ArticlePDF Available

The Energetic brain -Students to student review

Authors:
Accepted Article
This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1111/jnc.14829
This article is protected by copyright. All rights reserved.
DR. MOOTAZ M. SALMAN (Orcid ID : 0000-0002-5683-1706)
DR. ARTEM V DIUBA (Orcid ID : 0000-0002-2885-4281)
MR. EMIL JAKOBSEN (Orcid ID : 0000-0001-6290-4222)
DR. ASHLEY P L MARSH (Orcid ID : 0000-0001-6049-6931)
DR. CONSTANZE I SEIDENBECHER (Orcid ID : 0000-0002-7433-2716)
Article type : Review
The energetic brain - A review from students to students
M.P. Bordone; Universidad de Buenos Aires, Facultad de Farmacia y Bioquímica, Instituto de
Investigaciones Farmacológicas (ININFA), Buenos Aires, Argentina.
M.M. Salman; Department of Cell Biology, Harvard Medical School, and Program in Cellular and Molecular
Medicine, Boston Children’s Hospital, Boston, MA, USA
H.E. Titus; Northwestern University, Feinberg School of Medicine, Chicago, IL, USA
E. Amini; Department of Medicine, University Kebangsaan Malaysia Medical Centre (HUKM), Cheras,
Kuala Lumpur, Malaysia
J. V. Andersen; Department of Drug Design and Pharmacology, Faculty of Health and Medical Sciences,
University of Copenhagen, Denmark.
B Chakraborti; Manovikas Kendra, Kolkata, India
A.V. Diuba; Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University,
119992, Moscow, Russia
T. G. Dubouskaya; Institute of Biophysics and Cell Engineering National Academy of Sciences of Belarus,
Minsk, Belarus
E. Ehrke; Centre for biomolecular interactions, University of Bremen, Germany
A.E. Freitas; Neurobiology Section, Biological Sciences Division, University of California, San Diego, La
Jolla, California, USA
G.B. Freitas; institute of medical biochemistry, Federal University of Rio de Janeiro, Brazil
R.A. Gonçalves; Centre for neuroscience studies, Queen's University, Kingston, ON, Canada
D. Gupta; BITS Pilani, Pilani, India
R. Gupta; CSIR- Indian Institute of Toxicology Research, Lucknow, India
S.R. Ha; Baylor College of Medicine, Houston, TX, USA
I.A. Hemming; The Harry Perkins Institute of Medical Research, Brain Growth and Disease Laboratory,
Nedlands, Australia 2 The University of Western Australia, School of Medicine and Pharmacology,
Crawley, Australia
M. Jaggar; Department of biological sciences, Tata Institute of Fundamental Research, Mumbai, India
E. Jakobsen; Department of Drug Design and Pharmacology, Faculty of Health and Medical Sciences,
University of Copenhagen, Denmark
P. Kumari; Defense Institute of Physiology and allied sciences, Defense research and development
organization, Lucknow Timarpur, Delhi, India
N. Lakkappa; Department of Pharmacology, JSS college of Pharmacy, Ooty, India
A.P.L Marsh; Bruce Lefroy Centre for Genetic Health Research, Murdoch Children's Research Institute,
Royal Children’s Hospital, Parkville, Victoria, Australia.
J. Mitlöhner; Department of Neurochemistry and Molecular Biology, Leibniz Institute for Neurobiology
Magdeburg, Germany
Y. Ogawa; The Jikei University School of Medicine, Japan
R.K. Paidi; CSIR-Indian Institute of Chemical Biology, Kolkata, India
F.C. Ribeiro; Federal University of Rio de Janeiro, Brazil
A. Salamian; Department of Molecular and Cellular Neurobiology, Nencki Institute of Experimental Biology,
Polish Academy of Sciences, Warsaw, Poland
S.Saleem; CSIR Indian Institute for Chemical Biology, India
Accepted Article
This article is protected by copyright. All rights reserved.
S. Sharma; Department of Pharmacy, Birla Institute of Technology and Science, Pilani, Rajasthan, India
J.M. Silva; Life and Health Sciences Research Institute (ICVS), Medical School, University of Minho,
Campus Gualtar, 4710-057 Braga, Portugal
S. Singh; CSIR-Indian Institute of Toxicology Research, India
K. Sulakhiya; Department of Pharmacy, Indira Gandhi National Tribal University, Amarkantak, India
T.W. Tefera; School of Biomedical Sciences, The University of Queensland, Brisbane, Australia
B. Vafadari; Institute of environmental medicine, UNIKA-T, Technical University of Munich, Germany
A. Yadav; CSIR- Indian Institute of Toxicology Research, India
R. Yamazaki; Tokyo University of Pharmacy and Life Sciences, Tokyo, Japan
C. I. Seidenbecher; Department of Neurochemistry and Molecular Biology, Leibniz Institute for
Neurobiology Magdeburg, Germany
Corresponding author: Prof. Dr. Constanze I. Seidenbecher
Department of Neurochemistry and Molecular Biology
Leibniz Institute for Neurobiology Magdeburg
Brenneckestraße 6, 39118 Magdeburg
E-mail: seidenc@lin-magdeburg.de
Phone: +49-391-6263-92401
Keywords: neuronal energetic cost, metabolism, energy sources, neurometabolic coupling,
energy homeostasis, ANLS hypothesis, synapse.
Abbreviations used:
Aβ amyloid-beta
AD Alzheimer’s disease
ADP adenosine diphosphate
AGE advanced glycation end-product
AIS axon initial segment
ALS amyotrophic lateral sclerosis
AMP adenosine monophosphate
AMPA α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid
AMPK adenosine monophosphate-activated kinase
ANLS astrocyte-neuron lactate shuttle
AP action potential
APP amyloid precursor protein
ATP adenosine triphosphate
ATPmit mitochondrial adenosine triphosphate
ATPpresyn presynaptic adenosine triphosphate
BBB blood brain barrier
BDNF Brain-derived neurotrophic factor
CBF cerebral blood flow
CMRglu cerebral metabolic rate of glucose
CNS central nervous system
CoA coenzyme A
DA dopaminergic
DJ1/PARK7 protein deglycase DJ-1/Parkinson disease protein 7
Drp1 dynamin-related protein 1
EAAT 1/2 excitatory amino acid transporter 1/2
EEG electroencephalography
ER endoplasmic reticulum
Accepted Article
This article is protected by copyright. All rights reserved.
ETC electron transport chain
FDG fluorodeoxyglucose
FRET fluorescence resonance energy transfer
G6P glucose-6-phosphate
GAPDH glyceraldehyde-3-phosphate dehydrogenase
GDH glutamate dehydrogenase
GDNF glial cell line-derived neurotrophic factor
GLS glutaminase
GLUT glutamate transporter
GPR120 G protein-coupled receptor 120
GS glutamine synthase
GSH glutathione
GSSG glutathione disulfide
GTP guanosine triphosphate
H-H model Hodgkin-Huxley model
IGF-1 insulin-like growth factor-1
α-KG α-ketoglutarate
Km Michaelis-Menten constant
LDH lactate dehydrogenase
LPL lipoprotein lipase
LRRK2 leucine-rich repeat kinase 2
LTP long-term potentiation
MAS malate-aspartate shuttle
MCH melanin-concentrating hormone
MCT monocarboxylic acid transporter
MCTG medium chain triglyceride
Mfn1/2 Mitofusin 1/2
mRNA messenger ribonucleic acid
MS multiple sclerosis
mtDNA mitochondrial deoxyribonucleic acid
NAD+ oxidized nicotinamide adenine dinucleotide
NADPH reduced nicotinamide adenine dinucleotide phosphate
NBCe1 electrogenic sodium bicarbonate cotransporter 1
NE norepinephrine
NMDA N-methyl-D-aspartate
NMR nuclear magnetic resonance
NO nitric oxide
β-OHB β-hydroxybutyrate
OPA1 optic atrophy type 1
OxPhos oxidative phosphorylation
PC pyruvate carboxylase
PD Parkinon’s disease
PDGFRβ platelet-derived growth factor receptor β
PDH pyruvate dehydrogenase
PET positron emission tomography
PFK phosphofructokinase
PGC-1α peroxisome proliferator-activated receptor gamma coactivator 1-alpha
Accepted Article
This article is protected by copyright. All rights reserved.
PINK1 PTEN-induced kinase 1
PPAR peroxisome proliferator-activated receptor
PPP pentose phosphate pathway
PSC postsynaptic current
REM rapid eye movement
ROS reactive oxygen species
SIRT1 sirtuin-1
SLC1A5 neutral amino acid transporter B(0)
SLC6A8 sodium- and chloride-dependent creatine transporter 1
SLC38A1 sodium-coupled neutral amino acid transporter 1
SLC family solute carrier family
SNpc substantia nigra pars compacta
SPECT single-photon emission computed tomography
SWS slow wave sleep
T2DM type 2 diabetes mellitus
TBI traumatic brain injury
TCA tricarboxylic acid
TGF-β transforming growth factor β
VIP vasoactive intestinal polypeptide
Vmax maximum rate in Michaelis-Menten kinetics
Abstract
The past 20 years have resulted in unprecedented progress in understanding brain energy
metabolism and its role in health and disease. In this review, which was initiated at the 14th
International Society for Neurochemistry Advanced School, we address the basic concepts
of brain energy metabolism and approach the question of why the brain has high energy
expenditure. Our review illustrates that the vertebrate brain has a high need for energy
because of the high number of neurons and the need to maintain a delicate interplay
between energy metabolism, neurotransmission, and plasticity. Disturbances to the
energetic balance, to mitochondria quality control or to glia-neuron metabolic interaction may
lead to brain circuit malfunction or even severe disorders of the central nervous system
(CNS). We cover neuronal energy consumption in neural transmission and basic
(‘housekeeping’) cellular processes. Additionally, we describe the most common (glucose)
and alternative sources of energy namely glutamate, lactate, ketone bodies and medium
chain fatty acids. We discuss the multifaceted role of non-neuronal cells in the transport of
energy substrates from circulation (pericytes and astrocytes) and in the supply (astrocytes
and microglia) and usage of different energy fuels. Finally, we address pathological
consequences of disrupted energy homeostasis in the CNS.
1. Why does the brain have high energy expenditure?
The brain is a highly active and plastic organ with outstanding energetic needs. Thus, the
title of this review shall reflect both, the dynamic and adaptive nature of the brain with its
intense parallel information processing as well as its specific metabolic energy demands.
Accepted Article
This article is protected by copyright. All rights reserved.
In a majority of adult vertebrate species (excluding primates), the central nervous system
(CNS) uses 2-8% of the energy of total basal metabolism (Mink et al. 1981). The human
brain, however, accounts for 20% of oxygen (O2) consumption and 25% of glucose
utilization, although it amounts to only 2% of body weight (Sokoloff et al. 1977). Is the high
energetic expense of the human brain related to the higher cognitive abilities and wider
range of behaviors expressed by humans as compared to other vertebrates? Is it a
consequence of brain growth during evolution? During humanization, multifaceted social
behaviors evolved, such as formation of complex social groups, long-term parental
investment and cooperative foraging strategies. These outstanding social skills correlated
with increased intelligence and are highlighted as necessary for more effective foraging and
the exploitation of high-energy food, but at the same time the evolvement of these
capabilities also imposes higher nutrient requirements (summarized in Dunbar & Shulz,
2017). Traditionally, comparative studies of brain scaling take into account brain size and/or
the body-brain mass ratio to delineate an evolutionary explanation for the supposed human
brain exceptionality. While some authors consider the human brain as an outlier because it
deviates from the expected value even if compared to anthropoid primates (Jerison 2012;
Marino 1998), a more recent view, based on data obtained with isotropic fractionation
(Herculano-Houzel & Lent 2005), puts in focus the absolute number of neurons relative to
body-brain mass ratios. Studies on scaling brain metabolism according to brain size across
species or to neuronal number and/or density in a given structure were carried out to
examine this human brain peculiarity which could have implications in brain evolution and
could have exerted constraints for wiring patterns. According to Karbowski´s estimates,
which are based on the assumption of uniform scaling of neuronal density across species,
cerebral energy per neuron increases with brain size (Karbowski 2009). However, more
recently Herculano-Hozuel based her calculations on available glucose and O2 metabolic
rates in awake animals (mouse, rat, squirrel, macaque monkey, baboon and human) and the
total number of neurons determined by her group. She found that neuronal density in the
whole brain does not scale with brain mass across species, and that the energy budget of
the whole brain per neuron is fixed across species and brain sizes. Thus, the total glucose
use by the brain is a linear function of the number of neurons, and the remarkable energy
use in humans may be explained simply by its large number of neurons (Herculano-Houzel
2011).
Why do neurons have a high energetic demand? In addition to basic (‘housekeeping’)
cellular activity such as turnover of macromolecules, axoplasmic transport and mitochondrial
proton leak, neurons are highly specialized cells to perform energy-demanding
electrochemical signaling processes. Generation of action potentials (AP), postsynaptic ion
fluxes, presynaptic Ca2+ entry, transmitter reuptake and vesicle cycling but also maintenance
of the resting potential are costly cellular neuronal functions. In the following sections we will
provide a detailed description and explanation of each of the above-mentioned processes,
which altogether account for the neuronal energy demand.
1.1 Energy use in synaptic transmission and synaptic plasticity
Synaptic transmission imposes a large metabolic demand which is met through activity-
driven regulation of glycolysis and mitochondrial function (Rangaraju et al. 2014). Plasticity
of synapses can be mediated by changes in Ca2+ concentration and/or the number of
neurotransmitter receptors. Upon synaptic activity or plasticity, most of the adenosine
Accepted Article
This article is protected by copyright. All rights reserved.
triphosphate (ATP) is consumed when pumping ions to maintain resting membrane potential,
vesicle filling, vesicle transport, vesicle recycling and enzymatic processing of synaptic
transmitter within synapses (Harris et al. 2012) (Figure 1).
1.1.1 Presynaptic terminals
Activated presynaptic terminals are expected to place high ATP demands on energy
supplies. The presynaptic ATP (ATPpresyn) levels are mainly consumed by the Na+/K+
pump (Na+/K+-ATPase), the Ca2-ATPases in the plasma membrane and endoplasmic
reticulum (ER), the vacuolar H+-ATPase, motor proteins (Attwell & Laughlin 2001; Lennie
2003), and protein disassembly machineries (Rangaraju et al. 2014; Ly & Verstreken 2006).
The Na+/K+-ATPase imports 2 K+ and exports 3 Na+ ions involved in generating the AP and
powers Ca2+ removal by Na+/Ca2+ exchange, while Ca2+-ATPases in the plasma membrane
and ER reduce the elevated cytosolic Ca2+ concentration after membrane depolarization.
Vacuolar H+-ATPase energizes vesicular transmitter uptake and the motor proteins are
involved in intracellular transport of mitochondria and vesicles (Figure 1).
ATPpresyn levels are reduced by insufficiency of either glycolysis or mitochondrial function,
indicating the requirement of activity-driven ATP synthesis to meet the energy demands of
synaptic function. Most ATP produced in response to increased neuronal activity is
generated by mitochondria (Lin et al. 2010; Hall et al. 2012), highlighted by the positive
correlation between the signaling-related energy usage predicted by Attwell and Laughlin
(2001) and the distribution of mitochondria inside a cell, which was demonstrated to be
higher in soma plus dendrites (62%) than in axon terminals (36%) (Wong-Riley 1989). In
light of the distal location of nerve terminals from cell bodies, synapses must rely on local
ATP supplies, leaving them highly susceptible to mitochondriopathies. Studies in Drosophila
mutants affecting the mitochondrial localization as well as the mitochondrial ATP/ADP
translocase, confirmed the importance of mitochondria in synaptic function (Trotta et al.
2004; Guo et al. 2005; Verstreken et al. 2005).
ATPpresyn concentration is about 1.4 mM (corresponding to ~ 106 molecules for a typical
presynaptic varicosity) and ATP synthesis occurs through feedforward stimulation of both
glycolysis and oxidative phosphorylation (OxPhos) via electrical activity-driven Ca2+ influx
(Rangaraju et al. 2014). Completely blocking exocytosis does not significantly affect Ca2+
influx during neuronal APs, suggesting that the synaptic vesicle cycle consumes most
ATPpresyn. The monitoring of ATPpresyn during AP firing in the absence of external Ca2+
revealed that ATP utilized by the Na+/K+-ATPase alone represents a relatively small energy
burden compared to downstream vesicle cycle processes at presynaptic terminals and that
AP firing can persist even during phases of acute ATP synthesis blockade. Inhibition of
glycolysis is accompanied by a reduction in endocytosis and a shift in remaining vesicular
pH to more alkaline values. The arrest of endocytosis at energy deficit conditions depends
on the mechanochemical enzyme dynamin, which mediates membrane fission and depends
cooperatively on synthesized GTP using ATP (Rangaraju et al. 2014). By contrast,
compromised ATP synthesis does not immediately impact exocytosis. Therefore, ATPpresyn
demand and activity-induced ATP synthesis should be synchronous and are essential for the
maintenance of normal synaptic plasticity function. Any discrepancy or pathological condition
disturbing these metabolic functions will ultimately affect synaptic strength and plasticity.
Accepted Article
This article is protected by copyright. All rights reserved.
1.1.2 Postsynaptic terminals
At postsynaptic sites, ATP is mainly used for counterbalancing the ion fluxes through
postsynaptic receptors and, to a lesser extent, on rebounding Ca2+ to intracellular stores and
on mitochondrial trafficking. In inhibitory synapses, less energy is utilized to reverse
postsynaptic Cl- fluxes as the chloride reversal potential is close to the resting potential
(Attwell & Laughlin 2001; Howarth et al. 2010; Howarth et al. 2012). The number of activated
receptors, channel open time, channel conductance, and consequently ATP usage are
different at synapses throughout different regions of the brain (Silver et al. 1996; Spruston et
al. 1995; Markram et al. 1997). Based on analysis of Attwell and Laughlin (2001), when a
glutamatergic vesicle is released on a dendritic spine from non-cortical neurons, the number
of activated N-methyl-D-aspartate (NMDA) receptors is less than that of non-NMDA
receptors (Silver et al. 1996; Spruston et al. 1995; Markram et al. 1997). However, because
NMDA receptors have longer open times and higher conductance than non-NMDA
receptors, the ion influxes through those activated channels lead to hydrolysis of 5 % more
ATP molecules by the Na+/K+-ATPase, including the 3 Na+/Ca2+ exchange, for extrusion.
Moreover, activation of postsynaptic G protein-coupled receptors, such as metabotropic
glutamate receptors (mGluR), triggers downstream events, which constrain ATP usage.
However, the level of ATP consumption during G protein signaling is approximately 95 %
smaller than during NMDA and non-NMDA receptor activation (Attwell & Laughlin 2001).
1.1.3 Synaptic plasticity
Synaptic plasticity, increasing or decreasing of synaptic strength, can impact energy
expenditure. For example, long-term potentiation (LTP), which can be triggered by activation
of the postsynaptic NMDA receptors and insertion of more AMPA (α-amino-3-hydroxy-5-
methyl-4-isoxazolepropionic acid) receptors into the postsynaptic membrane, enhances ATP
supplement to potentiated synapses and energy consumption at postsynapses (Wieraszko
1982). It is suggested that induction of LTP, which is assumed as a fundamental mechanism
of learning and memory, is accompanied by increased energy usage at activated synapses.
Additionally, the level of lactate derived from glycogen in astrocytes is increased in response
to energy demand in animals performing learning and memory tasks (Suzuki et al. 2011;
Newman et al. 2011).
Mitochondrial ATP (ATPmit) production is noted to be essential for synaptic plasticity.
Inhibition of ATPmit production interfered with synaptic accumulation of mitochondria and
subsequently abolished synaptic potentiation at the Drosophila neuromuscular junction
(Tong 2007). Additionally, mutations in Drosophila Drp1 (dynamin-related protein 1) depleted
mitochondria from motor nerve terminals and interfered with mobilization of the reserve pool
of synaptic vesicles and maintenance of neurotransmission (Verstreken et al. 2005).
Moreover, transmission failure was rescued at high stimulation frequencies by adding ATP
exogenously and provided evidence that the reserve pool recruitment depends on ATPmit
production downstream of PKA (protein kinase A) signaling.
1.2. Action potential (AP) generation and propagation
The brain uses rapid electrical signaling in the form of APs as the primary means of
communication between neurons. This signaling is associated with substantial energetic
costs (Kole et al. 2008; Hallermann et al. 2012). Two sites, important for AP generation and
maintenance, due to their lowered excitation threshold, are the axon initial segment (AIS)
and the nodes of Ranvier (Figure 1), the sites between adjacent myelinated axonal
Accepted Article
This article is protected by copyright. All rights reserved.
segments (Palay et al. 1968; Peters 1966). It has been estimated that after one AP, rodent
cortical neurons require between ~ 4 8 x 108 ATP molecules to restore the Na+ and K+
gradient through the Na+/K+-ATPase, suggesting the metabolic cost of AP signaling is the
second largest after synaptic transmission (Lennie 2003; Hallermann et al. 2012). However,
energy usage within the brain depends in part on the AP rate. Moreover, because synaptic
energy cost is proportional to the transmitter release probability by an AP and to the number
of postsynaptic channels activated by transmitters, the estimated numbers for the synaptic
cost vary in the literature. The estimation of relative contributions of AP generation and
postsynaptic currents (PSC) to energy expenditure in the mammalian brain has varied in the
literature: while the classical Hodgkin-Huxley (H-H) model for AP generation in the squid
axon and a number of assumptions such as treating all neurons as identical (Attwell &
Laughlin 2001) led to an assumed energy expenditure ratio for AP and PSC of 58 % : 42 %,
direct experimental data in rat hippocampal non-myelinated mossy fibers showed a minimal
temporal overlap between the entry of sodium and the outflow of potassium during an AP
which results in one third less energy necessary to elicit depolarization than in the H-H
model, pointing to an AP : PSC ratio of 20 % : 80 % (Alle et al. 2009).
The approximately 50-fold higher densities of voltage-gated Na+ channels at AIS and nodes
of Ranvier, when compared to the soma and dendrites, are vital for producing the adequate
local current necessary to overcome membrane resistance and capacitance as well as to
initiate and propagate self-regenerating APs (Zhou et al. 1998; Kole et al. 2008). Moreover, it
has been shown that excess axonal Na+ influx at the AIS and nodes of Ranvier is critical for
AP conduction at high frequencies (Kole et al. 2008). Therefore, the high metabolic cost of
AP initiation by the AIS and propagation down the axon by nodes of Ranvier is seen as a
trade-off between minimizing energy costs and maximizing the conduction velocity of APs
(Hallermann et al. 2012).
1.3. Maintaining resting potential
Beyond spiking activity, a large portion of the energy sources in the brain is spent on
maintaining resting potential and non-signaling (‘housekeeping’) processes. Na+/K+-ATPases
are recruited to compensate non-zero cell membrane conductance for K+ and Na+. Atwell
and Laughlin (2001) calculated that at physiological spiking rate of 4 Hz, 15 % of the total
ATP in the grey matter is needed to maintain the resting potential of a typical neuron and an
associated glial cell (Attwell & Laughlin 2001). Lennie (2003) estimated that the spike-related
energy expenditure is only 13 % of total energy usage in human neocortex, attributing 28 %
and 10 % to maintain the resting potential in neurons and glia, respectively, and the
remaining to non-signaling (‘housekeeping’) processes (Lennie 2003). Maintaining resting
potential is estimated to take 30 % and ~ 44 % of the total energy used in partially and fully
myelinated white matter, respectively (Harris and Atwell, 2012), 34 % in the cerebellum
(Howarth et al. 2010), and 16 % to ~ 68 % of total energy used in the olfactory glomerulus,
depending on the number of activated olfactory receptor neurons (Nawroth et al. 2007).
1.4. Non-signaling (‘housekeeping’) processes
In the vertebrate brain, a majority of the total energy is spent on the removal of Na+ ions from
cells by Na+/K+-ATPases (Atwell and Laughlin, 2001; Lennie, 2003; Harris and Atwell, 2012).
Metabolic turnover is reduced by both, Na+/K+-ATPase inhibition (Whittam, 1961; Shibuki,
1989; Astrup et al., 1981) and deep anesthesia (Astrup et al. 1981; Sibson et al., 1998; Du
et al., 2008). The residual energy expenditure persisting after the suppression of spiking
Accepted Article
This article is protected by copyright. All rights reserved.
and/or Na+/K+-ATPase was attributed to non-signaling, or ‘housekeeping’, processes (Atwell
and Laughlin, 2001; Harris and Atwell, 2012; Engl and Atwell, 2015; Du et al., 2008). Using
13C NMR (nuclear magnetic resonance) spectroscopy, Sibson et al. (1998) estimated that
roughly 16 % of total glucose oxidation in the rat cortex are independent of synaptic activity.
The cerebral ATP metabolic rate is roughly reduced by half when rats are deeply
anesthetized and the electroencephalogram reads as “silent”, as compared with light
anesthesia using isoflurane (Du et al. 2008). Calculations for rodent brains attribute 25 % of
energy used in the grey matter to ‘housekeeping’ processes (Attwell and Laughlin 2001;
Lennie 2003; Harris et al. 2012), and ~ 63 % and ~ 56 % in partially or fully myelinated white
matter, respectively (Harris and Attwell 2012). The ‘housekeeping’ functions account for
19 % in the cerebellum (Howarth et al. 2010), and between 3 to 25 % in the olfactory
glomerulus (Nawroth et al. 2007). Particular non-signaling energy-consuming processes are
now under debate.
Actin cytoskeleton (re-)modeling underlies neuron morphology and modulates synaptic
function and structural plasticity (Luo 2002; Cingolani & Goda 2008), requiring ATP
hydrolysis (Wegner 1976; Carlier et al. 1988). Modeling and experimental estimates for
energy costs of actin treadmilling range from less than 1 % (Engl & Attwell 2015) of the total
brain energy usage to half of the energy used in neuronal culture (Bernstein & Bamburg
2003). The dynamic instability of microtubules instead requires energy in the form of GTP
(Margolis 1981; Zakharov et al. 2015).
Protein and phospholipid synthesis were estimated theoretically to account for no more than
2 % (Rolfe & Brown 1997; Attwell & Laughlin 2001) and 2 – 25 % (McKenna et al. 2012;
Purdon & Rapoport 2007; Purdon & Rapoport 1998) of the total ATP consumption in the
brain, respectively. Purdon and Rapoport (2007) estimated that ~ 26 % of the energy taken
by phospholipid metabolism in the rat brain is spent on fatty acid turnover,
phosphatidylinositol phosphorylation state and phospholipid bilayer asymmetry maintenance,
and on de novo synthesis of phosphatidylinositol and ether phospholipids (Purdon &
Rapoport 2007). Engl et al. (2017) provided first experimental single-assay measurements of
energy consumption by actin cycling, microtubule restructuring, and protein as well as
phospholipid metabolism (Engl et al. 2017). They applied specific blockers of these
processes to resting (i.e., without evoked signaling activity) hippocampal slices of developing
rat brains and traced the resulting changes in O2 consumption. O2 consumption decreased
by 25 % after actin cycling blockade (much higher than previously modelled by the same
group but lower than previous estimates), by 22 % after microtubule turnover blockade, and
by 18 % after lipid and protein metabolism blockade. However, blocking protein metabolism
alone did not change the O2 consumption significantly.
Proton leak across the inner mitochondrial membrane (IMM) uncoupled from ATP synthesis
was documented in isolated brain mitochondria (Rolfe et al. 1994); though, its contribution
was not quantified. The proton leak mechanism remains under debate. It may be attributed
to the proton conductance through the lipid bilayer, which depends on the fatty acid
composition of the mitochondrial membrane, to the non-specific proton conductance of IMM
carrier proteins such as adenine nucleotide translocase (Brand et al. 2005), or to the
functioning of the brain-specific uncoupling proteins UCP4 and UCP5 (Mao et al. 1999;
Sanchis et al. 1998). Deficits in mitophagy genes like in the PINK1 knockout mouse also
result in increased proton leak (Villeneuve et al. 2016).
Accepted Article
This article is protected by copyright. All rights reserved.
Finally, axonal transport contributes to ‘housekeeping’ energy expenditure (Harris et al.
2012; Lennie 2003) (Figure 1). Maday et al. (2014) calculated, as one kinesin-1 motor
spends one ATP molecule for every elementary 8 nm-step (Maday et al. 2014), the
anterograde transport of a vesicle along an average 40 mm axon would require ~ 5 x 106
ATP, which is small compared to the ~ 1 x 108 ATP consumed by the propagation of a single
spike along the same axon. However, this estimated cost of the axonal transport does not
account for tug-of-war events (Hendricks et al. 2010; Soppina et al. 2009) and dynein
retrograde transport. Further modeling is needed to estimate the actual energetic cost of
axonal transport processes.
The negligible gluconeogenesis activity of the brain raises the question of how the brain
uses energy in different physiological states and the importance of additional sources of
energy other than glucose.
2. Brain metabolism in sleep vs. wakefulness
Energy sources in the body are stored mainly in skeletal muscle, liver and adipose tissue
and maintain the reserve of energy during wakefulness (Brown & Ransom 2007). The
neurons of the lateral hypothalamus that express melanin-concentrating hormone (MCH)
and orexin/ hypocretin regulate body energy metabolism and the sleep-wakeful cycle. During
a high level of energy resources, MCH neurons are activated and promote conservation of
energy by inducing sleep, whereas a low level of glucose diminishes the excitability of MCH
neurons and promotes wakefulness. Alternatively, excitation of orexin neurons induces
wakefulness, while inhibition or loss of orexin neurons promotes sleep. Thus, MCH and
orexin neurons have opposing effects on sleep and wakefulness (Burdakov et al. 2005;
González et al. 2016).
Neuronal tissues are supplied with different sources of energy, such as glucose, lactate, and
acetate, through regulated mechanisms. Neurometabolic coupling is the mechanism by
which the brain energy metabolism and cerebral blood supply are modulated locally to meet
the needs of neuronal activity. The energy requirement of neurons can change rapidly during
wakefulness after sensory or motor stimulations, in sleep/waking and waking/sleep
transitions (Petit & Magistretti 2016). Different sleep stages can be distinguished based on
cortical activity, muscle tone, and eye movements. Deep, also called slow wave sleep
(SWS), is characterized by decreased heartbeat, breathing, and body temperature and
reduced cortical activity typically showing high-amplitude, low-frequency (0.5–4 Hz) EEG
activity. In rapid eye movements (REM) sleep cortical theta activity (4-11 Hz) is dominant
(Niethard et al. 2016). Studies have shown that the energy demand diminishes during SWS
sleep; thus, the utilization of glucose and O2 is reduced. Moreover, non-REM sleep promotes
anabolic processes such as biosynthesis of proteins, glycogen and fatty acids (Dworak et al.
2010). In contrast, sleep deprivation affects metabolic coupling and reduces glucose uptake
in all brain areas in humans and rodents and decreases ATP levels by enhancing energy
expenditure. Moreover, chronic sleep fragmentation reduced the uptake of 2-deoxyglucose
in cortex and hippocampus and decreased lactate levels in the cortex (Baud et al. 2016).
These results highlight the importance of non-REM sleep to restore brain energy levels
(Dworak et al. 2010) because it is during the non-REM sleep when the increase in ATP and
glycogen biosynthesis occurs; while REM sleep utilizes similar cerebral glucose as during
wakefulness (Maquet et al. 1990).
Accepted Article
This article is protected by copyright. All rights reserved.
Thus, synaptic potentiation, during the wakeful condition, increases consumption of ATP by
activating glycolysis, fatty acid oxidation, and glucose uptake; while during non-REM sleep,
ATP consumption decreases and leads to energy conservative processes, such as synaptic
scaling.
3. Sources of brain energy
Several regulatory mechanisms operate to regulate the production and usage of energy in
the brain. Though the brain requires a high amount of energy, it possesses minimal energy
reserve which can only satisfy a small portion of its energy demand and is dependent on the
supply of energy substrates from the blood through the blood brain barrier (BBB). Under
normal physiological conditions, the major energy fuel for the brain is glucose (Dienel,
2012a). However, several studies have shown that the brain can use alternative energy
substrates such as lactate, medium chain triglycerides (MCTGs), as well as ketone bodies,
during development and when glucose availability is limited (Owen et al. 1967; Hasselbalch
et al. 1994; Ebert et al. 2003).
Specific transporters allow the uptake of energy substrates, such as glucose and
monocarboxylic acids (lactate, pyruvate, β-hydroxybutyrate, acetoacetate and acetate) from
the circulation across the endothelial cell membrane and into brain cells. The uptake of
energy substrates by brain cells is dependent on the type and distribution of transporters
unique to each cell type and transport rate, on the number of transporters, and the catalytic
activity of each transporter. In the brain, several isoforms of the glucose transporter (GLUT)
and the monocarboxylic acid transporter (MCT) have been identified.
3.1 Glucose
Glucose metabolism provides the fuel for physiological brain function through the synthesis
of ATP via glycolysis, the pentose phosphate pathway (PPP), and the tricarboxylic acid
(TCA) cycle which serves as the basis for neuronal and non-neuronal cellular maintenance
as well as neurotransmitter and gliotransmitter production. Cells of the BBB, mainly
astrocytes and pericytes, act as the gatekeepers for glucose entry into the brain. Glucose
enters the brain tissue from the plasma by transport across the BBB mediated by the
GLUTs. The cytoarchitectural presence of astrocytic end feet enriched in the 45 kDa-isoform
of GLUT1 makes astrocytes an ideal cell type for the uptake of glucose. Glucose enters
neurons trans-cellularly through astrocytes via GLUT1 or directly via GLUT3, a neuronal
GLUT (Maher et al. 1994) (Figure 2).
Astrocytes are considered highly glycolytic due to low expression and activity of the E3
ubiquitin ligase APC/C-Cdh1, which in neurons mediates proteasomal degradation of the
glycolytic key enzyme phosphofructokinase (PFK) (Herrero-Mendez et al. 2009). Thus, high
expression and activity of PFK in astrocytes results in glucose and glucose-6-phosphate
(G6P) mainly being metabolized via the glycolytic pathway (Bélanger et al. 2011). Moreover,
the expression of lactate dehydrogenase (LDH) 5, which favors the conversion of
glycolytically derived pyruvate to lactate regenerating NAD+ (oxidized nicotinamide adenine
nucleotide) that is required as substrate in the reaction of glyceraldehyde-3-phosphate
dehydrogenase (GAPDH), also contributes to the high glycolytic activity of astrocytes
(Pellerin & Magistretti 2004; Hirrlinger & Dringen 2010).
Accepted Article
This article is protected by copyright. All rights reserved.
Unlike glycolytic astrocytes which synthetize lactate in aerobic conditions, neurons rely on
oxidative metabolism through the TCA cycle for their high energy needs. Neuronal activity
increases neurotransmitter concentration in the synaptic cleft (i.e glutamate) and
extracellular K+. This is met by intense mitochondrial activity putting neurons in risk of
oxidative stress, glutamate excitotoxicity, and apoptotic death. Increased glycolysis in the
brain can induce glutathione oxidized (GSSG) accumulation. Glutathione is a tripeptide that
serves not only as a scavenging antioxidant but also has been proposed as a reservoir for
glutamate (Koga et al. 2011). Therefore, as a neuroprotective mechanism, neurons
downregulate glycolysis and the use of glucose to maintain an antioxidant reduced
glutathione (GSH) pool. Additionally, glucose is used by neurons to restore reducing
equivalents of NADPH (reduced nicotinamide adenine dinucleotide phosphate) used in GSH
regeneration via the PPP which cannot be fueled by lactate (Gavillet et al. 2008; Bolaños et
al. 2010).
Moreover, glutamate released from presynaptic terminals into the synapse during neuronal
activation is taken up by astrocytes together with Na+ ions. Furthermore, it stimulates both
glucose uptake and glycolytic processing as well as lactate release in an ATP-dependent
manner, and as such plays a significant role in neuroenergetics (Magistretti et al. 1999).
Astrocytes convert glutamate to glutamine by glutamine synthetase (Norenberg & Martinez-
Hernandez 1979), which is then taken up by neurons and converted back to glutamate by
phosphate-activated glutaminase (GLS) (Figure 2), completing the glutamate-glutamine
cycle (Laake et al. 1999). Conversely, glutamate can be oxidized to enter into the TCA cycle
(Bak et al. 2006). Interestingly, astrocytes have the anaplerotic enzyme pyruvate
carboxylase (PC) (Yu et al. 1983), which carboxylates pyruvate to generate oxaloacetate.
This process is crucial to replace lost TCA cycle intermediates used in the synthesis of
neurotransmitters. Although sodium-coupled uptake of glutamate by astrocytes and the
ensuing activation of the Na+/K+-ATPase may trigger glycolysis (Fox et al. 1988; Magistretti
2006) in a time scale of minutes, extracellular K+ was demonstrated to stimulate astrocytic
glycolysis in vitro within seconds using a genetically encoded fluorescence resonance
energy transfer (FRET) -based nanosensor for glucose (Bittner et al. 2011), an effect
mediated by the Na+/HCO3- cotransporter NBCe1(Ruminot et al. 2011). This K+-mediated
mechanism was suggested to be a general hallmark of neurometabolic coupling, as K+ is
released not only by the postsynaptic terminal during glutamatergic neurotransmission but
also at nodes of Ranvier during AP propagation, at serotonergic synapses, or at the
cholinergic neuromuscular junction.
3.2 Glutamate
Glutamate acts as the major excitatory neurotransmitter in brain. As mentioned above, it
plays a role as a trigger in stimulating glucose utilization when taken up by astrocytes
leading to lactate production, acts as a recycled precursor for the neuronal neurotransmitter
pool and as an energy substrate in astrocytes. Hans Krebs was the first who recognized
glutamate as a brain energy substrate capable of increasing respiration in rabbit brain cortex
in the absence of glucose (Krebs, 1935). Later his group discovered that, in the absence of
glucose, extracellular glutamate was metabolized and mostly converted into aspartate in rat
brain homogenate (Haslam and Krebs, 1963). Further research discovered that glutamate is
not only the co-substrate in the aminotransferase catalyzed interconversion reaction into
aspartate but also into alanine, leucine, isoleucine, and valine (McKenna 2012; Schousboe
et al. 2014). Astrocytes convert the glutamate taken up from the extracellular milieu to α-
Accepted Article
This article is protected by copyright. All rights reserved.
ketoglutarate (α-KG) either primarily by the glutamate dehydrogenase (GDH)-catalyzed
energy-producing reaction or by transamination reactions in astrocytes (Bergles et al. 1999;
Zaganas et al. 2012). Further, the formed α-KG is metabolized into the four-carbon
compound oxaloacetate through sequential reactions of the TCA cycle and eventually
harvests nine ATP molecules (McKenna 2013).
Uptake of extracellular glutamate is an expensive process as astrocytes utilize one ATP for
the transfer of one glutamate. Evidence from in vitro and in vivo studies suggests that
mitochondrial mechanisms and multiprotein complexes are tightly associated with glutamate
uptake by astrocytes. Glial glutamate transporters are coupled to the first step in glycolysis
mediated by hexokinase, which ensures the oxidative energy metabolism of glutamate in
mitochondria yielding ample energy that pays the cost of glutamate uptake from the synaptic
cleft (Genda et al. 2011; Whitelaw & Robinson 2013). Oxidative metabolism of glutamate by
astrocytes, which results in complete glutamate oxidation and energy generation, has been
reported by several in vitro and in vivo studies (McKenna MC., 2013). Studies based on
substrate competition indicated that oxidation of glutamate in astrocytes as a source of
energy is preferred compared to glucose and other substrates (Figure 2), which suggests the
robustness of glutamate use as a substrate for energy in these cells (McKenna et al. 2012).
The steady and dynamic connection between the pre- and postsynaptic neuron and
astrocytes is efficiently involved in maintaining the low extracellular resting glutamate
concentration of ~1-10 μM to avoid excitotoxicity and maintain healthy brain states
(Waagepetersen et al. 1999; Hertz 1979; Schousboe 1981). In vitro experiments in primary
cultured neurons indicate that 10 mM of extracellular glutamate causes profound neuronal
cell death under normoglycemic and normoxic conditions (Khanna et al. 2003; Murphy et al.
1990). However, when glucose is removed from media, extracellular glutamate (10 mM) no
longer induces cell death (Rink et al. 2011).
3.3 Lactate and the Astrocyte-Neuron Lactate Shuttle (ANLS) hypothesis
Despite glucose being the major source of energy in the brain, under various circumstances,
lactate serves as an alternative energy substrate. Since lactate cannot diffuse passively
across the BBB, it needs to be produced in situ within the brain and efficiently transported
between cells (Figure 2). Additionally, as lactate cannot be utilized directly, conversion of
lactate to pyruvate, catalyzed by LDH, is essential as it then provides 18 ATPs via OxPhos
(Pellerin & Magistretti 1994; Magistretti 2006). The identification of MCTs further strengthens
the presence of lactate release and uptake in brain cells (Pellerin et al. 2005; Bröer et al.
1999). MCTs operate depending on the association with the glycoproteins basigin (MCTs 1,
3, 4) or embigin (MCT2) and the concentration gradients of monocarboxylates across the
plasma membrane (Halestrap 2013). While MCT2 and MCT4 are selectively expressed by
neurons and astrocytes, respectively, MCT1 is expressed in astrocytes, oligodendrocytes,
and endothelial cells of blood vessels (Debernardi et al. 2003; Pierre & Pellerin 2005;
Rinholm et al. 2011). Hence, lactate secreted by astrocytes through MCT4 and MCT1 is
assumed to be transported by MCT2 into neurons, where it can be converted to pyruvate
(Figure 2). Then, pyruvate can either enter the TCA cycle via pyruvate dehydrogenase
(PDH) and be metabolized via OxPhos in mitochondria to generate ATP or be converted to
lactate or alanine by LDH or aminotransferase, respectively. The distribution pattern of the
LDH isoforms (high expression of LDH5 in astrocytes and LDH1 in neurons) further provides
evidence for this concept. Furthermore, the muscle-type LDH5 having a greater Vmax, is
Accepted Article
This article is protected by copyright. All rights reserved.
better equipped at converting pyruvate to lactate and supports higher glycolytic rates, but
LDH1 exhibits a lower Km and is inhibited by low concentrations of pyruvate and by lactate.
Thus, efficient conversion of lactate to pyruvate and clearance of pyruvate are essential for
the enzymes to function properly. In summary, higher expression of LDH5 suggests a higher
rate of glycolysis in astrocytes and higher expression of LDH1 suggests lactate utilization as
an energy source in neurons. These data support the astrocyte-neuron lactate shuttle
(ANLS) hypothesis postulated in 1994 (Pellerin & Magistretti 1994). According to this,
astrocytes serve as a ‘lactate source’ whereas neurons serve as a ‘lactate sink’ (Figure 2).
In this context, when neuronal activity intensifies, astrocytes increase their glucose uptake,
thus increasing the rate of glycolysis and lactate release into the extracellular space.
Increased neuronal activity corresponds to an increased release of glutamate from
presynaptic vesicles into the synapse. Excessive glutamate is sensed and taken up by
astroglial glutamate transporters (i.e. EAAT1 and EAAT2). A sodium gradient drives this
glutamate exchange, where one glutamate is co-transported with three Na+ ions, thereby
increasing the concentration of Na+ within astrocytes. Glutamate uptake triggers glucose
uptake by the astrocytes in a stoichiometric ratio of 1:1. A higher Na+ concentration within
astrocytes leads to the activation of the α2 subunit of the Na+/K+ -ATPase which results in
glycolysis stimulation (Mason 2017). This stimulation leads to the production of lactate which
is used as a substitute energy substrate by neurons. Thus, lactate plays a role as an
alternative to meet the energetic demands of the CNS.
In opposition to the ANLS hypothesis, Bak and colleagues argue that oxidative metabolism
of lactate within neurons only occurs during repolarization (and in the period between
depolarizations) rather than during neurotransmission activity (Bak et al. 2009); and that
neurons use lactate as an ‘opportunistic’ substrate when it is present (Bak & Walls 2018). In
synaptic terminals, the use of lactate as energy source is tightly coupled to the activity of the
malate-aspartate shuttle (MAS), since its inhibition decreases the rate of lactate oxidation
(McKenna et al. 1993). According to the model proposed by Bak et al., elevated
neurotransmission may not increase oxidative metabolism of lactate; it decreases possibly
because of a depolarization-induced increase in intracellular Ca2+ concentration and a
putative limitation of the MAS, which transfers the reducing equivalents from the NADH
produced during glycolysis into mitochondria. Thus, when the activity of the MAS is limited
(due to Ca2+-induced activation of the α-KG dehydrogenase, which competes with the
malate/α-KG carrier for substrate), NAD+ cannot be regenerated for glycolysis and NADH is
not further oxidized in the electron transport chain (ETC), leading to an increase in cytosolic
NADH concentration and a decrease in glycolysis and OxPhos. At this point, pyruvate is
further converted into lactate with the concomitant regeneration of NAD+ for glycolysis; and it
is during repolarization (when cytosolic Ca2+ is low and the MAS is no longer limited) when
accumulated lactate is oxidized (Bak et al. 2009).
The ANLS concept explains the role of lactate as an important energy source for brain
function as well as defines the strong metabolic association between astrocytes and
neurons. It is suggested that most of the neurodegenerative diseases, as well as any other
adverse changes in the brain, have notable changes in the ANLS and lead to imbalances in
neurometabolic coupling (Bélanger et al. 2011). However, the ANLS concept has been
critically viewed and challenged due in part to contradictory results obtained in different
Accepted Article
This article is protected by copyright. All rights reserved.
experimental settings, and to the lack of a consensus-based method for real-time monitoring
of metabolic dynamics at cellular resolution. For example, while some authors argue that
neurons prefer lactate over glucose (Bouzier-Sore et al. 2003); others claim neurons are well
equipped to metabolize glucose in an activity-dependent manner (Bak et al. 2009), which
clearly opposes to the tight coupling of astrocytes to synaptic activity (Ruminot et al. 2017).
To further build upon these findings, for both research and diagnose purposes, positron
emission tomography (PET) with the glucose analogue 18F fluorodeoxyglucose (FDG) is
commonly used as clinical diagnostic measure for local glucose metabolism. Further
development of fluorescent protein-based sensors for specific, real-time readouts of
metabolites will fill this current technological gap (Zhang et al. 2018) and shed light on this
hypothesis which has long been a subject of debate (Chih & Roberts 2003; Dienel 2012;
Barros & Weber 2018a; Bak & Walls 2018; Barros & Weber 2018b). However, based on the
accumulated evidence in favor compared to that against this postulation, the ANLS
hypothesis seems now to be broadly accepted.
3.3.1 Glucose vs. Lactate at rest and while exercising
At rest, the predominant energy source for the brain is glucose (Dienel 2012; McEwen &
Reagan 2004). Strenuous physical activity increases O2 demand by the skeletal muscles
leading to increased heart rate and respiration. However, due to unmet O2 demand, plasma
lactate levels are enhanced by conversion of pyruvate to lactate, which is an important step
to regenerate NAD+, an essential substrate to carry out glycolysis and release ATP in the
muscle (Lucas-Cuevas et al. 2015). Rigorous physical activity increases lactate plasma
levels from 0.6 mmol/l to around 2 - 3 mmol/l, which then crosses the BBB via MCTs (Ide et
al. 1999). The cerebral uptake of lactate is thought to be 2-fold higher than that of glucose.
While performing extensive physical exercise, the cerebral metabolic ratio for carbohydrates,
defined as cerebral molar uptake of (O2/ (glucose +1/2 lactate)), decreases from a resting
value of 6 to <2 (Quistorff et al. 2008; Smith et al. 2003). Further, a clinical study involving
administration of sodium lactate reported a 3-4 fold increase in plasma lactate levels along
with an average of 17 % reduction in the rate of cerebral glucose uptake, indicative of a
preferential cerebral uptake of lactate over glucose (Smith et al. 2003). Lactate is thought to
be a favored cerebral energy source since the conversion of lactate to pyruvate does not
require ATP and is thermodynamically preferable compared to glucose that needs two
molecules of ATP (Dienel 2012; Quistorff et al. 2008).
Induced alteration in metabolic substrate concentration, through voluntary exercise, has
been shown to enhance metabolic enzymes involved in glycolysis, ATP synthesis, ATP
transduction, and glutamate turnover (Ding et al. 2006). Preclinical studies report exercise
training in mice augmented the expression of metabolically relevant genes (i.e. PGC-1α,
SIRT1, and citrate synthase) as well as increased mitochondrial DNA (mtDNA), suggestive
of mitochondrial biogenesis in most brain regions studied (Steiner et al. 2011). These
results, together with several converging lines of evidence suggest a critical role for
alterations in global and regional brain metabolism in the pathogenesis of neurodegenerative
diseases, indicating that physical activity could provide clinical benefit. However, the
mechanisms involved in exercise-induced cerebral mRNA expression of metabolically
relevant genes and mitochondrial biogenesis are not fully understood and need further
investigation.
Accepted Article
This article is protected by copyright. All rights reserved.
3.4. Ketone bodies
Ketone bodies provide the brain with an alternative source of energy during periods of
prolonged fasting and starvation. Under normal physiological conditions, monocarboxylates
cross the BBB with poor efficiency, but under starvation, the amount of ketones present in
the blood and expression of MCTs in cells forming the BBB are increased (Hasselbalch et al.
1995). Additionally, mild elevation of blood ketone bodies occurs during the process of
normal aging (Sengupta et al. 2010).
β-Hydroxybutyrate (β-OHB) is a metabolic intermediate that constitutes ~ 70 % of ketone
bodies produced in liver mitochondria mainly from the oxidation of fatty acids released from
adipose tissue (Persson 1970). The concentration of β-OHB in plasma under healthy, fasted
conditions is relatively low with reference values reported ~ 0.04 mM - 0.08 mM (Hansen &
Freier 1978). This may increase by fasting or starvation to 5 – 6 mM (Owen et al. 1967), and
to ~ 25 mM by dietary intervention or diabetic ketosis (Mitchell et al. 1995; Garber et al.
1974; Bonnefont et al. 1990; Saudubray et al. 1981). In non-diabetic subjects, a three-day
fast increases the concentration of β-OHB from an average of 0.03 to 3.15 mM in plasma,
and from 0.05 to 0.98 mM in the brain (Pan et al. 2000). Normal serum levels of ketone
bodies can be defined as <0.5 mM, hyperketonemia in excess of 1.0 mM, and ketoacidosis
over 3.0 mM (Mitchell et al. 1995).
Regulation of ketone body metabolism is different in astrocytes and synaptic terminals.
Inhibition of the MAS and other transaminase reactions on the oxidation of energy substrates
increases the oxidation of lactate and β-OHB in astrocytes, but has no significant effect on
the rates of β-OHB oxidation and decreases the rate of lactate oxidation by synaptic
terminals (McKenna et al. 1993). Additionally, ketone bodies neither alter the plasma
membrane potential of presynaptic terminals nor the pH of synaptic vesicles. β-OHB
supports synaptic vesicle recycling; however, reduces both endocytosis and, to a smaller
extent, exocytosis (Hrynevich et al. 2016).
3.5 Medium chain triglycerides (MCTGs) as alternative brain energy fuel
MCTGs are fatty acids with 7-12 carbon chain length; the most common MCTGs include
heptanoate (C7), octanoate (C8) and decanoate (C10) (Schönfeld & Wojtczak 2016). Owing
to their smaller size, MCTGs can readily diffuse into the brain as they do not require carnitine
transporters unlike long chain fatty acids (Oldendorf 1973; Kuge et al. 1995). Additionally,
even MCTGs can be degraded in the liver to the 4 carbon (C4) ketone β-OHB. C4 ketones
and MCTGs can be further metabolized to acetyl-CoA and enter into the TCA cycle, primarily
in astrocytes (Edmond et al. 1987; Marin-Valencia et al. 2013). Octanoate was shown to
promote ketogenesis in astrocytes (Thevenet et al. 2016) and to inhibit brain glycolysis in
mice (McDonald et al. 2014) by reducing the maximal activity of the rate-limiting enzyme
PFK (Tan et al. 2017). On the other hand, decanoate facilitates glycolysis and lactate
formation in astrocytes thereby providing fuel to neurons (Thevenet et al. 2016) and
improving mitochondrial energy metabolism (Tan et al. 2017).
The odd chain MCTGs are unique because they can be β-oxidized to deliver propionyl-CoA
and acetyl-CoA. Additionally, they can be metabolized in the liver to 5 carbon (C5) ketone
bodies β-ketopentanoate and β-hydroxypentanoate (Kinman et al. 2006; Deng et al. 2009).
Triheptanoin, a triglyceride of heptanoate, generates three molecules of heptanoate upon
hydrolysis. Heptanoate can directly enter into the brain or can be degraded into C5 ketones
Accepted Article
This article is protected by copyright. All rights reserved.
in the liver (Marin-Valencia et al. 2013). The C5 ketones get converted into propionyl CoA,
an anaplerotic molecule which can be carboxylated to succinyl coenzyme A (CoA) to feed
the TCA cycle. Anaplerosis is the process by which TCA cycle intermediates are restored to
permit continuous function which is necessary during neurotransmission. Triheptanoin was
shown to partially replenish TCA cycle intermediates in epileptic mice (Hadera et al. 2014).
Additionally, numerous studies have demonstrated the beneficial effects of MCTGs as
energy substrates in several neurological disorders such as epilepsy (Willis et al. 2010; Wlaź
et al. 2012; Chang et al. 2013; Tan et al. 2017), amyotrophic lateral sclerosis (Zhao et al.
2012; Tefera et al. 2016), stroke (Schwarzkopf et al. 2015), and cognitive impairments
(Wang & Mitchell 2016). Overall, these studies corroborate the importance of alternative
fuels as a source of energy to improve brain energy metabolism with respect to
pathoconditions.
4. The role of non-neuronal cells in brain energy metabolism
4.1 Pericytes
The perivascular location and morphology of pericytes led to the suggestion that they may
be contractile cells involved in regulation of capillary blood flow in response to vasoactive
agents and neural activity (Sweeney et al. 2016). They may also be found around the
lymphatic capillaries in cases of developmental abnormalities (Petrova et al. 2004).
Additionally, pericytes exhibit macrophage-like activity, as shown by the presence of
numerous lysosomes within their cytoplasm (Allt & Lawrenson 2001), their efficient uptake
capacity for soluble tracer compounds (delivered into the blood, in the ventricular
cerebrospinal fluid, or in the extracellular fluid by direct injection into the tissue) (Rucker et
al. 2000), their phagocytic activity (Thomas 1999), and their capability to present antigens
(Rustenhoven et al. 2017).
4.1.1 Role of pericytes in the BBB and glucose homeostasis
Pericytes are imperative for normal CNS functioning and have important roles in
angiogenesis, vessel stabilization, endothelial cell regulation, and maintenance of the BBB
(Fisher 2009; Hill et al. 2014). These functions of pericytes are vital in maintaining the
homeostasis of the perivascular environment (Vezzani et al. 2016). Moreover, glucose intake
of pericytes is four times higher than of endothelial cells, but similar to that of astrocytes,
suggesting potential astrocyte/pericyte complementary roles in maintaining glucose
homeostasis in the brain. Furthermore, astrocytes and pericytes express comparable
amounts of GLUT1 and GLUT4 (Castro et al. 2018).
As endothelial cells comprise a portion of the BBB that regulates CNS transport of energy
metabolites, ions, and clearance of neurotoxic metabolites (Zhao et al. 2015), and since
there is a strong interaction between pericytes and endothelial cells by means of tight and
gap junctions found at their contact sites (Cuevas et al. 1984) allowing physical
communication and molecule exchange between this cells, pericytes have also been
reported to play an important role in the BBB (Daneman et al. 2010). Additionally, the
stabilization of these contact sites by adhesion plaques between the cells and fibronectin
from the extracellular matrix supports the fine-tuned distribution of the contractile force
caused by vascular smooth muscle cells (Vezzani et al. 2016). Reports suggest that the
pericyte/endothelial interaction regulates the basement membrane (Stratman et al. 2009;
Stratman et al. 2010) and the anatomical proximity between pericytes and endothelial cells
Accepted Article
This article is protected by copyright. All rights reserved.
indicates a probable role of pericytes in paracrine or juxtacrine signaling; as pericytes have
been proposed to be involved in several signaling pathways including Angiopoietin-1, TGF-
β, and PDGFRβ (Gaengel et al. 2009).
4.2. Astrocytes
Astrocytes are the most abundant glial cells in the brain and display a number of active roles
critical for CNS function (Nedergaard et al. 2003; Barros et al. 2018b), including regulation of
neurotransmitters (Anderson & Swanson 2000), supplying substrates to neurons for OxPhos
(Pellerin et al. 1998), maintaining water homeostasis (Simard & Nedergaard 2004; Salman
et al. 2017a) and regulating blood supply to meet neuronal energy demand (Zonta et al.
2003; Takano et al. 2006; Gordon et al. 2008). Additionally, astrocytes play a major role in
synapse formation, maintenance, and plasticity during brain development and in adulthood
(Araque et al. 1999; Han et al. 2013; Kim et al. 2017). Astrocytes have extensive processes
originating from the soma; some of these ensheath synapses (Ventura & Harris 1999)
forming the ‘tripartite synapse’, while others known as ‘endfeet’ surround the brain arteries
and capillaries (Araque et al. 1999; Oberheim et al. 2006; Barros et al. 2018b).
An intricate link exists between blood flow, glucose utilization, synaptic plasticity and
neuronal activity. This neurometabolic coupling is a salient physiological characteristic of the
brain function and has formed the basis for understanding neuroenergetics. Astrocytes take
up glucose via GLUT1 located at the endfeet covering brain microvessels (Mathiisen et al.
2010) (Figure 2). Following this, glucose is rapidly phosphorylated to G6P by hexokinase I
(Tabernero et al. 2006; Brown & Ransom 2007). G6P can then be channeled into the
glycolytic pathway to produce pyruvate, metabolized to glucose-1-phosphate for the
synthesis of glycogen (Cataldo & Broadwell 1986; Dienel & Cruz 2015), or used in the PPP
(Dringen et al. 2007). While the PPP is only a minor contributor to the total glucose
oxidation, it generates NADPH, an important molecule for maintaining the antioxidant
reduced glutathione, as well as precursors for nucleotide synthesis (Dringen et al. 2007). In
addition to glucose, astrocytes can efficiently use alternative energy substrates, such as
mannose (Dringen et al. 1993); however, other carbohydrates like fructose or galactose are
considered to be poor substrates in astrocytes (Dringen et al. 1993).
4.2.1 Glycogen metabolism
Alternatively, G6P can be used for the synthesis of glycogen, the main storage form of
glucose in the brain (Hertz & Dienel 2002). Glycogen is predominantly found in astrocytes
and seems to be connected to specific organelles and non-randomly distributed (Calì et al.
2016). The glycogen pool is dynamic and rapidly responds to changes in cerebral energetic
demands by undergoing constant degradation and re-synthesis (Swanson 1992), facilitated
by glycogen synthase and glycogen phosphorylase expressed by astrocytes (Brown &
Ransom 2007; Pellegri et al. 1996). Utilization of glycogen stores allows astrocytes to quickly
increase glycolytic flux independent of glucose availability and hexokinase activity in
response to local energy demands (Brown & Ransom 2015). Astrocytic glycogenolysis is
activated by extracellular increase in neurotransmitter concentrations as well as changes in
the K+ homeostasis following neuronal activity (Hof et al. 1988; Sickmann et al. 2009; Wang
et al. 2012). Interestingly, vasoactive intestinal polypeptide (VIP) as well as norepinephrine
(NE) induce glycogenolysis in murine cortex (Magistretti et al. 1981). This signal-induced
regulation of glucose supply is thought to act in a complementary way; while VIP acts locally
Accepted Article
This article is protected by copyright. All rights reserved.
within cortical columns, NE can control glycogen metabolism across adjacent columns.
Glycogen can also be mobilized under conditions such as glucose deprivation (Dringen &
Hamprecht 1992) or elevated cellular Ca2+ levels (Hamprecht et al. 1993). Functionally,
glycogenolysis in astrocytes and the subsequent release of lactate have been shown to play
a crucial role in formation of LTP and memory (Suzuki et al. 2011; Duran et al. 2013). The
important role of glycogen in supporting neuronal signaling is highlighted by the glycogen
shunt, where part of the glucose that enters the astrocyte is converted into glycogen before
entering the glycolytic pathway, despite this being energetically unfavorable compared to
classical glycolysis (Walls et al. 2009). Furthermore, astrocyte glycogen plays an important
role in maintaining neuronal survival during conditions of hypoglycemia in vitro (Swanson &
Choi 1993) and in vivo (Suh et al. 2007).
4.3 Oligodendrocytes
Oligodendroglia are specialized cells in the CNS that are responsible for generation and
maintenance of myelin sheath that surrounds CNS axons (Bradl & Lassmann 2010; Nave
2010). Myelin acts as an electrical insulator by increasing the membrane resistance and
decreasing membrane capacitance, resulting in increased conduction velocity while reducing
axonal size requirements and neuronal metabolic demand. Moreover, myelin enables rapid
saltatory propagation between nodes of Ranvier, allowing fast and efficient transduction of
electrical signals in CNS (Ransom & Sontheimer 1992; Edgar & Garbern 2004; Harris et al.
2012).
While myelination is the primary function of oligodendrocytes, they also provide trophic
support to neurons by secreting a wide variety of neurotrophins, including insulin-like growth
factor-1 (IGF-1), glial cell line-derived neurotrophic factor (GDNF), and brain-derived
neurotrophic factor (BDNF) (Bradl & Lassmann 2010; Saab et al. 2013). Moreover, though
astrocytes play a critical role in sustaining energy substrates through their glycogen stores
and production of lactate through glycolysis, recent evidence suggests that oligodendroglia
are a prominent site of lactate export to neuronal axons (Saab et al. 2013).
4.3.1 Oligodendrocytes as energy consumers
As outlined above, CNS white matter, primarily composed of myelinated axons, is estimated
to consume one third of the energy of grey matter. Conversely, myelination is energetically
costly and the metabolic costs of generating enough lipid and protein for myelin synthesis
may be higher than the energy saved by accelerated axonal conduction (Nave & Werner
2014). It has been reported that during peak myelination, oligodendrocytes increase their
weight three-fold per day (Ludwin 1997; McLaurin & Yong 1995). Remarkably, in the optic
nerve it has been estimated the initial energetic costs of myelin formation during
development can be repaid by 1-2 months normal activity (Harris & Attwell 2012). However,
the ATP needed to maintain myelin throughout life, including the cost of maintaining
oligodendroglia resting potential, likely negates the energy saved (Harris & Attwell 2012).
Therefore, the primary function of myelination is not to save energy, but rather to allow fast
nerve conduction, thereby increasing information processing and improving cognitive power
(Harris & Attwell 2012; Nave & Werner 2014).
Accepted Article
This article is protected by copyright. All rights reserved.
4.4. Microglia
Microglia are cells of mesodermal origin that migrate into the CNS during embryonic
development (Hickey & Kimura 1988). As immune cells, microglia serve both a supportive
and a protective function within the CNS. In their resting state, microglia act as the CNS’s
surveillance system, equipped with a wide range of receptors (i.e. neurotransmitter, cytokine,
chemokine, pattern recognition) (Kettenmann et al. 2011). Microglia interact with neurons
(i.e. scanning synapses) and thus contribute to the structure of neuronal networks and
connectivity (Kettenmann et al. 2011). When activated, as indicated by a change in
morphology towards an increased soma size and thicker proximal ramifications, microglia
can migrate to the site of injury and proliferate (Bernhart et al. 2010). Microglia can mount a
molecular defense through production of bioactive molecules that can be beneficial in some
circumstances (i.e. phagocytosis of aberrant cells posing a threat to the CNS) and
detrimental in others (i.e. when remaining primed in a disease state and being dysfunctional
in response to a secondary injury, which can lead to loss of neural circuits) (Kettenmann et
al. 2011; Daneman 2012; Koss et al. 2019). Proteomic changes in activated microglia
involve several glycolytic enzymes leading to enhanced ATP production; highlighting the
necessity of enhanced cellular metabolism to regulate their adaptability (Bernhart et al.
2010).
4.4.1 Microglial energy sources and metabolism
Microglia and macrophages within the CNS have a high energetic demand to function as
they monitor for abnormalities and make connections with neurons. The three major energy
substrates in microglia are glucose, fatty acids and glutamine. Glucose is imperative for
microglia survival and can enter microglia via transporters GLUT1, GLUT3, GLUT4 and
GLUT5 (Wang et al. 2019; Payne et al. 1997). Conversely, GLUT5, exclusively expressed in
microglial cells of the human and rat brain, is a very poor transporter of glucose and has
been shown to facilitate the passage of fructose across the plasma membrane (Payne et al.
1997; Horikoshi et al. 2003). After uptake, glucose undergoes glycolysis and full aerobic
breakdown or is used for formation and secretion of lactate.
Fatty acids are an alternative source of energy for microglia. After uptake by lipoprotein
lipase (LPL) long chain fatty acyl-CoA synthetase catalyzes the formation of fatty acyl-CoA,
which can only enter the mitochondria together with the carrier protein carnitine. Once inside
the mitochondria, fatty acyl CoA is β-oxidized into acetyl-CoA, which enters the TCA cycle
and subsequently the ETC in order to generate ATP. G protein-coupled receptor 120
(GPR120) is known to bind unsaturated fatty acids and is possibly involved in their anti-
inflammatory effects (Kalsbeek et al. 2016). Microglia highly express all components of the
NADPH oxidase complex, which has been shown to be stimulated by fatty acids to increase
production of reactive oxygen species (ROS) in macrophages.
Finally, glutamine is used as an energy substrate by microglia (Kalsbeek et al. 2016). The
glutamine transporters SLC1A5 and SLC38A1 are expressed by microglia and enable
microglia to take up glutamine. Inside the mitochondria glutamine is converted to glutamate
and ammonia (NH4+) by the GLS, glutamate is further metabolized by GDH1 to α-KG, which
Accepted Article
This article is protected by copyright. All rights reserved.
can enter the TCA cycle. Both microglial-produced glutamate and NH4+ have been shown to
have neurotoxic effects, which may contribute to neuronal cell death in inflammatory,
infectious, ischemic, and neurodegenerative diseases. Infection, for example, increases the
demand of energy of microglia (Kalsbeek et al. 2016). Tissue damage releases ATP which
attracts microglia at the site of infection and consumes a vast amount of energy (Engl &
Attwell 2015). Activation of microglia by ROS, especially nitric oxide (NO), modulates
metabolic assembly based on glucose uptake and up-regulates both anaerobic glycolysis
and OxPhos of PPP. Thus, NO plays a central role in the energy metabolism of microglia
(Gimeno-Bayón et al. 2014).
5. Disturbed energy homeostasis as a hallmark of CNS disorders and brain aging
Energy homeostasis is essential in maintaining a healthy state of the brain. Disrupted energy
homeostasis may either be a cause or an effect of a disease or a disease-like condition (i.e.
unhealthy aging). Metabolic changes contribute to the pathology of neurodegenerative
diseases, traumatic brain injury or stroke, but also cause neurological symptoms of diabetes
mellitus, and accompany normal aging.
5.1 Aging
During human brain development, energy metabolism declines with aging. Studies
calculating brain glucose utilization over the course of human development, from birth to
adulthood, have identified that energy metabolism peaks during early childhood (Kuzawa et
al. 2014a) and declines during aging (Kuzawa et al. 2014a; Skoyles 2014; Kuzawa et al.
2014b). A recent study from Goyal et al. showed that age-related decline in brain glucose
uptake exceeds that of O2 utilization, causing a loss of brain aerobic glycolysis (Goyal et al.
2017).
A suggestion for the decline in energy metabolism is the progression of metabolic deficiency
resulting in the age-associated cognitive decline and general brain function disturbance (Yin
et al. 2014; Ivanisevic et al. 2016). Metabolomics analysis-based studies have identified
compromised cellular energy status with metabolic imbalances suggesting a failure to
maintain metabolite homeostasis (Yin et al. 2014). These studies have identified increased
adenosine monophosphate (AMP), ATP, purine, and pyrimidine levels (Ivanisevic et al.
2016) with the accumulation of these metabolites as hallmarks of multiple neurodegenerative
diseases (Yin et al. 2014; Nyhan 2005).
A recent study has identified metabolism-regulating peroxisome proliferator-activated
receptor (PPAR) transcription factors as possible energetic metabolic switches during adult
neurogenesis (Di Giacomo et al. 2017). PPAR transcription factors have been shown to be
widely distributed within the mammalian brain and to be involved in regulating the expression
of genes involved in energy metabolism (Woods et al. 2003; Cimini et al. 2005; Cimini &
Cerù 2008); making them strong candidates for possible key regulators of metabolic
pathways impacted by brain aging.
Of note, impaired energy metabolism accompanying aging is a distinguishing factor of
neurodegeneration, highlighting aging as a predisposition destabilizing the “healthy” brain
energetics and making it more prone to neurodegenerative diseases.
Accepted Article
This article is protected by copyright. All rights reserved.
5.2 Type 2 Diabetes mellitus (T2DM)
Type 2 diabetes mellitus (T2DM) is a complex metabolic disease more prevalent with aging.
It is characterized by chronic elevated blood glucose levels. Although glucose is the main
energy substrate of the brain, chronic elevated blood glucose is not advantageous for the
brain, thus untreated T2DM can lead to severe cognitive dysfunction, termed diabetic
encephalopathy. Impeded cerebral glucose metabolism has been found in patients with
T2DM; a condition commonly observed in Alzheimer patients (see below) (Mosconi et al.
2008; Baker et al. 2011). This inadequate glucose metabolism can arise from disruptions in
supply, transport, or utilization, which likely all contribute to T2DM (Wardelmann et al. 2019).
Furthermore, cognitive impairment in T2DM patients is suspected to arise due to
hippocampal insulin resistance, which leads to several deleterious effects (Sims-Robinson et
al. 2010; Biessels & Reagan 2015; Correia et al. 2012). Interestingly, a specific hippocampal
decrease in glutamate and glutamine metabolism was recently described in a mouse model
of T2DM (Andersen et al. 2017b), supporting that metabolic changes at the interface of
glucose and neurotransmitter conversion may mediate cognitive hippocampal deficits in
T2DM.
Brain mitochondrial function was also found to be altered in T2DM (Sims-Robinson et al.
2010; Correia et al. 2012). Several studies have documented deleterious effects on
mitochondrial bioenergetics in mice models of T2DM; including altered activity and
expression of components of the ETC (Ernst et al. 2013; Andersen et al. 2017a); highlighting
mitochondrial deficits play a role in the decline of cerebral health in T2DM.
In addition to elevated plasma levels of glucose, increased amounts of ketone bodies,
acetoacetate, and β-OHB, are observed in T2DM model mice (Vannucci et al. 1997;
Poplawski et al. 2011). Along this line, it has been shown that cerebral fatty acids and β-
OHB metabolism are elevated in a mouse model of T2DM (Makar et al. 1995; Andersen et
al. 2017a). These observations correlate with an increased capacity of ketone body transport
into the brain (Pierre et al. 2007) and indicate that alternative substrates might be able to
compensate for the diminished glucose metabolism in T2DM.
Diabetic conditions can have a disease-modifying effect for progressive neurodegenerative
disorders of the brain like Alzheimer’s disease, Parkinson’s disease, Amyotrophic lateral
sclerosis or Multiple sclerosis, all of which involve diverse molecular and cellular
mechanisms, yet despite unrelated aetiology, they all share a metabolic component which
may be a consequence of the disorder or even causally contribute to its occurrence.
5.3 Alzheimer’s disease (AD)
AD is the most common chronic progressive neurodegenerative disorder that causes
dementia. There are two hypotheses of AD pathogenesis. On one hand, the amyloid
hypothesis proposes that the presence of extracellular amyloid-beta (Aβ) plaques and
intracellular neurofibrillary tangles causes brain atrophy and leads to nerve cell death
(Bhardwaj et al. 2017). On the other hand, the mitochondrial cascade hypothesis unifies the
biochemical, histological, and clinical features of sporadic AD (Swerdlow & Khan 2004).
T2DM is a risk factor for the development of AD (Ott et al. 1996; Exalto et al. 2012; Crane et
al. 2013). Indeed, increased brain Tau phosphorylation has been shown in T2DM mouse
models (Kim et al. 2009; Ramos-Rodriguez et al. 2013). It has been suggested that T2DM
Accepted Article
This article is protected by copyright. All rights reserved.
accelerates AD development through perturbations in brain energy metabolism, advanced
glycation end-product (AGE) and ROS production, and synaptic degeneration (Correia et al.
2012; Duarte 2015).
As dysfunctions in cerebral energy metabolism are considered a hallmark of AD, their
detection represents a tool for AD diagnosis and understanding of pathophysiological
mechanisms (Ferris et al. 1980; Mosconi 2013). As mentioned above, glucose metabolism
can be assessed by 18FDG uptake by PET. This technique has revealed a marked reduction
in FDG labeling that correlates with AD progression and is more prominent in brain areas
most affected by the disease, including the temporal and occipital lobes (Marcus et al. 2014).
One possible explanation for brain energetic impairment in AD could be reduced glucose
uptake. It has been observed that expression of GLUT1 and GLUT3 is reduced in AD brains
(Simpson et al. 1994). Furthermore, extracellular Aβ binds to several membrane receptors,
such as the NMDA receptor (De Strooper & Karran 2016). This interaction leads to an
impaired activity of AMP-activated kinase (AMPK), a protein that is normally induced by high
AMP/ATP ratios. Further, AMPK inhibition promotes a reduction of GLUT3 and GLUT4 in
plasma membranes of hippocampal neurons, reducing ATP production (da Silva et al. 2017).
Additionally, Aβ oligomers reduce hexokinase activity and ATP levels in neuronal cultures
(da Silva et al. 2017). Thus, extracellular Aβ contributes to the impairment of the energetic
metabolism.
On the other hand, as aging is a major risk factor for the onset of neurodegenerative
diseases like AD, the mitochondrial cascade hypothesis (Swerdlow & Khan 2004) suggests
that accumulated mutations in mtDNA caused by ROS play a pivotal role in mitochondrial
dysfunction (Lin & Beal 2006), which can be an early event in AD (Nunomura et al. 2001).
Amyloid precursor protein (APP) forms can accumulate in protein import channels of
mitochondria of human AD brains and thus contribute to mitochondrial dysfunction (Devi et
al. 2006). These alterations promote a reduction in energetic metabolism in neurons
(Caspersen et al. 2005). Though, some of these effects could be associated with variations
in protein levels related with mitochondrial dynamics (Itoh et al. 2013). It has been
demonstrated that protein levels of Drp1 are increased in AD and potentiate cell death (Park
et al. 2014). This protein is important for mitochondrial quality surveillance and induces
mitochondrial fission, which as a consequence reduces the protein levels of mitofusin 1/2
(Mfn1/2) and optic atrophy type 1 (OPA1); GTPases that promote the fusion process (Baek
et al. 2017). Additionally, the inhibition of Drp1 ameliorates Aβ deposition and synaptic
impairment. Therefore, reduction of mitochondrial fission leads to increased occurrences of
mitochondrial fusion and results in neuroprotective effects in AD models (Park et al. 2014;
Itoh et al. 2013; Baek et al. 2017).
5.4 Parkinson's disease (PD)
PD is a progressive neurodegenerative disorder characterized by a loss of dopaminergic
(DA) neurons in the substantia nigra pars compacta (SNpc) and the presence of Lewy
bodies (with α-synuclein inclusions) in the substantia nigra (Sveinbjornsdottir 2016).
Interestingly, in a large retrospective cohort study a higher prevalence of PD in patients with
T2DM was detected (De Pablo-Fernandez et al. 2018).
Accepted Article
This article is protected by copyright. All rights reserved.
However, dysfunctions in energy and/or redox homeostasis, along with oxidative stress
leading to mitochondrial dysfunction, are considered the initiators of a chain of events that
result in synaptic dysfunction, neuronal degeneration, and motor disabilities in PD
(Saravanan et al. 2006; Schapira & Jenner 2011). But, the observation that Lewy bodies are
not found in all neurons or are randomly expressed in the brain, as less than 1 % of brain
neurons are affected through the mid-stages of the disease, raised the question of why
SNpc DA neurons are the most vulnerable. It was proposed that there are intrinsic neuronal
features of generating and handling APs that render DA SNpc neurons more susceptible to
mitochondrial dysfunction related to PD (Surmeier et al. 2012).
Indeed, candidate gene and genome-wide association studies identified many genetic
mutations/polymorphisms associated with PD (i.e. in genes like PINK1, Parkin, DJ1/PARK7
and LRRK2) that compromise mitochondrial function and dynamics (Coleman 2012). DA
neurons lacking the mitochondrial fusion gene MFN2 display fragmented mitochondria and
their transport within the axon is hindered (Pham et al. 2012). Mutation in MFN2 leads to
severe locomotory behavioral deficits, which is accompanied by loss of striatal DA efferents.
Moreover, altered mitochondrial fission may result in synaptic loss and DA neuronal cell loss
(Berthet et al. 2014). In human patients, two heterozygous missense mutations in the
mitochondrial fusion promoter OPA1 protein show slow symptoms of PD and dementia
(Carelli et al. 2015).
Studies in various animal models of PD reveal that mitochondrial dysfunction in the brain is
linked to alterations in mitochondrial morphology, dynamics, mutation of mtDNA, increased
proton leak (Villeneuve et al. 2016) and decreased rates of electron transfer (Golpich et al.
2017; van der Merwe et al. 2017). These impairments in mitochondria are associated with:
accumulation of oxidation products of phospholipids and proteins, increased ROS
production, increased lipid peroxidation, decreased respiration and membrane potential,
decreased capacity for ATP production, and neuronal degeneration due to oxidative damage
and energy defects in the aged mammalian brain (Navarro & Boveris 2010).
5.5 Amyotrophic lateral sclerosis (ALS)
ALS is a progressive neurologic disorder, primarily characterized by the selective death of
lower and upper motor neurons in the spinal cord and cortical regions which finally leads to
muscle denervation, weakness and paralysis. Several pathogenic mechanisms are believed
to contribute to motor neuron death, including abnormal protein aggregation (Bruijn et al.
1997), oxidative stress (Barber et al. 2006), glutamate excitotoxicity (Rothstein 1995), and
impaired energy metabolism (Dupuis et al. 2011).
Numerous pre-clinical investigations in vitro and in vivo as well as in patients have revealed
metabolic irregularities in ALS brains which include reductions in glucose uptake (Miyazaki
et al. 2012) and reduced gene expression of enzymes involved in glycolysis (Ferraiuolo et al.
2011), PPP (Kirby et al. 2005), and TCA cycle (D’Arrigo et al. 2010), as well as the ETC
(Ferraiuolo et al. 2011). These metabolic defects, which comprise also disrupted metabolic
interactions between neurons and glial cells (reviewed in Tefera & Borges 2017), result in
impaired mitochondrial oxidative phosphorylation, declined generation of ATP, and
subsequent death of neurons as well as non-neuronal cells.
Accepted Article
This article is protected by copyright. All rights reserved.
Given the impairments in energy metabolism in ALS, several metabolic agents aimed at
correcting metabolic irregularities have been examined in ALS mouse models and/or
patients, including the ketogenic diet (Zhao et al. 2006), dichloroacetate (Miquel et al. 2012),
caprylic triglyceride (Zhao et al. 2012), triheptanoin (Tefera et al. 2016), and others
(reviewed in (Tefera & Borges 2017). These metabolic substrates were able to modify ALS
disease progression to varying degrees, signifying the contribution of CNS energy
metabolism towards the pathogenesis of the disease and the need for further studies to
correct metabolic defects.
5.6 Multiple sclerosis (MS)
MS is a chronic autoimmune neurodegenerative disease characterized by axonal
demyelination and impaired remyelination (Aslani et al. 2017). Activated T cell-mediated
autoimmune destruction of CNS myelin has been highlighted as the major underlying cause,
yet genetic predisposition, oxidative stress, mitochondrial dysfunction, and energy failure
likely contribute to accelerate the disease. Axonal energy failure and disrupted mitochondrial
energetics, especially in white matter astrocytes, have been implicated in MS (Cambron et
al. 2012); highlighted by the absence of β2-adrenoreceptors in the astrocytes of MS patients
(Cambron et al. 2012), which impairs the noradrenaline-mediated glycogenolysis necessary
to exploit this energy reservoir for neurons and their axons, thus leading to disrupted ion
gradients that disturb the excitability of axons. Also, disrupted metabolic support from
oligodendrocytes contributes to axonal damage and disease progression (summarized in
(Philips & Rothstein 2017).
Functionally, energy crisis-driven impaired ATP generation may lead to the failure of
important pumps and exchangers like the Na+/K+-ATPase maintaining the Na+ concentration
within the axon, which is ATP-dependent, and the Na+/Ca2+ exchanger, which further
sustains calcium homeostasis. This leads to an increased Na+ and Ca2+ influx in axons and
disrupted membrane potential. Intra-axonal accumulation of calcium activates
phospholipases and proteases like calpain, which bring about axonal disintegration, and
further lead to energy failure via mitochondrial disruption (Pennisi et al. 2011).
5.7 Traumatic brain injury (TBI) and stroke
Brain injury occurring under ischemic, hemorrhagic, or traumatic conditions is characterized
by metabolic and energy imbalances within cerebral cells. Structural damage causes
reduced cerebral blood flow and cell membrane disruption and is accompanied by
mitochondrial dysfunction, inflammatory responses and oxidative stress, ionic gradient
breakdown, glutamate-mediated excitotoxicity, stress signaling, and ultimately cell death (Lo
et al. 2003; Malik & Dichgans 2018). In addition, the integrity of the BBB is adversely
affected due to the disruption of tight junctions (Chodobski et al. 2011).
Autoradiographic studies have shown a transient increase followed by a prolonged decrease
in cerebral glucose metabolism (measured as cerebral metabolic rate of glucose, CMRglu) in
rodent models as well as in human patients with head injury (Glenn et al. 2003). It is
postulated that the acute phase of hyperglycolysis occurs in order to compensate the
increased cellular energy requirement to restore the ionic imbalance and membrane
potential (Karelina & Weil 2016).The following long-term depression of CMRglu, indicating
reduced glucose uptake and utilization, suggests the inability of the injured brain to meet the
increased metabolic demands which could be due to reduced glucose supply (as a
Accepted Article
This article is protected by copyright. All rights reserved.
consequence of hindered cerebral blood flow, CBF), defects in glucose transporter systems,
and/or reduced need for glucose. However, changes in CBF in patients with TBI show a
strong heterogeneity with some patients showing significantly reduced CBF, while in others
normal CBF values can be seen that persist for days to weeks after injury (Marion et al.
1991; Bouma et al. 1991). Nonetheless, the increased glucose metabolism despite the low
CBF during acute phase after TBI indicates that the metabolic crisis is not likely caused by
the reduced cerebral glucose supply. The other possibility includes the reduced glucose
uptake due to impaired glucose transport systems following TBI. Preclinical studies have
shown a decreased GLUT1 immunoreactivity indicating less transport of glucose molecules
into the brain. However, the data are highly variable requiring further studies. Finally,
shunting of glucose to other metabolic pathways within the cell may contribute to the energy
crisis after TBI. An animal study has shown an upregulation of the PPP with a 12 % increase
in cerebral glucose levels that is shunted towards the synthesis of pentose 24h after the
injury (Bartnik et al. 2005). Moreover, a reduction in glucose-metabolizing enzymes (such as
PDH) and cofactors (NAD+) indicating decreased efficiency of mitochondria to process
glucose for oxidative metabolism in rats (Xing et al. 2009).
Altered ionic balance results in a calcium influx (Kitchen et al. 2015b), which in turn leads to
the depolarization of the mitochondrial membrane due to the opening of the mitochondrial
permeability transition pore and cytochrome c-mediated caspase-dependent apoptosis.
Elevated levels of NO have been observed in animal models of head trauma (Görlach et al.
2015). NO hampers the enzymatic activity of the mitochondrial complex IV and thus
interferes with energy metabolism, too.
Brain edema, which is a hallmark of TBI and stroke, occur when water enters the CNS via
astrocytes at the BBB, primarily through the aquaporin 4 channel (Amiry-Moghaddam &
Ottersen 2003; Kitchen et al. 2015a). There is an increased interest in applying therapeutic
measures that could target TBI and stroke through decreasing the energy need of the brain
as well as act on limiting the devastating effects of edema (Salman et al. 2017b).
Therapeutic hypothermia is gaining popularity to prevent or improve a wide range of
neurological morbidities (Yenari & Han 2012; Salman et al. 2017a). The neuroprotective
effect of therapeutic hypothermia, which potentially involves the restoration of the BBB,
preservation of high-energy phosphate compounds and cellular metabolism, has been
confirmed in a number of clinical trials investigating the outcomes of patients suffering from
neonatal hypoxia-ischemia (Gluckman et al. 2005; Shankaran et al. 2005) and cardiac arrest
(Bernard et al. 2002).
Brain metabolism produces oxygen molecules that are converted into hydrogen peroxides
and finally to hydroxyl radicals (·OH). During normal physiological conditions, these highly
reactive free radicals are quenched by the cellular antioxidant defense systems. However,
during brain injury the levels of free radicals overwhelm the scavenging systems and result
in oxidative damage. ·OH has been shown to increase immediately after injury and peak at
30 minutes in severely injured mice (Hall et al. 1993). In addition, lipid peroxidation, a marker
of increased oxidative metabolism, has been shown to increase up to 24h post-injury (Hall et
al. 1993) .
Accepted Article
This article is protected by copyright. All rights reserved.
Moreover, TBI is characterized by the presence of high lactate concentrations in the
extracellular fluid, clearly indicating increased rates of glycolysis, mitochondrial impairment,
and/or hypoxia. A high lactate/pyruvate ratio (i.e. <25) has been well reported for TBI
patients (Patet et al. 2016). TBI is frequently accompanied by ischemia and hypoxia, major
but not the only causes of mitochondrial dysfunction after injury (Patet et al. 2016; Carpenter
et al. 2015).
6. Concluding remarks
6.1 Costly human brain: evolutionary implications
In humans, the brain consumes more energy relative to the rest of the body and has a higher
relative size than in other animals. However, the relative basal metabolism does not
correlate with the relative size of the brain (Navarrete et al. 2011). Some studies connected
the possibility of having such a costly brain with the relative reduction of gut (Aiello &
Wheeler 1995) or skeletal muscles (Leonard et al. 2003) in volume or with the decrease of
energy spent on locomotive tasks (Navarrete et al. 2011). According to this, the high
energetic demands of the human brain are met not due to elevated basal metabolism, but
rather due to the energy redirection from other metabolically expensive tissues. New findings
highlight the existence of molecular mechanisms of energy redirection. As shown by
Pfefferle et al. (2011) comparing GLUT1 and 4 expression patterns in apes, monkeys and
humans, the energy trade-off between the brain and other organs may be achieved by
redistribution of glucose uptake systems. GLUT1 expression in human brain was found to be
much higher as compared to chimpanzee and macaque brains, but GLUT4 expression is
higher in chimpanzee skeletal muscle than in human and macaque muscle. Another system
of fueling energy to tissue that likely facilitated the evolution of the costly human brain is the
phosphocreatine pathway (Pfefferle et al. 2011). The expression of the active creatine
transporter SLC6A8 and the brain-type cytosolic creatine kinase are higher in humans than
in chimpanzees and rhesus macaques in cerebral cortex and cerebellum, but not in skeletal
muscles (Pfefferle et al. 2011). Thus, shifts in the expression pattern of metabolic pathway
players could account for primate brain divergence in evolution.
6.2 The importance of maintaining brain energy homeostasis
The existence of many alternative energy sources to fuel the brain as well as the
pathological consequences of disturbed regional/global brain metabolism and energetics
highlight the importance of maintaining brain energy homeostasis. Glucose metabolism is
connected to cell death pathways by glucose-metabolizing enzymes. Thus, disruption of
glucose delivery and metabolism can lead to debilitating brain diseases. Furthermore, other
molecules that provide fuel or metabolic intermediates for the brain, for example ketones,
are used under pathological (diabetic ketosis) or physiological conditions such as fasting,
starvation or aging; and lactate during extensive physical exercise. A complex interplay
exists between the brain, in particular the hypothalamus, and peripheral systems that control
glucose uptake and supply to the brain (Lam et al. 2009), utilization, and feeding
(JolyAmado et al. 2012; Wu et al. 2009). When the body undergoes glucose deprivation,
fatty acids are broken down into ketones in the liver and the permeability of the BBB to
monocarboxylic acids increases in parallel; all of which allow ketone body utilization. Long
chain fatty acids, in contrast, do not serve as fuel for the brain because they are bound to
albumin in the plasma and do not traverse the BBB.
Accepted Article
This article is protected by copyright. All rights reserved.
6.3 A non-neurocentric view for neuroenergetics
Neuronal cells together with glia orchestrate the metabolic reactions to maintain the energy
demand and redox balance of neuronal activity. Glial cells are highly specialized in their
roles within the brain. Moreover, neurodegenerative disorders are not exclusively restricted
to neurons but glial cells are indeed a pivotal player. Future research is needed in order to
illuminate the roles of the individual glial cells in brain health and disease as there are still
many molecular mechanisms to define. For example, utilizing a combination of optogenetic
stimulation, electrophysiological recordings, and bioenergetic readouts (Barros et al. 2018a)
as well as further development of non-invasive imaging of brain metabolism with high-field
NMR spectroscopy, PET or SPECT (Hyder & Rothman 2017). Further studies are
warranted, not only because of the high prevalence of metabolic syndrome, but additionally
for a detailed understanding of brain energy metabolism and the link between mental and
physical health.
Acknowledgments
This review was initiated at the 14th International Society for Neurochemistry (ISN)
Advanced School held in Varennes Jarcy, France, in August 2017. We thank the ISN for the
financial and educational support provided, as well as all ISN School faculty members, who
devoted their time to share their views on neuroenergetics and to encourage young
scientists. Work in the lab of C.I.S. is funded by the DFG (SFB779-B14 and RTG2413
“SynAge” TP08), BMBF (IB-049 “PrePLASTic”) and EU (MC-ITN “ECMED”; LSA ESF
ZS/2016/08/80645 “ABINEP”). C.I.S. is an editor for Journal of Neurochemistry. All authors
declare no conflict of interest.
Brain energetics is a huge and growing field, which can hardly be covered in a single review
with satisfying depth. We would like to thank all the scientists and labs who contributed to
our current knowledge and we would also like to apologize to those whom we missed to
refer to, due to space limitations.
References
Aiello, L. C. and Wheeler, P. (1995) The expensive-tissue hypothesis: the brain and the
digestive system in human and primate evolution. Current anthropology 36, 199-221.
Alle, H., Roth, A. and Geiger, J. R. (2009) Energy-efficient action potentials in hippocampal
mossy fibers. Science 325, 1405-1408.
Allt, G. and Lawrenson, J. (2001) Pericytes: cell biology and pathology. Cells tissues organs
169, 1-11.
Amiry-Moghaddam, M. and Ottersen, O. P. (2003) The molecular basis of water transport in
the brain. Nature Reviews Neuroscience 4, 991-1001.
Andersen, J. V., Christensen, S. K., Nissen, J. D. and Waagepetersen, H. S. (2017a)
Improved cerebral energetics and ketone body metabolism in db/db mice. J Cereb
Blood Flow Metab 37, 1137-1147.
Andersen, J. V., Nissen, J. D., Christensen, S. K., Markussen, K. H. and Waagepetersen, H.
S. (2017b) Impaired Hippocampal Glutamate and Glutamine Metabolism in the db/db
Mouse Model of Type 2 Diabetes Mellitus. Neural Plast 2017, 2107084.
Anderson, C. M. and Swanson, R. A. (2000) Astrocyte glutamate transport: review of
properties, regulation, and physiological functions. Glia 32, 1-14.
Accepted Article
This article is protected by copyright. All rights reserved.
Araque, A., Parpura, V., Sanzgiri, R. P. and Haydon, P. G. (1999) Tripartite synapses: glia,
the unacknowledged partner. Trends in neurosciences 22, 208-215.
Aslani, S., Jafari, N., Javan, M. R., Karami, J., Ahmadi, M. and Jafarnejad, M. (2017)
Epigenetic modifications and therapy in multiple sclerosis. Neuromolecular medicine
19, 11-23.
Attwell, D. and Laughlin, S. B. (2001) An energy budget for signaling in the grey matter of
the brain. Journal of Cerebral Blood Flow & Metabolism 21, 1133-1145.
Baek, S.-H., Park, S. J., Jeong, J. I. et al. (2017) Inhibition of Drp1 Ameliorates Synaptic
Depression, Aβ Deposition and Cognitive Impairment in Alzheimer's Disease Model.
Journal of Neuroscience, 2385-2316.
Bak, L. K., Schousboe, A. and Waagepetersen, H. S. (2006) The glutamate/GABAglutamine
cycle: aspects of transport, neurotransmitter homeostasis and ammonia transfer.
Journal of neurochemistry 98, 641-653.
Bak, L. K. and Walls, A. B. (2018) CrossTalk opposing view: lack of evidence supporting an
astrocytetoneuron lactate shuttle coupling neuronal activity to glucose utilisation in
the brain. The Journal of physiology 596, 351-353.
Bak, L. K., Walls, A. B., Schousboe, A., Ring, A., Sonnewald, U. and Waagepetersen, H. S.
(2009) Neuronal glucose but not lactate utilization is positively correlated with
NMDAinduced neurotransmission and fluctuations in cytosolic Ca2+ levels. Journal
of neurochemistry 109, 87-93.
Baker, L. D., Cross, D. J., Minoshima, S., Belongia, D., Watson, G. S. and Craft, S. (2011)
Insulin resistance and Alzheimer-like reductions in regional cerebral glucose
metabolism for cognitively normal adults with prediabetes or early type 2 diabetes.
Archives of neurology 68, 51-57.
Barber, S. C., Mead, R. J. and Shaw, P. J. (2006) Oxidative stress in ALS: a mechanism of
neurodegeneration and a therapeutic target. Biochimica et Biophysica Acta (BBA)-
Molecular Basis of Disease 1762, 1051-1067.
Barros, L. and Weber, B. (2018a) CrossTalk proposal: an important astrocytetoneuron
lactate shuttle couples neuronal activity to glucose utilisation in the brain. The Journal
of physiology 596, 347-350.
Barros, L. and Weber, B. (2018b) Rebuttal from LF Barros and B. Weber. The Journal of
physiology 596, 355-356.
Barros, L. F., Bolaños, J. P., Bonvento, G. et al. (2018a) Current technical approaches to
brain energy metabolism. Glia 66, 1138-1159.
Barros, L. F., Brown, A. and Swanson, R. A. (2018b) Glia in brain energy metabolism: A
perspective. Glia 66, 1134-1137.
Bartnik, B. L., Sutton, R. L., Fukushima, M., Harris, N. G., Hovda, D. A. and Lee, S. M.
(2005) Upregulation of pentose phosphate pathway and preservation of tricarboxylic
acid cycle flux after experimental brain injury. Journal of neurotrauma 22, 1052-1065.
Baud, M. O., Parafita, J., Nguyen, A., Magistretti, P. J. and Petit, J. M. (2016) Sleep
fragmentation alters brain energy metabolism without modifying hippocampal
electrophysiological response to novelty exposure. Journal of sleep research 25,
583-590.
Bélanger, M., Allaman, I. and Magistretti, P. J. (2011) Brain energy metabolism: focus on
astrocyte-neuron metabolic cooperation. Cell metabolism 14, 724-738.
Bergles, D. E., Diamond, J. S. and Jahr, C. E. (1999) Clearance of glutamate inside the
synapse and beyond. Current opinion in neurobiology 9, 293-298.
Bernard, S. A., Gray, T. W., Buist, M. D., Jones, B. M., Silvester, W., Gutteridge, G. and
Smith, K. (2002) Treatment of comatose survivors of out-of-hospital cardiac arrest
with induced hypothermia. New England Journal of Medicine 346, 557-563.
Bernhart, E., Kollroser, M., Rechberger, G. et al. (2010) Lysophosphatidic acid receptor
activation affects the C13NJ microglia cell line proteome leading to alterations in
glycolysis, motility, and cytoskeletal architecture. Proteomics 10, 141-158.
Bernstein, B. W. and Bamburg, J. R. (2003) Actin-ATP hydrolysis is a major energy drain for
neurons. Journal of Neuroscience 23, 1-6.
Accepted Article
This article is protected by copyright. All rights reserved.
Berthet, A., Margolis, E. B., Zhang, J. et al. (2014) Loss of mitochondrial fission depletes
axonal mitochondria in midbrain dopamine neurons. Journal of Neuroscience 34,
14304-14317.
Bhardwaj, D., Mitra, C., Narasimhulu, C. A., Riad, A., Doomra, M. and Parthasarathy, S.
(2017) Alzheimer's Disease-Current Status and Future Directions. J Med Food 20,
1141-1151.
Biessels, G. J. and Reagan, L. P. (2015) Hippocampal insulin resistance and cognitive
dysfunction. Nature reviews. Neuroscience 16, 660-671.
Bittner, C. X., Valdebenito, R., Ruminot, I. et al. (2011) Fast and reversible stimulation of
astrocytic glycolysis by K+ and a delayed and persistent effect of glutamate. Journal
of Neuroscience 31, 4709-4713.
Bolaños, J. P., Almeida, A. and Moncada, S. (2010) Glycolysis: a bioenergetic or a survival
pathway? Trends in biochemical sciences 35, 145-149.
Bonnefont, J., Specola, N., Vassault, A. et al. (1990) The fasting test in paediatrics:
application to the diagnosis of pathological hypo-and hyperketotic states. European
journal of pediatrics 150, 80-85.
Bouma, G. J., Muizelaar, J. P., Choi, S. C., Newlon, P. G. and Young, H. F. (1991) Cerebral
circulation and metabolism after severe traumatic brain injury: the elusive role of
ischemia. Journal of neurosurgery 75, 685-693.
Bouzier-Sore, A.-K., Voisin, P., Canioni, P., Magistretti, P. J. and Pellerin, L. (2003) Lactate
is a preferential oxidative energy substrate over glucose for neurons in culture.
Journal of Cerebral Blood Flow & Metabolism 23, 1298-1306.
Bradl, M. and Lassmann, H. (2010) Oligodendrocytes: biology and pathology. Acta
neuropathologica 119, 37-53.
Brand, M. D., Pakay, J. L., Ocloo, A., Kokoszka, J., Wallace, D. C., Brookes, P. S. and
Cornwall, E. J. (2005) The basal proton conductance of mitochondria depends on
adenine nucleotide translocase content. Biochemical Journal 392, 353-362.
BRÖER, S., BRÖER, A., SCHNEIDER, H.-P., STEGEN, C., HALESTRAP, A. P. and
DEITMER, J. W. (1999) Characterization of the high-affinity monocarboxylate
transporter MCT2 in Xenopus laevis oocytes. Biochemical Journal 341, 529-535.
Brown, A. M. and Ransom, B. R. (2007) Astrocyte glycogen and brain energy metabolism.
Glia 55, 1263-1271.
Brown, A. M. and Ransom, B. R. (2015) Astrocyte glycogen as an emergency fuel under
conditions of glucose deprivation or intense neural activity. Metabolic brain disease
30, 233-239.
Bruijn, L., Becher, M., Lee, M. et al. (1997) ALS-linked SOD1 mutant G85R mediates
damage to astrocytes and promotes rapidly progressive disease with SOD1-
containing inclusions. neuron 18, 327-338.
Burdakov, D., Gerasimenko, O. and Verkhratsky, A. (2005) Physiological changes in glucose
differentially modulate the excitability of hypothalamic melanin-concentrating
hormone and orexin neurons in situ. Journal of Neuroscience 25, 2429-2433.
Calì, C., Baghabra, J., Boges, D. J., Holst, G. R., Kreshuk, A., Hamprecht, F. A., Srinivasan,
M., Lehväslaiho, H. and Magistretti, P. J. (2016) Threedimensional immersive virtual
reality for studying cellular compartments in 3D models from EM preparations of
neural tissues. Journal of Comparative Neurology 524, 23-38.
Cambron, M., D'haeseleer, M., Laureys, G., Clinckers, R., Debruyne, J. and De Keyser, J.
(2012) White-matter astrocytes, axonal energy metabolism, and axonal degeneration
in multiple sclerosis. Journal of Cerebral Blood Flow & Metabolism 32, 413-424.
Carelli, V., Musumeci, O., Caporali, L. et al. (2015) Syndromic parkinsonism and dementia
associated with OPA 1 missense mutations. Annals of neurology 78, 21-38.
Carlier, M.-F., Pantaloni, D., Evans, J. A., Lambooy, P. K., Korn, E. D. and Webb, M. R.
(1988) The hydrolysis of ATP that accompanies actin polymerization is essentially
irreversible. FEBS letters 235, 211-214.
Carpenter, K. L., Jalloh, I. and Hutchinson, P. J. (2015) Glycolysis and the significance of
lactate in traumatic brain injury. Frontiers in neuroscience 9, 112.
Accepted Article
This article is protected by copyright. All rights reserved.
Caspersen, C., Wang, N., Yao, J. et al. (2005) Mitochondrial Aβ: a potential focal point for
neuronal metabolic dysfunction in Alzheimer’s disease. The FASEB Journal 19,
2040-2041.
Castro, V., Skowronska, M., Lombardi, J., He, J., Seth, N., Velichkovska, M. and Toborek,
M. (2018) Occludin regulates glucose uptake and ATP production in pericytes by
influencing AMP-activated protein kinase activity. Journal of Cerebral Blood Flow &
Metabolism 38, 317-332.
Cataldo, A. M. and Broadwell, R. D. (1986) Cytochemical identification of cerebral glycogen
and glucose6phosphatase activity under normal and experimental conditions: I.
Neurons and glia. Journal of Electron Microscopy Technique 3, 413-437.
Chang, P., Terbach, N., Plant, N., Chen, P. E., Walker, M. C. and Williams, R. S. (2013)
Seizure control by ketogenic diet-associated medium chain fatty acids.
Neuropharmacology 69, 105-114.
Chih, C. P. and Roberts, E. L. (2003) Energy substrates for neurons during neural activity: a
critical review of the astrocyte-neuron lactate shuttle hypothesis. J Cereb Blood Flow
Metab 23, 1263-1281.
Chodobski, A., Zink, B. J. and Szmydynger-Chodobska, J. (2011) Blood–brain barrier
pathophysiology in traumatic brain injury. Translational stroke research 2, 492-516.
Cimini, A., Benedetti, E., Cristiano, L., Sebastiani, P., D'amico, M., D'angelo, B. and Di
Loreto, S. (2005) Expression of peroxisome proliferator-activated receptors (PPARs)
and retinoic acid receptors (RXRs) in rat cortical neurons. Neuroscience 130, 325-
337.
Cimini, A. and Cerù, M. P. (2008) Emerging roles of peroxisome proliferator-activated
receptors (PPARs) in the regulation of neural stem cells proliferation and
differentiation. Stem cell reviews 4, 293-303.
Cingolani, L. A. and Goda, Y. (2008) Actin in action: the interplay between the actin
cytoskeleton and synaptic efficacy. Nature Reviews Neuroscience 9, 344.
Coleman, M. P. (2012) Mitochondria as a central sensor for axonal degenerative stimuli.
Trends in neurosciences 35, 364-372.
Correia, S. C., Santos, R. X., Carvalho, C., Cardoso, S., Candeias, E., Santos, M. S.,
Oliveira, C. R. and Moreira, P. I. (2012) Insulin signaling, glucose metabolism and
mitochondria: major players in Alzheimer's disease and diabetes interrelation. Brain
Res 1441, 64-78.
Crane, P. K., Walker, R. and Larson, E. B. (2013) Glucose levels and risk of dementia. The
New England journal of medicine 369, 1863-1864.
Cuevas, P., Gutierrez-Diaz, J. A., Reimers, D., Dujovny, M., Diaz, F. G. and Ausman, J. I.
(1984) Pericyte endothelial gap junctions in human cerebral capillaries. Anat Embryol
(Berl) 170, 155-159.
D’Arrigo, A., Colavito, D., Peña-Altamira, E., Fabris, M., Dam, M., Contestabile, A. and Leon,
A. (2010) Transcriptional profiling in the lumbar spinal cord of a mouse model of
amyotrophic lateral sclerosis: a role for wild-type superoxide dismutase 1 in sporadic
disease? Journal of molecular neuroscience 41, 404-415.
da Silva, G. S. S., Melo, H. M., Lourenco, M. V. et al. (2017) Amyloid-β oligomers transiently
inhibit AMP-activated kinase and cause metabolic defects in hippocampal neurons.
Journal of Biological Chemistry, jbc. M116. 753525.
Daneman, R. (2012) The blood–brain barrier in health and disease. Annals of neurology 72,
648-672.
Daneman, R., Zhou, L., Kebede, A. A. and Barres, B. A. (2010) Pericytes are required for
blood–brain barrier integrity during embryogenesis. Nature 468, 562.
De Pablo-Fernandez, E., Goldacre, R., Pakpoor, J., Noyce, A. J. and Warner, T. T. (2018)
Association between diabetes and subsequent Parkinson disease: A record-linkage
cohort study. Neurology 91, e139-e142.
De Strooper, B. and Karran, E. (2016) The cellular phase of Alzheimer’s disease. Cell 164,
603-615.
Accepted Article
This article is protected by copyright. All rights reserved.
Debernardi, R., Pierre, K., Lengacher, S., Magistretti, P. J. and Pellerin, L. (2003) Cell-
specific expression pattern of monocarboxylate transporters in astrocytes and
neurons observed in different mouse brain cortical cell cultures. J Neurosci Res 73,
141-155.
Deng, S., Zhang, G.-F., Kasumov, T., Roe, C. R. and Brunengraber, H. (2009) Interrelations
between C4 ketogenesis, C5 ketogenesis, and anaplerosis in the perfused rat liver.
Journal of Biological Chemistry 284, 27799-27807.
Devi, L., Prabhu, B. M., Galati, D. F., Avadhani, N. G. and Anandatheerthavarada, H. K.
(2006) Accumulation of amyloid precursor protein in the mitochondrial import
channels of human Alzheimer’s disease brain is associated with mitochondrial
dysfunction. Journal of Neuroscience 26, 9057-9068.
Di Giacomo, E., Benedetti, E., Cristiano, L. et al. (2017) Roles of PPAR transcription factors
in the energetic metabolic switch occurring during adult neurogenesis. Cell Cycle 16,
59-72.
Dienel, G. A. (2012) Brain lactate metabolism: the discoveries and the controversies. J
Cereb Blood Flow Metab 32, 1107-1138.
Dienel, G. A. and Cruz, N. F. (2015) Contributions of glycogen to astrocytic energetics during
brain activation. Metab Brain Dis 30, 281-298.
Ding, Q., Vaynman, S., Souda, P., Whitelegge, J. P. and Gomez-Pinilla, F. (2006) Exercise
affects energy metabolism and neural plasticity-related proteins in the hippocampus
as revealed by proteomic analysis. Eur J Neurosci 24, 1265-1276.
Dringen, R. and Hamprecht, B. (1992) Glucose, insulin, and insulin-like growth factor I
regulate the glycogen content of astroglia-rich primary cultures. J Neurochem 58,
511-517.
Dringen, R., Hoepken, H., Minich, T. and Ruedig, C. (2007) 1.3 Pentose Phosphate Pathway
and NADPH Metabolism. In: Handbook of neurochemistry and molecular
neurobiology, pp. 41-62. Springer.
Dringen, R., Wiesinger, H. and Hamprecht, B. (1993) Uptake of L-lactate by cultured rat
brain neurons. Neuroscience letters 163, 5-7.
Duarte, J. M. (2015) Metabolic alterations associated to brain dysfunction in diabetes. Aging
and disease 6, 304.
Dupuis, L., Pradat, P.-F., Ludolph, A. C. and Loeffler, J.-P. (2011) Energy metabolism in
amyotrophic lateral sclerosis. The Lancet Neurology 10, 75-82.
Duran, J., Saez, I., Gruart, A., Guinovart, J. J. and Delgado-García, J. M. (2013) Impairment
in long-term memory formation and learning-dependent synaptic plasticity in mice
lacking glycogen synthase in the brain. J Cereb Blood Flow Metab 33, 550-556.
Dworak, M., McCarley, R. W., Kim, T., Kalinchuk, A. V. and Basheer, R. (2010) Sleep and
brain energy levels: ATP changes during sleep. Journal of Neuroscience 30, 9007-
9016.
Ebert, D., Haller, R. G. and Walton, M. E. (2003) Energy contribution of octanoate to intact
rat brain metabolism measured by 13C nuclear magnetic resonance spectroscopy. J
Neurosci 23, 5928-5935.
Edgar, J. M. and Garbern, J. (2004) The myelinated axon is dependent on the myelinating
cell for support and maintenance: molecules involved. J Neurosci Res 76, 593-598.
Edmond, J., Robbins, R. A., Bergstrom, J. D., Cole, R. A. and de Vellis, J. (1987) Capacity
for substrate utilization in oxidative metabolism by neurons, astrocytes, and
oligodendrocytes from developing brain in primary culture. J Neurosci Res 18, 551-
561.
Engl, E. and Attwell, D. (2015) Non-signalling energy use in the brain. The Journal of
physiology 593, 3417-3429.
Engl, E., Jolivet, R., Hall, C. N. and Attwell, D. (2017) Non-signalling energy use in the
developing rat brain. J Cereb Blood Flow Metab 37, 951-966.
Ernst, A., Sharma, A. N., Elased, K. M., Guest, P. C., Rahmoune, H. and Bahn, S. (2013)
Diabetic db/db mice exhibit central nervous system and peripheral molecular
alterations as seen in neurological disorders. Transl Psychiatry 3, e263.
Accepted Article
This article is protected by copyright. All rights reserved.
Exalto, L. G., Whitmer, R. A., Kappele, L. J. and Biessels, G. J. (2012) An update on type 2
diabetes, vascular dementia and Alzheimer's disease. Exp Gerontol 47, 858-864.
Ferraiuolo, L., Higginbottom, A., Heath, P. R., Barber, S., Greenald, D., Kirby, J. and Shaw,
P. J. (2011) Dysregulation of astrocyte–motoneuron cross-talk in mutant superoxide
dismutase 1-related amyotrophic lateral sclerosis. Brain 134, 2627-2641.
Ferris, S. H., de Leon, M. J., Wolf, A. P. et al. (1980) Positron emission tomography in the
study of aging and senile dementia. Neurobiol Aging 1, 127-131.
Fisher, M. (2009) Pericyte signaling in the neurovascular unit. Stroke 40, S13-15.
Fox, P. T., Raichle, M. E., Mintun, M. A. and Dence, C. (1988) Nonoxidative glucose
consumption during focal physiologic neural activity. Science 241, 462-464.
Gaengel, K., Genové, G., Armulik, A. and Betsholtz, C. (2009) Endothelial-mural cell
signaling in vascular development and angiogenesis. Arteriosclerosis, thrombosis,
and vascular biology 29, 630-638.
Garber, A. J., Menzel, P. H., Boden, G. and Owen, O. E. (1974) Hepatic ketogenesis and
gluconeogenesis in humans. The Journal of clinical investigation 54, 981-989.
Gavillet, M., Allaman, I. and Magistretti, P. J. (2008) Modulation of astrocytic metabolic
phenotype by proinflammatory cytokines. Glia 56, 975-989.
Genda, E. N., Jackson, J. G., Sheldon, A. L. et al. (2011) Co-compartmentalization of the
astroglial glutamate transporter, GLT-1, with glycolytic enzymes and mitochondria. J
Neurosci 31, 18275-18288.
Gimeno-Bayón, J., López-López, A., Rodríguez, M. J. and Mahy, N. (2014) Glucose
pathways adaptation supports acquisition of activated microglia phenotype. J
Neurosci Res 92, 723-731.
Glenn, T. C., Kelly, D. F., Boscardin, W. J. et al. (2003) Energy dysfunction as a predictor of
outcome after moderate or severe head injury: indices of oxygen, glucose, and
lactate metabolism. Journal of Cerebral Blood Flow & Metabolism 23, 1239-1250.
Gluckman, P. D., Wyatt, J. S., Azzopardi, D. et al. (2005) Selective head cooling with mild
systemic hypothermia after neonatal encephalopathy: multicentre randomised trial.
The Lancet 365, 663-670.
Golpich, M., Amini, E., Mohamed, Z., Azman Ali, R., Mohamed Ibrahim, N. and Ahmadiani,
A. (2017) Mitochondrial dysfunction and biogenesis in neurodegenerative diseases:
pathogenesis and treatment. CNS neuroscience & therapeutics 23, 5-22.
González, J. A., Iordanidou, P., Strom, M., Adamantidis, A. and Burdakov, D. (2016) Awake
dynamics and brain-wide direct inputs of hypothalamic MCH and orexin networks.
Nat Commun 7, 11395.
Gordon, G. R., Choi, H. B., Rungta, R. L., Ellis-Davies, G. C. and MacVicar, B. A. (2008)
Brain metabolism dictates the polarity of astrocyte control over arterioles. Nature 456,
745-749.
Görlach, A., Bertram, K., Hudecova, S. and Krizanova, O. (2015) Calcium and ROS: a
mutual interplay. Redox biology 6, 260-271.
Goyal, M. S., Vlassenko, A. G., Blazey, T. M. et al. (2017) Loss of brain aerobic glycolysis in
normal human aging. Cell metabolism 26, 353-360. e353.
Guo, X., Macleod, G. T., Wellington, A. et al. (2005) The GTPase dMiro is required for
axonal transport of mitochondria to Drosophila synapses. Neuron 47, 379-393.
Hadera, M. G., Smeland, O. B., McDonald, T. S., Tan, K. N., Sonnewald, U. and Borges, K.
(2014) Triheptanoin partially restores levels of tricarboxylic acid cycle intermediates
in the mouse pilocarpine model of epilepsy. J Neurochem 129, 107-119.
Halestrap, A. P. (2013) The SLC16 gene family - structure, role and regulation in health and
disease. Mol Aspects Med 34, 337-349.
Hall, C. N., Klein-Flügge, M. C., Howarth, C. and Attwell, D. (2012) Oxidative
phosphorylation, not glycolysis, powers presynaptic and postsynaptic mechanisms
underlying brain information processing. Journal of Neuroscience 32, 8940-8951.
Hall, E. D., Andrus, P., Althaus, J. and Von Voigtlander, P. (1993) Hydroxyl radical
production and lipid peroxidation paralles selective postischemic vulnerability in
gerbil brain. Journal of neuroscience research 34, 107-112.
Accepted Article
This article is protected by copyright. All rights reserved.
Hallermann, S., de Kock, C. P., Stuart, G. J. and Kole, M. H. (2012) State and location
dependence of action potential metabolic cost in cortical pyramidal neurons. Nat
Neurosci 15, 1007-1014.
Hamprecht, B., Dringen, R., Pfeiffer, B. and Kurz, G. (1993) The possible roles of astrocytes
in energy metabolism of the brain. In: Biology and Pathology of Astrocyte-Neuron
Interactions, pp. 83-91. Springer.
Han, X., Chen, M., Wang, F. et al. (2013) Forebrain engraftment by human glial progenitor
cells enhances synaptic plasticity and learning in adult mice. Cell Stem Cell 12, 342-
353.
Hansen, J. L. and Freier, E. F. (1978) Direct assays of lactate, pyruvate, beta-
hydroxybutyrate, and acetoacetate with a centrifugal analyzer. Clin Chem 24, 475-
479.
Harris, J. J. and Attwell, D. (2012) The energetics of CNS white matter. Journal of
Neuroscience 32, 356-371.
Harris, J. J., Jolivet, R. and Attwell, D. (2012) Synaptic energy use and supply. Neuron 75,
762-777.
Hasselbalch, S. G., Knudsen, G. M., Jakobsen, J., Hageman, L. P., Holm, S. and Paulson,
O. B. (1994) Brain metabolism during short-term starvation in humans. Journal of
Cerebral Blood Flow & Metabolism 14, 125-131.
Hasselbalch, S. G., Knudsen, G. M., Jakobsen, J., Hageman, L. P., Holm, S. and Paulson,
O. B. (1995) Blood-brain barrier permeability of glucose and ketone bodies during
short-term starvation in humans. The American journal of physiology 268, E1161-
1166.
Hendricks, A. G., Perlson, E., Ross, J. L., Schroeder, H. W., Tokito, M. and Holzbaur, E. L.
(2010) Motor coordination via a tug-of-war mechanism drives bidirectional vesicle
transport. Curr Biol 20, 697-702.
Herculano-Houzel, S. (2011) Scaling of brain metabolism with a fixed energy budget per
neuron: implications for neuronal activity, plasticity and evolution. PloS one 6,
e17514.
Herculano-Houzel, S. and Lent, R. (2005) Isotropic fractionator: a simple, rapid method for
the quantification of total cell and neuron numbers in the brain. Journal of
Neuroscience 25, 2518-2521.
Herrero-Mendez, A., Almeida, A., Fernández, E., Maestre, C., Moncada, S. and Bolaños, J.
P. (2009) The bioenergetic and antioxidant status of neurons is controlled by
continuous degradation of a key glycolytic enzyme by APC/C-Cdh1. Nat Cell Biol 11,
747-752.
Hertz, L. (1979) Functional interactions between neurons and astrocytes I. Turnover and
metabolism of putative amino acid transmitters. Progress in neurobiology 13, 277-
323.
Hertz, L. and Dienel, G. A. (2002) Energy metabolism in the brain. International review of
neurobiology 51, 1-IN4.
Hickey, W. F. and Kimura, H. (1988) Perivascular microglial cells of the CNS are bone
marrow-derived and present antigen in vivo. Science 239, 290-292.
Hill, J., Rom, S., Ramirez, S. H. and Persidsky, Y. (2014) Emerging roles of pericytes in the
regulation of the neurovascular unit in health and disease. J Neuroimmune
Pharmacol 9, 591-605.
Hirrlinger, J. and Dringen, R. (2010) The cytosolic redox state of astrocytes: Maintenance,
regulation and functional implications for metabolite trafficking. Brain Res Rev 63,
177-188.
Hof, P. R., Pascale, E. and Magistretti, P. J. (1988) K+ at concentrations reached in the
extracellular space during neuronal activity promotes a Ca2+-dependent glycogen
hydrolysis in mouse cerebral cortex. J Neurosci 8, 1922-1928.
Horikoshi, Y., Sasaki, A., Taguchi, N., Maeda, M., Tsukagoshi, H., Sato, K. and Yamaguchi,
H. (2003) Human GLUT5 immunolabeling is useful for evaluating microglial status in
neuropathological study using paraffin sections. Acta Neuropathol 105, 157-162.
Accepted Article
This article is protected by copyright. All rights reserved.
Howarth, C., Gleeson, P. and Attwell, D. (2012) Updated energy budgets for neural
computation in the neocortex and cerebellum. J Cereb Blood Flow Metab 32, 1222-
1232.
Howarth, C., Peppiatt-Wildman, C. M. and Attwell, D. (2010) The energy use associated with
neural computation in the cerebellum. J Cereb Blood Flow Metab 30, 403-414.
Hrynevich, S. V., Waseem, T. V., Hébert, A., Pellerin, L. and Fedorovich, S. V. (2016) β-
Hydroxybutyrate supports synaptic vesicle cycling but reduces endocytosis and
exocytosis in rat brain synaptosomes. Neurochemistry international 93, 73-81.
Hyder, F. and Rothman, D. L. (2017) Advances in imaging brain metabolism. Annual Review
of Biomedical Engineering 19, 485-515.
Ide, K., Horn, A. and Secher, N. H. (1999) Cerebral metabolic response to submaximal
exercise. Journal of applied physiology 87, 1604-1608.
Itoh, K., Nakamura, K., Iijima, M. and Sesaki, H. (2013) Mitochondrial dynamics in
neurodegeneration. Trends Cell Biol 23, 64-71.
Ivanisevic, J., Stauch, K. L., Petrascheck, M. et al. (2016) Metabolic drift in the aging brain.
Aging (Albany NY) 8, 1000.
Jerison, H. (2012) Evolution of the brain and intelligence. Elsevier.
JolyAmado, A., Denis, R. G., Castel, J. et al. (2012) Hypothalamic AgRPneurons control
peripheral substrate utilization and nutrient partitioning. The EMBO journal 31, 4276-
4288.
Kalsbeek, M. J., Mulder, L. and Yi, C. X. (2016) Microglia energy metabolism in metabolic
disorder. Mol Cell Endocrinol 438, 27-35.
Karbowski, J. (2009) Thermodynamic constraints on neural dimensions, firing rates, brain
temperature and size. Journal of computational neuroscience 27, 415-436.
Karelina, K. and Weil, Z. M. (2016) Neuroenergetics of traumatic brain injury. Concussion 1.
Kettenmann, H., Hanisch, U.-K., Noda, M. and Verkhratsky, A. (2011) Physiology of
microglia. Physiological reviews 91, 461-553.
Khanna, S., Roy, S., Ryu, H., Bahadduri, P., Swaan, P. W., Ratan, R. R. and Sen, C. K.
(2003) Molecular basis of vitamin E action tocotrienol modulates 12-lipoxygenase, a
key mediator of glutamate-induced neurodegeneration. Journal of Biological
Chemistry 278, 43508-43515.
Kim, B., Backus, C., Oh, S., Hayes, J. M. and Feldman, E. L. (2009) Increased tau
phosphorylation and cleavage in mouse models of type 1 and type 2 diabetes.
Endocrinology 150, 5294-5301.
Kim, S. K., Nabekura, J. and Koizumi, S. (2017) Astrocyte-mediated synapse remodeling in
the pathological brain. Glia 65, 1719-1727.
Kinman, R. P., Kasumov, T., Jobbins, K. A. et al. (2006) Parenteral and enteral metabolism
of anaplerotic triheptanoin in normal rats. Am J Physiol Endocrinol Metab 291, E860-
866.
Kirby, J., Halligan, E., Baptista, M. J. et al. (2005) Mutant SOD1 alters the motor neuronal
transcriptome: implications for familial ALS. Brain 128, 1686-1706.
Kitchen, P., Day, R. E., Salman, M. M., Conner, M. T., Bill, R. M. and Conner, A. C. (2015a)
Beyond water homeostasis: Diverse functional roles of mammalian aquaporins.
Biochim Biophys Acta 1850, 2410-2421.
Kitchen, P., Day, R. E., Taylor, L. H., Salman, M. M., Bill, R. M., Conner, M. T. and Conner,
A. C. (2015b) Identification and molecular mechanisms of the rapid tonicity-induced
relocalization of the aquaporin 4 channel. Journal of Biological Chemistry 290,
16873-16881.
Koga, M., Serritella, A. V., Messmer, M. M., Hayashi-Takagi, A., Hester, L. D., Snyder, S. H.,
Sawa, A. and Sedlak, T. W. (2011) Glutathione is a physiologic reservoir of neuronal
glutamate. Biochemical and biophysical research communications 409, 596-602.
Kole, M. H., Ilschner, S. U., Kampa, B. M., Williams, S. R., Ruben, P. C. and Stuart, G. J.
(2008) Action potential generation requires a high sodium channel density in the
axon initial segment. Nat Neurosci 11, 178-186.
Accepted Article
This article is protected by copyright. All rights reserved.
Koss, K., Churchward, M. A., Tsui, C. and Todd, K. G. (2019) In Vitro Priming and Hyper-
Activation of Brain Microglia: an Assessment of Phenotypes. Molecular neurobiology,
1-17.
Kuge, Y., Yajima, K., Kawashima, H., Yamazaki, H., Hashimoto, N. and Miyake, Y. (1995)
Brain uptake and metabolism of [1-11C]octanoate in rats: pharmacokinetic basis for
its application as a radiopharmaceutical for studying brain fatty acid metabolism. Ann
Nucl Med 9, 137-142.
Kuzawa, C. W., Chugani, H. T., Grossman, L. I. et al. (2014a) Metabolic costs and
evolutionary implications of human brain development. Proceedings of the National
Academy of Sciences 111, 13010-13015.
Kuzawa, C. W., Chugani, H. T., Grossman, L. I. et al. (2014b) Reply to Skoyles: Decline in
growth rate, not muscle mass, predicts the human childhood peak in brain
metabolism. Proceedings of the National Academy of Sciences 111, E4910-E4910.
Laake, J. H., Takumi, Y., Eidet, J., Torgner, I. A., Roberg, B., Kvamme, E. and Ottersen, O.
P. (1999) Postembedding immunogold labelling reveals subcellular localization and
pathway-specific enrichment of phosphate activated glutaminase in rat cerebellum.
Neuroscience 88, 1137-1151.
Lam, C. K., Chari, M. and Lam, T. K. (2009) CNS regulation of glucose homeostasis.
Physiology 24, 159-170.
Lennie, P. (2003) The cost of cortical computation. Curr Biol 13, 493-497.
Leonard, W. R., Robertson, M. L., Snodgrass, J. J. and Kuzawa, C. W. (2003) Metabolic
correlates of hominid brain evolution. Comparative Biochemistry and Physiology Part
A: Molecular & Integrative Physiology 136, 5-15.
Lin, A. L., Fox, P. T., Hardies, J., Duong, T. Q. and Gao, J. H. (2010) Nonlinear coupling
between cerebral blood flow, oxygen consumption, and ATP production in human
visual cortex. Proceedings of the National Academy of Sciences of the United States
of America 107, 8446-8451.
Lin, M. T. and Beal, M. F. (2006) Mitochondrial dysfunction and oxidative stress in
neurodegenerative diseases. Nature 443, 787.
Lo, E. H., Dalkara, T. and Moskowitz, M. A. (2003) Neurological diseases: Mechanisms,
challenges and opportunities in stroke. Nature reviews neuroscience 4, 399.
Ludwin, S. K. (1997) The pathobiology of the oligodendrocyte. Journal of Neuropathology &
Experimental Neurology 56, 111-124.
Luo, L. (2002) Actin cytoskeleton regulation in neuronal morphogenesis and structural
plasticity. Annu Rev Cell Dev Biol 18, 601-635.
Ly, C. V. and Verstreken, P. (2006) Mitochondria at the synapse. The Neuroscientist 12,
291-299.
Maday, S., Twelvetrees, A. E., Moughamian, A. J. and Holzbaur, E. L. (2014) Axonal
transport: cargo-specific mechanisms of motility and regulation. Neuron 84, 292-309.
Magistretti, P. J. (2006) Neuron-glia metabolic coupling and plasticity. J Exp Biol 209, 2304-
2311.
Magistretti, P. J., Morrison, J. H., Shoemaker, W. J., Sapin, V. and Bloom, F. E. (1981)
Vasoactive intestinal polypeptide induces glycogenolysis in mouse cortical slices: a
possible regulatory mechanism for the local control of energy metabolism.
Proceedings of the National Academy of Sciences 78, 6535-6539.
Magistretti, P. J., Pellerin, L., Rothman, D. L. and Shulman, R. G. (1999) Energy on demand.
Science 283, 496-497.
Maher, F., Vannucci, S. J. and Simpson, I. A. (1994) Glucose transporter proteins in brain.
Faseb J 8, 1003-1011.
Makar, T. K., Hungund, B. L., Cook, G. A., Kashfi, K. and Cooper, A. J. (1995) Lipid
metabolism and membrane composition are altered in the brains of type II diabetic
mice. J Neurochem 64, 2159-2168.
Malik, R. and Dichgans, M. (2018) Challenges and opportunities in stroke genetics.
Cardiovascular research 114, 1226-1240.
Accepted Article
This article is protected by copyright. All rights reserved.
Mao, W., Yu, X. X., Zhong, A., Li, W., Brush, J., Sherwood, S. W., Adams, S. H. and Pan, G.
(1999) UCP4, a novel brain-specific mitochondrial protein that reduces membrane
potential in mammalian cells. FEBS letters 443, 326-330.
Maquet, P., Dive, D., Salmon, E., Sadzot, B., Franco, G., Poirrier, R., von Frenckell, R. and
Franck, G. (1990) Cerebral glucose utilization during sleep-wake cycle in man
determined by positron emission tomography and [18F]2-fluoro-2-deoxy-D-glucose
method. Brain Res 513, 136-143.
Marcus, C., Mena, E. and Subramaniam, R. M. (2014) Brain PET in the diagnosis of
Alzheimer’s disease. Clinical nuclear medicine 39, e413.
Margolis, R. L. (1981) Role of GTP hydrolysis in microtubule treadmilling and assembly.
Proceedings of the National Academy of Sciences of the United States of America
78, 1586-1590.
Marin-Valencia, I., Good, L. B., Ma, Q., Malloy, C. R. and Pascual, J. M. (2013) Heptanoate
as a neural fuel: energetic and neurotransmitter precursors in normal and glucose
transporter I-deficient (G1D) brain. J Cereb Blood Flow Metab 33, 175-182.
Marino, L. (1998) A comparison of encephalization between odontocete cetaceans and
anthropoid primates. Brain, Behavior and Evolution 51, 230-238.
Marion, D. W., Darby, J. and Yonas, H. (1991) Acute regional cerebral blood flow changes
caused by severe head injuries. Journal of neurosurgery 74, 407-414.
Markram, H., Lübke, J., Frotscher, M., Roth, A. and Sakmann, B. (1997) Physiology and
anatomy of synaptic connections between thick tufted pyramidal neurones in the
developing rat neocortex. The Journal of physiology 500 ( Pt 2), 409-440.
Mason, S. (2017) Lactate shuttles in neuroenergetics—homeostasis, allostasis and beyond.
Frontiers in neuroscience 11, 43.
Mathiisen, T. M., Lehre, K. P., Danbolt, N. C. and Ottersen, O. P. (2010) The perivascular
astroglial sheath provides a complete covering of the brain microvessels: an electron
microscopic 3D reconstruction. Glia 58, 1094-1103.
McDonald, T. S., Tan, K. N., Hodson, M. P. and Borges, K. (2014) Alterations of
hippocampal glucose metabolism by even versus uneven medium chain triglycerides.
J Cereb Blood Flow Metab 34, 153-160.
McEwen, B. S. and Reagan, L. P. (2004) Glucose transporter expression in the central
nervous system: relationship to synaptic function. European journal of pharmacology
490, 13-24.
McKenna, M. C. (2012) Substrate competition studies demonstrate oxidative metabolism of
glucose, glutamate, glutamine, lactate and 3-hydroxybutyrate in cortical astrocytes
from rat brain. Neurochem Res 37, 2613-2626.
McKenna, M. C. (2013) Glutamate pays its own way in astrocytes. Frontiers in endocrinology
4, 191.
McKenna, M. C., Dienel, G. A., Sonnewald, U., Waagepetersen, H. S. and Schousboe, A.
(2012) Energy metabolism of the brain. In: Basic Neurochemistry (Eighth Edition), pp.
200-231. Elsevier.
McKenna, M. C., Tildon, J. T., Stevenson, J. H., Boatright, R. and Huang, S. (1993)
Regulation of energy metabolism in synaptic terminals and cultured rat brain
astrocytes: differences revealed using aminooxyacetate. Dev Neurosci 15, 320-329.
McLaurin, J. and Yong, V. W. (1995) Oligodendrocytes and myelin. Neurologic clinics 13,
23-49.
Mink, J. W., Blumenschine, R. J. and Adams, D. B. (1981) Ratio of central nervous system
to body metabolism in vertebrates: its constancy and functional basis. The American
journal of physiology 241, R203-212.
Miquel, E., Cassina, A., Martínez-Palma, L., Bolatto, C., Trías, E., Gandelman, M., Radi, R.,
Barbeito, L. and Cassina, P. (2012) Modulation of astrocytic mitochondrial function by
dichloroacetate improves survival and motor performance in inherited amyotrophic
lateral sclerosis. PloS one 7, e34776.
Accepted Article
This article is protected by copyright. All rights reserved.
Mitchell, G. A., Kassovska-Bratinova, S., Boukaftane, Y., Robert, M. F., Wang, S. P.,
Ashmarina, L., Lambert, M., Lapierre, P. and Potier, E. (1995) Medical aspects of
ketone body metabolism. Clin Invest Med 18, 193-216.
Miyazaki, K., Masamoto, K., Morimoto, N., Kurata, T., Mimoto, T., Obata, T., Kanno, I. and
Abe, K. (2012) Early and progressive impairment of spinal blood flow—glucose
metabolism coupling in motor neuron degeneration of ALS model mice. Journal of
Cerebral Blood Flow & Metabolism 32, 456-467.
Mosconi, L. (2013) Glucose metabolism in normal aging and Alzheimer's disease:
Methodological and physiological considerations for PET studies. Clin Transl Imaging
1.
Mosconi, L., Pupi, A. and De Leon, M. J. (2008) Brain glucose hypometabolism and
oxidative stress in preclinical Alzheimer's disease. Ann N Y Acad Sci 1147, 180-195.
Murphy, T. H., Schnaar, R. L. and Coyle, J. T. (1990) Immature cortical neurons are uniquely
sensitive to glutamate toxicity by inhibition of cystine uptake. Faseb J 4, 1624-1633.
Navarrete, A., van Schaik, C. P. and Isler, K. (2011) Energetics and the evolution of human
brain size. Nature 480, 91.
Navarro, A. and Boveris, A. (2010) Brain mitochondrial dysfunction in aging,
neurodegeneration, and Parkinson's disease. Frontiers in aging neuroscience 2, 34.
Nave, K.-A. (2010) Myelination and support of axonal integrity by glia. Nature 468, 244.
Nave, K.-A. and Werner, H. B. (2014) Myelination of the nervous system: mechanisms and
functions. Annual review of cell and developmental biology 30, 503-533.
Nawroth, J. C., Greer, C. A., Chen, W. R., Laughlin, S. B. and Shepherd, G. M. (2007) An
energy budget for the olfactory glomerulus. J Neurosci 27, 9790-9800.
Nedergaard, M., Ransom, B. and Goldman, S. A. (2003) New roles for astrocytes: redefining
the functional architecture of the brain. Trends Neurosci 26, 523-530.
Newman, L. A., Korol, D. L. and Gold, P. E. (2011) Lactate produced by glycogenolysis in
astrocytes regulates memory processing. PloS one 6, e28427.
Niethard, N., Hasegawa, M., Itokazu, T., Oyanedel, C. N., Born, J. and Sato, T. R. (2016)
Sleep-stage-specific regulation of cortical excitation and inhibition. Current biology
26, 2739-2749.
Norenberg, M. D. and Martinez-Hernandez, A. (1979) Fine structural localization of
glutamine synthetase in astrocytes of rat brain. Brain Res 161, 303-310.
Nunomura, A., Perry, G., Aliev, G. et al. (2001) Oxidative damage is the earliest event in
Alzheimer disease. J Neuropathol Exp Neurol 60, 759-767.
Nyhan, W. L. (2005) Disorders of purine and pyrimidine metabolism. Molecular genetics and
metabolism 86, 25-33.
Oberheim, N. A., Wang, X., Goldman, S. and Nedergaard, M. (2006) Astrocytic complexity
distinguishes the human brain. Trends Neurosci 29, 547-553.
Oldendorf, W. H. (1973) Carrier-mediated blood-brain barrier transport of short-chain
monocarboxylic organic acids. American Journal of Physiology-Legacy Content 224,
1450-1453.
Ott, A., Stolk, R. P., Hofman, A., van Harskamp, F., Grobbee, D. E. and Breteler, M. M.
(1996) Association of diabetes mellitus and dementia: the Rotterdam Study.
Diabetologia 39, 1392-1397.
Owen, O., Morgan, A., Kemp, H., Sullivan, J., Herrera, M. and Cahill, G. J. (1967) Brain
metabolism during fasting. The Journal of clinical investigation 46, 1589-1595.
Palay, S. L., Sotelo, C., Peters, A. and Orkand, P. M. (1968) The axon hillock and the initial
segment. The Journal of cell biology 38, 193-201.
Pan, J. W., Rothman, T. L., Behar, K. L., Stein, D. T. and Hetherington, H. P. (2000) Human
brain beta-hydroxybutyrate and lactate increase in fasting-induced ketosis. J Cereb
Blood Flow Metab 20, 1502-1507.
Park, S. J., Shin, J. H., Jeong, J. I. et al. (2014) Down-regulation of mortalin exacerbates Aβ-
mediated mitochondrial fragmentation and dysfunction. Journal of Biological
Chemistry 289, 2195-2204.
Accepted Article
This article is protected by copyright. All rights reserved.
Patet, C., Suys, T., Carteron, L. and Oddo, M. (2016) Cerebral lactate metabolism after
traumatic brain injury. Current neurology and neuroscience reports 16, 31.
Payne, J., Maher, F., Simpson, I., Mattice, L. and Davies, P. (1997) Glucose transporter Glut
5 expression in microglial cells. Glia 21, 327-331.
Pellegri, G., Rossier, C., Magistretti, P. J. and Martin, J. L. (1996) Cloning, localization and
induction of mouse brain glycogen synthase. Brain Res Mol Brain Res 38, 191-199.
Pellerin, L., Bergersen, L. H., Halestrap, A. P. and Pierre, K. (2005) Cellular and subcellular
distribution of monocarboxylate transporters in cultured brain cells and in the adult
brain. J Neurosci Res 79, 55-64.
Pellerin, L. and Magistretti, P. J. (1994) Glutamate uptake into astrocytes stimulates aerobic
glycolysis: a mechanism coupling neuronal activity to glucose utilization. Proceedings
of the National Academy of Sciences of the United States of America 91, 10625-
10629.
Pellerin, L. and Magistretti, P. J. (2004) Neuroenergetics: calling upon astrocytes to satisfy
hungry neurons. Neuroscientist 10, 53-62.
Pellerin, L., Pellegri, G., Bittar, P. G., Charnay, Y., Bouras, C., Martin, J.-L., Stella, N. and
Magistretti, P. J. (1998) Evidence supporting the existence of an activity-dependent
astrocyte-neuron lactate shuttle. Developmental neuroscience 20, 291-299.
Pennisi, G., Cornelius, C., Cavallaro, M. et al. (2011) Redox regulation of cellular stress
response in multiple sclerosis. Biochemical Pharmacology 82, 1490-1499.
Persson, B. (1970) Determination of plasma acetoacetate and D-beta-hydroxybutyrate in
new-born infants by an enzymatic fluorometric micro-method. Scand J Clin Lab
Invest 25, 9-18.
Peters, A. (1966) The node of Ranvier in the central nervous system. Quarterly journal of
experimental physiology and cognate medical sciences 51, 229-236.
Petit, J.-M. and Magistretti, P. J. (2016) Regulation of neuron–astrocyte metabolic coupling
across the sleep–wake cycle. Neuroscience 323, 135-156.
Petrova, T. V., Karpanen, T., Norrmén, C. et al. (2004) Defective valves and abnormal mural
cell recruitment underlie lymphatic vascular failure in lymphedema distichiasis.
Nature medicine 10, 974-981.
Pfefferle, A. D., Warner, L. R., Wang, C. W., Nielsen, W. J., Babbitt, C. C., Fedrigo, O. and
Wray, G. A. (2011) Comparative expression analysis of the phosphocreatine circuit in
extant primates: Implications for human brain evolution. Journal of human evolution
60, 205-212.
Philips, T. and Rothstein, J. D. (2017) Oligodendroglia: metabolic supporters of neurons. The
Journal of clinical investigation 127, 3271-3280.
Pierre, K., Parent, A., Jayet, P. Y., Halestrap, A. P., Scherrer, U. and Pellerin, L. (2007)
Enhanced expression of three monocarboxylate transporter isoforms in the brain of
obese mice. The Journal of physiology 583, 469-486.
Pierre, K. and Pellerin, L. (2005) Monocarboxylate transporters in the central nervous
system: distribution, regulation and function. J Neurochem 94, 1-14.
Poplawski, M. M., Mastaitis, J. W., Isoda, F., Grosjean, F., Zheng, F. and Mobbs, C. V.
(2011) Reversal of diabetic nephropathy by a ketogenic diet. PloS one 6, e18604.
Purdon, A. and Rapoport, S. (2007) 4.6 Energy consumption by phospholipid metabolism in
mammalian brain. In: Handbook of Neurochemistry and Molecular Neurobiology, pp.
401-427. Springer.
PURDON, A. D. and RAPOPORT, S. I. (1998) Energy requirements for two aspects of
phospholipid metabolism in mammalian brain. Biochemical Journal 335, 313-318.
Quistorff, B., Secher, N. H. and Van Lieshout, J. J. (2008) Lactate fuels the human brain
during exercise. Faseb J 22, 3443-3449.
Ramos-Rodriguez, J. J., Ortiz, O., Jimenez-Palomares, M. et al. (2013) Differential central
pathology and cognitive impairment in pre-diabetic and diabetic mice.
Psychoneuroendocrinology 38, 2462-2475.
Rangaraju, V., Calloway, N. and Ryan, T. A. (2014) Activity-driven local ATP synthesis is
required for synaptic function. Cell 156, 825-835.
Accepted Article
This article is protected by copyright. All rights reserved.
Ransom, B. R. and Sontheimer, H. (1992) The neurophysiology of glial cells. Journal of
clinical neurophysiology: official publication of the American Electroencephalographic
Society 9, 224-251.
Rinholm, J. E., Hamilton, N. B., Kessaris, N., Richardson, W. D., Bergersen, L. H. and
Attwell, D. (2011) Regulation of oligodendrocyte development and myelination by
glucose and lactate. J Neurosci 31, 538-548.
Rink, C., Gnyawali, S., Peterson, L. and Khanna, S. (2011) Oxygen-inducible glutamate
oxaloacetate transaminase as protective switch transforming neurotoxic glutamate to
metabolic fuel during acute ischemic stroke. Antioxid Redox Signal 14, 1777-1785.
Rolfe, D. F. and Brown, G. C. (1997) Cellular energy utilization and molecular origin of
standard metabolic rate in mammals. Physiological reviews 77, 731-758.
Rolfe, D. F., Hulbert, A. J. and Brand, M. D. (1994) Characteristics of mitochondrial proton
leak and control of oxidative phosphorylation in the major oxygen-consuming tissues
of the rat. Biochimica et biophysica acta 1188, 405-416.
Rothstein, J. D. (1995) Excitotoxic mechanisms in the pathogenesis of amyotrophic lateral
sclerosis. Advances in neurology 68, 7-20; discussion 21-27.
Rucker, H. K., Wynder, H. J. and Thomas, W. E. (2000) Cellular mechanisms of CNS
pericytes. Brain research bulletin 51, 363-369.
Ruminot, I., Gutiérrez, R., Peña-Münzenmayer, G., Añazco, C., Sotelo-Hitschfeld, T.,
Lerchundi, R., Niemeyer, M. I., Shull, G. E. and Barros, L. F. (2011) NBCe1 mediates
the acute stimulation of astrocytic glycolysis by extracellular K+. Journal of
Neuroscience 31, 14264-14271.
Ruminot, I., Schmälzle, J., Leyton, B., Barros, L. F. and Deitmer, J. W. (2017) Tight coupling
of astrocyte energy metabolism to synaptic activity revealed by genetically encoded
FRET nanosensors in hippocampal tissue. Journal of Cerebral Blood Flow &
Metabolism, 0271678X17737012.
Rustenhoven, J., Jansson, D., Smyth, L. C. and Dragunow, M. (2017) Brain pericytes as
mediators of neuroinflammation. Trends in pharmacological sciences 38, 291-304.
Saab, A. S., Tzvetanova, I. D. and Nave, K. A. (2013) The role of myelin and
oligodendrocytes in axonal energy metabolism. Curr Opin Neurobiol 23, 1065-1072.
Salman, M. M., Kitchen, P., Woodroofe, M. N., Bill, R. M., Conner, A. C., Heath, P. R. and
Conner, M. T. (2017a) Transcriptome Analysis of Gene Expression Provides New
Insights into the Effect of Mild Therapeutic Hypothermia on Primary Human Cortical
Astrocytes Cultured under Hypoxia. Front Cell Neurosci 11, 386.
Salman, M. M., Kitchen, P., Woodroofe, M. N., Brown, J. E., Bill, R. M., Conner, A. C. and
Conner, M. T. (2017b) Hypothermia increases aquaporin 4 (AQP4) plasma
membrane abundance in human primary cortical astrocytes via a calcium/transient
receptor potential vanilloid 4 (TRPV4)and calmodulinmediated mechanism.
European Journal of Neuroscience.
Sanchis, D., Fleury, C., Chomiki, N. et al. (1998) BMCP1, a novel mitochondrial carrier with
high expression in the central nervous system of humans and rodents, and
respiration uncoupling activity in recombinant yeast. The Journal of biological
chemistry 273, 34611-34615.
Saravanan, K. S., Sindhu, K. M., Senthilkumar, K. S. and Mohanakumar, K. P. (2006) L-
deprenyl protects against rotenone-induced, oxidative stress-mediated dopaminergic
neurodegeneration in rats. Neurochemistry international 49, 28-40.
Saudubray, J. M., Marsac, C., Limal, J. M., Dumurgier, E., Charpentier, C., Ogier, H. and
Coudè, F. X. (1981) Variation in plasma ketone bodies during a 24-hour fast in
normal and in hypoglycemic children: relationship to age. J Pediatr 98, 904-908.
Schapira, A. H. and Jenner, P. (2011) Etiology and pathogenesis of Parkinson's disease.
Movement disorders 26, 1049-1055.
Schönfeld, P. and Wojtczak, L. (2016) Short- and medium-chain fatty acids in energy
metabolism: the cellular perspective. J Lipid Res 57, 943-954.
Schousboe, A. (1981) Transport and metabolism of glutamate and GABA in neurons and
glial cells. In: International review of neurobiology, Vol. 22, pp. 1-45. Elsevier.
Accepted Article
This article is protected by copyright. All rights reserved.
Schousboe, A., Scafidi, S., Bak, L. K., Waagepetersen, H. S. and McKenna, M. C. (2014)
Glutamate metabolism in the brain focusing on astrocytes. Adv Neurobiol 11, 13-30.
Schwarzkopf, T. M., Koch, K. and Klein, J. (2015) Reduced severity of ischemic stroke and
improvement of mitochondrial function after dietary treatment with the anaplerotic
substance triheptanoin. Neuroscience 300, 201-209.
Sengupta, S., Peterson, T. R., Laplante, M., Oh, S. and Sabatini, D. M. (2010) mTORC1
controls fasting-induced ketogenesis and its modulation by ageing. Nature 468,
1100-1104.
Shankaran, S., Laptook, A. R., Ehrenkranz, R. A. et al. (2005) Whole-body hypothermia for
neonates with hypoxic–ischemic encephalopathy. New England Journal of Medicine
353, 1574-1584.
Sickmann, H. M., Walls, A. B., Schousboe, A., Bouman, S. D. and Waagepetersen, H. S.
(2009) Functional significance of brain glycogen in sustaining glutamatergic
neurotransmission. J Neurochem 109 Suppl 1, 80-86.
Silver, R. A., Cull-Candy, S. G. and Takahashi, T. (1996) Non-NMDA glutamate receptor
occupancy and open probability at a rat cerebellar synapse with single and multiple
release sites. The Journal of physiology 494 ( Pt 1), 231-250.
Simard, M. and Nedergaard, M. (2004) The neurobiology of glia in the context of water and
ion homeostasis. Neuroscience 129, 877-896.
Simpson, I. A., Chundu, K. R., Davies-Hill, T., Honer, W. G. and Davies, P. (1994)
Decreased concentrations of GLUT1 and GLUT3 glucose transporters in the brains
of patients with Alzheimer's disease. Ann Neurol 35, 546-551.
Sims-Robinson, C., Kim, B., Rosko, A. and Feldman, E. L. (2010) How does diabetes
accelerate Alzheimer disease pathology? Nature Reviews Neurology 6, 551.
Skoyles, J. R. (2014) Skeletal muscle-induced hypoglycemia risk, not life history energy
trade-off, links high child brain glucose use to slow body growth. Proceedings of the
National Academy of Sciences of the United States of America 111, E4909.
Smith, D., Pernet, A., Hallett, W. A., Bingham, E., Marsden, P. K. and Amiel, S. A. (2003)
Lactate: a preferred fuel for human brain metabolism in vivo. J Cereb Blood Flow
Metab 23, 658-664.
Sokoloff, L., Reivich, M., Kennedy, C., Rosiers, M. D., Patlak, C. S., Pettigrew, K., et al.,
Sakurada, O. and Shinohara, M. (1977) The [14C] deoxyglucose method for the
measurement of local cerebral glucose utilization: theory, procedure, and normal
values in the conscious and anesthetized albino rat 1. Journal of neurochemistry 28,
897-916.
Soppina, V., Rai, A. K., Ramaiya, A. J., Barak, P. and Mallik, R. (2009) Tug-of-war between
dissimilar teams of microtubule motors regulates transport and fission of endosomes.
Proceedings of the National Academy of Sciences of the United States of America
106, 19381-19386.
Spruston, N., Jonas, P. and Sakmann, B. (1995) Dendritic glutamate receptor channels in rat
hippocampal CA3 and CA1 pyramidal neurons. The Journal of physiology 482 ( Pt
2), 325-352.
Steiner, J. L., Murphy, E. A., McClellan, J. L., Carmichael, M. D. and Davis, J. M. (2011)
Exercise training increases mitochondrial biogenesis in the brain. Journal of applied
physiology 111, 1066-1071.
Stratman, A. N., Malotte, K. M., Mahan, R. D., Davis, M. J. and Davis, G. E. (2009) Pericyte
recruitment during vasculogenic tube assembly stimulates endothelial basement
membrane matrix formation. Blood 114, 5091-5101.
Stratman, A. N., Schwindt, A. E., Malotte, K. M. and Davis, G. E. (2010) Endothelial-derived
PDGF-BB and HB-EGF coordinately regulate pericyte recruitment during
vasculogenic tube assembly and stabilization. Blood 116, 4720-4730.
Suh, S. W., Bergher, J. P., Anderson, C. M., Treadway, J. L., Fosgerau, K. and Swanson, R.
A. (2007) Astrocyte glycogen sustains neuronal activity during hypoglycemia: studies
with the glycogen phosphorylase inhibitor CP-316,819 ([RR*, S*]-5-chloro-N-[2-
Accepted Article
This article is protected by copyright. All rights reserved.
hydroxy-3-(methoxymethylamino)-3-oxo-1-(phenylmethyl) propyl]-1H-indole-2-
carboxamide). Journal of Pharmacology and Experimental Therapeutics 321, 45-50.
Surmeier, D. J., Guzman, J. N., Sanchez, J. and Schumacker, P. T. (2012) Physiological
phenotype and vulnerability in Parkinson’s disease. Cold Spring Harbor perspectives
in medicine, a009290.
Suzuki, A., Stern, S. A., Bozdagi, O., Huntley, G. W., Walker, R. H., Magistretti, P. J. and
Alberini, C. M. (2011) Astrocyte-neuron lactate transport is required for long-term
memory formation. Cell 144, 810-823.
Sveinbjornsdottir, S. (2016) The clinical symptoms of Parkinson's disease. Journal of
neurochemistry 139, 318-324.
Swanson, R. A. (1992) Physiologic coupling of glial glycogen metabolism to neuronal activity
in brain. Can J Physiol Pharmacol 70 Suppl, S138-144.
Swanson, R. A. and Choi, D. W. (1993) Glial glycogen stores affect neuronal survival during
glucose deprivation in vitro. J Cereb Blood Flow Metab 13, 162-169.
Sweeney, M. D., Ayyadurai, S. and Zlokovic, B. V. (2016) Pericytes of the neurovascular
unit: key functions and signaling pathways. Nat Neurosci 19, 771-783.
Swerdlow, R. H. and Khan, S. M. (2004) A “mitochondrial cascade hypothesis” for sporadic
Alzheimer's disease. Medical hypotheses 63, 8-20.
Tabernero, A., Medina, J. M. and Giaume, C. (2006) Glucose metabolism and proliferation in
glia: role of astrocytic gap junctions. J Neurochem 99, 1049-1061.
Takano, T., Tian, G. F., Peng, W., Lou, N., Libionka, W., Han, X. and Nedergaard, M. (2006)
Astrocyte-mediated control of cerebral blood flow. Nat Neurosci 9, 260-267.
Tan, K. N., Carrasco-Pozo, C., McDonald, T. S., Puchowicz, M. and Borges, K. (2017)
Tridecanoin is anticonvulsant, antioxidant, and improves mitochondrial function. J
Cereb Blood Flow Metab 37, 2035-2048.
Tefera, T. W. and Borges, K. (2017) Metabolic dysfunctions in amyotrophic lateral sclerosis
pathogenesis and potential metabolic treatments. Frontiers in neuroscience 10, 611.
Tefera, T. W., Wong, Y., Barkl-Luke, M. E., Ngo, S. T., Thomas, N. K., McDonald, T. S. and
Borges, K. (2016) Triheptanoin protects motor neurons and delays the onset of motor
symptoms in a mouse model of amyotrophic lateral sclerosis. PloS one 11,
e0161816.
Thevenet, J., De Marchi, U., Domingo, J. S. et al. (2016) Medium-chain fatty acids inhibit
mitochondrial metabolism in astrocytes promoting astrocyte-neuron lactate and
ketone body shuttle systems. Faseb J 30, 1913-1926.
Thomas, W. E. (1999) Brain macrophages: on the role of pericytes and perivascular cells.
Brain Research Reviews 31, 42-57.
Tong, J. J. (2007) Mitochondrial delivery is essential for synaptic potentiation. Biol Bull 212,
169-175.
Trotta, N., Rodesch, C. K., Fergestad, T. and Broadie, K. (2004) Cellular bases of activity-
dependent paralysis in Drosophila stress-sensitive mutants. J Neurobiol 60, 328-347.
van der Merwe, C., van Dyk, H. C., Engelbrecht, L., van der Westhuizen, F. H., Kinnear, C.,
Loos, B. and Bardien, S. (2017) Curcumin rescues a PINK1 knock down SH-SY5Y
cellular model of Parkinson’s disease from mitochondrial dysfunction and cell death.
Molecular neurobiology 54, 2752-2762.
Vannucci, S. J., Gibbs, E. M. and Simpson, I. A. (1997) Glucose utilization and glucose
transporter proteins GLUT-1 and GLUT-3 in brains of diabetic (db/db) mice. The
American journal of physiology 272, E267-274.
Ventura, R. and Harris, K. M. (1999) Three-dimensional relationships between hippocampal
synapses and astrocytes. J Neurosci 19, 6897-6906.
Verstreken, P., Ly, C. V., Venken, K. J., Koh, T. W., Zhou, Y. and Bellen, H. J. (2005)
Synaptic mitochondria are critical for mobilization of reserve pool vesicles at
Drosophila neuromuscular junctions. Neuron 47, 365-378.
Vezzani, B., Pierantozzi, E. and Sorrentino, V. (2016) Not All Pericytes Are Born Equal:
Pericytes from Human Adult Tissues Present Different Differentiation Properties.
Stem Cells Dev.
Accepted Article
This article is protected by copyright. All rights reserved.
Villeneuve, L. M., Purnell, P. R., Boska, M. D. and Fox, H. S. (2016) Early expression of
Parkinson’s disease-related mitochondrial abnormalities in PINK1 knockout rats.
Molecular neurobiology 53, 171-186.
Waagepetersen, H. S., Sonnewald, U. and Schousboe, A. (1999) The GABA paradox:
multiple roles as metabolite, neurotransmitter, and neurodifferentiative agent. J
Neurochem 73, 1335-1342.
Walls, A. B., Heimbürger, C. M., Bouman, S. D., Schousboe, A. and Waagepetersen, H. S.
(2009) Robust glycogen shunt activity in astrocytes: Effects of glutamatergic and
adrenergic agents. Neuroscience 158, 284-292.
Wang, D. and Mitchell, E. S. (2016) Cognition and Synaptic-Plasticity Related Changes in
Aged Rats Supplemented with 8- and 10-Carbon Medium Chain Triglycerides. PloS
one 11, e0160159.
Wang, F., Smith, N. A., Xu, Q., Fujita, T., Baba, A., Matsuda, T., Takano, T., Bekar, L. and
Nedergaard, M. (2012) Astrocytes modulate neural network activity by Ca²+-
dependent uptake of extracellular K+. Sci Signal 5, ra26.
Wang, L., Pavlou, S., Du, X., Bhuckory, M., Xu, H. and Chen, M. (2019) Glucose transporter
1 critically controls microglial activation through facilitating glycolysis. Molecular
neurodegeneration 14, 2.
Wardelmann, K., Blümel, S., Rath, M. et al. (2019) Insulin action in the brain regulates
mitochondrial stress responses and reduces diet-induced weight gain. Molecular
metabolism.
Wegner, A. (1976) Head to tail polymerization of actin. Journal of molecular biology 108,
139-150.
Whitelaw, B. S. and Robinson, M. B. (2013) Inhibitors of glutamate dehydrogenase block
sodium-dependent glutamate uptake in rat brain membranes. Front Endocrinol
(Lausanne) 4, 123.
Wieraszko, A. (1982) Changes in the hippocampal slices energy metabolism following
stimulation and long-term potentiation of Schaffer collaterals-pyramidal cell synapses
tested with the 2-deoxyglucose technique. Brain Res 237, 449-457.
Willis, S., Stoll, J., Sweetman, L. and Borges, K. (2010) Anticonvulsant effects of a
triheptanoin diet in two mouse chronic seizure models. Neurobiol Dis 40, 565-572.
Wlaź, P., Socała, K., Nieoczym, D., Łuszczki, J. J., Zarnowska, I., Zarnowski, T., Czuczwar,
S. J. and Gasior, M. (2012) Anticonvulsant profile of caprylic acid, a main constituent
of the medium-chain triglyceride (MCT) ketogenic diet, in mice. Neuropharmacology
62, 1882-1889.
Wong-Riley, M. T. (1989) Cytochrome oxidase: an endogenous metabolic marker for
neuronal activity. Trends Neurosci 12, 94-101.
Woods, J. W., Tanen, M., Figueroa, D. J., Biswas, C., Zycband, E., Moller, D. E., Austin, C.
P. and Berger, J. P. (2003) Localization of PPARδ in murine central nervous system:
expression in oligodendrocytes and neurons. Brain research 975, 10-21.
Wu, Q., Boyle, M. P. and Palmiter, R. D. (2009) Loss of GABAergic signaling by AgRP
neurons to the parabrachial nucleus leads to starvation. Cell 137, 1225-1234.
Xing, G., Ren, M., Watson, W. A., O’Neil, J. T. and Verma, A. (2009) Traumatic brain injury-
induced expression and phosphorylation of pyruvate dehydrogenase: a mechanism
of dysregulated glucose metabolism. Neuroscience letters 454, 38-42.
Yenari, M. A. and Han, H. S. (2012) Neuroprotective mechanisms of hypothermia in brain
ischaemia. Nature reviews. Neuroscience 13, 267-278.
Yin, F., Boveris, A. and Cadenas, E. (2014) Mitochondrial energy metabolism and redox
signaling in brain aging and neurodegeneration. Antioxidants & redox signaling 20,
353-371.
Yu, A. C., Drejer, J., Hertz, L. and Schousboe, A. (1983) Pyruvate carboxylase activity in
primary cultures of astrocytes and neurons. J Neurochem 41, 1484-1487.
Zaganas, I., Spanaki, C. and Plaitakis, A. (2012) Expression of human GLUD2 glutamate
dehydrogenase in human tissues: functional implications. Neurochem Int 61, 455-
462.
Accepted Article
This article is protected by copyright. All rights reserved.
Zakharov, P., Gudimchuk, N., Voevodin, V., Tikhonravov, A., Ataullakhanov, F. I. and
Grishchuk, E. L. (2015) Molecular and Mechanical Causes of Microtubule
Catastrophe and Aging. Biophys J 109, 2574-2591.
Zhang, Z., Chen, W., Zhao, Y. and Yang, Y. (2018) Spatiotemporal Imaging of Cellular
Energy Metabolism with Genetically-Encoded Fluorescent Sensors in Brain.
Neuroscience bulletin, 1-12.
Zhao, W., Varghese, M., Vempati, P. et al. (2012) Caprylic triglyceride as a novel therapeutic
approach to effectively improve the performance and attenuate the symptoms due to
the motor neuron loss in ALS disease. PloS one 7, e49191.
Zhao, Z., Lange, D. J., Voustianiouk, A. et al. (2006) A ketogenic diet as a potential novel
therapeutic intervention in amyotrophic lateral sclerosis. BMC neuroscience 7, 29.
Zhao, Z., Nelson, A. R., Betsholtz, C. and Zlokovic, B. V. (2015) Establishment and
Dysfunction of the Blood-Brain Barrier. Cell 163, 1064-1078.
Zhou, D., Lambert, S., Malen, P. L., Carpenter, S., Boland, L. M. and Bennett, V. (1998)
AnkyrinG is required for clustering of voltage-gated Na channels at axon initial
segments and for normal action potential firing. The Journal of cell biology 143, 1295-
1304.
Zonta, M., Angulo, M. C., Gobbo, S., Rosengarten, B., Hossmann, K. A., Pozzan, T. and
Carmignoto, G. (2003) Neuron-to-astrocyte signaling is central to the dynamic control
of brain microcirculation. Nat Neurosci 6, 43-50.
Figure legends
Figure 1: Schematic representation of the main ATP consuming processes in neurons. (A) At the
synapse, most of neuronal signaling-related ATP is utilized by the Na+/K+-ATPase and in the synaptic
vesicle cycle. The Na+/K+-ATPase restores membrane potential after depolarization and maintains the
electrochemical gradient of Na+ used in the removal of Ca2+ by the Na+/Ca2+ exchanger and in the
uptake of neurotransmitters (i.e monoamines, GABA, glycine and glutamate). The synaptic vesicle
cycle uses ATP for neurotransmitter loading by vacuolar H+-ATPase and for protein disassembly
machineries during docking and priming. Also, extrusion of elevated cytosolic Ca2+ after exocytosis by
the Ca2+-ATPase (not shown) needs ATP. (B) AP generation and propagation at the axon initial
segment and the nodes of Ranvier as well as reversing the ion fluxes after depolarization are also
linked to high signaling-related ATP expenditure by Na+/K+-ATPase. (C) Basic cellular processes,
such as axonal transport involved in the trafficking of vesicles, mitochondria, and other cargo within
the neurites require motor proteins that hydrolyze ATP.
Figure 2: Schematic representation of the neurometabolic coupling illustrating the energy expenditure
in a glutamatergic synapse and the ANLS hypothesis. Glutamatergic vesicular release activates
postsynaptic receptors (AMPA and NMDA, blue and red, respectively) depolarizes the membrane by
opening channels, in turn triggering the influx of Na+ ions that constrain the use of ATP when they are
pumped out. Glutamate in the synaptic cleft is taken up into astrocytes via EAAT2 and metabolized
via glutamine synthase (GS). Glial glutamine is released via the SN-1 transporter, taken up by
neurons via SAT to replenish glutamate pools via GLS1 and vesicular reuptake via vGLUT. Most of
the energy required by those presynaptic and postsynaptic processes comes from glucose
metabolism. However, neurons can use alternative sources of energy. According to the ANLS
hypothesis, astrocytes take up glucose from the blood via GLUT1 and convert it into lactate through
LDH5 that is released and taken up by neurons through MCT transporters. In neurons, lactate is
converted to pyruvate by LDH1, enters the TCA cycle and is metabolized via oxidative
phosphorylation in the mitochondria.
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Microglia are the resident immune cells of the central nervous system that mediate the life and death of nervous tissue. During normal function, they exhibit a surveying phenotype and maintain vital functions in nervous tissue. In the event of injury or disease, chronic inflammation can result, wherein microglia develop a hyper-activated phenotype, shed their regenerative function, actively kill contiguous cells, and can partition injured tissue by initiating scar formation. With recoverable injury, microglia can develop a primed phenotype, where they appear to recover from an inflammatory event, but are limited in their support functions and show inappropriate responses to future injury often associated with neurodegenerative disorders. These microglial phenotypes were acutely recreated in vitro with potent pro- and anti-inflammatory treatments. Primary cultured microglia or mixed glia (microglia, astrocytes, and oligodendrocytes) were treated for 6 h with lipopolysaccharide (LPS). Recovery from an inflammatory state was modeled with 18-h treatment of the anti-inflammatory steroid dexamethasone. The cells were then treated for 24 h with interferon gamma (IFNγ) to detect inflammatory memory after recovery. Surveying was best represented in the untreated vehicle (Veh) cases and was characterized by negligible secretion of pro-inflammatory factors, limited expression of immune proteins such as induced nitric oxide synthase (iNOS), major histocompatibility complex class II (MHCII), relatively high expression of brain-derived and glial-derived neurotrophic factors (BDNF and GDNF), and thinly branched smaller microglia. Activation was noted in the LPS- and IFNγ-treated microglia with increased cytokines, NO, NGF, iNOS, proliferation, phagocytosis, reduced BDNF, and flattened round amoeboid-shaped microglia. Priming was observed to be an incomplete surveying restoration using dexamethasone from an activation comparison of LPS, IFNγ, and LPS/IFNγ. Dexamethasone treatments resulted in the most profound dysregulation of expression of NO, TNF, IL-1β, NGF, CD68, and MHCII as well as ramified morphology and uptake of myelin. These findings suggest microglial priming and hyper-activation may be effectively modeled in vitro to allow mechanistic investigations into these key cellular phenotypes.
Article
Full-text available
Background: Uncontrolled microglial activation contributes to the pathogenesis of various neurodegenerative diseases. Previous studies have shown that proinflammatory microglia are powered by glycolysis, which relays on high levels of glucose uptake. This study aimed to understand how glucose uptake is facilitated in active microglia and whether microglial activation can be controlled by restricting glucose uptake. Methods: Primary murine brain microglia, BV2 cells and the newly established microglial cell line B6M7 were treated with LPS (100 ng/ml) + IFNγ (100 ng/ml) or IL-4 (20 ng/ml) for 24 h. The expression of glucose transporters (GLUTs) was examined by PCR and Western blot. Glucose uptake by microglia was inhibited using the GLUT1- specific inhibitor STF31. The metabolic profiles were tested using the Glycolysis Stress Test and Mito Stress Test Kits using the Seahorse XFe96 Analyser. Inflammatory gene expression was examined by real-time RT-PCR and protein secretion by cytokine beads array. The effect of STF31 on microglial activation and neurodegeneraion was further tested in a mouse model of light-induced retinal degeneration. Results: The mRNA and protein of GLUT1, 3, 4, 5, 6, 8, 9, 10, 12, and 13 were detected in microglia. The expression level of GLUT1 was the highest among all GLUTs detected. LPS + IFNγ treatment further increased GLUT1 expression. STF31 dose-dependently reduced glucose uptake and suppressed Extracellular Acidification Rate (ECAR) in naïve, M(LPS + IFNγ) and M(IL-4) microglia. The treatment also prevented the upregulation of inflammatory cytokines including TNFα, IL-1β, IL-6, and CCL2 in M(LPS + IFNγ) microglia. Interestingly, the Oxygen Consumption Rates (OCR) was increased in M(LPS + IFNγ) microglia but reduced in M(IL-4) microglia by STF31 treatment. Intraperitoneal injection of STF31 reduced light-induced microglial activation and retinal degeneration. Conclusion: Glucose uptake in microglia is facilitated predominately by GLUT1, particularly under inflammatory conditions. Targeting GLUT1 could be an effective approach to control neuroinflammation.
Article
Full-text available
Objective Insulin action in the brain controls metabolism and brain function, which is linked to proper mitochondrial function. Conversely, brain insulin resistance associates with mitochondrial stress and metabolic and neurodegenerative diseases. In the present study, we aimed to decipher the impact of hypothalamic insulin action on mitochondrial stress responses, function and metabolism. Methods To investigate the crosstalk of insulin action and mitochondrial stress responses (MSR), namely the mitochondrial unfolded protein response (UPRmt) and integrated stress response (ISR), qPCR, western blotting, and mitochondrial activity assays were performed. These methods were used to analyze the hypothalamic cell line CLU183 treated with insulin in the presence or absence of the insulin receptor as well as in mice fed a high fat diet (HFD) for three days and STZ-treated mice without or with insulin therapy. Intranasal insulin treatment was used to investigate the effect of acute brain insulin action on metabolism and mitochondrial stress responses. Results Acute HFD feeding reduces hypothalamic mitochondrial stress responsive gene expression of Atf4, Chop, Hsp60, Hsp10, ClpP, and Lonp1 in C57BL/6N mice. We show that insulin via ERK activation increases the expression of MSR genes in vitro as well as in the hypothalamus of streptozotocin-treated mice. This regulation propagates mitochondrial function by controlling mitochondrial proteostasis and prevents excessive autophagy under serum deprivation. Finally, short-term intranasal insulin treatment activates MSR gene expression in the hypothalamus of HFD-fed C57BL/6N mice and reduces food intake and body weight development. Conclusions We define hypothalamic insulin action as a novel master regulator of MSR, ensuring proper mitochondrial function by controlling mitochondrial proteostasis and regulating metabolism.
Article
Full-text available
Early views of glia as relatively inert, housekeeping cells have evolved, and glia are now recognized as dynamic cells that not only respond to neuronal activity but also sense metabolic changes and regulate neuronal metabolism. This evolution has been aided in part by technical advances permitting progressively better spatial and temporal resolution. Recent advances in cell-type specific genetic manipulation and sub-cellular metabolic probes promise to further this evolution by enabling study of metabolic interactions between intertwined fine neuronal and glial processes in vivo. Views of glia in disease processes have also evolved. Long considered purely reactive, glia and particularly microglia are now seen to play active roles in both promoting and limiting brain injury. At the same time, established concepts of glial energetics are now being linked to areas such as learning and neural network function, topics previously considered far removed from glial biology.
Article
Full-text available
Hypothermia is increasingly used as a therapeutic measure to treat brain injury. However, the cellular mechanisms underpinning its actions are complex and are not yet fully elucidated. Astrocytes are the most abundant cell type in the brain and are likely to play a critical role. In this study, transcriptional changes and the protein expression profile of human primary cortical astrocytes cultured under hypoxic conditions for 6 h were investigated. Cells were treated either with or without a mild hypothermic intervention 2 h post-insult to mimic the treatment of patients following traumatic brain injury (TBI) and/or stroke. Using human gene expression microarrays, 411 differentially expressed genes were identified following hypothermic treatment of astrocytes following a 2 h hypoxic insult. KEGG pathway analysis indicated that these genes were mainly enriched in the Wnt and p53 signaling pathways, which were inhibited following hypothermic intervention. The expression levels of 168 genes involved in Wnt signaling were validated by quantitative real-time-PCR (qPCR). Among these genes, 10 were up-regulated and 32 were down-regulated with the remainder unchanged. Two of the differentially expressed genes (DEGs), p38 and JNK, were selected for validation at the protein level using cell based ELISA. Hypothermic intervention significantly down-regulated total protein levels for the gene products of p38 and JNK. Moreover, hypothermia significantly up-regulated the phosphorylated (activated) forms of JNK protein, while downregulating phosphorylation of p38 protein. Within the p53 signaling pathway, 35 human apoptosis-related proteins closely associated with Wnt signaling were investigated using a Proteome Profiling Array. Hypothermic intervention significantly down-regulated 18 proteins, while upregulating one protein, survivin. Hypothermia is a complex intervention; this study provides the first detailed longitudinal investigation at the transcript and protein expression levels of the molecular effects of therapeutic hypothermic intervention on hypoxic human primary cortical astrocytes. The identified genes and proteins are targets for detailed functional studies, which may help to develop new treatments for brain injury based on an in-depth mechanistic understanding of the astrocytic response to hypoxia and/or hypothermia.
Article
Objective: To investigate the association between type 2 diabetes mellitus (T2DM) and subsequent Parkinson disease (PD). Methods: Linked English national Hospital Episode Statistics and mortality data (1999-2011) were used to conduct a retrospective cohort study. A cohort of individuals admitted for hospital care with a coded diagnosis of T2DM was constructed, and compared to a reference cohort. Subsequent PD risk was estimated using Cox regression models. Individuals with a coded diagnosis of cerebrovascular disease, vascular parkinsonism, drug-induced parkinsonism, and normal pressure hydrocephalus were excluded from the analysis. Results: A total of 2,017,115 individuals entered the T2DM cohort and 6,173,208 entered the reference cohort. There were significantly elevated rates of PD following T2DM (hazard ratio [HR] 1.32, 95% confidence interval [CI] 1.29-1.35; p < 0.001). The relative increase was greater in those with complicated T2DM (HR 1.49, 95% CI 1.42-1.56) and when comparing younger individuals (HR 3.81, 95% CI 2.84-5.11 in age group 25-44 years). Conclusions: We report an increased rate of subsequent PD following T2DM in this large cohort study. These findings may reflect shared genetic predisposition and/or disrupted shared pathogenic pathways with potential clinical and therapeutic implications.
Article
Stroke, ischemic stroke and subtypes of ischemic stroke display substantial heritability. Compared to related vascular conditions, the number of established risk loci reaching genome-wide significance for association with stroke is still in the lower range, particularly for etiological stroke subtypes such as large artery atherosclerotic stroke or small vessel stroke. Nevertheless, for individual loci substantial progress has been made in determining the specific mechanisms mediating stroke risk. In this review, we present a roadmap for functional follow-up of common risk variants associated with stroke. First, we discuss in-silico strategies for characterizing signals in non-coding regions and highlight databases providing information on quantitative trait loci (QTL) for mRNA and protein expression, as well as methylation, focusing on those with presumed relevance for stroke. Next, we discuss experimental strategies for following up on non-coding risk variants and regions such as massively parallel reporter assays (MPRA), proteome-wide association studies (PWAS) and chromatin conformation capture (3C) assays. These and other approaches are relevant for gaining insight into the specific variants and mechanisms mediating genetic stroke risk. Finally, we discuss how genetic findings could influence clinical practice by adding to diagnostic algorithms and eventually improve treatment options for stroke.
Article
We will argue here that there is net lactate transfer from astrocytes to neurons and that this transfer is important for brain function. Following CrossTalk guidelines, we will focus on data published over the last decade.
Article
Alzheimer's disease (AD) is an age-related neurodegenerative disorder of the brain. The presence of amyloid-beta (Aβ) plaques, neurofibrillary tangles (NFTs), loss of neurons, synapses, and altered sensory perceptions, including memory loss, delineate AD. However, the cause of AD is not clearly known. Several genetic and nongenetic factors have been implicated in the disease. Of the genes, the ɛ4 allele of apolipoprotein E is the largest known genetic risk factor of AD. This review article focuses on the various genetic and other predisposing factors that account for AD, pathophysiology of the disease, and the mechanisms by which Aβ plaques and NFTs are formed and could affect AD brain. In addition, recent advances and current diagnostics available for AD patients are detailed. As oxidative stress has been implicated in the etiology of the disease, special emphasis is given for nutrition based antioxidant therapies and interventional strategies for reducing/treating AD.