Locus coeruleus là gì

November 17, 2009; 73 [20] Clinical Implications of Neuroscience Research

Functional organization and potential clinical significance

First published November 16, 2009, DOI: //doi.org/10.1212/WNL.0b013e3181c2937c

This article requires a subscription to view the full text. If you have a subscription you may use the login form below to view the article. Access to this article can also be purchased.

The locus ceruleus [LC] contains norepinephrine [NE]-synthesizing neurons that send diffuse projections throughout the CNS. The LC-NE system has a major role in arousal, attention, and stress response. In the brain, NE may also contribute to long-term synaptic plasticity, pain modulation, motor control, energy homeostasis, and control of local blood blow. The LC is severely affected in neurodegenerative disorders such as Alzheimer disease [AD] and Parkinson disease [PD]. Dysregulation of LC-NE system has been implicated in sleep and arousal disorders, attention deficit hyperactivity disorder, and posttraumatic stress disorder and constitutes a target for pharmacologic treatment of these conditions. The neurobiology of the LC–noradrenergic system has been the subject of several excellent reviews.1–9

FUNCTIONAL ANATOMY OF THE LOCUS CERULEUS

The LC is a cluster of NE-containing neurons located in the upper dorsolateral pontine tegmentum [figure 1]. These neurons have extensively branched axons that project throughout the neuraxis and provide the sole source of NE to the neocortex, hippocampus, cerebellum, and most of the thalamus.1,2 Despite its widespread distribution, noradrenergic innervation shows regional specificity. For example, brain areas involved in spatial attention [such as the prefrontal and parietal cortices] receive particularly dense LC-NE inputs. In general, individual LC neurons send axon collaterals to multiple targets that process the same sensory information. Norepinephrine is released both at typical synapses and at nonsynaptic release sites; extrasynaptic NE mediates paracrine effects on neurons, glial cells, and microvessels.1–4,8

Figure 1 Anatomic organization of the locus ceruleus–norepinephrine system

The norepinephrine [NE] neurons of the locus ceruleus [LC] are located in the upper dorsolateral pontine tegmentum and can be identified by their immunoreactivity for tyrosine hydroxylase, the rate-limiting enzyme for catecholamine synthesis. These neurons have extensively branched axons that project throughout the neuraxis and provide the sole source of NE to the neocortex, hippocampus, cerebellum, and most of the thalamus. Modified, …

View Full Text

AAN Members

We have changed the login procedure to improve access between AAN.com and the Neurology journals. If you are experiencing issues, please log out of AAN.com and clear history and cookies. [For instructions by browser, please click the instruction pages below]. After clearing, choose preferred Journal and select login for AAN Members. You will be redirected to a login page where you can log in with your AAN ID number and password. When you are returned to the Journal, your name should appear at the top right of the page.

Google Safari Microsoft Edge Firefox

Click here to login

AAN Non-Member Subscribers

Purchase access

You May Also be Interested in

Abstract

Increased tonic activity of locus coeruleus noradrenergic [LC-NE] neurons induces anxiety-like and aversive behavior. While some information is known about the afferent circuitry that endogenously drives this neural activity and behavior, the downstream receptors and anatomical projections that mediate these acute risk aversive behavioral states via the LC-NE system remain unresolved. Here we use a combination of retrograde tracing, fast-scan cyclic voltammetry, electrophysiology, and in vivo optogenetics with localized pharmacology to identify neural substrates downstream of increased tonic LC-NE activity in mice. We demonstrate that photostimulation of LC-NE fibers in the BLA evokes norepinephrine release in the basolateral amygdala [BLA], alters BLA neuronal activity, conditions aversion, and increases anxiety-like behavior. Additionally, we report that β-adrenergic receptors mediate the anxiety-like phenotype of increased NE release in the BLA. These studies begin to illustrate how the complex efferent system of the LC-NE system selectively mediates behavior through distinct receptor and projection-selective mechanisms.

//doi.org/10.7554/eLife.18247.001

Introduction

The locus coeruleus noradrenergic system [LC-NE] comprises a widespread projection network throughout the central nervous system capable of modulating a diverse range of behaviors including arousal, learning, pain modulation, and stress-induced negative affective states [Berridge and Waterhouse, 2003; Sara, 2009]. Understanding the neural circuit basis for how this nearly ubiquitous neuromodulatory network exerts influence on negative affect is a critical step towards therapeutically targeting stress-induced neuropsychiatric disorders [Schwarz and Luo, 2015; Schwarz et al., 2015; Reyes et al., 2015; McCall et al., 2015; Arnsten et al., 2015; Kebschull et al., 2016]. One particular efferent projection from the LC is to the basolateral amygdala [BLA]. The BLA is an important candidate anatomical substrate for the widely known role of norepinephrine [NE] in affective behaviors [Berridge and Waterhouse, 2003; Schwarz et al., 2015; Davis, 1992; Valentino and Aston-Jones, 2010; Robertson et al., 2016, 2013; Grissom and Bhatnagar, 2011; Siuda et al., 2015a, 2016; Plummer et al., 2015]. The BLA is notable for integrating sensory information to encode and drive diverse and opposing affective behaviors including anxiety, fear, aversive, and reward behaviors [Kim et al., 2013; Stuber et al., 2011; Tye et al., 2011; Namburi et al., 2015; Gore et al., 2015; Belova et al., 2007; Beyeler et al., 2016; Wolff et al., 2014; Bermudez and Schultz, 2010; Bruchas et al., 2009; Knoll et al., 2011; Sugase-Miyamoto and Richmond, 2005; Crowley et al., 2016; Sears et al., 2013; Roozendaal et al., 2008, 2006; Miranda et al., 2007]. Notable efforts to uncover the role of the BLA and adrenergic signaling in consolidation of fear memories have been reported [Sears et al., 2013; Roozendaal et al., 2008, 2006], as well as recent studies showing that acute stress activates BLA adrenergic receptors [agnostic to the source of NE] to promote anxiety and other stress-related behaviors [Miranda et al., 2007; Buffalari and Grace, 2009a, 2009b; Chang and Grace, 2013]. Similarly acute stress paradigms cause selective activation of LC-NE neurons [McCall et al., 2015]. Together, there have been significant efforts to examine how source-independent noradrenergic [importantly, there are multiple sources of NE innervating the BLA [Robertson et al., 2016; Plummer et al., 2015] signaling in the BLA can alter synaptic plasticity, fear encoding, and memory consolidation, yet few studies have directly examined how the neuromodulatory LC-NE system utilizes BLA output to alter acute risk averse behaviors, such as anxiety [Grissom and Bhatnagar, 2011; Buffalari and Grace, 2009a, 2009b, 2007].

Downstream and independent of this projection, recent studies have demonstrated that direct activation of both basolateral amygdala [BLA] cell bodies or their projections is both anxiogenic and socially aversive [Siuda et al., 2015a, 2016; Tye et al., 2011; Felix-Ortiz et al., 2013, 2016; Felix-Ortiz and Tye, 2014]. Furthermore, it has also been demonstrated that increasing BLA excitatory output through Gαs G-protein activation, and more specifically, β-adrenergic receptor signaling causes acute social anxiety [Siuda et al., 2016]. Separately, noradrenergic cell firing in the LC has been shown to increase in the context of stressful stimuli [McCall et al., 2015; Abercrombie and Jacobs, 1987a, 1987b; Aston-Jones et al., 1999; Mana and Grace, 1997]. While the anatomical projections from the LC and their cell types have been studied for several years, the precise mechanisms by which fibers from the LC can directly influence BLA function to promote negative affective behaviors are not understood. Specifically, how LC-BLA projections generate affective behavioral responses through specific receptor systems and modulation of cell activity is unknown.

To determine the role of locus coeruleus noradrenergic influence on BLA function and negative affective behavior we optogenetically manipulated LC-NE inputs into the BLA, directly testing whether NE is released from LC terminals into the BLA and whether this terminal stimulation can drive anxiety-like and aversive behavioral responses. We demonstrate that photostimulation of LC projections to the basolateral amygdala releases NE and that this photostimulation evokes downstream modulation of neuronal activity in BLA neurons that project to anxiogenic brain regions. Stimulation of these fibers is sufficient to produce conditioned aversion and the stimulation-induced increased noradrenergic tone is sufficient to produce anxiety-like behavior mediated by local β-adrenergic receptor activity [β-ARs] in the BLA. Taken together, we report a previously undefined role for LC-BLA projections in mediating negative affective behavior through activation of β-ARs.

Results

Genetic and anatomical isolation of BLA-projection LC-NE neurons

Our previous work demonstrated that increased tonic LC-NE activity induces anxiety-like and aversive behavioral responses [McCall et al., 2015; Siuda et al., 2015a], we next sought to test whether these same behaviors can be generated by stimulating LC-NE fibers at localized projections in the BLA. To examine the potential sources of NE from the LC to the BLA we identified and isolated the projection using two distinct retrograde tracing approaches. First, we injected the tracer Fluorogold into the BLA of wild-type mice [Figure 1A]. Consistent with previous studies, this non-selective retrograde tracing approach revealed known inputs into the BLA from the LC [Robertson et al., 2016; Asan, 1998; Fallon et al., 1978] [Figure 1B]. We next used a dual injection strategy to anatomically isolate BLA-projecting LC-NE+ neurons. To do so, we used mice expressing Cre under the promoter for tyrosine hydroxylase, the rate-limiting enzyme for catecholamine synthesis [ThIRES-Cre mice][Savitt et al., 2005]. Here, we injected a red retrobead tracer into the BLA and the green-labeled adeno-associated virus, AAV5-DIO-Ef1α-ChR2[H134]-eYFP, into the LC [ThIRES-Cre::LC-BLA:ChR2; Figure 1C and D]. In these animals, we identified BLA-projecting Th+ LC neurons with the presence of both fluorophores in the same cells [Figure 1D]. Next, to examine Th+ terminal innervation of the BLA, we injected the cell-filling, Cre-dependent reporter AAV5-DIO-Ef1α-eYFP in the LC of ThIRES-Cre mice [ThIRES-Cre::LC-BLA:eYFP] [Figure 1E]. Here we clearly observed labeled LC neurons [Figure 1F and G] and their projection fibers terminating in the BLA [Figure 1H]. To further corroborate these findings, we examined recent projection experiments performed by the Allen Brain Institute Mouse Connectivity [ABIMC] project that also genetically and anatomically isolate this projection [Oh et al., 2014]. In three different experiments from three different genetic models, we observed LC-BLA projections that are qualitatively similar to our own observations [Figure 1—figure supplement 1A–L]. We present these findings here for clarity and ease of independent comparison, but this work was performed by ABIMC. Together, these anatomical studies identify a discrete projection of Th+ neurons from the LC that potentially release endogenous NE into the BLA.

Identifying a LC input to the BLA.

[A] Cartoon depicting fluorgold tracing strategy. [B] Representative image [selected from three injected mice] shows robust retrograde labeling of the LC from injection in the BLA [green = pseudocolored Fluorgold, tyrosine hydroxylase = red]. Arrowhead indicates example co-localization. Scale bar = 100 µm. 4V = 4th ventricle. The TH- cells dorsal and ventral to the LC are likely part of the medial parabrachial nucleus which has previously identified projections to the BLA [Saper and Loewy, 1980]. [C] Cartoon depicting dual injection tracing strategy for CTB-594 and DIO-ChR2-eYFP. [D] Representative images [selected from three injected mice] shows retrograde labeling in LC of red retrobeads and anterograde labeling of TH+ cells [green] [Nissl=blue]. Arrowhead indicates example co-localization. Scale bar = 100 µm [E] Cartoon depicting anterograde tracing strategy. [F–H] Coronal images depict robust eYFP [yellow] labeling in the LC [F and G] and BLA [H] of the same mouse [scale bars = [F] 50 µm, [G] 10 µm, [H] 20 µm. Inset [F], tyrosine hydroxylase = red, scale bar = 25 µm.

//doi.org/10.7554/eLife.18247.002

Optogenetic activation of LC-BLA terminals releases norepinephrine into the BLA

Using the same viral optogenetic strategy as above, we examined whether photostimulation of ThIRES-Cre::LC-BLA:ChR2 projections resulted in NE release at terminals. We validated functional ChR2 expression using whole-cell current clamp recordings of Th+ LC neurons. As, we previously demonstrated [McCall et al., 2015], this targeting method and photostimulation protocol was sufficient to generate action potentials at the LC cell bodies [Figure 2A]. Next, we used a carbon fiber microelectrode [CFME] to perform fast-scan cyclic voltammetry [FSCV] in the BLA during LC-NE terminal stimulation. Using the extended waveform method [see methods] in acute brain slices of ThIRES-Cre::LC-BLA:ChR2 animals, we stimulated slices with 30 5 ms light pulses from a 473 nm LED, at 10 Hz [Figure 2B]. Photostimulation of BLA slices, produced characteristic cyclic voltammograms and uptake consistent with NE [McElligott et al., 2013; Herr et al., 2012] [t1/2 = 2.0 ± 0.2 s, Figure 2C and D]. Following a 20 min baseline, 1 µM reserpine [an inhibitor of vesicular monoamine transporters] was perfused on the slices to deplete catecholamines from the axon terminals [Dahlstroem et al., 1965]. Reserpine treatment significantly attenuated the measured oxidative currents [31.6 ± 1.0% of baseline, Figure 2E and F] further confirming optically-evoked catecholamine release in this isolated ThIRES-Cre::LC-BLA:ChR2 projection. These findings suggest that optogenetic manipulation of LC-BLA terminals causes release of endogenous NE from the LC into the BLA.

Photostimulation of LC terminals in the BLA releases norepinephrine.

[A] LC neuron firing reliably to 10 Hz optical stimulation [CC=whole cell current clamp]. [B] Fast scan cyclic voltammetry [FSCV] schematic. [C–D] Oxidative and reductive currents [scale bar 2 s by 0.4 nA], with representative cyclic voltammograms [inset] and representative color plots [below] in response to photostimulation are attenuated by reserpine [1 μM]. Color plots for baseline and after reserpine [1 µM]: Files were collected over 15 s [X-axis] where the carbon fiber microelectrode was ramped with a triangular waveform from −0.4V to 1.3V and back to −0.4V at 400 V/S [Y-axis] and sampled at 10 Hz. 10 Hz, 473 nm blue LED stimulation onset at 2 s. Oxidative currents [nA] are positive in direction and reductive currents are negative [see color coded scale bar on right]. [E] Attenuation in NE oxidative current in response to reserpine [1 μM] n = 3 pairs; mean ± S.E.M]. [F] Average of first 20 min and last 15 min in [E] [Data represented as mean ± SEM, Paired Student’s t-tests to baseline, Mean difference = 68.56, t[2] = 18.75, **p=0.0028, 95% CI [52.82 to 84.29]].

//doi.org/10.7554/eLife.18247.004

In vivo photostimulation of LC-BLA terminals modulates BLA activity

BLA neurons have well reported responses to exogenous application of NE [Buffalari and Grace, 2007], but their response to endogenous NE release, explicitly from the LC, has not been previously defined. We next determined whether the optically-evoked endogenous catecholamine release would mimic the responses of BLA neurons to exogenous NE. To do so, we examined BLA single-unit activity in ThIRES-Cre::LC-BLA:ChR2 mice using 16-channel microelectrode arrays coupled to a fiber optic implant [optrode arrays] [McCall et al., 2015; Sparta et al., 2011]. These optrode arrays were used to isolate and record BLA single-unit activity before, during, and after photostimulation of ThIRES-Cre::LC-BLA:ChR2 projections [473 nm, 5 Hz, 10 ms pulse width] [Figure 3A and B]. In these experiments photostimulation of ThIRES-Cre::LC-BLA:ChR2 terminals caused an increase in firing frequency in 21.4% of units recorded in the BLA [Figure 3C–E,H; Figure 3—figure supplement 1A], while some cells [9.5%] displayed inhibitory responses [Figure 3D,F,I; Figure 3—figure supplement 1B]. The remaining neurons [69.0%] appeared to not change in response to photostimulation [Figure 3D,G,J; Figure 3—figure supplement 1C]. Furthermore, in a subset of cells blindly selected [without knowledge of the increase/decrease/static response to photostimulation] following the 5 Hz recordings, we also observed similar increases in firing rates to constant photostimulation, where the overall population of neurons increased firing during stimulation [Figure 3—figure supplement 1D and E]. In cases where the firing rate increased, the mean ± SD latency to fire following each light pulse was 129.3 ± 108.9 ms, suggesting that this change is not due to fast, monosynaptic neurotransmission, but likely a polysynaptic and/or neuromodulatory effect [Figure 3K]. Likewise, in cases where the firing rate decreased the mean latency was slower at 172.6 ± 77.01 ms, and neither case was significantly different to neurons whose firing rate did not change following photostimulation [260.4 ± 165.7] [Figure 3K].

Photostimulation of LC terminals in the BLA alters neuronal activity.

[A] Schematic illustrating single-unit extracellular recording paradigm of BLA neurons modulated by ChR2-expressing LC-BLA terminals. [B] Representative principal component analysis plot showing the first two principal components with clear clustering of a single unit [maroon] from the noise [grey]. Inset shows the waveform and spikes making up the isolated unit. Y-scale is 150 microvolts and x-scale is 500 ms. [C] Recordings from eight hemispheres of six Th-CreLC-BLA:ChR2 mice show the distribution of firing rates present in BLA neurons prior to and following LC-BLA terminal photostimulation [473 nm, 5 Hz, 3 min]. [D] Average normalized firing rate of neurons that increase [maroon], decrease [grey], or do not change [black] firing rate in response to photostimulation. Inset, shows number of neurons in each group. Representative histograms [1 s bins] of isolated single-units showing increase [E] or decrease [F], or no change [G] in neuronal firing in response to photostimulation [473 nm, 5 Hz, 3 min]. Z-scored population responses of neurons showing increase [H] or decrease [I], or no change [J] in neuronal firing in response to photostimulation. [K] Response latency following onset of photostimulation for cells that did not alter firing [=] [n = 29], increased firing [+] [n = 9], or decreased firing [-] [n = 4]. [Data represented as mean ± SD]. [L] The same cells sorted by baseline firing rate. [Data represented as mean ± SEM. Kruskal-Wallis test one-way ANOVA for non-parametric data with Dunn’s multiple comparisons test, Kruskal-Wallis statistic = 6.536, p=0.0381; + vs. – Mean rank difference = −18.75, adjusted *p=0.0329; + vs. + Mean rank difference = −6.828, adjusted p=0.4341; - vs. = Mean rank difference = 11.92, adjusted p=0.2053.] [M] Waveform similarity, within group distribution of linear correlations. Inset, every average waveform for each recorded unit separated by response profile [= black, + maroon, - grey]. [N–P] Bursting profiles for each recorded neuron. [N] Number of bursts per second. [Data represented as mean ± SEM]. [O] Mean firing rate within bursts for each neuron. [Data represented as mean ± SEM]. [P] Proportion of recorded spikes that occurred during bursts. [Data represented as mean ± SEM].

//doi.org/10.7554/eLife.18247.005

We next sought to determine whether these differences in neuronal response arise from distinct subsets of BLA neurons. To do so, we examined the baseline firing rate, waveforms, and bursting properties of recorded units. Neurons that increase firing in response to LC-BA terminal stimulation have a significantly lower mean basal firing rate [2.194 ± 2.22 SD] than those that decrease firing [13.06 ± 11.82 SD], however neither group is distinguishable from non-responsive neurons [Figure 3L]. To quantifiably assess the waveform shapes of the recorded neurons, we calculated the average waveform for each neuron and performed a linear correlation on these values within each group of neurons. These analyses demonstrate that all neurons that increase firing to LC-BLA photostimulation have an r value greater than 0.835 and all neurons that decrease have an r value above 0.820, while the population of neurons that did not significantly respond to photostimulation have an r above 0.515. These results indicate that neurons that do not respond to photostimulation are less internally similar, while the excited and inhibited cells are more similar within groups, suggesting that these modulated neurons are more likely to each be part of single class of neurons [Figure 3M]. Finally, we examined the bursting properties of the recorded neurons. While most neurons exhibited some bursting properties, no differences were found between groups in terms of frequency of bursting [Figure 3N], mean firing rate during bursts [Figure 3O], or the percentage of spikes from the recording sessions that occurred within a burst [Figure 3P]. Together, these heterogeneous firing properties are consistent with previous studies using iontophoresis of NE into the BLA, and further highlight the complex pharmacological and circuit activity that NE modulates within the BLA [Buffalari and Grace, 2007; Ferry et al., 1997; Huang et al., 1996]. These results suggest that photostimulation of LC terminals in the BLA causes varied neuronal firing rate responses, with the majority of photostimulation-responsive neurons increasing firing in response to the LC-NE terminal photostimulation.

LC-BLA terminal stimulation biases activation towards anxiety-promoting BLA neurons

The BLA is thought to be a heterogeneous hub for emotional processing containing separable populations for regulating either positive or negative affect. Recent studies have suggested that these opposing populations are distinct in either their projection target or their cell-type [Stuber et al., 2011; Namburi et al., 2015; Beyeler et al., 2016; Felix-Ortiz et al., 2016; Burgos-Robles et al., 2017; Kim et al., 2016, 2017; Correia et al., 2016]. Given the role of BLA adrenergic receptors in modulating anxiety-like and aversion behaviors [Siuda et al., 2015a, 2016], we hypothesized that LC-NE innervation of BLA neurons may preferentially bias activation of neurons that promote anxiety-like behavior such as the ventral hippocampus [vHPC]- and central amygdala [CeA]- projecting BLA neurons as opposed to projections that promote positive affect and anxiolysis such as nucleus accumbens [NAc]-projecting BLA neurons. Using a combination of retrograde viral tracing, immunohistochemistry, and optogenetic stimulation in Th-CreLC-BLA:ChR2 mice, we assessed these potential tri-synaptic circuits that may underlie anxiety-like behaviors [Figure 4A]. Photostimulation of LC-NE terminals in the BLA [5 Hz, 10 ms] significantly increases the number of cFos-expressing BLA neurons in ThIRES-Cre::LC-BLA:ChR2 animals compared to ChR2-lacking [Figure 4B; Figure 4—figure supplement 1A–D] and contralateral BLA controls following photostimulation [Figure 4—figure supplement 1E]. To assess whether the cFos+ BLA neurons resulting from LC-NE terminal activation were biased toward a particular class of BLA projection neurons we next repeated the experiment with the retrograde tracer Cholera toxin subunit B [CTB] injected into BLA projection regions [vHPC, CeA, and NAc]. These immunohistochemistry studies reveal that the cFos present following LC-NE terminal activation in the BLA, overlaps significantly more with vHPC- and CeA-targeted compared to NAc-targeted CTB in the BLA [Figure 4C–G; Figure 4—figure supplement 1F–H]. These results suggest that LC-NE terminals in the BLA preferentially activate vHPC- and CeA- projecting BLA neurons thought to be involved in modulating negative valence and affective behaviors.

Photostimulation of LC terminals in the BLA preferentially activates BLA circuitry associated with anxiety-like behavior.

[A] Diagram of viral and optogenetic strategy. [B] 5 Hz photostimulation increases cFos expression within the BLA in ThIRES-Cre::LC-BLA:ChR2+ animals compared to ThIRES-Cre::LC-BLA:ChR2- controls [Data represented as mean ± SEM, n = 9 ChR2, n = 4 Ctrl; average of 3 sections/mouse; Student’s t-test, Mean difference = 19.17, t[10] = 4.005, **p=0.0040, 95% CI [−35.47 to −10.11]. [C] 5 Hz photostimulation increases cFos expression significantly more in BLA neurons projecting to the vHPC and CeA compared to NAc in ThIRES-Cre::LC-BLA:ChR2 animals [Data represented as mean ± SEM, n = 9 vHPCCTB, n = 6 CeACTB, n = 9 CeACTB, 3 sections per mouse; One-Way ANOVA, Bonferroni’s Multiple Comparison Test, F2,20 = 7.199, **p=0.0044; ThIRES-Cre::LC-BLA:ChR2:CTB-vHPC vs. ThIRES-Cre::LC-BLA:ChR2:CTB-NAc Mean difference = 12.95, t[20] = 3.585, **p

Chủ Đề