Here we demonstrate that pPVT neurons in males and females display different synaptic, passive membrane, and active membrane properties in response to restraint stress. Identification of these differences in PVT neuron function represents important progress in our understanding of sex differences in the stress response as the pPVT is necessary for habituation to repeated restraint stress (7, 33). In general, our findings demonstrate that electrophysiological properties of pPVT neurons are altered by restraint in males, but females only display moderate changes in response to restraint. Changes in the mRNA of specific voltage-gated ion channel transcripts within the pPVT are consistent with electrophysiological properties of pPVT neurons that are altered by restraint and/or different between males and females. Together, these results are the first to identify sex differences in pPVT neuron function in the context of stress.
Males and females display differences in sEPSCs in the pPVT at baseline and following restraint.
We examined the effects of restraint on electrophysiological properties of pPVT neurons in males and females. Electrophysiological properties were assessed 24 hours following 1 or 5 days of restraint rather than immediately following restraint as we were primarily interested in investigating stable changes induced by stress. Examination of pPVT neurons proximal to restraint is likely to identify properties of these cells that are impacted by the immediate stress or rapid recovery from the stress. While the short-term impact of stress is important to study, we chose to examine those properties of pPVT cells that are likely to influence the response to stress the next day as habituation is influenced by prior experience and predictability of stress experiences (1). We found that sEPSC amplitude was lower in non-restrained males compared to non-restrained females. Restraint, either 1 or 5 days, increased sEPSC amplitude in males to be similar to that of non-restrained females and had no effect on sEPSC amplitude in females. Increased EPSC amplitude can be primarily attributed to increased post-synaptic glutamate receptors. However, other factors may also contribute to increases in EPSC amplitude including increases in other neurotransmitter receptors with high cation conductance or increased pre-synaptic release of glutamate. Although we cannot rule out these possibilities, the observed decay times suggest that sEPSCs in the pPVT are consistent with those of AMPA receptors (40). Therefore, we hypothesize that restraint increases AMPA receptor expression or trafficking to the synapse in male, but not female, pPVT neurons.
The sEPSC decay times of 4–10 msec that we observed in pPVT neurons were consistent with those of AMPA receptors (40), whereas those of NMDA, kainate, and metabotropic receptors are significantly longer and in the 100 msec range (41–46). We observed that sEPSC decay time is increased in non-restrained females compared to non-restrained males. sEPSC decay time is increased by 1 and 5 days of restraint in males, but decreased by 1 and 5 days of restraint in females. Increased EPSC decay time increases excitatory post-synaptic potential amplitude and increases the probability of action potential firing (47). Therefore, compared to baseline conditions, restraint is predicted to increase action potential firing in pPVT neurons of males, but reduce action potential firing in pPVT neurons of females. Longer sEPSC decay times may be attributed to asynchronous glutamate release in pre-synaptic axon terminals, increased glutamate release, or slower glutamate reuptake and clearance from the synaptic cleft (48, 49). Thus, glutamate transporter expression or function may be decreased by restraint in males, but increased by restraint in females. Together, the effects of increased EPSC amplitude and EPSC decay time in response to restraint in males indicate that restraint increases both the amplitude and duration of each individual excitatory quantum in the pPVT. This is expected to cause a compounding effect on pPVT output as both EPSC amplitude and EPSC decay time increase the probability of action potential firing (47). Therefore, restraint increases responsiveness to glutamatergic inputs and the probability of action potential firing in the pPVT of males. Because restraint has no effect on EPSC amplitude and reduces EPSC decay time in females, restraint decreases the responsiveness of the pPVT to glutamatergic inputs in females, which should reduce the probability of glutamate-induced action potential firing. There were no effects of sex or restraint on sEPSC frequency, which would suggest changes in the frequency of presynaptic glutamatergic vesicle release. Our findings of restraint-induced increases in EPSC amplitude and decay time are consistent with our recent finding that expression of the neuronal activity markers c-Fos and Arc is increased in the pPVT of male rats following 1 or 5 days of restraint (17). Thus, increased sEPSC amplitude and EPSC decay time represent two important electrophysiological properties that are consistent with increases in pPVT action potential firing caused by restraint in males. Because pPVT activity is necessary for habituation (17, 50), these electrophysiological properties may contribute to habituation in male rats. In female rats, sEPSC amplitude is unchanged by restraint and sEPSC decay time is reduced by restraint. These findings are predicted to have no effect on pPVT firing or reduce pPVT firing, respectively, and may at least partially explain why habituation is impaired in female rats.
Males and females display differences in passive membrane properties in the pPVT at baseline and following restraint.
Resting membrane potential was more hyperpolarized in female pPVT neurons compared to males. Although the difference in mean resting membrane potential was modest, even subtle changes in this basic electrophysiological property could affect action potential firing. Resting membrane potential is regulated by differences in extracellular and intracellular ion concentrations and the permeability of those ions (51, 52). Therefore, more hyperpolarized pPVT neurons in females may be due to changes in ion transporters or ion leak channels. Resting membrane potential in pPVT neurons is complex and regulated by a wide variety of conductances including those mediated by inwardly rectifying potassium channels and TWIK-related acid sensitive potassium channels (53). Thus, increases in the expression and/or membrane trafficking of these ion channels and others may underlie the more hyperpolarized resting membrane potential observed in the pPVT neurons of females compared to males. Input resistance was increased by 1 and 5 days of restraint in pPVT neurons of male, but not female, rats. Increased input resistance is primarily attributed to reduced potassium leakage in response to depolarizing current. Restraint may reduce the expression, function, and/or membrane trafficking of potassium leak channels in pPVT neurons of male, but not female, rats. Reduced potassium leakage in response to membrane depolarization is predicted to facilitate EPSC integration and thus increase the likelihood of action potential firing. Together, these findings suggest that, compared to females, the passive membrane properties of pPVT neurons of male rats make them more likely to fire action potentials in response to excitatory inputs. Therefore, in addition to restraint-induced increases in sEPSC amplitude and decay time, a more depolarized resting membrane potential and restraint-induced increases in input resistance may further contribute to the observation that pPVT activity is increased by restraint male rats (17). Resting membrane potential is more hyperpolarized in females compared to males and restraint has no effect on input resistance in females. These findings are predicted to render pPVT activity less responsive to restraint and may contribute to habituation impairments in females.
Males and females display differences in active membrane properties in the pPVT at baseline and following restraint.
Males displayed higher percentages of single spike and burst-firing pPVT neurons than females whereas females displayed a higher percentage of sustained-firing neurons. These findings were consistent regardless of restraint group. Sustained-firing neurons have a short refractory period, allowing them to fire continuously in response to prolonged experimental depolarization. Because sustained-firing neurons fire action potentials with a higher frequency in response to continuous depolarization, a higher percentage of sustained firing pPVT neurons in females might suggest increased pPVT output in females compared to males. However, this would only be true if the pPVT is continuously depolarized in vivo. Excitatory inputs that depolarize pPVT neurons past the action potential threshold may occur less frequently in vivo than the refractory period required for action potential firing in burst-firing pPVT neurons. If this was the case, action potential frequency in pPVT neurons might be similar in males and females because the limiting factor in firing an action potential would be frequency and amplitude of EPSCs. The in vivo refractory period of burst-firing pPVT neurons and in vivo EPSC frequency are unknown to the best of our knowledge. Therefore, we cannot accurately predict the effect of neuron firing type percentages on overall pPVT action potential firing frequency in vivo.
Within each sex, restraint had little effect on the percentage of pPVT neurons displaying properties of each firing type. Compared to the other female groups, the 5-day restraint females exhibited an increased percentage of sustained-firing neurons and a decrease in bursting neurons. However, we should note that chi-square analysis was only significant when all groups were included in the analysis. We did not observe differences between any two groups within or between sexes. Firing patterns of pPVT neurons change from day to night phases. pPVT neurons during the light phase are more hyperpolarized and primarily single-spiking, but during the dark phase they are more depolarized and tend to be bursting or sustained-firing. These changes during the dark phase are regulated by reduced potassium currents and increased T-type calcium channel currents (53, 54). We should note that these recordings were taken from anterior pPVT neurons whereas our recordings were taken from posterior pPVT neurons and our recordings were conducted during the animal’s light phase. This is notable because anterior and posterior subdivisions of the pPVT have different functions and anatomical connectivity (20, 29–31). Further, the effects of sex on circadian influence of pPVT neuron firing types is unknown to the best of our knowledge.
Following 1, but not 5, days of restraint, the threshold for firing an action potential in pPVT neurons of males was modestly depolarized compared to non-restrained controls. We also observed an overall sex effect indicating that action potential firing threshold was more depolarized in male pPVT neurons compared to females. These findings suggest that pPVT neurons of 1-day restraint males may require more excitatory input in order to fire an action potential compared to non-restrained males. However, the resting membrane potential of male pPVT neurons is depolarized compared to females, so similar excitatory inputs may induce similar action potential firing in males and females in vivo. This is because the relative difference between resting membrane potential and the membrane potential required for firing action potentials (action potential firing threshold) is similar in males and females (~ 24 mV for each sex regardless of restraint group). Action potential half-width in pPVT neurons was increased in males restrained for 1, but not 5, days compared to non-restrained males. This indicates that 1 day of restraint increased the duration of action potentials in males, which may increase voltage-sensitive calcium transients in axon terminals and promote neurotransmitter release from pPVT neurons. Afterhyperpolarization potentials (AHPs) were greater in females compared to males, regardless of restraint group. This suggests that voltage-gated potassium channels may remain open longer in female pPVT neurons and the action potential refractory period may be longer compared to that of males. Although female pPVT neurons are more likely to display sustained firing patterns, greater AHPs may reduce action potential firing frequency in female pPVT neurons compared to male pPVT neurons with the same firing pattern. Together, these findings suggest that active membrane properties of pPVT neurons are different in males and females at baseline and in response to restraint stress.
Restraint stress has different effects on the expression of voltage-gated ion channel mRNA transcripts in the pPVT of males and females.
Similar to the timeframes of our studies on the electrophysiological properties of male and female pPVT neurons, we examined the mRNA of voltage-gated ion channel transcripts 24 hours following 1 and 5 days of restraint compared to non-restrained controls. This allowed us to investigate stable changes in gene expression that temporally correlate with changes in pPVT neuron electrophysiology. All data were presented relative to mean housekeeping gene expression of non-restrained controls so that relative levels of each transcript, and therefore their influence on neuron function, could be inferred qualitatively. Male rats displayed reduced expression of Kcnj6 transcripts 24 hours following a 5th daily restraint. Kcnj6 encodes the G-protein-activated inward rectifier potassium channel 2. Inwardly rectifying potassium currents are important regulators of the resting membrane potential of pPVT neurons (53). Therefore, reduced expression of Kcnj6 may contribute to depolarization of resting membrane potential in males following 5 days of restraint. Although resting membrane potential of pPVT neurons is not changed by restraint in males, reduced Kcnj6 expression may be an important factor that compensates for other factors that would otherwise drive hyperpolarization of resting membrane potential following restraint. Wild-type Kcnj6 inhibits dopaminergic tone (55–57). A Kcnj6 variant has been linked to alcohol dependence in individuals exposed to psychosocial stress early in life (58). Single nucleotide polymorphisms (SNPs) in the Kcnj6 gene are risk factors for developing attention-deficit/hyperactivity disorder (ADHD) (59). Addiction (60, 61) and ADHD (62) are both heavily influenced by the rewarding and reinforcing effects of dopaminergic neurotransmission. The PVT receives dopaminergic inputs (63) and is an important regulator of reward (64–66) and attention (17). Therefore, reduced Kcnj6 expression could contribute to addiction and/or attention deficits in stressed animals.
Male rats displayed increased expression of Kcnh3 24 hours following a 5th restraint. Kcnh3 encodes Kv12.2, a voltage-gated potassium channel subunit. Kv12.2 deletions reduce action potential firing threshold (67). Thus, increased expression of Kv12.2 following restraint may contribute to the more depolarized action potential firing threshold observed in the PVT of restrained male rats. Kcnh3 knockout mice display improved memory and PFC-mediated attention (68). Because restraint (33) and the pPVT (17) regulate PFC-mediated cognitive flexibility and attention, restraint-induced increases in Kcnh3 in the PVT may impair cognitive function. Following 1, but not 5, restraints, male rats displayed reduced expression of Kcnk1 mRNA, which encodes the two-pore domain potassium channel TWIK-1. Little is known about the function of TWIK-1 in the brain (69). However, reduced Kcnk1 expression in 1-day restraint males may represent an early response to restraint and contribute to increased action potential half-width in 1-day restraint males as reduced Kcnk1 expression may impair potassium efflux. Kcnk1 expression may return to baseline expression after 5 restraints in response to changes in the electrophysiological properties of pPVT neurons and/or reduced Kcnj6 expression. Together, these findings suggest that changes in the mRNA of certain voltage-gated ion channels in the pPVT of male rats may contribute to restraint-induced changes in the electrophysiological properties observed in the pPVT of male rats.
Although restraint did not affect active membrane properties in female pPVT neurons, mRNA transcripts encoding 10 different voltage-gated ion channels were altered by 1 or 5 days of restraint in the female pPVT. Some of these channels regulate similar functions, but restraint has opposite effects on their expression. Therefore, changes in the expression of these channels may be countered by compensatory changes in the expression of other transcripts encoding similar functions. For example, twenty-four hours following a 5th daily restraint, female rats displayed increased mRNA expression of Kcnb2 transcripts, but decreased expression of Kcnd2. Kcnb2 and Kcnd2 encode the voltage-gated potassium channels Kv2.2 and Kv4.2, respectively. These channels regulate delayed rectifier currents during the action potential (70). Additionally, the expression of Kcnn1 and Kcnn2 transcripts, which encode the calcium-activated potassium channel subfamily N members KCa2.1 and KCa2.2 (71), respectively, were both increased 24 hours following a 5th restraint in the female pPVT. These channels augment the afterhyperpolarization phase of the action potential and thereby have inhibitory effects in the neurons they are expressed in (72). Kcnn2 overexpression in the amygdala reduces anxiety-like behavior, presumably by reducing amygdala output (73). Restraint-induced increases in Kcnn2 might impair habituation in females by countering restraint-induced effects in the pPVT that would otherwise increase its excitability. The expression of Kcnmb4, which encodes the calcium-activated potassium channel subfamily M beta subunit 4, was decreased 24 hours following 1 and 5 days of restraint in female pPVT neurons. It is possible that the opposing expression patterns of these calcium-activated potassium channels counter one another to prevent restraint-induced changes in afterhyperpolarization potential. If these changes are compensatory, reduced Kcnmb4 may initiate changes in the expression of other voltage-gated ion channels as it is the only ion channel to be altered by a single restraint.
Compared to non-restrained controls, the expression of Scn2a1, Scn2b, and Scn8a were all increased in the PVT of females 24 hours following 5 days of restraint. Scn2a1, Scn2b, and Scn8a encode the voltage-gated sodium channel subunits Nav1.2, Navβ2, and Nav1.6, respectively (74). Both Nav1.1 and Nav1.6 contain voltage-sensing and pore-forming domains of the voltage-gated sodium channel (75). Navβ2 is an auxiliary subunit for voltage-gated sodium channels involved in the trafficking of voltage-sensitive subunits to the plasma membrane and stabilizing it there (75–77). Although 3 different transcripts that are predicted to increase voltage-dependent sodium channel currents were increased in the pPVT of females restrained for 5 days compared to non-restrained females, no active membrane properties were affected by restraint in females. This may be because voltage-gated sodium channel translation was impaired, trafficking of the channels to the plasma membrane was impaired, or compensatory changes in the functional expression of other channels that regulate active membrane properties negated voltage-gated sodium channel function. Compared to non-restrained controls, the expression of Kcnk1 was increased in females 24 hours following a 5th daily restraint. The expression of Slc12a5, which encodes the potassium-chloride cotransporter KCC2, was increased in the pPVT of females following 5 restraints compared to non-restrained controls. This channel is the major extruder of intracellular chloride in mature neurons, allowing for chloride influx during GABAergic neurotransmission (78). Therefore, increased expression of KCC2 could enhance GABAergic chloride currents that inhibit PVT activity in vivo. Together, these findings indicate that although restraint did not affect active membrane properties of PVT neurons in females, the expression of 10 voltage-gated ion channels were altered by restraint. Effects of altered mRNA expression may be negated by impaired translation, impaired protein regulation, or countered by compensatory changes in the expression or function of other voltage-gated ion channels as has been reported in rodent models of epilepsy (79). Further studies are needed to fully understand how these stress-induced changes in expression of voltage gated ion channels may regulate functional consequences of pPVT activity in male and female animals.
Perspectives And Significance
These findings are the first to characterize the effects of sex and stress on electrophysiological properties and voltage-gated ion channel expression in pPVT neurons. We found that restraint altered EPSCs, input resistance, and active membrane properties that are predicted to increase pPVT activity in males, but restraint did not have these effects in females. In pPVT neurons of males, restraint-induced changes in the expression of certain voltage-gated ion channels may contribute to some of the changes in active membrane properties that were altered by restraint. Restraint modulated the expression of 10 different voltage-gated ion channels in pPVT neurons of females, but there were few effects of stress on electrophysiological properties. Together, these findings suggest that restraint increases pPVT activity in males, but only modestly affects pPVT activity in females. These findings identify mechanisms through which males may habituate to 5 days of repeated restraint, but females do not. Sex differences in electrophysiological properties of pPVT neurons, under baseline conditions and following restraint, may help to explain the impaired habituation in females to repeated mild restraint. The increased number of voltage-gated ion channels that are altered by restraint in females might reflect compensatory changes in channel expression that prevent altered electrophysiological function, which might have been caused by restraint-induced changes in the expression of a few key voltage-gated ion channels. Restraint-induced changes in ion channel expression might also represent flux in membrane properties that could become apparent once females begin to habituate, which is likely around day 9 or 10. These changes may also be due to differences in estrous cyclicity, which we did not assess here. Future studies should determine the role of estradiol on voltage-gated ion channel expression using ovariectomized females and estradiol replacement. These findings are an important step in developing a comprehensive understanding of the genes and electrophysiological processes that underlie habituation in male rats to repeated restraint and the mechanisms contributing to delayed habituation in females.