PreprintPDF Available

Abstract and Figures

Whole brain ionic and metabolic imaging has potential as a powerful tool for the characterization of brain diseases. In this study we combined sodium MRI ( ²³ Na MRI) and ¹ H-MR Spectroscopic Imaging ( ¹ H-MRSI) and compared ionic/metabolic changes probed by this multimodal approach to intracerebral stereotactic-EEG (SEEG) recordings. We applied a multi-echo density adapted 3D projection reconstruction pulse sequence at 7T ( ²³ Na MRI) and a 3D echo planar spectroscopic imaging sequence at 3T ( ¹ H-MRSI) in 19 patients suffering from drug-resistant focal epilepsy who underwent presurgical SEEG. We investigated ²³ Na MRI parameters including total sodium concentration (TSC) and the sodium signal fraction associated of with the short component of T 2 * decay ( f ), alongside the level of metabolites N-acetyl aspartate (NAA), choline compounds (Cho) and total creatine (tCr). All measures were extracted from spherical regions of interest (ROIs) centered between two adjacent SEEG electrode contacts and z-scored against the same ROI in controls. Group comparison showed a significant increase in f only in the epileptogenic zone (EZ) compared to controls and compared to patients propagation zone (PZ) and non-involved zone (NIZ). TSC was significantly increased in all patients’ regions compared to controls. Conversely, NAA levels were significantly lower in patients compared to controls, and lower in the EZ compared to PZ and NIZ. Multiple regression analyzing the relationship between sodium and metabolites levels revealed significant relations in PZ and in NIZ but not in EZ. Our results are in agreement with the energetic failure hypothesis in epileptic regions associated with widespread tissue reorganization.
Content may be subject to copyright.
Combining Sodium MRI, Proton MR Spectroscopic
Imaging and Intracerebral EEG in Epilepsy
Mikhael Azilinon1,2,3, Julia Scholly3,4, Wafaa Zaaraoui1,3, Samuel Medina Villalon2,4 , Patrick
Viout1,3, Tangi Roussel1,3, Mohamed Mounir El Mendili1,3, Ben Ridley5, Jean-Philippe
Ranjeva1,3, Fabrice Bartolomei2,4 ,Viktor Jirsa2and Maxime Guye1,3
Abstract
Whole brain ionic and metabolic imaging has potential as a powerful tool for the
characterization of brain diseases. In this study we combined sodium MRI (23Na MRI) and
1H-MR Spectroscopic Imaging (1H-MRSI) and compared ionic/metabolic changes probed by
this multimodal approach to intracerebral stereotactic-EEG (SEEG) recordings.
We applied a multi-echo density adapted 3D projection reconstruction pulse sequence at 7T
(23Na MRI) and a 3D echo planar spectroscopic imaging sequence at 3T (1H-MRSI) in 19
patients suffering from drug-resistant focal epilepsy who underwent presurgical SEEG. We
investigated 23Na MRI parameters including total sodium concentration (TSC) and the
sodium signal fraction associated of with the short component of T2* decay (f), alongside the
level of metabolites N-acetyl aspartate (NAA), choline compounds (Cho) and total creatine
(tCr). All measures were extracted from spherical regions of interest (ROIs) centered between
two adjacent SEEG electrode contacts and z-scored against the same ROI in controls.
Group comparison showed a significant increase in fonly in the epileptogenic zone (EZ)
compared to controls and compared to patients propagation zone (PZ) and non-involved zone
(NIZ). TSC was significantly increased in all patients’ regions compared to controls.
Conversely, NAA levels were significantly lower in patients compared to controls, and lower
in the EZ compared to PZ and NIZ. Multiple regression analyzing the relationship between
sodium and metabolites levels revealed significant relations in PZ and in NIZ but not in EZ.
Our results are in agreement with the energetic failure hypothesis in epileptic regions
associated with widespread tissue reorganization.
Author affiliations:
1 Aix Marseille Univ, CNRS, CRMBM, Marseille, France.
2 Aix Marseille Univ, INSERM, INS, Inst Neurosci Syst, Marseille, France.
3 APHM, Timone Hospital, CEMEREM, Marseille, France.
4 APHM, Timone Hospital, Epileptology Department, Marseille, France.
5 IRCCS Istituto delle Scienze Neurologiche di Bologna, Bologna, Italy
1
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
NOTE: This preprint reports new research that has not been certified by peer review and should not be used to guide clinical practice.
Correspondence to: Prof Maxime Guye, MD, PhD,
Centre de Résonance Magnétique Biologique et Médicale
Faculté des Sciences Médicales et paramédicales
27, Bd Jean Moulin, 13385 Marseille Cedex 5, France
Email: maxime.guye@univ-amu.fr
Running title: Sodium MRI and spectroscopy in focal epilepsy
Keywords : Sodium MRI ; 1H-Spectroscopic Imaging ; Epilepsy ; Intracranial EEG ;
Multimodal Imaging ; 7T MRI
Abbreviations:
AC-PC = Anterior Commissure - Posterior Commissure; ASL = Achieved Significance
Level; Cho = Choline Compounds; CRLB = Cramer-Rao Lower Bound; DNET =
Dysembryoplastic Neuroepithelial Tumor; EAAT = Glutamate Transporter; ECV =
Extracellular Volume; EI = Epileptogenicity Index; EZ = Epileptogenic Zones; f = fraction of
sodium signal with short T2* decays; FCD = Focal Cortical Dysplasia ; FDR = False
Discovery Rate; 18F-FDG-PET = Fluorodeoxyglucose PET; FOV = Field of View; GRAPPA
= GeneRalized Autocalibrating Partial Parallel Acquisition; 1H = Hydrogen; HC = Healthy
Controls; MIDAS = Metabolite Imaging and Data Analyses System; MNI = Montreal
Neurological Institute; MP2RAGE = Magnetization Prepared 2 Rapid Acquisition Gradient
Echoes; MPRAGE = Magnetization Prepared Rapid Acquisition Gradient Echoes; EPSI =
Echo Planar Spectroscopic Imaging; 23Na = Sodium; NAA = N-Acetyl Aspartate; NaLF =
apparent concentrations of sodium with long T2* decays; NaSF = apparent concentrations of
sodium with short apparent T2* decay; NBC =Bicarbonate Sodium Cotransporter; NCX =
Sodium Calcium Exchanger; NHE = Sodium Hydrogen Antiporter; NIZ = Non Involved
Zones; Na+/K+ pump = Sodium Potassium pump; NKCC1 = Sodium Potassium Chloride
cotransporter; pH = Potential of Hydrogen; PZ = Propagation Zones; QED = Quality
Electrodynamics; ROI = Region of Interest; SEEG = Stereotactic EEG; T2*long = Sodium
longT2* relaxation time; T2*short = Sodium short T2* relaxation time; tCr = Total Creatine; TE
= Echo Time; TI = Inversion Time; TR = Repetition Time; TSC = Total Sodium
Concentration; UTE = Ultrashort Echo Time; VGNC = Voltage Gated Sodium Channels
2
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
1. Introduction
Recent advances in neuroimaging have challenged the concept of focal epilepsy as a brain
disorder strictly limited to the regions responsible for seizure generation and propagation
(Larivière et al., 2021). Various MRI modalities (i.e. structural, functional, metabolic) have
consistently demonstrated structural and functional alterations that extend beyond the
epileptogenic regions and affect areas non-involved in ictal discharges or interictal
epileptiform activity. These complex changes affect both structure and function at different
spatial and temporal scales and can reflect seizure-induced alterations, neuronal plasticity as
well as the underlying etiology. These factors, and the precise anatomical location of
generators of epileptiform activity, are subject to variation even between individuals with the
same epileptic syndrome. Thus, a systematic comparison of imaging data with the gold
standard of electrophysiological data derived from intracerebral
stereoelectroencephalography (SEEG) recording is essential to test the potential contribution
of any imaging modality towards better definition of the epileptogenic zone (EZ) in patients
suffering from drug-resistant focal epilepsy who are candidates for surgery. Moreover, the
need to characterize the variable manifestations of pathology suggest a clear need to exploit
multimodal imaging, combining metrics from different modalities. Insight into alterations of
ionic homeostasis and metabolic function can be probed by sodium (23Na) MRI and 1H-MR
spectroscopic imaging (MRSI), respectively.
23Na-MRI provides a unique opportunity to non-invasively image sodium signal in the brain
(Madelin et al., 2014). To date, the only 23Na-MRI study performed in a group of human focal
epilepsy, has demonstrated an increase of the total sodium concentration (TSC) in patients’
brain compared to controls, which was greater in the epileptogenic zone (EZ) compared to
the propagation zone (PZ) and the non-involved regions (NIZ) (Ridley et al., 2017). The
3
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
usefulness of TSC as a potential epileptogenicity marker may remain limited due to its
limited specificity, as it likely reflects different underlying phenomena at the cellular level,
such as changes in intracellular sodium concentration, changes in extracellular volume and
cells density and/or organization among others.
For 23Na-MRI the increase in signal to noise available at 7T allows novel approach based on a
3D-multi-echoes density-adapted radial sequence which exploits the biexponential T2* decay
of the 23Na MR signal (Ridley et al., 2018). This approach permits a multiparametric
investigation of variation in T2* decay behavior related to the quadrupolar interactions of the
3/2 spin of 23Na with the electric field gradient of surrounding molecules (Rooney &
Springer, 1991), as an indicator of tissue organization and molecular environment. The
biexponential fit model can be used as a probe to determine the motional regimes of sodium
nuclei within the surrounding environment. Thus, by quantifying the sodium signal fraction
with the short T2* decay component (f) this approach may offer a more relevant metric for
studying tissue alterations and potentially provide a better link between sodium homeostasis
and neuronal excitability in human epilepsy. In the present study, we implemented this
multiparametric approach of sodium MRI for the first time in the assessment of profiles
within and outside the epileptogenic and propagation networks in focal drug-resistant
epilepsy.
As a further objective, we assessed metabolic alterations accompanying sodium concentration
changes. To do so, we explored metabolic status by using whole-brain 1H-Echo Planar
spectroscopic imaging (1H-EPSI) in the same subjects at 3T. Three main metabolites were
quantified : (i) N-acetyl Aspartate (NAA) reflecting neuronal viability, mitochondrial
dysfunction or neuronal loss (Moffett et al., 2013; Stefano et al., 1995), (ii) Choline
4
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
compounds (Cho) reflecting membrane turnover and inflammatory processes (Achten, 1998;
Urenjak et al., 1993), and (iii) total-creatine compounds (tCr) reflecting intracellular energy
states, and energy-dependent systems in the brain (Kreis et al., 1992; Kreis & Ross, 1992),
and considered as a cellularity index (Kreis et al., 1993). Importantly, NAA has been
consistently shown to be decreased in the epileptic brain, particularly in the EZ and PZ,
compared to non-involved regions, and is thus considered a potential epileptogenicity marker
in focal epilepsy (Guye et al., 2002, 2005; Hugg et al., 1993; Kuzniecky et al., 1998;
Lundbom et al., 2001; Simister et al., 2002). Furthermore, it has been hypothesized that the
observed increase in total sodium concentration would mainly reflect energetic failures due to
mitochondrial dysfunction affecting the Na+/K+pump activity (Ridley et al., 2017; Stys et al.,
1992). Thus, measuring both sodium and NAA in the same regions provides clues with
regard to the mitochondrial defect hypothesis (Donadieu et al., 2019; Paling et al., 2011).
However, the use of this metabolite has been limited by poor resolution and spatial coverage
of routinely performed 1H-MRSI, as well as by its insufficient specificity. In this study, we
benefited from the whole brain coverage with a relatively high spatial resolution allowing
comparison between multimodal MRI and electrophysiological metrics.
Therefore, through this trimodal approach, we aimed to characterize ionic and metabolic
changes within epileptogenic networks in comparison with electrophysiologically normal
appearing brain networks. For this purpose, we analyzed differences in 23Na-MRI (TSC and f)
and 1H-MRSI (Cho, NAA and tCr) metrics between patients and controls as well as between
regions of interest (ROI) defined by quantitative SEEG signal analysis (Figure 1). We then
investigated the association between 23Na-MRI and 1H-MRSI metrics in EZ, PZ and NIZ in
order to link the homeostatic and metabolic mechanisms to the SEEG recorded electrical
alterations.
5
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
2. Materials and Methods
2.1. Subjects
Among all patients who underwent stereotactic intracerebral EEG (SEEG) recording in the
context of presurgical evaluation for drug-resistant focal epilepsy at our center between
January 2017 and February 2020, 19 consecutive patients with available 3D 1H-MRSI and
23Na-MRI were retrospectively included (Table 1). All patients had detailed non-invasive
presurgical evaluation including medical history, neurological examination,
neuropsychological assessment, 18F-FDG-PET, high-resolution structural 7T and 3T MRI,
and a long-term scalp-video-EEG. All these steps were necessary for patient enrollment. The
SEEG was indicated in all patients to localize the EZ and to precisely determine its relation
with eloquent areas. SEEG was performed as a part of the routine clinical management in line
with the French national guidelines on stereoelectroencephalography (SEEG) (Isnard et al.,
2018). SEEG implantation was planned individually for each patient, according to
anatomo-electro-clinical hypotheses about the localization of the EZ based on non-invasive
investigations. All SEEG explorations were bilateral and systematically sampled temporal,
insular, frontal, and parietal regions of at least one hemisphere. Follow-up information was
collected from a review of the medical records.
MRI data were always acquired before SEEG implantation. After quality checks (see below),
seventeen spectroscopic 1H-MRSI datasets, and fifteen 23Na-MRI datasets were used for
further analysis. Thirteen out of the nineteen patients fulfilled data quality in sufficient ROIs
for both modalities. For 1H-MRSI, we used a control database of 25 healthy controls (HC;
mean age 30.5 ± 9.7 years, range 20-60 years, 14 women). For 23Na-MRI, we used a control
database of 18 HC for (mean age 30.5 ± 8.36 years, range 21-54 years, 10 women).
6
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Participants provided informed consent in compliance with the ethical requirements of the
Declaration of Helsinki and the protocol was approved by the local Ethics Committee
(Comité de Protection des Personnes sud Méditerranée 1).
2.2. MRI Acquisitions
The protocol was conducted for all subjects on the two same MR scanners. The
Spectroscopic imaging protocol on a 3-Tesla Magnetom Verio MR system (Siemens,
Erlangen, Germany) and the 23Na MRI protocol on a whole-body 7-Tesla Magnetom Step 2
MR system (Siemens, Erlangen, Germany).
At 3T, 1H-MRI and 1H-EPSI were performed with a thirty-two channel phased-array head
coil and included a sagittal high resolution 3D-MPRAGE protocol (TE/TR/TI = 3/2300/900
ms, 160 sections, 256×256 mm², FOV 256×256 matrix, resolution = 1 mm³). Whole brain 3D
1H-EPSI was acquired as described in (Lecocq et al., 2015) using two axial acquisitions with
two different orientations that are the AC-PC plane and the AC-PC + 15° plane (TE/TR/TI =
20/1710/198 ms, nominal voxel size = 5.6 × 5.6 × 10 mm3, FOV = 280 × 280 × 180 mm3, flip
angle = 73°, 50 × 50 × 18 k-space points, GeneRalized Autocalibrating Partial Parallel
Acquisition (GRAPPA) factor = 2, acquisition time 17 minutes). The two angles of EPSI
orientations were chosen in order to obtain good quality spectra on as large brain area as
possible on at least one acquisition with a reasonable angulation to permit accurate automatic
normalization procedure.
At 7T, a high-resolution proton MRI 3D-MP2RAGE (TR = 5000 ms/TE = 3 ms/TI1 = 900
ms/TI2 = 2750 ms, 256 slices, 0.6 mm isotropic resolution, acquisition time = 10 min) was
7
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
obtained using a 32-element (32Rx/1Tx) 1H head coil (Nova Medical). 23Na-MRI was
acquired using a dual-tuned 23Na/ 1H QED birdcage coil and a multi-echo density adapted 3D
projection reconstruction pulse sequence (TR = 120 ms, 5000 spokes, 384 radial samples per
spoke, 3 mm nominal isotropic resolution, 24 echoes (8 per run with 3 runs, acquisition time
10 min per run (30 min in total)). Three dimension 23Na MRI volumes were obtained at 24
different TEs ranging from 0.2 ms to 70.78 ms (Run 1: 0.2 - 9.7 - 19.2 - 28.7 - 38.2 - 47.7 -
57.2 - 66.7 ms; Run 2: 1.56 - 11.06 - 20.56 - 30.06 - 39.56 - 49.06 - 58.56 - 68.06 ms; Run 3:
4.28 - 13.78 - 23.28 - 32.78 - 42.28 - 51.78 - 61.28 - 70.78 ms). This ensured a sufficient
number and distribution of TEs while taking into account the 5 ms readout of the sequence,
especially for measuring 23Na signal with short T2*. For quantitative calibration of brain
sodium concentrations we used as external reference six tubes (80mm length, 15mm
diameter) filled with a mixture of 2% agar gel and sodium at different concentrations: two
tubes at 25mM, one at 50mM, two at 75mM and one at 100mM. Tubes were positioned in the
field of view in front of the subject’s head and maintained using a cap.
2.3. MRI Data Processing
Three dimension 1H-EPSI images were post-processed with the Metabolite Imaging and Data
Analyses System (MIDAS, Trac, MRIR, Miami) (Maudsley et al., 2006). This software
ensures B0map correction, lipid suppression, tissue volume fraction through T1segmentation,
spectral fitting, exclusion of outlier voxels based on Cramer Rao lower bounds (CRLB) and
signal normalization with the interleaved water signal acquired. MIDAS provided AC-PC and
AC-PC 15° oriented maps, including metabolite maps (NAA, Cho, tCr), quality maps, CRLB
maps and linewidth maps among others that were used in further processing and quality
check steps. In this study we only analyzed NAA, tCr and Cho maps because m-Ino
(Myo-Inositol) and Glx (for glutamate, glutamine and glutathione) maps did not fulfill CRLB
8
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
criteria in the majority of the ROIs (Figure 2). ROI selection process is detailed in the next
paragraph. The maps that we used for the ROIs signal extraction are the average maps of
realigned AC-PC and AC-PC 15° oriented maps (Donadieu et al., 2019; Lecocq et al., 2015)
(SPM12, (Statistical Parametric Mapping: The Analysis of Functional Brain Images - 1st
Edition, n.d.)).
Metabolic profiles of each ROI (for details on ROI definition see section 2.6) were
determined by extracting water signal normalized values (arbitrary unit) from corrected
quantitative NAA, tCr and Cho maps derived from MIDAS. We applied quality assurance
criteria for spectra in each ROI based on a combination of Cramer-Rao minimum variance
bound (CRLB < 15%, or CRLB lower than half of all HC), and water peak linewidth (<16
Hz) thresholds. Supplementary Fig.1. illustrates the spectra that are in the permitted range or
not. If, after ROIs rejection in a patient, there was no ROI in EZ, the patient was entirely
discarded from further analyzes.
The 23Na-MRI data were processed according to (Grimaldi et al., 2021) using a homemade
adjusted procedure. After brain extraction on ²³Na-MRI images (ANTS, (Avants et al.,
2011)), a denoising filter was applied to the resulting 23Na-MRI volumes (Aja-Fernandez et
al., 2008; Rajan et al., 2010). The first echo time (TE) volumes from 23Na MRI were used as
reference for coregistration of the other TE volumes to correct for potential motion between
acquisitions. Hand drawn ROIs were placed in the center of each agar tube to extract signal
intensities from the twenty-four TE volumes of 23Na-MRI acquisitions for signal calibration
procedure. Linear fitting of 23Na signal decay from tubes ROI (Matlab) provided the slope (a)
and intercept (b), which is used for calibration purposes (see below).
9
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
For each ROI (for details on ROI definition see section 2.6), we fitted the mean signal
intensity across each of the 24 TEs with a biexponential model using the equation:
where Ais an amplitude scaling term, fis the sodium signal fraction of the short T2* decay
component and Ric refers to a Rician noise-related scaling parameter (Ridley et al., 2017).
From the model we estimated a sodium signal fraction for short (f) and long (1-f) T2* decay.
Note that these short and long fractions of the sodium signal are only present when the Na+
ions are present in an organized non-isotropic molecular environment. Indeed, in these
environments, the electric quadrupolar moment of the sodium nucleus interacts strongly with
the surrounding electric field gradient leading to a residual quadrupolar interaction
responsible for the bi-exponential T2relaxation with a short T2* and a long T2* depending on
the motional regime of the environment. In contrast, in an isotropic non-restricted
environment such as the cerebrospinal fluid, there is no residual quadrupolar interaction and
all energetic transitions are equal and lead to a monoexponential decay with only one (long)
T2* (Burstein & Springer, 2019). We calculated magnetization (M0) corresponding to the
signal fraction estimated by the model in terms of the intercepts of the signal fraction
components of the model, obtaining M0SF = fand M0LF = A·(1-f).Then, NaSF and NaLF
were calculated with raw M0 signal values and the linear fit estimated over the tube
phantoms, i.e. slope (a) and intercept (b):
Finally, we calculated the total sodium concentration (TSC) in each ROI as follows:
10
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Each ROI had to be almost totally (> 99.9%) included in the brain mask of the individual
patient and in the brain masks of at least half of all HC. Finally, to limit the partial volume
effect related to CSF, patient ROIs with high CSF contains - estimated with structural image
segmentation - relative to HC (|Z-scoreCSF| > 1.96) were also discarded. If after this final step
a patient had no remaining EZ ROIs, the patient was entirely discarded from the analyzes,
reducing the number of investigated patients from 22 to 19.
2.4. SEEG Recordings
Recordings were performed using intracerebral multiple contact electrodes (10–18 contacts
with length 2 mm, diameter 0.8 mm, and 1.5 mm apart, Alcis, France). The electrodes were
implanted using a stereotactic surgical robot ROSA™. Cranial CT scan was performed to
verify the absence of any complication and the spatial accuracy of the implantation. CT/MRI
data co-registration and 3D-reconstructions of patients' brains with electrodes was performed
using an in-house open-source software (EpiTools, (Medina Villalon et al., 2018)) to
automatically localize the position of each electrode contact and display the results of signal
analysis in each patient's anatomy.
Signals were recorded on a 256-channel Natus system, sampled at 512 Hz and saved on a
hard disk (16 bits/ sample) using no digital filter. Two hardware filters were present in the
acquisition procedure: a high-pass filter (cut-off frequency equal to 0.16 Hz at -3 dB), and an
anti-aliasing low-pass filter (cut-off frequency equal to 170 Hz at 512Hz).
2.5. SEEG-signal analysis
All signal analyzes were performed in a bipolar montage and computed using the
open-source AnyWave software (Colombet et al., 2015) available at
11
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
https://meg.univ-amu.fr/wiki/AnyWave.The epileptogenicity of different brain structures was
assessed by quantitative SEEG-signal analysis using the Epileptogenicity Index (EI)
(Bartolomei et al., 2008).31 The EI combines analysis of both spectral and temporal features
of SEEG signals, respectively, related to the propensity of a brain area to generate fast
discharges (12.4 127 Hz), and to how early this area becomes involved in seizure. A
normalized EI value is used, ranging from 0 to 1. If there is no involvement of the brain
structure, the EI = 0 (no epileptogenicity) whereas if the brain structure generates a rapid
discharge and the time to seizure onset is minimal, the EI = 1 (maximal epileptogenicity).
In each patient, maximal EI values from at least three representative seizures were computed.
We labeled each pair of bipolar SEEG contacts as belonging to the EZ, propagation zone (PZ)
or non-involved zone (NIZ), as defined by EI, based on previous studies (Aubert et al., 2009;
Lagarde et al., 2019). An EI value of 0.4 and higher was set as a threshold to define a
structure as belonging to the EZ. The PZ was defined as brain areas with 0.1 < EI < 0.4, with
sustained discharge during the seizure course. The NIZ was defined as all other brain
structures.
2.6. ROI Definition
GARDEL software (Medina Villalon et al., 2018) provided the MRI voxel coordinates of
electrode contacts allowing the definition of spherical ROIs for sodium and metabolite
quantification in the patient’s native space. As SEEG recordings processed layout
corresponds to the signal differential of two adjacent electrode contacts, a five mm radius
spherical ROIs were positioned with a center between two adjacent electrode contacts (Ridley
et al., 2017). ROIs corresponding to a poor SEEG signal quality - those located in white
matter for instance - were discarded by expert (J.S.) inspection. To deal with possible
12
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
contamination of the brain sodium signal by an over-representation of CSF in patients, we
discarded ROIs with overly high Z-scoreCSF (see section 2.3. MRI Data Processing).
Anatomical references were the MPRAGE volumes for 1H-MRSI (3T) and the MP2RAGE
volumes for 23Na-MRI (7T). ANTS brain extraction function was applied to all anatomical
and 23Na-MRI images. Each subject’s (patients and controls) anatomical images were
coregistered onto their respective 1H-MRSI and 23Na-MRI maps. Each individual anatomical
volume was also spatially normalized onto the MNI 152 template to obtain the direct and
reverse spatial transforms for each subject. ROIs were projected from the patient native space
to MNI space, and back-projected from MNI to each HC native space (ANTS). This
procedure allowed us to extract normative values of 23Na MRI and 1H-MRSI data from the
control group for each ROI defined in each patient. Details about the numbers of ROIs for
each category were summarized in supplementary Table 1.
2.7. Statistical Analysis
We performed a group comparison of sodium and metabolites levels across the three ROI
classes, EZ, PZ and NIZ, to decipher subtle and specific homeostatic and metabolic
modifications among patients compared to healthy controls. To account for brain sodium and
metabolite variability across participants, sodium and metabolite levels had to be normalized
across participants. Thus, each patient’s sodium concentrations and metabolites levels in each
ROI were expressed as z-scores with respect to the same mean quantities from the
corresponding ROI in healthy controls. Normalization of healthy controls’ sodium
concentrations and metabolites levels was done using a leave-one-out procedure, z-scoring
control’s quantities relative to the same ROI in other controls’ sodium concentrations and
metabolites levels.
13
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
We compared patients and healthy controls, looking for different sodium concentration and/or
metabolic change profiles between them, and between different region categories (described
in ‘SEEG-signal analysis’). Significant group differences of z-scores were analyzed using a
bootstrap two-tailed Welch’s t-test procedure (Efron & Tibshirani, 1993). Achieved
significance levels (ASL) (equivalent to exact p-values) were obtained after one million
random samplings, then corrected for multiple comparisons using false-discovery rate (FDR)
(Benjamini & Hochberg, 1995).
In order to investigate the relationship between sodium concentrations and metabolites levels,
we evaluated how well sodium concentration can be predicted from metabolite profiles.
Hence, we performed a multiple linear regression analysis (Jobson, 1991) using Statsmodels
(Seabold & Perktold, 2010) on all ROIs, with NAA, Cho and tCr z-scores as predictors to
compute linear model for each 23Na-MRI measures (fand TSC) z-scores.
3. Results
3.1. Clinical features
Patients’ clinical characteristics are summarized in Table 1. Mean age at epilepsy onset was
13.2 years (range 0.1-40), mean duration of epilepsy was 18 years (range 4-31). Mean seizure
frequency was 26 per month (range 2-120). Anatomical MRI was normal in fifteen (71%)
and showed a structural abnormality in four (29%) cases. Thirteen out of twenty-one patients
underwent curative surgical procedure following SEEG exploration. From the remaining
eight, six patients were recused from surgery because of bilaterality of the epileptogenic zone
or of a high risk of post-surgical functional impairment, one became seizure-free after
SEEG-guided thermocoagulations and one refused surgery. The post-surgical outcome was
14
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
favorable in eleven (Engel class I (seizure-free) and class II (almost seizure-free), n=11, 85%)
and with worthwhile improvement in two cases (Engel class III, 15%). Epilepsy etiology,
according to histopathological findings, was: focal cortical dysplasia (FCD) type I
(non-detectable on MRI) in six patients, FCD type II in two, and slight gliosis in two. One
patient underwent laser-guided interstitial thermosurgery for an MRI-diagnosed DNET, with
no histology available.
From the initial sample of nineteen patients and according to the quality thresholds
previously described, data analyzes were conducted in fifteen patients for 23Na-MRI and
seventeen patients for 1H-MRSI (Table 1). Patients and healthy controls (HC) did not
significantly differ in terms of age and sex, neither in the 23Na-MRI associated patient group
(Wilcoxon rank sum, w = -0.36, p = 0.72; χ2(1, N=33) = 0.26, p = 0.88) nor 1H-MRSI
associated patient group (Wilcoxon rank sum, w = 0.58, p = 0.56; χ2(1, N=42) = 0.038, p =
0.98).
3.2. Ion homeostasis and metabolic profiles in patients
TSC was increased in epileptic patients relative to corresponding ROIs in healthy controls in
all types of electrophysiologically defined regions, (i.e. EZ , PZ and NIZ) with no significant
differences between these three region categories within patients (see Table 2 and Figure
3.A). The short fraction fwas significantly increased in EZ relative to corresponding ROIs in
healthy controls, but was not significantly different compared to healthy controls in regions
corresponding to PZ and NIZ. In patients, fwithin EZ was also significantly increased
relative to PZ and to NIZ.
15
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
In patients, NAA levels were significantly decreased in all types of regions relative to
controls (see Table 2 and Figure 3.B). In addition, NAA was significantly decreased in EZ
compared to PZ and NIZ. In patients relative to controls, tCho levels were significantly
increased in EZ, decreased in NIZ and not significantly different in PZ; tCr levels were not
significantly different in EZ and PZ but significantly decreased in NIZ.
3.3. Association between ionic and metabolic parameters
To study the relationship between sodium concentrations and metabolite levels, a multiple
linear regression analysis was conducted. We evaluated the prediction of each 23Na-MRI
measure from all 1H-MRSI measures in each region category, namely EZ, PZ and NIZ,
setting Bonferroni corrected p for coefficient t-test was set at 0.05/ 2*3*3 = 0.0028.
Significant regression equations were found for the model predicting TSC from metabolites
predictors alone in PZ (F(3, 80) = 10.76, p < 0.0028) and in NIZ (F(3, 283) = 16.25, p <
0.0028) but not in EZ (F(3, 34) = 1.29, p = 0.295). As mentioned previously, the significant
regressions only partially thinly explained variations in TSC, with of .29 in PZ and a of
.15 in NIZ. We observed a negative association between TSC and NAA in PZ = -0.87, p <
0.0028) and in NIZ = -0.45, p < 0.0028). We also found a negative association between
TSC and Cho and a trend towards a positive association with tCr in NIZ (Table 3). Multiple
linear regression analysis of fdoes not provide significant results in any ROI.
4. Discussion
This work aimed to explore both in vivo ion homeostasis and metabolic alterations in focal
drug-resistant epilepsy investigated by SEEG. We observed global brain impact in patients
reflected by consistent multimodal alterations including an increase of TSC as well as
16
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
decreases in NAA levels within all categories of electrically defined regions, namely EZ, PZ
and NIZ. Interestingly, the EZ showed a characteristic pattern with significantly higher f, as
well as significantly lower NAA and higher Cho levels. TSC exhibits strong association with
metabolites, especially in the PZ and NIZ. These imaging features likely reflect both
hyperexcitability and tissue alterations with a specific pattern of ffor the epileptogenic
tissues.
4.1. Deciphering sodium homeostasis processes with 7T 23Na MRI
Accumulation of TSC is a feature described in a number of neurological diseases including
multiple sclerosis (Maarouf et al., 2017; Zaaraoui et al., 2012), amyotrophic lateral sclerosis
(Grapperon et al., 2019), Hungtington’s disease (Reetz et al., 2012) and epilepsy (Ridley et
al., 2017). Attempts to model processes involved in TSC increases have been proposed
considering in vitro experiments showing increases of intracellular sodium concentrations in
multiple sclerosis (Waxman, 2006). In this model, an increase in TSC the total sodium
concentration - is usually associated with an elevation of intracellular sodium concentration
due to an influx of sodium resulting from a dysfunction of the sodium potassium pump
(Na+/K+ pump) (Pike et al., 1985). However, though sensitive, TSC is not specific to altered
homeostasis between sodium compartments.
TSC alterations in humans in vivo have largely been demonstrated at 3T. Here, to improve
our understanding of the dysregulation mechanisms affecting the sodium ion homeostasis in
epilepsy, we used 7T 23Na MRI. Ultra-high field offers the opportunity to study the
complexity of the T2*decays of the 23Na MR signal influenced by the quadrupolar relaxation.
Indeed, multi-exponential relaxations of this three-half spin reflect the time-dependent
relative position of the quadrupolar moments of sodium nuclei and the electric field gradients
17
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
of charged molecules at their vicinity. The multi-TE density adapted radial 23Na, ultra-short
echo time (UTE) approach used here permits the characterization of the signal decay
components through a bi-exponential fit. While related, the signal fraction of the short T2*
decay component (f), and the total sodium concentration (TSC) may reflect different
consequences of ion homeostasis dysregulation (Ridley et al., 2018).
Normal physiological neuronal activity has been observed to dynamically alter sodium signal
across different TEs, in a manner consilient with fMRI and consistent with physiological
mechanisms that should dominate at different points of ‘activation’ (Bydder et al., 2019)
Using dynamic sodium MRI acquisition compared to BOLD functional MRI, variations in
sodium signals recorded at different TE (0.2ms, 10ms, 19ms) during a right finger tapping
task showed a slightly increased TSC (TE=0.2ms) in the activated left contralateral motor
area interpreted as increased cerebral blood volume, and a more drastic signal decrease at
longer TEs considered, at least in part, to a decrease in extracellular contributions due a
reduced extracellular volume fraction (Antonio et al., 2016; Dietzel et al., 1982; Lux et al.,
1986), while reverse signal variations were observed in the deactivated motor area ipsilateral
to movement. The fmetric, reflecting the variations in the ratio of apparent short and long
T2* sodium signal decays, enables to pool in a single parameter these effects seen at different
TEs, and has the potential to locate regions with abnormal excitability as shown in the present
study.
4.2. Sodium changes in epilepsy
In the context of epilepsy, abnormal TSC increase has been found in EZ and to a lesser extent
also in other regions of the brain at 3T (Ridley et al., 2017). In addition, rat models of
acquired epilepsy have reported persistent TSC increases in affected cortices in response to
kainate-induced epileptogenesis (Mori et al., 2000; Y. Wang et al., 1996). TSC may
18
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
incorporate various processes including both structural reorganization and dysregulation of
ionic homeostasis. Indeed, several structural modifications can impact 23Na-MRI signals,
particularly changes in the extracellular volume (ECV). Cell shrinkage and cerebral atrophy
could lead to ECV increase, and subsequently to increases in TSC. Both have been reported
in epilepsy, localized in the epileptic areas, but also often extended to the non-epileptogenic
regions (Bernhardt et al., 2009; Dingledine et al., 2014; Liu et al., 2005; Voets et al., 2017).
Thus, even though excessively high CSF levels (Z-scoreCSF > 1.95) led to the removal of a
ROI from consideration, its contribution cannot be entirely ruled out. In addition, ECV
increase was recently shown to be related to neuronal excitability (Colbourn et al., 2019).
Moreover, increases of perivascular space were also reported in epilepsy (Feldman et al.,
2018, 2019), and could contribute to TSC accumulation due to the CSF surrounding the
vessels. Recently, venous blood was also demonstrated to have an impact on total sodium
signal (Driver et al., 2020), leading to overestimation of sodium concentration measures in
case of atrophy. In addition, reactive astrogliosis and microgliosis (Devinsky et al., 2013;
Seifert & Steinhäuser, 2013; Sofroniew & Vinters, 2010) can be associated to epilepsy
implying changes of astrocytes proportion, size (Boscia et al., 2016) and ion homeostasis,
leading to surrounding cell homeostasis disruption (Karus et al., 2015). Interestingly, TSC
was also shown to correlate with conductivity at 3T (Liao et al., 2019).
Beyond general explanations for alterations of TSC in epilepsy, cortices subject to different
epileptiform manifestations (EZ, PZ; NIZ) are likely to differ in underlying pathological
mechanisms, something that was probed in the current work through the use of a
multiparametric approach. While keeping in mind that both sodium signal fractions (i.e. short
and long) will contain contributions from extracellular sodium albeit with potentially
different weightings a plausible reason for the concomitant EZ-specific increase relative to
19
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
both HC and other cortices of both fand TSC is an increase of intracellular sodium
concentration. This would be consistent with a range of known mechanisms associated with
the EZ while TSC increase outside the EZ in the absence of changes in f may could be related
to structural changes combined with preserved perfusion, which is usually decreased in the
EZ during the interictal period (Kojan et al., 2021; Y.-H. Wang et al., 2018).
Sodium homeostasis dysregulation in the EZ could be induced by several mechanisms
affecting ion channels such as (i) alterations in type II and III voltage gated sodium channel
(VGNC) properties (Bartolomei et al., 1997; Gastaldi et al., 1997; Gorter et al., 2010;
Lombardo et al., 1996), (ii) incomplete inactivation of sodium channels and a consequent
increase in persistent sodium currents (Mantegazza et al., 2010; Oliva et al., 2012) and (iii)
the reduced efficiency of clearance by the Na+/K+ pump induced by a lack in ATP supplies
(Folbergrová & Kunz, 2012; Grisar et al., 1992; Kovac et al., 2017).
Other mechanisms secondary to hyperexcitability or energy failure affecting astrocytes could
also lead to an alteration of sodium homeostasis and an increase in intracellular sodium
(Gerkau et al., 2017; Kirischuk et al., 2012; Rose & Karus, 2013). Indeed, as a consequence
of hyperexcitability, astrocytes may uptake sodium through various
tranporters(sodium-potassium-chloride co-transporter (NKCC1), sodium-bicarbonate
co-transporter (NBC), sodium-proton exchanger (NHE) and sodium-calcium exchanger
(NCX)) increasing the intracellular sodium concentrations. Heightened intracellular sodium
concentrations reduce the Electrochemical gradient for glutamate uptake (Karus et al., 2015)
by the excitatory amino acid transporter (EAAT), a transmembrane transporter of sodium and
glutamate from peri-synaptic extracellular space into astrocytes, ensuring ion and
neurotransmitter clearance. Then extracellular glutamate concentration increases and
20
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
eventually downregulates EAAT (Rose & Karus, 2013); as astrocytes store energy supplies,
astrocytic energy failure is critical for ion homeostasis of both astrocytes and the surrounding
neurons. The reduced availability of ATP associated with increases in intracellular sodium
also leads to the dysregulation of homeostasis for other ions such as potassium, calcium and
proton, resulting in excitotoxicity (Gerkau et al., 2017). During hyperexcitability, intracellular
pH decreases which leads to sodium uptake via NBC, VGNC or Na+/K+ pump among
others. This reduced intracellular pH was suggested to result from lack of NHE1 (Zhao et al.,
2016). The relationship between reduced intracellular pH and increased intracellular sodium
was also shown in epilepsy during hepatic encephalitis (Kelly et al., 2009) with ammonium
intoxication. The intoxication of cerebral tissue promotes intracellular pH increase (in
particular in astrocytes) resulting in intracellular sodium increase.
4.3. Metabolic changes
The multimodal nature of our investigation provides further evidence of an alteration in ionic
homeostasis in epilepsy due to a disruption of metabolic energy supply. The decreased NAA
we observed has been consistently associated with neuronal death, mitochondrial dysfunction
(Stefano et al., 1995) and subsequent ATP decrease (Vagnozzi et al., 2007). NAA decrease in
the EZ has been widely reported since the 90’s (Guye et al., 2005; Hugg et al., 1993;
Kuzniecky et al., 1998; Petroff et al., 2003; Simister et al., 2002; van der Hel et al., 2013).
While more recent research indicates that decreases in NAA extends to other regions, those
involved in seizures remain the most affected (Guye et al., 2005; Lundbom et al., 2001;
Mueller et al., 2011) as is the case for TSC as well as NAA in this study. This finding is in
line with our energetic failure hypothesis, however, it should be noted we were unable to
observe a significant association between these two measures when explored by multiple
linear regression.
21
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Increased Cho has also been reported in temporal lobe epilepsy (Achten et al., 1997; Simister
et al., 2009) and is frequently associated with cellular proliferation and increased membrane
turnover (Miller, 1991; Urenjak et al., 1993). These processes are likely to be linked to
structural changes associated with epileptogenic lesions such as tumors, tumor-like tissue and
gliosis (van der Hel et al., 2013). The Cho decrease in NIZ was rather unexpected and has not
to our knowledge been previously reported. It could result from reduced cellular density or
neuronal loss extending beyond the EZ.
In our cohort, tCr was significantly decreased in NIZ only. Heighted tCr was previously
related to decreased intracellular energy status or reactive astrocytes in the litterature (Achten,
1998; Urenjak et al., 1993). This result also points to alterations beyond EZ potentially
accompanying structural changes and cognitive co-morbidities (Kreis et al., 1992; Kreis &
Ross, 1992).
4.4. Technical limitations
For both 23Na-MRI and 1H-MRSI, CSF signal contribution is critical as it can clearly bias the
data. MIDAS software handled this bias for 1H-MRSI data (Lecocq et al., 2015). For
23Na-MRI we corrected for partial volume effect and removed ROIs with remaining CSF after
segmentation. We designed a quality check procedure inspired by (Ridley et al., 2017) in
order to get rid of this CSF contribution. Despite this, it is impossible to totally delimit the
partial volume effect completely for the moment, neither for 1H-MRSI nor 23Na-MRI.
Another limitation is inherent to the SEEG procedure, which suffers from limited spatial
sampling, whereas MRI gives access to information across the entire brain. Conversely, the
22
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
pathological specificity offered by SEEG is considered gold-standard and beyond what is
possible with anatomically defined and atlas derived ROIs.
4.5. Conclusion and Perspectives
An increase of the signal fraction of the short T2* decay component (f) was found to be
associated with the EZ whereas increased TSC was not limited to epileptogenic regions.
Taken together with the patterns of metabolite changes our results are in line with the
energetic failure hypothesis in epileptic regions associated with widespread tissue changes
beyond electrically abnormal areas. Here, a multimodal approach has allowed parallel
insights supporting pathological processes in epilepsy. This and additional combinations -
including for example 31P-MRSI, which provides information about energy metabolism via
ATP quantification and pH estimation - could further strengthen the delineation of the EZ as
this non-invasive information would complement the current presurgical evaluation in
patients suffering from drug resistant focal epilepsy.
Acknowledgements
The authors would like to thank L. Pini, C. Costes, and V. Gimenez for data acquisition and
study logistics. We would also like to thank A. Ivanov and C. Bernard for helpful discussions.
Funding
This work has received support from the French government under the “Programme
Investissements d’Avenir”, Excellence Initiative of Aix–Marseille University –A*MIDEX
(AMX-19-IET-004), 7TEAMS Chair, EPINOV (Grant ANR-17-RHUS-0004) and ANR
(ANR-17-EURE-0029); and from the European Union’s Horizon 2020 Framework Program
23
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
for Research and Innovation under the Specific Grant Agreement No. 785907 (Human Brain
Project SGA2) and No. 945539 (Human Brain Project SGA3)
Declaration of Competing Interest
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.
Data Availability Statement
The data that support the findings of this study are available on request from the
corresponding author. The data are not publicly available due to sensitive information that
could compromise the privacy of research participants.
CRediT authorship contribution statement
Mikhael Azilinon: Investigation, Methodology, Data Curation, Formal Analysis, Writing -
original draft. Julia Scholly: Resources, Data Curation, Formal Analysis, Writing - review
and editing. Wafaa Zaaraoui: Funding Acquisition, Writing - review and editing. Samuel
Medina Villalon: Methodology. Patrick Viout: Formal Analysis. Tangi Roussel:
Methodology, Writing - review and editing. Mohamed Mounir El Mendili: Methodology,
Writing - review and editing. Ben Ridley: Methodology, Writing - review and editing.
Jean-Philippe Ranjeva: Supervision, Validation, Methodology, Writing - original draft.
Fabrice Bartolomei: Funding Acquisition, Resources, Writing - review and editing. Viktor
Jirsa: Funding Acquisition, Supervision, Writing - review and editing. Maxime Guye:
Project Administration, Funding Acquisition, Resources, Conceptualization, Supervision,
Validation, Writing - original draft.
24
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
References
Achten, E. (1998). Aspects of proton MR spectroscopy in the seizure patient. Neuroimaging
Clinics of North America,8(4), 849–862.
Achten, E., Boon, P., Kerckhove, T. V. D., Caemaert, J., Reuck, J. D., & Kunnen, M. (1997).
Value of Single-Voxel Proton MR Spectroscopy in Temporal Lobe Epilepsy. 9.
Aja-Fernandez, S., Alberola-Lopez, C., & Westin, C. (2008). Noise and Signal Estimation in
Magnitude MRI and Rician Distributed Images: A LMMSE Approach. IEEE
Transactions on Image Processing,17(8), 1383–1398.
https://doi.org/10.1109/TIP.2008.925382
Antonio, L. L., Anderson, M. L., Angamo, E. A., Gabriel, S., Klaft, Z.-J., Liotta, A., Salar, S.,
Sandow, N., & Heinemann, U. (2016). In vitro seizure like events and changes in
ionic concentration. Journal of Neuroscience Methods,260, 33–44.
https://doi.org/10.1016/j.jneumeth.2015.08.014
Aubert, S., Wendling, F., Regis, J., McGonigal, A., Figarella-Branger, D., Peragut, J.-C.,
Girard, N., Chauvel, P., & Bartolomei, F. (2009). Local and remote epileptogenicity in
focal cortical dysplasias and neurodevelopmental tumours. Brain,132(11),
3072–3086. https://doi.org/10.1093/brain/awp242
Avants, B. B., Tustison, N. J., Song, G., Cook, P. A., Klein, A., & Gee, J. C. (2011). A
reproducible evaluation of ANTs similarity metric performance in brain image
registration. NeuroImage,54(3), 2033–2044.
https://doi.org/10.1016/j.neuroimage.2010.09.025
Bartolomei, F., Chauvel, P., & Wendling, F. (2008). Epileptogenicity of brain structures in
human temporal lobe epilepsy: A quantified study from intracerebral EEG. Brain,
131(7), 1818–1830. https://doi.org/10.1093/brain/awn111
25
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Bartolomei, F., Gastaldi, M., Massacrier, A., Planells, R., Nicolas, S., & Cau, P. (1997).
Changes in the mRNAs encoding subtypes I, II and III sodium channel alpha subunits
following kainate-induced seizures in rat brain. Journal of Neurocytology,26(10),
667–678. https://doi.org/10.1023/a:1018549928277
Benjamini, Y., & Hochberg, Y. (1995). Controlling the False Discovery Rate: A Practical and
Powerful Approach to Multiple Testing. Journal of the Royal Statistical Society.
Series B (Methodological),57(1), 289–300.
Bernhardt, B. C., Rozen, D. A., Worsley, K. J., Evans, A. C., Bernasconi, N., & Bernasconi,
A. (2009). Thalamo–cortical network pathology in idiopathic generalized epilepsy:
Insights from MRI-based morphometric correlation analysis. NeuroImage,46(2),
373–381. https://doi.org/10.1016/j.neuroimage.2009.01.055
Boscia, F., Begum, G., Pignataro, G., Sirabella, R., Cuomo, O., Casamassa, A., Sun, D., &
Annunziato, L. (2016). Glial Na +-dependent ion transporters in pathophysiological
conditions: Glial Na +-dependent Transporters in CNS Diseases. Glia,64(10),
1677–1697. https://doi.org/10.1002/glia.23030
Burstein, D., & Springer, C. S. (2019). Sodium MRI revisited. Magnetic Resonance in
Medicine,82(2), 521–524. https://doi.org/10.1002/mrm.27738
Bydder, M., Zaaraoui, W., Ridley, B., Soubrier, M., Bertinetti, M., Confort-Gouny, S., Schad,
L., Guye, M., & Ranjeva, J.-P. (2019). Dynamic 23Na MRI - A non-invasive window
on neuroglial-vascular mechanisms underlying brain function. NeuroImage,184,
771–780. https://doi.org/10.1016/j.neuroimage.2018.09.071
Colbourn, R., Naik, A., & Hrabetova, S. (2019). ECS Dynamism and Its Influence on
Neuronal Excitability and Seizures. Neurochemical Research,44(5), 1020–1036.
https://doi.org/10.1007/s11064-019-02773-w
Colombet, B., Woodman, M., Badier, J. M., & Bénar, C. G. (2015). AnyWave: A
26
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
cross-platform and modular software for visualizing and processing
electrophysiological signals. Journal of Neuroscience Methods,242, 118–126.
https://doi.org/10.1016/j.jneumeth.2015.01.017
Devinsky, O., Vezzani, A., Najjar, S., De Lanerolle, N. C., & Rogawski, M. A. (2013). Glia
and epilepsy: Excitability and inflammation. Trends in Neurosciences,36(3),
174–184. https://doi.org/10.1016/j.tins.2012.11.008
Dietzel, I., Heinemann, U., Hofmeier, G., & Lux, H. D. (1982). Stimulus-induced changes in
extracellular Na+ and Cl- concentration in relation to changes in the size of the
extracellular space. Experimental Brain Research,46(1), 73–84.
https://doi.org/10.1007/BF00238100
Dingledine, R., Varvel, N. H., & Dudek, F. E. (2014). When and How Do Seizures Kill
Neurons, and Is Cell Death Relevant to Epileptogenesis? In H. E. Scharfman & P. S.
Buckmaster (Eds.), Issues in Clinical Epileptology: A View from the Bench (Vol. 813,
pp. 109–122). Springer Netherlands. https://doi.org/10.1007/978-94-017-8914-1_9
Donadieu, M., Le Fur, Y., Maarouf, A., Gherib, S., Ridley, B., Pini, L., Rapacchi, S.,
Confort-Gouny, S., Guye, M., Schad, L. R., Maudsley, A. A., Pelletier, J., Audoin, B.,
Zaaraoui, W., & Ranjeva, J.-P. (2019). Metabolic counterparts of sodium
accumulation in multiple sclerosis: A whole brain 23Na-MRI and fast 1H-MRSI
study. Multiple Sclerosis Journal,25(1), 39–47.
https://doi.org/10.1177/1352458517736146
Driver, I. D., Stobbe, R. W., Wise, R. G., & Beaulieu, C. (2020). Venous contribution to
sodium MRI in the human brain. Magnetic Resonance in Medicine,83(4),
1331–1338. https://doi.org/10.1002/mrm.27996
Efron, B., & Tibshirani, R. J. (1993). An Introduction to the Bootstrap. Springer US.
https://doi.org/10.1007/978-1-4899-4541-9
27
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Feldman, R. E., Delman, B. N., Pawha, P. S., Dyvorne, H., Rutland, J. W., Yoo, J., Fields, M.
C., Marcuse, L. V., & Balchandani, P. (2019). 7T MRI in epilepsy patients with
previously normal clinical MRI exams compared against healthy controls. PLOS
ONE,14(3), e0213642. https://doi.org/10.1371/journal.pone.0213642
Feldman, R. E., Rutland, J. W., Fields, M. C., Marcuse, L. V., Pawha, P. S., Delman, B. N., &
Balchandani, P. (2018). Quantification of perivascular spaces at 7 T: A potential MRI
biomarker for epilepsy. Seizure,54, 11–18.
https://doi.org/10.1016/j.seizure.2017.11.004
Folbergrová, J., & Kunz, W. S. (2012). Mitochondrial dysfunction in epilepsy.
Mitochondrion,12(1), 35–40. https://doi.org/10.1016/j.mito.2011.04.004
Gastaldi, M., Bartolomei, F., Massacrier, A., Planells, R., Robaglia-Schlupp, A., & Cau, P.
(1997). Increase in mRNAs encoding neonatal II and III sodium channel a-isoforms
during kainate-induced seizures in adult rat hippocampus. 12.
Gerkau, N. J., Rakers, C., Petzold, G. C., & Rose, C. R. (2017). Differential effects of energy
deprivation on intracellular sodium homeostasis in neurons and astrocytes: Effects of
energy deprivation on cellular sodium. Journal of Neuroscience Research,95(11),
2275–2285. https://doi.org/10.1002/jnr.23995
Gorter, J. A., Zurolo, E., Iyer, A., Fluiter, K., Van Vliet, E. A., Baayen, J. C., & Aronica, E.
(2010). Induction of sodium channel Nax (SCN7A) expression in rat and human
hippocampus in temporal lobe epilepsy: SCN7A Expression in TLE. Epilepsia,51(9),
1791–1800. https://doi.org/10.1111/j.1528-1167.2010.02678.x
Grapperon, A.-M., Ridley, B., Verschueren, A., Maarouf, A., Confort-Gouny, S., Fortanier,
E., Schad, L., Guye, M., Ranjeva, J.-P., Attarian, S., & Zaaraoui, W. (2019).
Quantitative Brain Sodium MRI Depicts Corticospinal Impairment in Amyotrophic
Lateral Sclerosis. Radiology,292(2), 422–428.
28
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
https://doi.org/10.1148/radiol.2019182276
Grimaldi, S., El Mendili, M. M., Zaaraoui, W., Ranjeva, J.-P., Azulay, J.-P., Eusebio, A., &
Guye, M. (2021). Increased Sodium Concentration in Substantia Nigra in Early
Parkinson’s Disease: A Preliminary Study With Ultra-High Field (7T) MRI. Frontiers
in Neurology,12, 1610. https://doi.org/10.3389/fneur.2021.715618
Grisar, T., Guillaume, D., & Delgado-Escuet, A. V. (1992). Contribution of Na+,K+-ATPase
to focal epilepsy: A brief review. Epilepsy Research,12(2), 141–149.
https://doi.org/10.1016/0920-1211(92)90034-Q
Guye, M., Le Fur, Y., Confort-Gouny, S., Ranjeva, J.-P., Bartolomei, F., Régis, J., Raybaud,
C. A., Chauvel, P., & Cozzone, P. J. (2002). Metabolic and Electrophysiological
Alterations in Subtypes of Temporal Lobe Epilepsy: A Combined Proton Magnetic
Resonance Spectroscopic Imaging and Depth Electrodes Study. Epilepsia,43(10),
1197–1209. https://doi.org/10.1046/j.1528-1157.2002.05102.x
Guye, M., Ranjeva, J. P., Le Fur, Y., Bartolomei, F., Confort-Gouny, S., Regis, J., Chauvel, P.,
& Cozzone, P. J. (2005). 1H-MRS imaging in intractable frontal lobe epilepsies
characterized by depth electrode recording. NeuroImage,26(4), 1174–1183.
https://doi.org/10.1016/j.neuroimage.2005.03.023
Hugg, J. W., Laxer, K. D., Matson, G. B., Maudsley, A. A., & Weiner, M. W. (1993). Neuron
loss localizes human temporal lobe epilepsy by in vivo proton magnetic resonance
spectroscopic imaging. Annals of Neurology,34(6), 788–794.
https://doi.org/10.1002/ana.410340606
Isnard, J., Taussig, D., Bartolomei, F., Bourdillon, P., Catenoix, H., Chassoux, F., Chipaux,
M., Clémenceau, S., Colnat-Coulbois, S., Denuelle, M., Derrey, S., Devaux, B.,
Dorfmüller, G., Gilard, V., Guenot, M., Job-Chapron, A.-S., Landré, E., Lebas, A.,
Maillard, L., Sauleau, P. (2018). French guidelines on
29
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
stereoelectroencephalography (SEEG). Neurophysiologie Clinique,48(1), 5–13.
https://doi.org/10.1016/j.neucli.2017.11.005
Jobson, J. D. (1991). Multiple Linear Regression. In J. D. Jobson (Ed.), Applied Multivariate
Data Analysis: Regression and Experimental Design (pp. 219–398). Springer.
https://doi.org/10.1007/978-1-4612-0955-3_4
Karus, C., Mondragão, M. A., Ziemens, D., & Rose, C. R. (2015). Astrocytes restrict
discharge duration and neuronal sodium loads during recurrent network activity:
Astrocytes Restrict Neuronal Sodium Loads. Glia,63(6), 936–957.
https://doi.org/10.1002/glia.22793
Kelly, T., Kafitz, K. W., Roderigo, C., & Rose, C. R. (2009). Ammonium-evoked alterations
in intracellular sodium and pH reduce glial glutamate transport activity. Glia,57(9),
921–934. https://doi.org/10.1002/glia.20817
Kirischuk, S., Parpura, V., & Verkhratsky, A. (2012). Sodium dynamics: Another key to
astroglial excitability? Trends in Neurosciences,35(8), 497–506.
https://doi.org/10.1016/j.tins.2012.04.003
Kojan, M., Gajdoš, M., Říha, P., Doležalová, I., Řehák, Z., & Rektor, I. (2021). Arterial Spin
Labeling is a Useful MRI Method for Presurgical Evaluation in MRI-Negative Focal
Epilepsy. Brain Topography,34(4), 504–510.
https://doi.org/10.1007/s10548-021-00833-5
Kovac, S., Dinkova Kostova, A., Herrmann, A., Melzer, N., Meuth, S., & Gorji, A. (2017).
Metabolic and Homeostatic Changes in Seizures and Acquired
Epilepsy—Mitochondria, Calcium Dynamics and Reactive Oxygen Species.
International Journal of Molecular Sciences,18(9), 1935.
https://doi.org/10.3390/ijms18091935
Kreis, R., Ernst, T., & Ross, B. D. (1993). Development of the human brain: In vivo
30
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
quantification of metabolite and water content with proton magnetic resonance
spectroscopy. Magnetic Resonance in Medicine,30(4), 424–437.
https://doi.org/10.1002/mrm.1910300405
Kreis, R., & Ross, B. D. (1992). Cerebral metabolic disturbances in patients with subacute
and chronic diabetes mellitus: Detection with proton MR spectroscopy. Radiology,
184(1), 123–130. https://doi.org/10.1148/radiology.184.1.1319074
Kreis, R., Ross, B. D., Farrow, N. A., & Ackerman, Z. (1992). Metabolic disorders of the
brain in chronic hepatic encephalopathy detected with H-1 MR spectroscopy.
Radiology. https://doi.org/10.1148/radiology.182.1.1345760
Kuzniecky, R., Hugg, J. W., Hetherington, H., Butterworth, E., Bilir, E., Faught, E., &
Gilliam, F. (1998). Relative utility of 1H spectroscopic imaging and hippocampal
volumetry in the lateralization of mesial temporal lobe epilepsy. Neurology,51(1),
66–71. https://doi.org/10.1212/WNL.51.1.66
Lagarde, S., Scholly, J., Popa, I., Valenti-Hirsch, M. P., Trebuchon, A., McGonigal, A., Milh,
M., Staack, A. M., Lannes, B., Lhermitte, B., Proust, F., Benmekhbi, M., Scavarda,
D., Carron, R., Figarella-Branger, D., Hirsch, E., & Bartolomei, F. (2019). Can
histologically normal epileptogenic zone share common electrophysiological
phenotypes with focal cortical dysplasia? SEEG-based study in MRI-negative
epileptic patients. Journal of Neurology,266(8), 1907–1918.
https://doi.org/10.1007/s00415-019-09339-4
Larivière, S., Bernasconi, A., Bernasconi, N., & Bernhardt, B. C. (2021). Connectome
biomarkers of drug-resistant epilepsy. Epilepsia,62(1), 6–24.
https://doi.org/10.1111/epi.16753
Lecocq, A., Le Fur, Y., Maudsley, A. A., Le Troter, A., Sheriff, S., Sabati, M., Donnadieu,
M., Confort-Gouny, S., Cozzone, P. J., Guye, M., & Ranjeva, J.-P. (2015).
31
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Whole-brain quantitative mapping of metabolites using short echo three-dimensional
proton MRSI: 3D- 1H-MRSI Covering the Whole Brain. Journal of Magnetic
Resonance Imaging,42(2), 280–289. https://doi.org/10.1002/jmri.24809
Liao, Y., Lechea, N., Magill, A. W., Worthoff, W. A., Gras, V., & Shah, N. J. (2019).
Correlation of quantitative conductivity mapping and total tissue sodium
concentration at 3T/4T. Magnetic Resonance in Medicine,82(4), 1518–1526.
https://doi.org/10.1002/mrm.27787
Liu, R. S. N., Lemieux, L., Bell, G. S., Sisodiya, S. M., Bartlett, P. A., Shorvon, S. D.,
Sander, J. W. A. S., & Duncan, J. S. (2005). Cerebral Damage in Epilepsy: A
Population-based Longitudinal Quantitative MRI Study. Epilepsia,46(9), 1482–1494.
https://doi.org/10.1111/j.1528-1167.2005.51603.x
Lombardo, A. J., Kuzniecky, R., Powers, R. E., & Brown, G. B. (1996). Altered brain sodium
channel transcript levels in human epilepsy. Molecular Brain Research,35(1–2),
84–90. https://doi.org/10.1016/0169-328X(95)00194-W
Lundbom, N., Gaily, E., Vuori, K., Paetau, R., Liukkonen, E., Rajapakse, J. C., Valanne, L.,
Häkkinen, A., & Granström, M. (2001). Proton Spectroscopic Imaging Shows
Abnormalities in Glial and Neuronal Cell Pools in Frontal Lobe Epilepsy. Epilepsia,
42(12), 1507–1514. https://doi.org/10.1046/j.1528-1157.2001.15301.x
Lux, H. D., Heinemann, U., & Dietzel, I. (1986). Ionic changes and alterations in the size of
the extracellular space during epileptic activity. Advances in Neurology,44, 619–639.
Maarouf, A., Audoin, B., Pariollaud, F., Gherib, S., Rico, A., Soulier, E., Confort-Gouny, S.,
Guye, M., Schad, L., Pelletier, J., Ranjeva, J.-P., & Zaaraoui, W. (2017). Increased
total sodium concentration in gray matter better explains cognition than atrophy in
MS. Neurology,88(3), 289–295. https://doi.org/10.1212/WNL.0000000000003511
Madelin, G., Lee, J.-S., Regatte, R. R., & Jerschow, A. (2014). Sodium MRI: Methods and
32
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
applications. Progress in Nuclear Magnetic Resonance Spectroscopy,79, 14–47.
https://doi.org/10.1016/j.pnmrs.2014.02.001
Mantegazza, M., Curia, G., Biagini, G., Ragsdale, D. S., & Avoli, M. (2010). Voltage-gated
sodium channels as therapeutic targets in epilepsy and other neurological disorders.
The Lancet Neurology,9(4), 413–424.
https://doi.org/10.1016/S1474-4422(10)70059-4
Maudsley, A. A., Darkazanli, A., Alger, J. R., Hall, L. O., Schuff, N., Studholme, C., Yu, Y.,
Ebel, A., Frew, A., Goldgof, D., Gu, Y., Pagare, R., Rousseau, F., Sivasankaran, K.,
Soher, B. J., Weber, P., Young, K., & Zhu, X. (2006). Comprehensive processing,
display and analysis for in vivo MR spectroscopic imaging. NMR in Biomedicine,
19(4), 492–503. https://doi.org/10.1002/nbm.1025
Medina Villalon, S., Paz, R., Roehri, N., Lagarde, S., Pizzo, F., Colombet, B., Bartolomei, F.,
Carron, R., & Bénar, C.-G. (2018). EpiTools, A software suite for presurgical brain
mapping in epilepsy: Intracerebral EEG. Journal of Neuroscience Methods,303,
7–15. https://doi.org/10.1016/j.jneumeth.2018.03.018
Miller, B. L. (1991). A review of chemical issues in1H NMR
spectroscopy:N-acetyl-l-aspartate, creatine and choline. NMR in Biomedicine,4(2),
47–52. https://doi.org/10.1002/nbm.1940040203
Moffett, J. R., Arun, P., Ariyannur, P. S., & Namboodiri, A. M. A. (2013). N-Acetylaspartate
reductions in brain injury: Impact on post-injury neuroenergetics, lipid synthesis, and
protein acetylation. Frontiers in Neuroenergetics,5.
https://doi.org/10.3389/fnene.2013.00011
Mori, Y., Kondziolka, D., Balzer, J., Fellows, W., Flickinger, J. C., Lunsford, L. D., &
Thulborn, K. R. (2000). Effects of stereotactic radiosurgery on an animal model of
hippocampal epilepsy. Neurosurgery,46(1), 157–165; discussion 165-168.
33
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Mueller, S. G., Ebel, A., Barakos, J., Scanlon, C., Cheong, I., Finlay, D., Garcia, P., Weiner,
M. W., & Laxer, K. D. (2011). Widespread extrahippocampal NAA/(Cr+Cho)
abnormalities in TLE with and without mesial temporal sclerosis. Journal of
Neurology,258(4), 603–612. https://doi.org/10.1007/s00415-010-5799-6
Oliva, M., Berkovic, S. F., & Petrou, S. (2012). Sodium channels and the neurobiology of
epilepsy: Sodium Channels and Epilepsy.Epilepsia,53(11), 1849–1859.
https://doi.org/10.1111/j.1528-1167.2012.03631.x
Paling, D., Golay, X., Wheeler-Kingshott, C., Kapoor, R., & Miller, D. (2011). Energy failure
in multiple sclerosis and its investigation using MR techniques. Journal of Neurology,
258(12), 2113–2127. https://doi.org/10.1007/s00415-011-6117-7
Petroff, O. A. C., Errante, L. D., Kim, J. H., & Spencer, D. D. (2003). N-acetyl-aspartate,
total creatine, and myo-inositol in the epileptogenic human hippocampus. Neurology,
60(10), 1646–1651. https://doi.org/10.1212/01.WNL.0000068020.85450.8B
Pike, M. M., Frazer, J. C., Dedrick, D. F., Ingwall, J. S., Allen, P. D., Springer, C. S., &
Smith, T. W. (1985). 23Na and 39K nuclear magnetic resonance studies of perfused
rat hearts. Discrimination of intra- and extracellular ions using a shift reagent.
Biophysical Journal,48(1), 159–173. https://doi.org/10.1016/S0006-3495(85)83769-3
Rajan, J., Poot, D., Junta, J., & Sijbers, J. (2010). Noise measurement from magnitude MRI
using local estimates of variance and skewness. Physics in Medicine and Biology,
55(22), 6973–6973. https://doi.org/10.1088/0031-9155/55/22/6973
Reetz, K., Romanzetti, S., Dogan, I., Saß, C., Werner, C. J., Schiefer, J., Schulz, J. B., &
Shah, N. J. (2012). Increased brain tissue sodium concentration in Huntington’s
Disease—A sodium imaging study at 4T. NeuroImage,63(1), 517–524.
https://doi.org/10.1016/j.neuroimage.2012.07.009
Ridley, B., Marchi, A., Wirsich, J., Soulier, E., Confort-Gouny, S., Schad, L., Bartolomei, F.,
34
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Ranjeva, J.-P., Guye, M., & Zaaraoui, W. (2017). Brain sodium MRI in human
epilepsy: Disturbances of ionic homeostasis reflect the organization of pathological
regions. NeuroImage,157, 173–183.
https://doi.org/10.1016/j.neuroimage.2017.06.011
Ridley, B., Nagel, A. M., Bydder, M., Maarouf, A., Stellmann, J.-P., Gherib, S., Verneuil, J.,
Viout, P., Guye, M., Ranjeva, J.-P., & Zaaraoui, W. (2018). Distribution of brain
sodium long and short relaxation times and concentrations: A multi-echo ultra-high
field 23Na MRI study. Scientific Reports,8(1), 4357.
https://doi.org/10.1038/s41598-018-22711-0
Rooney, W. D., & Springer, C. S. (1991). A comprehensive approach to the analysis and
interpretation of the resonances of spins 3/2 from living systems. NMR in
Biomedicine,4(5), 209–226. https://doi.org/10.1002/nbm.1940040502
Rose, C. R., & Karus, C. (2013). Two sides of the same coin: Sodium homeostasis and
signaling in astrocytes under physiological and pathophysiological conditions:
Sodium Homeostasis in Astrocytes. Glia,61(8), 1191–1205.
https://doi.org/10.1002/glia.22492
Seabold, S., & Perktold, J. (2010). Statsmodels: Econometric and Statistical Modeling with
Python. 92–96. https://doi.org/10.25080/Majora-92bf1922-011
Seifert, G., & Steinhäuser, C. (2013). Neuron–astrocyte signaling and epilepsy. Experimental
Neurology,244, 4–10. https://doi.org/10.1016/j.expneurol.2011.08.024
Simister, R. J., McLean, M. A., Barker, G. J., & Duncan, J. S. (2009). Proton MR
spectroscopy of metabolite concentrations in temporal lobe epilepsy and effect of
temporal lobe resection. Epilepsy Research,83(2), 168–176.
https://doi.org/10.1016/j.eplepsyres.2008.11.006
Simister, R. J., Woermann, F. G., McLean, M. A., Bartlett, P. A., Barker, G. J., & Duncan, J.
35
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
S. (2002). A Short-echo-time Proton Magnetic Resonance Spectroscopic Imaging
Study of Temporal Lobe Epilepsy. Epilepsia,43(9), 1021–1031.
https://doi.org/10.1046/j.1528-1157.2002.50701.x
Sofroniew, M. V., & Vinters, H. V. (2010). Astrocytes: Biology and pathology. Acta
Neuropathologica,119(1), 7–35. https://doi.org/10.1007/s00401-009-0619-8
Statistical Parametric Mapping: The Analysis of Functional Brain Images—1st Edition.
(n.d.). Retrieved November 3, 2021, from
https://www.elsevier.com/books/statistical-parametric-mapping-the-analysis-of-functi
onal-brain-images/penny/978-0-12-372560-8
Stefano, N. D., Matthews, P. M., Ford, B., Genge, A., Karpati, G., & Arnold, D. L. (1995).
Shortterm dichloroacetate treatment improves indices of cerebral metabolism in
patients with mitochondrial disorders. Neurology,45(6), 1193–1198.
https://doi.org/10.1212/WNL.45.6.1193
Stys, P., Waxman, S., & Ransom, B. (1992). Ionic mechanisms of anoxic injury in
mammalian CNS white matter: Role of Na+ channels and Na(+)-Ca2+ exchanger. The
Journal of Neuroscience,12(2), 430–439.
https://doi.org/10.1523/JNEUROSCI.12-02-00430.1992
Urenjak, J., Williams, S. R., Gadian, D. G., & Noble, M. (1993). Proton nuclear magnetic
resonance spectroscopy unambiguously identifies different neural cell types. The
Journal of Neuroscience: The Official Journal of the Society for Neuroscience,13(3),
981–989.
Vagnozzi, R., Tavazzi, B., Signoretti, S., Amorini, A. M., Belli, A., Cimatti, M., Delfini, R.,
Di Pietro, V., Finocchiaro, A., & Lazzarino, G. (2007). TEMPORAL WINDOW OF
METABOLIC BRAIN VULNERABILITY TO CONCUSSIONS. Neurosurgery,
61(2), 379–389. https://doi.org/10.1227/01.NEU.0000280002.41696.D8
36
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
van der Hel, W. S., van Eijsden, P., Bos, I. W. M., de Graaf, R. A., Behar, K. L., van
Nieuwenhuizen, O., de Graan, P. N. E., & Braun, K. P. J. (2013). In vivo MRS and
histochemistry of status epilepticus-induced hippocampal pathology in a juvenile
model of temporal lobe epilepsy: STATUS EPILEPTICUS-INDUCED
HIPPOCAMPAL PATHOLOGY. NMR in Biomedicine,26(2), 132–140.
https://doi.org/10.1002/nbm.2828
Voets, N. L., Hodgetts, C. J., Sen, A., Adcock, J. E., & Emir, U. (2017). Hippocampal MRS
and subfield volumetry at 7T detects dysfunction not specific to seizure focus.
Scientific Reports,7(1), 16138. https://doi.org/10.1038/s41598-017-16046-5
Wang, Y., Majors, A., Najm, I., Xue, M., Comair, Y., Modic, M., & Ng, T. C. (1996).
Postictal alteration of sodium content and apparent diffusion coefficient in epileptic
rat brain induced by kainic acid. Epilepsia,37(10), 1000–1006.
https://doi.org/10.1111/j.1528-1157.1996.tb00539.x
Wang, Y.-H., An, Y., Fan, X.-T., Lu, J., Ren, L.-K., Wei, P.-H., Cui, B.-X., Du, J.-L., Lu, C.,
Wang, D., Zhang, H.-Q., Shan, Y.-Z., & Zhao, G.-G. (2018). Comparison between
simultaneously acquired arterial spin labeling and 18F-FDG PET in mesial temporal
lobe epilepsy assisted by a PET/MR system and SEEG. NeuroImage. Clinical,19,
824–830. https://doi.org/10.1016/j.nicl.2018.06.008
Waxman, S. G. (2006). Axonal conduction and injury in multiple sclerosis: The role of
sodium channels. Nature Reviews. Neuroscience,7(12), 932–941.
https://doi.org/10.1038/nrn2023
Zaaraoui, W., Konstandin, S., Audoin, B., Nagel, A. M., Rico, A., Malikova, I., Soulier, E.,
Viout, P., Confort-Gouny, S., Cozzone, P. J., Pelletier, J., Schad, L. R., & Ranjeva,
J.-P. (2012). Distribution of brain sodium accumulation correlates with disability in
multiple sclerosis: A cross-sectional 23Na MR imaging study. Radiology,264(3),
37
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
859–867. https://doi.org/10.1148/radiol.12112680
Zhao, H., Carney, K. E., Falgoust, L., Pan, J. W., Sun, D., & Zhang, Z. (2016). Emerging
roles of Na+/H+ exchangers in epilepsy and developmental brain disorders. Progress
in Neurobiology,138–140, 19–35. https://doi.org/10.1016/j.pneurobio.2016.02.002
38
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Figures and Tables
Figure 1:3D representation of an example of SEEG electrode implantation scheme (top)
and f, TSC and NAA maps slices corresponding to the transversal green section on the
3D view. Red spherical ROIs correspond to regions belonging to the epileptogenic zone (EZ),
yellow ROIs correspond to regions belonging to the propagation zones (PZ), blue ROIs
correspond to regions non-involved by electrical abnormalities (NIZ) and black ROIs
correspond to regions not explored in this study. Black spheres correspond to regions
excluded from the analyses based on the SEEG signal. The red dot in the transverse planes
(i.e. T1w image and f, TSC and NAA maps) represents the ROI corresponding to electrode
OF’ (left orbitofrontal location) contact 5 and 6 (or bipolar contact 5-6). OF’5-OF’6 ROI is
circled in red in the implantation scheme. Red arrows are pointing to the mean ROI value for
each map.
39
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Figure 2: Spectrogram from a healthy control cortical voxel. Here we highlighted the
three metabolites we focused on in this study, namely total Choline (Cho), N-Acetyl
Aspartate (NAA) and total creatine (tCr).
Glx: glutamate, glutamine and glutathione; m-Ino: myo-Inositol.
40
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Figure 3: Violin plot of mean Z-scores of sodium (A) and metabolite estimates (B) observed
in patients within epileptogenic zones (EZ, red), propagation zones (PZ, yellow) and
non-involved zones (NIZ, black), compared to healthy controls (HC, gray). Asterisks indicate
significant differences between patients and controls when under violins, and between
regions within patients when above violins. *: p-uncorrected < 0.05. **: p-FDR < 0.05
41
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Table 1. Patient clinical demography
Patient
Gender
Epilepsy
duration
(y)
Seizure
frequency
(per month)
Epilepsy type
Side
Surgical procedure
Engel
class
Histology
23Na-MRI/
1H-MRSI
1
F
14
2
Temporal latero-mesial
L
TC, SR
II
FCD1
Both
2
F
26
20
Temporo-parietal mesial
R
TC, VNS, contra-indication for SR
NA
NA
Both
3
M
31
2
Insulo-parietal
L
TC, VNS contra-indication for SR
NA
NA
Both
4
F
9
4
Temporal mesial
R
TC, SR (ATL)
I
FCD3a
Both
5
F
27
3
Temporo-insular
L
TC, SR
I
FCD1
Both
6
M
26
120
Temporal mesial
L
TC, SR
II
slight gliosis
Both
7
M
11
8
Temporal mesial
L
TC, SR
II
FCD1b
Both
8
M
12
4
Temporo-insular
L
TC, contra-indication for SR
NA
NA
Both
9
M
24
10
Temporal lateral
L
TC, SR
III
slight gliosis
Both
10
F
9
30
Temporal mesial
R
TC, SR (ATL)
I
FCD1
Both
11
F
29
60
Orbitofronto-insular
L
TC, SR
III
FCD2a
Both
12
F
10
30
L parietal mesial
Bilateral temporal mesial
R&L
TC, LITT (aiming NDT)
III
NA
Both
13
F
18
100
Bilateral temporal mesial,
R insulo-operculo-frontal
R>L
TC, contra-indication for SR
NA
NA
Both
14
M
4
5
Temporal latero-mesial
L
TC, SR
I
FCD2a
23Na-MRI
15
F
47
12
Temporo-parietal
L
TC, contra-indication for SR
NA
NA
1H-MRSI
16
M
26
12
Temporal mesial
L
TC, patient refused SR
NA
NA
1H-MRSI
17
F
10
5
Bilateral temporal mesial
R&L
TC, DBS, contra-indication for SR
NA
NA
Both
18
M
10
12
Insulo-opercular
L
TC, seizure free
NA
NA
Both
19
M
7
60
Temporal mesial
R
TC, SR (ATL)
III
FCD1c
Both
Abbreviations: ATL, anterior temporal lobectomy; DBS, deep brain stimulation; DNET, dysembryoplastic neuroepithelial tumor; FCD, focal
cortical dysplasia; L, left; LITT, laser-guided interstitial thermal therapy; NA, not applicable; NDT, neuro-developmental tumor; PMG,
polymicrogyria; PNH, periventricular nodular heterotopia; R, right; SR, surgical resection; TC, thermocoagulation; VNS, vagus nerve
stimulation.
42
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
Table 2. Bootstrapped t-value followed by p (or ASL, the bootstrapped achieved
significance level).
EZ Vs HC
PZ Vs HC
NIZ Vs HC
EZ Vs PZ
EZ Vs NIZ
PZ Vs NIZ
f
2.96, 3.8∙10-3
-
-
2.51, 0.012
2.43, 0.015
-
TSC
3.73, 2.5∙10-4
5.35, 1.0∙10-6
12.83, p < 1.0∙10-6
-
-
-
NAA
-6.03,p < 1.0∙10-6
-4.51, 8.0∙10-6
-9.46, p < 1.0∙10-6
-
-2.32, 0.022
-
Cho
3.31, 1.8∙10-3
-
-2.75, 6.4∙10-3
-
3.94, 1.3∙10-4
2.47, 0.015
tCr
-
-
-7.42, p < 1.0∙10-6
-
-
-
FDR-correction yielded p < 0.016 as threshold for tested 23Na-MRI measurements, and p <
0.022 for tested 1H-MRSI.
Table 3. Summary of multiple linear regression analysis for TSC and f.
Parameters
Coefficients
p
TSC
EZ
Const
0,43
0,191
Df Residuals
34
Cho
-0,05
0,837
Df Model
3
NAA
-0,59
0,0743
0.1
tCr
0,39
0,229
F (p)
1.29 (0.295)
PZ
Const
0,86
8,89 ∙ 10-5
Df Residuals
80
Cho
-0,33
0,0787
Df Model
3
NAA
-0,87
9,02 ∙ 10-4*
0.29
tCr
0,56
0,0408
F (p)
10.76 (5.13 ∙ 10-6)
NIZ
Const
0,94
1,60 ∙ 10-17
Df Residuals
283
Cho
-0,29
6,53 ∙ 10-3*
Df Model
3
NAA
-0,45
1,90 ∙ 10-6*
0.15
tCr
0,25
0,0392
F (p)
16.25 (8.96 ∙ 10-10)
f
EZ
Const
0,51
6,15 ∙ 10-3
Df Residuals
34
Cho
-0,33
0,0114
Df Model
3
NAA
0,01
0,967
0.18
tCr
0,24
0,17
F (p)
2.48 (0.0781)
PZ
Const
-0,12
0,448
Df Residuals
80
Cho
0,03
0,832
Df Model
3
NAA
-0,28
0,154
0.03
tCr
0,12
0,572
F (p)
0.94 (0.425)
NIZ
Const
0,02
0,791
Df Residuals
283
Cho
0,11
0,158
Df Model
3
NAA
0,04
0,587
0.01
tCr
-0,16
0,0615
F (p)
1.4 (0.244)
Bonferroni corrected p-values are highlighted in bold followed by an asterisk, while
uncorrected p-values are in bold without asterisk.
43
. CC-BY-NC-ND 4.0 International licenseIt is made available under a perpetuity.
is the author/funder, who has granted medRxiv a license to display the preprint in(which was not certified by peer review)preprint The copyright holder for thisthis version posted August 3, 2022. ; https://doi.org/10.1101/2022.08.03.22278332doi: medRxiv preprint
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Pathophysiology of idiopathic Parkinson's disease (iPD) is complex and still misunderstood. At a time when treatments with disease-modifying potential are being developed, identification of early markers of neurodegeneration is essential. Intracerebral sodium accumulation could be one of them. Indeed, it may be in relation to the mitochondrial dysfunction that early exists in iPD. For the first time, we used brain sodium (²³Na) MRI to explore sodium concentration changes that have already been reported to be related to neurodegeneration in other diseases. We prospectively included 10 iPD patients (mean age 52.2 ± 5.9 years-old) with motor symptoms that started <36 months before inclusion and 12 healthy subjects (mean age 53 ± 6.4 years-old). Patients were scanned in OFF medication state by using proton (¹H) and ²³Na MRI at 7T. We then extracted quantitative Total Sodium Concentration (TSC) from five regions of interest known to be early impaired in iPD [substantia nigra (SN), putamen, caudate nucleus, pallidum, thalamus] and in one region supposed to be relatively spared in the first stages of the disease [cortical gray matter (neocortex)]. Potential atrophy in these structures was also investigated with ¹H MRI. Relative to healthy subjects, iPD patients showed higher TSC in the SN (43.73 ± 4.64 vs. 37.72 ± 5.62, p = 0.006 after Bonferroni correction). A trend of increase in sodium concentrations was found within the pallidum (45.80 ± 4.19 vs. 41.07 ± 4.94, p = 0.017), putamen (48.65 ± 4.58 vs. 43.66 ± 5.04, p = 0.041) and the cortical gray matter (56.34 ± 3.92 vs. 50.81 ± 5.50, p = 0.021). No significant brain atrophy was found in patients compared to controls. Thus, alteration of sodium homeostasis in the SN in the absence of atrophy could be considered as a potential early marker of cellular dysfunction in iPD.
Article
Full-text available
Arterial spin labeling (ASL) is an MRI technique measuring brain perfusion using magnetically labeled blood as a tracer. The clinical utility of ASL for presurgical evaluation in non-lesional epilepsy as compared with the quantitative analysis of interictal [18F] fluorodeoxyglucose PET (FDG-PET) was studied. In 10 patients (4 female; median age 29 years) who underwent a complete presurgical evaluation followed by surgical resection, the presurgical FDG-PET and ASL scans were compared with the resection masks using asymmetry index (AI) maps. The positive predictive value (PPV) and sensitivity (SEN), were calculated from the number of voxels inside the mask (true positive), and outside the mask (false positive). The comparison of the PPVs showed better PPV in 6 patients using ASL and in 2 patients with PET. SEN was better in 4 patients using ASL and in 5 patients with PET. According to the Wilcoxon signed rank test for PPV (p = 0.74) and for SEN (p = 0.43), these methods have similar predictive power. ASL is a useful method for presurgical evaluation in non-lesional epilepsy. The main benefits of ASL over PET are that it avoids radiation exposure for patients, and it offers lower costs, higher availability, and better time efficiency.
Article
Full-text available
Purpose To investigate the correlation between electrical conductivity and sodium concentration, both measured in vivo, in the human brain. Methods Conductivity measurements were performed on samples with different sodium (Na⁺) and agarose concentrations using a dielectric probe, and the correlation between conductivity and Na+ content was evaluated. Subsequently, brain conductivity and total Na+ content maps were measured in 8 healthy subjects using phase‐based MREPT and sodium MRI, respectively. After co‐registration and spatial normalization to the 1 mm 152 MNI brain atlas, the relationship between conductivity and tissue sodium concentration (TSC) was examined within different brain regions. Results The conductivities of agarose gels increased linearly with NaCl concentration, while remaining almost independent of agarose content. When measured in healthy subjects, conductivities showed positive correlation with total tissue sodium concentration (R = 0.39, P < 0.005). The same trend was found in gray matter (R = 0.36, P < 0.005) and in white matter (R = 0.28, P < 0.05). Conclusion Tissue conductivity shows a positive correlation with total sodium concentration. Conductivity might serve as a novel technique to visualize the total tissue electrolyte concentration, although refinements in the consideration of e.g., tissue water content, would be necessary to improve the quantitative value.
Article
Full-text available
Objective To compare by 7 Tesla (7T) magnetic resonance imaging (MRI) in patients with focal epilepsy who have non-lesional clinical MRI scans with healthy controls. Methods 37 patients with focal epilepsy, based on clinical and electroencephalogram (EEG) data, with non-lesional MRIs at clinical field strengths and 21 healthy controls were recruited for the 7T imaging study. The MRI protocol consisted of high resolution T1-weighted, T2-weighted and susceptibility weighted imaging sequences of the entire cortex. The images were read by two neuroradiologists, who were initially blind to clinical data, and then reviewed a second time with knowledge of the seizure onset zone. Results A total of 25 patients had findings with epileptogenic potential. In five patients these were definitely related to their epilepsy, confirmed through surgical intervention, in three they co-localized to the suspected seizure onset zone and likely caused the seizures. In seven patients the imaging findings co-localized to the suspected seizure onset zone but were not the definitive cause, and ten had cortical lesions with epileptogenic potential that did not localize to the suspected seizure onset zone. There were multiple other findings of uncertain significance found in both epilepsy patients and healthy controls. The susceptibility weighted imaging sequence was instrumental in guiding more targeted inspection of the other structural images and aiding in the identification of cortical lesions. Significance Information revealed by the improved resolution and enhanced contrast provided by 7T imaging is valuable in noninvasive identification of lesions in epilepsy patients who are non-lesional at clinical field strengths.
Article
Full-text available
Objective In the detection of seizure onset zones, arterial spin labeling (ASL) can overcome the limitations of positron emission tomography (PET) with ¹⁸F-fluorodeoxyglucose (¹⁸F-FDG), which is invasive, expensive, and radioactive. PET/magnetic resonance (MR) systems have been introduced that allow simultaneous performance of ASL and PET, but comparisons of these techniques with stereoelectroencephalography (SEEG) and comparisons among the treatment outcomes of these techniques are still lacking. Here, we investigate the effectiveness of ASL compared with that of SEEG and their outcomes in localizing mesial temporal lobe epilepsy (MTLE) and assess the correlation between simultaneously acquired PET and ASL. Methods Between October 2016 and August 2017, we retrospectively studied 12 patients diagnosed with pure unilateral MTLE. We extracted and quantitatively computed values for ASL and PET in the bilateral hippocampus. SEEG findings and outcome were considered the gold standard of lateralization. Finally, the bilateral asymmetry index (AI) was calculated to assess the correlation between PET and ASL. Results Our results showed that hypoperfusion in the hippocampus detected using ASL matched the SEEG-defined epileptogenic zone in this series of patients. The mean normalized voxel value of ASL in the contralateral hippocampus was 0.97 ± 0.19, while in the ipsilateral hippocampus, it was 0.84 ± 0.14. Meanwhile, significantly decreased perfusion and metabolism were observed in these patients (Wilcoxon, p < 0.05), with a significant positive correlation between the AI values derived from PET and ASL (Pearson's correlation, r = 0.74, p < 0.05). Significance In our SEEG- and outcome-defined patients with MTLE, ASL could provide significant information during presurgical evaluation, with the hypoperfusion detected with ASL reliably lateralizing MTLE. This non-invasive technique may be used as an alternative diagnostic tool for MTLE lateralization.
Article
Full-text available
White matter of the mammalian CNS suffers irreversible injury when subjected to anoxia/ischemia. However, the mechanisms of anoxic injury in central myelinated tracts are not well understood. Although white matter injury depends on the presence of extracellular Ca2+, the mode of entry of Ca2+ into cells has not been fully characterized. We studied the mechanisms of anoxic injury using the in vitro rat optic nerve, a representative central white matter tract. Functional integrity of the nerves was monitored electrophysiologically by quantitatively measuring the area under the compound action potential, which recovered to 33.5 +/- 9.3% of control after a standard 60 min anoxic insult. Reducing Na+ influx through voltage-gated Na+ channels during anoxia by applying Na+ channel blockers (TTX, saxitoxin) substantially improved recovery; TTX was protective even at concentrations that had little effect on the control compound action potential. Conversely, increasing Na+ channel permeability during anoxia with veratridine resulted in greater injury. Manipulating the transmembrane Na+ gradient at various times before or during anoxia greatly affected the degree of resulting injury; applying zero-Na+ solution (choline or Li+ substituted) before anoxia significantly improved recovery; paradoxically, the same solution applied after the start of anoxia resulted in more injury than control. Thus, ionic conditions that favored reversal of the normal transmembrane Na+ gradient during anoxia promoted injury, suggesting that Ca2+ loading might occur via reverse operation of the Na+)-Ca2+ exchanger. Na(+)- Ca2+ exchanger blockers (bepridil, benzamil, dichlorobenzamil) significantly protected the optic nerve from anoxic injury. Together, these results suggest the following sequence of events leading to anoxic injury in the rat optic nerve: anoxia causes rapid depletion of ATP and membrane depolarization leading to Na+ influx through incompletely inactivated Na+ channels. The resulting rise in the intracellular [Na+], coupled with membrane depolarization, causes damaging levels of Ca2+ to be admitted into the intracellular compartment through reverse operation of the Na(+)-Ca2+ exchanger. These observations emphasize that differences in the pathophysiology of gray and white matter anoxic injury are likely to necessitate multiple strategies for optimal CNS protection.
Article
Full-text available
Sodium (23Na) MRI proffers the possibility of novel information for neurological research but also particular challenges. Uncertainty can arise in in vivo23Na estimates from signal losses given the rapidity of T2* decay due to biexponential relaxation with both short (T2*short) and long (T2*long) components. We build on previous work by characterising the decay curve directly via multi-echo imaging at 7 T in 13 controls with the requisite number, distribution and range to assess the distribution of both in vivo T2*shortand T2*longand in variation between grey and white matter, and subregions. By modelling the relationship between signal and reference concentration and applying it to in vivo23Na-MRI signal,23Na concentrations and apparent transverse relaxation times of different brain regions were measured for the first time. Relaxation components and concentrations differed substantially between regions of differing tissue composition, suggesting sensitivity of multi-echo23Na-MRI toward features of tissue composition. As such, these results raise the prospect of multi-echo23Na-MRI as an adjunct source of information on biochemical mechanisms in both physiological and pathophysiological states.
Article
Full-text available
Ultra high-field 7T MRI offers sensitivity to localize hippocampal pathology in temporal lobe epilepsy (TLE), but has rarely been evaluated in patients with normal-appearing clinical MRI. We applied multimodal 7T MRI to assess if focal subfield atrophy and deviations in brain metabolites characterize epileptic hippocampi. Twelve pre-surgical TLE patients (7 MRI-negative) and age-matched healthy volunteers were scanned at 7T. Hippocampal subfields were manually segmented from 600μm isotropic resolution susceptibility-weighted images. Hippocampal metabolite spectra were acquired to determine absolute concentrations of glutamate, glutamine, myo-inositol, NAA, creatine and choline. We performed case-controls analyses, using permutation testing, to identify abnormalities in hippocampal imaging measures in individual patients, for evaluation against clinical evidence of seizure lateralisation and neuropsychological memory test scores. Volume analyses identified hippocampal subfield atrophy in 9/12 patients (75%), commonly affecting CA3. 7/8 patients had altered metabolite concentrations, most showing reduced glutamine levels (62.5%). However, neither volume nor metabolite deviations consistently lateralized the epileptogenic hippocampus. Rather, lower subiculum volumes and glutamine concentrations correlated with impaired verbal memory performance. Hippocampal subfield and metabolic abnormalities detected at 7T appear to reflect pathophysiological processes beyond epileptogenesis. Despite limited diagnostic contributions, these markers show promise to help elucidate mnemonic processing in TLE.
Article
Purpose: 7T (7T) magnetic resonance imaging (MRI) facilitates the visualization of the brain with resolution and contrast beyond what is available at conventional clinical field strengths, enabling improved detection and quantification of small structural features such as perivascular spaces (PVSs). The distribution of PVSs, detected in vivo at 7T, may act as a biomarker for the effects of epilepsy. In this work, we systematically quantify the PVSs in the brains of epilepsy patients and compare them to healthy controls. Methods: T2-weighted turbo spin echo images were obtained at 7T on 21 epilepsy patients and 17 healthy controls. For all subjects, PVSs were manually marked on Osirix image analysis software. Marked PVSs with diameter≥0.5mm were then mapped by hemisphere and lobe. The asymmetry index (AI) was calculated for each region and the maximum asymmetry index (|AImax|) was reported for each subject. The asymmetry in epilepsy subjects was compared to that of controls, and the region with highest asymmetry was compared to the suspected seizure onset zone. Results: There was a significant difference between the |AImax| in epilepsy subjects and in controls (p=0.016). In 72% of patients, the region or lobe of the brain showing maximum PVS asymmetry was the same as the region containing the suspected seizure onset zone. Conclusion: These findings suggest that epilepsy may be associated with significantly asymmetric distribution of PVSs in the brain. Furthermore, the region of maximal asymmetry of the PVSs may help provide localization or confirmation of the seizure onset zone.