Skip to main content
  • Research article
  • Open access
  • Published:

Effects of prostaglandin E2 on the electrical properties of thermally classified neurons in the ventromedial preoptic area of the rat hypothalamus

Abstract

Background

Physiological and morphological evidence suggests that activation of the ventromedial preoptic area of the hypothalamus (VMPO) is an essential component of an intravenous LPS-dependent fever. In response to the endogenous pyrogen prostaglandin E2 (PGE2), the majority of temperature insensitive neurons in the VMPO show an increase in firing rate, while warm sensitive neurons are inhibited. We have hypothesized that these PGE2 dependent effects on firing rate are due to changes in the inherent electrical properties of VMPO neurons, which are regulated by the activity of specific ionic currents.

Results

To characterize the electrical properties of VMPO neurons, whole-cell recordings were made in tissue slices from male Sprague-Dawley rats. Our results indicate that PGE2 dependent firing rate responses were not the result of changes in resting membrane potential, action potential amplitude and duration, or local synaptic input. However, PGE2 reduced the input resistance of all VMPO neurons, while increasing the excitability of temperature insensitive neurons and decreasing the excitability of warm sensitive neurons. In addition, the majority of temperature insensitive neurons responded to PGE2 with an increase in the rate of rise of the depolarizing prepotential that precedes each action potential. This response to PGE2 was reversed for warm sensitive neurons, in which the prepotential rate of rise decreased.

Conclusion

We would therefore suggest that PGE2 is having an effect on the ionic currents that regulate firing rate by controlling how fast membrane potential rises to threshold during the prepotential phase of the action potential.

Background

Fever, an elevation in body temperature, is thought to play an adaptive role in the immune system's ability to fight infection [1]. A suggested mechanism for its production and maintenance is a shifting of the thermostatic set-point into the hyperthermic range [1, 2]. Through the integration of both central and afferent thermal information, this set-point is established by the activity of neurons in the preoptic and anterior regions of the hypothalamus (PO/AH) that can be thermally classified on the basis of their inherent ability to respond to changes in temperature [3]. The majority of PO/AH neurons are considered temperature insensitive, showing little or no temperature dependent changes in firing rate. Approximately 30% of PO/AH neurons can be classified as warm sensitive, responding to local warming with an increase in firing rate [4]. While there has been considerable debate as to the criteria that should be used to classify a neuron as warm sensitive, we have used a regression coefficient of at least 0.8 impulses·s-1·°C-1. This criterion is based on previous studies that indicate a functional difference for neurons which show this degree of inherent thermosensitivity [3, 4]. In addition to responding to local changes in temperature, some of these warm sensitive neurons are also responsive to changes in skin or spinal temperature, while others show thermally dependent changes in their firing rates that may directly correlate with the activation of specific thermoregulatory responses. Although this integrative ability seems to be restricted to warm sensitive neurons in the PO/AH, temperature insensitive neurons may also play an important role in determining the set-point temperature through their synaptic interactions with thermoregulatory effector neurons [3]. Regardless of thermosensitivity, many PO/AH neurons may respond to adjustments in other homeostatic conditions or the presence of endogenous pyrogens such as prostaglandin E2 (PGE2), which could shift the thermostatic set-point and alter the activation of thermoregulatory mechanisms [2].

In response to stimulation of the immune system, changes in the activity of neurons in specific regions of the PO/AH may be responsible for the adjustment of the thermostatic set-point that results in an elevation in body temperature. Physiologic evidence suggests that in response to endotoxins such as lipopolysaccharide (LPS), this shift in set-point is mediated by the activation of afferent neural pathways or the production of systemic pyrogens, which ultimately leads to the local production of PGE2 within the hypothalamus [5, 6]. Early microinjection studies clearly established a role for prostaglandins in the production of a fever and later identified the importance of the region surrounding the OVLT in this response [710]. More recently, it has been shown that fever in response to intravenous LPS is dependent on the presence of the PGE2 producing enzyme cyclooxygenase-2 in the ventromedial preoptic area of the hypothalamus (VMPO) [11]. In addition, it has now been demonstrated that unlike other regions of the PO/AH, PGE2 has a selective effect on the firing rates of VMPO neurons, based on thermosensitivity, with PGE2 increasing the firing rates of temperature insensitive neurons and inhibiting the firing rates of warm sensitive neurons [12]. Anatomical studies also support the importance of the VMPO in the production of a fever, demonstrating that either the intravenous injection of LPS or microinjection of PGE2 directly into the VMPO will produce a fever that can be correlated with an increase in the cellular activation of neurons within the VMPO [13, 14].

Using a functional criterion for determining the thermosensitivity of hypothalamic neurons, a clear difference in the effects of PGE2 on the firing rates of VMPO neurons has been demonstrated [12]. Based on current models of set-point temperature regulation, this PGE2 dependent increase in the firing rates of temperature insensitive neurons or decrease in the firing rates of warm sensitive neurons could lead to a hyperthermic shift in the thermostatic set-point and production of a fever [3]. Yet, little is known about the electrical responses by which PGE2 regulates the firing rates of VMPO neurons. We have hypothesized that these PGE2 dependent changes in firing rate are not the result of a change in the frequency of synaptic input to these neurons, but a selective effect on specific electrical properties of VMPO neurons. To characterize these responses, whole-cell recordings were made from VMPO neurons in tissue slices from male Sprague-Dawley rats, in response to changes in temperature and PGE2.

Results

For forty two VMPO neurons, temperature sensitivity and PGE2 dependent changes in firing rate, electrical activity, and the frequency of synaptic input were determined. The majority of these neurons were classified as temperature insensitive (n = 32). The remaining ten neurons were classified as warm sensitive. With respect to thermosensitivity or responses to PGE2, there was no specific pattern to the distribution of these neurons throughout the VMPO (Fig. 1).

Figure 1
figure 1

Intracellular whole-cell recordings from VMPO Neurons. The positions of all recorded neurons in this study are indicated on the above coronal diagrams (temperature insensitive = , warm sensitive = ▲ modified from Figures 18–20, Paxinos & Watson, 1998). A corresponds to the anterior regions of VMPO, approximately 400 μm rostral to the areas shown in B, which correspond to the midpoint of the VMPO. C corresponds to the posterior regions of VMPO, approximately 400 μm caudal to the areas shown in B. 3 V, third ventricle; AVPe, anteroventral periventricular nucleus; MPA, medial preoptic area; Pe, periventricular hypothalamic nucleus; ox, optic chiasm; VMPO, ventromedial preoptic nucleus.

Using the cellularly invasive procedure of whole cell recording, the PGE2 dependent changes in firing rate that were recorded from VMPO neurons were similar to those reported in an earlier extracellular single-unit recording study [12]. In response to PGE2, fourteen temperature insensitive neurons showed significant increases in firing rate. Thirteen of these neurons had thermosensitivities ≤ 0.4 impulses·s·°C-1. The firing rates of temperature insensitive neurons, having thermosensitivities ≤ 0.4 impulses·s-1.°C-1, significantly increased in response to PGE2, from 5.75 ± 1.31 impulses·s-1 to 6.5 ± 1.31 impulses·s-1 (paired T test, P = 0.04; firing rates at least 10 minutes into the following washout period = 6.37 ± 1.32 impulses·s-1). Of the thirteen temperature insensitive neurons with thermosensitivities of 0.41 – 0.79 impulses·s-1.°C-1, the majority (n = 10) showed little or no change in firing rate in response to PGE2. The firing rates of these neurons did not significantly change from a baseline of 8.4 ±1.21 impulses·s-1 (PGE2 = 8.55 ± 1.24 impulses·s-1 (paired T test, P = 0.38); washout = 8.34 ± 1.24 impulses·s-1). In contrast to the responses of temperature insensitive neurons, the majority (n = 8) of VMPO warm sensitive neurons showed a significant decrease in firing rate during perfusion with PGE2. The firing rates of warm sensitive neurons significantly decreased from 15.28 ± 4.93 impulses·s-1 to 12.35 ± 4.85 impulses·s-1 in response to PGE2 (paired T test, P = 0.003; firing rates at least 10 minutes into the following washout period = 12.87 ± 4.12 impulses·s-1).

Electrical properties

All VMPO neurons recorded in this study had resting membrane potentials of -45.0 ± 1.1 mV (n = 42). There was no significant difference between the resting membrane potentials of temperature insensitive neurons (-45.6 ± 1.2 mV; n = 32) and warm sensitive neurons (-43.1 ± 2.4 mV; n = 10). In addition, resting membrane potential did not change in response to PGE2 and was not responsible for PGE2 dependent changes in firing rate.

The top panels of Figure 2 show the action potential activity of a temperature insensitive neuron during baseline conditions, perfusion with PGE2, and the washout period. While the resting membrane potential did not change from a baseline mean of -43.94 mV, firing rate increased 43.4% in response to PGE2, from a mean of 5.11 impulses·s-1 to 7.33 impulses·s-1. The onset of this response occurred several minutes after perfusion with PGE2 had begun and lasted approximately 15 minutes beyond the point when perfusion with PGE2 ended. This was typical of all temperature insensitive neurons which had a significant change in firing rate in response to PGE2. These neurons showed response latencies of 3.5 ± 0.69 minutes and durations that ranged from 7 to 40 minutes before firing rate returned towards the baseline level.

Figure 2
figure 2

The effects of temperature and PGE 2 on the activity of a VMPO temperature insensitive neuron and a VMPO warm sensitive neuron. The top panels show one second records of the action potential activity of a VMPO temperature insensitive neuron during all three experimental conditions (resting membrane potential = -43.94 mV). The firing rate significantly increased from a baseline of 5.11 ± 0.12 impulses·s-1 to 7.33 ± 0.13 impulses·s-1 during perfusion with 1 μM PGE2 (washout = 4.90 ± 0.14 impulses·s-1). The lower panels show one second records of the action potential activity of a VMPO warm sensitive neuron during all three experimental conditions (resting membrane potential = -51.62 mV). The firing rate significantly decreased from a baseline of 10.07 ± 0.13 impulses·s-1 to 8.40 ± 0.12 impulses·s-1 during perfusion with 1 μM PGE2 (washout = 10.20 ± 0.16 impulses·s-1).

The lower panels of Figure 2 show the action potential activity of a warm sensitive neuron during baseline conditions, perfusion with PGE2, and the washout period. While the resting membrane potential did not change from a baseline mean of -51.62 mV, the firing rate of this neuron decreased in response to PGE2, from a mean of 10.07 impulses·s-1 to 8.40 impulses·s-1. The onset of this response occurred several minutes after perfusion with PGE2 had begun and lasted approximately 25 minutes beyond the point when perfusion with PGE2 was stopped. Similar changes were recorded for the other warm sensitive neurons that were inhibited by PGE2, which showed response latencies of 3.90 ± 0.95 minutes and durations that ranged from 10 to 25 minutes.

Throughout the entire length of a recording (46.7 ± 1.8 minutes), the amplitude and duration of action potentials recorded from VMPO neurons slowly decreased by an average of 4.6 mV and 0.16 milliseconds, respectively (n = 42). It is presumed that this was due to minor changes in ionic gradients, resulting from the technique of whole-cell recording. Based on the PGE2 dependent changes in firing rate reported in this study and previous extracellular recordings [12], these small changes in action potential amplitude and duration did not affect the ability of VMPO neurons to respond to PGE2. There were also no significant changes in the amplitudes or durations of action potentials recorded from temperature insensitive and warm sensitive neurons in response to PGE2.

The input resistance of VMPO neurons recorded in this study significantly decreased in response to PGE2 from 367.1 ± 24.5 MΩ. to 339.7 ± 23.7 MΩ (n = 42; paired T test, P = 0.0001). This PGE2 dependent decrease in resistance was independent of thermosensitivity. Figure 3 shows current-voltage plots from a temperature insensitive neuron (Fig. 3A) and a warm sensitive neuron (Fig. 3B). For both of these neurons, input resistance decreased similarly in response to PGE2, with a reversal potential at or near the resting membrane potential.

Figure 3
figure 3

The effects of PGE 2 on the input resistance of a VMPO temperature insensitive neuron (A) and a VMPO warm sensitive neuron (B). For each, changes in membrane potential are plotted in response to a series of hyperpolarizing current pulses (-10 to -100 pA, 200 ms), during baseline conditions (), perfusion with PGE2 (▲), and after PGE2 has been washed out of the chamber (■). For the temperature insensitive neuron in A (m = 0.05), input resistance decreased from a baseline of 406.33 MΩ, to 332.40 MΩ during perfusion with PGE2. After PGE2 had been washed out of the chamber, input resistance increased to 372.32 MΩ. For the warm sensitive neuron in B (m = 1.30), input resistance decreased from a baseline of 388.00 MΩ, to 244.86 MΩ during perfusion with PGE2. After PGE2 had been washed out of the chamber, input resistance increased to 327.63 MΩ.

During the application of a depolarizing current, VMPO neurons did not show PGE2 dependent changes in the frequency of action potentials (frequency response). However, the majority of VMPO neurons did show PGE2 dependent changes in another characteristic of neuronal excitability, the first spike latency. Temperature insensitive neurons with thermosensitivities ≤ 0.4 impulses·s-1.°C-1showed a significant PGE2 dependent decrease in the first spike latency, from 7.4 ± 1.2 ms during baseline conditions to 5.2 ± 0.6 ms in response to PGE2 (n = 18; paired T test, P = 0.008; Figure 4). The first spike latency for these temperature insensitive neurons returned to baseline levels during the washout period (7.6 ± 1.4 ms). In contrast, warm sensitive neurons showed a significant increase in this latency response during perfusion with PGE2(n = 7), similar to the response of the warm sensitive neuron in Figure 5. (baseline = 4.9 ± 0.5 ms; PGE2 = 7.9 ± 1.1 ms (paired T test, P = 0.046); washout = 4.2 ± 0.5 ms). Temperature insensitive neurons with thermal coefficients of 0.41 to 0.79 impulses·s-1.°C-1 showed little change in this measurement of neuronal excitability (n = 8), in response to PGE2 (baseline = 7.0 ± 1.3 ms; PGE2 = 6.8 ± 1.1 ms (paired T test, P = 0.69); washout = 6.8 ± 1.1 ms).

Figure 4
figure 4

The effect of depolarizing current and PGE 2 on the activity of a VMPO temperature insensitive neuron. A shows the initial action potential that was produced during baseline conditions, in response to a depolarizing current injection (50 pA, 500 msec). The onset of the current is indicated on the line at the bottom of the figure and only the initial 34 mSec are shown. The first spike latency was 8.0 mSec. In B, the initial action potential that was produced in response to a depolarizing current injection is shown during perfusion with PGE2 (1 μM). The first spike latency was 3.0 mSec.

Figure 5
figure 5

The effect of depolarizing current and PGE 2 on the activity of a VMPO warm sensitive neuron. A shows the initial action potential that was produced during baseline conditions, in response to a depolarizing current injection (50 pA, 500 msec). The onset of the current is indicated on the line at the bottom of the figure and only the initial 22.0 mSec are shown. The first spike latency was 4.0 mSec. In B, the initial action potential that was produced in response to a depolarizing current injection is shown during perfusion with PGE2 (200 nM). The first spike latency was 6.6 mSec.

In Figure 6, the averaged pre- and post-spike activity of a temperature insensitive neuron (A) and a warm sensitive neuron (B) are shown during baseline conditions and in response to PGE2. In a similar manner to the neuron shown in Figure 6A, the majority of temperature insensitive neurons with thermal coefficients ≤ 0.4 impulses·s-1.°C-1 showed a significant increase in the rate of rise of the depolarizing prepotential in response to PGE2 (n = 12; baseline = 0.46 ± 0.04 mV·ms-1, PGE2 = 0.55 ± 0.05 mV·ms-1; paired T test, P = 0.01). In contrast, the majority of warm sensitive neurons showed a significant decrease in the rate of rise of the depolarizing prepotential in response to PGE2, similar to the response of the neuron in Figure 6B (n = 8; baseline= 0.61 ± 0.08 mV·ms-1, PGE2 = 0.50 ± 0.08 mV·ms-1; paired T test, P = 0.012). Although small changes in the voltage deflections that occurred at the end of an action potential showed some degree of change in response to PGE2, these changes were inconsistent in both temperature insensitive or warm sensitive neurons. Temperature insensitive neurons with thermosensitivities of 0.41 – 0.79 impulses·s-1.°C-1 did not show a significant change in the rate of rise of the depolarizing prepotential in response to PGE2.

Figure 6
figure 6

Averaged pre- and post-spike activity of a temperature insensitive VMPO neuron (A) and warm sensitive VMPO neuron (B). For each, averages of 10 action potentials (truncated) during baseline conditions and during perfusion with PGE2 (1 μM), are superimposed on the spike threshold (post-pike activity during PGE2 is indicated by an asterisk). These averages do not include pre-spike activity that contained putative postsynaptic potentials. For the temperature insensitive neuron in A, the rate of rise of the depolarizing prepotential increased from 0.17 mV·mSec-1 to 0.27 mV·mSec-1. For the warm sensitive neurons in B, the rate of rise of the depolarizing prepotential decreased from 0.38 mV·mSec-1 to 0.18 mV·mSec-1.

Synaptic input

The frequency of post synaptic potentials (PSPs) recorded from VMPO neurons in the localized environment of coronal tissue slices was predominately inhibitory and insensitive to changes in temperature (Table 1). In 92% of the recordings, temperature had little or no effect on the frequencies of either inhibitory post synaptic potentials (IPSPs; m = 0.25 ± 0.06 PSPs·s-1.°C-1) or excitatory post synaptic potentials (EPSPs; m = 0.23 ± 0.06 PSPs·s-1.°C-1). In response to PGE2, the frequency of synaptic input recorded from VMPO neurons did not change (Table 1).

Table 1 Effect of PGE2on the Frequency of IPSPs and EPSPs recorded from VMPO Neurons.

Discussion

PGE2dependent changes in the firing rates of VMPO neurons

As current models for temperature regulation suggest, the integrated responses of hypothalamic warm sensitive neurons and temperature insensitive neurons play an important role achieving and maintaining a discrete set-point for temperature control [2]. While their importance in thermoregulatory pathways has yet to be determined, in vitro recordings from tissue slices have also identified neurons in the PO/AH that lack spontaneously generated activity (silent neurons), are predominantly driven to produce action potentials by synaptic input (EPSP-driven neurons), or produce action potentials in a bursting pattern [2, 15]. However, only prepotential driven warm sensitive and temperature insensitive neurons were identified in the VMPO.

In the generation of a fever, current thermoregulatory models of neural networking suggest that either an increase in the activity of temperature insensitive neurons or the inhibition of warm sensitive neurons would shift the set-point to a more hyperthermic temperature [2]. Through competing synaptic inputs, either of these responses would alter the temperature at which thermoeffector neurons begin to activate thermoregulatory responses. However, previous studies have not been able to show a correlation between the thermosensitivity of hypothalamic neurons and firing rate responses to PGE2 or other endogenous pyrogens [1619]. This may be the result of the various criteria that were used to define thermosensitivity in many of these studies, as well as recordings from more general areas of the PO/AH. In a recent study of the extracellular single-unit activity of neurons in the VMPO, a clear thermal distinction in firing rate responses to PGE2 was reported, with temperature insensitive neurons excited by PGE2 and warm sensitive neurons inhibited [12]. In this previous study, a functionally significant criterion was used to define warm sensitivity (m ≥ 0.8 impulses·s-1.°C-1) [3, 15]. As this criterion provides a method for identifying integrative warm sensitive neurons in the tissue slice preparation, it allows the experimental findings to have a functional significance in modeling thermoregulation in vivo. In the present study, which uses whole-cell recording techniques to record the intracellular properties of VMPO neurons and the same functional criterion for defining warm sensitivity, similar firing rate responses were recorded. As suggested above, either an increase in the firing rates of temperature insensitive neurons or a decrease in the firing rates of warm sensitive neurons could lead to a hyperthermic shift in the thermostatic set-point and a fever.

It is also important to note that while the majority of PGE2 responsive neurons showed a significant recovery in firing rate during the washout period, some did not. This lack of complete recovery may have resulted from responses to PGE2 that were considerably long in duration, lasting up to 40 minutes, and minor changes in ionic gradients that can occur over time during whole-cell recordings. As mentioned in the methods, all recordings were closely monitored for any changes that would make them unacceptable.

Cellular properties of VMPO neurons and responses to PGE2

In addition to the direct effects of PGE2 on the electrical properties of VMPO neurons, the frequency of localized synaptic input was characterized. In the present study, we were specifically interested in monitoring firing rate responses to PGE2 and characterizing the resulting voltage potential changes (i.e., action potentials and membrane potentials). In order to ensure that we did not interrupt the firing rate responses, we did not perform any voltage clamp measurements and did all recordings in current clamp, with current = 0 pA. This prevented any detailed characterization of synaptic potentials, other than frequency. As has been shown in other regions of the PO/AH [15], the majority of this synaptic input to VMPO neurons was inhibitory and predominantly insensitive to temperature. In response to PGE2, the frequency of PSPs recorded from most VMPO neurons either did not change or showed an increase. However, PGE2 dependent increases in the frequency of PSPs were not recorded from neurons that showed PGE2 dependent changes in firing rate. This may have resulted from a smaller degree of synaptic influence on firing rate, due to a change in the input resistance [20]. Additionally, the lack of local synaptic input from warm sensitive neurons would suggest that the axonal projections of these neurons may not terminate within the VMPO, but form efferent projections to other hypothalamic nuclei such as the paraventricular nucleus.

As shown in Figure 3, the input resistance recorded from most of the VMPO neurons decreased in response to PGE2, regardless of temperature sensitivity or PGE2 dependent changes in firing rate. Since this response had little or no influence on firing rate, we would suggest that decreases in resistance resulted from changes in the activity of multiple currents, which continued to maintain equilibrium at or near the resting membrane potential. While we have also suggested that PGE2 may cause cAMP concentrations to either increase or decrease within these neurons [12, 21], the net effect on hyperpolarizing K+ currents and depolarizing Na+ or Ca++ currents may still lead to an increase in conductance that does not result in a change in resting membrane potential. Therefore, although these changes in the input resistance may be an important response to PGE2, they are not directly responsible for PGE2 dependent changes in firing rate.

In response to a depolarizing current pulse, PGE2 dependent changes in the excitability of VMPO neurons were recorded. Using two primary measurements of excitability, the frequency response and first spike latency (see: Results & Methods), our recordings indicate that PGE2 selectively increased the excitability of temperature insensitive neurons, while decreasing the excitability of warm sensitive neurons. Although there was no significant PGE2 dependent changes in the frequency responses of VMPO neurons, temperature insensitive neurons showed a significant decrease in first spike latency in response to PGE2 (Fig. 4), while warm sensitive neurons showed a significant increase (Fig. 5). This would suggest that PGE2 is having a direct effect of the activity of transient voltage dependent currents, which are responsible for regulating the excitability and rhythmic firing rates of these neurons. While these current pulses were not matched so that they would result in the same level of depolarization, making it difficult to discern which voltage-gated conductances may be responding to PGE2, this data does provide support for our finding of PGE2 dependent changes in the prepotential [22].

Several transient voltage dependent currents have been identified in hypothalamic neurons which regulate excitability and rhythmic firing rate activity. These include a slow component sodium current and a calcium dependent potassium current (IK, Ca) in suprachiasmatic neurons [23], a low voltage activated calcium current and a IK, Ca current in posterior hypothalamic neurons [24], a non-inactivating potassium current in neurhypophyseal nerve terminals [25], and a IA current in PO/AH neurons [22]. Although any or all of these currents may be present in VMPO neurons, only the IA has been implicated in temperature dependent changes in firing rate, through a mechanism in which the thermally dependent inactivation rate of this current influences the rate of rise of the prepotential that precedes each action potential [22]. In a similar manner, our data suggests that PGE2 has a direct effect on the firing rates of VMPO neurons through changes in the prepotential, increasing the rate of rise in temperature insensitive neurons, while decreasing the rate of rise of the prepotential in warm sensitive neurons (Fig. 6). Therefore, PGE2 dependent changes in the firing rates of VMPO neurons may also depend on an ability to influence the inactivation rate of the IA type current.

Within the VMPO, overlapping expression of EP3 and EP4 receptors may provide PGE2 with the ability to selectively affect the activity of neurons in this region [26]. Activation of either receptor subtype is known to influence cellular activity through the regulation of intracellular cAMP concentrations, with EP4 activation leading to an increase in cAMP and EP3 activation leading to a decrease [2628]. Support for this mechanism is provided by evidence showing that cAMP plays a role in thermoregulation and more specifically, in the generation of a fever [29]. With respect to excitability and rhythmic firing rate activity, several studies have demonstrated that cAMP modulates PGE2 dependent changes the activity of certain potassium currents, including the IA [25, 30, 31]. Therefore, selective activation of either EP3 or EP4 receptors may be responsible for the changes in firing rate and electrical activity we have recorded from VMPO neurons.

Conclusion

In response to intravenous LPS, the local production of PGE2 within the PO/AH region results in a thermally dependent change in the firing rates of VMPO neurons. Through a direct effect on the rate of rise of the depolarizing prepotential, which is also a determinant of thermosensitivity, PGE2 increases the firing rates of temperature insensitive neurons, while decreasing the firing rates of warm sensitive neurons in the VMPO. While the results of this study provide a clear focus for future studies into the conductance mechanisms of these responses, it also supports a functional and important role for the VMPO region and both groups of thermally classified neurons in the production of this type of fever.

Methods

Anterior hypothalamic tissue slices containing the VMPO were prepared from male Sprague-Dawley rats (100 – 150 grams), which were housed under standard conditions and given food and water ad lib. Prior to each recording session, an animal was anesthetized (isoflurane) and sacrificed by quick decapitation, according to procedures approved by the National Science Foundation and the Animal Care and Use Committee of the College of William and Mary. Following removal of the brain, a tissue block of the hypothalamus was cut using a vibratome into 400 μm thick coronal slices. Two or three slices containing the VMPO were then placed in a recording chamber and allowed to equilibrate for 1–2 hours. Tissue slices were continually perfused with pyrogen free artificial cerebral spinal fluid (aCSF), which consisted of (in mM), 124 NaCl, 26 NaHCO3, 10 glucose, 5 KCl, 2.4 CaCl2, 1.3 MgSO4, and 1.24 KH2PO4. This nutrient medium was oxygenated (95% O2 – 5% CO2), warmed to a stable temperature of ~36°C by a thermoelectric assembly, and allowed to gravity flow into the chamber at 1 – 1.5 ml· min-1 (chamber volume = 1.2 ml) [32]. A small thermocouple, positioned in the recording chamber just below the tissue slices, was used to continuously monitor tissue temperature.

Tight-seal whole-cell recordings were made using glass microelectrodes with tip inner diameters of ~2 μm (3–5 MΩ), filled with a solution that consisted of (in mM) 130 K-gluconate, 10 EGTA, 2 ATP, 1 MgCl2, 1 CaCl, having a pH of 7.2 and an osmolarity of approximately 295 mOsmols/liter. As described previously [33], a liquid junction potential of 12.0 mV was subtracted from all recorded potentials. All recordings were made using an integrated patch-clamp amplifier (Axopatch 200B, Axon Instruments) and along with temperature, were stored on digital tape for later analysis. Firing rate was continuously recorded as a voltage measurement from a rate/interval monitor and was determined by action potential triggered input from a window discriminator (FHC Inc.). Specific stimulation protocols to measure input resistance and excitability were generated by a computer that was interfaced with the amplifier (pClamp software, Axon Instruments). Acceptable recordings consisted of action potential amplitudes through 0 mV and stable recordings of at least 20 minutes.

Using a stereomicroscope, the recording electrode was positioned in the VMPO, which encompasses 700 μm of the hypothalamus just rostral to the suprachiasmatic nucleus. It extends laterally 900 μm from the third ventricle and dorsally 500 μm from the ventral brain surface [34]. Once a tight seal (> 2 GΩ) was achieved between the electrode and the surface of a neuron, the cell membrane was ruptured by suction, establishing an intracellular recording. When the activity of a neuron was stable for several minutes, temperature in the recording chamber was varied 2–3°C above and below 36°C, by changing the input voltage to the thermoelectric assembly. Neuronal thermosensitivity (impulses· s-1·°C-1) was characterized by plotting firing rate as a function of temperature to determine the regression coefficient (m) of this plot. As in previous studies [12, 15, 35], warm sensitivity was defined as a regression coefficient of at least 0.8 impulses· s-1·°C-1. All other neurons in this study were defined as temperature insensitive.

After thermosensitivity had been characterized, each neuron was tested for its response to PGE2. At a stable temperature (~36°C), the perfusion medium was switched to one containing PGE2 (200 nM or 1 μM, Sigma Chemical Co.). The duration of exposure to PGE2 ranged from 5 – 15 minutes, with durations of less than 10 minutes occurring only when there was a clear indication of a response. Exposure to PGE2 was followed by perfusion with aCSF for a washout period of at least 10 minutes.

To determine if PGE2 had a significant effect on firing rate, one-minute segments of stable activity were digitized for comparison (60 Hz; pClamp Software, Axon Instruments). These segments were collected during baseline conditions (just prior to perfusion with PGE2) at the end of perfusion with PGE2, and at the end of a 10-minute washout period (or when firing rate returned to baseline levels). For each segment, a mean and standard error were calculated (Sigmaplot software, SPSS Inc.). A significant response to PGE2 was determined by comparison to the baseline level using a standard T-test (P ≤ 0.05).

Changes in resting membrane potential and the action potential waveform in response to PGE2 were also characterized. As in a previous study [32], the resting membrane potential was continuously recorded. After rapid changes in membrane potential (including action potentials) were filtered out (half amplitude response setting at 0.5 Hz), one minute segments of voltage activity were digitized (60 Hz) during baseline conditions, perfusion with PGE2, and the washout period. To characterize changes in the action potential, an averaged action potential from ten inherently activated action potentials was produced for each of the experimental conditions (sampling rate = 66.7 kHz). Measurements were made of action potential amplitude (resting membrane potential to peak) and duration (at one half peak). Measurements were also made of the rate of rise of the depolarizing prepotential that precedes each action potential [22]. For resting membrane potential and action potential measurements, significant responses to PGE2 were determined by comparison to baseline using a standard T-test (P ≤ 0.05).

Throughout each recording, input resistance and excitability were measured every 5 – 10 minutes to determine changes in response to PGE2 and to insure that recordings remained stable. Large changes that did not return towards baseline levels were a characteristic of a deteriorating recording and marked the end of legitimate data. Input resistance was determined by the slope of a current-voltage plot obtained from a computer generated protocol in which a series of ten current pulses (-10 to -100 pA) were administered [32]. In addition, two primary characteristics of neuronal excitability were measured in response to a depolarizing current [36]. The frequency response was determined by the number of action potentials produced during the depolarizing current. The first spike latency was also measured and was defined as the duration from the start of the depolarizing current to the peak of the first action potential. All significant responses to PGE2 were again determined by comparison to baseline using a standard T-test (P ≤ 0.05).

The frequencies of synaptic potentials (EPSPs and IPSPs) recorded from each VMPO neuron were determined during baseline conditions, perfusion with PGE2, and the washout period. As in previous whole-cell recordings, individual potentials were identified as rapid changes in membrane potential of at least 1 mV greater than background noise [15, 20]. For each experimental condition, blind counts of PSPs were made over a 20 second duration to produce frequency averages (PSPs·s-1). As with the other measurements, all significant responses to PGE2 were determined by comparison to baseline using a standard T-test (P ≤ 0.05). In addition, the thermosensitivities of EPSPs and IPSPs were characterized by calculating frequency averages (using the same methods as detailed above) at different temperatures (≥ 3°C range) and plotting the results as a function of temperature [15].

Once a recording had been completed, a stereomicroscope was used to visually confirm the location of the recording electrode. The ventral edge of the third ventricle was used as a reference to determine the lateral-medial and dorsal-ventral coordinates. The coronal position was also specified by the depth of the electrode from the surface of the tissue slice. In addition, the side of the brain from which the recording was made was identified by preparing the tissue slices with more lateral tissue on the left. Tissue slices were then removed from the recording chamber, fixed in a 10% formalin solution, and sectioned again to a thickness of 50 μm. Sections were then stained with giemsa to identify specific hypothalamic areas so that the location of each recording within VMPO could be reconfirmed [12] [38].

References

  1. Saper CB, Breder CD: The neurological basis of fever. N Eng J Med. 1994, 330: 1880-1886. 10.1056/NEJM199406303302609.

    Article  CAS  Google Scholar 

  2. Boulant JA: Hypothalamic neurons: Mechanism of sensitivity to temperature. Ann N Y Acad Sci. 1998, 856: 108-115.

    Article  CAS  PubMed  Google Scholar 

  3. Boulant JA: Thermoregulation In Fever: Basic Mechanisms and Management. 1991, New York: Raven Press, 1-22.

    Google Scholar 

  4. Boulant JA, Hardy JD: The effect of spinal and skin temperatures on the firing rate activity and thermosensitivity of preoptic neurons. J Physiol. 1974, 240: 639-660.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  5. Blatties CM, Sehic E, Li S: Pyrogen sensing and signaling: Old views and new concepts. Clinical Infectious Disease. 2000, 31: 168-177. 10.1086/317522.

    Article  Google Scholar 

  6. Elmquist JK, Scammell TE, Saper CB: Mechanisms of CNS response to systemic immune challenge; the febrile response. Trends in Neuroscience. 1997, 20: 565-570. 10.1016/S0166-2236(97)01138-7.

    Article  CAS  Google Scholar 

  7. Blatteis CM, Bealer WS, Hunter J, Llanos-Q RA, Mashburn TA: Suppression of fever after lesions of the anteroventral third ventricle in guinea pigs. Brain Res Bull. 1983, 11: 519-526. 10.1016/0361-9230(83)90124-7.

    Article  CAS  PubMed  Google Scholar 

  8. Feldberg W, Saxena P: Further studies on prostaglandin E1 fever in cats. J Physiol. 1971, 219: 739-745.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  9. Stitt JT: Prostaglandin E1 fever in rabbits. J Physiol. 1973, 232: 163-179.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  10. Stitt JT: Differential sensitivity in the sites of fever production by prostaglandin E1 within the hypothalamus of the rat. J Physiol. 1991, 432: 99-110.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Scammell TE, Griffin JD, Elmquist JK, Saper CB: Microinjection of a cyclooxygenase inhibitor into the anteroventral preoptic region attenuates LPS fever. Am J Physiol. 1998, 274 (3 Pt 2): R783-789.

    CAS  PubMed  Google Scholar 

  12. Ranels HJ, Griffin JD: The effects of prostaglandin E2 on the firing rate activity of thermosensitive and temperature insensitive neurons in the ventromedial preoptic area of the rat hypothalamus. Brain Res. 2003, 964: 42-50. 10.1016/S0006-8993(02)04063-5.

    Article  CAS  PubMed  Google Scholar 

  13. Elmquist JK, Scammell TE, Jacobson CD, Saper CB: Distribution of Fos-like immunoreactivity in the rat brain following intravenous lippopolysaccharide administration. J Comp Neurology. 1996, 371: 1-19. 10.1002/cne.903710102.

    Article  Google Scholar 

  14. Scammell TE, Elmquist JK, Griffin JD, Saper CB: Ventromedial preoptic prostaglandin E2 activates fever-producing autonomic pathways. J Neurosci. 1996, 16: 6246-6254.

    CAS  PubMed  Google Scholar 

  15. Griffin JD, Saper CB, Boulant JA: Synaptic and morphological characteristics of temperature sensitive and insensitive rat hypothalamic neurons. J Physiol. 2001, 537: 521-535. 10.1111/j.1469-7793.2001.00521.x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Hori T, Shibata M, Nakashima M, Yakashima M, Asami A, Asami T, Koga H: Effects of interleukin-1 and arachiodonate on the preoptic and anterior hypothalamic neurons. Brain Res Bull. 1988, 20: 75-82. 10.1016/0361-9230(88)90010-X.

    Article  CAS  PubMed  Google Scholar 

  17. Matsuda T, Hori T, Nakashima T: Thermal and PGE2 sensitivity of the organum vasculosum lamina terminalis region and preoptic area in rat brain slices. J Physiol. 1992, 454: 197-212.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  18. Morimoto A, Murakami N, Watanabe T: Effect of prostaglandin E2 on thermoresponsive neurons in the preoptic and ventromedial hypothalamic regions of rats. J Physiol. 1988, 405: 713-725.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Nakashima T, Hori T, Kuriyama K, Matsuda T: Effects of interferon-α on the activity of preoptic thermosensitive neurons in tissue slices. Brain Res. 1988, 454: 361-367. 10.1016/0006-8993(88)90838-4.

    Article  CAS  PubMed  Google Scholar 

  20. Burgoon PW, Boulant JA: Temperature sensitive properties of rat suprachiasmatic nucleus neurons. Am J Physiol Regul Integr Comp Physiol. 2001, 281: R706-715.

    CAS  PubMed  Google Scholar 

  21. Thompson ED, Griffin JD: The effects of cAMP dependent protein kinase activation on the firing rates of thermosensitive and temperature insensitive neurons in the ventromedial preoptic area of the rat hypothalamus [Abstract]. FASEB J. 2003, A28.

    Google Scholar 

  22. Griffin JD, Kaple ML, Chow AR, Boulant JA: Cellular mechanisms for neuronal thermosensitivity in the rat hypothalamus. J Physiol. 1996, 492: 231-242.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  23. Pennartz CMA, Bos NPA, De Jeu MTG, Geurtsen AMS, Mirmiran M, Sluiter AA, Buijs RM: Membrane properties and morphology of vasopressin neurons in slices of rat suprachiasmatic nucleus. J Neurophysiol. 1998, 80: 2710-2717.

    CAS  PubMed  Google Scholar 

  24. Fan Y-P, Horn EM, Waldrop TG: Biophysical characterization of rat caudal hypothalamic neurons; calcium channel contribution to excitability. J Neurophysiol. 2000, 84: 2896-2903.

    CAS  PubMed  Google Scholar 

  25. Zhang L, Karpinski E, Benishin CG: Prostaglandin E2 modulates a non-inactivation potassium current in rat neurohypophyseal nerve terminals. Neurochem Intern. 1999, 35: 345-355. 10.1016/S0197-0186(99)00073-X.

    Article  CAS  Google Scholar 

  26. Oka T, Oka K, Scammell TE, Lee C, Kelly JF, Nantel F, Elmquist JK, Saper CB: Relationship of EP1–4 prostaglandin receptors with rat hypothalamic cell groups involved in lippolysaccharide fever response fever response. J Comp Neurology. 2000, 428: 20-32. 10.1002/1096-9861(20001204)428:1<20::AID-CNE3>3.0.CO;2-X.

    Article  CAS  Google Scholar 

  27. Ek M, Arias C, Sawchenko P, Ericsson-Dahlstrand A: Distribution of the EP3 prostaglandin E2 receptor subtypes in the rat brain: Relationship to sites of Interleukin-1 induced cellular responsiveness. J Comp Neurology. 2000, 428: 5-20. 10.1002/1096-9861(20001204)428:1<5::AID-CNE2>3.0.CO;2-M.

    Article  CAS  Google Scholar 

  28. Oka T, Aou S, Hori T: Intracerebroventricular injection of prostaglandin E2 induces thermal hyperalgesia in rats: the possible involvement of EP3 receptors. Brain Res. 1994, 663: 287-292. 10.1016/0006-8993(94)91275-0.

    Article  CAS  PubMed  Google Scholar 

  29. Steiner AA, Branco LGS: Fever and anapyrexia in systemic inflammation: intracellular signaling by cyclic nucleotides. Frontiers in Bioscience. 2003, 8: s1398-1408.

    Article  CAS  PubMed  Google Scholar 

  30. Evans AR, Vasko MR, Nicol GD: The cAMP transduction cascade mediates the PGE2 induced inhibition of potassium currents in rat sensory neurons. J Physiol. 1999, 516: 163-168. 10.1111/j.1469-7793.1999.163aa.x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  31. Lopshire JC, Nicol GD: The cAMP transduction cascade mediates the PGE2 enhancement of the capsaicin-elicited current in rat sensory neurons: whole cell and single unit studies. J Neurosci. 1998, 18: 6081-6092.

    CAS  PubMed  Google Scholar 

  32. Kelso SR, Nelson DO, Silva NL, Boulant JA: A slice chamber for intracellular and extracellular recording during continuous perfusion. Brain Res Bull. 1983, 10: 853-857. 10.1016/0361-9230(83)90219-8.

    Article  CAS  PubMed  Google Scholar 

  33. Griffin JD, Boulant JA: Temperature effects on membrane potential and input resistance in rat hypothalamic neurones. J Physiol. 1995, 488: 407-418.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Paxinos G, Watson C: The Rat Brain In Stereotaxic Coordinates. 1998, New York: Academic Press

    Google Scholar 

  35. Kelso SR, Perlmutter MN, Boulant JA: Thermosensitive single-unit activity in vitro hypothalamic neurons. Am J Physiol. 1982, 242: R77-84.

    CAS  PubMed  Google Scholar 

  36. Russier M, Carlier E, Ankri N, Fronzaroli L, Debanne D: A-, T-, and H-type currents shape intrinsic firing of developing rat abducens motoneurons. J Physiol. 2003, 549 (Pt 1): 21-36. 10.1113/jphysiol.2002.037069. DOI: 10.1113/jphysiol.2002.037069

    Article  PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We want to thank Erin Waller, Sarah Norcross and Elizabeth Wallis for their assistance in the histological processing of neural tissue. This work was supported by NSF grant IBN-9983624, and in part by a Howard Hughes Medical Institute grant through the Undergraduate Biological Sciences Program to the College of William and Mary.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to John D Griffin.

Additional information

Authors' contributions

HJR, carried out the majority of cellular recordings and data analysis. JDG conceived of the study and participated in its design, coordination and completion. Both authors contributed equally to the drafting of this manuscript.

Authors’ original submitted files for images

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ranels, H.J., Griffin, J.D. Effects of prostaglandin E2 on the electrical properties of thermally classified neurons in the ventromedial preoptic area of the rat hypothalamus. BMC Neurosci 6, 14 (2005). https://doi.org/10.1186/1471-2202-6-14

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2202-6-14

Keywords