Skip to main content
Erschienen in: Sleep and Biological Rhythms 2/2017

Open Access 01.04.2017 | Original Article

Activation of the ventral tegmental area increased wakefulness in mice

verfasst von: Huan-Xin Sun, Dian-Ru Wang, Chen-Bo Ye, Zhen-Zhen Hu, Chen-Yao Wang, Zhi-Li Huang, Su-Rong Yang

Erschienen in: Sleep and Biological Rhythms | Ausgabe 2/2017

Abstract

The ventral tegmental area (VTA) is crucial for brain functions, such as voluntary movement and cognition; however, the role of VTA in sleep-wake regulation when directly activated or inhibited remains unknown. In this study, we investigated the effects of activation or inhibition of VTA neurons on sleep-wake behavior using the pharmacogenetic “designer receptors exclusively activated by designer drugs (DREADD)” approach. Immunohistochemistry staining was performed to confirm the microinjection sites, and combined with electrophysiological experiments, to determine whether the VTA neurons were activated or inhibited. The hM3Dq-expressing VTA neurons were excited confirmed by clozapine-N-oxide (CNO)-driven c-Fos expression and firing in patch-clamp recordings; whereas the hM4Di-expressing VTA neurons inhibited by reduction of firing. Compared with controls, the activation of VTA neurons at 9:00 (inactive period) produced a 120.1% increase in the total wakefulness amount for 5 h, whereas NREM and REM sleep were decreased by 62.5 and 92.2%, respectively. Similarly, when VTA neurons were excited at 21:00 (active period), the total wakefulness amount increased 81.5%, while NREM and REM sleep decreased 64.6 and 93.8%, respectively, for 8 h. No difference of the amount and EEG power density of the NREM sleep was observed following the arousal effects of CNO. The inhibition of VTA neurons during active or inactive periods gave rise to no change in the time spent in the wakefulness, REM, and NREM sleep compared with control. The results indicated that VTA neurons activated pharmacogentically played important roles in promoting wakefulness.
Hinweise
An erratum to this article is available at https://​doi.​org/​10.​1007/​s41105-017-0115-9.

Introduction

The ventral tegmental area (VTA) contains dopamine (DA) containing neurons and is traversed by many blood vessels and nerve fibers. Close to its poorly defined borders are found DA and 5-HT containing neurons [1]. The VTA comprises several subdivisions and neurons synthesizing neurotransmitters, such as DA, GABA, or glutamate [2]. The DA neurons arises in the VTA and projects to: (1) the septal area, olfactory tubercle, nucleus accumbens, amygdaloid complex, and piriform cortex (mesolimbic projection) and (2) the medial prefrontal, cingulate, and entorhinal areas (mesocortical projection) [1]. The VTA is crucial for brain functions, such as voluntary movement and goal-directed behavior, as well as cognition, emotion, reward, working memory, and decision making [36].
Accordingly, DA midbrain system dysfunction is associated with neurologic and psychiatric diseases, such as schizophrenia, addiction, attention deficit hyperactivity disorder, and Parkinson’s disease [7]. Several lines of evidence suggest that DA signals are closely associated with regulation of sleep-wake activity [8, 9]. In humans, malfunction of the brain DA system is associated with neuropsychopathy, with disruption of sleep as one symptom [10]. Systemic administration of a selective D1 receptor agonist induced an increase of wakefulness (W) and a reduction of slow wave sleep (SWS) and REM sleep. On the other hand, low doses of D2 receptor agonist reduced W and increased SWS and REM, whereas large doses induced the opposite effect [11]. Psycho-stimulants promote wakefulness by inhibiting the DA reuptake transporter (DAT) and thereby enhancing the extracellular DA levels in mice [12]. The results are often controversial by drug effects on targets outside the VTA area or by potential disruption of neural circuits during development after gene knockout. The effects of activation or inhibition of VTA neurons on sleep/wake behavior remains to be directly tested.
The VTA is a heterogeneous area where dopaminergic and non-dopaminergic neurons are intermixed [2]. The pharmacogenetic technique “designer receptors exclusively activated by a designer drug” DREADD [13] utilizes extrinsic muscarinic receptors (hM3Dq for excitation and hM4Di for inhibition) that have lost their affinity for endogenous acetylcholine but can still be activated by a synthetic inert ligand (clozapine-N-oxide [CNO]). Activation of Gq-coupled hM3Dq by CNO has previously been shown to activate neurons through a phospholipase C-dependent mechanism [14]. CNO can also stimulate the Gi/o-coupled hM4Di receptor, thereby activating the inwardly rectifying potassium 3 channel, resulting in membrane hyperpolarization and neuronal silencing [13]. In this way, the activity of VTA neurons can be manipulated temporarily and reversibly. Our study established a direct link between the activity of midbrain VTA neurons and arousal promotion. The results will contribute for understanding the roles of the VTA area in regulation of sleep and wake states, which will provide a therapeutic basis for sleep disorders involved in VTA neurons.

Materials and methods

Animals

Pathogen-free adult male S129 mice (25–30 g) were obtained from the Laboratory Animal Center of Chinese Academy of Sciences (Shanghai, China). The animals were housed in individual cages at a constant temperature (24 ± 1 °C) with a relative humidity (60 ± 2%) on an automatically controlled 12:12 light/dark cycle (lights on at 7:00). The mice had free access to food and water. The experimental protocols were approved by the Committee on the Ethics of Animal Experiments of the University of Fudan, Shanghai Medical College (Shanghai, China). Every effort was made to minimize the number of animals used and any pain and discomfort experienced by the animals.

Surgery

Virus injection

Male mice were anesthetized with chloral hydrate (5% in saline [360 mg/kg]), then placed in a stereotaxic frame, so that the head was fixed. A burr hole was made, and a fine glass pipette (1 mm glass stock, tapering slowly to a 10–20 μm tip) containing adeno-associated viral (AAV) was bilaterally lowered to the VTA (coordinates relative to bregma: anterioposterior = − 3.4 mm, mediolateral = ± 0.3 mm, dorsoventral = −4.0 mm) [15]. Then, the AAV vectors (0.05 µl per side) were injected with nitrogen gas pulses. After ten additional min, the pipette was slowly withdrawn. In our study, we employed an evolved G-protein-coupled muscarinic receptor (hM3Dq or hM4Di) that is selectively activated or inhibited by the exogenous ligand CNO. The animals recovered for 2 weeks before electrodes were implanted for electroencephalogram (EEG) and electromyogram (EMG) recordings.

Sleep recording and vigilance state analysis

To monitor EEG signals, two stainless steel screws were positioned 1 mm anterior to the bregma, both of which were 1.5 mm laterals to the midline [15]. Two Teflon-coated, stainless steel wires were placed into the trapezius muscles bilaterally for EMG recording. All electrodes were attached to a micro-connector and fixed to the skull with dental cement. After surgery, each mouse was allowed recover for 1 week before polygraphic recording. The recordings were started at 7:00 (onset of light period) or 19:00 (onset of dark period). We recorded EEG and EMG for two consecutive days. On day 1, the mice were treated with vehicle (intrape ritoneal [i.p.] saline) at 9:00 or 21:00, and the recordings served as the baseline data. On the next day, CNO (1 mg/kg) was injected at the same time on day 1. The same mice were used four times for different injection time and drug (9:00 saline/CNO; 21:00 saline/CNO). The interval for each experiment was 2 days.
The EEG and EMG signals were amplified and filtered (EEG, 0.5–30 Hz; EMG, 20–200 Hz), then digitized at a sampling rate of 128 Hz and recorded by VitalRecorder (Kissei Comtec, Nagano, Japan) [16]. Vigilant states were automatically classified off-line under 10-s epochs into three stages, i.e., wakefulness, rapid eye movement (REM), and non-rapid eye movement (NREM) sleep, using SLEEPSIGN (Kissei Comtec) according to standard criteria [17]. Wakefulness is defined by a low-amplitude and high-frequency EEG with a high activity of EMG. REM sleep is characterized by a low-amplitude, high-frequency EEG associated with the absence of EMG activity; the presence of EEG theta-activity (6–9 Hz) in the recording can be used to confirm this state. NREM sleep is commonly defined by a high-amplified EEG associated with a low-voltage EMG. The presence of high EEG delta activity (0.65–4 Hz) is also employed to characterize this state. As a final step, defined sleep-wake stages were examined visually and corrected if necessary [18].

Drugs and administration

CNO (LKT Labs C4759, USA) was dissolved in saline to a concentration of 1 mg/10 ml before treatment. CNO was administered by i.p.to each mouse (0.1 ml/10 g body weight).

Immunohistochemistry

One group of mice was injected with saline and the other with CNO at a dose of 1 mg/kg. Two hours later, the animals were deeply anesthetized with 5% chloral hydrate (360 mg/kg) and were immediately perfused transcardially with 60 ml of saline, followed by 50 ml of ice-cold 4% paraformaldehyde (PFA). The brains were removed and post-fixed for 24 h at 4 °C in 4% PFA, then equilibrated in phosphate buffer containing 30% sucrose at 4 °C. Coronal Sect. (30 μm) were cut serially on a cryostat. For double immunohistochemistry analysis, sections were thrice washed with phosphate-buffered saline (PBS) and incubated with 0.3% hydrogen peroxide in 0.1 M PBS for 30 min. After washing in PBS, the sections were incubated with a primary rabbit anti-c-Fos (1:10000, Millipore 2239640, USA) antibody diluted in PBS containing 0.3% Triton X-100 (PBST) for 48 h. On the second day, the sections were rinsed and incubated in biotinylated anti-rabbit secondary antiserum (1:1000, Jackson ImmunoResearch BA-100, West Grove, PA, USA) for 2 h. After thrice washing in PBS, all sections were then treated with avidin-biotin-peroxidase complex (1:1:1000, Vectastain ABC Elite kit; Vector Laboratories PK-6100, San Francisio, CA, USA) for 90 min. The peroxidase reactions were visualized with 0.05% 3, 3- diamino-benzidine (DAB; Sigma SK-4100, St. Louis, MO, USA) in 0.1 M PB and 0.01% hydrogen peroxide. After three washes in PBS, the sections were incubated in another primary rabbit anti-DsRed (1:10000, Takara 632496, USA) antibody diluted in PBST overnight. The procedure was similar as c-Fos staining previously described above with the exception of the last step; the color reagents of DAB did not include nickel. All sections were mounted onto gelatin-coated slides, dehydrated in graded ethanol, placed into xylene, and cover-slipped.
The sections were examined under bright-field illumination with a microscope (Leica Microsystems, Wetzlar, Germany). Using light microscopy, neurons positive for c-Fos were identified by dense black nuclear staining; the mCherry were identified by brown cytomembrane staining. Locations in the brain were confirmed by staining and reference to the primary literature and a mouse brain atlas.

Electrophysiologic experiments

Preparation for brain slices

Electrophysiologic experiments were performed on slices retrieved from the mice injected with virus 3 weeks earlier. Briefly, the mice were sacrificed by decapitation, and the VTA and SNc were identified according to stereotaxic coordinates [15]. Coronal midbrain slices (300 μm thick) containing VTA and SNc were cut using a vitratome (Leica VT 1000 S, USA) in ice-cold glycerol-based artificial cerebrospinal fluid containing: glycerol (260 mM), KCL (5 mM), KH2PO4 (1.25 mM), MgSO4 (1.3 mM), CaCL2 (0.5 mM), NaHCO3(20 mM) and glucose (10 mM), and saturated with 95% O2–5% CO2. Slices were allowed to recover for at least 1 h in a holding chamber at a water bath (32 °C) before recording.

Patch-clamp recordings in the whole-cell configuration

This recording configuration was used to study the effects of CNO application on a single VTA/SNc cell. The path electrodes were pulled from borosilicate glass capillaries (1.5 mm outside diameter, 0.86 inside diameter, Sutter Instrument, Novato, CA) on a Brown-Flaming micropipette (Model P-97, Sutter Instrument, Novato, CA, USA). The path electrode had a resistance of 4–6 MΩ when filled with pipette solution containing potassium gluconate (120 mM), KCl (20 mM), MgCl2 (1 mM), CaCl2 (0.16 mM), HEPES (10 mM), EGTA (0.5 mM), MgATP (2 mM), and NaGTP (0.5 mM); with the pH was adjusted to 7.4 with KOH. A single coronal slice was transferred to the recording chamber, where the slice was held down with a platinum ring. Artificial cerebrospinal fluid (ACSF) gassed with 95% O2–5% CO2 flowed through the bath (2 ml/min). VTA/SNc neurons were identified under visual guidance using upright microscope (BX-51, Olympus, Japan) with a 40× water immersion objective lens. The cells were recorded using mCherry fluorescent signals. The image was detected with a CCD camera (U-TV1X-2, Olympus, Japan) and displayed on a monitor. The mCherry-positive neurons were clamped to record spontaneous action potentials and/or membrane potentials. The series resistance (3–5 MΩ) and input resistance (300–400 MΩ) were monitored throughout the cell recording, and data were discarded when either of the two resistances changed by >20% [19].

Statistical analysis

All data were presented as the mean ± standard error. Histograms of the amounts of sleep and wakefulness after vehicle or CNO injection were compared using a Student’s two-tailed paired t test. The hourly amounts of each stage for sleep–wake profiles in mice treated with vehicle or CNO were compared using repeated-measures analysis of variance followed by Fisher’s probable least-squares difference (PLSD) test, p < 0.05 was considered statically significant.

Results

Injection sites confirmed by mCherry expression of AAV

To control the VTA neurons in behaving animals, we bilaterally injected the AAV containing excitatory (hSyn-hM3Dq-mCherry-AAV10) or inhibitory (hSyn-hM4Di-mCherry-AAV10) modified muscarinic G-protein-coupled receptors which can be activated by CNO into the VTA (Fig. 1AB). AAV vectors without hM3Dq or hM4Di receptors (hSyn-mCherry-AAV10) were used as control (Fig. 1Cc). Robust cell-surface expression of the hM3Dq (Fig. 1Dd) or hM4Dq receptors (Fig. 1Ee) was observed in the VTA.

Modulation of the activities of VTA neurons by CNO

CNO was given by i.p. with minimal perturbation. The effects of excitatory hM3Dq and inhibitory hM4Di were examined during both the light and dark periods. After EEG/EMG recording, following treatment of CNO or saline in mice expressing mCherry, hM3Dq, or hM4Di, the animals were sacrificed and fixed 2 h later for staining. The VTA slices of these mice were examined by double staining with anti-mCherry and anti-Fos antibodies to assess the activity of these VTA neurons.
In wild-type mice with control AAV (hSyn-mCherry-AAV10) microinjection, in which VTA neurons did not express hM3Dq or hM4Di, the double-labeled neurons (mCherry-positive neurons with c-Fos-positive nuclei) were not seen (Fig. 2a). There was an apparent increase in the number of double-labeled neurons in the mice microinjected with hSyn-hM3Dq-AAV10 (Fig. 2b). Because the neuron activity was inhibited by hM4Di, the mice microinjected with hSyn-hM4Di-AAV10 showed no expression of the double-labeled neurons (Fig. 2c). Then, we carried out target recordings from hM3Dq/mCherry or hM4Di/mCherry neurons in brain slices to determine the effects of CNO on VTA neurons.
Whole-cell current–clamp recordings demonstrated that CNO (5 μmol/l) elicited vigorous firing of action potentials (Fig. 2d) in hM3Di/mCherry-positive neurons, but inhibited firing of VTA hM4Di/mCherry neurons (Fig. 2e).

Modulation of activities of VTA neurons altered behavioral states in mice

After confirming the effectiveness of the pharmacogenetic approach, we tested whether or not the selective activation or inhibition of VTA neurons affects sleep-wake states. When CNO was given at 9:00 or 21:00, the amount of wakefulness, REM, and NREM sleep in mice microinjected hSyn-mCherry-AAV10 or hSyn-hM4Di-AAV10 did not change compared with the control group (Figs. 3a, c, 4a, c).
However, the mice microinjected with hSyn-hM3Dq- AAV10 showed a remarkable increase in wakefulness accompanied by reduction in REM and NREM sleep after CNO administration (Fig. 3b, d, e). The sum of wakefulness in 5 h following CNO injection on light period increased 144.9% (174.2 ± 3.1 versus 71.1 ± 8.9 min, p < 0.01); meanwhile, the sum of REM sleep decreased 100% (0 versus 13.9 ± 1.5 min, p < 0.01), and NREM sleep decreased 94.0% (5.8 ± 3.1 versus 94.8 ± 7.9 min, p < 0.01) (Fig. 3f). When CNO was administered at 21:00 (active period), a similar phenomenon was noted in the mice microinjected with hSyn-hM3Dq- AAV10 as the wakefulness increased, while REM and NREM sleep decreased (Fig. 4b, d, e). The sum of wakefulness in 8 h following CNO injection increased 81.5% (396.3 ± 16.8 versus 218.4 ± 12.4 min, p < 0.01), meanwhile, the sum of REM sleep decreased 93.8% (1.9 ± 0.9 versus 30.8 ± 2.4 min, p < 0.01) and NREM sleep decreased 64.6% (81.8 ± 16.3 versus 230.9 ± 10.4 min, p < 0.01) (Fig. 4f).
Next, we analyze the EEG power density of NREM sleep after the VTA neurons’ activities were modified. In the mice microinjected with AAV-hM4Di or AAV- control, EEG power density of NREM sleep did not change (data were not shown). As for the hM3Dq positive mice, although the activation of the VTA neurons by CNO injection at 9:00 (Fig. 3g) or 21:00 (Fig. 4g) increased remarkable wakefulness, but no difference of the amount and EEG power density of the NREM sleep was observed during the time period following the arousal effects of CNO.

Discussion

The first occurrence of the VTA in the literature is due to Tsai in 1925 [20]. In its description of the opossum brain, he identified with Nissl and Golgi staining a region lateral to the interpeduncular nucleus as the trigonum interpeduncular. Recently, the VTA DA neurons were thought to be homogeneous in their properties and behavioral functions, such that they express characteristic phasic excitatory responses to rewards and cue that predict rewards while being inhibited by omission of rewards [21]. A prior study wherein the catecholamine-containing neurons in mesencephalon were given electrolytic lesion in cats and the animals had an obvious decrease in behavioral waking indicated the catecholamine-containing neurons played a role in sleep/wake regulation [22]. However, this method produced irreversible death of neurons and may also lead to some compensatory reactions of other brain areas. Despite higher dopamine levels usually associated with arousal [12, 23], the previous studies did not establish a direct link between the activity of VTA neurons and wakefulness.
In the present study, we applied DREADD to pharmacogenetically manipulate the activity of VTA neurons. DREADD utilizes G-protein-coupled receptor signaling, which can affect neuronal activity in a relatively chronic and consistent manner and allow investigation of the behavioral effects of activation or inhibition of VTA neurons. Besides, the AAV has proven to be very effective for studying neuronal function and behavior, it is long-lived, highly neurotropic, non-pathogenic and, importantly, it can drive transgene expression at significantly high levels necessary to alter neuronal function and produce a behavioral phenotype [24]. In this study, we confirmed its effect by in vitro electrophysiological experiments, in which the spontaneous firing rates of the hM3Dq-positive neurons were increased obviously, while the hM4Di-positive neurons were inhibited by CNO administration. In this way, we can selectively manipulate the neurons in the VTA acutely and reversibly.
Considering that the firing activity of DA neurons peaked between 07:00 and 11:00 and between 19:00 and 23:00 [1], we examined the effects of the pharmacogenetic activation of the VTA neurons either at 9:00 or at 21:00 on sleep-wake behaviors in mice. We found no matter the level of sleep pressure was high or low [25], and the activation of VTA produced remarkable increase in wakefulness, which lasted longer when CNO was given during the active period in mice. The findings demonstrate for the first time that injection of CNO in mice expressing hM3Dq in the VTA is able to significantly increase arousal. Our results were consistent with previous reports that DAT knockout mice exhibited an increase in wakefulness [23], selectively activating midbrain dopaminergic neurons using DREADD induced hyperactivity in DAT-Cre mice [6], and optogenetic stimulation of VTA dopaminergic neurons initiated and maintained wakefulness in TH-Cre mice [26].
Although the VTA is generally known as the source of dopaminergic projection neurons, the VTA contains multiple cell types: dopaminergic (about 60%); GABAergic (about 35%); and glutamatergic neurons (about 5%) [2, 27]. Therefore, we think that all types of VTA neurons were simultaneously activated by this nonspecific pharmacogenetic DREADD system. VTA DA neurons project heavily to several limbic structures, including the nucleus accumbens, amygdala, and prefrontal cortex. In the rat, both GABAergic and glutamatergic neurons form local synapses in the VTA [28, 29] and project in parallel with the DA neurons to limbic regions [30, 31]. As a result, the increased wakefulness induced by activation of VTA may be caused not only by activated dopamine neurons but also by glutamatergic neurons. It may increase the dopamine and glutamate levels in targeted area which may also contribute to the wakefulness.
It is recognized that sleep is under the control of circadian and homeostatic processes, and animals will attempt to regain or compensate for sleep that was previously depleted [32, 33]. Previous studies showed that sleep deprivation often accompanied with the increase of EEG power density of NREM [34, 35]. Our results showed that the activation of VTA no matter during the day or night is responsible for wakefulness. However, the NREM sleep amount or delta power in mice expressing hM3Dq receptors treated with CNO following the induced extended wakefulness did not increase. The possible reasons may be involved the different methods of sleep deprivation, in which gentle handling or water platform techniques were employed in the previous studies, whereas the arousal was induced by VTA activation using DREADD approach. The present phenomena was in agreement with Qiu’s study, in which they did not find EEG or behavioral sleep rebound even after 4 days of induced wakefulness by chemogenetic stimulation of the pontine parabrachial nucleus [36].
To investigate whether the VTA has some roles in physiological sleep regulation, the sleep analysis in mice with hSyn-hM4Di-AAV10 was carried out. The inhibition of VTA neurons was confirmed with patch-clamp recording, in which CNO bath application decreased the firing of the mCherry-hM4Di positive neurons. We found there were no differences in the mice microinjected with hSyn-hM4Di-AAV10 between CNO and saline group in sleep-wake behavior. The reason may be a comprehensive effect by inhibition of different type neurons in VTA. When the VTA nucleus was inhibited, on one hand, the dopaminergic neurons were inhibited; on the other hand, the inhibition of the GABAergic neurons could produce disinhibition of local dopaminergic neurons. However, we cannot exclude some roles of single cell type neurons in VTA in sleep-wake regulation under physiological conditions. In fact, a recent study has shown the VTA dopaminergic neurons are necessary for wakefulness since the inhibition of these neurons suppressed wakefulness in TH-Cre mice [26].
In this study, we investigated the role of the VTA in sleep/wake regulation; however, viruses indiscriminately transduce most cell types within the introduced region. Next, we will manipulate neuronal subtypes independently to understand the genetic basis of the sleep-wake phenomena for therapeutic purposes.

Compliance with ethical standards

Conflict of interest

This was not an industry supported study. None of the authors has any financial interest or conflicts of interest related to this work.

Source of funding

This study was supported by the National Natural Science Foundation of China (81420108015 and 81571296) and Shanghai Leading Academic Discipline Project (B119).

Ethical committee permission

Experimental protocols were approved by the Shanghai Medical Experimental Animal Administrative Committee.

Animals

Male S129 mice were obtained from the Laboratory Center, Chinese Academy of Sciences (Shanghai, China).
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

e.Med Neurologie

Kombi-Abonnement

Mit e.Med Neurologie erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes, den Premium-Inhalten der neurologischen Fachzeitschriften, inklusive einer gedruckten Neurologie-Zeitschrift Ihrer Wahl.

Literatur
1.
Zurück zum Zitat Oades RD, Halliday GM. Ventral tegmental (A10) system: neurobiology. 1. Anatomy and connectivity. Brain Res. 1987; 434(2):117–65.CrossRefPubMed Oades RD, Halliday GM. Ventral tegmental (A10) system: neurobiology. 1. Anatomy and connectivity. Brain Res. 1987; 434(2):117–65.CrossRefPubMed
2.
Zurück zum Zitat Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam JP, Ungless MA. Stereological estimates of dopaminergic, GABAergic and glutamatergic neurons in the ventral tegmental area, substantia nigra and retrorubral field in the rat. Neuroscience. 2008; 152(4):1024–1031.CrossRefPubMedPubMedCentral Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam JP, Ungless MA. Stereological estimates of dopaminergic, GABAergic and glutamatergic neurons in the ventral tegmental area, substantia nigra and retrorubral field in the rat. Neuroscience. 2008; 152(4):1024–1031.CrossRefPubMedPubMedCentral
3.
Zurück zum Zitat Grace AA, Floresco SB, Goto Y, Lodge DJ. Regulation of firing of dopaminergic neurons and control of goal-directed behaviors. Trends Neurosci. 2007; 30(5):220–7.CrossRefPubMed Grace AA, Floresco SB, Goto Y, Lodge DJ. Regulation of firing of dopaminergic neurons and control of goal-directed behaviors. Trends Neurosci. 2007; 30(5):220–7.CrossRefPubMed
4.
Zurück zum Zitat D’Ardenne K, Eshel N, Luka J, Lenartowicz A, Nystrom LE, Cohen JD. Role of prefrontal cortex and the midbrain dopamine system in working memory updating. Proc Natl Acad Sci USA. 2012; 109(49):19900–19909.CrossRefPubMedPubMedCentral D’Ardenne K, Eshel N, Luka J, Lenartowicz A, Nystrom LE, Cohen JD. Role of prefrontal cortex and the midbrain dopamine system in working memory updating. Proc Natl Acad Sci USA. 2012; 109(49):19900–19909.CrossRefPubMedPubMedCentral
5.
Zurück zum Zitat Chowdhury R, Guitart-Masip M, Lambert C, Dayan P, Huys Q, Duzel E et al. Dopamine restores reward prediction errors in old age. Nat Neurosci. 2013; 16(5):648–53.CrossRefPubMedPubMedCentral Chowdhury R, Guitart-Masip M, Lambert C, Dayan P, Huys Q, Duzel E et al. Dopamine restores reward prediction errors in old age. Nat Neurosci. 2013; 16(5):648–53.CrossRefPubMedPubMedCentral
6.
Zurück zum Zitat Wang S, Tan Y, Zhang JE, Luo M. Pharmacogenetic activation of midbrain dopaminergic neurons induces hyperactivity. Neurosci Bull. 2013; 29(5):517–24.CrossRefPubMed Wang S, Tan Y, Zhang JE, Luo M. Pharmacogenetic activation of midbrain dopaminergic neurons induces hyperactivity. Neurosci Bull. 2013; 29(5):517–24.CrossRefPubMed
7.
Zurück zum Zitat Brisch R, Saniotis A, Wolf R, Bielau H, Bernstein HG, Steiner J, et al. The role of dopamine in schizophrenia from a neurobiological and evolutionary perspective: old fashioned, but still in vogue. Front. Psychiatry. 2014;5:47. Brisch R, Saniotis A, Wolf R, Bielau H, Bernstein HG, Steiner J, et al. The role of dopamine in schizophrenia from a neurobiological and evolutionary perspective: old fashioned, but still in vogue. Front. Psychiatry. 2014;5:47.
8.
Zurück zum Zitat Qu WM, Xu XH, Yan MM, Wang YQ, Urade Y, Huang ZL. Essential role of dopamine D2 receptor in the maintenance of wakefulness, but not in homeostatic regulation of sleep, in mice. J Neurosci. 2010; 30(12):4382–4389.CrossRefPubMed Qu WM, Xu XH, Yan MM, Wang YQ, Urade Y, Huang ZL. Essential role of dopamine D2 receptor in the maintenance of wakefulness, but not in homeostatic regulation of sleep, in mice. J Neurosci. 2010; 30(12):4382–4389.CrossRefPubMed
9.
Zurück zum Zitat Qiu MH, Qu WM, Xu XH, Yan MM, Urade Y, Huang ZL. D(1)/D(2) receptor-targeting L-stepholidine, an active ingredient of the Chinese herb Stephonia, induces non-rapid eye movement sleep in mice. Pharmacol Biochem Behav. 2009 Nov;94(1):16–23.CrossRefPubMed Qiu MH, Qu WM, Xu XH, Yan MM, Urade Y, Huang ZL. D(1)/D(2) receptor-targeting L-stepholidine, an active ingredient of the Chinese herb Stephonia, induces non-rapid eye movement sleep in mice. Pharmacol Biochem Behav. 2009 Nov;94(1):16–23.CrossRefPubMed
10.
Zurück zum Zitat Telarovic S, Mijatovic D, Telarovic I. Effects of various factors on sleep disorders and quality of life in Parkinson’s disease. Acta Neurol Belg. 2015; 115(4):615–21.CrossRefPubMed Telarovic S, Mijatovic D, Telarovic I. Effects of various factors on sleep disorders and quality of life in Parkinson’s disease. Acta Neurol Belg. 2015; 115(4):615–21.CrossRefPubMed
11.
Zurück zum Zitat Monti JM, Monti D. The involvement of dopamine in the modulation of sleep and waking. Sleep Med Rev. 2007; 11(2):113–33.CrossRefPubMed Monti JM, Monti D. The involvement of dopamine in the modulation of sleep and waking. Sleep Med Rev. 2007; 11(2):113–33.CrossRefPubMed
12.
Zurück zum Zitat Schenk JO. The functioning neuronal transporter for dopamine: kinetic mechanisms and effects of amphetamines, cocaine and methylphenidate. Prog Drug Res. 2002;59:111–31.CrossRefPubMed Schenk JO. The functioning neuronal transporter for dopamine: kinetic mechanisms and effects of amphetamines, cocaine and methylphenidate. Prog Drug Res. 2002;59:111–31.CrossRefPubMed
13.
Zurück zum Zitat Armbruster BN, Li X, Pausch MH, Herlitze S, Roth BL. Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc Natl Acad Sci USA. 2007; 104(12):5163–5168.CrossRefPubMedPubMedCentral Armbruster BN, Li X, Pausch MH, Herlitze S, Roth BL. Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc Natl Acad Sci USA. 2007; 104(12):5163–5168.CrossRefPubMedPubMedCentral
14.
Zurück zum Zitat Alexander GM, Rogan SC, Abbas AI, Armbruster BN, Pei Y, Allen JA, et al. Remote control of neuronal activity in transgenic mice expressing evolved G protein-coupled receptors. Neuron. 2009; 63(1):27–39.CrossRefPubMedPubMedCentral Alexander GM, Rogan SC, Abbas AI, Armbruster BN, Pei Y, Allen JA, et al. Remote control of neuronal activity in transgenic mice expressing evolved G protein-coupled receptors. Neuron. 2009; 63(1):27–39.CrossRefPubMedPubMedCentral
15.
Zurück zum Zitat Paxinos G, Franklin KBJ. The mouse brain in stereotaxic coordinates, 2nd edn (2001). Paxinos G, Franklin KBJ. The mouse brain in stereotaxic coordinates, 2nd edn (2001).
16.
Zurück zum Zitat Oishi Y, Huang ZL, Fredholm BB, Urade Y, Hayaishi O. Adenosine in the tuberomammillary nucleus inhibits the histaminergic system via A1 receptors and promotes non-rapid eye movement sleep. Proc Natl Acad Sci USA. 2008 ;105(50):19992–19997.CrossRefPubMedPubMedCentral Oishi Y, Huang ZL, Fredholm BB, Urade Y, Hayaishi O. Adenosine in the tuberomammillary nucleus inhibits the histaminergic system via A1 receptors and promotes non-rapid eye movement sleep. Proc Natl Acad Sci USA. 2008 ;105(50):19992–19997.CrossRefPubMedPubMedCentral
17.
Zurück zum Zitat Huang ZL, Qu WM, Eguchi N, Chen JF, Schwarzschild MA, Fredholm BB, et al. Adenosine A2A, but not A1, receptors mediate the arousal effect of caffeine. Nat Neurosci. 2005; 8(7):858–9.CrossRefPubMed Huang ZL, Qu WM, Eguchi N, Chen JF, Schwarzschild MA, Fredholm BB, et al. Adenosine A2A, but not A1, receptors mediate the arousal effect of caffeine. Nat Neurosci. 2005; 8(7):858–9.CrossRefPubMed
18.
Zurück zum Zitat Chen L, Yin D, Wang TX, Guo W, Dong H, Xu Q, et al. Basal forebrain cholinergic neurons primarily contribute to inhibition of electroencephalogram delta activity, rather than inducing behavioral wakefulness in mice. Neuropsychopharmacology. 2016; 41(8):2133–46.CrossRefPubMedPubMedCentral Chen L, Yin D, Wang TX, Guo W, Dong H, Xu Q, et al. Basal forebrain cholinergic neurons primarily contribute to inhibition of electroencephalogram delta activity, rather than inducing behavioral wakefulness in mice. Neuropsychopharmacology. 2016; 41(8):2133–46.CrossRefPubMedPubMedCentral
19.
Zurück zum Zitat Anaclet C, Ferrari L, Arrigoni E, Bass CE, Saper CB, Lu J, et al. The GABAergic parafacial zone is a medullary slow wave sleep-promoting center. Nat Neurosci. 2014; 17(9):1217–24.CrossRefPubMedPubMedCentral Anaclet C, Ferrari L, Arrigoni E, Bass CE, Saper CB, Lu J, et al. The GABAergic parafacial zone is a medullary slow wave sleep-promoting center. Nat Neurosci. 2014; 17(9):1217–24.CrossRefPubMedPubMedCentral
20.
Zurück zum Zitat Tsai C. The optic tracts and centers of the opossum, Didelphis virginiana. J Comp Neurol. 1925;39(2):173–216.CrossRef Tsai C. The optic tracts and centers of the opossum, Didelphis virginiana. J Comp Neurol. 1925;39(2):173–216.CrossRef
21.
Zurück zum Zitat Schultz W. Predictive reward signal of dopamine neurons. J Neurophysiol. 1998; 80(1):1–27.PubMed Schultz W. Predictive reward signal of dopamine neurons. J Neurophysiol. 1998; 80(1):1–27.PubMed
22.
Zurück zum Zitat Jones BE, Bobillier P, Pin C, Jouvet M. The effect of lesions of catecholamine-containing neurons upon monoamine content of the brain and EEG and behavioral waking in the cat. Brain Res. 1973; 58(1):157–177.CrossRefPubMed Jones BE, Bobillier P, Pin C, Jouvet M. The effect of lesions of catecholamine-containing neurons upon monoamine content of the brain and EEG and behavioral waking in the cat. Brain Res. 1973; 58(1):157–177.CrossRefPubMed
23.
Zurück zum Zitat Wisor JP, Nishino S, Sora I, Uhl GH, Mignot E, Edgar DM. Dopaminergic role in stimulant-induced wakefulness. J Neurosci. 2001; 21(5):1787–1794.PubMed Wisor JP, Nishino S, Sora I, Uhl GH, Mignot E, Edgar DM. Dopaminergic role in stimulant-induced wakefulness. J Neurosci. 2001; 21(5):1787–1794.PubMed
24.
Zurück zum Zitat McCown TJ. Adeno-associated virus (AAV) vectors in the CNS. Curr Gene Ther. 2011; 11(3):181–8.CrossRefPubMed McCown TJ. Adeno-associated virus (AAV) vectors in the CNS. Curr Gene Ther. 2011; 11(3):181–8.CrossRefPubMed
25.
Zurück zum Zitat Vyazovskiy VV, Cirelli C, Pfister-Genskow M, Faraguna U, Tononi G. Molecular and electrophysiological evidence for net synaptic potentiation in wake and depression in sleep. Nat Neurosci. 2008; 11(2):200–8.CrossRefPubMed Vyazovskiy VV, Cirelli C, Pfister-Genskow M, Faraguna U, Tononi G. Molecular and electrophysiological evidence for net synaptic potentiation in wake and depression in sleep. Nat Neurosci. 2008; 11(2):200–8.CrossRefPubMed
26.
Zurück zum Zitat Eban-Rothschild A, Rothschild G, Giardino WJ, Jones JR, de Lecea L. VTA dopaminergic neurons regulate ethologically relevant sleep-wake behaviors. Nat Neurosci. 2016; 19(10):1356–66.CrossRefPubMedPubMedCentral Eban-Rothschild A, Rothschild G, Giardino WJ, Jones JR, de Lecea L. VTA dopaminergic neurons regulate ethologically relevant sleep-wake behaviors. Nat Neurosci. 2016; 19(10):1356–66.CrossRefPubMedPubMedCentral
27.
Zurück zum Zitat Taylor SR, Badurek S, Dileone RJ, Nashmi R, Minichiello L, Picciotto MR. GABAergic and glutamatergic efferents of the mouse ventral tegmental area. J Comp Neurol. 2014; 522(14):3308–3334.CrossRefPubMedPubMedCentral Taylor SR, Badurek S, Dileone RJ, Nashmi R, Minichiello L, Picciotto MR. GABAergic and glutamatergic efferents of the mouse ventral tegmental area. J Comp Neurol. 2014; 522(14):3308–3334.CrossRefPubMedPubMedCentral
28.
Zurück zum Zitat Omelchenko N, Sesack SR. Ultrastructural analysis of local collaterals of rat ventral tegmental area neurons: GABA phenotype and synapses onto dopamine and GABA cells. Synapse. 2009; 63(10):895–906.CrossRefPubMedPubMedCentral Omelchenko N, Sesack SR. Ultrastructural analysis of local collaterals of rat ventral tegmental area neurons: GABA phenotype and synapses onto dopamine and GABA cells. Synapse. 2009; 63(10):895–906.CrossRefPubMedPubMedCentral
29.
Zurück zum Zitat Dobi A, Margolis EB, Wang HL, Harvey BK, Morales M. Glutamatergic and nonglutamatergic neurons of the ventral tegmental area establish local synaptic contacts with dopaminergic and nondopaminergic neurons. J Neurosci. 2010; 30(1):218–229.CrossRefPubMedPubMedCentral Dobi A, Margolis EB, Wang HL, Harvey BK, Morales M. Glutamatergic and nonglutamatergic neurons of the ventral tegmental area establish local synaptic contacts with dopaminergic and nondopaminergic neurons. J Neurosci. 2010; 30(1):218–229.CrossRefPubMedPubMedCentral
30.
Zurück zum Zitat Carr DB, Sesack SR. GABA-containing neurons in the rat ventral tegmental area project to the prefrontal cortex. Synapse. 2000; 38(2):114–23.CrossRefPubMed Carr DB, Sesack SR. GABA-containing neurons in the rat ventral tegmental area project to the prefrontal cortex. Synapse. 2000; 38(2):114–23.CrossRefPubMed
31.
Zurück zum Zitat Yamaguchi T, Wang HL, Li X, Ng TH, Morales M. Mesocorticolimbic glutamatergic pathway. J Neurosci. 2011; 31(23):8476–8490.CrossRefPubMed Yamaguchi T, Wang HL, Li X, Ng TH, Morales M. Mesocorticolimbic glutamatergic pathway. J Neurosci. 2011; 31(23):8476–8490.CrossRefPubMed
32.
Zurück zum Zitat Kim Y, Laposky AD, Bergmann BM, Turek FW. Repeated sleep restriction in rats leads to homeostatic and allostatic responses during recovery sleep. Proc Natl Acad Sci USA. 2007; 104(25):10697–10702.CrossRefPubMedPubMedCentral Kim Y, Laposky AD, Bergmann BM, Turek FW. Repeated sleep restriction in rats leads to homeostatic and allostatic responses during recovery sleep. Proc Natl Acad Sci USA. 2007; 104(25):10697–10702.CrossRefPubMedPubMedCentral
33.
Zurück zum Zitat Deboer T. Behavioral and electrophysiological correlates of sleep and sleep homeostasis. Curr Top Behav Neurosci. 2015;25:1–24.PubMed Deboer T. Behavioral and electrophysiological correlates of sleep and sleep homeostasis. Curr Top Behav Neurosci. 2015;25:1–24.PubMed
34.
Zurück zum Zitat Qiu MH, Liu W, Qu WM, Urade Y, Lu J, Huang ZL. The role of nucleus accumbens core/shell in sleep-wake regulation and their involvement in modafinil-induced arousal. PLoS One. 2012;7(9):e45471.CrossRefPubMedPubMedCentral Qiu MH, Liu W, Qu WM, Urade Y, Lu J, Huang ZL. The role of nucleus accumbens core/shell in sleep-wake regulation and their involvement in modafinil-induced arousal. PLoS One. 2012;7(9):e45471.CrossRefPubMedPubMedCentral
35.
Zurück zum Zitat Yang SR, Sun H, Huang ZL, Yao MH, Qu WM. Repeated sleep restriction in adolescent rats altered sleep patterns and impaired spatial learning/memory ability. Sleep. 2012; 35(6):849–859.CrossRefPubMedPubMedCentral Yang SR, Sun H, Huang ZL, Yao MH, Qu WM. Repeated sleep restriction in adolescent rats altered sleep patterns and impaired spatial learning/memory ability. Sleep. 2012; 35(6):849–859.CrossRefPubMedPubMedCentral
36.
Zurück zum Zitat Qiu MH, Chen MC, Fuller PM, Lu J. Stimulation of the pontine parabrachial nucleus promotes wakefulness via extra-thalamic forebrain circuit nodes. Curr Biol. 2016; 26(17):2301–2312.CrossRefPubMed Qiu MH, Chen MC, Fuller PM, Lu J. Stimulation of the pontine parabrachial nucleus promotes wakefulness via extra-thalamic forebrain circuit nodes. Curr Biol. 2016; 26(17):2301–2312.CrossRefPubMed
Metadaten
Titel
Activation of the ventral tegmental area increased wakefulness in mice
verfasst von
Huan-Xin Sun
Dian-Ru Wang
Chen-Bo Ye
Zhen-Zhen Hu
Chen-Yao Wang
Zhi-Li Huang
Su-Rong Yang
Publikationsdatum
01.04.2017
Verlag
Springer Japan
Erschienen in
Sleep and Biological Rhythms / Ausgabe 2/2017
Print ISSN: 1446-9235
Elektronische ISSN: 1479-8425
DOI
https://doi.org/10.1007/s41105-017-0094-x

Weitere Artikel der Ausgabe 2/2017

Sleep and Biological Rhythms 2/2017 Zur Ausgabe

Leitlinien kompakt für die Neurologie

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Stuhltransfusion könnte Fortschreiten von Parkinson-Symptomen bremsen

03.05.2024 Parkinson-Krankheit Nachrichten

Kann eine frühzeitige Stuhltransplantation das Fortschreiten von Parkinson-Symptomen verlangsamen? Die Ergebnisse einer randomisierten Phase-2-Studie scheinen dafür zu sprechen.

Frühe Tranexamsäure-Therapie nützt wenig bei Hirnblutungen

02.05.2024 Hirnblutung Nachrichten

Erhalten Personen mit einer spontanen Hirnblutung innerhalb von zwei Stunden nach Symptombeginn eine Tranexamsäure-Therapie, kann dies weder die Hämatomexpansion eindämmen noch die Mortalität senken.

Sind Frauen die fähigeren Ärzte?

30.04.2024 Gendermedizin Nachrichten

Patienten, die von Ärztinnen behandelt werden, dürfen offenbar auf bessere Therapieergebnisse hoffen als Patienten von Ärzten. Besonders scheint das auf weibliche Kranke zuzutreffen, wie eine Studie zeigt.

Akuter Schwindel: Wann lohnt sich eine MRT?

28.04.2024 Schwindel Nachrichten

Akuter Schwindel stellt oft eine diagnostische Herausforderung dar. Wie nützlich dabei eine MRT ist, hat eine Studie aus Finnland untersucht. Immerhin einer von sechs Patienten wurde mit akutem ischämischem Schlaganfall diagnostiziert.

Update Neurologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.