterations and the age of onset in the brain and peripheral organs remain unclear. Using untargeted NMR-based metabolomics in SMA mice, we identify cerebral and hepatic abnormalities related to energy homeostasis pathways and amino acid metabolism, emerging already at postnatal day 3 (P3) in the liver. Through HPLC, we find that SMN deficiency induces a drop in cerebral nor-epinephrine levels in overt symptomatic SMA mice at P11, affecting the mRNA and protein expression of key genes regulating monoamine metabolism, including aromatic L-amino acid decarboxylase (AADC), dopamine beta-hydroxylase (DβH) and monoamine oxidase A (MAO-A). In support of the translational value of our preclinical observations, we also discovered that SMN upregulation increases cerebrospinal fluid norepinephrine concentration in Nusinersen-treated SMA1 patients. Our findings highlight a previously unrecognized harmful influence of low SMN levels on the expression of critical enzymes involved in monoamine metabolism, suggesting that SMN-inducing therapies may modulate catecholamine neurotransmission. These results may also be relevant for setting therapeutic approaches to counteract peripheral metabolic defects in SMA.
... In other species, such as cats, rabbits or dogs, area A5 is less clearly defined [15,16]. In humans, its location is very similar to that of rats, and its alterations seem to be involved in some of the cardiorespiratory manifestations observed in some pathologies, such as Rett syndrome and Ondine disease syndrome, which are currently among the aetiological factors responsible for sudden infant death syndrome [17][18][19] or being involved in the neuroinflammation and neuromodulation that mediates various neurodegenerative diseases [20][21][22][23]. ...
Article
Full-text available
Area A5 is a noradrenergic cell group in the brain stem characterised by its important role in triggering sympathetic activity, exerting a profound influence on the sympathetic outflow, which is instrumental in the modulation of cardiovascular functions, stress responses and various other physiological processes that are crucial for adaptation and survival mechanisms. Understanding the role of area A5, therefore, not only provides insights into the basic functioning of the sympathetic nervous system but also sheds light on the neuronal basis of a number of autonomic responses. In this review, we look deeper into the specifics of area A5, exploring its anatomical connections, its neurochemical properties and the mechanisms by which it influences sympathetic nervous system activity and cardiorespiratory regulation and, thus, contributes to the overall dynamics of the autonomic function in regulating body homeostasis.
Article
Full-text available
Changes in dopaminergic neuromodulation play a key role in adult memory decline. Recent research has also implicated noradrenaline in shaping late-life memory. However, it is unclear whether these two neuromodulators have distinct roles in age-related cognitive changes. Here, combining longitudinal MRI of the dopaminergic substantia nigra–ventral tegmental area (SN-VTA) and noradrenergic locus coeruleus (LC) in younger (n = 69) and older (n = 251) adults, we found that dopaminergic and noradrenergic integrity are differentially associated with memory performance. While LC integrity was related to better episodic memory across several tasks, SN-VTA integrity was linked to working memory. Longitudinally, we found that older age was associated with more negative change in SN-VTA and LC integrity. Notably, changes in LC integrity reliably predicted future episodic memory. These differential associations of dopaminergic and noradrenergic nuclei with late-life cognitive decline have potential clinical utility, given their degeneration in several age-associated diseases.
Article
Full-text available
The noradrenergic system shows pathological modifications in aging and neurodegenerative diseases and undergoes substantial neuronal loss in Alzheimer’s disease and Parkinson’s disease. While a coherent picture of structural decline in post-mortem and in vivo MRI measures seems to emerge, whether this translates into a consistent decline in available noradrenaline levels is unclear. We conducted a meta-analysis of noradrenergic differences in Alzheimer’s disease dementia and Parkinson’s disease using CSF and PET biomarkers. CSF noradrenaline and 3-methoxy-4-hydroxyphenylglycol levels as well as noradrenaline transporters availability, measured with PET, were summarized from 26 articles using a random-effects model meta-analysis. Compared to controls, individuals with Parkinson’s disease showed significantly decreased levels of CSF noradrenaline and 3-methoxy-4-hydroxyphenylglycol, as well as noradrenaline transporters availability in the hypothalamus. In Alzheimer’s disease dementia, 3-methoxy-4-hydroxyphenylglycol but not noradrenaline levels were increased compared to controls. Both CSF and PET biomarkers of noradrenergic dysfunction reveal significant alterations in Parkinson’s disease and Alzheimer’s disease dementia. However, further studies are required to understand how these biomarkers are associated to the clinical symptoms and pathology.
Article
Full-text available
Pupil size variations have been associated with changes in brain activity patterns related with specific cognitive factors, such as arousal, attention, and mental effort. The locus coeruleus (LC), a key hub in the noradrenergic system of the brain, is considered to be a key regulator of cognitive control on pupil size, with changes in pupil diameter corresponding to the release of norepinephrine (NE). Advances in eye-tracking technology and open-source software have facilitated accurate pupil size measurement in various experimental settings, leading to increased interest in using pupillometry to track the nervous system activation state and as a potential biomarker for brain disorders. This review explores pupillometry as a non-invasive and fully translational tool for studying cortical plasticity starting from recent literature suggesting that pupillometry could be a promising technique for estimating the degree of residual plasticity in human subjects. Given that NE is known to be a critical mediator of cortical plasticity and arousal, the review includes data revealing the importance of the LC-NE system in modulating brain plasticity and pupil size. Finally, we will review data suggesting that pupillometry could provide a quantitative and complementary measure of cortical plasticity also in pre-clinical studies.
Article
Full-text available
Clinical and pathological evidence revealed that α-synuclein (α-syn) pathology seen in PD patients starts in the gut and spreads via anatomically connected structures from the gut to the brain. Our previous study demonstrated that depletion of central norepinephrine (NE) disrupted brain immune homeostasis, producing a spatiotemporal order of neurodegeneration in the mouse brain. The purpose of this study was 1) to determine the role of peripheral noradrenergic system in the maintenance of gut immune homeostasis and in the pathogenesis of PD and 2) to investigate whether NE-depletion induced PD-like α-syn pathological changes starts from the gut. For these purposes, we investigated time-dependent changes of α-synucleinopathy and neuronal loss in the gut following a single injection of DSP-4 (a selective noradrenergic neurotoxin) to A53T-SNCA (human mutant α-syn) over-expression mice. We found DPS-4 significantly reduced the tissue level of NE and increased immune activities in gut, characterized by increased number of phagocytes and proinflammatory gene expression. Furthermore, a rapid-onset of α-syn pathology was observed in enteric neurons after 2 weeks and delayed dopaminergic neurodegeneration in the substantia nigra was detected after 3-5 months, associated with the appearance of constipation and impaired motor function, respectively. The increased α-syn pathology was only observed in large, but not in the small, intestine, which is similar to what was observed in PD patients. Mechanistic studies reveal that DSP-4-elicited upregulation of NADPH oxidase (NOX2) initially occurred only in immune cells during the acute intestinal inflammation stage, and then spread to enteric neurons and mucosal epithelial cells during the chronic inflammation stage. The upregulation of neuronal NOX2 correlated well with the extent of α-syn aggregation and subsequent enteric neuronal loss, suggesting that NOX2-generated reactive oxygen species play a key role in α-synucleinopathy. Moreover, inhibiting NOX2 by diphenyleneiodonium or restoring NE function by salmeterol (a β2-receptor agonist) significantly attenuated colon inflammation, α-syn aggregation/propagation, and enteric neurodegeneration in the colon and ameliorated subsequent behavioral deficits. Taken together, our model of PD shows a progressive pattern of pathological changes from the gut to the brain and suggests a potential role of the noradrenergic dysfunction in the pathogenesis of PD.
Article
Full-text available
Parkinson’s disease (PD) is diagnosed many years after its onset, under a significant degradation of the nigrostriatal dopaminergic system, responsible for the regulation of motor function. This explains the low effectiveness of the treatment of patients. Therefore, one of the highest priorities in neurology is the development of the early (preclinical) diagnosis of PD. The aim of this study was to search for changes in the blood of patients at risk of developing PD, which are considered potential diagnostic biomarkers. Out of 1835 patients, 26 patients were included in the risk group and 20 patients in the control group. The primary criteria for inclusion in a risk group were the impairment of sleep behavior disorder and sense of smell, and the secondary criteria were neurological and mental disorders. In patients at risk and in controls, the composition of plasma and the expression of genes of interest in lymphocytes were assessed by 27 indicators. The main changes that we found in plasma include a decrease in the concentrations of l-3,4-dihydroxyphenylalanine (L-DOPA) and urates, as well as the expressions of some types of microRNA, and an increase in the total oxidative status. In turn, in the lymphocytes of patients at risk, an increase in the expression of the DA D3 receptor gene and the lymphocyte activation gene 3 (LAG3), as well as a decrease in the expression of the Protein deglycase DJ-1 gene (PARK7), were observed. The blood changes we found in patients at risk are considered candidates for diagnostic biomarkers at the prodromal stage of PD.
Article
Full-text available
Background: Advances in ultrasensitive detection of phosphorylated tau (p-tau) in plasma has enabled the use of blood tests to measure Alzheimer's disease (AD) biomarker changes. Examination of postmortem brains of participants with antemortem plasma p-tau levels remains critical to understanding comorbid and AD-specific contribution to these biomarker changes. Methods: We analyzed 35 population-based Mayo Clinic Study of Aging participants with plasma p-tau at threonine 181 and threonine 217 (p-tau181, p-tau217) available within 3 years of death. Autopsied participants included cognitively unimpaired, mild cognitive impairment, AD dementia, and non-AD neurodegenerative disorders. Global neuropathologic scales of tau, amyloid-β, TDP-43, and cerebrovascular disease were examined. Regional digital pathology measures of tau (phosphorylated threonine 181 and 217 [pT181, pT217]) and amyloid-β (6F/3D) were quantified in hippocampus and parietal cortex. Neurotransmitter hubs reported to influence development of tangles (nucleus basalis of Meynert) and amyloid-β plaques (locus coeruleus) were evaluated. Results: The strongest regional associations were with parietal cortex for tau burden (p-tau181 R = 0.55, p = 0.003; p-tau217 R = 0.66, p < 0.001) and amyloid-β burden (p-tau181 R = 0.59, p < 0.001; p-tau217 R = 0.71, p < 0.001). Linear regression analysis of global neuropathologic scales explained 31% of variability in plasma p-tau181 (Adj. R2 = 0.31) and 59% in plasma p-tau217 (Adj. R2 = 0.59). Neither TDP-43 nor cerebrovascular disease global scales independently contributed to variability. Global scales of tau pathology (β-coefficient = 0.060, p = 0.016) and amyloid-β pathology (β-coefficient = 0.080, p < 0.001) independently predicted plasma p-tau217 when modeled together with co-pathologies, but only amyloid-β (β-coefficient = 0.33, p = 0.021) significantly predicted plasma p-tau181. While nucleus basalis of Meynert neuron count/mm2 was not associated with plasma p-tau levels, a lower locus coeruleus neuron count/mm2 was associated with higher plasma p-tau181 (R = -0.50, p = 0.007) and higher plasma p-tau217 (R = -0.55, p = 0.002). Cognitive scores (Adj. R2 = 0.25-0.32) were predicted by the global tau scale, but not by the global amyloid-β scale or plasma p-tau when modeled simultaneously. Conclusions: Higher soluble plasma p-tau levels may be the result of an intersection between insoluble deposits of amyloid-β and tau accumulation in brain, and may be associated with locus coeruleus degeneration.
Article
Full-text available
Locus coeruleus (LC) noradrenergic (NE) neurons supply the main adrenergic input to the forebrain. NE is a dual modulator of cognition and neuroinflammation. NE neurons of the LC are particularly vulnerable to degeneration both with normal aging and in neurodegenerative disorders. Consequences of this vulnerability can be observed in both cognitive impairment and dysregulation of neuroinflammation. LC NE neurons are pacemaker neurons that are active during waking and arousal and are responsive to stressors in the environment. Chronic overactivation is thought to be a major contributor to the vulnerability of these neurons. Here we review what is known about the mechanisms underlying this neuronal vulnerability and combinations of environmental and genetic factors that contribute to confer risk to these important brainstem neuromodulatory and immunomodulatory neurons. Finally, we discuss proposed and potential interventions that may reduce the overall risk for LC NE neuronal degeneration.
Article
The noradrenergic locus coeruleus (LC) is one of the protein pathology epicenters in neurodegenerative diseases. In contrast to PET (positron emission tomography), MRI (magnetic resonance imaging) offers the spatial resolution necessary to investigate the 3-4 mm wide and 1.5 cm long LC. However, standard data postprocessing is often too spatially imprecise to allow investigating the structure and function of the LC at the group level. Our analysis pipeline uses a combination of existing toolboxes (SPM12, ANTs, FSL, FreeSurfer), and is tailored towards achieving suitable spatial precision in the brainstem area. Its effectiveness is demonstrated using 2 datasets comprising both younger and older adults. We also suggest quality assessment procedures which allow to quantify the spatial precision obtained. Spatial deviations below 2.5 mm in the LC area are achieved, which is superior to current standard approaches. Relevant for ageing and clinical researchers interested in brainstem imaging, we provide a tool for more reliable analyses of structural and functional LC imaging data which can be also adapted for investigating other nuclei of the brainstem.
Article
The locus coeruleus (LC) is a small noradrenergic brainstem nucleus that plays a central role in regulating arousal, attention, and performance. In the mammalian brain, individual LC neurons make divergent axonal projections to different brain regions, which are distinguished in part by which noradrenaline (NA) receptor subtypes they express. Here, we sought to determine whether similar organizational features characterize LC projections to corticobasal ganglia (CBG) circuitry in the zebra finch song system, with a focus on the basal ganglia nucleus Area X, the thalamic nucleus DLM, as well as the cortical nuclei HVC, LMAN, and RA. Single and dual retrograde tracer injections reveal that single LC-NA neurons make divergent projections to LMAN and Area X, as well as to the dopaminergic VTA/SNc complex that innervates this CBG circuit. Moreover, in situ hybridization revealed that differential expression of mRNA encoding α2A and α2C adrenoreceptors distinguishes LC-recipient CBG song nuclei. Therefore, LC-NA signaling in the zebra finch CBG circuit employs a similar strategy as in mammals, which could allow a relatively small number of LC neurons to exert widespread yet distinct effects across multiple brain regions.