Skip to main content

Sex-dependent neuronal effects of α-synuclein reveal that GABAergic transmission is neuroprotective of sleep-controlling neurons

Abstract

Background

Sleep disorders (SDs) are a symptom of the prodromal phase of neurodegenerative disorders that are mechanistically linked to the protein α-synuclein (α-syn) including Parkinson’s disease (PD). SDs during the prodromal phase could result from neurodegeneration induced in state-controlling neurons by accumulation of α-syn predominant early in the disease, and consistent with this, we reported the monomeric form of α-syn (monomeric α-syn; α-synM) caused cell death in the laterodorsal tegmental nucleus (LDT), which controls arousal as well as the sleep and wakefulness state. However, we only examined the male LDT, and since sex is considered a risk factor for the development of α-syn-related diseases including prodromal SDs, the possibility exists of sex-based differences in α-synM effects. Accordingly, we examined the hypothesis that α-synM exerts differential effects on membrane excitability, intracellular calcium, and cell viability in the LDT of females compared to males.

Methods

Patch clamp electrophysiology, bulk load calcium imaging, and cell death histochemistry were used in LDT brain slices to monitor responses to α-synM and effects of GABA receptor acting agents.

Results

Consistent with our hypothesis, we found differing effects of α-synM on female LDT neurons when compared to male. In females, α-synM induced a decrease in membrane excitability and heightened reductions in intracellular calcium, which were reliant on functional inhibitory acid transmission, as well as decreased the amplitude and frequency of spontaneous excitatory postsynaptic currents (sEPSCs) with a concurrent reduction in action potential firing rate. Cell viability studies showed higher α-synM-mediated neurodegeneration in males compared to females that depended on inhibitory amino acid transmission. Further, presence of GABA receptor agonists was associated with reduced cell death in males.

Conclusions

When taken together, we conclude that α-synM induces a sex-dependent effect on LDT neurons involving a GABA receptor-mediated mechanism that is neuroprotective. Understanding the potential sex differences in neurodegenerative processes, especially those occurring early in the disease, could enable implementation of sex-based strategies to identify prodromal PD cases, and promote efforts to illuminate new directions for tailored treatment and management of PD.

Background

Parkinson’s disease (PD) is among one of the most widespread neurodegenerative disorders [1, 2], and sex is a risk-factor in the development of this disease, as PD is more common in men than in women with an approximated odds ratio of 2:1 [3,4,5,6,7]. Although PD is clinically diagnosed by the cardinal motor symptoms, evidence emerging over the past two decades has established sleep disorders (SDs) such as REM sleep behavior disorder (RBD), which is a sleeping disorder characterized by excessive motor behavior during what is normally a period of atonia, and excessive daytime sleepiness (EDS) as markers of the prodromal phase of PD, and these SDs can precede the motor symptoms by years to decades [8,9,10,11,12,13]. Sex has been acknowledged as an important determinant of both the susceptibility to neurodegenerative diseases and whether SDs co-occur following diagnosis, and while not well studied, it is likely that sex differences are also present in the prodromal phase of the disease prior to diagnosis, which could include the expression of SDs.

We hypothesized that the appearance of SDs prodromal to the motor symptoms in PD could be due to alteration of cellular function and neurodegeneration in sleep controlling nuclei and that sex differences in cellular effects could be present. Neurodegeneration in several brain nuclei in PD is associated with aggregation of the protein α-synuclein (α-syn), which is the histological hallmark of PD [14]. Pathological studies have shown that aggregated α-syn can interfere with several cellular functions in addition to promoting cell toxicity [15]. Heightened cell death has been reported in patients with α-synucleinopathies in the laterodorsal tegmentum (LDT) [16], which is a heterogenous nucleus comprised of cholinergic, glutamatergic and GABAergic neurons [17] that are importantly involved in the control of motor atonia during sleep, and arousal during wakefulness [18, 19]. Further, the co-occurrence of RBD and α-syn pathology was strongly correlated with brainstem cholinergic dysfunction in a predominantly male cohort consistent with α-syn-mediated degeneration of LDT cholinergic systems underlying aberrant behavioral state behavior [16]. Taken together, this suggests that α-syn-mediated PD processes include cellular actions in SD-controlling neurons already from abnormal levels of α-syn.

Although several studies have investigated the neuronal effects induced by α-syn, their focus was on actions of forms of the protein which aggregate (oligomeric and fibril). These forms are suspected to be the most damaging to neurons and neural transmission, and so very few studies have reported on cellular effects induced by the disordered, native monomeric form (α-synM), which is widely considered to be relatively benign. The aggregation and fibrillation of α-syn may be preceded by dysregulation of expression e.g. as a result of SNCA gene multiplication [20]) and/or clearance [21] leading to abnormal levels of α-syn in the CNS, which then in turn may lead to nucleation and fibril formation [22]. We recently found that α-synM has cellular effects on neurons in the LDT that were associated with heightened cell death, which we speculated could play a role in SDs prodromal to, as well as following, diagnosis of PD. Specifically, we found that α-syn in monomeric form induced excitation, increased intracellular calcium, and heightened neuronal death of neurons of the LDT. In contrast,  different effects were induced in the substantia nigra (SN) in that α-synM elicited membrane inhibition, and greater decreases of intracellular calcium with no evidence of α-synM-induced cell death [23]. However, that investigation was conducted solely in LDT of males. Because the appearance of sex-based differences in α-syn-related disease symptoms, including differences between SDs, suggest different mechanistic actions are involved in disease processes in males and females, we wished to determine whether excitatory cellular effects of α-synM on LDT, and heightened cell death in males were also present in females. Accordingly, in the present report, we have used electrophysiological and calcium imaging techniques ex vivo to investigate cellular effects of highly purified α-synM on LDT and SN neurons from females and compared effects to those in male LDT neurons.

Methods

Animals

The protocols for animal experiments used in this study were approved in concordance with the European Communities Council Directive (86/609/EEC). Brain slices of 250 μm thickness from female and male Naval Medical Research Institute (NMRI) mice aged 12 to 30 days (Harlan Mice Laboratories, Denmark) were used in electrophysiological, calcium imaging and cell viability studies. The animals were housed under the following conditions: temperature (22–23 °C), humidity (45–65%), light-dark cycle 12:12 h, water and food were available ad libitum.

Brain slice preparations

A state of anesthesia was induced via inhalation of isoflurane (Baxter A/S, Denmark) and decapitation was conducted when anesthesia had been achieved as assayed by failure to react to a paw pinch. A block of the brain which contained LDT or SN was rapidly removed and submerged in ice-cold artificial cerebrospinal fluid (ACSF). The ACSF solution which contained 124 NaCl, 5 KCl, 1.2 Na2HPO4•2H2O, 2.7 CaCl2•2H2O, 1.2 MgSO4 (anhydrous), 10 dextrose, 26 NaHCO3 in mM was adjusted to a pH of 7.4 and an osmolarity of 298–302 mOsm/kg following saturation with carbogen (95% O2/5% CO2). The brain was sectioned in 250 μm thick slices containing the LDT or the SN with a vibratome (Leica VT1200S, Leica Biosystems, Germany). Brain slices were collected and placed in a chamber containing oxygenated ACSF, and incubated at 37 oC for 15 min. To allow the tissue to equilibrate after the incubation period, the slices were kept at room temperature, and carbogen was continuously supplied for at least 1 h prior to further procedures, including exposure of the slice to the monomeric form of α-syn.

Monomeric α-syn

Human α-syn was recombinantly expressed and purified as described in more detail in our earlier published work using this peptide [23]. Briefly, α-syn was cloned into E. Coli BL21DE3 cells using a pET-11a vector construct. Harvested cells were lysed by osmotic shock. Subsequently, boiling and centrifugation were conducted to remove non-heat-stable proteins. Ion-exchange chromatography was used to isolate α-syn, and the monomeric fraction was isolated by size exclusion chromatography SEC; thereafter, the monomers were pooled and kept in PBS buffer stored at -80oC until application to the slices.

α-syn application

The monomeric form of α-syn (α-synM) was stored in solution at -20oC in aliquots of 10 µl (150 µM) until use at which time it was applied via the bath. To reach a final concentration of 100 nM, an aliquot (150 µM) of 10 µl of α-synM was diluted in ACSF. After the establishment of baseline holding currents or baseline fluorescence, α-synM was applied for 3–4 min to monitor effects on membrane holding currents, synaptic activity, action potential firing, and intracellular calcium. For cell viability studies, incubation of the 250 μm brain slice in α-synM for 7 h was conducted in protocols described below.

Drugs

Action potentials within the slice were blocked by 0.5 mM tetrodotoxin (TTX, Tocris, UK). Glycinergic receptors were blocked with strychnine (2.5 µM; Sigma, Denmark). GABAA and GABAB receptor-mediated responses were blocked by SR-95,531 (gabazine, 10 µM, Sigma, Demark) and CGP 55,845 (10 µM, Tocris, UK), respectively. Muscimol (30 µM, Sigma, Denmark) and baclofen (10 µM, Sigma, Denmark) were used as agonists of the GABAA and GABAB receptors, respectively. Stock solutions were stored in appropriate aliquots at -20oC prior and diluted in ACSF to final concentrations before use, and all drugs were applied via the bath.

Patch-clamp recordings to monitor changes in membrane currents, synaptic activity, and action potential firing

Whole cell patch clamp recordings were conducted in 250 μm thick brain slices from neurons in the LDT (19 LDT brain slices; 14 mice) or the SN (7 SN brain slices; 6 mice). For recordings done in the LDT, we wished to target cholinergic neurons. Therefore, we selected the recorded neurons based on soma size (medium-to-large cells) and location within the central LDT wherein the concentration of cholinergic neurons is highest [24]. To fabricate patch pipette electrodes for recording neuronal electrical activity in LDT and SN brain slices, borosilicate filamented glass capillary tubes (1.5 mm, Sutter Instruments, USA) were pulled after heating to a sharp tip in a horizontal Flaming/Brown micropipette puller (P-97, Sutter Instruments, USA). These glass electrodes were filled with an intracellular solution (144 K-gluconate; 2 KCl; 10 HEPES; 0.2 EGTA; 5 Mg-ATP and 0.3 Na-GTP; in mM), which resulted in a pipette resistance of 6–11 MΩ. A brain slice containing LDT or SN was placed in the recording chamber that was situated in a microscope stage and ACSF saturated with a mixture of 95% oxygen/5% carbon dioxide (carbogen) was continuously perfused over the slice (flow rate 1.2 mL/min). A water immersion objective (60x) coupled to an upright microscope (BX50WI, Olympus; Japan) with an infrared Dodt gradient contrast system (IR-DGC; Luigs & Neumann, Germany) and a CCD camera (CCD-300ETRC; DAGE-MTI, Michigan City, IN) were used to visualize the cells. The software, Patchmaster (HEKA; version v2 × 91), was used to control a patch-clamp EPC9 amplifier (HEKA, Germany). Recordings were initiated in voltage-clamp mode to establish high resistance seals (> 1 MΩ) between the patch pipette and the cell membrane, and the holding voltage was kept at -60 mV. After at least a stabilization period of 7 min following membrane breakthrough, data were collected. AxoScope 10.2 (Molecular Devices Corporation, USA) and an Axon miniDigi 1B digitizer (Molecular Devices Corporation) were used to sample membrane effects. To quantify relative changes in holding currents, the holding currents in pA averaged from at least 30 sec of recording before application of α-synM and the holding currents averaged from at least 30 sec during the maximum effect of α-synM were subtracted. A change in amplitude of 2 pA from baseline was used as criteria to be considered a response. For firing frequency studies, action potentials were recorded in current-clamp mode before and after α-synM application. Current was applied to depolarize the neuron sufficiently to induce a sustained firing of action potentials (-45 mV). A period of firing in an epoch of 30 sec immediately prior to and 30 sec after application of α-synM was selected and interspike intervals were measured and averaged. Interspike intervals were determined by measuring the period of time (in msec) from the initiation of a spike to initiation of the next spike.

Calcium imaging to monitor changes in intracellular calcium

Intracellular loading of cells with a calcium binding dye was performed in 42 LDTF (25 mice) and 15 LDTM (7 mice) 250 μm thick brain slices following standard protocols [25] for single-photon calcium imaging. Recordings were conducted utilizing ratiometric fluorescent calcium indicator dye Fura-2 acetoxymethyl ester (Fura-2 AM). Prior to recordings, slices were rinsed for over 15 min in the recording chamber by continuous perfusion of oxygenated ACSF at a flow rate of 1.2 ml/min to wash out free dye debris and allow temperature equilibration. To localize the LDT, an upright microscope (BX50WI, Olympus, Germany) was used under bright field illumination and visualization guided by characteristic landmarks located close to these nuclei. Individual cells were viewed with a water immersion objective (40x). A cooled CCD fluorescence camera (12-bit Sensicam, PCO Imaging, Germany) attached to the microscope and controlled by the imaging software Live Acquisition (TILL Photonics, Germany) was used to collect paired images that were binned at 2 × 2 pixels on the camera chip in order to optimize spatial and temporal resolution. Live Acquisition controlled rapid switching between the excitation wavelengths of 340 and 380 nm in order to ensure a minimal amount of time passing between collection of the image at each wavelength, which allowed ratiometric calculations. The acquisition interval between each frame pair of 340 and 380 nm was 4 sec. Regions of interest (ROI) were drawn around each Fura-2 loaded cell, as well as around a region of the field of view that contained no dye loaded cells, which was used to measure background fluorescence. Analysis software (Offline Analysis, TILL Photonics, Germany) was used to quantify fluorescence in each ROI by averaging the intensity of the 2 × 2 binned pixels, and background in each channel was subtracted. The 340 and 380 nm channels were ratioed (340 nm:380 nm). Baseline fluorescence was calculated from an average of intensity taken from 10 frames collected before drug application. Following application of α-synM, changes in fluorescence from that measured at baseline were noted in the majority of cells, and the fluorescence of the peak deflection from baseline was measured by averaging across 10 frames. Data are presented as DF/F% which represents the subtraction of the baseline fluorescence from the maximum change in fluorescence induced by α-synM application divided by the baseline, followed by conversion to a percentage. Ascendent deflection of fluorescence indicates intracellular calcium elevations with descendent deflections reflecting decreases in intracellular calcium. Actual calcium levels were not calculated due to well-known complications with converting changes in fluorescence in brain slices to calcium concentrations [26, 27].

Neurotoxicity assay to evaluate cell viability

Propidium iodide (PI; Sigma-Aldrich) was used to identify dead cells and 4′,6-diamidino-2-phenylindole (DAPI; Sigma-Aldrich) was used to mark live cells from LDTF and LDTM slices. A total of 116 slices from 56 mice were used in 3 different protocols in which 250 μm slices were bisected with each half being exposed for 7 h under oxygenation to: (1) ACSF or α-synM (100 nM), (2) α-synM (100 nM) or α-synM (100 nM) plus GABAA, GABAB and glycine receptor antagonists, or (3) α-synM (100 nM) or α-synM (100 nM) plus GABAA, GABAB, and glycine receptor agonists. The bisection of the slices ensured that data sets could be compared from tissue taken from the same group of animals, and that protocols were run side by side under the same laboratory conditions. Following incubation, slices were fixed in 4% paraformaldehyde overnight, cryprotected with sucrose saturation (30%), and resectioned to a thickness of 40 μm on a cryostat (Leica CM3050, Triolab, DK). Resectioned slices of 40 μm were incubated for 3 periods of 5 min in a solution which contained 1 µg/ml of both PI and DAPI with a pH of 7.4. To detect fluorescence signals from PI and DAPI, an upright Zeiss microscope coupled to a monochrome CCD camera (Axiocam MRM, Zeiss, Germany) controlled by Axioskop 2 software (AxioVision 4.6, Zeiss) and required filter cubes were used (Zeiss 59 fluorescent filter cube sets, wavelengths PI: 472–578 nm; DAPI: 358–463 nm).

To conduct analysis of collected images, a macro written for ImageJ (National Institutes of Health, Bethesda, MD) was used to automatically count the number of DAPI and PI-labeled cells following background subtraction, application of thresholding, and separation of objects via a watershed algorithm. Cells were selected using the batch processing macro, Analyze Particle; however, selections were manually confirmed. Cell survival was quantified by the proportion of live cells (DAPI positive cells) to the total cell count, which was calculated as the addition of PI positive cells and DAPI positive cells. Cell viability was evaluated in up to 5 areas within the LDT in the 40 μm resectioned slices. For this data set, na = the number of areas examined/n40 = the number of resectioned slices. Data was normalized in each of the 3 protocols to the control condition, which was either exposure to ACSF or α-synM without GABAA, GABAB, and glycine receptor agonists or antagonists. However, to compare cell death between LDTF and LDTM, cell survival was evaluated without normalization. In figure panels depicting PI positive and DAPI positive cells, contrast has been applied equally across entire images.

Data analysis and statistics

Amplitudes of membrane holding currents were measured (the difference between baseline and maximum deflection) using Axoscope 10.5 (Molecular Devices, USA). Spontaneous excitatory synaptic events (sEPSCs) were detected and analyzed using MiniAnalysis (Synaptosoft, USA). Analysis of synaptic events was conducted by selecting 30 sec of the recording just before application of α-synM and at the peak amplitude of the effect on membrane current. Inter-event intervals and the amplitude of events were averaged across a population of cells and statistically analyzed. Firing frequency was analyzed selecting 30 sec epochs before and after α-synM application, and intervals between action potentials were measured and averaged. Calcium imaging data including the amplitudes of DF/F% changes were analyzed in Graphpad Prism (version 7.0). The numbers of observations in the data sets analyzed for cell viability are reflected as na. Results are presented as mean values ± SEM. The figures were prepared using Igor Pro software (Wavemetrics, USA) and GraphPad Prism. Differences in numerical data were tested using a Paired or Unpaired Student’s T-test, and differences in categorical data were examined using the Fisher’s Exact test. P values are reported in text as 4 decimal points, and a significant difference was determined if alpha was less than 0.05.

Results

α-syn M effects on membrane currents, and synaptic transmission in neurons of sleep and motor controlling nuclei in the female

α-synM induced outward currents in sleep and motor control nuclei in the female LDT

Our previous study showed that α-synM induced an inward current in LDT neurons in brain slices from male mice (LDTM) [23]. Unexpectedly, in neurons in LDT brain slices from female mice (LDTF; Fig. 1A), α-synM (100 nM, 3 min) induced an outward membrane current in all cells examined (amplitude: 54.8 ± 11.3 pA, n = 14). To compare the effect of α-synM in neurons of a sleep controlling nucleus vs. neurons of a motor control nucleus, we next investigated the neuronal effect of α-synM in the SN of females (SNF). Similar to our previous report that showed α-synM induced inhibitory membrane responses in neurons of the male SN (SNM), we observed that α-synM induced inhibitory responses in the membrane of 100% of the neurons examined in SNF (amplitude: 47.2 ± 16.0 pA, n = 11). The average amplitude of the outward current induced in neurons of the LDTF did not differ from that induced in SNF neurons (p = 0.3847; Unpaired Student’s T-test; Fig. 1B).

Fig. 1
figure 1

α-synM induced an inhibitory outward current and modulated synaptic transmission in LDT neurons in the female. A) (Left) Cartoon schematic of the sagittal mouse brain to show the block of the brain containing the LDT. A LDT coronal brain slice taken from this block is shown below in inset. (A) (Right) Coronal brain slice cartoon modified from [72] to show in greater detail in insets to the sides the location of the LDT (indicated by white arrow in panel to the left). (B) Sample of membrane response to α-synM, which induced inhibitory, outward currents in the female in the LDT (B1). An outward inhibitory current was also elicited in SN neurons (B2). (B) Graphs of holding currents before and after application of α-synM to LDTF and SNF neurons showed a significant increase in positive holding current indicating that α-synM induced outward currents. The amplitude of the outward current evoked by α-synM in LDTF and SNF neurons was not different (LDTF: n = 14, SNF: n = 11; p = 0.3847; Unpaired Student’s T-test) as shown by the plots of the individual amplitude of current induced in both nuclei. Bar chart showing that the proportion of recorded cells responding to α-synM with induction of outward current did not differ significantly between the LDTF and SNF (LDTF: n = 14 sampled/14 responded, SNF: n = 11 sampled/11 responded; p = 1.000; Fisher’s Exact Test). (C) α-synM modulated synaptic events in neurons recorded within LDTF and SNF. (C1- C2) samples of recordings showing frequency of synaptic events in control and in presence of α-synM in both LDTF and SNF. (Rightmost panels) Single sEPSCs (spontaneous excitatory postsynaptic currents) in a LDTF and in a SNF neuron are shown with a high-gain time and amplitude scale under control conditions and in presence of α-synM illustrating the reduction in amplitude in both nuclei when α-synM was present. Data presented in paired plots summarize findings from the population of recorded cells, which revealed that α-synM induced a significant decrease in amplitude of EPSCs in LDTF (n = 5; p = 0.0253; Paired T-test) and SNF neurons (n = 4; p = 0.0483; Paired T-test) and elicited a significant decrease in the frequency of sEPSCs in LDTF neurons (n = 5; p = 0.0475; Paired T-test), which was a change not seen in sEPSCs in the SN (n = 4; p = 0.4735; Paired T-test). LDT: Laterodorsal tegmental nucleus; 4 V: 4th ventricle; IC: Inferior colliculus; DTgP: Dorsal tegmental nucleus, pericentral: DRN: dorsal raphe nucleus; LC: Locus coeruleus; LDTF: Laterodorsal tegmental nucleus of female; SNF: Substantia nigra of female. * Indicates p < 0.05, *** Indicates p < 0.001

α-synM alters synaptic events in sleep and motor control nuclei in the female LDT

Our previous study showed that α-synM altered synaptic transmission in LDTM neurons producing increases in frequency as well as amplitude of spontaneous excitatory postsynaptic currents (sEPSCs). However, in the present study, the opposite effect was seen in LDTF as α-synM induced a significant decrease of nearly 20% in sEPSC frequency (Control: 7.2 ± 2.7 Hz; α-synM: 5.8 ± 2.3 Hz; p = 0.0475; n = 5, Paired T-test) and a 21% decrease in the amplitude of sEPSCs was noted which was significant (control: 10.8 ± 1.7 pA; α-synM: 8.5 ± 1.3 pA; n = 5; p = 0.0253; Paired T-test; Fig. 1C).

In SNF, α-synM induced a decrease of 14% in amplitude of sEPSCs which was significant when compared to control (control: 6.1 ± 0.8pA; α-synM: 5.2 ± 0.5pA; n = 4; p = 0.0483; Paired T-test); however, changes induced in the frequency were not significantly different (control: 5.9 ± 1.9 Hz; α-synM: 7.4 ± 2.8 Hz; n = 4; p = 0.4735; Paired T-test; Fig. 1C). In our previous work, we did not examine the effect of α-synM on sEPSCs in the SN in the male (SNM). Therefore, in order to examine sex-based potential differences of α-synM on sEPSCs in this motor control nucleus, we determined whether α-synM had effects on synaptic activity in the SNM and found that changes in sEPSCs were qualitatively similar to those seen in SNF. α-synM induced a decrease of 10% in the amplitude of the current of the sEPSCs (Ctrl: 6.0 ± 1.7 pA; α-synM: 5.4 ± 1.6 pA; n = 4; p = 0.0449; Paired T-test), and changes induced in the frequency were not significantly different (Ctrl: 11.8 ± 2.2 Hz; α-synM: 11.7 ± 2.4 Hz; n = 4; p = 0.9162; Paired T-test). In summary, these data indicate that α-synM induced inhibitory membrane effects on neurons of LDTF, and a reduction in the amplitude and frequency of sEPSCs, which were opposite effects from those we have reported before in LDTM [23]. In SNF, α-synM also induced inhibitory effects on the membrane, which were similar to those effects we have reported before in SNM. Further, α-synM had similar effects on synaptic events in SNM to those seen in SNF as in both sexes we observed a reduction in the amplitude of sEPSCs with no effect on frequency. Taken together, our findings show that the examined effects of α-synM on LDT neurons are sex-dependent, whereas, α-synM effects on SN are independent of sex.

Sex differences in LDT neurons of α-synM-mediated alteration of intracellular calcium

As α-synM-induced neuronal effects have been hypothesized to lead to calcium dysregulation [23, 28], we previously examined actions of this protein on intracellular calcium levels and found that α-synM induced changes in calcium in LDTM. We repeated those experiments here and confirmed our earlier findings. Using multiple-cell calcium imaging to monitor changes in Fura 2-AM fluorescence which were induced by α-synM (100 nM, 3 min), we observed changes in fluorescence in the majority of LDTM cells (97.4%; n = 38/39; Fig. 2B1), and the majority of responses were indicative of an increase in calcium (76.3%; n = 29/38; Fig. 2A1a, B2). We then examined responses in LDTF and found that α-synM induced changes in intracellular calcium in 100% (n = 89/89) of the examined cells (Fig. 2B1). However, interestingly, the majority of responding LDTF neurons exhibited changes in fluorescence indicative of a decrease in intracellular calcium levels (58.4%, n = 52/89; Fig. 2A2b, B2). When we compared alterations in intracellular calcium induced by α-synM in LDTM to LDTF, there was no significant difference in the numbers of responding or non-responding neurons (p = 0.3047; Fischer’s Exact test; Fig. 2B1). In contrast, there was a difference between males and females in the ratio of increases in calcium to decreases in calcium in response to α-synM. We observed a significantly higher proportion of cells responding with a decrease in calcium in LDTF when compared to the proportion of cells exhibiting decreases in LDTM (p = 0.0004; Fisher’s Exact test; Fig. 2B2).

Fig. 2
figure 2

Sample of changes in fluorescence (DF/F%) induced by α-synM, which are indicative of alterations in intracellular calcium levels in LDTM and LDTF. (A) In both sexes, changes in response of the fluorescence to α-synM exhibited two different polarities, which suggested increases (A1a, A2a) or decreases (A1b, A2b) in intracellular calcium levels, respectively. Inset in A2 is a fluorescent image under 380 nm wavelength light of one of the LDTF brain slices used in this study in which two Fura 2-AM filled cells indicated with red arrows can be seen. Regions of interest were drawn around each cell and average fluorescent intensity (F) within each region of interest was plotted against time. White scale bar indicates 20 μm. (B) Histograms summarizing the data from the population of recorded cells indicating that (B1) the frequency of responses to α-synM did not significantly differ between the two sexes (LDTM: n = 38/39, LDTF: n = 89/89; p = 0.3047; Fischer’s Exact test), (B2) whereas the distribution of response polarity differed significantly between the sexes with a greater proportion of responses suggesting decreases in calcium being elicited in females than males (LDTM: n of decreases = 9/38, LDTF: n of decreases = 52/89; p = 0.0004; Fisher’s Exact test). LDTM: Laterodorsal tegmental nucleus of male; LDTF: Laterodorsal tegmental nucleus of female. *** Indicates p < 0.001

Evaluation of the mechanism of α-syn M inhibition induced in the membrane of LDT neurons

To gain more information regarding the sex-based difference in the mechanism underlying α-synM membrane effects, we examined α-synM actions during blockade of the generation of Na+-dependent action potentials by inclusion in the bath of tetrodotoxin (TTX, 500 nM). Unexpectedly, in presence of TTX, α-synM induced an inward current (-7.7 ± 2.0 pA, n = 4), and this effect was present in all neurons tested (Fig. 3A, B).

Fig. 3
figure 3

An excitatory inward current was revealed when α-synM was applied during blockade of presynaptic transmission, or antagonism of GABAA, GABAB and glycine receptors, suggesting that α-synM induces an outward membrane current in LDTF neurons due to actions at presynaptic inhibitory neurons. A) Sample of membrane responses to α-synM in which an inward current is revealed when α-synM is applied in presence of (A1) TTX, (A2) low calcium solution, or (A3) a cocktail of SR-95,531 + CGP-55,845 + strychnine, which block GABAA, GABAB and glycine receptors, respectively). (B) Histograms from a population of cells recorded in which there were significant changes in the (B1) polarity of the evoked membrane current responses to α-synM in presence of TTX, low calcium solution or GABA and glycine receptor antagonists when compared to control responses (Fisher’s Exact test). (B2) The amplitude of the current evoked by α-synM in control conditions and under conditions of synaptic blockade and inhibitory receptor antagonists is shown revealing the change in polarity of the α-synM induced current. (B2, Inset) Graphs of holding currents before and after application of α-synM to LDTF showed that in presence of TTX, low calcium solution, and GABAA, GABAB, and glycine receptor antagonists, holding currents became significantly more negative after application of α-synM, which reflected the α-synM-mediated induction of inward currents. (C) The decrease in intracellular calcium in LDTF is mediated, at least in part, by inhibitory receptors as illustrated in this histogram showing data from the population of LDTF cells recorded that showed a significantly smaller decrease in the amplitude of intracellular calcium induced by α-synM in presence of SR-95,531, CGP-55,845 and strychnine when compared to control conditions (Control: n = 52, GABA/Gly antagonists: n = 49; p = 0.0001; Paired T-test). LDTF: Laterodorsal tegmental nucleus of female. * Indicates p < 0.05, ** Indicates p < 0.01, *** Indicates p < 0.001

Our findings with TTX indicated that α-synM-induced outward currents relied on a presynaptic mechanism. To confirm the involvement of a presynaptically-mediated mechanism in the induction of outward currents in the postsynaptic membrane of LDTF neurons, we next applied α-synM in a reduced calcium solution, which effectively eliminates calcium-dependent synaptic transmission. Because we wished to use a within cell control, we first verified whether a second application of α-synM to the same cell could result in similar effects on the membrane as those elicited in a first application. Thus, in a subset of LDTF neurons, we reapplied α-synM following a first application, and we observed that a second inhibitory response was elicited in all neurons tested, which did not significantly vary in amplitude from the outward current obtained in the first application (n = 3). Consistent with the TTX data, in all the cells tested in which the first application elicited an outward current, in the second application in presence of low calcium solution, we did not observe an induction of an outward current, and instead, an inward current was elicited (-16.1 ± 5.5 pA; n = 3; Fig. 3A2, B). Taken together, these data indicate that the inhibitory effect induced by α-synM on the membrane of LDTF involves presynaptic-dependent mechanisms.

LDT neurons receive a heavy inhibitory input from presynaptic GABAergic terminals both from local LDT neurons but also from projections sourcing from outside the nucleus [29], and, accordingly, GABAergic presynaptic mechanisms could be involved in α-synM-mediated inhibitory responses in LDTF. Therefore, we investigated α-synM-induced membrane responses in LDTF in presence of the GABA receptor antagonists, SR-95,531 (10 µM) and CGP-55,845 (10 µM), which block GABAA and GABAB receptors, respectively. Although no evidence has been presented of glycine-mediated inhibition of LDT cells, and we have not noted any glycinergically-mediated spontaneous inhibitory currents (sIPSCs) in our own studies under our recording conditions, we also included strychnine (2.5 µM) in the ACSF to ensure the blockade of any glycine-mediated events. In a population of cells in which the first application of α-synM induced an outward current, we found that in the presence of GABA and glycine receptor blockade, an inward current was elicited in all tested cells (-16.3 ± 5.2 pA; n = 3; Fig. 3A3, B1, 2).

When taken together, our data indicate that induction of outward current in LDTF neurons is reliant on inhibitory transmission from presynaptic neurons. Blockade of inhibitory transmission revealed an inward membrane current similar to what has been seen in LDTM. Although not tested in the present study, we showed in our previous work that inward currents in LDTM were mediated by a G-protein receptor coupled mechanism in postsynaptic neurons. Although we conducted experiments to examine a role for receptors previously implicated in α-synM effects, we could not identify the specific receptor involved; however, we speculate that this same excitatory mechanism is being activated in in LDTF but masked by the concurrent induction of outward current induced by α-synM-mediated stimulation of inhibitory presynaptic transmission.

Inhibitory amino acids are involved in the decrease of intracellular calcium

As we had seen that membrane current effects of α-synM involved inhibitory transmission, we examined whether similar mechanisms were also involved in the decrease of intracellular calcium seen in response to α-synM in the majority of neurons of the LDTF. During blockade of GABA and glycine receptors, while decreases in fluorescence indicative of reductions in calcium were still elicited, the amplitude of the decrease in fluorescence was significantly smaller (36%) compared to that elicited in control conditions (control: 57.1 ± 3.0% DF/F, n = 52; blockers: 36.2 ± 2.3% DF/F, n = 49; p = 0.0001; Paired T-test; Fig. 3C). Taken together, while they suggest that other mechanisms might be involved in the reductions in intracellular calcium induced by α-synM, these data provide evidence that inhibitory amino acids, most likely GABA contribute to α-synM-mediated calcium decreases in LDTF and provide further support that inhibitory transmission targeting postsynaptic LDT cells is activated by α-synM.

α-synM Reduces the excitability of LDT neurons in the female

We previously reported that α-synM enhanced the firing frequency of neurons within LDTM. However, the inhibitory effect induced by α-synM on the membrane of LDTF in conjunction with the reduction in amplitude and frequency of EPSCs could reduce neuronal excitability in the female. To directly investigate the functional effect of α-synM-mediated actions on LDTF neurons which could affect the output of these cells, we examined the firing frequency in current clamp mode following depolarization of the membrane of LDTF neurons sufficiently to induce action potentials (VM: -45.0 ± 5.0 mV) before and after application of α-synM. Under these conditions, α-synM reduced the firing frequency by 36.5% from baseline levels (control: 0.41 ± 0.05 Hz; α-synM: 0.26 ± 0.08 Hz; n = 3; p = 0.0498; Paired T-test; Fig. 4A, B). These data suggest that in direct contrast to findings in LDTM, functional actions of α-synM include reductions in neuronal excitability of neurons in LDTF, which would be expected to alter output of these cells to target regions.

Fig. 4
figure 4

α-synM induces significant changes in the firing frequency in neurons recorded within LDTF. (A) Representative examples of current-clamp recordings of LDT neurons in which action potentials were induced by holding the cell at -45 mV under control conditions (top) and in presence of α-synM. (bottom). (B) The reduction in firing frequency induced by α-synM was significant as shown in the bar graphs depicting the average firing rate from the population of recorded LDTF neurons (n = 3; p = 0.0498; Paired T-test). LDTF: Laterodorsal tegmental nucleus of female

* Indicates p < 0.05

α-syn M induces a lower cell death of LDT neurons in females compared to males

In our previous report, α-synM-mediated excitation of the membrane of neurons with a concurrent rise of intracellular calcium was suspected to induce excitotoxicity, which was supported by a heightened cell death over control in LDTM [34]. As α-synM-induced inhibitory membrane current, and increases in calcium were less prominent in LDTF neurons, we hypothesized that neurodegeneration induced by α-synM would also exhibit a sex-based difference. First, we needed to determine whether α-synM induced cell death in LDTF above control. Accordingly, we evaluated cell survival in LDTF hemi slices in which one half had been exposed for 7 h to ACSF and the other half to 7 h in α-synM. We found a relatively lower cell survival in the half of the slice exposed to α-synM when normalized to survival seen in control (Cell Survival: Control: ACSF: 100 ± 0.9%, na = 217/n40 = 57; α-synM: 92.8 ± 1.5%, na = 183/n40 = 48; Fig. 5A).

Fig. 5
figure 5

(A-C, left panels) DAPI and PI immunohistochemistry conducted in LDTF and LDTM from brain slices exposed to α-synM under 3 different treatment protocols is shown in representative fluorescent images. The first column represents living cells visualized by DAPI (blue). The second column represents dead cells visualized by PI (red), and the last column is a merged image of DAPI and PI labeled cells. (A) The presence of DAPI and PI in the LDTF following incubation of one half of a slice in control solution (ACSF) and the other half in ACSF containing α-synM for 7 h is shown and indicates relatively lower cell survival in the half of the slice exposed to α-synM, which was reflected in the population data. The bar graph to the right shows that following α-synM exposure, cell survival in LDTF was significantly greater than that seen in LDTM (Cell Survival Female: na = 183/n40 = 48, Cell Survival Male: na=168/n40 = 66; p < 0.0001; Unpaired Student’s T-test). In this and subsequent panels, red points represent observations from LDTF, blue represent data from LDTM, and cell counts within each area (na) represent one data point or observation in the bar chart columns. (B) Fluorescent images showing DAPI and PI presence in LDTF cells treated for 7 h with α-synM or with α-synM in presence of GABAA, GABAB and glycine receptors antagonists (G-ANT). As shown in the bar graph to the right, a reduced cell survival indicative of greater cell mortality was observed in the population of LDTF slices exposed to α-synM when GABA and glycine receptor antagonists were present (Cell survival: α-synM: na = 199/n40 = 44, Cell survival α-synM + G-ANT: na = 169/n40 = 35; p = 0.0001; Mann-Whitney Test). To compare the population data, bisected slices were used, and the proportion of surviving cells observed in the half of the bisected slice exposed to α-synM was considered the baseline, and the number of surviving cells in the other half of the bisected slice exposed to α-synM + G-ANT was normalized to this baseline. (C) Fluorescent images of LDTM slices exposed to ⍺-synM or to α-synM in presence of 7 h of GABAA and GABAB receptor agonists (G-AGO). As can be seen from the population data shown in bar graphs to the right, the presence of the GABA and glycine receptor agonists in the LDTM was associated with significantly greater cell survival following exposure to α-synM (Cell Survival α-synM: na = 168/n40 = 66, Cell Survival α-synM + G-AGO: na = 127 /n40 = 48; p = 0.0334; Mann-Whitney Test). In this protocol, the proportion of surviving cells observed in the half of the bisected slice exposed to α-synM was considered the baseline, and the number of surviving cells in the other half of the bisected slice exposed to α-synM + G-AGO was normalized to this baseline. LDTF: the laterodorsal tegmental nucleus of female; LDTM: the laterodorsal tegmental nucleus of male. G-ANT: contains SR-95,531 (gabazine, 10 µM), CGP 55,845 (10 µM) and strychnine (2.5 µM) to block GABAA, GABAB, and glycine receptor-mediated responses, respectively. G-AGO: contains muscimol (30 µM) and baclofen (10 µM), which are agonists of GABAA and GABAB receptors, respectively. The scale bar in all images corresponds to 50 μm under 40x magnification. Contrast has been added equally across all the images. *p < 0.05, **** p < 0.0001

Next, we compared cell survival in the halves of the LDTF exposed to α-synM for 7 h to halves of brain slices of the LDTM that had been similarly exposed to 7 h of α-synM. Supporting our hypothesis of a sex difference in α-synM cellular effects, we noted a significantly greater cell survival in the LDTF when compared to cell survival seen in the LDTM (Cell Survival: Female: 87.4 ± 0.6%, na = 183/n40 = 48; Males: 79.9 ± 0.6%, na = 168/n40 = 66, p < 0.0001; Unpaired Student’s T-test; Fig. 5A). These results indicate that α-synM induces toxic effects in neurons of the LDTF to a lesser degree than in LDTM.

Endogenous neuroprotection against α-synM induced degeneration in LDT neurons from female involved inhibitory transmission

As we showed a sex-based differential effect of α-synM on neuronal mortality in the LDT which was reminiscent of the sex-based difference in membrane excitability and changes in intracellular calcium levels induced by this protein which were affected by GABA and glycine receptor antagonists, we reasoned that the differential effect on neurodegeneration could involve inhibitory amino acid activity which was protective in the LDTF. To examine this hypothesis, we exposed LDTF cells to α-synM for 7 h in the presence of antagonists of GABAA, GABAB and glycine receptors (G-ANT) and normalized cell viability to that in the other halves of the bisected slices that were exposed only to α-synM. The relative degree of LDTF of cell survival associated with α-synM in the presence of G-ANT was significantly lower than in absence of blockers of receptors of inhibitory amino acids (Cell survival: α-synM: 100.0 ± 2.4%, na = 199/n40 = 44, α-synM + G-ANT: 88.8 ± 1.5%, na = 169/n40 = 35; p = 0.0001; Mann-Whitney Test; Fig. 5B). These results indicate that the neuroprotective effects seen in the female brain against α-synM-mediated toxicity involve functional inhibitory amino acid transmission.

Activation of GABAergic transmission in male brain induces neuroprotection against α-synM -induced neurodegeneration

Next, we hypothesized that activation of GABA receptors could be neuroprotective against α-synM-mediated toxicity in LDTM. To examine this hypothesis, we exposed LDTM neurons to α-synM for 7 h in the presence of the GABAA, and GABAB agonists, muscimol and baclofen (G-AGO), and cell death was compared to that in the other halves of the bisected slices exposed only to α-synM. Remarkably, when exposed to α-synM the degree of cell survival seen in the halves of the slice treated with GABA receptor agonists was significantly higher than that in the other halves of the bisected slices exposed only to α-synM (Cell Survival: ⍺-synM: 100.0 ± 2.3%, na = 168/n40 = 66, α-synM + G-AGO; 106.0 ± 2.7%, na = 127 /n40 = 48, p = 0.0334; Mann-Whitney Test; Fig. 5C). These results provide further support for the conclusion that in LDTF, a GABAergic-mediated mechanism protects against α-synM-induced toxicity, and, excitingly, indicate that activation of GABA mechanisms could protect neurons of the LDTM from α-synM-mediated neurodegeneration.

Discussion

We found that in contrast to what we previously saw in LDTM, effects of α-synM on the membrane of LDTF neurons were inhibitory, and we recorded decreases in excitatory synaptic events, reductions in firing rate, and relatively more decreases in intracellular calcium. Further, cell death associated with α-synM was lower in females than in males. Changes in membrane currents and synaptic excitability noted in the SN did not exhibit a sex-based difference, suggesting nucleus specificity of α-synM-mediated effects. Inhibitory membrane currents and reductions in calcium induced in LDTF were found to involve inhibitory transmission, which when blocked revealed membrane excitation similar to that seen in LDTM. Finally, consistent with a protective role of inhibitory signaling, blockade of GABAA, GABAB, and glycine neurotransmission in the LDT of the female resulted in greater cell death and activation of GABAergic receptors reduced α-synM-mediated neurodegeneration in the LDT of the male.

While the polarity of α-synM-induced membrane effects on LDTF neurons was opposite from that seen in our earlier study conducted in LDTM, when presynaptic input was blocked, an excitatory membrane response similar to that seen in LDTM neurons was revealed. This leads us to the interpretation that α-synM induces a dual effect on the membrane of neurons of the LDTF with the summation resulting in inhibition of membrane currents of the postsynaptic neuron. The mechanism underlying the occlusion of excitatory membrane actions putatively involved GABA-releasing, presynaptic neurons, although glycinergic mechanisms were not ruled out. Sex-based differences were also found in the polarity of calcium responses between male and female in that a decrease of intracellular calcium was observed in the majority of neurons of the LDTF, whereas in LDTM, the majority of responses were indicative of rises in intracellular calcium. Similar to the membrane responses, the decrease in intracellular calcium induced by α-synM in female involved inhibitory amino acids, suggesting sex-based differences in inhibitory transmission in the LDT. Cell death associated with α-synM was lower in females but increased when GABA was blocked suggesting that GABA is involved in inhibiting neurodegenerative processes. Consistent with this, the presence of GABA receptor acting agents was able to prevent neurodegeneration in the male LDT.

Besides suggesting potential targets to inhibit neurodegeneration, our data suggest presence of a GABAergic system in the female LDT, which leads to reductions in excitatory cellular effects and cell death and that this system does not function similarly in the male LDT. While not examined specifically within the LDT, the GABAergic system has shown sexual dimorphism in the brain. In both young and adult mice, expression of proteins involved in GABA synthesis and metabolism, as well as presence of GABAA receptors have shown sex-based differences [30,31,32,33,34,35]. Further, the numbers of GABAergic neurons, as well as the responsiveness to GABA-acting drugs, have been shown to be associated with sex [36, 37]. The sex-specific phenotypic and functional differences in the GABAergic system may play key roles in the differential sensitivity of the LDTM and LDTF to α-synM [31, 34, 38, 39].

Our study has several limitations. Patch clamp recordings and multiple-cell calcium imaging with Fura-AM is difficult in slices from old animals, and thus it remains unknown if the sex-dependent effect of α-synM continues across ontogeny, which is relevant to neurodegeneration which is expected to increase across age. Further, we did not identify the phenotype of cells that were protected in presence of GABAA, and GABAB receptor agonists. Since loss of cholinergic cells in the LDT and the neighboring pedunculopontine tegmentum has been one neuropathological feature noted in α-syn-related diseases [40, 41], we tried to target large neurons with the cholinergic profile [24], however, while we do not believe we recorded from many, if any, GABA cells as they are much smaller [17, 24], non-cholinergic neurons could have been included. Nevertheless, while we did not identify the LDTF cell phenotypes exhibiting inhibitory membrane responses to α-synM, we did show in our earlier work that excitatory cellular responses in the LDTM were transmitter phenotype-independent [42]. Finally, while we did examine a nucleus which was central in SDs, given the global nature of sleep, it is almost certain that networks of nuclei, and not just activity in one nucleus mediate aberrant sleep behaviors seen in neurodegenerative diseases. Accordingly, future studies should examine effects across a larger age span, identify cellular phenotype, and conduct recordings across multiple sleep-controlling nuclei.

Despite the limitations, our work is based on multiple strengths, which differ from other investigations. Most studies of cellular effects of α-syn have been conducted using the oligomeric form of α-syn, and thus our data provides important information about effects of the monomeric form. Further, while the focus of much work has been on targeting α-syn intracellular exposure, we have utilized extracellular exposure. Additionally, in many studies, concentrations applied have been higher than those seen during pathological conditions (from 0.5 µM to 5 µM) [43]; however, we have used nanomolar concentrations of highly purified α-synM, which more accurately reflects the clinical condition. Moreover, few studies have used ex vivo brain tissue but rather cultured cells; thus, our findings more directly add to the body of knowledge of effects in native mammalian tissue. Finally, to the best of our knowledge, no one has previously reported a sex-based difference in cellular effects of α-synM on any neuronal type in ex vivo studies. α-synM was shown to induce an inhibitory effect in synaptic transmission in Calyx of Held; however, while both male and female rats were used, data were not analyzed for potential sex-based differences [43]. Taken together, this constitutes the first report to show that α-synM at a concentration reflective of clinical exposures induces a sex-dependent, cellular effect in mammalian neurons. Furthermore, as no difference in membrane effects was observed between LDTF and SNF, but we did see differences between the LDTM and SNM in earlier work [23], this suggests that α-synM induces sex-dependent effects in specific brain nuclei. Future studies of α-synM effects should consider sex as a factor as well as regional differences.

Functional implications

The observed sex-based, different cellular responses likely have pervasive functional implications for behaviors and symptoms when neurons of the LDT are exposed to α-synM. We have a working hypothesis that prodromal SDs in PD could be due to early dysfunction of sleep controlling nuclei, including the LDT, which has been implicated in RBD and EDS [23, 42, 44, 45].

Central to this hypothesis, we suspect that as levels of the monomeric form of α-syn rise, cellular effects are exerted on LDT neurons, which include enhancement of cellular excitability and increases in levels of intracellular calcium. Such effects are not elicited in the SN by the monomeric form, and it is believed to be the later appearing forms of aggregated oligomeric and fibril α-syn which produces neurodegeneration in the SN [23, 46,47,48,49,50]. Sustained elevation of excitability and calcium levels in the LDT induced by α-synM could trigger neurodegenerative processes when cells are excited for extended periods, and when calcium levels remain high [51, 52]. Consistent with this, we have shown that the effects induced by α-synM were associated with neuronal death in LDTM neurons. However, in SNM neurons in which α-synM induced an inhibitory membrane effect, and the predominant response was a decrease in intracellular calcium, no differences in neuronal survival were noted, which shows that α-synM cellular actions do not necessarily lead to degeneration as we saw in the LDT [23].

Interestingly, in the LDTF, α-synM induced very similar effects on the membrane and synaptic events to those seen in the SNM leading us to suggest that these effects are neuroprotective in the female LDT. This tenant is supported by greater cell viability in the LDT of the female compared to that in the male following α-synM exposure. The neuroprotective mechanisms involved inhibitory amino acid transmission as the blockade of GABA and glycine receptors revealed an excitatory effect in the LDTF similar to that seen in LDTM, which we hypothesize could underlie cellular degeneration. The outward current was sufficient to mask the concurrent excitatory effect and presumably limit putative damage from α-synM-mediated excitation. Further, treatment with GABA receptor agonists resulted in reductions in cell death in the LDTM lending further support to the interpretation that GABA signaling is neuroprotective. We did not identify the source of the putative, protective GABAergic effect in LDTF; however, elucidation of the source of this GABA tone in LDTF is of great interest and will be a focus of future studies. Also likely contributing to α-synM-induced neurodegeneration in the male LDT were the sex-dependent differences in the firing frequency, as a reduction in firing was induced in the LDTF by α-synM, whereas an enhancement in firing was seen in our earlier study in the LDTM [42]. Over the long-term, increases of neuronal discharge can produce overexcitability-induced cell death since high-levels of excitability and firing elevates glutamate exposure, which results in alterations of intracellular calcium levels. This can trigger apoptotic events and collapse of mitochondrial functions which are all processes contributing to neuronal death [53,54,55,56,57,58]. Accordingly, the α-synM-induced reduction in neuronal firing seen in LDTF could exert a protective effect.

One implication of our findings is that processes controlled by the LDT that could be affected in PD are less likely to be affected in females. While speculative, our data do support clinical findings related to occurrence of LDT-involved sleeping disorders seen prodromal to PD. RBD and EDS appear to be more common in PD males as well as in males during the prodromal phase. The majority of patients diagnosed with RBD are male, with the reported percentage of females in these studies ranging from 10 to 17.5% of all diagnosed cases [9, 10, 45, 59,60,61,62]. Although very few studies have focused on differences in SD symptoms between men and women, sex differences in RBD symptomatology have been reported with aggressive and violent motor active RBD behaviors appearing more commonly in men than in women [63, 64]. EDS is characterized by the incapacity of the individual to stay awake during the circadian day due to excessive sleepiness. Several studies have examined the risk of development of PD in EDS patients; however, the majority of investigations which have shown an association between EDS in the prodromal phase of PD have been conducted in males [13, 65]. In one of the few studies to include both sexes, a higher risk of development of PD was documented in those expressing sleepiness during the day; however, the data were not analyzed to compare the risk in males vs. females, and the majority of the cohort who exhibited EDS were males, which reflects the sex-based odds ratio of PD in the general population [66]. In several studies of PD diagnosed patients, EDS has been shown to be more common among PD-affected men than women [65, 67, 68].

Conclusion

Taken together, our data lead us to conclude that the cellular effects exerted by α-synM are neuroprotective in the LDTF and could be sufficient to delay α-synM-mediated cell death in this nucleus, perhaps ceasing when the relevant GABAergic neurons perish as neuronal loss proceeds throughout the PD affected brain. Output from the LDT to rostral and caudal targets is implicated in control of arousal during wakefulness, governance of the sleep and wakefulness cycle, and maintenance of motor atonia, which is a hallmark of REM sleep [18, 69]. Thus, if GABAergic neurotransmission plays a neuroprotective role by leading to outward currents and reductions in intracellular calcium, thereby counterbalancing negative effects induced by PD processes, this could block loss of cells in the LDT that produce motor atonia during REM sleep and an aroused EEG during wakefulness and sleep, and thereby lead to the lower frequency of RBD and EDS symptoms see in female when compared to those seen in male patients with PD [10, 70, 71].

As we observed no sex-based difference in cellular responses in the SN, our findings cannot account for the higher incidence of PD in males compared to females; however, we do suggest a mechanistic basis for the higher prevalence of SDs among male vs. female PD patients. Thus, our findings represent an important step toward the identification of sex differences in the mechanisms underlying the pathology of α-syn-associated neurogenerative diseases. Such identification offers the potential of targeting inhibitory mechanisms as neuroprotective strategies in neurodegenerative diseases, and to speed efforts for development of new directions for PD treatment and management in the prodromal phase of these diseases.

Data Availability

All data are available upon reasonable request made to the corresponding author.

References

  1. Goetz CG. The history of Parkinson’s Disease: early clinical descriptions and neurological therapies, Cold Spring Harb. Perspect Med. 2011;1:a008862–2. https://doi.org/10.1101/cshperspect.a008862.

    Article  CAS  Google Scholar 

  2. Fahn S. Description of Parkinson’s Disease as a clinical syndrome. Ann N Y Acad Sci. 2006;991:1–14. https://doi.org/10.1111/j.1749-6632.2003.tb07458.x.

    Article  Google Scholar 

  3. Poewe W, Seppi K, Tanner CM, Halliday GM, Brundin P, Volkmann J, Schrag A-E, Lang AE. Parkinson disease. Nat Rev Dis Prim. 2017;3:17013. https://doi.org/10.1038/nrdp.2017.13.

    Article  PubMed  Google Scholar 

  4. Elbaz A, Bower JH, Maraganore DM, McDonnell SK, Peterson BJ, Ahlskog JE, Schaid DJ, Rocca WA. Risk tables for parkinsonism and Parkinson’s disease. J Clin Epidemiol. 2002;55:25–31. https://doi.org/10.1016/S0895-4356(01)00425-5.

    Article  PubMed  Google Scholar 

  5. de Lau LML, Giesbergen PCLM, de Rijk MC, Hofman A, Koudstaal PJ, Breteler MMB. Incidence of parkinsonism and Parkinson disease in a general population: the Rotterdam Study, Neurology. 63 (2004) 1240–4. https://doi.org/10.1212/01.WNL.0000140706.52798.BE.

  6. Van Den Eeden SK, Ethnicity. Am J Epidemiol. 2003;157:1015–22. https://doi.org/10.1093/aje/kwg068.

    Article  PubMed  Google Scholar 

  7. Heller J, Dogan I, Schulz JB, Reetz K. Evidence for gender differences in cognition, emotion and quality of life in Parkinson’s disease? Aging Dis. 2014;5:63–75. https://doi.org/10.14366/AD.2014.050063.

    Article  PubMed  Google Scholar 

  8. Schenck CH. Expanded insights into idiopathic REM sleep behavior disorder. Sleep. 2016;39:7–9. https://doi.org/10.5665/sleep.5300.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Fernández-Arcos A, Iranzo A, Serradell M, Gaig C, Santamaria J. The clinical phenotype of idiopathic Rapid Eye Movement Sleep Behavior Disorder at Presentation: a study in 203 consecutive patients. Sleep. 2016;39:121–32. https://doi.org/10.5665/sleep.5332.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Postuma RB, Iranzo A, Hu M, Högl B, Boeve BF, Manni R, Oertel WH, Arnulf I, Ferini-Strambi L, Puligheddu M, Antelmi E, De Cock VC, Arnaldi D, Mollenhauer B, Videnovic A, Sonka K, Jung K-Y, Kunz D, Dauvilliers Y, Provini F, Lewis SJ, Buskova J, Pavlova M, Heidbreder A, Montplaisir JY, Santamaria J, Barber TR, Stefani A, St.Louis EK, Terzaghi M, Janzen A, Leu-Semenescu S, Plazzi G, Nobili F, Sixel-Doering F, Dusek P, Bes F, Cortelli P, Ehgoetz Martens K, Gagnon J-F, Gaig C, Zucconi M, Trenkwalder C, Gan-Or Z, Lo C, Rolinski M, Mahlknecht P, Holzknecht E, Boeve AR, Teigen LN, Toscano G, Mayer G, Morbelli S, Dawson B, Pelletier A. Risk and predictors of dementia and parkinsonism in idiopathic REM sleep behaviour disorder: a multicentre study. Brain. 2019;142:744–59. https://doi.org/10.1093/brain/awz030.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Postuma RB, Gagnon JF, Vendette M, Fantini ML, Massicotte-Marquez J, Montplaisir J. Quantifying the risk of neurodegenerative disease in idiopathic REM sleep behavior disorder. Neurology. 2009;72:1296–300. https://doi.org/10.1212/01.wnl.0000340980.19702.6e.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Sixel-Döring F, Trautmann E, Mollenhauer B, Trenkwalder C. Rapid Eye Movement Sleep behavioral events: a new marker for Neurodegeneration in Early Parkinson Disease? Sleep. 2014;37:431–8. https://doi.org/10.5665/sleep.3468.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Abbott RD, Ross GW, White LR, Tanner CM, Masaki KH, Nelson JS, Curb JD, Petrovitch H. Excessive daytime sleepiness and subsequent development of Parkinson disease. Neurology. 2005;65:1442–6. https://doi.org/10.1212/01.wnl.0000183056.89590.0d.

    Article  CAS  PubMed  Google Scholar 

  14. Spillantini MG, Crowther RA, Jakes R, Hasegawa M, Goedert M. Alpha-synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proc Natl Acad Sci U S A. 1998;95:6469–73. https://doi.org/10.1073/pnas.95.11.6469.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Decressac M, Mattsson B, Weikop P, Lundblad M, Jakobsson J, Björklund A. TFEB-mediated autophagy rescues midbrain dopamine neurons from α-synuclein toxicity. Proc Natl Acad Sci. 2013;110. https://doi.org/10.1073/pnas.1305623110.

  16. Dugger BN, Murray ME, Boeve BF, Parisi JE, Benarroch EE, Ferman TJ, Dickson DW. Neuropathological analysis of brainstem cholinergic and catecholaminergic nuclei in relation to rapid eye movement (REM) sleep behaviour disorder. Neuropathol Appl Neurobiol. 2012;38:142–52. https://doi.org/10.1111/j.1365-2990.2011.01203.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wang H-L, Morales M. Pedunculopontine and laterodorsal tegmental nuclei contain distinct populations of cholinergic, glutamatergic and GABAergic neurons in the rat. Eur J Neurosci. 2009;29:340–58. https://doi.org/10.1111/j.1460-9568.2008.06576.x.

    Article  PubMed  Google Scholar 

  18. Mena-Segovia J. Structural and functional considerations of the cholinergic brainstem. J Neural Transm. 2016;123:731–6. https://doi.org/10.1007/s00702-016-1530-9.

    Article  CAS  PubMed  Google Scholar 

  19. Webster HH, Jones BE. Neurotoxic lesions of the dorsolateral pontomesencephalic tegmentum-cholinergic cell area in the cat. II. Effects upon sleep-waking states. Brain Res. 1988;458:285–302. https://doi.org/10.1016/0006-8993(88)90471-4.

    Article  CAS  PubMed  Google Scholar 

  20. Chartier-Harlin MC, Kachergus J, Roumier C, Mouroux V, Douay X, Lincoln S, Levecque C, Larvor L, Andrieux J, Hulihan M, Waucquier N, Defebvre L, Amouyel P, Farrer M, Destée A. α-synuclein locus duplication as a cause of familial Parkinson’s disease. Lancet. 2004;364:1167–9. https://doi.org/10.1016/S0140-6736(04)17103-1.

    Article  CAS  PubMed  Google Scholar 

  21. Stefanis L, Emmanouilidou E, Pantazopoulou M, Kirik D, Vekrellis K, Tofaris GK. How is alpha-synuclein cleared from the cell? J Neurochem. 2019;150:577–90. https://doi.org/10.1111/JNC.14704.

    Article  CAS  PubMed  Google Scholar 

  22. Lashuel HA, Overk CR, Oueslati A, Masliah E. The many faces of α-synuclein: from structure and toxicity to therapeutic target. Nat Rev Neurosci. 2013;14:38–48. https://doi.org/10.1038/nrn3406.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Dos Santos AB, Skaanning LK, Mikkelsen E, Romero-Leguizamón CR, Kristensen MP, Klein AB, Thaneshwaran S, Langkilde AE, Kohlmeier KA. α-Synuclein responses in the Laterodorsal Tegmentum, the Pedunculopontine Tegmentum, and the Substantia Nigra: implications for early appearance of Sleep Disorders in Parkinson’s Disease. J Parkinsons Dis. 2021;11:1773–90. https://doi.org/10.3233/JPD-212554.

    Article  CAS  PubMed  Google Scholar 

  24. Boucetta S, Jones BE. Activity profiles of cholinergic and intermingled GABAergic and putative glutamatergic neurons in the Pontomesencephalic Tegmentum of urethane-anesthetized rats. J Neurosci. 2009;29:4664–74. https://doi.org/10.1523/JNEUROSCI.5502-08.2009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ipsen TH, Polli FS, Kohlmeier KA. Calcium rises induced by AMPA and nicotine receptors in the ventral tegmental area show differences in mouse brain slices prenatally exposed to nicotine. Dev Neurobiol. 2018;78:828–48. https://doi.org/10.1002/dneu.22607.

    Article  CAS  Google Scholar 

  26. Grynkiewicz G, Poenie M, Tsien RY. A new generation of Ca2 + indicators with greatly improved fluorescence properties. J Biol Chem. 1985;260:3440–50. https://doi.org/10.1016/S0021-9258(19)83641-4.

    Article  CAS  PubMed  Google Scholar 

  27. Connor JA, Cormier RJ. Cumulative effects of glutamate microstimulation on Ca2 + responses of CA1 hippocampal pyramidal neurons in slice. J Neurophysiol. 2000;83:90–8. https://doi.org/10.1152/JN.2000.83.1.90/ASSET/IMAGES/LARGE/9K0100736007.JPEG.

    Article  CAS  PubMed  Google Scholar 

  28. Angelova PR, Ludtmann MHR, Horrocks MH, Negoda A, Cremades N, Klenerman D, Dobson CM, Wood NW, Pavlov EV, Gandhi S, Abramov AY. Ca 2 + is a key factor in α-synuclein-induced neurotoxicity. J Cell Sci. 2016;129:1792–801. https://doi.org/10.1242/jcs.180737.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Luquin E, Huerta I, Aymerich MS, Mengual E. Stereological estimates of glutamatergic, GABAergic, and cholinergic neurons in the Pedunculopontine and Laterodorsal Tegmental nuclei in the rat. Front Neuroanat. 2018;12. https://doi.org/10.3389/fnana.2018.00034.

  30. Sherif F, Eriksson L, Oreland L. GABA-transaminase activity in rat and human brain: regional, age and sex-related differences. J Neural Transm. 1991;84:95–102. https://doi.org/10.1007/BF01249113.

    Article  CAS  Google Scholar 

  31. Cooke B, Hegstrom CD, Villeneuve LS, Breedlove SM. Sexual differentiation of the Vertebrate Brain: principles and mechanisms. Front Neuroendocrinol. 1998;19:323–62. https://doi.org/10.1006/frne.1998.0171.

    Article  CAS  PubMed  Google Scholar 

  32. Kokka N, Sapp DW, Witte U, Olsen RW. Sex differences in sensitivity to pentylenetetrazol but not in GABAA receptor binding. Pharmacol Biochem Behav. 1992;43:441–7. https://doi.org/10.1016/0091-3057(92)90174-E.

    Article  CAS  PubMed  Google Scholar 

  33. Segovia S, Guillamón A. Sexual dimorphism in the vomeronasal pathway and sex differences in reproductive behaviors. Brain Res Rev. 1993;18:51–74. https://doi.org/10.1016/0165-0173(93)90007-M.

    Article  CAS  PubMed  Google Scholar 

  34. Segovia S, Del Cerro MCR, Ortega E, Pérez-Laso C, Rodriguez-Zafra M, Izquierdo MAP, Guillamón A. Role of GABA(A) receptors in the organization of brain and behavioural sex differences, Neuroreport. (1996). https://doi.org/10.1097/00001756-199611040-00030.

  35. Davis AM, Grattan DR, Selmanoff M, Mccarthy MM. Sex differences in glutamic acid decarboxylase mRNA in neonatal rat brain: implications for sexual differentiation. Horm Behav. 1996;30:538–52. https://doi.org/10.1006/hbeh.1996.0057.

    Article  CAS  PubMed  Google Scholar 

  36. Peričić D, Bujas M. Sex differences in the response to GABA antagonists depend on the route of drug administration, exp. Brain Res. 1997;115:187–90. https://doi.org/10.1007/PL00005681.

    Article  Google Scholar 

  37. Peričić D, Manev H, Geber J. Sex related differences in the response of mice, rats and cats to administration of picrotoxin. Life Sci. 1986;38:905–13. https://doi.org/10.1016/0024-3205(86)90258-4.

    Article  PubMed  Google Scholar 

  38. Becú-Villalobos D, Iglesias AG, Díaz-Torga G, Hockl P, Libertun C. Brain sexual differentiation and gonadotropins secretion in the rat. Cell Mol Neurobiol. 1997. https://doi.org/10.1023/A:1022542221535.

    Article  PubMed  Google Scholar 

  39. Knickmeyer RC, Baron-Cohen S. Topical review: fetal testosterone and sex differences in typical Social Development and in Autism. J Child Neurol. 2006;21:825–45. https://doi.org/10.1177/08830738060210101601.

    Article  PubMed  Google Scholar 

  40. Benarroch EE. Pedunculopontine nucleus: functional organization and clinical implications. Neurology. 2013;80:1148–55. https://doi.org/10.1212/WNL.0b013e3182886a76.

    Article  PubMed  Google Scholar 

  41. Schmeichel AM, Buchhalter LC, Low PA, Parisi JE, Boeve BW, Sandroni P, Benarroch EE. Mesopontine cholinergic neuron involvement in Lewy body dementia and multiple system atrophy. Neurology. 2008;70:368–73. https://doi.org/10.1212/01.WNL.0000298691.71637.96.

    Article  CAS  PubMed  Google Scholar 

  42. Dos Santos AB, Skaanning LK, Thaneshwaran S, Mikkelsen E, Romero-Leguizamón CR, Skamris T, Kristensen MP, Langkilde AE, Kohlmeier KA. Sleep-controlling neurons are sensitive and vulnerable to multiple forms of α-synuclein: implications for the early appearance of sleeping disorders in α-synucleinopathies. Cell Mol Life Sci. 2022;79:450. https://doi.org/10.1007/s00018-022-04467-z.

    Article  CAS  PubMed  Google Scholar 

  43. Eguchi K, Taoufiq Z, Thorn-Seshold O, Trauner D, Hasegawa M, Takahashi T. Wild-type Monomeric α-Synuclein can impair vesicle endocytosis and synaptic Fidelity via Tubulin polymerization at the Calyx of held. J Neurosci. 2017;37:6043–52. https://doi.org/10.1523/JNEUROSCI.0179-17.2017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Boeve BF, Silber MH, Saper CB, Ferman TJ, Dickson DW, Parisi JE, Benarroch EE, Ahlskog JE, Smith GE, Caselli RC, Tippman-Peikert M, Olson EJ, Lin SC, Young T, Wszolek Z, Schenck CH, Mahowald MW, Castillo PR, Del Tredici K, Braak H. Pathophysiology of REM sleep behaviour disorder and relevance to neurodegenerative disease. Brain. 2007. https://doi.org/10.1093/brain/awm056.

    Article  PubMed  Google Scholar 

  45. Postuma RB, Gagnon J-F, Vendette M, Montplaisir JY. Idiopathic REM sleep behavior disorder in the transition to degenerative disease. Mov Disord. 2009;24:2225–32. https://doi.org/10.1002/mds.22757.

    Article  PubMed  Google Scholar 

  46. Bengoa-Vergniory N, Roberts RF, Wade-Martins R, Alegre-Abarrategui J. Alpha-synuclein oligomers: a new hope. Acta Neuropathol. 2017;134:819–38. https://doi.org/10.1007/s00401-017-1755-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Thakur P, Breger LS, Lundblad M, Wan OW, Mattsson B, Luk KC, Lee VMY, Trojanowski JQ, Björklund A. Modeling Parkinson’s disease pathology by combination of fibril seeds and α-synuclein overexpression in the rat brain. Proc Natl Acad Sci. 2017;114. https://doi.org/10.1073/pnas.1710442114.

  48. Thakur P, Chiu WH, Roeper J, Goldberg JA. α-Synuclein 2.0 — moving towards cell type specific pathophysiology. Neuroscience. 2019;412:248–56. https://doi.org/10.1016/j.neuroscience.2019.06.005.

    Article  CAS  PubMed  Google Scholar 

  49. Tsigelny IF, Sharikov Y, Wrasidlo W, Gonzalez T, Desplats PA, Crews L, Spencer B, Masliah E. Role of α-synuclein penetration into the membrane in the mechanisms of oligomer pore formation. FEBS J. 2012. https://doi.org/10.1111/j.1742-4658.2012.08489.x.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Pacheco CR, Morales CN, Ramírez AE, Muñoz FJ, Gallegos SS, Caviedes PA, Aguayo LG, Opazo CM. Extracellular α-synuclein alters synaptic transmission in brain neurons by perforating the neuronal plasma membrane. J Neurochem. 2015;132:731–41. https://doi.org/10.1111/jnc.13060.

    Article  CAS  PubMed  Google Scholar 

  51. Lau A, Tymianski M. Glutamate receptors, neurotoxicity and neurodegeneration. Pflügers Arch - Eur J Physiol. 2010;460:525–42. https://doi.org/10.1007/s00424-010-0809-1.

    Article  CAS  Google Scholar 

  52. Wang Y, Qin Z. Molecular and cellular mechanisms of excitotoxic neuronal death. Apoptosis. 2010;15:1382–402. https://doi.org/10.1007/s10495-010-0481-0.

    Article  CAS  PubMed  Google Scholar 

  53. Vergun O, Sobolevsky AI, Yelshansky MV, Keelan J, Khodorov BI, Duchen MR. Exploration of the role of reactive oxygen species in glutamate neurotoxicity in rat hippocampal neurones in culture. J Physiol. 2001;531:147–63. https://doi.org/10.1111/j.1469-7793.2001.0147j.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Abramov AY, Duchen MR. Impaired mitochondrial bioenergetics determines glutamate-induced delayed calcium deregulation in neurons, Biochim. Biophys. Acta - Gen Subj. 2010. https://doi.org/10.1016/j.bbagen.2009.08.002.

    Article  Google Scholar 

  55. Vergun O, Keelan J, Khodorov BI, Duchen MR. Glutamate-induced mitochondrial depolarisation and perturbation of calcium homeostasis in cultured rat hippocampal neurones. J Physiol. 1999;519:451–66. https://doi.org/10.1111/j.1469-7793.1999.0451m.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Stout AK, Reynolds IJ. High-affinity calcium indicators underestimate increases in intracellular calcium concentrations associated with excitotoxic glutamate stimulations. Neuroscience. 1999;89:91–100. https://doi.org/10.1016/S0306-4522(98)00441-2.

    Article  CAS  PubMed  Google Scholar 

  57. Stout AK, Raphael HM, Kanterewicz BI, Klann E, Reynolds IJ. Glutamate-induced neuron death requires mitochondrial calcium uptake. Nat Neurosci. 1998;1:366–73. https://doi.org/10.1038/1577.

    Article  CAS  PubMed  Google Scholar 

  58. Abramov AY, Duchen MR. Mechanisms underlying the loss of mitochondrial membrane potential in glutamate excitotoxicity, Biochim. Biophys. Acta - Bioenerg. 2008;1777:953–64. https://doi.org/10.1016/j.bbabio.2008.04.017.

    Article  CAS  Google Scholar 

  59. Kotagal V, Albin RL, Müller MLTM, Koeppe RA, Chervin RD, Frey KA, Bohnen NI. Symptoms of rapid eye movement sleep behavior disorder are associated with cholinergic denervation in Parkinson disease. Ann Neurol. 2012;71:560–8. https://doi.org/10.1002/ana.22691.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Iranzo A, Molinuevo JL, Santamaría J, Serradell M, Martí MJ, Valldeoriola F, Tolosa E. Rapid-eye-movement sleep behaviour disorder as an early marker for a neurodegenerative disorder: a descriptive study. Lancet Neurol. 2006;5:572–7. https://doi.org/10.1016/S1474-4422(06)70476-8.

    Article  PubMed  Google Scholar 

  61. Wing YK, Lam SP, Li SX, Yu MWM, Fong SYY, Tsoh JMY, Ho CKW, Lam VKH. REM sleep behaviour disorder in Hong Kong Chinese: clinical outcome and gender comparison. J Neurol Neurosurg Psychiatry. 2008;79:1415–6. https://doi.org/10.1136/jnnp.2008.155374.

    Article  CAS  PubMed  Google Scholar 

  62. Olson EJ, Boeve BF, Silber MH. Rapid eye movement sleep behaviour disorder: demographic, clinical and laboratory findings in 93 cases. Brain. 2000;123:331–9. https://doi.org/10.1093/brain/123.2.331.

    Article  PubMed  Google Scholar 

  63. Bodkin CL, Schenck CH. Rapid Eye Movement Sleep Behavior Disorder in Women: relevance to General and Specialty Medical Practice, J Women’s Heal 18 (2009) 1955–63. https://doi.org/10.1089/jwh.2008.1348.

  64. Bjørnarå KA, Dietrichs E, Toft M. REM sleep behavior disorder in Parkinson’s disease – is there a gender difference? Parkinsonism Relat Disord. 2013;19:120–2. https://doi.org/10.1016/j.parkreldis.2012.05.027.

    Article  PubMed  Google Scholar 

  65. Picillo M, Erro R, Amboni M, Longo K, Vitale C, Moccia M, Pierro A, Scannapieco S, Santangelo G, Spina E, Orefice G, Barone P, Pellecchia MT. Gender differences in non-motor symptoms in early Parkinson’s disease: a 2-years follow-up study on previously untreated patients. Parkinsonism Relat Disord. 2014;20:850–4. https://doi.org/10.1016/j.parkreldis.2014.04.023.

    Article  PubMed  Google Scholar 

  66. Gao J, Huang X, Park Y, Hollenbeck A, Blair A, Schatzkin A, Chen H. Daytime Napping, Nighttime sleeping, and Parkinson Disease. Am J Epidemiol. 2011;173:1032–8. https://doi.org/10.1093/aje/kwq478.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Martinez-Martin P, Falup Pecurariu C, Odin P, Hilten JJ, Antonini A, Rojo-Abuin JM, Borges V, Trenkwalder C, Aarsland D, Brooks DJ, Ray K, Chaudhuri. Gender-related differences in the burden of non-motor symptoms in Parkinson’s disease. J Neurol. 2012;259:1639–47. https://doi.org/10.1007/s00415-011-6392-3.

    Article  PubMed  Google Scholar 

  68. Kovács M, Makkos A, Aschermann Z, Janszky J, Komoly S, Weintraut R, Karádi K, Kovács N. Impact of sex on the nonmotor symptoms and the Health-Related Quality of Life in Parkinson’s Disease. Parkinsons Dis. 2016;2016:1–12. https://doi.org/10.1155/2016/7951840.

    Article  CAS  Google Scholar 

  69. Kroeger D, Ferrari LL, Petit G, Mahoney CE, Fuller PM, Arrigoni E, Scammell TE. Cholinergic, glutamatergic, and GABAergic neurons of the Pedunculopontine Tegmental Nucleus have distinct Effects on Sleep/Wake Behavior in mice. J Neurosci. 2017;37:1352–66. https://doi.org/10.1523/JNEUROSCI.1405-16.2016.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Schenck CH, Boeve BF, Mahowald MW. Delayed emergence of a parkinsonian disorder or dementia in 81% of older men initially diagnosed with idiopathic rapid eye movement sleep behavior disorder: a 16-year update on a previously reported series. Sleep Med. 2013;14:744–8. https://doi.org/10.1016/j.sleep.2012.10.009.

    Article  PubMed  Google Scholar 

  71. Barber TR, Lawton M, Rolinski M, Evetts S, Baig F, Ruffmann C, Gornall A, Klein JC, Lo C, Dennis G, Bandmann O, Quinnell T, Zaiwalla Z, Ben-Shlomo Y, Hu MTM. Prodromal parkinsonism and neurodegenerative risk stratification in REM sleep behavior disorder. Sleep. 2017;40. https://doi.org/10.1093/sleep/zsx071.

  72. Van Dort CJ, Zachs DP, Kenny JD, Zheng S, Goldblum RR, Gelwan NA, Ramos DM, Nolan MA, Wang K, Weng F-J, Lin Y, Wilson MA, Brown EN. Optogenetic activation of cholinergic neurons in the PPT or LDT induces REM sleep. PNAS 11(2):584–9. https://doi.org/10.1073/pnas.1423136112.

Download references

Acknowledgements

Christel Ammitzböll Halberg is acknowledged for technical assistance.

Funding

Open access funding provided by Royal Library, Copenhagen University Library. Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), Brazil is acknowledged for funding support in the form of a Ph.D Grant to Altair Brito dos Santos. AEL acknowledges funding from the Lundbeck Foundation Initiative BRAINSTRUC (2015–2666).

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the study conception and design. Material preparation, data collection and analysis of calcium imaging and electrophysiological data were performed by ABS, ST, YW, MPK, AEL, and KAK. Material preparation, data collection and analysis of cell death data was conducted by ABS, LKA, CRRL, AEL, and KAK. The first draft of the manuscript was written by ABS and KAK and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Kristi A. Kohlmeier.

Ethics declarations

Ethics approval and consent to participate

All animal use was approved in concordance with the European Communities Council Directive (86/609/EEC). No human sourced material was included in this study.

Consent for publication

All authors read and approved the final manuscript.

Competing interests

All authors disclose that they have no relevant financial or non-financial interests to disclose.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Santos, A.B.D., Thaneshwaran, S., Ali, L.K. et al. Sex-dependent neuronal effects of α-synuclein reveal that GABAergic transmission is neuroprotective of sleep-controlling neurons. Cell Biosci 13, 172 (2023). https://doi.org/10.1186/s13578-023-01105-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13578-023-01105-4

Keywords