ArticlePDF Available

Variability by region and method in human brain sodium concentrations estimated by 23Na magnetic resonance imaging: a meta-analysis

Authors:
  • IRCCS - Istituto delle Scienze Neurologiche di Bologna (Italy)

Abstract and Figures

Sodium imaging (²³Na-MRI) is of interest in neurological conditions given potential sensitivity to the physiological and metabolic status of tissues. Benchmarks have so far been restricted to parenchyma or grey/white matter (GM/WM). We investigate (1) the availability of evidence, (2) regional pooled estimates and (3) variability attributable to region/methodology. MEDLINE literature search for tissue sodium concentration (TSC) measured in specified ‘healthy’ brain regions returned 127 reports, plus 278 retrieved from bibliographies. 28 studies met inclusion criteria, including 400 individuals. Reporting variability led to nested data structure, so we used multilevel meta-analysis and a random effects model to pool effect sizes. The pooled mean from 141 TSC estimates was 40.51 mM (95% CI 37.59–43.44; p < 0.001, I²Total=99.4%). Tissue as a moderator was significant (F²14 = 65.34, p-val < .01). Six sub-regional pooled means with requisite statistical power were derived. We were unable to consider most methodological and demographic factors sought because of non-reporting, but each factor included beyond tissue improved model fit. Significant residual heterogeneity remained. The current estimates provide an empirical point of departure for better understanding in ²³Na-MRI. Improving on current estimates supports: (1) larger, more representative data collection/sharing, including (2) regional data, and (3) agreement on full reporting standards.
²³Na magnetic resonance imaging. (a) An exemplar ²³Na-MRI brain image, with external calibration phantoms of differing concentrations visible on the bottom left axial image. Calibration can also be done relative to internal references such as vitreous humor or cerebrospinal fluid in the ventricles. (b) Schematic Venn diagram (following the nomenclature in Springer²⁸) showing physical domains (Blue) where nuclei share similar nuclear magnetic resonance properties, and biological compartments (Red) found in brain tissue. These are arrayed along two axes to indicate their orthogonality: neither physical domain is unique to any given biological compartment. ²³Na-MRI TSC estimates are a weighted average influenced by concentrations, volumes and microstructure in a range of environments including the intracellular (neuronal and glial cytosol, and organelles), extracellular (interstitial and vascular spaces.) and membranous (cyto/axolemmal) contributions²⁸. The vast majority of in vivo sodium is present in metal-aquo complexes with a tetrahedral hydration shell surrounding the ²³Na ion, with a much smaller population bound to macromolecular loci⁸⁰. The domain on the left corresponds to the situation in bulk solution, where magnetic and electric fields average out to become isotropic. An anisotropic domain (right) pertains at the interface of/with macromolecules and/or lipid assemblies, where the surface experienced by a diffusing ion is not randomly orientated and the resulting electric field gradient (EFG) fluctuations do not average to zero. ²³Na is a quadropolar ion (spin = 3/2) that, under the influence of a magnetic field, exhibits four energy levels with three possible single quantum transitions, one central and two satellites each contributing to relaxation⁸¹. In the context of isotropic domains, i.e. aqueous environments with rapid motions, quadrupole interactions are minimal and all transitions occur approximately at the same decay time resulting in a MRI-visible monoexponential decay curve⁸¹. In anisotropic domains, where motions are slowed, the non-spherical distribution of the electric charge of the sodium nucleus permits interaction with anisotropic electric fields of the charged groups on the macromolecular anions. Thus, quadrupole interactions are non-zero and biexponential relaxation is observed66,80–82 with the satellite transitions showing faster decay than the central transition.
… 
This content is subject to copyright. Terms and conditions apply.
1
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports
Variability by region and method
in human brain sodium
concentrations estimated by 23Na
magnetic resonance imaging:
a meta‑analysis
Ben Ridley
1,4*, Filomena Morsillo
1, Wafaa Zaaraoui
2,3 & Francesco Nonino
1
Sodium imaging (23Na‑MRI) is of interest in neurological conditions given potential sensitivity to the
physiological and metabolic status of tissues. Benchmarks have so far been restricted to parenchyma
or grey/white matter (GM/WM). We investigate (1) the availability of evidence, (2) regional pooled
estimates and (3) variability attributable to region/methodology. MEDLINE literature search for
tissue sodium concentration (TSC) measured in specied ‘healthy’ brain regions returned 127 reports,
plus 278 retrieved from bibliographies. 28 studies met inclusion criteria, including 400 individuals.
Reporting variability led to nested data structure, so we used multilevel meta‑analysis and a random
eects model to pool eect sizes. The pooled mean from 141 TSC estimates was 40.51 mM (95% CI
37.59–43.44; p < 0.001, I2Total=99.4%). Tissue as a moderator was signicant (F214 = 65.34, p‑val < .01).
Six sub‑regional pooled means with requisite statistical power were derived. We were unable to
consider most methodological and demographic factors sought because of non‑reporting, but each
factor included beyond tissue improved model t. Signicant residual heterogeneity remained. The
current estimates provide an empirical point of departure for better understanding in 23Na‑MRI.
Improving on current estimates supports: (1) larger, more representative data collection/sharing,
including (2) regional data, and (3) agreement on full reporting standards.
Sodium magnetic resonance imaging (23Na-MRI) is of interest as a ‘quantitative’ imaging modality. Using a refer-
ence of known concentration (Fig.1a), measured signal (M0) can be converted from arbitrary signal intensity
to a quantitative scale (millimolars, mM). As a candidate for metabolic imaging in particular, the advantages of
23Na-MRI include: (1) it natively produces 3D, whole-brain, voxel-based data and is not restricted to pre-dened
volume-of-interest analyses, and (2) the fact it requires no contrast agents, meaning contraindications are the
same as for conventional proton (1H) MRI. Ionic homeostasis is a pre-requisite for proper cellular functioning,
with sodium in the nervous system being critical in trans-membrane transport, osmotic and electrostatic regula-
tion and the generation/propagation of action potentials13. As such, the non-invasive, invivo measurement of
sodium concentration by 23Na-MRI is of interest in the context of neuro-oncological46, neurodegenerative79,
demyelinating1019 and cerebrovascular20 conditions, and in both physiological and pathological neuronal
activity2125.
In practice, assigning a single imaging parameter to a voxel in an MRI image of biological tissue is an over-
simplication. is is the case both for the weighted average referred to as ‘total’ or ‘tissue sodium concentration
(TSC) in 23Na-MRI, as well as ‘conventional’ MRI contrasts targeting tissue water protons (1H) such as diusion
or T2 measurements26,27. Image sampling/tissue fraction eects are one reason, where diverse tissue types such
as white and grey matter (WM and GM) and cerebrospinal uid (CSF) contribute to measured signal within a
single voxel. Even within a given tissue type MRI cannot resolve the sub-cellular compartments/organelles that,
in the case of 23Na-MRI, can be said to actually have a specic concentration28. e measurement of a given
OPEN
1IRCCS Istituto Delle Scienze Neurologiche di Bologna, Bologna, Italy. 2Aix Marseille Univ, CNRS, CRMBM,
Marseille, France. 3APHM, Hôpital de La Timone, CEMEREM, Marseille, France. 4Ben Ridley, Epidemiologia e
Statistica, IRCCS Istituto Delle Scienze Neurologiche di Bologna, Padiglione G, Via Altura, 3, 40139 Bologna,
Italy. *email: ben.ridley@ausl.bologna.it
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
voxel will include, at a minimum, intracellular and extracellular compartments with their own concentrations,
volumes and microstructure (Fig.1b).
Physical behaviour of sodium atoms in biological media and the complexities of measuring them with MRI
present other challenges. Relative to tissue water protons, the lower MR sensitivity and abundance of 23Na result
in lower signal to noise ratios and larger voxel sizes, exacerbating tissue fraction eects29,30. 23Na-MRI pulse
sequences with ultra-fast echo times (TE) that compensate for the short, biexponential transverse relaxation of
23Na nuclei2932, oen use non-cartesian sampling schemes which can have broader point spread functions (PSF)
and greater inter-voxel spill-over eects in comparison to 1H-MRI. ese partial volume eect (PVE) issues are
the target of growing attempts to develop or import PVE correction techniques, such as those adapted from
positron emission tomography imaging (PET)12,2932. Correction techniques beg the question of benchmarks
for correction algorithms to target.
A range of tissue volume models to understand and validate 23Na-MRI-derived concentrations have been
proposed. e ‘canonical’ model describes two compartments: a large volume/low concentration (variously
10–15 mM5,6,10,20,33,34) intracellular space and an extracellular space of smaller volume but higher concentration
(140 mM5,6,10,20 or 145 mM33,34). ese gures lead to a general estimate for overall brain tissue (parenchyma) of
about 37–45 mM6,12,20. Model-based estimates beyond a general gure for parenchyma or gross tissue divisions
like GM or WM are largely lacking35, particularly because the cellular data required to elaborate beyond this are
for the most part not available. More broadly, histology is an invaluable tool but should not be naively taken as
the absolute gold-standard for MR features both because it is not necessarily sensitive to the same properties as
Figure1.
23Na magnetic resonance imaging. (a) An exemplar 23Na-MRI brain image, with external calibration
phantoms of diering concentrations visible on the bottom le axial image. Calibration can also be done
relative to internal references such as vitreous humor or cerebrospinal uid in the ventricles. (b) Schematic
Venn diagram (following the nomenclature in Springer28) showing physical domains (Blue) where nuclei share
similar nuclear magnetic resonance properties, and biological compartments (Red) found in brain tissue.
ese are arrayed along two axes to indicate their orthogonality: neither physical domain is unique to any
given biological compartment. 23Na-MRI TSC estimates are a weighted average inuenced by concentrations,
volumes and microstructure in a range of environments including the intracellular (neuronal and glial
cytosol, and organelles), extracellular (interstitial and vascular spaces.) and membranous (cyto/axolemmal)
contributions28. e vast majority of invivo sodium is present in metal-aquo complexes with a tetrahedral
hydration shell surrounding the 23Na ion, with a much smaller population bound to macromolecular loci80.
e domain on the le corresponds to the situation in bulk solution, where magnetic and electric elds average
out to become isotropic. An anisotropic domain (right) pertains at the interface of/with macromolecules
and/or lipid assemblies, where the surface experienced by a diusing ion is not randomly orientated and the
resulting electric eld gradient (EFG) uctuations do not average to zero. 23Na is a quadropolar ion (spin = 3/2)
that, under the inuence of a magnetic eld, exhibits four energy levels with three possible single quantum
transitions, one central and two satellites each contributing to relaxation81. In the context of isotropic domains,
i.e. aqueous environments with rapid motions, quadrupole interactions are minimal and all transitions occur
approximately at the same decay time resulting in a MRI-visible monoexponential decay curve81. In anisotropic
domains, where motions are slowed, the non-spherical distribution of the electric charge of the sodium nucleus
permits interaction with anisotropic electric elds of the charged groups on the macromolecular anions. us,
quadrupole interactions are non-zero and biexponential relaxation is observed66,8082 with the satellite transitions
showing faster decay than the central transition.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
MR27 and because various techniques suer from their own limitations regarding compartmental concentra-
tions, ecological validity given preparation eects, and limited spatial, temporal and cross-species sampling28.
e positioning of 23Na-MRI as a putative ‘quantitative’ method, implies that concentration measures should
converge toward a ‘true’ estimate for a given sample, modulo statistical and methodological eects. If 23Na-MRI is
sensitive to the physiological state of tissues—a key assumption motivating its use in the context of neurological
conditions—regional variability in measured TSC concentrations should also be expected. Conversely, positing
a single parenchymal concentration as sucient to characterise all regions implies a limit on detectable dier-
ences between individuals with and without neurological conditions, given known variation in factors like uid
fractions36,37, distributions of cellular types and architectures38,39 and macromolecular content4042.
In this context, meta-analytic approaches are another means to synthesize evidence and identify impediments
and progress towards consensus. Meta-analysis aims to estimate the true eect size (including central tendency
measures like the mean) based on the combination of observed eect sizes taken from several empirical samples,
while trying to account for sample and study variability43. As such, we sought to apply a meta-analytic approach
to investigate the existing literature on estimates of TSC in human brain regions. We aimed to address (1) the
range of available evidence in the form of regional TSC estimates in the literature, (2) the possibility of consensus
estimates of concentration in various brain regions, and (3) the extent to which methodological and anatomical
factors contribute to variation in measured TSC.
Results
Search results. A search (see Methods) of MEDLINE dated 12/7/2021 returned 127 records, and we identi-
ed an additional 278 records by examining the bibliographies of recovered records and the ‘cited by’ function
on the PubMed website. ese records underwent screening of titles and abstracts, and the remaining texts
underwent full text assessment for inclusion (see Preferred Reporting Items for Systematic Reviews and Meta-
Analyses (PRISMA) ow diagram Fig.2, and Supplementary Tables1, 2 for PRISMA checklists). From an overall
total of 405 records, 44 records were screened from further consideration based on title/abstract. 361 records
were sought for full text retrieval, of which 329 were excluded with the main reasons being a focus on non-brain
tissue or the lack of a sodium measurement in the form of a concentration estimate. We were unable to access
four records. See Supplementary Table3 for a full list of records identied and reasons for exclusion. Inclusion
criteria were met by 28 reports.
Included studies. We included 28 reports containing measurements of total sodium concentration in
healthy controls of specied human brain regions or tissue divisions (other than ‘parenchyma’) (Table1). Nomi-
nally, 400 healthy controls in total were included in these reports, the mean number per report being 14.3 indi-
viduals (range: 4–45, S.D: 11.1). All but three included reports were published aer 2010 (Fig.3a).
We used a modied version of the checklists associated with the Committee on Best Practice in Data Analy-
sis and Sharing (COBIDAS)44, appropriate to the context of 23Na-MRI: specically, the sections on descriptive
statistics, image acquisition reporting and pre-processing reporting. A list of reporting domains included in
the modied version can be found in Supplementary Table4, along with the coded results from the included
reports and a summary can be seen in Fig.3b. Information relating to the numbers of included participants,
whether informed consent was given, MRI scanner used, repetition time (TR), echo time (TE), pulse sequence
and nominal resolution was provided by all included studies. No included report provided information of the
distribution of handedness in the included groups. e remaining domains were reported by varying numbers
of included reports.
Figure2. PRISMA ow diagram for search performed 12/7/2021. Generated with the PRISMA 2020 app83.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
Descriptive. From the 28 included studies, a total of 162 eect sizes in the form of means and SD of meas-
ured TSC (mM) in healthy controls were extracted. In addition, relevant information relating to the COBIDAS
domains for which all included studies had relevant data were extracted, except for ‘Scanner’ where the report-
ing was too variable and insucient to permit reclassication. To account for variability in nomenclature, ‘Pulse
sequence’ and ‘Tissue’ were re-coded according to Table1 and Supplementary Table5, respectively. Tissue re-
coding was informed by the wish to maximise the data per region, tissue homogeneity and physiological plausi-
bility and minimise non-independence, and we excluded twenty eect sizes from regions too sparsely sampled
to satisfy these considerations. e remaining 141 eects sizes across 14 updated Tissue regions in were taken
forward for meta-analysis (Fig.3c): 22 eect sizes for both GM and WM regions; 14 in the Brainstem and Pons
combined; 10 each in Central WM, alamus, and ‘GM, Temporal’ regions; eight each in ‘Deep WM’ and WM
in the cerebellum and dentate nucleus (WM, Cb + DN); seven each in ‘GM, Parietal’ and Putamen; six each in
Caudate, Globus Pallidus and ‘GM, Frontal’ and ve in ‘GM, Occipital’.
We also recorded the type of calibration method used—an external phantom, or internal references in the
vitreous humour of the eyes or the ventricles of the brain. Since important demographic factors like age and sex
were not completely reported COBIDAS domains (Fig.3b), we also recorded ‘Comparison group’ in the hopes
that this might capture some of the variability associated with missing demographic information. Where controls
were selected based on matching for age and sex to a patient group, the characteristic range for these factors
in certain patient groups could represent a sampling bias varying between conditions e.g. controls matched to
patients with Huntingdons disease versus multiple sclerosis. Reports without a patient-dened comparison group
can oen be technical MR methodology papers, which may tend to sample from a dierent population, such as
for example the authors/students themselves.
Multi‑level meta‑analysis: model‑t. We used a multilevel/multivariable approach with four levels
(participant, eect size, tissue regions and study), and a random eects model to pool eect sizes. e pooled
Table 1. Characteristics of included studies. a Int. internal, Ext. external, CSF cerebrospinal uid, VH vitreous
humour (eyes). b Repetition time. c Echo time, where multiple the shortest TE was used in analysis. d x,y,z
product of nominal resolution, see Supplementary Table7 for specic values. e NR = not reported, other values
are presented as per original reports in ranges (XX-XX), mean ± standard deviation, except for Zaaraoui etal.
2012 and Maarouf etal. 2014 which report median and range and Inglese etal. 2010 and Eisele etal. 2016
which report mean and range.
Paper Ref Tesla Sequence CalibrationaTRb (ms) TEc(ms) Voxel volume (mm3)dN N. Female Age (years)eComparison group
Winkler etal. 1989 45 1.5 GRE Int. VH 98 3 490 4 NR 20–35 None
Ouwerkerk etal. 2003 51.5 TPI External 120 0.37 39.3 9 3 22–63 Tumour
ulborn etal. 2005 87 3TPI External 100 0.3 125 5 NR NR Stroke
Inglese etal. 2010 10 3 Radial External 120 0.05 64 13 10 36.7, 26–60 MS
Lu etal. 2010 33 3 FlexTPI External 160 0.36 125 5 1 32.4 ± 8.9 None
Reetz etal. 2012 84 SPRITE External 10 0.3 64 13 6 44.9 ± 9.9 HD
Qian etal. 2012 71 7 AWSOS Int. CSF 100 0.5 2.9 5 5 20–48 None
Zaaraoui et 2012 11 3 DA radial External 120 0.2 46.6 15 12 30.20–54 MS
Paling etal. 2013 12 3 Radial External 120 0.27 64 27 16 42.9 ± 11.3 MS
Maarouf etal. 2014 13 3 DA radial External 120 0.2 46.6 15 NR 30, 21–54 MS
Mirkes etal. 2015 73 9.4 AWSOS External 150 0.3 5 5 1 29 ± 4 None
Niesporak etal. 2015 88 7 DA radial External 150 0.45 27 4 1 26 ± 2 None
Eisele etal. 2016 15 3 DA radial External 60 0.22 46.6 10 5 33.23–53 MS
Petracca etal. 2016 16 7 GRE External 150 6.8 125 17 8 46.16 ± 11.65 MS
ulborn etal. 2016 34 9.4 FlexTPI External 160 0.26 42.9 45 NR 48 ± 19 None
Maarouf etal. 2017 14 3 DA radial External 120 0.2 46.7 31 15 35.7 ± 12.4 MS
Eisele etal. 2017 17 3 DA radial External 60 0.22 46.7 6 5 42 ± 10 MS
Ridley etal. 2018 63 7 DA radial External 120 0.3 42.9 13 5 23.9 ± 3.6 None
Wortho etal. 2018 46 4 SISTINA Int. VH 150 0.36 216 40 16 19–70 None
Reimer etal. 2019 72 3Cones External 100 0.5 64 11 3 32 ± 6 None
Driver etal. 2019 89 4.7 TPI Int. VH 85 0.11 65.5 9 5 30 ± 6 None
Liao etal. 2019 90 3TPI Int. CSF 160 0.4 40.7 8 3 25–32 None
Meyer etal. 2019a 91 3 DA radial External 120 0.2 46.7 12 8 31 ± 8.3 None
Meyer etal. 2019b 24 3 DA radial External 120 0.2 64 12 12 34.3 ± 10.7 Migraine
Kim etal. 2020 32 7 GRE Int. CSF 100 4 64 8 0 20–35 None
Gerhalter etal. 2021 92 3 FLORET Int. VH 100 0.2 216 19 12 31.4 ± 7.5 TBI
Brownlee etal. 2019 19 3Cones External 120 0.22 27 34 23 35.5 ± 10.1 MS
Schneider etal. 2021 59 7 DA radial Int. CSF 100 0.35 8 5 3 28.4 ± 6.5 None
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
mean of TSC across all 141 estimates based on the multilevel meta-analytic model was 40.51mM (95% CI
37.59–43.44; p < 0.001). We identied considerable heterogeneity (I2Tota l=99.4%), with the results of the standard
random-eects model suggesting that most of the total variance is due to between-study heterogeneity (i.e., vari-
ance in the ‘true’ means), while the remaining (0.6%) can be attributed to sampling variance.
Variance components: tissue and study factors. e factors Tissue and Study were included in the
model to account for the nested structure of dependencies within the data: there are multiple eect sizes per
paper and the eect sizes are not independent—a given individual may have contributed to multiple levels of the
factor Tissue, and a given region may be estimated based on dierent numbers of eect sizes from multiple and
varying numbers of papers. e estimated variance components were t = 45.81 for the Study level, 9.93 for Tissue
level and 28.96 for Eect Size level. In terms of the distribution of variance across levels as a percentage of total
variance, 53.75% of the total variation in our data can be attributed to between-study heterogeneity, 11.65% to
between-tissue heterogeneity and 33.99% to the eect size level (i.e. within-factor heterogeneity for Tissue), and
only 0.6% is due to sampling variance. e high heterogeneity both between Studies and within Tissue, suggest
that a subgroup analysis by anatomical region is appropriate.
Model comparison: tissue sub‑group versus reduced model. In Fig.4 we report the forest plots for
anatomical regions with at least ten eect sizes only (following general statistical power guidelines for meta-
analytic sub-analyses43). ese included (pooled mean [95% CI]): GM (45.92mM [42.27; 49.57]), Temporal GM
Figure3. Descriptive plots for identied and included studies. (a) Literature search results 1980–2021 by
year of publication. Chart includes total papers published for a given year (orange), publications identied by
searching bibliographies (cyan), papers identied through the MEDLINE search (red) and number of studies
included (green). (b) Overview of COBIDAS Domains reported in included studies. FOV, Field of view; PSF,
Point spread function; ROIs, regions of interest. (c) Scatterplot of 141 eect sizes used in meta-analysis by
published report in alphabetical order of rst author surname. Error bars correspond to standard deviation,
except for Zaaraoui etal. 2012 who reported range and Driver etal. 2019 who reported standard error. Cb
cerebellum, DN dentate nucleus, GM grey matter, WM white matter. Images created in R8486.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
(45.48mM [40.92; 50.05]), alamus (42.08mM [36.38; 47.77]), Brainstem + Pons (40.99mM [33.28; 48.70]),
WM (37.29mM [33.04; 41.53]) and ‘Central’ WM (34.67mM [27.89; 41.44]). We found that the subgroup mul-
tilevel model provided a signicantly better t compared to a reduced multilevel model, as indicated by lower
Akaike (AIC) and Bayesian Information Criterion (BIC). e likelihood ratio test (LRT) comparing both models
is signicant (χ213 = 145.50, p < 0.01). A test of the moderators ‘Tissue’ was signicant, F214 = 65.34, p-val < 0.01),
indicating that in this model the mean TSC is dierent between each anatomical region. However, the results
indicate that high heterogeneity remains overall even with the inclusion of ‘Tissue’ level (χ2127 = 9537.26,
p-val < 0.01).
Precision, small study eects and publication bias. We investigated further factors impacting the
distribution of results with respect to pooled means via Egger’s tests and funnel plots (Fig.5). e egger’s test
used standard error, a measure of precision, as predictor. ings being equal, there should an inverse relationship
between standard error and the probability of a given study’s estimate being dierent from the actual value in the
population, with an expected symmetry in over- and under-estimates. Overall, the distribution of all eect sizes
did not show the expected distribution, showing substantial asymmetry (Egger’s test, t = 5.24, p < 0.001) which
remained statistically signicant when outliers (identied by sensitivity analysis eliminating one by one the
extreme points of the distribution) are removed (t = 4.89, p < 0.001). Asymmetry can be evidence for small study
Figure4. Forest plots for anatomical regions with at least ten eect sizes: greymatter;Temporalgreymatter;
thalamus;white matter;central white matter;brainstem + pons. Each forest plot contains the eect size data,
represented by grey squares scaled to their weight in the meta-analytic model and error bars corresponding to
95% condence intervals. e regional pooled estimate for each plot is represented by a grey diamond scaled
in length to the condence interval of the pooled estimates, and a dotted reference line. e pooled overall
mean of all 141 included eect sizes is represented by a solid reference line on each plot. Plots generated in R,
in the metafor package74.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
eects and publication bias, based on the assumption that small studies are at greatest risk of non-signicant
results and biasing the published literature toward the high eect sizes that are most likely to be signicant with
Figure5. Funnel plots comparing eect sizes (mean TSC) with their precision (standard error): All tissues
(141 eect sizes) and with outliers removed (136 eect sizes); GM eect sizes with (22, le) and without
outliers (20, right); WM eect sizes with (22, le) and without outliers (19, right); GM, Temporal; alamus;
Brainstem + Pons, and Central WM. Images created in R74,84.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
small N43. In our case we are investigating a measure of central tendency: mean TSC, as opposed to a standard-
ized mean dierence, compared to standard error. As such, we investigated the possibility that our data does not
show the expected symmetry because it is constituted by sub-groups, and the ’pooled mean’ is not the best refer-
ence given the subgroup analysis, above.
In GM the test of asymmetry was at the threshold of signicance (t = 2.1, p = 0.05), though aer removing
some outliers (60 mM45; 69 mM5) the gures become more symmetrical and the tests become non-signicant
(t = 1.73, p = 0.1) and the remaining regions do not show signicant skewness. Similarly, in WM the presence of
asymmetry is indicated (t = 2.48, p = 0.02), but aer removal of some outliers (19.38 mM10, 25 mM46, 69 mM5)
distribution was no longer signicantly asymmetrical. e remaining regions did not show signicant asymmetry.
Eect of moderators. Papers diered in their methodology across domains. We sought to understand the
eect of additional methodological moderators by adding them individually and exploring their association with
dierent mean TSC estimates. Relative to a reduced model including the factor Tissue, all additional individual
factor added to the model produced a signicantly better t (Table2), in terms of reducing overall heterogeneity.
Moderator eects on mean TSC. Within factors, we explored the levels associated with signicant dier-
ences in mean TSC between levels independent of Tissue (Fig.6). For studies using Sequence Type as “Radial”
or “SISTINA, the mean TSC is signicantly lower than that of studies using the “DA Radial” type (t118 = −2.95,
p < 0.01 and t118 = −2, p < 0.05 respectively). Field strength of 1.5T is associated with higher concentrations of
Sodium (t122 = 2.54, p = 0.01) relative to 3T. In studies with “Brain Tumour” as the Comparison Group the mean
TSC is higher in controls (t118 = 2.68, p < 0.01) than in studies where there was no comparison group. A single
study is sampled for the SISTINA level of the factor Sequence while another single study was sampled both at
the 1.5 Tesla level of the factor Field Strength and the Brain Tumour level of Comparison Group, which was also
identied as an outlier in the assessment of asymmetry in standard error distribution for both GM and WM5
(Figs.5, 6). Tests for Calibration method, Voxel Volume, TR and TE were non-signicant, indicating no associa-
tion between the level of the two moderators and the mean measured sodium level, regardless of region.
Intra‑regional heterogeneity. Given the high heterogeneity within the factor Tissue, we extended our
analysis to identifying where inclusion of methodological moderators reduces the heterogeneity of estimates
within anatomical regions (with at least 10 eects sizes) suggesting the pooled estimates in the reduced model
are impacted by dierences in a given factor. We compared the specic heterogeneity (tau) of a region in the
reduced model compared to the model with the methodological factor using Hedges’ g, identifying ‘signicant’
reductions in the form of a standardized mean dierence whose 95% CIs did not cross zero (Supplementary
Table6, Supplementary Fig.1).
‘Sequence’ as moderator reduces the heterogeneity for GM, GM-Temporal and WM regions. Adding “Com-
parison Group” reduces heterogeneity in GM, Brainstem + Pons, and WM. “Calibration method” reduces hetero-
geneity within GM. Residual heterogeneity is reduced in “Brainstem + Pons” when “Voxel Volume” is used as a
moderator. “TR” reduced residual heterogeneity when used as a moderator in Temporal GM and the thalamus.
“TE” reduced heterogeneity in Temporal GM and Central WM. Field strength did not contribute to explaining
the variability within anatomical regions. Note that for all methodological factors, the test of residual heteroge-
neity for the model overall remained signicant (Table2).
Discussion
Data from 28 studies, identied by literature search, were explored via meta-analysis—to our knowledge the
rst such attempt in the context of data from 23Na-MRI. e overall pooled estimate from all 141 across all 28
studies was 40.51mM (95% CI 37.59–43.44), well within the ranges suggested by parenchymal volume models
(37–45 mM6,12,20). Meta-analytic estimates were associated with high heterogeneity, which further analysis sug-
gested was largely associated with between-study heterogeneity. is supports the idea that there is underlying
dierences in the ‘true means’ the dierent studies are trying to measure—and that a parenchymal estimate is not
sucient to characterise the range of empirical values obtained from dierent brain regions. Pooled estimates
Table 2. Tests comparing models including each methodological moderator to areduced model. DF degrees
of freedom.
Moderator
Likelihood ratio test Test of residual
heterogeneity
DF χ2pDF χ2p
Sequence 9 78.07 < 0.01 118 4772.90 < 0.01
Comparison group 6 50.73 < 0.01 121 9081.33 < 0.01
Calibration method 2 16.82 < 0.01 125 8636.05 < 0.01
Voxel volume 1 7.93 < 0.01 125 7963.39 < 0.01
Field strength (Tesla) 5 26.81 < 0.01 126 9216.67 < 0.01
Repetition time (TR) 1 29.74 < 0.01 126 8697.18 < 0.01
Echo time (TE) 1 16.01 < 0.01 126 9529.9 < 0.01
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
based on the extant literature of TSC estimates in human samples are 45.92mM [42.27; 49.57] for GM and
37.29mM [33.04; 41.53] for WM. is is noticeably higher than tissue volume model-based estimates for both
tissue types, with examples including estimates of 20–33mM in WM and for 30–35 for GM33,47.
A number of potential sources may account for discrepancy between theoretical tissue volume models and
the empirical estimates. One is partial volume eects, and other methodological factors impacting acquisition
of the estimates making up the literature explored here. e models use simplifying assumptions, most notably
that they have adequately captured the relevant inuences on sodium 23Na-MRI measurement with a limited
Figure6. Scatterplot of 141 eect sizes used in meta-analysis ordered by moderators with a mean eect on TSC
between levels independent of Tissue. Error bars correspond to standard deviation, except for Zaaraoui etal.
2012 (DA Radial, 3 Tesla, Multiple Sclerosis) who reported range and Driver etal. 2019 (TPI, 4.7 Tesla, None)
who reported standard error. Cb, cerebellum, DN dentate nucleus, GM grey matter, WM white matter. Images
created in R8486.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
number of compartmental volume contributions, and that these compartments are internally homogenous. For
example, the extracellular compartment is generally (but not always35) considered to include everything outside
cells membranes, thereby subsuming the interstitial extracellular matrix48 and vascular spaces and attributing
the 23Na concentration of ‘pure’ CSF to the entire compartment20,33,35,49. Similarly, the contribution of mem-
branes, lipids like myelin and other ‘solids’ are assumed to be captured by a single volume contribution with no
sodium contribution (on the basis they exclude sodium and should reduce overall measured signal for a given
volume35,49), and can be summarised by a single fractional variable (e.g. 0.7–0.933,35,37) for a given tissue type. It
would be interesting to see how the modication of any or all these foctors impacts interact to produce expected
values and how this might be applied to regional estimates.
More granular tissue models are currently lacking, as the requisite cellular data is not available. A potentially
relevant factor is regional variation in the ratios of dierent cell types in the context of divergent sodium con-
centrations, for example astrocytes have approximately twice the cytosolic sodium concentration (~ 15–20mM)
compared to neurons (~ 10mM) in rodent samples5054. Recent automated immunocytochemical techniques
have provided much needed information, including correcting widespread misconceptions about neuronal ver-
sus non-neuronal populations and masses, however precise data on cell volumes and their variation—which
would be relevant for building regionally specic volume models for 23Na-MRI—are not yet available38,39. In the
absence of complete histologic information, another source of insight could come from comparing 23Na-MRI
to other imaging indices that might capture relevant features, with other quantitative imaging modalities being
of particular interest. Parallel changes in 23Na-MRI measures and diusion imaging15,5557 and proton density58
are suggestive, but extending these to direct evaluations of the redundancy and complementarity with other
quantitative modalities59, especially in healthy controls, in a range of regions and tissue structures would be a
welcome development.
While the empirical values are higher than the model-based estimates for a given tissue, the relative values
of dierent tissues concentrations (e.g. GM > WM) is preserved. e dierence in myelin content between grey
and white matter—captured in the dierences in the solid fractions that are usually assigned—may account for
some of this dierence. Indeed, among the subcortical regions we were able to provide pooled estimates for (> 10
eect sizes), it is noticeable that intermediate values were produced. e ostensibly GM nucleus though highly
myelinated alamus has a lower value (42.08mM [36.38; 47.77]) than some other GM estimates (45.48mM
[40.92; 50.05] for Temporal, GM), while ROIs sampling regions that are likely to be predominantly WM but with
contributions from GM nuclei like the brainstem and pons indicate higher values (40.99mM [33.28; 48.70])
than some other regions (34.67mM [27.89; 41.44] for ‘Central WM’). While the overlap between condence
intervals and the remaining heterogeneity limits the certainty of these precise pooled means, which should not
be taken as denitive given the limitations of the available literature as represented in this meta-analysis, the
importance of considering the variation in apparent 23Na associated with regional dierences is reected in the
signicant improvement in the model when including the tissue factor and the nding of signicant dierences
in mean TSC between levels/regions.
Given the remaining unexplained heterogeneity even aer the Tissue factor was involved, we explored addi-
tional methodological factors where full reporting made this possible. e fundamentally most limiting property
of 23Na-MRI is the reduced nuclear MR sensitivity and relative abundance of sodium and other non-1H based
contrasts60, leading to reduced signal to noise ratios and resolution, and consequently to partial volume eects.
A given ‘Sequence’ is an attempt mitigate between trade-os in acquisition parameters with impacts on available
signal and resolution (e.g. TE, TR, Flip angle, Voxel Volume, Field Strength). For example, the ideal sequence
would entail a spin density weighting with a minimum of relaxation eects, however in practice studies will
dier in the degree they are aected by T1 or T2 weighting and thus vary in quantication. Dierent internal
and external calibration methods provide scope for dierent degrees of experimenter error as well as spatial and
physiological variability61. Comparison group may reect demographic factors of potential relevance62 reecting
the target patient group and other variable experimenter/location factors.
We were unable to consider the majority of the methodological and demographic factors sought by the modi-
ed COBIDAS checklist because the information was not reported (Fig.3b). Explicit reference to and information
pertaining to the following COBIDAS domain were not identied for age, handedness, sex, coil information,
acquisition time, processing soware used, ip angle, segmentation and ROI denition, details of normalisa-
tion/registration, eld of view, information pertaining to smoothing or point spread function. Working towards
consensus reporting standards could facilitate comparability of studies in the future. e COBIDAS standards
in general, and the adapted subset used here (Fig.3, Supplementary Table4) could represent a starting point, to
which further 23Na-MRI specic parameters could be added such as relaxation and B1 correction methods, as
well as phantom calibration.
We found each fully-reported methodological factor with included beyond tissue improved the meta-analytic
model t (Table2), but that signicant residual heterogeneity remained regardless. Only Field Strength, Sequence,
and Comparison Group diered in mean TSC between specic levels independently of tissue. Interpretations
of these results should take into consideration the risk of bias due to sampling issues given the limited data for
various levels of these factors. Considered in combination with Tissue, all methodological factors except for Field
Strength reduced heterogeneity in some regions when included (Supplementary Fig.1). Collectively these results
stress the importance of methodological factors but also the limitations of the available literature and underline
the need for more and completely reported data covering multiple acquisition schemes and brain regions.
We analysed estimates of Total/Tissue Sodium Concentration, as the most common measure available. Other
23Na-MRI derived metrics are possible, for example there are approaches that measure63 or lter sodium signal
based on relaxation behaviour (e.g. inversion recovery, IR64), or multiple quantum ltering (MQF)16. In principle,
any specic measurement of invivo sodium by 23Na-MRI—TSC, IR, MQF or other – cannot be said to derive
from a single cellular-level tissue compartment65. However, while attribution to dierent sources is a subject
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
of longstanding and ongoing investigation3,66,67, it is legitimate to discuss a dierence or change in measured
parameters (between conditions and across spatial/temporal domains) in terms of changes in concentration or
structure in sub-compartments that may have contributed, even when the latter are below the limit of resolu-
tion. In practice, precisely attributing changes in empirical MR-level estimates to compartmental micro-features
is unlikely to be denitive because these factors rarely alter in isolation. For example, while pathological TSC
alterations may be related to metabolic impairments of transmembrane23Na exchange, they may also reect
changes in cellular death, swelling, proliferation etc6870. Fundamentally, 23Na-MRI appears to be a more sensitive
than specic measure, and a claim that it is sensitive to variation in a particular structural, functional or patho-/
physiological context is an empirical question to answered by further appropriate data and not modelling nor a
priori arguments from incomplete biophysical data alone.
Some further limitations should be noted and considered in future analysis. Variable reporting required
accommodations to be made in several factors that were included in the analysis. Allocation of a given data point
to a particular level of Tissue was based on explicit textual references in the included studies, but the method of
segmentation, precise anatomic boundaries, use of atlases and precise coordinates were not always clear. is
may be another source of heterogeneity in the results, and again highlights the need for clear reporting. We also
considered only published reports, and not ‘grey literature’ (e.g. dissertations, preprints, government reports,
or conference proceedings)43 which could potentially improve sampling. We produced and examined a central
tendency measure via mean estimates of TSC, but if sucient data were available meta-analytic analysis could be
applied to other 23Na-MRI metrics as well as combined estimates of dierences between groups, structures, and
states. Finally, we considered the impact of methodological parameters in isolation, while to fully characterise
their impact it will likely be necessary to investigate their interactions.
Conclusions
Data from 28 studies, identied by literature search, were explored via meta-analysis—to our knowledge the rst
such attempt in the context of data from 23Na-MRI. e nested nature of the data, due in part to accommoda-
tions made to the variability of reporting in the published studies, lead to the use of a multi-level meta-analytic
approach. We produced pooled meta-analytic estimates of brain TSC, but signicant remaining heterogeneity
limits the certainty and precision associated with the estimates. Consideration of tissue dierences explains part
of that heterogeneity, but not all. Where they were fully reported, methodological moderators were explored.
While their inclusion reduces heterogeneity within certain tissue regions, and eects the measured TSC levels,
substantial residual heterogeneity remains. e current estimates provide an empirical point of departure for
better understanding of variability in 23Na-MRI. Improving on current estimates supports: (1) larger, more rep-
resentative data collection/sharing, including (2) regional data, and (3) agreement on full reporting standards.
Methods
Literature search. e following MEDLINE search was run by BR via pubmed.ncbi.nlm.nih.gov on
12/07/2021: “Brain” [Title/Abstract] AND ((((sodium MRI [Title/Abstract]) OR 23Na MRI [Title/Abstract])
OR sodium imaging [Title/Abstract]) OR 23Na imaging [Title/Abstract])”. Bibliographies of potentially eligible
studies were consulted and studies of potential relevance to 23Na-MRI were included in the screening.
Recovered records were excluded based on the abstract or full text if they were non-experimental, non-
original reports (review/commentary), conference proceedings, phantom-only studies, concerned with cultured
tissue or organs other than the brain, non-human subjects, or did not include estimates of sodium concentrations
determined by quantitative 23Na-MRI in healthy subjects without known neurological conditions. Studies con-
sidering only estimates in overall parenchyma, or where it was not possible to attribute estimates to a specied
anatomical region were also excluded. Screening and full text review were performed by BR with reference to
other authors as necessary.
Data extraction. Initial extraction of data pertaining to 23Na-MRI concentrations and methodological
domains was performed by BR, with verication and consultation with WZ. Data re-coding, as discussed in the
Descriptive” section in Results, was based on consensus decisions by WZ/BR for “Sequence” and FN/BR for
“Tissue” (See Table1 and Supplementary Table5). Where a given study included multiple potentially relevant
samples a context-based decision was made: where the same individuals were sampled with diering methods
we took the highest eld / highest resolution observation for Qian etal.71; where the same individual was sam-
pled multiple times with the same methods (reproducibility studies) we took the overall mean across samples for
Riemer etal.72 and Meyer etal.24; where PVC-corrected values by dierent methods were used in Kim etal.32 we
used the spill-over and ventricular CSF-based PVC-corrected values. Where multiple TEs were reported45,63,73
we used the acquisition with the fastest TE.
Meta‑analysis. Meta-analyses were conducted in R (R-4.1.2) using the “metafor” package74. e restricted
maximum likelihood estimator75 was used to calculate the heterogeneity variance (τ2) and we used Knapp-
Hartung adjustments76 to calculate the condence interval around the pooled eect. Multi-level models were
investigated to account for any correlations induced by the multi-level structure of the data, whereby a given
individual may have contributed to multiple levels of the factor Tissue, and a given region may be estimated
based on dierent numbers of eect sizes from multiple and varying numbers of papers. To account for cor-
related sampling errors due to dierent eect sizes being based on the same sample of patients we used a Corre-
lated and Hierarchical Eects (CHE) model77: an extension of the multilevel model that considers the correlation
of eect sizes within clusters, in this case the factor ‘Paper’. A robust Sandwich covariate estimator was used to
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
estimate condence intervals and relative p-values78. Egger’s tests79 were used to evaluate asymmetry of funnel
plots, based on weighted regression models with multiplicative dispersion, with standard error as the predictor.
Data availability
All data generated or analysed during this study are included in this published article and its Supplementary
Information les.
Received: 4 November 2022; Accepted: 21 February 2023
References
1. Madelin, G. & Regatte, R. R. Biomedical applications of sodium MRI invivo. J. Magn. Reson. Imaging JMRI 38, 511–529 (2013).
2. Chatton, J.-Y., Magistretti, P. J. & Barros, L. F. Sodium signaling and astrocyte energy metabolism. Glia 64, 1667–1676 (2016).
3. ulborn, K. R. Quantitative sodium MR imaging: A review of its evolving role in medicine. Neuroimage 168, 250–268 (2018).
4. Leslie, T. K. et al. Sodium homeostasis in the tumour microenvironment. Biochim. Biophys. Acta Rev. Cancer 1872, 188304 (2019).
5. Ouwerkerk, R., Bleich, K. B., Gillen, J. S., Pomper, M. G. & Bottomley, P. A. Tissue sodium concentration in human brain tumors
as measured with 23Na MR imaging. Radiology 227, 529–537 (2003).
6. ulborn, K. R., Davis, D., Adams, H., Gindin, T. & Zhou, J. Quantitative tissue sodium concentration mapping of the growth of
focal cerebral tumors with sodium magnetic resonance imaging. Magn. Reson. Med. 41, 351–359 (1999).
7. Grapperon, A.-M. et al. Quantitative brain sodium MRI depicts corticospinal impairment in amyotrophic lateral sclerosis. Radiol-
ogy 292, 422–428 (2019).
8. Reetz, K. et al. Increased brain tissue sodium concentration in Huntington’s Disease—A sodium imaging study at 4 T. Neuroimage
63, 517–524 (2012).
9. Mohamed, S. A. et al. Evaluation of sodium (23Na) MR-imaging as a biomarker and predictor for neurodegenerative changes in
patients with Alzheimer’s disease. Vivo Athens Greece 35, 429–435 (2021).
10. Inglese, M. et al. Brain tissue sodium concentration in multiple sclerosis: A sodium imaging study at 3 tesla. Brain J. Neurol. 133,
847–857 (2010).
11. Zaaraoui, W. et al. Distribution of brain sodium accumulation correlates with disability in multiple sclerosis: A cross-sectional
23Na MR imaging study. Radiology 264, 859–867 (2012).
12. Paling, D. et al. Sodium accumulation is associated with disability and a progressive course in multiple sclerosis. Brain J. Neurol.
136, 2305–2317 (2013).
13. Maarouf, A. et al. Topography of brain sodium accumulation in progressive multiple sclerosis. Magma N. Y. N 27, 53–62 (2014).
14. Maarouf, A. et al. Increased total sodium concentration in gray matter better explains cognition than atrophy in MS. Neurology
88, 289–295 (2017).
15. Eisele, P. et al. Heterogeneity of acute multiple sclerosis lesions on sodium (23Na) MRI. Mult. Scler. Houndmills Basingstoke Engl.
22, 1040–1047 (2016).
16. Petracca, M. et al. Brain intra- and extracellular sodium concentration in multiple sclerosis: A 7 T MRI study. Brain J. Neurol. 139,
795–806 (2016).
17. Eisele, P. et al. Sodium MRI of T1 high signal intensity in the dentate nucleus due to gadolinium deposition in multiple sclerosis.
J. Neuroimaging O. J. Am. Soc. Neuroimaging 27, 372–375 (2017).
18. Grist, J. T. et al. Imaging intralesional heterogeneity of sodium concentration in multiple sclerosis: Initial evidence from 23Na-MRI.
J. Neurol. Sci. 387, 111–114 (2018).
19. Brownlee, W. J. et al. Cortical grey matter sodium accumulation is associated with disability and secondary progressive disease
course in relapse-onset multiple sclerosis. J. Neurol. Neurosurg. Psychiatry 90, 755–760 (2019).
20. ulborn, K. R., Gindin, T. S., Davis, D. & Erb, P. Comprehensive MR imaging protocol for stroke management: Tissue sodium
concentration as a measure of tissue viability in nonhuman primate studies and in clinical studies. Radiology 213, 156–166 (1999).
21. Bydder, M. et al. Dynamic 23Na MRI - A non-invasive window on neuroglial-vascular mechanisms underlying brain function.
Neuroimage 184, 771–780 (2019).
22. Gandini Wheeler-Kingshott, C. A. M. et al. Challenges and perspectives of quantitative functional sodium imaging (fNaI). Front.
Neurosci. 12, 810 (2018).
23. Ridley, B. et al. Brain sodium MRI in human epilepsy: Disturbances of ionic homeostasis reect the organization of pathological
regions. Neuroimage 157, 173–183 (2017).
24. Meyer, M. M. et al. Cerebral sodium (23Na) magnetic resonance imaging in patients with migraine—A case-control study. Eur.
Radiol. 29, 7055–7062 (2019).
25. Azilinon, M. et al. Combining sodium MRI, proton MR spectroscopic imaging, and intracerebral EEG in epilepsy. Hum. Brain
Mapp. https:// doi. org/ 10. 1002/ hbm. 26102 (2022).
26. Dowell, N. & Wood, T. T2: Transverse relaxation time. In Quantitative MRI of the Brain: Principles of Physical Measurement 83–96
(CRC Press, 2018). https:// doi. org/ 10. 1201/ b21837-6.
27. Novikov, D. S., Kiselev, V. G. & Jespersen, S. N. On modeling. Magn. Reson. Med. 79, 3172–3193 (2018).
28. Springer, C. S. Measurement of metal cation compartmentalization in tissue by high-resolution metal cation NMR. Annu. Rev.
Biophys. Biophys. Chem. 16, 375–399 (1987).
29. Niesporek, S. C. et al. Partial volume correction for invivo 23Na-MRI data of the human brain. Neuroimage 112, 353–363 (2015).
30. Gnahm, C. & Nagel, A. M. Anatomically weighted second-order total variation reconstruction of 23Na MRI using prior informa-
tion from 1H MRI. Neuroimage 105, 452–461 (2015).
31. Niesporek, S. C. et al. Improved T2* determination in (23)Na, (35)Cl, and (17)O MRI using iterative partial volume correction
based on (1)H MRI segmentation. Magma N. Y. N. https:// doi. org/ 10. 1007/ s10334- 017- 0623-2 (2017).
32. Kim, S.-Y. et al. Voxel-wise partial volume correction method for accurate estimation of tissue sodium concentration in 23 Na-MRI
at 7 T. NMR Biomed. 34, e4448 (2021).
33. Lu, A., Atkinson, I. C., Claiborne, T. C., Damen, F. C. & ulborn, K. R. Quantitative sodium imaging with a exible twisted
projection pulse sequence. Magn. Reson. Med. 63, 1583–1593 (2010).
34. ulborn, K. et al. Quantitative sodium MRI of the human brain at 9.4 T provides assessment of tissue sodium concentration and
cell volume fraction during normal aging. NMR Biomed. 29, 137–143 (2016).
35. Gilles, A., Nagel, A. M. & Madelin, G. Multipulse sodium magnetic resonance imaging for multicompartment quantication:
Proof-of-concept. Sci. Rep. 7, 17435 (2017).
36. Neeb, H., Schenk, J. & Weber, B. Multicentre absolute myelin water content mapping: Development of a whole brain atlas and
application to low-grade multiple sclerosis. NeuroImage Clin. 1, 121–130 (2012).
37. Filo, S. & Mezer, A. PD: Proton density of tissue water. In Quantitative MRI of the Brain: Principles of Physical Measurement 55–72
(CRC Press, 2018). https:// doi. org/ 10. 1201/ b21837-4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
13
Vol.:(0123456789)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
38. Herculano-Houzel, S. & Dos Santos, S. E. You do not mess with the glia. Neuroglia 1, 193–219 (2018).
39. Dos Santos, S. E. et al. Similar microglial cell densities across brain structures and mammalian species: Implications for brain
tissue function. J. Neurosci. O. J. Soc. Neurosci. 40, 4622–4643 (2020).
40. Harris, A. D., Puts, N. A. J., Barker, P. B. & Edden, R. A. E. Spectral-editing measurements of GABA in the human brain with and
without macromolecule suppression. Magn. Reson. Med. 74, 1523–1529 (2015).
41. Rooney, W. D. et al. Mapping human brain capillary water lifetime: high-resolution metabolic neuroimaging. NMR Biomed. 28,
607–623 (2015).
42. Považan, M. et al. Simultaneous mapping of metabolites and individual macromolecular components via ultra-short acquisition
delay 1 H MRSI in the brain at 7T. Magn. Reson. Med. 79, 1231–1240 (2018).
43. Harrer, M., Cuijpers, P., Furukawa, T. A. & Ebert, D. D. Doing Meta-Analysis with R: A Hands-On Guide (Chapman & Hall/CRC
Press, 2021).
44. Nichols, T. E. et al. Best practices in data analysis and sharing in neuroimaging using MRI. Nat. Neurosci. https:// doi. org/ 10. 1101/
054262 (2016).
45. Winkler, S. S., omasson, D. M., Sherwood, K. & Perman, W. H. Regional T2 and sodium concentration estimates in the normal
human brain by sodium-23 MR imaging at 1.5 T. J. Comput. Assist. Tomogr. 13, 561–566 (1989).
46. Wortho, W. A., Shymanskaya, A. & Shah, N. J. Relaxometry and quantication in simultaneously acquired single and triple
quantum ltered sodium MRI. Magn. Reson. Med. 81, 303–315 (2019).
47. Fleysher, L. et al. Noninvasive quantication of intracellular sodium in human brain using ultrahigh-eld MRI. NMR Biomed. 26,
9–19 (2013).
48. Syková, E. & Nicholson, C. Diusion in brain extracellular space. Physiol. Rev. 88, 1277–1340 (2008).
49. Madelin, G., Babb, J., Xia, D. & Regatte, R. R. Repeatability of quantitative sodium magnetic resonance imaging for estimating
pseudo-intracellular sodium concentration and pseudo-extracellular volume fraction in brain at 3 T. PLoS ONE 10, e0118692
(2015).
50. Knöpfel, T., Guatteo, E., Bernardi, G. & Mercuri, N. B. Hyperpolarization induces a rise in intracellular sodium concentration in
dopamine cells of the substantia nigra pars compacta. Eur. J. Neurosci. 10, 1926–1929 (1998).
51. Pisani, A., Calabresi, P., Tozzi, A., Bernardi, G. & Knöpfel, T. Early sodium elevations induced by combined oxygen and glucose
deprivation in pyramidal cortical neurons. Eur. J. Neurosci. 10, 3572–3574 (1998).
52. Rose, C. R. & Ransom, B. R. Intracellular sodium homeostasis in rat hippocampal astrocytes. J. Physiol. 491(Pt 2), 291–305 (1996).
53. Rose, C. R. & Ransom, B. R. Gap junctions equalize intracellular Na+ concentration in astrocytes. Glia 20, 299–307 (1997).
54. Unichenko, P., Myakhar, O. & Kirischuk, S. Intracellular Na+ concentration inuences short-term plasticity of glutamate trans-
porter-mediated currents in neocortical astrocytes. Glia 60, 605–614 (2012).
55. El Mendili, M. M. et al. Alterations of microstructure and sodium homeostasis in fast amyotrophic lateral sclerosis progressors:
A brain DTI and sodium MRI study. AJNR Am. J. Neuroradiol. 43, 984–990 (2022).
56. Kolbe, S. C. et al. Microstructural correlates of 23Na relaxation in human brain at 7 Tesla. Neuroimage 211, 116609 (2020).
57. Collorone, S. et al. Brain microstructural and metabolic alterations detected invivo at onset of the rst demyelinating event. Brain
J. Neurol. https:// doi. org/ 10. 1093/ brain/ awab0 43 (2021).
58. Rodriguez, G. G. et al. Repeatability of simultaneous 3D 1H MRF/23Na MRI in brain at 7 T. Sci. Rep. 12, 14156 (2022).
59. Schneider, T. M. et al. Multiparametric MRI for characterization of the basal ganglia and the midbrain. Front. Neurosci. https://
doi. org/ 10. 3389/ fnins. 2021. 661504 (2021).
60. Niesporek, S. C., Nagel, A. M. & Platt, T. Multinuclear MRI at ultrahigh elds. Top. Magn. Reson. Imaging 28, 173–188 (2019).
61. Adlung, A. et al. Quantication of tissue sodium concentration in the ischemic stroke: A comparison between external and internal
references for 23Na MRI. J. Neurosci. Methods 382, 109721. https:// doi. org/ 10. 1016/j. jneum eth. 2022. 109721 (2022).
62. ulborn, K. R. Gender dierences in cell volume fraction (CVF): a structural parameter reecting the energy eciency of main-
taining the resting membrane potential. NMR Biomed. 35, e4693 (2022).
63. Ridley, B. et al. Distribution of brain sodium long and short relaxation times and concentrations: a multi-echo ultra-high eld
23Na MRI study. Sci. Rep. 8, 4357 (2018).
64. Nagel, A. M. et al.e potential of relaxation-weighted sodium magnetic resonance imaging as demonstrated on brain tumors.
Invest. Radiol. 46, 539–547 (2011).
65. Burstein, D. & Springer, C. S. Sodium MRI revisited. Magn. Reson. Med. 82, 521–524 (2019).
66. Hutchison, R. B. & Shapiro, J. I. Measurement of intracellular sodium with NMR methods. Concepts Magn. Reson. 3, 215–236
(1991).
67. Hoesl, M. A. U. et al. 23 Na Triple-quantum signal of invitro human liver cells, liposomes, and nanoparticles: Cell viability assess-
ment vs. separation of intra- and extracellular signal. J. Magn. Reson. Imaging JMRI https:// doi. org/ 10. 1002/ jmri. 26666 (2019).
68. Inglese, M., Oesingmann, N., Zaaraoui, W., Ranjeva, J. P. & Fleysher, L. Sodium imaging as a marker of tissue injury in patients
with multiple sclerosis. Mult. Scler. Relat. Disord. 2, 263–269 (2013).
69. Paling, D., Golay, X., Wheeler-Kingshott, C., Kapoor, R. & Miller, D. Energy failure in multiple sclerosis and its investigation using
MR techniques. J. Neurol. 258, 2113–2127 (2011).
70. Petracca, M., Fleysher, L., Oesingmann, N. & Inglese, M. Sodium MRI of multiple sclerosis. NMR Biomed. 29, 153–161 (2016).
71. Qian, Y., Zhao, T., Zheng, H., Weimer, J. & Boada, F. E. High-resolution sodium imaging of human brain at 7 T. Magn. Reson. Med.
68, 227–233 (2012).
72. Riemer, F. et al. Measuring tissue sodium concentration: Cross-vendor repeatability and reproducibility of 23 Na-MRI across two
sites. J. Magn. Reson. Imaging JMRI 50, 1278–1284 (2019).
73. Mirkes, C. C., Homann, J., Shajan, G., Pohmann, R. & Scheer, K. High-resolution quantitative sodium imaging at 9.4 Tesla.
Magn. Reson. Med. 73, 342–351 (2015).
74. Viechtbauer, W. Conducting meta-analyses in R with the metafor package. J. Stat. Sow. 36, 1–48 (2010).
75. Viechtbauer, W. Bias and eciency of meta-analytic variance estimators in the random-eects model. J. Educ. Behav. Stat. 30,
261–293 (2005).
76. Knapp, G. & Hartung, J. Improved tests for a random eects meta-regression with a single covariate. Stat. Med. 22, 2693–2710
(2003).
77. Pustejovsky, J. E. & Tipton, E. Meta-analysis with robust variance estimation: Expanding the range of working models. Prev. Sci.
23, 425–438 (2022).
78. Pustejovsky, J. clubSandwich: Cluster-Robust (Sandwich) Variance Estimators with Small-Sample Corrections. (2022). https://
CRAN.R- proje ct. org/ packa ge= clubS andwi ch.
79. Egger, M., Davey Smith, G., Schneider, M. & Minder, C. Bias in meta-analysis detected by a simple, graphical test. BMJ 315, 629–634
(1997).
80. Springer, C. S. Biological systems: Spin-3/2 nuclei. In eMagRes (American Cancer Society, 2007). https:// doi. org/ 10. 1002/ 97804
70034 590. emrst m0029.
81. Woessner, D. E. NMR relaxation of spin-3/2 nuclei: Eects of structure, order, and dynamics in aqueous heterogeneous systems.
Concepts Magn. Reson. 13, 294–325 (2001).
82. Eliav, U. & Navon, G. Sodium NMR/MRI for anisotropic systems. NMR Biomed. 29, 144–152 (2016).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
14
Vol:.(1234567890)
Scientic Reports | (2023) 13:3222 | https://doi.org/10.1038/s41598-023-30363-y
www.nature.com/scientificreports/
83. Haddaway, N. R., Page, M. J., Pritchard, C. C. & McGuinness, L. A. PRISMA2020: An R package and Shiny app for producing
PRISMA 2020-compliant ow diagrams, with interactivity for optimised digital transparency and Open Synthesis. Campbell Syst.
Rev. 18, e1230 (2021).
84. R Core Team. R: A language and environment for statistical computing. (2021).
85. Wickham, H. et al. Welcome to the Tidyverse. J. Open Source Sow. 4, 1686 (2019).
86. Wickham, H. et al. ggplot2: Create Elegant Data Visualisations Using the Grammar of Graphics. (2021).
87. ulborn, K. R., Davis, D., Snyder, J., Yonas, H. & Kassam, A. Sodium MR imaging of acute and subacute stroke for assessment of
tissue viability. Neuroimaging Clin. N. Am. 15(639–653), xi–xii (2005).
88. Niesporek, S. C. et al. Partial volume correction for invivo (23)Na-MRI data of the human brain. Neuroimage 112, 353–363 (2015).
89. Driver, I. D., Stobbe, R. W., Wise, R. G. & Beaulieu, C. Venous contribution to sodium MRI in the human brain. Magn. Reson.
Med. 83, 1331–1338 (2020).
90. Liao, Y. et al. Correlation of quantitative conductivity mapping and total tissue sodium concentration at 3T/4T. Magn. Reson. Med.
82, 1518–1526 (2019).
91. Meyer, M. M. et al. Repeatability and reproducibility of cerebral 23Na imaging in healthy subjects. BMC Med. Imaging 19, 26
(2019).
92. Gerhalter, T. et al. Global decrease in brain sodium concentration aer mild traumatic brain injury. Brain Commun. 3, fcabo51
(2021).
Author contributions
B.R. conceived and planned the work, designed, and ran the literature search. B.R., W.Z. and F.N. contributed
to data extraction. B.R. and F.M. designed and ran the analysis and prepared gures. All authors contributed to
interpretation of the results. B.R. draed the manuscript, which was revised by all authors, who agreed to the
nal version.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary Information e online version contains supplementary material available at https:// doi. org/
10. 1038/ s41598- 023- 30363-y.
Correspondence and requests for materials should be addressed to B.R.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access is article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. e images or other third party material in this
article are included in the articles Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http:// creat iveco mmons. org/ licen ses/ by/4. 0/.
© e Author(s) 2023
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Such heterogeneity has been reported in other studies analyzing brain myelination, 39 iron 40 and water 29 content, and sodium concentration. 41 Since water content and sodium concentration are the main factors that determine electrical properties, we anticipate that variation in the former would be reflected in the latter. For example, for deep GM ROIs such as the caudate nucleus, putamen, and thalamus, which consist of GM structures with embedded myelinated axons, we found intermediate conductivity and relative permittivity between WM and GM theoretical predictions. ...
... This is consistent with previous reports for water content 29 and sodium concentration. 41 Conventional phase-only HH-EPT acquisition also provided conductivity maps of the tissues. Measured conductivity T A B L E 1 Average across subject population of median EPT values within WM/GM/CSF masks obtained with bSSFP-based HH-EPT (conductivity only) and QTI-based wEPT (conductivity and relative permittivity), in comparison with their theoretical predictions based on a fourth order Cole-Cole model. ...
Article
Full-text available
In this study, we aimed to develop a fast and robust high‐resolution technique for clinically feasible electrical properties tomography based on water content maps (wEPT) using Quantitative Transient‐state Imaging (QTI), a multiparametric transient state‐based method that is similar to MR fingerprinting. Compared with the original wEPT implementation based on standard spin‐echo acquisition, QTI provides robust electrical properties quantification towards B 1 ⁺ inhomogeneities and full quantitative relaxometry data. To validate the proposed approach, 3D QTI data of 12 healthy volunteers were acquired on a 1.5 T scanner. QTI‐provided T 1 maps were used to compute water content maps of the tissues using an empirical relationship based on literature ex‐vivo measurements. Assuming that electrical properties are modulated mainly by tissue water content, the water content maps were used to derive electrical conductivity and relative permittivity maps. The proposed technique was compared with a conventional phase‐only Helmholtz EPT (HH‐EPT) acquisition both within whole white matter, gray matter, and cerebrospinal fluid masks, and within different white and gray matter subregions. In addition, QTI‐based wEPT was retrospectively applied to four multiple sclerosis adolescent and adult patients, compared with conventional contrast‐weighted imaging in terms of lesion delineation, and quantitatively assessed by measuring the variation of electrical properties in lesions. Results obtained with the proposed approach agreed well with theoretical predictions and previous in vivo findings in both white and gray matter. The reconstructed maps showed greater anatomical detail and lower variability compared with standard phase‐only HH‐EPT. The technique can potentially improve delineation of pathology when compared with conventional contrast‐weighted imaging and was able to detect significant variations in lesions with respect to normal‐appearing tissues. In conclusion, QTI can reliably measure conductivity and relative permittivity of brain tissues within a short scan time, opening the way to the study of electric properties in clinical settings.
Article
Full-text available
Whole brain ionic and metabolic imaging has potential as a powerful tool for the characterization of brain diseases. We combined sodium MRI (23 Na MRI) and 1 H-MR Spectroscopic Imaging (1 H-MRSI), assessing changes within epileptogenic networks in comparison with electrophysiologically normal networks as defined by stereotactic EEG (SEEG) recordings analysis. We applied a multi-echo density adapted 3D projection reconstruction pulse sequence at 7 T (23 Na-MRI) and a 3D echo-planar spectroscopic imaging sequence at 3 T (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 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 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.
Article
Full-text available
Proton MRI can provide detailed morphological images, but it reveals little information about cell homeostasis. On the other hand, sodium MRI can provide metabolic information but cannot resolve fine structures. The complementary nature of proton and sodium MRI raises the prospect of their combined use in a single experiment. In this work, we assessed the repeatability of normalized proton density (PD), T1, T2, and normalized sodium density-weighted quantification measured with simultaneous 3D 1H MRF/23Na MRI in the brain at 7 T, from ten healthy volunteers who were scanned three times each. The coefficients of variation (CV) and the intra-class correlation (ICC) were calculated for the mean and standard deviation (SD) of these 4 parameters in grey matter, white matter, and cerebrospinal fluid. As result, the CVs were lower than 3.3% for the mean values and lower than 6.9% for the SD values. The ICCs were higher than 0.61 in all 24 measurements. We conclude that the measurements of normalized PD, T1, T2, and normalized sodium density-weighted from simultaneous 3D 1H MRF/23Na MRI in the brain at 7 T showed high repeatability. We estimate that changes > 6.6% (> 2 CVs) in mean values of both 1H and 23Na metrics could be detectable with this method.
Article
Full-text available
Background Reporting standards, such as PRISMA aim to ensure that the methods and results of systematic reviews are described in sufficient detail to allow full transparency. Flow diagrams in evidence syntheses allow the reader to rapidly understand the core procedures used in a review and examine the attrition of irrelevant records throughout the review process. Recent research suggests that use of flow diagrams in systematic reviews is poor and of low quality and called for standardised templates to facilitate better reporting in flow diagrams. The increasing options for interactivity provided by the Internet gives us an opportunity to support easy‐to‐use evidence synthesis tools, and here we report on the development of a tool for the production of PRISMA 2020‐compliant systematic review flow diagrams. Methods and Findings We developed a free‐to‐use, Open Source R package and web‐based Shiny app to allow users to design PRISMA flow diagrams for their own systematic reviews. Our tool allows users to produce standardised visualisations that transparently document the methods and results of a systematic review process in a variety of formats. In addition, we provide the opportunity to produce interactive, web‐based flow diagrams (exported as HTML files), that allow readers to click on boxes of the diagram and navigate to further details on methods, results or data files. We provide an interactive example here; https://prisma-flowdiagram.github.io/. Conclusions We have developed a user‐friendly tool for producing PRISMA 2020‐compliant flow diagrams for users with coding experience and, importantly, for users without prior experience in coding by making use of Shiny (https://estech.shinyapps.io/prisma_flowdiagram/). This free‐to‐use tool will make it easier to produce clear and PRISMA 2020‐compliant systematic review flow diagrams. Significantly, users can also produce interactive flow diagrams for the first time, allowing readers of their reviews to smoothly and swiftly explore and navigate to further details of the methods and results of a review. We believe this tool will increase use of PRISMA flow diagrams, improve the compliance and quality of flow diagrams, and facilitate strong science communication of the methods and results of systematic reviews by making use of interactivity. We encourage the systematic review community to make use of the tool, and provide feedback to streamline and improve their usability and efficiency.
Article
Full-text available
Objectives To characterize subcortical nuclei by multi-parametric quantitative magnetic resonance imaging. Materials and Methods: The following quantitative multiparametric MR data of five healthy volunteers were acquired on a 7T MRI system: 3D gradient echo (GRE) data for the calculation of quantitative susceptibility maps (QSM), GRE sequences with and without off-resonant magnetic transfer pulse for magnetization transfer ratio (MTR) calculation, a magnetization−prepared 2 rapid acquisition gradient echo sequence for T1 mapping, and (after a coil change) a density-adapted 3D radial pulse sequence for ²³Na imaging. First, all data were co-registered to the GRE data, volumes of interest (VOIs) for 21 subcortical structures were drawn manually for each volunteer, and a combined voxel-wise analysis of the four MR contrasts (QSM, MTR, T1, ²³Na) in each structure was conducted to assess the quantitative, MR value-based differentiability of structures. Second, a machine learning algorithm based on random forests was trained to automatically classify the groups of multi-parametric voxel values from each VOI according to their association to one of the 21 subcortical structures. Results The analysis of the integrated multimodal visualization of quantitative MR values in each structure yielded a successful classification among nuclei of the ascending reticular activation system (ARAS), the limbic system and the extrapyramidal system, while classification among (epi-)thalamic nuclei was less successful. The machine learning-based approach facilitated quantitative MR value-based structure classification especially in the group of extrapyramidal nuclei and reached an overall accuracy of 85% regarding all selected nuclei. Conclusion Multimodal quantitative MR enabled excellent differentiation of a wide spectrum of subcortical nuclei with reasonable accuracy and may thus enable sensitive detection of disease and nucleus-specific MR-based contrast alterations in the future.
Article
Full-text available
The pathological cascade of tissue damage in mild traumatic brain injury is set forth by a perturbation in ionic homeostasis. However, whether this class of injury can be detected in vivo and serve as a surrogate marker of clinical outcome is unknown. We employ sodium MRI to test the hypotheses that regional and global total sodium concentrations: (i) are higher in patients than in controls and (ii) correlate with clinical presentation and neuropsychological function. Given the novelty of sodium imaging in traumatic brain injury, effect sizes from (i), and correlation types and strength from (ii), were compared to those obtained using standard diffusion imaging metrics. Twenty-seven patients (20 female, age 35.9 ± 12.2 years) within 2 months after injury and 19 controls were scanned with proton and sodium MRI at 3 Tesla. Total sodium concentration, fractional anisotropy and apparent diffusion coefficient were obtained with voxel averaging across 12 grey and white matter regions. Linear regression was used to obtain global grey and white matter total sodium concentrations. Patient outcome was assessed with global functioning, symptom profiles and neuropsychological function assessments. In the regional analysis, there were no statistically significant differences between patients and controls in apparent diffusion coefficient, while differences in sodium concentration and fractional anisotropy were found only in single regions. However, for each of the 12 regions, sodium concentration effect sizes were uni-directional, due to lower mean sodium concentration in patients compared to controls. Consequently, linear regression analysis found statistically significant lower global grey and white matter sodium concentrations in patients compared to controls. The strongest correlation with outcome was between global grey matter sodium concentration and the composite z-score from the neuropsychological testing. In conclusion, both sodium concentration and diffusion showed poor utility in differentiating patients from controls, and weak correlations with clinical presentation, when using a region-based approach. In contrast, sodium linear regression, capitalizing on partial volume correction and high sensitivity to global changes, revealed high effect sizes and associations with patient outcome. This suggests that well-recognized sodium imbalances in traumatic brain injury are (i) detectable non-invasively; (ii) non-focal; (iii) occur even when the antecedent injury is clinically mild. Finally, in contrast to our principle hypothesis, patients’ sodium concentrations were lower than controls, indicating that the biological effect of traumatic brain injury on the sodium homeostasis may differ from that in other neurological disorders. Note: This figure has been annotated.
Article
Full-text available
Background/aim: Sodium (23Na) MR imaging is a noninvasive MRI technique that has been shown to be sensitive to visualize biochemical information about tissue viability, their cell integrity, and cell function in various studies. The aim of this study was to evaluate differences in regional brain 23Na signal intensity between Alzheimer's disease (AD) and healthy controls to preliminarily evaluate the capability of 23Na imaging as a biomarker for AD. Patients and methods: A total of 14 patients diagnosed with AD were included: 12 in the state of dementia and 2 with mild cognitive impairment (MCI), and 12 healthy controls (HC); they were all scanned on a 3T clinical scanner with a double tuned 1H/23Na birdcage head coil. After normalizing the signal intensity with that of the vitreous humor, relative tissue sodium concentration (rTSC) was measured after automated segmentation in the hippocampus, amygdala, basal ganglia, white matter (WM) and grey matter (GM) in both cerebral hemispheres. Results: Patients with AD showed a significant increase in rTSC in comparison to healthy controls in the following brain regions: WM 13.6%; p=0.007, hippocampus 12.9%; p=0.003, amygdala 18.9%; p=0.0007. Conclusion: 23Na-MRI has the potential to be developed as a useful biomarker for the diagnosis of AD.
Article
Background and purpose: While conventional MR imaging has limited value in amyotrophic lateral sclerosis, nonconventional MR imaging has shown alterations of microstructure using diffusion MR imaging and recently sodium homeostasis with sodium MR imaging. We aimed to investigate the topography of brain regions showing combined microstructural and sodium homeostasis alterations in amyotrophic lateral sclerosis subgroups according to their disease-progression rates. Materials and methods: Twenty-nine patients with amyotrophic lateral sclerosis and 24 age-matched healthy controls were recruited. Clinical assessments included disease duration and the Revised Amyotrophic Lateral Sclerosis Functional Rating Scale. Patients were clinically differentiated into fast (n = 13) and slow (n = 16) progressors according to the Revised Amyotrophic Lateral Sclerosis Functional Rating Scale progression rate. 3T MR imaging brain protocol included 1H T1-weighted and diffusion sequences and a 23Na density-adapted radial sequence. Quantitative maps of diffusion with fractional anisotropy, mean diffusivity, and total sodium concentration were measured. The topography of diffusion and sodium abnormalities was assessed by voxelwise analyses. Results: Patients with amyotrophic lateral sclerosis showed significantly higher sodium concentrations and lower fractional anisotropy, along with higher sodium concentrations and higher mean diffusivity compared with healthy controls, primarily within the corticospinal tracts, corona radiata, and body and genu of the corpus callosum. Fast progressors showed wider-spread abnormalities mainly in the frontal areas. In slow progressors, only fractional anisotropy measures showed abnormalities compared with healthy controls, localized in focal regions of the corticospinal tracts, the body of corpus callosum, corona radiata, and thalamic radiation. Conclusions: The present study evidenced widespread combined microstructural and sodium homeostasis brain alterations in fast amyotrophic lateral sclerosis progressors.
Article
The cell volume fraction (CVF) of the human brain is high (~ 82%) and preserved across healthy aging while the brain declines in volume. These two observations, supported by several independent techniques, suggest that cell volume fraction is an important structural parameter. A new biophysical model is presented that incorporates CVF into the Goldman equation of classical membrane electrophysiology. The Goldman equation contains few structural constraints beyond two compartments separated by a semipermeable membrane supporting ion gradients. As potassium is the most permeable ion in the resting state, the resting membrane potential is determined by the potassium ion gradient. This biophysical model indicates that the sodium-potassium ion pumps use less energy at high CVF to maintain the resting membrane potential, explaining the high value of CVF and its conservation with healthy aging. Cell volume fraction is measured to be statistically significantly higher in the brains of males compared to females, suggesting a structural requirement for higher energy efficiency in the larger male brain to support the greater number of neurons and synapses. As CVF can be measured in humans using quantitative sodium MR imaging and has potential implications for brain health, CVF may be a quantitative parameter that is useful for assessment of brain health, especially in patients with diseases such as dementia and psychiatric disease that do not have anatomical correlates detectable by clinical proton MR imaging.
Article
In prevention science and related fields, large meta-analyses are common, and these analyses often involve dependent effect size estimates. Robust variance estimation (RVE) methods provide a way to include all dependent effect sizes in a single meta-regression model, even when the exact form of the dependence is unknown. RVE uses a working model of the dependence structure, but the two currently available working models are limited to each describing a single type of dependence. Drawing on flexible tools from multilevel and multivariate meta-analysis, this paper describes an expanded range of working models, along with accompanying estimation methods, which offer potential benefits in terms of better capturing the types of data structures that occur in practice and, under some circumstances, improving the efficiency of meta-regression estimates. We describe how the methods can be implemented using existing software (the "metafor" and "clubSandwich" packages for R), illustrate the proposed approach in a meta-analysis of randomized trials on the effects of brief alcohol interventions for adolescents and young adults, and report findings from a simulation study evaluating the performance of the new methods.