You are on page 1of 10

Epilepsy Research 145 (2018) 63–72

Contents lists available at ScienceDirect

Epilepsy Research
journal homepage: www.elsevier.com/locate/epilepsyres

Carbamazepine-induced suppression of repetitive firing in CA1 pyramidal T


neurons is greater in the dorsal hippocampus than the ventral hippocampus

Madeline Collette Evans, Kelly Ann Dougherty
Department of Biology, Rhodes College, 2000 N Parkway, Memphis, TN, 38112, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Medial temporal lobe epilepsy (mTLE)–the most common form of focal epilepsy–is defined by recurrent partial
Carbamazepine seizures originating within the medial temporal lobe. Such seizures are commonly associated with the anterior
Dorsal hippocampus hippocampus (as opposed to the posterior hippocampus), and refractory to the currently available anti-epileptic
Intrinsic excitability drugs (AED) for about one third of patients. Unfortunately, the mechanisms driving seizure generation and AED
Anti-Epileptic drug
efficacy along the longitudinal hippocampal axis remain poorly understood. Recently, several groups in-
CA1 pyramidal neuron
vestigating differences in excitability along the rodent longitudinal hippocampal axis have demonstrated that
CA1 pyramidal neurons from the rodent ventral hippocampus (the rodent homolog of the human anterior
hippocampus) are intrinsically more excitable than their dorsal counterparts (the rodent homolog of the human
posterior hippocampus). This phenotypic difference is accompanied by significant differences in gene expression
along the longitudinal hippocampal axis, which include gene products–such as voltage-gated sodium channel β-
subunits–known to influence AED efficacy. Given this phenotypic heterogeneity, and the differential expression
of gene products known to influence anti-epileptic drug efficacy, we sought to investigate the efficacy of the
classical use-dependent sodium channel blocker, carbamazepine, in CA1 pyramidal neurons across the long-
itudinal hippocampal axis. Accordingly, we performed whole-cell current-clamp recordings on CA1 pyramidal
neurons from acute hippocampal slices prepared from the dorsal and ventral hippocampus, and found that acute
exposure to 100 μM carbamazepine induced a significantly greater suppression of repetitive firing for dorsal
neurons relative to ventral neurons by inducing profound spike frequency adaptation (SFA). Moreover, we
observed a small, but significant depolarization of resting membrane potential (RMP) for dorsal neurons (but not
ventral neurons), following exposure to carbamazepine. Together, these observations demonstrate that carba-
mazepine’s effect is concentrated in the dorsal hippocampus, which could provide meaningful insight into the
side effect profile of carbamazepine (and related anti-epileptic drugs) in non-epileptic tissue, and inform future
work investigating the mechanisms of carbamazepine resistance in epileptic tissue.

1. Introduction seizures in about one-third of mTLE patients, thereby necessitating the


surgical resection of the anterior hippocampus (Kwan and Sander,
Medial temporal lobe epilepsy (mTLE)–the most common form of 2004). Although the specific mechanisms driving the hyperexcitability
focal epilepsy–is defined by partial seizures that spontaneously origi- of the human anterior hippocampus remain poorly understood, recent
nate within the medial temporal lobe. Specifically, seizures often ori- work exploring the rodent ventral hippocampus (the rodent homolog of
ginate in the anterior region of the hippocampal formation before the human anterior hippocampus) has begun to yield insight into the
spreading to other regions of the hippocampus and brain, which is as- special excitability of this brain region (Dougherty et al., 2012;
sociated with cell death and hippocampal sclerosis within the anterior Hönigsperger et al., 2015; Marcelin et al., 2012; Kim and Johnston,
hippocampus (Engel et al., 1990; Babb et al., 1984). Unfortunately, 2015; Malik and Johnston, 2017). Accordingly, several groups have
currently available anti-epileptic drugs (AED) do not adequately control identified significant heterogeneity in anatomy, synaptic connectivity,

Abbreviations: mTLE, medial temporal lobe epilepsy; AED, anti-epileptic drug; DHC, dorsal hippocampus; dIHC, dorsal intermediate hippocampus; vIHC, ventral intermediate hip-
pocampus; VHC, ventral hippocampus; ISA, somatodendritic A-type current; IM, m-current; GIRK, channel, G-protein coupled inwardly recitifying potassium channel; CBZ, carbamazepine;
aCSF, artificial cerebrospinal fluid; DNQX, 6,7-dinitroquinoxaline 2,3-dione; DMSO, dimethyl sulfoxide; Rin, input resistance; τM, membrane time constant; Vthr, threshold voltage; APPeak,
action potential peak; APsize, action potential size; APhw, action potential half width; ISI, interspike interval; SFA, spike frequency adaptation; PTSD, post-traumatic stress disorder; FPKM,
fragments per kilobase of exon per million fragments mapped

Corresponding author at: 2000 N Parkway, Memphis, TN, 38112, USA.
E-mail address: doughertyk@rhodes.edu (K.A. Dougherty).

https://doi.org/10.1016/j.eplepsyres.2018.05.014
Received 7 March 2018; Received in revised form 11 May 2018; Accepted 29 May 2018
Available online 09 June 2018
0920-1211/ © 2018 Elsevier B.V. All rights reserved.
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

network-level phenomena, cellular morphology, cellular physiology, Table 1


and gene expression along the longitudinal hippocampal axis (dor- Expression of neuronal voltage-gated sodium channel α- and β-subunits across
sal–ventral axis; (Strange et al., 2014). An emerging theme from this the longitudinal hippocampal axis.
work is that the trisynaptic circuits of the dorsal hippocampus (DHC) Sodium Channel HippoSeq Gene CA1 DHC (FPKM) CA1 VHC (FPKM)
and ventral hippocampus (VHC) show quantitative differences, despite or β-subunit ID
their qualitative similarity. This “tuning” of the trisynaptic circuit likely
Nav1.1 Scn1a 46.8 ± 24.5 33.3 ± 20.6
underlies the hyperexcitability of the VHC, when compared with the
Nav1.2 Scn2a1 111.5 ± 41.5 136.2 ± 45.0
DHC. Nav1.3 Scn3a 11.6 ± 4.2 10.1 ± 4.2
Recently, a more nuanced picture of these dorsal and ventral phe- Nav1.6 Scn8a 104.8 ± 22.2 84.2 ± 18.3
notypes has begun to emerge, as several groups have reported sig- β1 Scn1b 81.7 ± 25.8a 38.8 ± 15.4a
nificant physiological differences between CA1 pyramidal neurons from β2 Scn2b 56.9 ± 15.2 44.1 ± 12.7
the dorsal and ventral hippocampus. Dorsal CA1 pyramidal neurons are β3 Scn3b 322.7 ± 75.4b 85.8 ± 26.4b
less intrinsically excitable than their ventral counterparts, due to the β4 Scn4b 10.4 ± 4.3c 0.0 ± 0.0c

increased activity of A-type potassium channels underlying the soma-


Relative transcript abundance was determined using HippoSeq, a publically
todendritic A-type potassium current (ISA), the increased activity of searchable hippocampal transcriptome database (http://hipposeq.janelia.org/;
KCNQ channels underlying the m-current (IM), and the hyperpolariza- Cembrowski et al., 2016b). Transcript abundance is presented as the average of
tion of resting membrane potential (RMP) stemming from decreased at least three replicates expressed in fragments per kilobase of exon per million
activity of the hyperpolarization-activated cation non-selective current fragments mapped (FPKM) bounded by their 95% confidence interval. Super-
(Ih) (Dougherty et al., 2012, 2013; Marcelin et al., 2012; Hönigsperger scripted letters indicate dorsal–ventral pairs with non-overlapping 95% con-
et al., 2015; Milior et al., 2016). Additionally, the expression of GIRK fidence intervals.
channels coupled with A1 adenosine receptors further suppresses the
intrinsic excitability of DHC neurons, thereby limiting the ability of 2. Materials and methods
these neurons to associate synaptic inputs over relatively long periods
of time (Kim and Johnston, 2015; Malik and Johnston, 2017). These 2.1. Acute hippocampal slice preparation
contrasting DHC and VHC phenotypes gradually blend together through
the intermediate hippocampus (IHC), as the intrinsic excitability of Acute hippocampal slices were prepared from 22 adult male
individual neurons from intermediate positions along the longitudinal Sprague-Dawley rats (4.7 ± 0.5 months; 403 ± 12 g) in accordance
hippocampal axis (i.e., the dorsal intermediate hippocampus, or dIHC, withe the rules and regulations of the Rhodes College Institutional
and the ventral intermediate hippocampus, or vIHC) lies between the Animal Care and Use Committee. Rats were deeply anesthetized with a
hypoexcitable DHC phenotype and the hyperexcitable VHC phenotype mixture of ketamine and xylazine immediately prior to transcardial
(Malik et al., 2016). This phenotypic transition along the longitudinal perfusion with a cutting saline of the following composition (in mM):
hippocampal axis is supported by several studies describing discrete 210 sucrose, 1.25 NaH2PO4, 25 NaHCO3, 2.5 KCl, 0.5 CaCl2, 7 MgCl2, 7
dorsal, intermediate, and ventral CA1 regions along the longitudinal dextrose, 1.3 ascorbic acid, and 3 sodium pyruvate (continuously
hippocampal axis defined by changes in gene expression (Dong et al., bubbled with 95% O2/5% CO2). Immediately thereafter, rats were de-
2009; Fanselow and Dong, 2010; Cembrowski et al., 2016a). Among capitated, their brains were removed, and hemisected along the long-
these region-specific genes are several determinants of intrinsic excit- itudinal fissure, and each hemisphere was prepared for slicing by
ability, which include, but are not limited to voltage-gated ion channels making a series of blocking cuts. Specifically, dorsal hippocampal slices
and their auxiliary subunits. Interestingly, one recent study by were prepared by resting the hemisected brain on its ventral surface
Cembrowski et al. (2016a) demonstrated a significant enrichment of the and making a ∼45° blocking cut (between the sagittal and coronal
sodium channel β4 subunit (Scn4b) in CA1 pyramidal neurons from the planes) at the level of the striatum, which subsequently served as the
DHC, thereby implying differences in the α–β subunit composition of mounting surface (for a more detailed description, please see
voltage-gated sodium channels in these neurons (Cembrowski et al., (Dougherty et al., 2012)). Slices from the ventral hippocampus were
2016a). Using a publically searchable hippocampal transcriptome da- prepared by resting the hemisected brain on its sagittal surface and
tabase (HippoSeq; http://hipposeq.janelia.org/), we examined the ex- making a single blocking cut along the horizontal plane at the level of
pression profiles of the remaining sodium channel subunits across the the dorsal hippocampus, which subsequently served as the mounting
longitudinal hippocampal axis, and found that three of the four sodium surface. 350 μm thick transverse slices were then prepared using a vi-
channel β-subunits (β1, β3, and β4) showed increased expression in the brating microtome (Vibratome 1000; McCormick Scientific, St Louis,
DHC relative to the VHC, whereas the β2 subunit was expressed simi- MO, USA) and allowed to recover in a holding saline of the following
larly across the longitudinal hippocampal axis (Table 1; Cembrowski composition (in mM): 125 NaCl, 2.5 KCl, 1.25 NaH2PO4, 25 NaHCO3, 2
et al., 2016b). CaCl2, 2 MgCl2, 12.5 dextrose, 1.3 ascorbic acid, and 3 sodium pyruvate
Given the physiological and molecular heterogeneity of CA1 pyr- (continuously bubbled with 95% O2/5% CO2), for 15 min at 35 °C, and
amidal neurons along the longitudinal hippocampal axis, the increased then at room temperature for an additional 45 min.
expression of sodium channel β-subunits in the DHC, and the differing
roles of the DHC and VHC in seizure generation, we hypothesized that 2.2. Electrophysiology
the classical anti-epileptic drug, carbamazepine (CBZ), which limits
excitability through a use-dependent sodium channel blocking me- Somatic whole-cell current-clamp recordings were performed on
chanism, would differentially influence intrinsic excitability across the CA1 pyramidal neurons from submerged slices bathed in a flowing
longitudinal hippocampal axis. Accordingly, we found that acute ex- (1–2 mL/min), heated (31–33 °C) artificial cerebrospinal solution
posure to 100 μM CBZ 1) induces a greater suppression of repetitive (aCSF) of the following composition (in mM): 125 NaCl, 3 KCl, 1.25
firing in the DHC than the VHC, and 2) results in a significant depo- NaH2PO4, 25 NaHCO3, 2 CaCl2, 1 MgCl2, 12.5 dextrose, 1.3 ascorbate,
larization of RMP for DHC neurons, but not VHC neurons. These find- and 3 sodium pyruvate (continuously bubbled with 95% O2/5% CO2).
ings underscore the molecular and physiological heterogeneity of CA1 Neurons were visually identified using a Zeiss Axioskop FS microscope
neurons across the longitudinal hippocampal axis, suggest a unique role equipped with infrared differential interference contrast optics coupled
for the DHC in the side effect profile of carbamazepine in non-epileptic with an infrared camera and television monitor (DAGE-MTI, Michigan
tissue, and prompt compelling questions concerning carbamazepine’s City, IN, USA). Electrodes (4–6 MΩ) were pulled from borosilicate ca-
efficacy along the longitudinal hippocampal axis in epileptic tissue. pillary glass using a P-87 Flaming–Brown micropipette puller (Sutter

64
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

Instruments, San Fransisco, CA, USA), and filled with an internal so- 2.5. Statistical evaluation
lution with the following composition (in mM): 120 potassium gluco-
nate, 20 KCl, 10 HEPES, 4 NaCl, 4 Mg-ATP, 0.3 Na-GTP, and 7 (Tris)2- Statistical significance between paired groups was determined using
phosphocreatine (pH 7.3, adjusted with KOH; 300 mOsm, adjusted with a Paired Student’s t test when n ≥ 7, and a Wilcoxon Signed-Rank test
sucrose). Whole-cell current-clamp recordings were performed using a when n < 7, the data were not normally distributed, or the variances
Dagan BVC-700A amplifier (Dagan, Minneapolis, MN, USA) under the were unequal. Statistical significance between independent groups was
control of a MacMini computer (Apple Computer Company, Cupertino, determined using Student’s t test when n ≥ 7, and a Wilcoxon Ranked-
CA, USA) through an ITC-18 computer interface (InstruTech, Port Sum test when n < 7, the data were not normally distributed, or the
Washington, NY, USA). Signals from were low-pass filtered at 3 kHz and variances were unequal. Differences between firing frequency versus
sampled at 10 kHz for subthreshold experiments, whereas signals from current injection (F–I) relationships were determined using a Two-Way
suprathreshold experiments were low-pass filtered at 5 kHz and sam- Repeated Measures ANOVA test, followed by a Bonferroni correction
pled at 40 kHz. Neurons were allowed to stabilize for at least five for multiple comparisons. Paired F–I relationships were evaluated using
minutes after break in, and the pipette capacitance and access re- repeated measures in two dimensions (time and current injection),
sistance were compensated throughout the experiment prior to each whereas unpaired F–I relationships were evaluated using repeated
recording. Experiments were terminated if the series resistance ex- measures in one dimension (current injection). Spearman’s Rank
ceeded 30 MΩ, and small DC current injections were often used to Correlation test was used to evaluate correlation between datasets.
maintain the membrane potential at −65 mV. Membrane potentials Unless otherwise stated, statistical analyses and graphical displays were
have not been adjusted to account for a liquid junction potential of accomplished using IGOR Pro v6.35 (Wavemetrics, Lake Oswego, OR,
approximately 8 mV. Neurobiotin (0.1–0.2%; Vector Laboratories, USA). Two-Way ANOVA tests and correlation analyses were performed
Burlingame, CA) was added to the internal solution in order to facilitate using Prism 7 (GraphPad Software, La Jolla, CA, USA). All hypothesis
the subsequent visualization of individual neurons and to verify intact testing was conducted with α = 0.05, and the specific statistical tests
apical and basal dendritic trees. After each experiment, slices were fixed used to evaluate each dataset are indicated in the corresponding figure
in 0.1 M phosphate buffer with 3% gluteraldehyde, stored at 4 °C for at legends. All data are presented as SEM, unless otherwise stated.
least 48 h, and processed using an avidin-HRP system activated by
diaminobenzidine (DAB; Vector Laboratories). Processed slices were 3. Results
mounted in glycerol and visualized using a compound microscope with
a 10X objective. 3.1. Baseline parameters

2.3. Drugs and solutions In order to evaluate the influence of carbamazepine along the
longitudinal hippocampal axis, we performed whole-cell current-clamp
All experiments were performed in an aCSF supplemented with recordings on CA1 pyramidal neurons from DHC and VHC slices pre-
20 μM 6,7-dinitroquinoxaline 2,3-dione (DNQX; Abcam, Cambridge, pared from adult male rats (4.7 ± 0.5 months old; 403 ± 12 g). DHC
MA, USA) and 2 μM Gabazine (Abcam) to block AMPA receptors and slices were sampled from a range of locations along the longitudinal
GABAA receptors, respectively, and thereby suppress spontaneous ac- hippocampal axis that included both the dorsal hippocampus and the
tivity within the slice. Carbamazepine (Sigma-Aldrich, St Louis, MO, dorsal intermediate hippocampus (dIHC; as defined by Malik et al.,
USA) was dissolved in DMSO (Sigma-Aldrich) to yield a 100 mM stock 2016), whereas VHC slices were taken mostly from the ventral inter-
solution, which was subsequently aliquoted and frozen until needed. mediate hippocampus (vIHC), rather than the ventral hippocampus
Carbamazepine aliquots were diluted 1000-fold into aCSF to yield near the temporal pole (Malik et al., 2016). CA1 pyramidal neurons
100 μM carbamazepine and 0.1% DMSO. This CBZ concentration is from this segment of the longitudinal hippocampal axis (DHC–dIHC–-
approximately 2–3X the plasma concentration required to suppress vIHC) are generally similar with regard to their morphology and
hind limb extension in the maximal electroshock seizure test in rats baseline intrinsic parameters, such as Rin (at −65 mV), τm (at −65 mV),
(Bialer et al., 2004). DMSO was diluted 1000-fold in aCSF to yield 0.1% and Vthr (at −65 mV) (Malik et al., 2016). This observation stands in
DMSO solutions for vehicle control experiments. contrast to the region of the VHC closest to the temporal pole, where
CA1 pyramidal neurons exhibit a significantly different morphological
2.4. Data analysis and electrophysiological phenotype (Malik et al., 2016; Dougherty
et al., 2012). Therefore, we limited our study to the DHC–dIHC–vIHC
Suprathreshold parameters, such as the threshold voltage (Vthr), segment of the longitudinal hippocampal axis, and will hereafter refer
maximal dV/dt, and action potential peak (APpeak), were determined to the vIHC as the VHC. Accordingly, representative DHC and VHC
through phase–plane analysis, where Vthr was defined as the membrane slices, with corresponding representative filled CA1 pyramidal neurons
potential were the dV/dt first exceeded 20 mV/ms. The action potential are displayed in Fig. 1A–B. Baseline parameters such as RMP, Rin (at
size was defined as the difference between the APpeak and Vthr, and the −65 mV), Rin (at RMP), and τm (at −65 mV) were not significantly
action potential half-width was defined as the width of the action po- different between DHC and VHC neurons (DHC RMP = −63.4 ±
tential waveform at one half of the action potential size. The spike 0.9 mV, n = 15, Rin (at −65 mV) = 58.0 ± 2.5 MΩ, n = 15, Rin (at
frequency adaptation (SFA) ratio was determined from action potential RMP) = 59.6 ± 3.9 MΩ, n = 11, τm (at –65 mV) = 20.0 ± 1.1 ms,
trains elicited by one second long current injections, where the current n = 10; VHC RMP = −62.0 ± 0.8 mV, n = 15, Rin (at
magnitude was varied to produce trains containing 10–12 action po- −65 mV) = 61.8 ± 3.4 MΩ, n = 15, Rin (at RMP) = 71.5 ± 8.9 MΩ,
tentials. The SFA was then calculated by dividing the duration of the n = 11, τm (at –65 mV) = 21.0 ± 1.9, n = 10; Fig. 1C–E). These values
final interspike interval (ISI) by the duration of the first ISI. The input were similar to those reported in a previous study using adult rats under
resistance (Rin) was determined as the slope of the voltage–current plot similar conditions (i.e., without blocking NMDA receptors), although
constructed from the steady-state voltage responses to a family of the Rin values for VHC neurons were slightly lower than previously
current injections ranging from −50 to 50 pA (only the linear region of reported (Clemens and Johnston, 2014). This is likely due to the
these plots were used for determining the Rin). The membrane time longitudinal location of our VHC slices (see above). The voltage
constant (τm) was determined by averaging the voltage responses to threshold (Vthr) of single action potentials elicited by current injections
100, 1 ms square current injections (100 pA), and fitting the average from −65 mV were also found to be indistinguishable between DHC
response with a double exponential curve. The slower of the two time and VHC neurons in both the 10 ms or 50 ms timeframes (Fig. 1F; DHC
constants was taken as τm. Vthr (10 ms) = −49.2 ± 1.0 mV, n = 13, Vthr (50 ms) = −47.0 ±

65
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

Fig. 1. Baseline properties of CA1 pyramidal neurons across the longitudinal hippocampal axis of adult rats. A–B) Representative slices (left) and filled CA1
pyramidal neurons (right) from the dorsal (A) and ventral (B) hippocampus. C–E) The resting membrane potential (C), Rin (at −65 mV) (D; circles), the Rin (at RMP)
(D; squares), and the τm (at −65 mV) (E) of DHC (blue) and VHC neurons (black) were not significantly different (Student’s t-test, p > 0.05). F) The voltage threshold
measured 10 ms (circles) and 50 ms (squares) after the onset of the current stimulus were not significantly different between DHC (blue) and VHC (black) neurons
(Student’s t-test, p > 0.05). The voltage threshold for VHC neurons was significantly more depolarized in the 50 ms timeframe than in the 10 ms timeframe (Students’s
t-test, p < 0.05), whereas the voltage threshold was not significantly different for DHC neurons between these timeframes (Student’s t-test, p > 0.05). Average
values are represented as horizontal lines. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

0.8 mV, n = 13; VHC Vthr (10 ms) −49.6 ± 0.6 mV, n = 12, Vthr firing frequency of both DHC neurons and VHC neurons, although VHC
(50 ms) = −46.5 ± 0.8 mV, n = 12). Although some subtle gradients neurons continued to fire throughout the current stimulus, whereas
in intrinsic excitability are likely present, these observations demon- DHC neurons often fired a few action potentials at the beginning of the
strate that several key factors of intrinsic excitability are generally si- current stimulus before ceasing firing altogether (Fig. 3A–C). As in the
milar along a significant portion of the longitudinal hippocamal axis. shorter, one second timeframe, 100 μM CBZ induced a significantly
This overall similarity serves as a foundation for investigating the in- greater suppression in the F–I relationships for DHC neurons than their
fluence of AEDs across this region of the longitudinal hippocampal axis. VHC counterparts, despite indistinguishable baseline F–I relationships
(Fig. 3D–E). There was no significant reduction of the F–I relationship
in 0.1% DMSO control experiments (Supplementary Fig. 2). Moreover,
3.2. Carbamazepine-induced suppression of repetitive firing is greater in the firing frequency in the presence of 0.1% DMSO increased over time,
DHC neurons than VHC neurons yielding a fold change in firing frequency (10 s, 250 pA current injec-
tions) greater than one for both DHC and VHC neurons (Fig. 4; DHC
We evaluated the effect of carbamazepine on repetitive firing in Fold Change (DMSO) = 1.63 ± 0.16, n = 6; VHC Fold Change
DHC and VHC neurons before, and 20–60 min after applying 100 μM (DMSO) = 1.37 ± 0.18, n = 5). When compared to 0.1% DMSO con-
CBZ through the bath solution (solution exchange was conducted at trol experiments, 100 μM CBZ significantly reduced the firing frequency
RMP, which was allowed to float throughout the wash-in period). of both DHC and VHC neurons (Fig. 4; DHC Fold Change
100 μM CBZ significantly reduced the firing frequency for action po- (CBZ) = 0.13 ± 0.06, n = 8; VHC Fold Change (CBZ) = 0.65 ± 0.13,
tential trains elicited by one-second long current injections from n = 7). The fold change in firing frequency in response to 100 μM CBZ
−65 mV (Fig. 2A–C). This effect was especially pronounced in DHC was significantly lower for DHC neurons than their VHC counterparts
neurons, which often ceased firing after just a few action potentials (Fig. 4).
(Fig. 2A, left column). In contrast, VHC neurons tended to fire
throughout the current injection, albeit at a lower frequency (Fig. 2A,
right column). Although the F–I baseline relationships for DHC and 3.3. Carbamazepine influences subthreshold and supratheshold properties
VHC neurons were indistinguishable (Fig. 2D), the F–I relationship for of CA1 pyramidal neurons from the dorsal hippocampus
DHC neurons was significantly lower than their VHC counterparts in
the presence of 100 μM CBZ (Fig. 2E). This differential suppression of In order to understand the mechanism(s) driving the dispropor-
the F–I relationship was not observed in vehicle control experiments, tionate effect of CBZ on DHC neurons, we measured carbamazepine’s
where only 0.1% DMSO was applied to DHC and VHC neurons (Sup- effect on single action potentials in DHC and VHC neurons. Single ac-
plementary Fig. 1). tion potentials were elicited by brief (10 ms) current injections of
We next extended the duration of the current injection from one variable amplitude, such that the first action potential occurred within
second to ten seconds in order to examine the influence of carbama- a ± 1 ms time window centered 8 ms following the onset of the current
zepine on firing for DHC and VHC neurons throughout prolonged action stimulus before and after 100 μM CBZ wash in (Fig. 5A; Table 2). This
potential trains. Accordingly, both DHC and VHC neurons fired short timeframe was chosen to minimize the influence of slowly acti-
throughout the 10 s, 250 pA current injection in the absence of carba- vating and inactivating ion channels. Under these conditions, baseline
mazepine (Fig. 3A). Exposure to 100 μM CBZ significantly reduced the parameters from DHC and VHC neurons were indistinguishable

66
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

Fig. 2. Carbamazepine-induced suppression of repetitive firing in response to short current injections. A) Representative traces from DHC (left) and VHC
(right) neurons elicited by 1 s, 330 pA current injections from −65 mV before (upper traces) and after (lower traces) exposure to 100 μM carbamazepine. B–C) F–I
relationships for DHC (B) and VHC (C) neurons before (hollow circles) and after (filled circles) exposure to 100 μM carbamazepine. Carbamazepine significantly
suppressed repetitive firing for DHC neurons and VHC neurons (Two-way repeated measures ANOVA; Bonferroni test for multiple comparisons, p < 0.05). D–E) F–I
relationships of DHC and VHC neurons before (D) and after (E) exposure to 100 μM carbamazepine. In the presence of 100 μM CBZ, the F–I relationship was
significantly lower for DHC neurons than VHC neurons (Two-way repeated measures ANOVA; Bonferroni test for multiple comparisons, p < 0.05).

(Table 2), except for the maximum dV/dt, which was significantly n = 8; VHC ΔVthr = −0.6 ± 0.7 mV, n = 6), action potential size
lower for VHC neurons than DHC neurons (DHC max dV/ (DHC ΔAPsize = −3.5 ± 1.6 mV, n = 8; VHC ΔAPsize = −3.8 ±
dt = 437.1 ± 16.5 mV/ms, n = 13; VHC max dV/dt = 371.2 ± 2.0 mV, n = 6), and action potential halfwidth (DHC ΔAPhw =
12.8 mV/ms, n = 12) Subsequent exposure to 100 μM CBZ did not 0.02 ± 0.02 ms, n = 8; VHC ΔAPhw = 0.04 ± 0.03 ms, n = 6) for ei-
significantly alter the voltage threshold (DHC ΔVthr = −1.2 ± 2.0 mV, ther DHC or VHC neurons (Fig. 5; Table 2). However, the maximal dV/

Fig. 3. Carbamazepine-induced suppression of repetitive firing in response to long current injections. A) Representative traces from DHC (left) and VHC
(right) neurons elicited by 10 s, 250 pA current injections from −65 mV before (upper traces) and after (lower traces) exposure to 100 μM carbamazepine. B–C) F–I
relationships for DHC (B) and VHC (C) neurons before (hollow circles) and after (filled circles) exposure to 100 μM carbamazepine. Carbamazepine significantly
suppressed repetitive firing for DHC neurons and VHC neurons (Two-way repeated measures ANOVA; Bonferroni test for multiple comparisons, p < 0.05). D–E) F–I
relationships of DHC and VHC neurons before (D) and after (E) exposure to 100 μM carbamazepine. In the presence of 100 μM CBZ, the F–I relationship was
significantly lower for DHC neurons than VHC neurons (Two-way repeated measures ANOVA; Bonferroni test for multiple comparisons, P < 0.05).

67
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

dt was significantly reduced for both DHC and VHC neurons in the
presence of 100 μM CBZ (Fig. 5D; DHC Δmax dV/dt = −41.6 ±
13.0 mV/ms, n = 8; VHC Δmax dV/dt = −27.7 ± 7.2 mV/ms, n = 6).
Despite this effect, single action potentials were mostly unaffected by
the presence of 100 μM CBZ, which is consistent with carbamazepine’s
classical mechanism of action as a use-dependent sodium channel
blocker (Ragsdale et al., 1991).
Although single action potentials were mostly unaffected, carba-
mazepine had a profound effect on action potential trains for both DHC
and VHC neurons (Figs. 2 and 3). Accordingly, we examined carba-
mazepine’s influence on spike frequency adaptation, by calculating the
SFA ratio from action potential trains containing 10–12 action poten-
tials (Fig. 6A). The baseline SFA ratio was significantly larger for DHC
neurons compared to VHC neurons (Fig. 6B; DHC SFA ratio = 2.95 ±
0.31, n = 11; VHC SFA ratio = 1.39 ± 0.19, n = 10), which is con-
sistent with previous reports (Malik et al., 2016). However, in the
presence of 100 μM CBZ, it became impossible to elicit the 10–12 action
potentials required to determine the SFA ratio in all but two DHC
Fig. 4. Summary of carbamazepine’s effect on firing frequency. The fold neurons–the SFA ratio for these two neurons increased 2.41 and 1.67
change in firing frequency (relative to the baseline firing frequency) for action fold. Alternatively, 100 μM CBZ did not significantly alter the SFA ratio
potential trains elicited by 10 s, 250 pA current injections, is significantly lower for VHC neurons (Fig. 6B; VHC ΔSFA ratio = 1.59 ± 0.13, n = 5), and
in the presence of 100 μM carbamazepine for both DHC and VHC neurons re- no significant effect was observed in 0.1% DMSO control experiments
lative to 0.1% DMSO controls (Wilcoxon’s Ranked Sum Test, p < 0.05). The (Fig. 6B; DHC ΔSFA ratio = 0.72 ± 0.06, n = 5; VHC ΔSFA
fold change in firing frequency is significantly lower for DHC neurons than VHC
ratio = 0.88 ± 0.24, n = 4). These observations demonstrate that
neurons following 100 μM CBZ exposure (Wilcoxon’s Ranked Sum Test,
carbamazepine induces a profound spike frequency adaptation in DHC
p < 0.05), whereas there was no difference in the fold change in firing fre-
quency following 0.1% DMSO exposure (Wilcoxon’s Ranked Sum Test,
neurons, but spares their VHC counterparts.
p > 0.05). Carbamazepine’s effect was not limited to supratheshold properties.
Surprisingly, carbamazepine influenced the subthreshold properties of

Fig. 5. Carbamazepine does not significantly alter single action potentials. A) Single action potentials elicited by 10 ms current injections before and after
exposure to 100 μM CBZ in DHC (upper traces) and VHC (lower traces) neurons. B–C) The threshold voltage (B) and action potential size (C) were not significantly
different following 100 μM CBZ exposure for DHC (blue circles; Paired Student’s t-test, p > 0.05) or VHC (black circles; Wilcoxon’s Signed Rank test, p > 0.05)
neurons. D) Carbamazepine significantly reduced the max dV/dt for both DHC (Paired Student’s t-test, p < 0.05) and VHC neurons (Wilcoxon’s Signed Rank test,
p < 0.05). E) The action potential halfwidth was not significantly different following carbamazepine exposure for either DHC (Paired Student’s t-test, p > 0.05) or
VHC neurons (Wilcoxon’s Signed Rank test, p > 0.05). Average values are represented as horizontal lines. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

68
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

Table 2
Single action potential parameters.
Voltage Threshold (mV) Maximum dV/dt (mV/ms) AP size (mV) AP halfwidth (ms)

DHC CBZ baseline −49.3 ± 1.3 (8) 438.9 ± ac


18.1 (8) 90.9 ± 1.4 (8) 0.87 ± 0.04 (8)
DHC CBZ wash-in −50.5 ± 1.8 (8) 397.2 ± 15.3a (8) 87.4 ± 1.6 (8) 0.88 ± 0.05 (8)
DHC DMSO baseline −49.0 ± 1.7 (5) 434.3 ± 34.8 (5) 86.1 ± 4.6 (5) 0.88 ± 0.03 (5)
DHC DMSO wash-in −52.3 ± 0.8 (5) 398.5 ± 21.2 (5) 87.3 ± 2.0 (5) 0.91 ± 0.04 (5)
VHC CBZ baseline −50.3 ± 0.6 (6) 370.2 ± 17.2bc (6) 88.7 ± 1.45 (6) 0.93 ± 0.04 (6)
VHC CBZ wash-in −50.9 ± 0.4 (6) 342.6 ± 18.9b (6) 84.9 ± 2.6 (6) 0.97 ± 0.05 (6)
VHC DMSO baseline −48.4 ± 1.1 (5) 383.5 ± 20.9 (5) 86.1 ± 2.44 (5) 0.87 ± 0.02 (5)
VHC DMSO wash-in −50.4 ± 1.3 (5) 374.7 ± 17.3 (5) 80.1 ± 2.7 (5) 0.79 ± 0.06 (5)

The parameters of single action potentials elicited approximately 8 ms after the onset of a square pulse from −65 mV are displayed before (baseline), and after
exposure to either 100 μM CBZ or 0.1% DMSO. Statistical significance was determined using either a paired Student’s t test (n ≥ 7) or a Wilcoxon’s Signed Rank test
(n < 7), with p < 0.05. Significant differences are indicated by superscripted letters, and n values are indicated in parentheses.

Fig. 6. Carbamazepine-induced enhancement of spike frequency adaptation across the longitudinal hippocampal axis. A) The SFA ratio was determined by
dividing the duration of the final ISI (dotted lines) by the first ISI (solid lines) for one second long action potential trains containing 10–12 action potentials. B) The
baseline SFA ratio was significantly higher for DHC neurons (blue symbols) than VHC neurons (black symbols) for both 0.1% DMSO control (squares; Wilcoxon’s
Ranked Sum test, p < 0.05) and 100 μM CBZ experiments (circles; Wilcoxon’s Ranked Sum test, p < 0.05). Exposure to 100 μM CBZ induced a profound spike
frequency adaptation in DHC neurons that made it impossible to elicit 10–12 action potentials in the one second time frame in all but two neurons, whereas VHC
neurons continued to fire throughout the one second interval, although 100 μM CBZ did not significantly alter the SFA ratio (Wilcoxon’s Signed Rank test, p > 0.05).
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

DHC neurons, as 100 μM CBZ induced a small, but significant depo- 4. Discussion
larization of RMP for DHC neurons, but not VHC neurons (Fig. 7A; DHC
ΔRMP = 3.0 ± 1.1 mV, n = 9; VHC ΔRMP = −0.2 ± 0.6 mV, n = 8). 4.1. Summary
However, the presence of 100 μM CBZ did not significantly alter the Rin
(at − 65 mV) (Fig. 7B–C; DHC ΔRin (at -65 mV) = -0.8 ± 5.6 MΩ, In this study, we investigated the influence of carbamazepine on CA1
n = 6; VHC ΔRin (at −65 mV) = 10.1 ± 9.7 MΩ, n = 7), and no pyramidal neurons across the longitudinal hippocampal axis, and have
changes in either the RMP nor the Rin (at −65 mV) were observed found that carbamazepine 1) significantly suppresses repetitive firing in
during 0.1% DMSO control experiments (DHC ΔRMP = 1.3 ± 0.8 mV, DHC neurons through an enhancement of spike frequency adaptation
ΔRin (at -65 mV) = 7.6 ± 7.2 MΩ, n = 6; VHC ΔRMP = −1.3 ± (Figs. 2–4 and 6), 2) induces a slight depolarization of the RMP in DHC
0.8 mV, ΔRin (at −65 mV) = 5.1 ± 8.6 MΩ, n = 7). Furthermore, neurons (Fig. 7), and 3), that these two effects were significantly corre-
100 μM CBZ did not alter τm (at −65 mV) for DHC neurons (Fig. 7D–E; lated across the longitudinal hippocampal axis (Fig. 8). Carbamazepine
Δτm (at −65 mV) = 0.8 ± 3.3 ms, n = 6), but induced a small, but also suppressed repetitive firing in VHC neurons, but this effect was
significant increase in the τm (at −65 mV) for VHC neurons (Fig. 7D–E; significantly smaller than their DHC counterparts (Figs. 2–4). Moreover,
VHC Δτm (at −65 mV) = 3.1 ± 2.0 ms, n = 6). No significant differ- no change in the RMP was observed following 100 μM CBZ exposure for
ences were observed for τm (at −65 mV) following exposure to 0.1% VHC neurons (Fig. 7A). Together, these observations suggest a unique
DMSO (DHC Δτm (at −65 mV) = 5.9 ± 2.1, n = 4; VHC Δτm (at sensitivity to carbamazepine for DHC neurons relative to their VHC
−65 mV) = 2.1 ± 0.8 ms, n = 4). Interestingly, cells that depolarized counterparts. The potential mechanisms and physiological implications
in the presence of 100 μM CBZ tended to show a greater suppression of of these findings are further discussed below.
their F–I relationships, as the ΔRMP was significantly correlated with
the fold change in firing frequency along the longitudinal hippocampal 4.2. Comparison with previous work
axis (Fig. 8A). Neither the suprathreshold nor the subthreshld effect of
carbamazepine significantly correlated with the original RMP, at which Interestingly, this regional difference in carbamazepine sensitivity
100 μM CBZ was applied to the slice (data not shown). Together, these emerged from a population of CA1 pyramidal neurons with similar
observations demonstrate that carbamazepine exerts a unique influence baseline intrinsic properties (e.g., RMP, Rin (at −65 mV), Rin (at RMP),
over the intrinsic excitability of CA1 pyramidal neurons from the DHC. τm (at −65 mV), and Vthr; Fig. 1). At first glance, this baseline similarity
between DHC and VHC neurons seemingly contradicts the significant
differences in intrinsic excitability reported across the longitudinal

69
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

Fig. 7. 100 μM carbamazepine influences subthreshold properties of CA1 pyramidal neurons. A) The RMP of DHC neurons (blue circles), but not VHC neurons
(black circles) was significantly more depolarized following exposure to 100 μM CBZ (Paired Student’s t-test, p < 0.05). B–C) the Rin (at −65 mV) was not sig-
nificantly different following exposure to 100 μM CBZ for either DHC (B blue circles; C blue traces), or VHC neurons (B black circles; C black traces; Paired Student’s t-
test, p > 0.05). D–E) the τm (at −65 mV) was not significantly different following exposure to 100 μM CBZ for DHC neurons (D blue circles; E left, blue/gray traces;
Wilcoxon’s Signed Rank test, p > 0.05) but showed a small, yet significant increase for VHC neurons (D black circles; E right, black/gray traces; Wilcoxon’s Sign Rank
test, p < 0.05). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

for this study were significantly older than those used in previous stu-
dies (Dougherty et al., 2012, 2013; Malik et al., 2016; Kim and
Johnston, 2015), 2) NMDA receptor blockers were omitted from our
cocktail of synaptic blockers, and 3) slices for this study were prepared
using a slightly different blocking strategy than previous studies, which
doesn’t sample the extreme ventral hippocampus near the temporal
pole. Recent work by Clemens and Johnston (2014), using adult rats
and a similar synaptic blocker cocktail (i.e., without NMDA blockers)
reported an ∼3 mV difference between the average RMPs, and an ∼
20 MΩ difference in the Rin (at −65 mV) of DHC and VHC neurons
(Clemens and Johnston, 2014). Our baseline RMP and Rin (at −65 mV)
values matched those previously reported by Clemens and Johnston
(2014) for DHC neurons, whereas our Rin (at −65 mV) was lower than
previously reported for VHC neurons (62 MΩ versus ∼80 MΩ from
Clemens and Johnston (2014)). This discrepancy likely stems from the
relative longitudinal position of the slices used for this study. Accord-
Fig. 8. Correlation between carbamazepine’s subthreshold and supra-
ingly, Malik et al. (2016) have recently reported tends in intrinsic ex-
threshold effects across the longitudinal hippocampal axis. DHC (blue
circles) and VHC neurons (black circles) were combined in order to evaluate the citability along the longitudinal axis of young adult rats, wherein the
correlation between the ΔRMP with the fold change in firing frequency for CA1 Rin (at −65 mV) remains constant around 60 MΩ along the dorsal
pyramidal neurons across the longitudinal hippocampal axis. This relationship portion of the longitudinal axis until jumping to about 80 MΩ near the
was described assuming a linear function with the following parameters: y- temporal pole (Malik et al., 2016). The slices used in this study were
intercept = 3.2 mV, slope = −4.1 mV/fold change, and correlation was eval- prepared mostly from the vIHC, rather than the extreme VHC, which
uated using Spearman’s Rank Correlation test (R = −0.54, p = < 0.05, n = 15). places our observations in good agreement with those previously re-
(For interpretation of the references to colour in this figure legend, the reader is ported by Malik et al. (2016). This reasoning also accounts for the si-
referred to the web version of this article.) milarity in Vthr that we observed across the longitudinal hippocampal
axis, as Vthr remains similar across the DHC, dIHC, and vIHC, before
hippocampal axis (Dougherty et al., 2012; Hönigsperger et al., 2015; depolarizing in the extreme VHC (Malik et al., 2016). The similarity in
Malik et al., 2016; Kim and Johnston, 2015; Milior et al., 2016). RIn (at −65 mV) and Vthr along this region of the longitudinal hippo-
However, our current experimental conditions differ from those re- campal axis spares our interpretation from the potentially confounding
ported in previous studies in several important ways: 1) The rats used influence of differences in baseline intrinsic excitability.

70
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

4.3. Potential mechanisms VHC; Figs. 7 and 8). Interestingly, the depolarization of RMP occurred
in DHC neurons, rather than in VHC neurons, where a depolarization of
The main finding of this work is the region-specific suppression of RMP might be expected if carbamazepine was augmenting the persis-
repetitive firing of CA1 pyramidal neurons along the longitudinal hip- tent sodium current. This suggests that the effect on RMP is either
pocampal axis in response to 100 μM CBZ. In the presence of carba- mechanistically unrelated to the decreased firing frequency in DHC
mazepine, the F–I relationships for DHC neurons were significantly neurons, or that our inferred mechanism involving the persistent so-
more suppressed than their VHC counterparts in both the one and 10 s dium current is incorrect, and that carbamazepine’s dual effects on DHC
time frames (Figs. 2 and 3), despite having similar baseline intrinsic neurons are coupled through an unknown mechanism. We favor the
properties (see above). Moreover, single action potentials were gen- former explanation, as DHC neurons are known to express secondary
erally unaffected by the presence of 100 μM CBZ, although there was a targets for carbamazepine: A1 adenosine receptors (Lee et al., 1983;
slight reduction in the maximal dV/dt for both DHC and VHC neurons Kim and Johnston, 2015). Carbamazepine interacts with A1 adenosine
(Fig. 5). Instead, carbamazepine influenced action potential trains by receptors, which, through their coupling with GIRK channels, hy-
augmenting spike frequency adaptation in DHC neurons (Figs. 2, 3 and perpolarize the membrane potential of DHC neurons (Weir et al., 1984;
6). This enhancement of spike frequency adaptation was so profound, Kim and Johnston, 2015). The slight depolarization of RMP observed in
that most DHC neurons fired only 1–5 action potentials shortly after the DHC neurons could plausibly stem from carbamazepine’s interaction
stimulus onset, which precluded the determination of the SFA ratio for with this system, although additional work will be required to fully
these neurons. This difference between carbamazepine’s effect on DHC explore this hypothesis.
and VHC neurons is striking, and raises questions about mechanistic
differences between these two populations of neurons. Recently, Malik 4.4. Physiological implications
et al. (2016) reported a gradient in spike frequency adaptation along
the longitudinal hippocampal axis, with DHC neurons showing the Although carbamazepine is most commonly used for seizure control
greatest degree of adaptation (Malik et al., 2016). Our work corrobo- in epileptic patients, it is often used to treat non-epileptic conditions,
rates this baseline difference in spike frequency adaptation, and sug- such as neuropathic pain (trigeminal neuralgia and diabetic neuro-
gests that carbamazepine exaggerates this pre-existing difference in pathy), bipolar disorder, and post-traumatic stress disorder (PTSD),
spike frequency adaptation across the longitudinal hippocampal axis, among others (Spina and Perugi, 2004). Because this study was per-
thereby implying some difference in the molecular targets present in formed using naive, non-epileptic tissue, the physiological ramifications
these neurons. of this work should be considered in the context of carbamazepine’s
Carbamazepine’s classical mechanism of action involves the use- uses in non-epileptic patients. Carbamazepine is generally associated
dependent blockade of voltage-gated sodium channels. This progres- with several negative side effects, including significant changes in
sively limits the fraction of available sodium channels throughout an cognitive function, such as reductions in information processing speed,
action potential train, thereby depolarizing threshold and prolonging attention, verbal fluency, and arithmetic (Eddy et al., 2011). Carba-
the interspike interval, which results in spike frequency adaptation. mazepine’s effect on memory is less clear, as some reports claim, im-
This effect relies primarily on the loss of the transient component of pairment, no effect, or even an enhancement in memory tasks (Eddy
sodium currents mediated by voltage-gated sodium channels, due to a et al., 2011). In adult rats similar to those used in this study, carba-
hyperpolarizing shift in the steady-state inactivation curve and corre- mazepine has been shown to impair performance in the Morris Water
spondingly slowed recovery from inactivation. However, Uebachs et al. Maze–a spatial learning and memory task known to rely on the DHC
(2010) have reported a paradoxical carbamazepine-induced effect on (Churchill et al., 2003). This observation is generally consistent with
the persistent sodium current that relies on sodium channel subunit our main finding that carbamazepine’s effects are concentrated within
composition (Uebachs et al., 2010). Specifically, carbamazepine in- the DHC, and suggests that carbamazepine’s influence over the DHC
duces a hyperpolarizing shift in the conductance vs voltage relationship could impact behavior by impairing spatial learning in adult rats.
(G–V curve) of the persistent sodium current in CA1 pyramidal neurons Finally, it is surprising that the carbamazepine-induced suppression
from mice lacking the β1 auxiliary subunit. This augmentation of the of repetitive firing is greater in the DHC, rather than in the VHC, where
subthreshold sodium current accelerates the depolarizing envelope seizures tend to originate, and where CA1 neurons show greater in-
between action potentials, thereby decreasing the interspike interval trinsic excitability. This observation generally resembles another recent
and counteracting the suppression of repetitive firing in these neurons. study by Hönigsperger et al. (2015), which demonstrated a DHC-spe-
The resulting F–I relationships for Scn1b knock-out mice are remarkably cific effect of retigabine–another anti-epileptic drug with a mechanism
unchanged by exposure to 100 μM CBZ (Uebachs et al., 2010). Our unrelated to carbamazepine (i.e., enhancement of IM). Together, these
results are generally consistent with such a mechanism, as the sig- observations suggest that the suppression of activity in the dorsal hip-
nificantly reduced carbamazepine sensitivity of VHC neurons mirrored pocampus could play a larger role in limiting seizure generation than
the effect of 100 μM CBZ on CA1 neurons from Scn1b knock-out mice, previously thought. However, additional experiments exploring the
whereas carbamazepine’s suppression of repetitive firing in DHC neu- effect of anti-epileptic drugs in the dorsal hippocampus of epileptic
rons resembled the Scn1b wild type mouse (Uebachs et al., 2010). Im- tissue will be required to fully explore this issue.
portantly, this potential mechanism is undergirded by the observation
that β1, β3, and β4 sodium channel subunits were expressed more 4.5. Conclusions
abundantly in the DHC relative to the VHC (Table 1), with no difference
in β2-subunit expression or any of the relevant sodium channel α-sub- In summary, we have demonstrated that acute exposure to the use-
unit genes (i.e., Scn1A/Nav1.1, Scn2a1/Nav1.2, Scn3 A/Nav1.3, and dependent sodium channel blocker carbamazepine induces greater
Scn8 A/Nav1.6; Table 1). This observation implies a gradient in β1- suppression of repetitive firing in CA1 neurons from the DHC than the
subunit expression along the longitudinal hippocampal axis, which VHC. Although single action potentials are generally unaffected, car-
could modify carbamazepine’s efficacy along this axis, in a manner bamazepine induces profound spike frequency adaptation in DHC
consistent with our results. The potential influence of the β3 and β4- neurons. Moreover, carbamazepine induces a small, but significant
subunits on carbamazepine efficacy in the DHC, however, remains depolarization of the RMP for DHC neurons, but not VHC neurons, and
unexplored. the magnitude of this depolarization correlates with the suppression of
In addition to the suppression of the F–I relationship, we also ob- repetitive firing across all regions of the longitudinal hippocampal axis
served a slight depolarization of the RMP in DHC neurons, which cor- under investigation. Together, these observations provide novel insight
related with the decreased firing frequency across all cells (DHC and into carbamazepine’s side effect profile in non-epileptic tissue, and

71
M.C. Evans, K.A. Dougherty Epilepsy Research 145 (2018) 63–72

could inform future work on carbamazepine’s (or related anti-epileptic Natl. Acad. Sci. U. S. A. 106, 11794–11799.
drugs) ability to suppress seizure generation in epileptic hippocampal Dougherty, K.A., Islam, T., Johnston, D., 2012. Intrinsic excitability of CA1 pyramidal
neurones from the rat dorsal and ventral hippocampus. J. Physiol. 590, 5707–5722.
tissue. Dougherty, K.A., Nicholson, D.A., Diaz, L., Buss, E.W., Neuman, K.M., Chetkovich, D.M.,
Johnston, D., 2013. Differential expression of HCN subunits alters voltage-dependent
Funding gating of h-channels in CA1 pyramidal neurons from dorsal and ventral hippocampus.
J. Neurophysiol. 109, 1940–1953.
Eddy, C.M., Rickards, H.E., Cavanna, A.E., 2011. The cognitive impact of antiepileptic
This research was supported through funding from Rhodes College. drugs. Ther. Adv. Neurol. Disord. 4, 385–407.
Engel, J., Henry, T.R., Risinger, M.W., Mazziotta, J.C., Sutherling, W.W., Levesque, M.F.,
Phelps, M.E., 1990. Presurgical evaluation for partial epilepsy: relative contributions
Conflicts of interest of chronic depth-electrode recordings versus FDG-PET and scalp-sphenoidal ictal
EEG. Neurology 40, 1670–1677.
None. Fanselow, M.S., Dong, H.-W., 2010. Are the dorsal and ventral hippocampus functionally
distinct structures? Neuron 65, 7–19.
Hönigsperger, C., Marosi, M., Murphy, R., Storm, J.F., 2015. Dorsoventral differences in
Acknowledgements Kv7/M-current and its impact on resonance, temporal summation and excitability in
rat hippocampal pyramidal cells. J. Physiol. 593, 1551–1580.
We would like to thank Dr Darrin Brager and members of the Kim, C.S., Johnston, D., 2015. A1 adenosine receptor-mediated GIRK channels contribute
to the resting conductance of CA1 neurons in the dorsal hippocampus. J.
Dougherty lab for their continuous support throughout this project. Neurophysiol. 113, 2511–2523.
Kwan, P., Sander, J.W., 2004. The natural history of epilepsy: an epidemiological view. J.
Appendix A. Supplementary data Neurol. Neurosurg. Psychiatry 75, 1376–1381.
Lee, K.S., Reddington, M., Schubert, P., Kreutzberg, G., 1983. Regulation of the strength
of adenosine modulation in the hippocampus by a differential distribution of the
Supplementary material related to this article can be found, in the density of A1 receptors. Brain Res. 260, 156–159.
online version, at doi:https://doi.org/10.1016/j.eplepsyres.2018.05. Malik, R., Johnston, D., 2017. Dendritic GIRK channels gate the integration window,
plateau potentials, and induction of synaptic plasticity in dorsal but not ventral CA1
014. neurons. J. Neurosci. 37, 3940–3955.
Malik, R., Dougherty, K.A., Parikh, K., Byrne, C., Johnston, D., 2016. Mapping the elec-
References trophysiological and morphological properties of CA1 pyramidal neurons along the
longitudinal hippocampal axis. Hippocampus 26, 341–361.
Marcelin, B., Liu, Z., Chen, Y., Lewis, A.S., Becker, A., McClelland, S., Chetkovich, D.M.,
Babb, T.L., Brown, W.J., Pretorius, J., Davenport, C., Lieb, J.P., Crandall, P.H., 1984. Migliore, M., Baram, T.Z., et al., 2012. Dorsoventral differences in intrinsic properties
Temporal lobe volumetric cell densities in temporal lobe epilepsy. Epilepsia 25, in developing CA1 pyramidal cells. J. Neurosci. 32, 3736–3747.
729–740. Milior, G., Di Castro, M.A., Pepe’Sciarria, L., Garofalo, S., Branchi, I., Ragozzino, D.,
Bialer, M., Twyman, R.E., White, H.S., 2004. Correlation analysis between anticonvulsant Limatola, C., Maggi, L., 2016. Electrophysiological properties of CA1 pyramidal
ED50 values of antiepileptic drugs in mice and rats and their therapeutic doses and neurons along the longitudinal axis of the mouse hippocampus. Sci. Rep. 6, 38242.
plasma levels. Epilepsy Behav. 5, 866–872. Ragsdale, D.S., Scheuer, T., Catterall, W.A., 1991. Frequency and voltage-dependent in-
Cembrowski, M.S., Bachman, J.L., Wang, L., Sugino, K., Shields, B.C., Spruston, N., hibition of type IIA Na++ channels, expressed in a mammalian cell line, by local
2016a. Spatial gene-expression gradients underlie prominent heterogeneity of CA1 anesthetic, antiarrhythmic, and anticonvulsant drugs. Mol. Pharmacol. 40, 756–765.
pyramidal neurons. Neuron 89, 351–368. Spina, E., Perugi, G., 2004. Antiepileptic drugs: indications other than epilepsy. Epileptic
Cembrowski, M.S., Wang, L., Sugino, K., Shields, B.C., Spruston, N., 2016b. HippoSeq: a Disord. 6, 57–75.
comprehensive RNA-Seq database of gene expression in hippocampal principal Strange, B.A., Witter, M.P., Lein, E.S., Moser, E.I., 2014. Functional organization of the
neurons. eLife 5, e14997. hippocampal longitudinal axis. Nat. Rev. Neurosci. 15, 655–669.
Churchill, J.D., Fang, P.-C., Voss, S.E., Besheer, J., Herron, A.L., Garraghty, P.E., 2003. Uebachs, M., Opitz, T., Royeck, M., Dickhof, G., Horstmann, M.-T., Isom, L.L., Beck, H.,
Some antiepileptic compounds impair learning by rats in a morris water maze. Integr. 2010. Efficacy loss of the anticonvulsant carbamazepine in mice lacking sodium
Physiol. Behav. Sci. 38, 91–103. channel beta subunits via paradoxical effects on persistent sodium currents. J.
Clemens, A.M., Johnston, D., 2014. Age- and location-dependent differences in store Neurosci. 30, 8489–8501.
depletion-induced h-channel plasticity in hippocampal pyramidal neurons. J. Weir, R.L., Padgett, W., Daly, J.W., Anderson, S.M., 1984. Interaction of anticonvulsant
Neurophysiol. 111, 1369–1382. drugs with adenosine receptors in the central nervous system. Epilepsia 25, 492–498.
Dong, H.-W., Swanson, L.W., Chen, L., Fanselow, M.S., Toga, A.W., 2009. Genomic-
anatomic evidence for distinct functional domains in hippocampal field CA1. Proc.

72

You might also like