Altered Functional Efficacy of Hippocampal Interneuron

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Brain Research Bulletin 131 (2017) 25–38

Contents lists available at ScienceDirect

Brain Research Bulletin


journal homepage: www.elsevier.com/locate/brainresbull

Research report

Altered functional efficacy of hippocampal interneuron during


epileptogenesis following febrile seizures
Yeon Hee Yu, Kahyun Lee, Dal Sik Sin, Kyung-Ho Park, Dae-Kyoon Park ∗ , Duk-Soo Kim ∗
Department of Anatomy, College of Medicine, Soonchunhyang University, Cheonan-Si, Chungcheongnam-Do, 31151, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: Febrile seizure (FS) is the most common seizure type in infants and young children. FS may induce
Received 7 October 2016 functional changes in the hippocampal circuitries. Abnormality of excitatory and inhibitory neurotrans-
Received in revised form 17 February 2017 missions was previously related to wide-spread seizure attack in the hippocampus following recurrent
Accepted 23 February 2017
seizure onset. To clarify the involvement of expressional changes and functional alterations of hippocam-
Available online 7 March 2017
pal interneurons with epileptogenesis following FS, we investigated long-term effects following recurrent
seizure in a hyperthermia-induced seizure animal model. At 12 weeks following FS, the recurrent seizure
Keywords:
time period, local field potentials (LFP) revealed high amplitude potential and a sharp wave characteris-
Febrile seizure
GABAA -␣1
tic of epilepsy. Mossy fiber reorganization in the hippocampus was also detected as abnormal synaptic
Calretinin connection at 8 weeks. Calretinin (CR) −positive interneurons were transiently enhanced during epilep-
Field excitatory postsynaptic potential togenic period at 7–9 weeks after FS in the CA1 and DG region and it is double labeled with VGLUT-1.
Epileptogenesis However, although GABAA -␣1 immunoreactivities were un-changed as similar to control hippocampus
Paired-pulse response at 7–9 weeks after seizure onset, its expression was significantly enhanced at 4 weeks and 12 weeks
and it is colocalized with GABA. Furthermore, the field excitatory postsynaptic potential (fEPSP) and
the paired-pulse responses including population spike (PS) latency, excitability ratio and PS2/PS1 ratio
were markedly altered in the CA1 and DG region at 12 weeks after FS. Therefore, our findings in present
study indicate that these time-dependent changes may be based on the persistent alterations of hip-
pocampal neuronal circuits in balance between excitatory and inhibitory responses, and may lead to the
epileptogenesis and spread of seizure activity following FS.
© 2017 Elsevier Inc. All rights reserved.

1. Introduction may sustain enhanced hippocampal excitatory circuits and con-


tribute toward the development of temporal lobe epilepsy (TLE)
Febrile seizure (FS) induced by fever is the most common seizure by enhanced hyperexcitatory neuronal circuit to epileptogenesis
type in infants and young children, which occurs in 3–5% of chil- during developmental stage in infant (Dubé et al., 2006).
dren between 6 months and 5 years of age (Toth et al., 1998; Virta ␥-Aminobutyric acid (GABA) is the major inhibitory neuro-
et al., 2002; Bender et al., 2003). Animal models have indicated that transmitter. It is affected by the intermediation of GABAA and
seizures of immature animals through neurogenesis in the dentate GABAB receptors in the brain (Brooks-Kayal et al., 1998; Matthews
gyrus (DG) at postnatal 1–2 weeks may more effectively induce et al., 1998). GABAA receptors are main inhibitory receptors that
the innervations of newly generated granule cells (Parent et al., mediate many pathophysiological aspects (Sieghart et al., 1999).
1997; Koyama et al., 2012; Scott and Holmes, 2012). Therefore, pro- However, calretinin (CR)-positive interneurons innervate diverse
longed FS may induce the alteration in hippocampal circuitry (Dube interneuronal subtypes in cortical areas and have an apprecia-
et al., 2000; Bender et al., 2003; Kwak et al., 2008). These changes ble disinhibitory effect on pyramidal neurons in the hippocampus
(Barinka and Druga, 2010; Barinka et al., 2012). CR-positive
interneurons control the contact of other interneuronal dendritic
and exodendritic with each other and selectively innervate several
Abbreviations: CR, calretinin; DG, dentate gyrus; LFP, Local field potentials; FS,
Febrile seizure; fEPSP, field excitatory postsynaptic potential; GABA, ␥-aminobutyric interneurons (Gulyás et al., 1999; Bausch, 2005; Barinka and Druga
acid; LTP, Long-term potentiation; PBS, phosphate-buffered saline; PS, population et al., 2010).
spike; SE, status epilepticus; sTPS, strong theta-patterned stimulation; TLE, temporal GABAergic interneurons are decreased in the hippocampus of
lobe epilepsy.
various epileptic models; these declines may involve epileptogen-
∗ Corresponding authors.
esis (Sloviter et al., 1991; Fritschy et al., 1999; Bouilleret et al., 2000;
E-mail addresses: mdeornfl@sch.ac.kr (D.-K. Park), dskim@sch.ac.kr (D.-S. Kim).

http://dx.doi.org/10.1016/j.brainresbull.2017.02.009
0361-9230/© 2017 Elsevier Inc. All rights reserved.
26 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

Bausch, 2005; Kwak et al., 2005; Sloviter et al., 2006; Kwak et al., scale criteria [30,stage 1, mouth and facial movements; stage 2,
2008). Moreover, interneuron CR and GABAergic cells have been head nodding; stage 3, forelimb clonus; stage 4, rearing; stage 5,
implicated in epileptic animal models and in humans (Maglóczky rearing and falling].
et al., 2000; André et al., 2001; Slézia et al., 2004; Van Vliet
et al., 2004; Tóth et al., 2010). GABAB receptor and parvalbumin- 2.3. Local field potentials (LFP)
mediated inhibitory alterations in the hippocampus have been
documented in an animal model of FS (Kwak et al., 2008). How- At the designated times (1–12 weeks following control and FS,
ever, the effects of other interneuronal alterations including GABAA n = 20 respectively), each animal was anesthetized by intraperi-
receptor subunits and CR-positive interneurons during epilepto- toneal injection of urethane 1.5 g/kg and placed in a stereotaxic
genic periods following FS in the hippocampus remain unclear. frame. Holes were drilled through the skull for the introduction of
We investigated the temporal involvement of expressional patterns electrodes. In young animals (1–6 weeks after FS), LFP was recorded
and functional alterations of GABAA -␣1 and CR in the hippocam- using a tungsten parylene electrode (0.005-inch outer diameter;
pal interneurons with epileptogenesis following recurrent seizure A-M Systems, USA). The coordinates (in mm) referenced to the
onset. bregma were as follows (to neocortex): 1.0 posterior to bregma,
1.0 lateral to midline, 0.7 depths. In older animals (7–12 weeks
after FS), glass microelectrodes (microfilament capillary 1.2 outer
2. Materials and methods
diameter; 5–10 M) filled with artificial cerebrospinal fluid (ACSF,
in mM; NaCl 126, KCl 5, CaCl2 2, MgCl2 2, NaH2 PO4 1.25, NaHCO3
2.1. Experimental animals
26, d-glucose 10, pH 7.2) were used. The coordinates (in mm) ref-
erenced to bregma were as follows (in mm to dentate gyrus): 3.8
All experiments utilized the progeny of Sprague-Dawley (SD)
posterior to bregma, 2.5 lateral to midline, 2.9 depths. Signals were
rats obtained from Experimental Animal Center, Soonchunhyang
recorded with QP511 AC amplifier (0.1–3000 Hz bandpass, GRASS
University (Cheonan, South Korea). All animals were provided with
Technologies, USA) and data were digitized (5 kHz) and recorded
a commercial diet and water ad libitum under controlled tempera-
to obtain the baseline value for 2 h. The single-channel acquisition
ture, humidity and lighting conditions (light/dark cycle 12:12, and
was performed using the Axoscope 10.2 (Axon Instruments, USA)
22 ± 2 ◦ C, 55 ± ; 5%). All animal protocols were approved by the
software. Analysis of the single-channel electrical traces was car-
Administrative Panel on Laboratory Animal Care of Soonchunhyang
ried out using the Clampfit 10.2 (Axon Instruments, USA) software.
University (permit No. SCH15-0002). All possible efforts were to
To analyze changes in normalized power of LFP, the amplitude
avoid suffering of the rats and to minimize the number used during
spectrum analysis of normalized power was estimated by event
the experiments.
frequency, and the root mean square (RMS) values were used to
derive estimates of spectral power (mV2 ) in the 1 Hz frequency
2.2. FS induction bins for each electrode site. Spectral power values were averaged
across all epochs within a single base-line, the resulting power
The hyperthermic seizure rat model of FS has been described was expressed as mV2 /Hz. For each subject, fast fourier transform
(Kwak et al., 2008). Briefly, rat pups (postnatal 11 days) were used, (FFT) of the epochs with a resolution of 0.61 Hz was computed
because hippocampal developmental at that age is generally equiv- for all electrodes and then averaged. Non-overlapping hamming
alent to human infants (Baram et al., 1997; Gottlieb et al., 1977; windows controlled spectral leakage. Moreover, The FFT power
Scantlebury et al., 2005). After punch-marking of their ears, pups value measurements within each frequency between 1 and 50 Hz
were warmed in a plastic chamber measuring 10 × 13 cm at the were averaged to create 50 non-overlapping <1 Hz frequency bins,
base and 12 cm in height. The chamber floor was covered with because the frequency bands of interest were defined as: ␦ (1–4 Hz),
paper towels. A 175W mercury vapor lamp was held 3 cm above ␥ (25.0–50 Hz) (Schutter et al., 2003; Lee et al., 2004; Scheffer-
the chamber. Core temperatures were measured before the induc- Teixeira et al., 2013; Fuggetta et al., 2014). After LFP recording, some
tion of hyperthermic seizure and checked every 5 min using an ear animals were used for paired-pulse responses, immunoreactivity,
thermoprobe. All animals showed generalized seizures at 5–10 min immunofluorescence and Timm’s staining.
after hyperthermic seizure induction, when core temperature was
41–43 ◦ C. Hyperthermic seizures were maintained in 40 min after 2.4. Paired-pulse responses
generalized seizure onset because the duration for febrile seizures
are maintained minimum 15 min more than one seizure in a 24 h The paired-pulse responses from hippocampus following hyper-
period, result in transient neuronal injury and associate adjust- thermic seizure were recorded (10–12 weeks after hyperthermic
ing epilepsy (Maytal et al., 2000; Dubé and Baram, 2006; Graves seizure and control, n = 10 respectively) using published protocols
et al., 2012). During maintenance of hyperthermic seizures, ani- (Kim et al., 2007, 2008), with some modifications. Briefly, animals
mals showed multiple generalized seizures (Racine, 1972; stage 4 were anesthetized (urethane, 1.5 g/kg, i.p.) and placed in a stereo-
and 5). After hyperthermic seizure, rats were moved to a cool sur- taxic frame. Holes were drilled through the skull for introducing
face, and were then returned to their mothers (mortality, n = 9 out electrodes. The coordinates (in mm) referenced to bregma were
of 315, 3% approximately). Siblings of the rats were used as controls, as follows. For the unipolar recording electrode (to the dentate
and they were placed in a chamber at room temperature. In order gyrus); 3.8 posterior to bregma, 2.0 lateral to midline, 3.5 depths.
to completely identify FS inductions, we were observed behavior For the bipolar stimulating electrode (to the angular bundle);
at recurrent seizures stage through video monitoring for 24 h. As a 8.0 posterior to bregma, 4.4 laterals to midline, 4.0 depths were
result, recurrent seizure was observed at interval of approximately placed. Electrode depths were finally determined by optimizing the
4 h each day in the vivarium for general behavior (Racin scale crite- evoked response. Glass microelectrodes (microfilament capillary
ria, 2.3 ± 0.07). In addition, we have double cross-checked under 1.2 outer diameters; 5–10 M◦ ) filled with artificial cerebrospinal
EEG monitoring for observation of FS induction. Thus, we have fluid (ACSF, in mM: NaCl 126, KCl 5, CaCl2 2, MgCl2 2, NaH2 PO4
utilized only completely inducted animal model for FS induction 1.25, NaHCO3 26, d-glucose 10, pH 7.2) and bipolar tungsten stimu-
animal groups in this study. After recurrent seizure onset (10–12 lating electrodes were used. Body temperature was monitored and
weeks after hyperthermic seizure induction, n = 95 out of 134, ∼71% maintained at 37 ± 0.3 ◦ C by thermostat during recording. Stimuli
approximately), behavioral seizures were scored based on Racine were applied as DC square pulses at 0.5 Hz with pairs of 150 ␮s
Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38 27

constant current stimuli at 30-ms inter-stimulus intervals (GABAA ware (WinLTP Ltd.). Responses were evoked by single pulse stimuli
receptor-mediated inhibition, stimulus intensity; 2 × thresholds) and were delivered at 20-s intervals. A stable baseline was recorded
using a Digital Stimulus Isolation unit (Getting Instruments, CA, for 30–60 min. Long-term potential (LTP) was induced by the strong
USA) depending on the threshold intensity evoking the popula- theta- patterned stimulus (sTPS, four trains of 10 bursts of 5 pulses
tion spike, whereas 70-ms and 150-ms of inter-stimulus interval at 400 Hz with a 200-ms inter-burst interval, 15-s inter-train inter-
revealed the measurements of facilitation and slow inhibition val), because sTPS (bursts of 400 Hz stimuli) are NMDA-receptor
(GABAB receptor-mediated inhibition, Kwak et al., 2008). Dur- dependent and robust LTP of CA1 and DG areas are evoked by
ing baseline recording, the population spike threshold of dentate sTPS in vivo (Cho et al., 2013; Farmer et al., 2004). To analyze the
gyrus was determined and stimulus intensity was increased until changes of fEPSP, slopes of fEPSP were averaged over 60-s intervals
a population spike was detected consistently. One hour after base- and expressed as percentages of the mean fEPSP slope measured
line recording, paired-pulse responses were recorded. To obtain during the 30-min baseline period, which was expressed as 100%.
input/output curve, stimulus intensities were changed from 0.5 to The responses of fEPSP for age-matched control animal were also
4 × thresholds. Then responses were recorded in each train with measured using the same method.
30 min inter-train intervals. Signals were recorded with a P55 A.C.
pre-amplifier (3–1000 Hz bandpass, Astro-Med Inc., USA) and data 2.6. Timm’s staining
were digitized (20 kHz) and analyzed using WinLTP ver 2.01 soft-
ware (WinLTP Ltd., USA). To analyze changes in evoked responses, For identifying mossy fiber reorganization, Timm’s staining was
population spike (PS) amplitude was normalized by the baseline performed. Briefly, one to 12 weeks of the FS and control animal
values (the average of the PS amplitude of the first response) model (n = 5 respectively) were intraperitoneally anesthetized with
from the same animal and the same recording session. The ratio urethane 150 mg/kg and perfused through the heart by 0.9% saline
of the second PS amplitude/the first PS amplitudes was deter- for 10 min, by 0.6% sodium sulfide and 0.6% sodium phosphate for
mined to measure GABAA -mediated inhibition. In addition, the 10 min, and finally by 4% paraformaldehyde for 10 min. The same
latency and amplitude of the PS following first evoked response method was used to perfuse age-matched control animals. The
were analyzed for measuring the extent of glutamatergic synap- brains were removed and postfixed in the same fixative for 12 h,
tic transmission in the dentate gyrus before and after recurrent and rinsed in phosphate buffer (pH 7.4, PB) containing 30% sucrose
seizure onset (Feng and Durand, 2004; Matzen et al., 2008). PS at room temperature for 1 day. Thereafter, the tissues were frozen
amplitude and PS latency in first evoked responses were normal- and sectioned with a cryostat at 30 ␮m and consecutive sections
ized by the baseline values from the same animal and the same were collected in six-well plates containing phosphate buffered
recording session. Furthermore, in order to investigate the effi- saline (PBS). Some hippocampal tissues attached to gelatin-coated
ciency of glutamatergic synaptic transmission in the dentate gyrus slides were incubated in the dark in a solution consisting of one-
the excitability ratio was calculated as the PS amplitude versus part solution A (1 M AgNO3 ), 20 parts solution B (2% hydroquinone
the field excitatory postsynaptic potential (fEPSP) slope in the first and 5% citric acid in water), and 100 parts solution C (20% gum
evoked response (Kwak et al., 2008). Briefly, the averages of the arabic in water). Development took about 3 h, was monitored by
PS amplitude and the fEPSP slope of the first response at stim- periodic evaluations under low light at 36.5 ◦ C. The reaction was
ulus intensity 2 × threshold were used to normalize all of the PS terminated by washing in water.
amplitudes and the fEPSP slope measurements during the record-
ing session from the same animal and the same recording session. 2.7. Immunohistochemistry and immunofluorescence
Paired-pulse responses for age-matched control animals were also
measured using the same method. After recording, animals were Anesthesized animals (urethane 1.5 kg/kg) were perfused tran-
used for immunohistochemistry, immunofluorescence or Timm’s scardially with PBS followed by 4% paraformaldehyde in 0.1 M PB
staining. (1–12 weeks after hyperthermic seizure and control, n = 15 respec-
tively). The brains were removed and postfixed in the same fixative
2.5. In vivo field EPSP recordings for 4 h, and rinsed in PB containing 30% sucrose at 4 ◦ C for 2 days.
Thereafter, the tissues were frozen and sectioned with a cryo-
Field potentials expressed as fEPSPs were recorded from the stat at 30 ␮m thickness and consecutive sections were collected
CA1 of the hippocampus as described (Wong et al., 2007; Ge et al., in six-well plates containing PBS. For cell counts, every sixth sec-
2010; Cho et al., 2013), with some modifications. At the recur- tion in the series throughout the entire hippocampus was used
rent seizure time period (>7 weeks after FS and control, n = 10 in the same animals. Sections were incubated with each primary
respectively), animals were intraperitoneally anesthetized using antibody in PBS containing 0.3% Triton X-100 overnight at room
urethane 1.5 g/kg and placed into a stereotaxic frame. Rectal tem- temperature: mouse anti-CR IgG (Chemicon, USA; diluted 1:1000)
perature was maintained at 37 ± 0.3 ◦ C during the course of the and rabbit anti-GABAA -␣1 IgG (Millipore, USA; diluted 1:400). The
surgery using a temperature controller (Harvard Instruments, USA). sections were washed three times for 10 min with PBS, incubated
The scalp was opened and separated. Holes were drilled through sequentially in biotinylated goat anti-mouse and goat anti-rabbit
the skull for introducing electrodes. The coordinates (in mm) ref- IgG (Vector, USA) and ABC complex (Vector), and diluted 1:200 in
erenced to bregma were as follows. For the recording electrode the same solution as the primary antiserum. Between the incuba-
(to the Schaffer collateral): 4.0 posterior to bregma, 3.0 lateral to tions, the tissues were washed with PBS three times for 10 min each.
midline, 2.5 depths. For the stimulating electrode (to the stratum The sections were visualized with 3,3 -diaminobenzidine (DAB)
radiatum of CA1): 3.5 posterior to bregma, 2.0 lateral to midline, 3.5 in 0.1 M Tris buffer and mounted on gelatin-coated slides. The
depths were positioned in the hippocampus. Electrode depths were immunoreactions were observed using DMRB microscope (Leica,
finally determined by optimizing the evoked response. fEPSPs were Germany) and images were captured using a model DP72 digital
adjusted to ∼60% of maximal response size for testing. Stimulation camera and DP2-BSW microscope digital camera software (Olym-
was generated by a BNC-2110 apparatus (National Instruments, pus, Japan). To establish the specificity of the immunostaining,
USA) and a Digital Stimulus Isolation unit (Getting Instruments, a negative control test was carried out with pre-immune serum
CA, USA). Pyramidal neuron responses to the Schaffer collateral instead of primary antibody. The negative control resulted in the
stimulation were recorded with a P55A.C. pre-amplifier (3–1000 Hz absence of immunoreactivity in any structures. To identify the
bandpass, Astro-Med Inc.) and analyzed using WinLTP ver 2.01 soft- morphological changes induced by hyperthermic seizure in the
28 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

same hippocampal tissue, double immunofluorescence staining for sections sampled or section sampling fraction [minimum six tis-
CR/vesicular glutamate transporter-1 (VGLUT-1) (7–9 weeks after sues per each animal models (1–12 weeks FS and control groups,
hyperthermic seizure and control, n = 15 respectively) and GABAA - n = 15 respectively) in this study, since we used all sections in
␣1/GABA (10–12 weeks after hyperthermic seizure and control, the entire hippocampus]. The immunoreactive cells were counted
n = 15 respectively) was performed. Brain tissues were incubated with a 40 × objective lens. All immunoreactive cells were counted
overnight at room temperature in mixture of rabbit anti-CR IgG regardless the intensity of labeling. Cell counts were performed
(Abcam, USA, diluted 1:500)/guinea pig anti-VGLUT-1 IgG (Chemi- by two different investigators who were blind to the classifica-
con, USA, diluted 1:2500) and rabbit anti-GABAA -␣1 IgG (Millipore, tion of tissues. For quantification of immunodensity [4,8 and 12
USA; diluted 1:200)/mouse anti-GABA IgG (Abcam, UK; diluted weeks after FS and control groups for WB (n = 5 respectively); 1–12
1:300). After washing three times for 10 min with PBS, sections weeks following FS and control groups for Timm’s staining (n = 5
were also incubated in a mixture of Cy2- and Cy3-conjugated sec- respectively)], CA1 and DG were delineated with a 2.5 × objective
ondary antisera (1:200, Amersham, USA) for 1 h room temperature. lens. Each image was normalized by adjusting the black and white
Sections were mounted in Vectashield mounting media with DAPI range of the image using Adobe PhotoShop v. 8.0. Thereafter, 10
(Vector, USA). All images were captured using a model Fluoview areas per rat (250 ␮m2 for each area) were selected and intensity
FV10i and FV10i software (Olympus, Japan). measurements represented as the mean number of a 256 gray scale
(using NIH Image 1.59 software). Values of background staining
2.8. Western blot were obtained from the corpus callosum. Optical density values
were corrected by subtracting the average values of background
Based on the immunohistochemical results, of the expression noise obtained from five image inputs. All data obtained from the
of CR and GABAA -␣1 protein was quantified in the control and FS quantitative measurements were analyzed using one-way analysis
hippocampus as previously described (Blancher and Jones, 2001; of variance (ANOVA) to determine statistical significance. Bonfer-
Mahmood and Yang, 2012). At designated times (4, 8 and 12 weeks roni’s test was used for post-hoc comparisons. A p-value < 0.01
after FS and control, n = 5 respectively), immunoblot analysis was or <0.05 was considered statistically significant (Kim et al., 2007,
performed. After sacrifice and removal of the hippocampus, the 2008).
tissues were homogenized in 50 mMTris-HCl (pH 8.0) contain-
ing 150 mM NaCl, 50 mM NaF, 0.5% NP-40, 5 mM EDTA, 1 mM
phenyl methyl sulfonyl fluoride (PMSF), and 1 mMA protinin. After
centrifugation at 14000 rpm, the protein concentrations in the 3. Results
supernatants were determined as previously described (Bradford,
1976). Proteins were separated by 12% SDS-PAGE. Equal amounts 3.1. Representative LFP profiles following FS
of protein (30 ␮g in all assays) were loaded in each lane with
loading buffer containing 1 M Tris-HCl (pH 6.8), 20% glycerol, 10% Distinguishing characteristic of electroencephalography in the
SDS, 25% ␤-mercaptoethanol and 0.002% bromphenol blue. Sam- control and epileptic hippocampus following FS were showed sig-
ples were heated at 100 ◦ C for 5 min before gel loading. After nificant differences. Although control rats showed a normal LFP in
electrophoresis, the gels were transferred to nitrocellulose trans- the 8 weeks (Fig. 1A1 and C1) and 12 weeks (Fig. 1A3 and C3),
fer membranes (Bio-Rad, USA). To reduce background staining, LFP signals at 12 weeks following FS showed epileptiform dis-
the filters were incubated with 5% non-fat skim milk in PBS for charges, which was characterized intermittently high amplitude
1 h and sequentially incubated with the primary antibody in PBS field potentials and sharp waves in DG and CA1 as compared to
overnight at 4 ◦ C. Antibody used was mouse anti-CR IgG (Chemicon, control (Fig. 1A4 and C4), whereas its signal at 8 weeks follow-
USA; diluted 1:1000) or rabbit anti-GABAA -␣1 IgG (Millipore, USA; ing FS revealed as similar to control groups (Fig. 1A2 and C2). In
diluted 1:200). After washing, the membranes were incubated with addition, the power spectral analysis from the LFP recording of
peroxidase-conjugated goat anti-rabbit IgG secondary antibody the hippocampus following FS showed strong power in a lower
for 1 h (Jackson ImmunoResearch Inc., USA; diluted 1:10000) and frequency of the hippocampus than the control (Fig. 1B1 and
anti-mouse IgG (Santa Cruz Biotechnology, USA; diluted 1:2000). D1), and its normalized power in DG and CA1 was more pow-
Immunoblots were visualized using an ECL kit (Advansta Inc., USA), erful in the FS hippocampus (FS 8 weeks, 0.414 ± 0.08 in CA1,
and the protein density was quantified by densitometric analysis 0.456 ± 0.02 in DG; FS 12 weeks, 0.99 ± 0.25 in CA1, 0.45 ± 0.02 in
using Image J software (NIH, USA). DG) as compared with the control (control 8 weeks, 0.248 ± 0.01
in CA1, 0.198 ± 0.03 in DG; control 12 weeks, 0.02 ± 0.25 in CA1,
2.9. Quantification of data and statistical analysis 0.02 ± 0.08 in DG; P < 0.01, Fig. 1B2 and 1D2). Especially, the delta
rhythm synchronous neuronal activity of hippocampus, which as
An optical fractionation was used to estimate the cell num- a unique LFP signaling marker of epilepsy (Di Gennaro et al.,
bers. The optical fractionators (combination of performing counting 2003; Gelisse et al., 2011), was significantly elevated in the hip-
with the optical dissector, with fractionators sampling) is a stereo- pocampus of 12 weeks following FS than in the control (FS 12
logical method based on a properly designed systematic random weeks, 67.40 ± 1.21 in CA1, 57.89 ± 2.97 in DG; control 12 weeks,
sampling method that by definition yields unbiased estimates 60.27 ± 1.38 in CA1, 41.78 ± 2.03 in DG; P < 0.01, Fig. 1B3 and D3),
of population number. The sampling procedure is accomplished whereas its signal at 8 weeks after FS was similar to control
by focusing through the depth of the tissue (the optical dissec- groups (FS 8 weeks, 41.18 ± 3.76 in DG, 60.75 ± 1.70 in CA1; con-
tor height, h; of 15 ␮m in all cases for this study). The number trol 8 weeks, 40.86 ± 2.80 in DG, 54.68 ± 3.84 in CA1). However,
of each cell type (C) in each of the subregion is estimated as: the gamma rhythm associated with hyper brain activity, which is
C = ˙Q− × t/h × 1/asf × 1/ssf, where Q− is the number of cells actu- linked to GABAergic interneurons (White et al., 2000; Wendling
ally counted in the dissectors that fell within the sectional profiles et al., 2002; Hughes, 2008), had a significantly declined rate in
of the subregion seen on the sampled sections, and asf is the areal the hippocampus of 12 weeks following FS than in the control
sampling fraction calculated by the area of the counting frame of the (FS 12 weeks, 2.17 ± 0.44 in CA1, 2.20 ± 0.55 in DG; control 12
dissector, a(frame) (50 × 50 ␮m2 in this study) and the area asso- weeks, 3.47 ± 0.31 in CA1, 10.02 ± 2.16 in DG; P< 0.01, P< 0.05,
ciated with each x, y movement grid (x, y step) (of 250 × 250 ␮m2 Fig. 1B4 and 1D4), although its signals at 8 weeks following FS
in this study) {asf = [a(frame)/a(x,y step)]}. ssf is the fraction of the shown similar result with control groups (FS 8 weeks, 3.042 ± 0.47
Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38 29

Fig. 1. Local field potential profiles in DG and CA1 between the control and the epileptic hippocampus following FS. Normal control rats show a general representative LFP
signals at 8 weeks (A1 and C1) and 12 weeks of control groups (A3 and C3). At the recurrent seizure time period (12 weeks following FS), the LFP signals show significant
epileptiform discharges, as large amplitude spikes of irregularly sharp wave and multi-spikes (A4 and C4), although its signal is similar to control group at 8 weeks after FS
(A2 and C2). Power spectral analysis after FS revealed more power in lower frequency than the control group (B1 and D1). Representative average normalized power of 12
weeks after FS shows strong power more than control (B2 and D2). At the same time period, although synchronous neuronal activity of delta wave is markedly elevated,
gamma wave declines compared with control (B3-B4 and D3-D4), whereas these signals at 8 weeks after FS show strong normalized power only in the DG of hippocampus
(B2-B4 and D2-D4). All data are presented as mean ± SEM. Significant differences from the control. *p < 0.05, **p < 0.01, vs. control.

in CA1, 10.70 ± 0.53 in DG; control 8 weeks, 5.880 ± 1.25 in CA1, than control levels as shown synaptic reorganizations (P < 0.01,
9.907 ± 2.95 in DG). Fig. 2C1–C2, D1–D2 and E).

3.3. Immunohistochemical characterizations of CR and


3.2. Mossy fiber reorganization following hyperthermic seizure GABAA -˛1 expression following FS

In order to investigate the anatomical synaptic alterations of To identify the functional alteration of interneurons in the
development and adult stage following FS, we identified the ter- hippocampus following hyperthermic seizure, immunohistochem-
minals of mossy fiber using Timm’s staining. Although diverse istry analyses for CR and GABAA -␣1 receptors were done. The
synaptic fibers of mossy fiber in the dentate gyrus were observed immunoreactivities of CR −positive interneurons in the control
in the development stage of control (control 2 weeks, Fig. 2A1 and were observed in the CA1 and DG of hippocampus in the develop-
A2), these fibers were declined on the hilar region following hyper- ing (Fig. 3A1 and A2) and adult hippocampus (Fig. 3C1 and C2). At
thermic seizure, especially in the granule cell layer as compared 4 weeks after FS, CR immunoreactivity is enhanced only in the DG
to controls (44.15 ± 2.03% in FS 2 weeks; 100 ± 2.53% in control 2 than control (14 ± 0.70 in FS 4 weeks; 8.5 ± 1.06 in control 4 weeks;
weeks; P < 0.01, Fig. 2B1–B2 and E). At the adult stage (8 weeks), the P < 0.05, Fig. 3A1-A2, B1-B2 and F). No differences of immunoblot
axonal processes of mossy fiber in the dentate gyrus following FS analysis were apparent at 4 weeks of development (99.56 ± 26.77%
were significantly enhanced in the inner molecular (272.05 ± 2.58% in FS 4 weeks; 100 ± 56.87%, in control 4 weeks; Fig. 3G and H). At 8
in FS 8 weeks; 100 ± 2.55% in control 8 weeks) and granule cell layer weeks following FS, CR-positive interneurons were markedly ele-
(210.31 ± 3.02% in FS 8 weeks; 100 ± 2.53% in control 8 weeks) more vated in each region more than control level (FS 8 weeks, 14.5 ± 0.35
30 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

Fig. 2. Timm’s staining in the dentate gyrus of control and epileptic hippocampus following FS during developmental (A1-B2) and adult stages (C1-D2). The staining reveals
dark labeling mainly in the hilar region of control animals including some of the granule cell layer (arrow, A1 and A2), while its labeling is reduced in the DG (GCL, HL and
IML) at 2 weeks following hyperthermic seizure (B1 and B2). At 8 weeks after recurrent seizure onset, the dense mossy fiber sprouting in the DG, including GCL and IML,
is observed (arrow, D1 and D2) more than in the control (C1 and C2). Rectangles in panels A1, B1, C1 and D1 indicate the high-magnification of panels A2, B2, C2 and D2.
Bar = 200 ␮m (panels A1, B1, C1 and D1), 25 ␮m (panels A2, B2, C2 and D2). Densitometric analysis for these staining is obtained same results in the GCL and IML (E). All data
are presented as mean ± SEM. Significant differences from the control. **p < 0.01, vs. control. Abbreviations: GCL, granule cell layer; HL, hilar region; IML, inner molecular
layer.

in DG, 14 ± 0.70 in CA1; control 8 weeks, 8 ± 0.70 in DG, 11 ± 0.17 15.5 ± 0.35 in CA1, 25 ± 0.77 in DG; P < 0.05, Fig. 5B1-B3 and G).
in CA1; P < 0.05, Fig. 3C1–C2, D1–D2 and F). WB analysis for CR Immunoblot analysis also revealed same results with immuno-
expression also revealed significantly enhancements at 8 weeks histochemical data at this time points (189.62 ± 19.66% in FS 4
following FS more than control groups (265.30 ± 44.62% in FS 8 weeks; 100 ± 14.54% in control 4 weeks; P < 0.01, Fig. 5I and J). How-
weeks; 100 ± 48.30% in control 8 weeks; P < 0.01, Fig. 3G and H). ever, GABAA -␣1 immuonreactive interneurons at 8 weeks after FS
In addition, at this time period, CR-positive interneurons were were significantly down-regulated both regions as similar to con-
double labeled with VGLUT-1 immunoreactivities in the CA1 and trol groups (FS 8 weeks, 20.75 ± 0.51 in CA1, 22.5 ± 0.77 in DG;
DG region following FS, and its number of colocalized interneu- control 8 weeks, 23 ± 1.41 in CA1, 19.5 ± 0.35 in DG; Fig. 5C1–C3,
rons were significantly enhanced more than control 8 weeks (FS D1-D3 and G), and western blot experiments also shown very sim-
8 weeks, 226.98 ± 53.99% in CA1, 244.44 ± 33.89% in DG; con- ilar results with immunohistochemistry data (98.74 ± 14.13% in FS
trol 8 weeks, 100 ± 25.25% in CA1, 100 ± 22.22% in DG; P < 0.05, 8 weeks; 100 ± 13.02% in control 8 weeks; Fig. 5I and J). At the
P < 0.01, Fig. 4A1–D4, E and F). However, the immunoreactivity of recurrent seizure time period (12 weeks following hyperthermic
CR-positive interneurons at recurrent seizure period (10–12 weeks seizure), the numbers of GABAA -␣1 positive interneurons were re-
following hyperthermic seizure) was significantly down-regulated enhanced in the CA1 and DG regions as compared to the control
in the CA1 (9.5 ± 0.62 in FS 12 weeks; 8.5 ± 0.35 in control 12 weeks) group (FS 12 weeks, 32 ± 0.71 in CA1, 30.75 ± 0.65 in DG; control
and DG regions (11.75 ± 0.23 in FS 12 weeks; 10 ± 0.70 in control 12 weeks, 22.5 ± 0.35 in CA1, 22.5 ± 0.35 in DG; P < 0.01, Fig. 5E1-
12 weeks), similar to control 8 and 12 weeks (Fig. 3E1–E2 and F). E3, F1-F3 and G). At the same time periods, immunofluorescence
Immunoblot analysis also revealed similar results with immuno- staining revealed that GABAA -␣1 positive interneurons were dou-
histochemistry data at 12 weeks after FS (95.00 ± 30.82% in FS 12 ble labeled with GABA immunereactivity in both regions, and
weeks; 100 ± 48.30% in control 8 weeks; Fig. 3G and H). its numbers of colocalized interneuronal population at 12 weeks
On the other hand, the immunoreactivity of GABAA -␣1 positive following recurrent seizure onset were significantly enhanced
interneurons was distinguished from the CR expressions follow- more than control groups (Fig. 5H1–H4). Immunoblot and den-
ing FS. Briefly, GABAA -␣1 immunoreactivities at 4 weeks were sitometric analysis of GABAA -␣1 protein also showed as similar
detected in the diverse interneruons of CA1 and DG regions in results with immunohistochemical data (214.55 ± 16.64% in FS
the control hippocampus (Fig. 5A1–A3), whereas its expression 12 weeks; 100 ± 13.02% in control 8 weeks; P < 0.01, Fig. 5I and
was remarkably elevated age-matched hippocampus following FS J).
(FS 4 weeks, 24.5 ± 1.76 in CA1, 32 ± 0.70 in DG; control 4 weeks,
Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38 31

Fig. 3. Calretinin (CR) immunoreactivities and quantitative analyses in the control and the epileptic hippocampus following FS (A1- F). At four weeks after FS, CR immunore-
activity is markedly enhanced in the DG than control (open arrow, A2 and B2), whereas its expression in CA1 similar to control groups (A1 and B1). CR-positive interneurons
at 8 weeks following hyperthermic seizure in hippocampus are enhanced in CA1 and DG (open arrow, D1 and D2) as compared to control levels (C1 and C2). At recurrent
seizure time period (12 weeks following FS), the expressions of CR-positive interneuron are significantly reduced as similar to controls groups (E1 and E2). Bar = 50 ␮m
(panels A1- E2). The results of average cell number for CR-positive interneurons are shows as similar with immunohistochemical data (F). The immunoblot (G) and optical
density analyses (H) are revealed although the CR protein unchanged at 4 weeks after FS as compared to control, its protein level markedly elevated in the hippocampus at
8 weeks following seizure onset. All data are presented as mean ± SEM. Significant differences from the control, *p < 0.05, **p < 0.01, vs. control.

3.4. Paired-pulse responses and fEPSP following FS PS amplitude (PS2/PS1), were calculated to evaluate the change
of the neuronal excitability, synaptic transmission, and GABAergic
To identify the glutamatergic synaptic transmissions by NMDA inhibition from interneurons (Feng and Durand, 2004). PS ampli-
receptors in the hippocampal CA1 region following FS because of tude after tetanic stimulation was increased at 12 weeks following
that NMDA receptor is an important role in regulating the GABAer- FS to 149.1 ± 5.7% (137.8 ± 7.0% in control 12 weeks; Fig. 6E),
gic input by GABA release (Maccaferri and Dingledine, 2002), we whereas the PS latency after tetanic stimulation was significantly
investigated the activity-dependent synaptic plasticity under the down-regulated as compared to the condition before tetanic stim-
induction of NMDA receptors-mediated LTP at the time of recur- ulation to 90.5 ± 0.5% (100 ± 1.5% in control 12 weeks; P < 0.01,
rent seizure onset. The fEPSPs evoked by electrical stimulation Fig. 6E).
of the Schaffer collateral pathway at 0.05 Hz were recorded from
the stratum radiatum of the CA1 region. After a stable baseline
recording was obtained, LTP was induced by sTPS (four trains of
10 bursts of 5 pulses at 400 Hz with a 200 ms inter-burst inter-
val, 15-s intertrain interval), since robust LTP of CA1 and DG
areas are evoked by sTPS in vivo (Cho et al., 2013; Farmer et al., 4. Discussion
2004). As shown in Fig. 6A, this sTPS stimulation reliably elicited
LTP of the slope of fEPSPs in the CA1. At recurrent seizure time Prior studies used diverse FS animal models to identify epilepto-
periods, the amplitudes and the slope of evoked fEPSP follow- genic factors, because FS is a potential etiologic factor in human TLE
ing LTP were significantly reduced as compared to control groups (Shinnar and Glauser, 2002; Dubé et al., 2006). These animal models
(126.9 ± 2.5% in FS 12 weeks; 169.7 ± 3.9% in control 12 weeks; have indicated that hippocampal cell death or damage could pre-
P < 0.01, Fig. 6A–C). In order to investigate the evoked poten- lude epileptic seizures and mesial temporal lobe epilepsy (Cendes
tials by paired-pulse stimulation from the dentate gyrus following et al., 1993; Kuks et al., 1993; VanLandingham et al., 1998; Sarkisian
hyperthermic seizure, paired-pulse responses were recorded using et al., 1999; Tanabe et al., 2001). However, the models were limited
the perforant path stimulation. At recurrent seizure time period by un-matched age with human infants of FS patients (Dobbing
(12week after FS), the PS2/PS1 ratio (the paired-pulse ratio) at and Sands, 1973), absence of neuronal death or morphological
the 30-ms inter-stimulus interval (GABAA receptor-mediated fast changes, and lower induction rate of FS (Baram et al., 1997; Bender
inhibition) was significantly reduced (0.62 ± 0.03) as compared et al., 2003; Scantlebury et al., 2005; Dubé et al., 2006). Kwak et al.
to control (1.18 ± 0.06 in control; P < 0.01, Fig. 6D and E), simi- (2008) developed a rat model to study the long-term effects of FS
lar to other studies (Tsai and Leung, 2006; Kwak et al., 2008), and epileptogenesis. Advantages of these models include the high
whereas 70-ms and 150-ms of inter-stimulus interval indicate a morbidity to recurrent seizure using single hyperthermia (∼68%
facilitation and an action of GABAB receptor-mediated slow inhi- of animals), morphological changes, lower mortality (∼2.9% of ani-
bition. In addition, the excitability ratio at this time period was mals), and the age matched with human infants. Therefore, we used
elevated to 191 ± 24.4% (100 ± 8.4% in control 12 weeks; P < 0.01, the rat hyperthermic seizure model to investigate the long-term
Fig. 6E). Moreover, the PS amplitude and the PS latency in the effect of FS and the functional and/or morphological alterations of
first response, as well as the ratio of the second with the first hippocampal interneurons involved epileptogenesis.
32 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

Fig. 4. Double staining for CR and VGLUT1 in the hippocampus at 8 weeks after FS (A1-D4). DAPI counterstaining (blue, A1-A4); CR (green, B1-B4); VGLUT-1 (red, C1-C4);
Merged-images (yellow, D1-D4). CR immunoreactivity is detected in some interneurons of hippocampus following FS. Both CR and VGLUT1 are colocalized in the same
interneurons of the CA1 and DG at 8 weeks after FS than control groups (arrow). Bar = 17 ␮m (panels A1-D4). The results of average cell number for CR and VGLUT double
labeled interneurons are shows as similar with immunofluorescence staining (E and F). All data are presented as mean ± SEM. Significant differences from the control,
*p < 0.05, **p < 0.01, vs. control. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

4.1. LFP alterations following FS tory neuronal transmission and inhibitory influences of GABAergic
interneurons at recurrent seizure time period following FS.
During recurrent seizure time period, LFP signals generally
showed an epileptic wave as a fast spike and high amplitude multi- 4.2. Morphological changes of synaptic connection following FS
spikes in the hippocampus; these signals were more noticeable in
the CA1 than in the DG region. The LFP signal in CA1 region appears In order to investigate the abnormal synaptic connections after
as an output signal of the local circuit, whereas these signals are an axonal reorganization, Timm’s staining was used from immature
input signal in the DG region incoming to the hippocampus (Takano (developmental stage) to mature (adult stage) brain following
and Coulter, 2012). Thus, the present results indicate that the CA1 hyperthermic seizure. The staining method labels zinc-rich ter-
region is more vulnerable area for the epileptogenesis and the minals of mossy fiber and can identify axon reorganization
seizure activity following FS, because of declined control abilities (Buckmaster et al., 2002). Activity-dependent anatomical reorgani-
to hyperexcitatory signal in the hippocampus. However, additional zation of mossy fiber is an important mechanism of development
experiments are necessary to verify the relationship between the and adult neural plasticity for recurrent connection of dentate
influences of input and output signals for epileptogenesis in this gyrus, principle neurons and interneurons (Cline, 2001; Wong and
animal model. On the other hand, delta rhythm synchronous neu- Ghosh, 2002; Ge et al., 2006). Moreover, axonal reorganization is
ronal activity in the hippocampus offers a unique LFP signal marker important in the excitatory neuronal circuit in epileptic hippocam-
in epilepsy (Di Gennaro et al., 2003; Gelisse et al., 2011). In addi- pus; a novel recurrent excitatory network involving glutamatergic
tion, gamma rhythm associated with hyper brain activities have neurotransmission has been demonstrated in patients and in ani-
been linked to the influences of GABAergic interneurons (White mal models of TLE (Victor Nadler et al., 1980; Tauck and Nadler,
et al., 2000; Wendling et al., 2002; Hughes, 2008). According to the 1985; Buckmaster et al., 2002). These facts may contribute to
properties of these brain waves, the present findings demonstrate epileptogenesis that occurs through excitation by abnormal synap-
that the delta rhythm during the recurrent seizure period was sig- tic connection in FS (Kwak et al., 2008). Similar to these previous
nificantly increased in the hippocampus, whereas gamma rhythm reports, the present data demonstrate axonal reorganizations in
was reduced in CA1 and DG region. Therefore, our findings suggest the DG region during development (2 weeks after FS). In addition,
that these changes of LFP signal result in the enhanced epileptiform the dense mossy fiber sprouting in the granule cell layer including
discharges accompanied by abnormalities between hyperexcita- inner molecular layer of DG was detected at the adult time period
Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38 33

Fig. 5. GABAA -␣1 immunoreactivities and quantitative analyses in CA1 and DG of the control and the epileptic hippocampus following FS (A1-G). At 4 weeks after FS,
GABAA -␣1 immunoreactivity is markedly increased in the hippocampal interneurons in both regions more than control (open arrow A1-B3). At 8 weeks following recurrent
seizure onset, GABAA -␣1 immunoreactivity is reduced in the hippocampal interneurons as similar to control levels (C1-D3). However, GABAA -␣1 positive interneurons are
significantly re-enhanced in the CA1 and DG region at recurrent seizure time period (12 weeks after FS) as compared to controls (open arrow E1-F3). Rectangles in panels
A1, B1, C1, D1, E1 and F1 indicate the high-magnification of panels A2-A3, B2-B3, C2-C3, D2-D3, E2-E3 and F2-F3. Moreover, rectangles in panels E2 and F2 indicate the right
low high-magnification of panels E2 and F2. Bar = 200 ␮m (panels A1, B1, C1, D1, E1 and F1), 50 ␮m (panels A2-A3, B2-B3, C2-C3, D2-D3, E2-E3 and F2-F3) and 25 ␮m (high-
magnification in panels E2 and F2). Quantitative analysis for average numbers of GABAA -␣1 positive interneurons is shown as similar with imuunohistochemical data (G).
All data are presented as mean ± SEM. Significant differences from the control, *p < 0.05, **p < 0.01, vs. control. Double immunofluorescence staining for GABAA -␣1 receptor
and GABA in the hippocampus at 8 weeks after FS (H1-H4). GABAA -␣1 receptor (green); GABA (red); Merged-images (yellow, H1-H4). GABAA -␣1 receptor immunoreactivity
is detected in some hippocampal interneurons following FS. Both GABAA -␣1 receptor and GABA are double labeled in the same interneurons of the CA1 and DG at 12 weeks
after recurrent seizure onset than control groups (arrow H3 and H4). Bar = 17 ␮m (panels H1-H4). Western blot (I) and optical density analysis (J) of GABAA -␣1 antibody are
revealed its protein levels at developmental and adult stages. Interestingly, at 4 weeks after recurrent seizure onset, GABAA -␣1 are transiently increased as compared to
control groups, while its expression is observed as similar between each groups at 8 weeks following seizure onset (I and J). According time course following FS, GABAA -␣1
protein levels also markedly re-enhanced at recurrent seizure time period (12 weeks after FS) as compared to age-matched control hippocampus (I and J). All data are
presented as mean ± SEM. Significant differences from the control, **p < 0.01, vs. control. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

(8 weeks after FS) following hyperthermic seizure. Inhibitory hilar interneurons in human temporal lobe epilepsy (TLE), including
interneurons in the DG during neuronal damage contribute to net- the CR-positive Cajal-Retzius cells (Maglóczky et al., 2000; Tóth
work instability by enhancing the relative impact in controlling et al., 2010). In order to verify this controversy, thus we primarily
granule cell excitability (Hunt et al., 2011) and alteration of mossy identified the expressional patterns of CR-positive interneuronal
fiber sprout is caused the malfunction of inhibitory interneurons population in the hippocampus following hyperthermic seizure.
by mossy cell degeneration (Sloviter et al., 2006). The present In present study, our results show that immunoreactivity of CR-
findings further suggest that reorganization of newly generated positive interneurons was transiently increased in the adult time
granule cells may result in the disruption of the gateway at the den- period (7–9 weeks following hyperthermic seizure) in the CA1
tate gyrus, which contributes to increased seizure susceptibility, in and DG regions, although these expressions were reduced similar
agreement with a prior study (Kwak et al., 2008). to control level during the recurrent seizure time period (10–12
weeks after FS). These findings are consistent with a previous
study that demonstrated the preservation of CR-containing cells in
4.3. Effects of hyperthermic seizure in functional efficacies of epilepsy (Blumcke et al., 1996) and the enhanced numbers of CR-
inhibitory interneuron positive Cajal-Retzius cells in FS (Blumcke et al., 1996; Thom et al.,
2002). Thus, these discrepancies may reflect the compensatory
Previously, some investigators reported that the preservation mechanisms including increased synaptic frequency of CR-positive
of CR-containing cells in epilepsy (Blumcke et al., 1996) and the interneuronal terminals, and the time dependent phenomenon
enhanced numbers of CR-positive Cajal-Retzius cells in FS (Blumcke induced by the developmental state of brain and the transient
et al., 1996; Thom et al., 2002). However, other investigators accelerations of CR-positive cell production. Since a previous study
demonstrated the increased vulnerability or loss of CR-positive
34 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

Fig. 6. Changes of the field excitatory post-synaptic potential (fEPSP) for synaptic plasticity in CA1 of the control and epileptic hippocampus. Representative traces for fEPSP
in the CA1 region at before and after tetanus in control and FS (A). The amplitude of evoked fEPSP (B) and the slope changes (C) following LTP at recurrent seizure time period
(12 weeks after FS) was reduced as compared to control. Arrows in panel B indicate the time point of sTPS application for LTP. The changes of paired-pulse responses for field
potentials in the DG of the control and FS animal hippocampus (D and E). Although the PS amplitude and the excitability ratio are increased at recurrent seizure time period
as compared to control levels, the PS latency and PS2/PS1 ratio are significantly declined in the DG (E). All data are presented as mean ± SEM. Significant differences from the
control, **p < 0.01, vs. control.

demonstrated, even with sensitivity and loss of CR-containing which are mostly granule cells (Hester and Danzer, 2013). Thus,
interneurons in the DG, an increased frequency of CR-positive more CR-positive cells might be seen in the epileptic mice after
interneuronal terminals in epilepsy (Maglóczky et al., 2000; Tóth pilocarpine-induced epilepsy, although their number is decreased
et al., 2010). This might mean that sprouting of the remaining CR- at the chronic phase with the decreasing tendency of neuron
containing inhibitory cells occurs as a compensatory mechanism production and weakening of the CR immunoreactivity of prin-
to offset the enhanced excitatory input on granule cell dendrites cipal cells with time (Brandt et al., 2003; Kralic et al., 2005).
(Tóth and Maglóczky, 2014). In addition, CR-positive interneurons Therefore, the present findings suggest that transient elevations
were colocalized with VGLUT-1 immunoreactivity at the adult time of the CR-positive interneuronal population may underlie one of
period (7–9 weeks following hyperthermic seizure) in the CA1 and the compensatory mechanisms according to the enhanced synap-
DG regions. Because CR-positive interneurons may control via the tic efficacies of CR-positive interneurons and the time dependent
synaptic connections with other interneuronal dendrites and the events during epileptogenic period following FS. Further studies
exodendritic synapses with each other, and may selectively inner- are needed to assess this hypothesis.
vate with several interneurons and calbindin-containing inhibitory On the other hand, although the slow inhibition is governed by
interneurons (Barinka et al., 2012; Tóth et al., 2010; Bausch, 2005; GABAB receptors, the fast inhibitory synaptic transmission in the
Gulyás et al., 1999), it seems to be modulated the inhibitory brain is modulated by the actions of GABAA receptors. The major-
interneuronal activities for controlling output of principal neurons. ity of GABAA receptors contain two ␣ subunits, two ␤ subunits,
In addition, VGLUT-1 positive neurons revealed the distinct exci- and a ␥ or ␦ subunit. These receptors are primarily responsi-
tatory neuronal population and fast-spiking interneurons in the ble for mediating phasic and tonic inhibitions (Baumann et al.,
cortical regions (Fattorini et al., 2015; Wouterlood et al., 2007; 2002; Hines et al., 2012). In the present study, immunoreactiv-
Moutsimilli et al., 2005; Berghuis et al., 2004), and fast-spiking ity of GABAA -␣1 receptor was markedly enhanced in the CA1
GABAergic interneuron is forming synapses to control the neuronal and DG regions 4 weeks after hyperthermic seizure compared to
excitability in the perisomatic inhibitions of principal cell (Berghuis control. Although these expressional patterns were recovered at
et al., 2004), which is synchronizing action potential firing to con- adult stage (7–9 weeks after FS) as similar to the control level,
trol output of principal cells (Freund 2003). Therefore, our findings GABAA -␣1 positive interneuron was significantly re-increased at
in this study suggest that temporary enhancements of excitatory the recurrent seizure time period (10–12 weeks following FS) in
neurotransmission via up-regulated VGLUT-1 expression in the both areas. GABAA -␣1 positive interneurons, at this time period,
CR-positive interneurons following hyperthermic seizure may be were revealed the colocalization with GABA, although the recep-
result in compensatory responses as the modulation of inhibitory tors containing GABAA -␣1 receptor subunit are located dendritic
interneuronal activities in the hippocampus. Moreover, granule areas (Schwarzer et al., 1997) and extrasynapse both of hip-
cells and mossy cells also contain CR in mice, especially young pocampal interneurones (Semyanov et al., 2004) and dentate gyrus
mice (Blasco-Ibáñez and Freund, 1997; Brandt et al., 2003), and granule cells (Nusser et al., 1995). These findings are consistent
the induction of status epilepticus may transiently accelerate with our paired-pulse responses data and the previous studies
the production of newly formed neurons (Kralic et al., 2005), that demonstrated the enhancements of paired-pulse inhibition at
Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38 35

30-ms inter-pulse interval as GABAA receptor-mediated fast inhi- enhancing feed-forward inhibitory pathway (Kairiss et al., 1987).
bition in the DG following hyperthermia-induced seizures (Tsai In addition to the arising of LTP by sufficient depolarization of
and Leung, 2006), and the increased GABAA receptor-mediated the membrane, this effect may due to the depression of inhibitory
inhibition in various seizure models including hippocampal kin- responses (Davies et al., 1990). A previous study also showed the
dling (Tuff et al., 1983; Leung et al., 1994) and status epilepticus block of excitatory synaptic transmission eliminated the afferent
(Kloosterman et al., 2001; Velišek and Velišková, 2002; Gorter et al., inputs to the interneurons from both feed-forward and recurrent
2002; Leung and Wu, 2003). In fact, the mRNA and protein lev- inhibition circuits, indicating that this block could cause a decrease
els of GABAA receptor subunits including ␣1- and ␣4-subunit in in the interneuronal firing and a decline of their inhibitory action
granule cells of DG were enhanced after status epilepticus (SE) in on the pyramidal neurons in generating of epileptiform activity
rats induced by kainate and electrical stimulation (Schwarzer et al., (Matthews et al., 1981; Alger and Nicoll, 1982; Arai et al., 1995;
1997; Tsunashima et al., 1997; Nishimura et al., 2005), and GABAA Feng and Durand, 2004; Liu et al., 2014; Houser, 2014). Therefore,
receptor ␣1 -subunit in the DG was also increased in the acute lease in part, the present findings suggest that prolonged reduced
and chronic epileptic hippocampus (Loup et al., 2000; Pirker et al., LTP induction after hyperthermic seizure may be accompanied by
2003; Sperk, 2007). In addition, Macdonald and their colleague declined functional efficacies of inhibitory interneuron due to the
(Macdonald and Rogawski, 2007) reported that the main therapeu- continuous changes in synaptic function after FS. However, it seems
tic action of diverse antiepileptic drugs, such as benzodiazepines to be need future experiments to investigate whether this declined
and barbiturates, is to enhance GABAA receptor currents. Although LTP induction following FS is closely involved to functional abnor-
hyperthermic seizure suppressed GABAergic transmission in hip- malities of inhibitory interneurons in synaptic transmission of the
pocampal neurons of immature rats for shifting the excitation and pyramidal neurons result in generating of epileptiform activities.
inhibition balance, the remaining interneurons can become hyper- On the other hand, our findings reveal increased PS amplitude
active to compensate for the loss of GABAergic inhibition (Rowley and the excitability ratio from DG region at the recurrent seizure
et al., 1995; Beau and Alger, 1998; Isokawa, 1998; Cossart et al., time period following hyperthermic seizure. These findings are
2001) and the newborn DG granule cells following FS are a cause of consistent with a previous study that demonstrated the enhanced
long-lasting increased GABAA receptor expression (Swijsen et al., PS amplitude in DG region after status epilepticus (SE) in an animal
2012). Thus, least in part, it is likely that the increased immunore- model (Matzen et al., 2008). In fact, the PS amplitude is propor-
activity of GABAA -␣1 receptor may be a common phenomenon in tional to the number of granule cells discharging an action potential
the recurrent seizure models, such as FS model. Furthermore, the (Andersen et al., 1971) and the fEPSP slope is believed to be related
present study also revealed the enhancement of GABAA -␣1 pos- to the total amount of excitatory transmitter release by the perfor-
itive interneurons in the hippocampus during recurrent seizure. mant path volley (Lomo, 1971). In addition, the excitability ratio
Therefore, because of the abnormal neuronal discharge by dys- was revealed the efficiency of glutamatergic synaptic transmission
function of GABAB receptor after FS, the present findings suggest in the DG as the PS amplitude versus the fEPSP slope in the first
that an increased long-term efficacy of GABAA -␣1 receptor after evoked response. Thus, the present results seem to be imply the
hyperthermic seizure may be related to a consequence of com- declined synaptic balances by the altered neuronal circuit and the
pensated functional responses of excessive increased inhibitory enhanced hyperexcitatory signal by the excessive increased gluta-
circuits in the hippocampus. Further studies are needed to confirm matergic synaptic transmission in the DG, because the PS amplitude
this hypothesis. and the excitability ratio indicate abnormalities of the excitability
and GABAergic inhibition in the hippocampus (Feng and Durand,
4.4. Alterations of activity-dependent synaptic plasticity and 2004). Furthermore, PS latency and the PS2/PS1 ratio in DG region
paired-pulse responses following hyperthermic seizure significantly down-regulated during the recurrent seizure time
period after FS. Although progressively enhanced paired-pulse
LTP is regulated dependent to activation of NMDR, metabotropic facilitation accompanied by impaired late paired-pulse inhibition
glutamate receptor (mGluR), acetylcholine (ACh) and many other has been observed in the DG following FS (Kwak et al., 2008), fast
factors (Cho et al., 2013; McCoy et al., 2008; Huang et al., 1997). paired-pulse inhibition enhanced in the DG has been described
Moreover, the LTP at glutamatergic synapses in the hippocam- in diverse animal models of epilepsy (Doherty and Dingledine,
pus may play significant in the encoding of memory that is 2001; Kobayashi and Buckmaster, 2003; Cohen et al., 2003; Leroy
activity-dependent effect of synaptic plasticity (Omrani et al., 2003; et al., 2004). Similarly, our study demonstrated the enhancement
Collingridge et al., 2004), and the postsynaptic calcium influx of GABAA receptor-mediated fast inhibition (30-ms inter-stimulus
was triggered for inducing LTP by activation of NMDA recep- interval) following FS. In previous studies, GABA-mediated inhi-
tors or voltage-dependent Ca+ channels (Malenka, 1991; Huber bition was measured by the ratio of the second population spike
et al., 1995; Chen et al., 2002; Omrani et al., 2003). Presently, the amplitude/the first population spike amplitude (Andersen et al.,
induction of NMDA receptor-mediated LTP demonstrated that glu- 1966; Tuff et al., 1983), and paired-pulse stimulation identifies the
tamatergic synaptic transmissions in the hippocampal CA1 region influences of feed-forward and feed-back inhibition in the principal
following hyperthermic seizure were significantly reduced at dur- cells. In addition, feed-forward inhibition is reflected by frequency-
ing recurrent seizures as compared to control. This suppressed LTP dependent inhibition of a single spike and feed-back inhibition is
induction agrees with previous studies demonstrating the declined showed by first spike amplitude-dependent inhibition of the sec-
induction of LTP by tetanic stimulation as consequence of network ond spike (Sloviter, 1991). Thus, our findings such as remarkable
changes from epileptogenesis in DG and amygdala of TLE patients up-regulated GABAA receptor-mediated fast inhibition in this study
and many epileptic animal models (Beck et al., 2000; Schubert may seem to be related to the abnormalities of GABAergic neu-
et al., 2005; Salmani et al., 2007; Schubert and Albrecht, 2008; rotransmission result in the persistent functional imbalances of
Salmani et al., 2011). Thus, it is likely that the reduced LTP induc- hippocampal neuronal circuits. However, it is need to investigate
tions may represent the maximal potential of excitatory synapse, whether these altered efficacies of GABAA receptor-mediated fast
because suppression of LTP in tetanic stimulation indicates that inhibition are involved to a cause of excitatory neuronal circuits to
synapses are already in maximum state (Rioult-Pedotti et al., 2000; recurrent seizure spreading or accompanied by a consequence of
Savic et al., 2003; Abegg et al., 2004). Interestingly, NMDAR is enhanced neuronal hyperexcitability as a compensatory response
an important role in regulating the GABAergic input by GABA in the DG following FS. Indeed, PS latency considered to be an
release (Maccaferri and Dingledine, 2002) and LTP may accompany assess parameters for the characterizing hyperexcitatory states in
36 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

DG region (Matzen et al., 2008), evaluating the increase of neu- Beau, F.E., Alger, B.E., 1998. Transient suppression of GABAA-receptor-mediated
ral excitability and the suppression of synaptic transmission (Feng IPSPs after epileptiform burst discharges in CA1 pyramidal cells. J.
Neurophysiol. 79, 659–669.
and Durand, 2004). The reduced PS latency after SE has been related Beck, H., Goussakov, I.V., Lie, A., Helmstaedter, C., Elger, C.E., 2000. Synaptic
to DG excitability changes following recurrent seizures in vivo by plasticity in the human dentate gyrus. J. Neurosci. 20, 7080–7086.
a higher release probability for glutamate on a presynaptic level Bender, R.A., Dubé, C., Gonzalez-Vega, R., Mina, E.W., Baram, T.Z., 2003. Mossy fiber
plasticity and enhanced hippocampal excitability, without hippocampal cell
(Scimemi et al., 2006) and by the enhanced expression and altered loss or altered neurogenesis, in an animal model of prolonged febrile seizures.
function of glutamate receptor (Blumcke et al., 1996; Mathern et al., Hippocampus 13, 399–412.
1998) or changes in sodium or calcium currents (Gorter et al., 2002; Berghuis, P., Dobszay, M.B., Sousa, K.M., Schulte, G., Mager, P.P., Härtig, W., Görcs,
T.J., Zilberter, Y., Ernfors, P., Harkany, T., 2004. Brain-derived neurotrophic
Ellerkmann et al., 2003). The prior and present findings suggest that
factor controls functional differentiation and microcircuit formation of
a long-term reduced PS latency at recurrent seizure time period selectively isolated fast-spiking GABAergic interneurons. Eur. J. Neurosci. 20,
may be a result of excessive up-regulated excitatory neurotrans- 1290–1306.
Blancher, C., Jones, A., 2001. SDS-PAGE and western blotting techniques. Methods
mission in the DG following hyperthermic seizure, and thus may
Mol. Med. 57, 145–162.
form the basis of epileptogenic changes that facilitate the genera- Blasco-Ibáñez, J.M., Freund, T.F., 1997. Distribution, ultrastructure, and
tion of spontaneous seizures in the FS hippocampus. connectivity of calretinin-immunoreactive mossy cells of the mouse dentate
In conclusion, abnormalities between glutamate-mediated exci- gyrus. Hippocampus 7, 307–320.
Blumcke, I., Beck, H., Nitsch, R., Eickhoff, C., Scheffler, B., Celio, M.R., Schramm, J.,
tation and GABA-mediated inhibition lead to hyperexcitability with Elger, C.E., Wolf, H.K., Wiestler, O.D., 1996. Preservation of
epileptogenesis and emergence of seizures, which can generate calretinin-immunoreactive neurons in the hippocampus of epilepsy patients
a reduction in inhibition coincident with subsequent formation with Ammon’s horn sclerosis. J. Neuropathol. Exp. Neurol. 55, 329–341.
Bouilleret, V., Loup, F., Kiener, T., Marescaux, C., Fritschy, J.M., 2000. Early loss of
of recurrent excitatory networks and in part axonal sprouting of interneurons and delayed subunit-specific changes in GABA(A)-receptor
excitatory principal cells (Bausch, 2005). Therefore, the present expression in a mouse model of mesial temporal lobe epilepsy. Hippocampus
findings reveal that time-dependent altered GABAA -␣1 and CR 10, 305–324.
Bradford, M.M., 1976. A rapid and sensitive method for the quantitation of
immunoreactivities, and chronological changes in paired-pulse microgram quantities of protein utilizing the principle ofprotein-dye binding.
responses in the FS hippocampus may be involved to the per- Anal. Biochem. 72, 248–254.
sistent functional alterations of hippocampal neuronal circuits as Brandt, M.D., Jessberger, S., Steiner, B., Kronenberg, G., Reuter, K., Bick-Sander, A.,
Von Der Behrens, W., Kempermann, G., 2003. Transient calretinin expression
an imbalance of GABAergic and glutamatergic neurotransmission
defines early postmitotic step of neuronal differentiation in adult hippocampal
by declined response abilities of excitatory and inhibitory actions. neurogenesis of mice. Mol. Cell. Neurosci. 24, 603–613.
These changes may lead to the epileptogenesis and the spreading Brooks-Kayal, A.R., Shumate, M.D., Jin, H., Rikhter, T.Y., Coulter, D.A., 1998.
Selective changes in single cell GABA(A) receptor subunit expression and
of seizure activity following hyperthermic seizure.
function in temporal lobe epilepsy. Nat. Med. 4, 1166–1172.
Buckmaster, P.S., Zhang, G.F., Yamawaki, R., 2002. Axon sprouting in a model of
temporal lobe epilepsy creates a predominantly excitatory feedback circuit. J.
Conflict of interest
Neurosci. 22, 6650–6658.
Cendes, F., Andermann, F., Dubeau, F., Gloor, P., Evans a Jones-Gotman, M., Olivier a
The authors declare that they have no competing financial inter- Andermann, E., Robitaille, Y., Lopes-Cendes, I., 1993. Early childhood prolonged
est. febrile convulsions, atrophy and sclerosis of mesial structures, and temporal
lobe epilepsy: an MRI volumetric study. Neurology 43, 1083–1087.
Chen, C., Magee, J.C., Bazan, N.G., 2002. Cyclooxygenase-2 regulates prostaglandin
E2 signaling in hippocampal long-term synaptic plasticity. J. Neurophysiol. 87,
Acknowledgement
2851–2857.
Cho, T., Ryu, J.K., Taghibiglou, C., Ge, Y., Chan, A.W., Liu, L., Lu, J., Mclarnon, J.G.,
This work was supported by the Soonchunhyang University Wang, Y.T., 2013. Long-term potentiation promotes proliferation/survival and
Research Fund (No. 20120683) and a Basic Science Research Pro- neuronal differentiation of neural stem/progenitor cells. Public Libr. Sci. One 8
(10), e76860.
gram (NRF-2013R1A1A1076058) through the National Research Cline, H.T., 2001. Dendritic arbor development and synaptogenesis. Curr. Opin.
Foundation of Korea funded by the Ministry of Science. Neurobiol. 11, 118–126.
Cohen, A.S., Lin, D.D., Quirk, G.L., Coulter, D.A., 2003. Dentate granule cell GABAA
receptors in epileptic hippocampus: enhanced synaptic efficacy and altered
References pharmacology. Eur. J. Neurosci. 17, 1607–1616.
Collingridge, G.L., Isaac, J.T.R., Wang, Y.T., 2004. Receptor trafficking and synaptic
Abegg, M.H., Savic, N., Ehrengruber, M.U., McKinney, R.A., Gähwiler, B.H., 2004. plasticity. Nat. Rev. Neurosci. 5, 952–962.
Epileptiform activity in rat hippocampus strengthens excitatory synapses. J. Cossart, R., Dinocourt, C., Hirsch, J.C., Merchan-Perez a De Felipe, J., Ben-Ari, Y.,
Physiol. 554, 439–448. Esclapez, M., Bernard, C., 2001. Dendritic but not somatic GABAergic inhibition
Alger, B.E., Nicoll, R.A., 1982. Feed-forward dendritic inhibition in rat hippocampal is decreased in experimental epilepsy. Nat. Neurosci. 4, 52–62.
pyramidal cells studied in vitro. J. Physiol. 328, 105–123. Davies, C.H., Davies, S.N., Collingridge, G.L., 1990. Paired-pulse depression of
Andersen, P., Gillow, M., Rudjord, T., 1966. Rhythmic activity in a stimulated monosynaptic GABA-mediated inhibitory postsynaptic responses in rat
neuronal network. J. Physiol. 185, 418–428. hippocampus. J. Physiol. 424, 513–531.
Andersen, P., Bliss, T.V.P., Skrede, K.K., 1971. Unit analysis of hippocampal Di Gennaro, G., Quarato, P.P., Onorati, P., Colazza, G.B., Mari, F., Grammaldo, L.G.,
population spikes. Exp. Brain Res. 13, 208–221. Ciccarelli, O., Meldolesi, N.G., Sebastiano, F., Manfredi, M., Esposito, V., 2003.
André, V., Marescaux, C., Nehlig, A., Fritschy, J.M., 2001. Alterations of hippocampal Localizing significance of temporal intermittent rhythmic delta activity
GABAergic system contribute to development of spontaneous recurrent (TIRDA) in drug-resistant focal epilepsy. Clin. Neurophysiol. 114, 70–78.
seizures in the rat lithium-pilocarpine model of temporal lobe epilepsy. Dobbing, J., Sands, J., 1973. Quantitative growth and development of human brain.
Hippocampus 11, 452–468. Arch. Dis. Child. 48, 757–767.
Arai, A., Silberg, J., Lynch, G., 1995. Differences in the refractory properties of two Doherty, J., Dingledine, R., 2001. Reduced excitatory drive onto interneurons in the
distinct inhibitory circuitries in field CA1 of the hippocampus. Brain Res. 704, dentate gyrus after status epilepticus. J. Neurosci. 21, 2048–2057.
298–306. Dubé, C.M., Baram, T.Z., 2006. Complex febrile seizures-an experimental model in
Baram, T.Z., Gerth, A., Schultz, L., 1997. Febrile seizures: an appropriate-aged immature rodents. models of seizures and epilepsy. Chapt 26, 333–340.
model suitable for long-term studies. Dev. Brain Res. 98, 265–270. Dubé, C., Richichi, C., Bender, R.A., Chung, G., Litt, B., Baram, T.Z., 2006. Temporal
Barinka, F., Druga, R., 2010. Calretinin expression in the mammalian neocortex. A lobe epilepsy after experimental prolonged febrile seizures: prospective
review. Physiol. Res. 59, 665–677. analysis. Brain 129, 911–922.
Barinka, F., Salaj, M., Rybv, J., Krajčovičová, E., Kubová, H., Druga, R., 2012. Dube, C., Chen, K., Eghbal-Ahmadi, M., Brunson, K., Soltesz, I., Baram, T.Z., 2000.
Calretinin, parvalbumin and calbindin immunoreactive interneurons in Prolonged febrile seizures in the immature rat model enhance hippocampal
perirhinal cortex and temporal area Te3V of the rat brain: qualitative and excitability long term. Ann. Neurol. 47, 336–344.
quantitative analyses. Brain Res. 1436, 68–80. Ellerkmann, R.K., Remy, S., Chen, J., Sochivko, D., Elger, C.E., Urban, B.W., Becker, A.,
Baumann, S.W., Baur, R., Sigel, E., 2002. Forced subunit assembly in ␣1␤2␥2 Beck, H., 2003. Molecular and functional changes in voltage-dependent Na+
GABAA receptors: insight into the absolute arrangement. J. Biol. Chem. 277, channels following pilocarpine-induced status epilepticus in rat dentate
46020–46025. granule cells. Neuroscience 119, 323–333.
Bausch, S.B., 2005. Axonal sprouting of GABAergic interneurons in temporal lobe
epilepsy. Epilepsy Behav. 7, 390–400.
Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38 37

Farmer, J., Zhao, X., van Praag, H., Wodtke, K., Gage, F.H., Christie, B.R., 2004. Effects Lee, J., Kim, D., Shin, H.S., 2004. Lack of delta waves and sleep disturbances during
of voluntary exercise on synaptic plasticity and gene expression in the dentate non-rapid eye movement sleep in mice lacking alpha1G-subunit of T-type
gyrus of adult male Sprague-Dawley rats in vivo. Neuroscience 124 (1), 71–79. calcium channels. Proc Natl. Acad. Sci. U. S. A. 101 (52), 18195–18199.
Fattorini, G., Antonucci, F., Menna, E., Matteoli, M., Conti, F., 2015. Co-expression of Leroy, C., Poisbeau, P., Keller a, F., Nehlig, A., 2004. Pharmacological plasticity of
VGLUT1 and VGAT sustains glutamate and GABA co-release and is regulated by GABA(A) receptors at dentate gyrus synapses in a rat model of temporal lobe
activity in cortical neurons. J. Cell Sci. 128, 1669–1673. epilepsy. J. Physiol. 557, 473–487.
Feng, Z., Durand, D.M., 2004. Suppression of excitatory synaptic transmission can Leung, L.S., Wu, C.P., 2003. Kindling suppresses primed-burst-induced long-term
facilitate low-calcium epileptiform activity in the hippocampus in vivo. Brain potentiation in hippocampal CA1. Neuroreport 14, 211–214.
Res. 1030, 57–65. Leung, L.S., Zhao, D., Shen, B., 1994. Long-lasting effects of partial hippocampal
Fritschy, J.M., Kiener, T., Bouilleret, V., Loup, F., 1999. GABAergic neurons and kindling on hippocampal physiology and function. Hippocampus 4, 696–704.
GABAA-receptors in temporal lobe epilepsy. Neurochem. Int. 34, 435–445. Liu, Y.Q., Yu, F., Liu, W.H., He, X.H., Peng, B.W., 2014. Dysfunction of hippocampal
Fuggetta, G., Bennett, M.A., Duke, P.A., Young, A.M., 2014. Quantitative interneurons in epilepsy. Neurosci. Bull. 30, 985–998.
electroencephalography as a biomarker for proneness toward developing Lomo, T., 1971. Potentiation of monosynaptic EPSPs in the perforant path-dentate
psychosis. Schizophr. Res. 153 (1-3), 68–77. granule cell synapse. Exp. Brain Res. 12, 46–63.
Ge, S., Goh, E.L., Sailor, K.A., Kitabatake, Y., Ming, G.L., Song, H., 2006. GABA Loup, F., Wieser, H.G., Yonekawa, Y., Aguzzi, A., Fritschy, J.M., 2000. Selective
regulates synaptic integration of newly generated neurons in the adult brain. alterations in GABAA receptor subtypes in human temporal lobe epilepsy. J.
Nature 439, 589–593. Neurosci. 20, 5401–5419.
Ge, Y., Dong, Z., Bagot, R.C., Howland, J.G., Phillips, A.G., Pan, T., 2010. Hippocampal Maccaferri, G., Dingledine, R., 2002. Control of feedforward dendritic inhibition by
long-term depression is required for the consolidation of spatial memory. Proc. NMDA receptor-dependent spike timing in hippocampal interneurons. J.
Natl. Acad. Sci. U. S. A. 107, 16697–16702. Neurosci. 22, 5462–5472.
Gelisse, P., Serafini, A., Velizarova, R., Genton, P., Crespel, A., 2011. Temporal Macdonald, R.L., Rogawski, M.A., 2007. Cellular effects of antiepileptic drugs.
intermittent delta activity: a marker of juvenile absence epilepsy? Seizure 20, Epilepsy: A Comprehensive Textbook, vol. II., pp. 1433–1445.
38–41. Maglóczky, Z.S., Wittner, L., Borhegyi, Z.S., Halász, P., Vajda, J., Czirják, S., Freund,
Gorter, J.A., Borgdorff, A.J., Van Vliet, E.A., Lopes da Silva, F.H., Wadman, W.J., 2002. T.F., 2000. Changes in the distribution and connectivity of interneurons in the
Differential and long-lasting alterations of high-voltage activated calcium epileptic human dentate gyrus. Neuroscience 96, 7–25.
currents in CA1 and dentate granule neurons after status epilepticus. Eur. J. Mahmood, T., Yang, P.C., 2012. Western blot: technique, theory, and trouble
Neurosci. 16, 701–712. shooting. N. Am. J. Med. Sci. 4, 429–434.
Gottlieb, A., Keydar, I., Epstein, H.T., 1977. Rodent brain growth stages: an Malenka, R.C., 1991. The role of postsynaptic calcium in the induction of long-term
analytical review. Biol. Neonate 32, 166–176. potentiation. Mol. Neurobiol. 5, 289–295.
Graves, R.C., Oehler, K., Tingle, L.E., 2012. Febrile seizures: risks, evaluation, and Mathern, G.W., Pretorius, J.K., Leite, J.P., Kornblum, H.I., Mendoza, D., Lozada, A.,
prognosis. Am. Fam. Phys. 85, 149–153. Bertram, E.H., 1998. Hippocampal AMPA and NMDA mRNA levels and subunit
Gulyás, A.I., Megías, M., Emri, Z., Freund, T.F., 1999. Total number and ratio of immunoreactivity in human temporal lobe epilepsy patients and a rodent
excitatory and inhibitory synapses converging onto single interneurons of model of chronic mesial limbic epilepsy. Epilepsy Res. 32, 154–171.
different types in the CA1 area of the rat hippocampus. J. Neurosci. 19, Matthews, W.D., McCafferty, G.P., Setler, P.E., 1981. An electrophysiological model
10082–10097. of GABA-mediated neurotransmission. Neuropharmacology 20, 561–565.
Hester, M.S., Danzer, S.C., 2013. Accumulation of abnormal adult-generated Matthews, D.B., Devaud, L.L., Fritschy, J.M., Sieghart, W., Morrow, A.L., 1998.
hippocampal granule cells predicts seizure frequency and severity. J. Neurosci. Differential regulation of GABA(A) receptor gene expression by ethanol in the
33, 8926–8936. rat hippocampus versus cerebral cortex. J. Neurochem. 70, 1160–1166.
Hines, R.M., Davies, P.A., Moss, S.J., Maguire, J., 2012. Functional regulation of GABA Matzen, J., Buchheim, K., Landeghem, F.K.H., Van Meierkord, H., Holtkamp, M.,
A receptors in nervous system pathologies. Curr. Opin. Neurobiol. 22, 552–558. 2008. Functional and morphological changes in the dentate gyrus after
Houser, C.R., 2014. Do structural changes in GABA neurons give rise to the epileptic experimental status epilepticus. Seizure 17 (1), 76–83.
state? Adv. Exp. Med. Biol. 813, 151–160. Maytal, J., Steele, R., Eviatar, L., Novak, G., 2000. The value of early postictal EEG in
Huang, L.Q., Rowan, M.J., Anwyl, R., 1997. mGluR II agonist inhibition of LTP children with complex febrile seizures. Epilepsia 41, 219–221.
induction, and mGluR II antagonist inhibition of LTD induction, in the dentate McCoy, P., Norton, T.T., McMahon, L.L., 2008. Layer 2/3 synapses in monocular and
gyrus in vitro. Neuroreport 8, 687–693. binocular regions of tree shrew visual cortex express mAChR-dependent
Huber, K.M., Mauk, M.D., Kelly, P.T., 1995. Distinct LTP induction mechanisms: long-term depression and long-term potentiation. J. Neurophysiol. 100,
contribution of NMDA receptors and voltage-dependent calcium channels. J. 336–345.
Neurophysiol. 73, 270–279. Moutsimilli, L., Farley, S., Dumas, S., El Mestikawy, S., Giros, B., Tzavara, E.T., 2005.
Hughes, J.R., 2008. Gamma, fast, and ultrafast waves of the brain: their Selective cortical VGLUT1 increase as a marker for antidepressant activity.
relationships with epilepsy and behavior. Epilepsy Behav. 13, 25–31. Neuropharmacology 49, 890–900.
Hunt, R.F., Scheff, S.W., Smith, B.N., 2011. Synaptic reorganization of inhibitory Nishimura, T., Schwarzer, C., Gasser, E., Kato, N., Vezzani, A., Sperk, G., 2005.
hilar interneuron circuitry after traumatic brain injury in mice. J. Neurosci. 31, Altered expression of GABAA and GABAB receptor subunit mRNAs in the
6880–6890. hippocampus after kindling and electrically induced status epilepticus.
Isokawa, M., 1998. Modulation of GABA(A) receptor-mediated inhibition by Neuroscience 134, 691–704.
postsynaptic calcium in epileptic hippocampal neurons. Brain Res. 810, Nusser, Z., Roberts, J.D., Baude, A., Richards, J.G., Sieghart, W., Somogyi, P., 1995.
241–250. Immunocytochemical localization of the alpha 1 and beta 2/3 subunits of the
Kairiss, E.W., Abraham, W.C., Bilkey, D.K., Goddard, G.V., 1987. Field potential GABAA receptor in relation to specific GABAergic synapses in the dentate
evidence for long-term potentiation of feed-forward inhibition in the rat gyrus. Eur. J. Neurosci. 7 (4), 630–646.
dentate gyrus. Brain Res. 401, 87–94. Omrani, A., Fathollahi, Y., Almasi, M., Semnanian, S., Mohammad, S., Firoozabadi, P.,
Kim, D.S., Kim, J.E., Kwak, S.E., Kim, D.W., Choi, S.Y., Kwon, O.S., Kang, T.C., 2007. 2003. Contribution of ionotropic glutamate receptors and voltage-dependent
Seizure activity selectively reduces 5-HT1A receptor immunoreactivity in CA1 calcium channels to the potentiation phenomenon induced by transient
interneurons in the hippocampus of seizure-prone gerbils. Brain Res. 1154, pentylenetetrazol in the CA1 region of rat hippocampal slices. Brain Res. 959,
181–193. 173–181.
Kim, D.S., Kim, J.E., Kwak, S.E., Choi, K.C., Kim, D.W., Kwon, O.S., Choi, S.Y., Kang, Parent, J.M., Yu, T.W., Leibowitz, R.T., Geschwind, D.H., Sloviter, R.S., Lowenstein,
T.C., 2008. Spatiotemporal characteristics of astroglial death in the rat D.H., 1997. Dentate granule cell neurogenesis is increased by seizures and
hippocampo-entorhinal complex following pilocarpine-induced status contributes to aberrant network reorganization in the adult rat hippocampus.
epilepticus. J. Comp. Neurol. 511, 581–598. J. Neurosci. 17, 3727–3738.
Kloosterman, F., Peloquin, P., Leung, L.S., 2001. Apical and basal orthodromic Pirker, S., Schwarzer, C., Czech, T., Baumgartner, C., Pockberger, H., Maier, H.,
population spikes in hippocampal CA1 in vivo show different origins and Hauer, B., Sieghart, W., Furtinger, S., Sperk, G., 2003. Increased expression of
patterns of propagation. J. Neurophysiol. 86, 2435–2444. GABAA receptor beta-subunits in the hippocampus of patients with temporal
Kobayashi, M., Buckmaster, P.S., 2003. Reduced inhibition of dentate granule cells lobe epilepsy. J. Neuropathol. Exp. Neurol. 62, 820–834.
in a model of temporal lobe epilepsy. J. Neurosci. 23, 2440–2452. Racine, R., 1972. Modification of seizure activity by electrical stimulation. II. Motor
Koyama, R., Tao, K., Sasaki, T., Ichikawa, J., Miyamoto, D., Muramatsu, R., Matsuki, seizure. Electroencephalogr. Clin. Neurophysiol. 32, 281–284.
N., Ikegaya, Y., 2012. GABAergic excitation after febrile seizures induces Rioult-Pedotti, M.-S.S., Friedman, D., Donoghue, J.P., 2000. Learning-induced LTP in
ectopic granule cells and adult epilepsy. Nat. Med. 18, 1271–1278. neocortex. Science 290, 533–536.
Kralic, J.E., Ledergerber, D.A., Fritschy, J.M., 2005. Disruption of the neurogenic Rowley, H.L., Martin, K.F., Marsden, C.A., 1995. Decreased GABA release following
potential of the dentate gyrus in a mouse model of temporal lobe epilepsy with tonic-clonic seizures is associated with an increase in extracellular glutamate
focal seizures. Eur. J. Neurosci. 22, 1916–1927. in rat hippocampus in vivo. Neuroscience 68, 415–422.
Kuks, J.B., Cook, M.J., Fish, D.R., Stevens, J.M., Shorvon, S.D., 1993. Hippocampal Salmani, M.E., Mirnajafizadeh, J., Fathollahi, Y., 2007. Offsetting of aberrations
sclerosis in epilepsy and childhood febrile seizures. Lancet 342, 1391–1394. associated with seizure proneness in rat hippocampus area CA1 by theta pulse
Kwak, S.E., Kim, J.E., Kim, D.S., Jung, J.Y., Moo, H.W., Kwon, O.S., Choi, S.Y., Kang, T.C., stimulation-induced activity pattern. Neuroscience 149, 518–526.
2005. Effects of GABAergic transmissions on the immunoreactivities of calcium Salmani, M.E., Fathollahi, Y., Mirnajafizadeh, J., Semnanian, S., 2011. Epileptogenic
binding proteins in the gerbil hippocampus. J. Comp. Neurol. 485, 153–164. insult alters endogenous adenosine control on long-term changes in synaptic
Kwak, S.E., Kim, J.E., Kim, S.C., Kwon, O.S., Choi, S.Y., Kang, T.C., 2008. Hyperthermic strength by theta pattern stimulation in hippocampus area CA1. Synapse 65,
seizure induces persistent alteration in excitability of the dentate gyrus in 189–197.
immature rats. Brain Res. 1216, 1–15.
38 Y.H. Yu et al. / Brain Research Bulletin 131 (2017) 25–38

Sarkisian, M.R., Holmes, G.L., Carmant, L., Liu, Z., Yang, Y., Stafstrom, C.E., 1999. Tóth, K., Eross, L., Vajda, J., Halász, P., Freund, T.F., Maglóczky, Z., 2010. Loss and
Effects of hyperthermia and continuous hippocampal stimulation on the reorganization of calretinin-containing interneurons in the epileptic human
immature and adult brain. Brain Dev. 21, 318–325. hippocampus. Brain 133, 2763–2777.
Savic, N., Luthi, A., Gahwiler, B.H., McKinney, R.A., 2003. N-methyl-D-aspartate Takano, H., Coulter, D.A., 2012. Imaging of hippocampal circuits in epilepsy.
receptor blockade during development lowers long-term potentiation Epilepsia 51, 20.
threshold without affecting dynamic range of CA3-CA1 synapses. Proc. Natl. Tanabe, T., Suzuki, S., Hara, K., Shimakawa, S., Wakamiya, E., Tamai, H., 2001.
Acad. Sci. U. S. A. 100, 5503–5508. Cerebrospinal fluid and serum neuron-specific enolase levels after febrile
Scantlebury, M.H., Gibbs, S.A., Foadjo, B., Lema, P., Psarropoulou, C., Carmant, L., seizures. Epilepsia 42, 504–507.
2005. Febrile seizures in the predisposed brain: a new model of temporal lobe Tauck, D.L., Nadler, J.V., 1985. Evidence of functional mossy fiber sprouting in
epilepsy. Ann. Neurol. 58, 41–49. hippocampal formation of kainic acid-treated rats. J. Neurosci. 5, 1016–1022.
Scheffer-Teixeira, R., Belchior, H., Leão, R.N., Ribeiro, S., Tort, A.B., 2013. On Thom, M., Sisodiya, S.M., Beckett, A., Martinian, L., Lin, W.-R., Harkness, W.,
high-frequency field oscillations (>100 Hz) and the spectral leakage of spiking Mitchell, T.N., Craig, J., Duncan, J., Scaravilli, F., 2002. Cytoarchitectural
activity. J. Neurosci. 33 (4), 1535–1539. abnormalities in hippocampal sclerosis. J. Neuropathol. Exp. Neurol. 61,
Schubert, M., Albrecht, D., 2008. Activation of kainate GLU(K5) transmission 510–519.
rescues kindling-induced impairment of LTP in the rat lateral amygdala. Toth, Z., Yan, X.X., Haftoglou, S., Ribak, C.E., Baram, T.Z., 1998. Seizure-Induced
Neuropsychopharmacology 33, 2524–2535. neuronal injury: vulnerability to febrile seizures in an immature rat model. J.
Schubert, M., Siegmund, H., Pape, H.-C., Albrecht, D., 2005. Kindling-induced Neurosci. 18, 4285–4294.
changes in plasticity of the rat amygdala and hippocampus. Learn. Mem. 12, Tsai, M.L., Leung, L.S., 2006. Decrease of hippocampal GABAB receptor-mediated
520–526. inhibition after hyperthermia-induced seizures in immature rats. Epilepsia 47,
Schutter, D.J., van Honk, J., d’Alfonso, A.A., Peper, J.S., Panksepp, J., 2003. High 277–287.
frequency repetitive transcranial magnetic over the medial cerebellum Tsunashima, K., Schwarzer, C., Kirchmair, E., Sieghart, W., Sperk, G., 1997. GABA(A)
induces a shift in the prefrontal electroencephalography gamma spectrum: a receptor subunits in the rat hippocampus III: Altered messenger rna
pilot study in humans. Neurosci. Lett. 336 (2), 73–76. expression in kainic acid-induced epilepsy. Neuroscience 80, 1019–1032.
Schwarzer, C., Tsunashima, K., Wanzenböck, C., Fuchs, K., Sieghart, W., Sperk, G., Tuff, L.P., Racine, R.J., Mishra, R.K., 1983. The effects of kindling on GABA-mediated
1997. GABA(A) receptor subunits in the rat hippocampus II: altered inhibition in the dentate gyrus of the rat. II. Receptor binding. Brain Res. 277,
distribution in kainic acid-induced temporal lobe epilepsy. Neuroscience 80, 91–98.
1001–1017. Van Vliet, E.A., Aronica, E., Tolner, E.A., Lopes Da Silva, F.H., Gorter, J.A., 2004.
Scimemi, A., Schorge, S., Kullmann, D.M., Walker, M.C., 2006. Epileptogenesis is Progression of temporal lobe epilepsy in the rat is associated with
associated with enhanced glutamatergic transmission in the perforant path. J. immunocytochemical changes in inhibitory interneurons in specific regions of
Neurophysiol. 95, 1213–1220. the hippocampal formation. Exp. Neurol. 187, 367–379.
Scott, R.C., Holmes, G.L., 2012. Febrile seizures and the wandering granule cell. Nat. VanLandingham, K.E., Heinz, E.R., Cavazos, J.E., Lewis, D.V., 1998. Magnetic
Med. 18, 1180–1182. resonance imaging evidence of hippocampal injury after prolonged focal
Semyanov, A., Walker, M.C., Kullmann, D.M., Silver, R.A., 2004. Tonically active febrile convulsions. Ann. Neurol. 43, 413–426.
GABAA receptors: modulating gain and maintaining the tone. Trends Neurosci. Velišek, L., Velišková, J., 2002. Estrogen treatment protects GABAb inhibition in the
27, 262–269. dentate gyrus of female rats after kainic acid-induced status epilepticus.
Shinnar, S., Glauser, T.A., 2002. Febrile seizures. J. Child Neurol. 17, S44–52. Epilepsia 43 (Suppl. 5), 146–151.
Sieghart, W., Fuchs, K., Tretter, V., Ebert, V., Jechlinger, M., Höger, H., 1999. Victor Nadler, J., Perry, B.W., Cotman, C.W., 1980. Selective reinnervation of
Structure and subunit composition of GABA(A) receptors. Neurochem. Int. 34, hippocampal area CA1 and the fascia dentata after destruction of CA3-CA4
379–385. afferents with kainic acid. Brain Res. 182, 1–9.
Slézia, A., Kékesi, A.K., Szikra, T., Papp, A.M., Nagy, K., Szente, M., Maglóczky, Z., Virta, M., Hurme, M., Helminen, M., 2002. Increased frequency of
Freund, T.F., Juhász, G., 2004. Uridine release during aminopyridine-induced interleukin-1((-511) allele 2 in febrile seizures. Pediatr. Neurol. 26, 192–195.
epilepsy. Neurobiol. Dis. 16, 490–499. Wendling, F., Bartolomei, F., Bellanger, J.J., Chauvel, P., 2002. Epileptic fast activity
Sloviter, R.S., Sollas a, L., Barbaro, N.M., Laxer, K.D., 1991. Calcium-binding protein can be explained by a model of impaired GABAergic dendritic inhibition. Eur. J.
(calbindin-D28K) and parvalbumin immunocytochemistry in the normal and Neurosci. 15, 1499–1508.
epileptic human hippocampus. J. Comp. Neurol. 308, 381–396. White, J., a Banks, M.I., Pearce, R., a Kopell, N.J., 2000. Networks of interneurons
Sloviter, R.S., Zappone, C.A., Harvey, B.D., Frotscher, M., 2006. Kainic acid-induced with fast and slow gamma-aminobutyric acid type A (GABAA) kinetics provide
recurrent mossy fiber innervation of dentate gyrus inhibitory interneurons: substrate for mixed gamma-theta rhythm. Proc. Natl. Acad. Sci. U. S. A. 97,
possible anatomical substrate of granule cell hyperinhibition in chronically 8128–8133.
epileptic rats. J. Comp. Neurol. 494, 944–960. Wong, R.O.L., Ghosh, A., 2002. Activity-dependent regulation of dendritic growth
Sloviter, R.S., 1991. Feedforward and feedback inhibition of hippocampal principal and patterning. Nat. Rev. Neurosci. 3, 803–812.
cell activity evoked by per-forant path stimulation: GABA-mediated Wong, T.P., Howland, J.G., Robillard, J.M., Ge, Y., Yu, W., Titterness, A.K., Brebner, K.,
mechanisms that regulate excitability in vivo. Hippocampus 1, 31–40. Liu, L., Weinberg, J., Christie, B.R., Phillips, A.G., Wang, Y.T., 2007. Hippocampal
Sperk, G., 2007. Changes in GABAA receptors in status epilepticus. Epilepsia 48, long-term depression mediates acute stress-induced spatial memory retrieval
11–13. impairment. Proc. Natl. Acad. Sci. U. S. A. 104, 11471–11476.
Swijsen, A., Brône, B., Rigo, J.M., Hoogland, G., 2012. Long-lasting enhancement of Wouterlood, F.G., Canto, C.B., Aliane, V., Boekel, A.J., Grosche, J., Härtig, W., Beliën,
GABAA receptor expression in newborn dentate granule cells after early-life J.A.M., Witter, M.P., 2007. Coexpression of vesicular glutamate transporters 1
febrile seizures. Dev. Neurobiol. 72, 1516–1527. and 2: glutamic acid decarboxylase and calretinin in rat entorhinal cortex.
Tóth, K., Maglóczky, Z., 2014. The vulnerability of calretinin-containing Brain Struct. Funct. 212, 303–319.
hippocampal interneurons to temporal lobe epilepsy. Front. Neuroanat. 8,
1–12.

You might also like