13
ORIGINAL PAPER Metabolism, Compartmentation, Transport and Production of Acetate in the Cortical Brain Tissue Slice Caroline Rae Aure ´lie D. Fekete Mohammed A. Kashem Fatima A. Nasrallah Stefan Bro ¨er Received: 30 January 2012 / Revised: 12 July 2012 / Accepted: 13 July 2012 / Published online: 1 August 2012 Ó Springer Science+Business Media, LLC 2012 Abstract Acetate is a two carbon intermediate in metab- olism. It is an accepted marker of astrocytic metabolism, and a substrate for production of metabolites such as glutamine, glutamate and GABA. However, anomalies exist in the current explanations of compartmentation and metabolism of acetate. Here, we investigated these anomalies by exam- ining transport, production and metabolism of acetate. Acetate is a good substrate for the neuronal monocarboxylate transporter MCT2 (K M = 2.58 ± 0.8) and the glial MCT1 but a poor substrate for the glial MCT4. Acetate is accu- mulated by brain cortical tissue slices to concentrations in excess of those in the media, suggesting active transport, possibly via the sodium dependent SMCT. [2- 13 C]Acetate is produced from [3- 13 C]pyruvate, [3- 13 C]lactate and [1- 13 C]glucose with the rate of production related to acetyl- CoA levels, which is likely generated in a ubiquitous cyto- solic compartment via acetyl-CoA hydrolase. Citrate breakdown occurs in response to demand for acetyl-CoA units; this citrate is not derived from acetate carbon but its fate is influenced by acetate levels. Finally, use of acetate is altered by levels of nicotinamide or NAD ? . This suggests that metabolism of acetate is controlled rigorously at the enzyme level, via changes in the acetylation status of acetyl- CoA synthetase and is not regulated by restriction of uptake. Keywords Acetate Á 13 C NMR spectroscopy Á Neuron-glia interactions Introduction Glucose is the mandatory substrate for energy supply in the brain. Brain can also use the glucose catabolites lactate, pyruvate and acetate. Both pyruvate [1] and lactate have been shown to be adequate substrates for brain metabolism (Fig. 1) although there is some debate about the full extent to which they can subserve brain activity [2, 3]. Acetate has long been considered a marker for glial metabolism where it is incorporated mostly into glutamine, with label also reported from a small glutamate pool [4]. Acetate must be metabolized by conversion to acetyl- CoA in a reaction catalysed by acetyl-CoA synthetase (E.C. 6.2.1.1); this reaction requires ATP and releases pyrophos- phate. Two isoforms are known in mammalian cells [5]. AceCS2 is a mitochondrial isoform channeling acetate into the TCA cycle, while AceCS1 is a cytosolic isoform, which provides acetyl-CoA for fatty acid biosynthesis. Acetate units (Fig. 1) are also used in the brain for lipid synthesis [6], for acetylation of histone proteins in the nucleus [7], for synthesis of the neurotransmitter acetylcholine [8] and for many other acetylated compounds such as N-acetylaspartate, N-acetylaspartylglutamate and acetylcarnitine. Both AceCS isoforms are subject to complex layers of regulation; the cytosolic form is increased by the presence of acetate which may come from a variety of sources, such as via blood from oxidation of ingested ethanol, from histone deacetylases, or by the activity of the ubiquitous acetyl-CoA hydrolase. AceCS2 levels are induced by ketogenic conditions but the mechanisms of this induction Special Issue: In Honor of Leif Hertz. C. Rae (&) Á A. D. Fekete Á M. A. Kashem Á F. A. Nasrallah Neuroscience Research Australia, and Brain Sciences, University of New South Wales, Barker St, Randwick, NSW 2031, Australia e-mail: [email protected] S. Bro ¨er Research School of Biology, The Australian National University, Canberra, ACT 0200, Australia 123 Neurochem Res (2012) 37:2541–2553 DOI 10.1007/s11064-012-0847-5

Metabolism, Compartmentation, Transport and Production of Acetate in the Cortical Brain Tissue Slice

  • Upload
    unsw

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

ORIGINAL PAPER

Metabolism, Compartmentation, Transport and Productionof Acetate in the Cortical Brain Tissue Slice

Caroline Rae • Aurelie D. Fekete •

Mohammed A. Kashem • Fatima A. Nasrallah •

Stefan Broer

Received: 30 January 2012 / Revised: 12 July 2012 / Accepted: 13 July 2012 / Published online: 1 August 2012

� Springer Science+Business Media, LLC 2012

Abstract Acetate is a two carbon intermediate in metab-

olism. It is an accepted marker of astrocytic metabolism, and

a substrate for production of metabolites such as glutamine,

glutamate and GABA. However, anomalies exist in the

current explanations of compartmentation and metabolism

of acetate. Here, we investigated these anomalies by exam-

ining transport, production and metabolism of acetate.

Acetate is a good substrate for the neuronal monocarboxylate

transporter MCT2 (KM = 2.58 ± 0.8) and the glial MCT1

but a poor substrate for the glial MCT4. Acetate is accu-

mulated by brain cortical tissue slices to concentrations in

excess of those in the media, suggesting active transport,

possibly via the sodium dependent SMCT. [2-13C]Acetate is

produced from [3-13C]pyruvate, [3-13C]lactate and

[1-13C]glucose with the rate of production related to acetyl-

CoA levels, which is likely generated in a ubiquitous cyto-

solic compartment via acetyl-CoA hydrolase. Citrate

breakdown occurs in response to demand for acetyl-CoA

units; this citrate is not derived from acetate carbon but its

fate is influenced by acetate levels. Finally, use of acetate is

altered by levels of nicotinamide or NAD?. This suggests

that metabolism of acetate is controlled rigorously at the

enzyme level, via changes in the acetylation status of acetyl-

CoA synthetase and is not regulated by restriction of uptake.

Keywords Acetate � 13C NMR spectroscopy �Neuron-glia interactions

Introduction

Glucose is the mandatory substrate for energy supply in the

brain. Brain can also use the glucose catabolites lactate,

pyruvate and acetate. Both pyruvate [1] and lactate have

been shown to be adequate substrates for brain metabolism

(Fig. 1) although there is some debate about the full extent

to which they can subserve brain activity [2, 3]. Acetate has

long been considered a marker for glial metabolism where

it is incorporated mostly into glutamine, with label also

reported from a small glutamate pool [4].

Acetate must be metabolized by conversion to acetyl-

CoA in a reaction catalysed by acetyl-CoA synthetase (E.C.

6.2.1.1); this reaction requires ATP and releases pyrophos-

phate. Two isoforms are known in mammalian cells [5].

AceCS2 is a mitochondrial isoform channeling acetate into

the TCA cycle, while AceCS1 is a cytosolic isoform, which

provides acetyl-CoA for fatty acid biosynthesis. Acetate

units (Fig. 1) are also used in the brain for lipid synthesis [6],

for acetylation of histone proteins in the nucleus [7], for

synthesis of the neurotransmitter acetylcholine [8] and for

many other acetylated compounds such as N-acetylaspartate,

N-acetylaspartylglutamate and acetylcarnitine.

Both AceCS isoforms are subject to complex layers of

regulation; the cytosolic form is increased by the presence

of acetate which may come from a variety of sources, such

as via blood from oxidation of ingested ethanol, from

histone deacetylases, or by the activity of the ubiquitous

acetyl-CoA hydrolase. AceCS2 levels are induced by

ketogenic conditions but the mechanisms of this induction

Special Issue: In Honor of Leif Hertz.

C. Rae (&) � A. D. Fekete � M. A. Kashem � F. A. Nasrallah

Neuroscience Research Australia, and Brain Sciences,

University of New South Wales, Barker St, Randwick,

NSW 2031, Australia

e-mail: [email protected]

S. Broer

Research School of Biology, The Australian National University,

Canberra, ACT 0200, Australia

123

Neurochem Res (2012) 37:2541–2553

DOI 10.1007/s11064-012-0847-5

remain unclear. In the brain AceCS activity is found in

both neurons and glia with a mostly mitochondrial distri-

bution, although a small but significant fraction of it is also

likely to be found in the cytosol [9, 10].

AceCS isoforms are also subjected to post-translational

regulation via acetylation under control of the deacetylat-

ing silent information regulator protein family, the Sirtuins.

Specifically, AceCS1 is regulated by SIRT1 and AceCS2

by SIRT3 [11]. Sirtuins are under cellular control via

NAD?; nicotinamide is an inhibitor of the deacetylase

reaction such that low-energy status may turn off energy

consuming reactions such as AceCS. In the brain, SIRT1 is

located mostly, if not exclusively, in neurons [12].

Acetate is taken up and metabolized by neurons in

culture [13] as well as by hippocampal nerve terminals [8]

and in whole rat brain [14]. It is a poor substrate for cortical

tissue slices and needs to be used in conjunction with

substrates such as glucose for adequate slice respiration

[15, 16]. In vivo, acetate is rapidly incorporated into glu-

tamine, GABA, glutamate and aspartate, with the highest

specific activity in glutamine.

Acetate is a substrate for the monocarboxylate trans-

porters (MCTs). These are members of the SLC16 family

and are proton symporters with a wide range of monocarb-

oxylate substrates. MCT1 is a low affinity transporter found

mainly in the blood–brain barrier. Lower levels are found in

astrocytes throughout the brain. MCT2 on the other hand is

generally a higher affinity transporter found in neurons in the

post-synaptic density [17, 18]. MCT4 is a low affinity

transporter and is selectively expressed in astrocytes [19].

Recently another lactate transporter family has been

described which is sodium dependent. SMCT1 (sodium-

dependent monocarboxylate transporter 1; SLC5A8) is a

high-affinity transporter found in neurons, while SMCT2

(SLC5A12) has been reported in glial cells [20].

Waniewski and Martin [21] suggested that the reason

acetate is a relatively poor substrate for neurons was due to

transporter affinity, based on poor uptake of acetate by syn-

aptosomal fractions compared to astrocytes. The uptake of

acetate by neurons was not measured in this work, and their

findings disagreed with the earlier results of Clarke et al. who

showed that acetate is taken up avidly by synaptosomes,

where, rather than being metabolized, it is used to acetylate

synaptosomal proteins. This acetate is released upon activa-

tion of sodium channels, suggesting that acetylation reactions

are involved in dynamic regulation of synaptic activity [22].

Here, we investigated the basis of the compartmentation of

acetate metabolism by studying its metabolic fates in Guinea

pig brain cortical tissue slices along with a range of competing

substrates, and its transport in Xenopus oocytes to clarify

which MCT isoforms are capable of transporting acetate.

Methods

Materials

Guinea pigs (Dunkin-Hartley), weighing 400–800 g, were fed

ad libitum on standard Guinea pig/rabbit pellets, with fresh

cabbage leaves and lucerne hay roughage. Animals were main-

tained on a 12 h light/dark cycle. All experiments were con-

ducted in accordance with the guidelines of the National Health

and Medical Research Council of Australia and were approved

by the institutional (UNSW) Animal Care Ethics Committee.

Sodium [3-13C]pyruvate, sodium [3-13C]L-lactate, sodium

[2-13C]acetate and [1-13C]D-glucose and sodium [13C]formate

Fig. 1 Cartoon showing metabolism of acetate in the brain

2542 Neurochem Res (2012) 37:2541–2553

123

were purchased from Cambridge Isotope Laboratories Inc

(Andover, MA, USA). Sodium pyruvate, sodium lactate,

sodium acetate, nicotinamide, NAD? and ethylenediamine-

tetraacetic acid (EDTA) were purchased from Sigma-Aldrich

(St Louis, MO). All other reagents were of Analytical Reagent

grade.

Other data were taken from experiments that have

already been published, including those with AR-C122982,

a potent, specific MCT inhibitor [23], those with ligands

active at GABA(B) receptors [24] and those examining the

effect of blocking transport of alanine [25]. 1H MRS

spectra obtained from these experiments were used to

measure the amount of 13C-labelled acetate present using

the resolved 13C-satellite resonance at d = 1.78 ppm. The

data from a previously published experiment examining

metabolism of 5 mM [1-13C]D-glucose in the presence of

the inhibitor 4-hydroxycyanocinnamate (7 mM) [23] was

also examined for acetate production.

Acetate Uptake in Xenopus Oocytes

Uptake of [14C]acetate in Xenopus laevis oocytes was

carried out as described in detail previously [26–28]. A

series of concentrations were used and the uptake rate

calculated using [14C]acetate standards to obtain KM. In

flux experiments the activity of non-injected oocytes was

subtracted at all substrate concentrations to correct for

endogenous uptake of monocarboxylates. Endogenous

uptake of monocarboxylates in oocytes is Na?-dependent

and therefore is not observed when monocarboxylate

uptake is monitored by pH-sensitive microelectrodes [29].

Preparation of Brain Cortical Tissue Slices

Following cervical dislocation, guinea pig brains were

removed from the cranial vault and 350 lm paraxial cortical

slices were obtained using a McIlwain tissue chopper. The

slices were then washed three times in a modified Krebs-

Henseleit buffer [15], resuspended for 1 h in fresh buffer

containing 10 mmol/L unlabelled glucose and gassed with

95 % O2/5 % CO2 in a shaking water bath, maintained at

37 �C, to allow metabolic recovery [30]. Slices were then

washed three times in glucose-free buffer and resuspended in

fresh buffer with the substrate of choice.

Synthesis of Acetate from Various Substrates in Brain

Cortical Tissue Slices

To determine labeling patterns from each substrate a

number of experiments were made where label fate and

pool sizes were monitored in the presence and absence of

competing monocarboxylates. The scope of these experi-

ments was limited by the requirement to maintain brain

slice respiration at maximal rates so that changes in

metabolism did not reflect simple limitations on mito-

chondrial respiration rates due to substrate unavailability.

The following experiments were performed:

To determine the effect of competing acetate with

pyruvate:

1. 2 mM sodium [3-13C]pyruvate

2. 2 mM sodium [3-13C]pyruvate plus 5 mM sodium

acetate

3. 2 mM sodium pyruvate plus 5 mM sodium

[2-13C]acetate

Five mM acetate was chosen to fully saturate both MCT1

and MCT2. Historically, this concentration of acetate has

been shown to be metabolized mostly in an astrocyte pool if

metabolism is limited to a short time frame [15].

To determine the effect of competing acetate with

lactate:

1. 0.74 mM sodium [3-13C]lactate and 2 mM D-glucose

2. 0.74 mM sodium [3-13C]lactate and 2 mM D-glucose

plus 5 mM sodium acetate

3. 0.74 mM sodium lactate and 2 mM D-glucose plus

5 mM sodium [2-13C]acetate.

This concentration of lactate was chosen to compete

mainly at MCT2 (KM = 0.74 mM, Table 1) and less so at

MCT1 (KM = 5.6 mM, Table 1). We added 2 mM glucose

to allow the slices to maintain their respiratory rate at

adequate levels as 0.74 mM lactate is an insufficient

amount of substrate to maintain respiration [30].

To determine the effect of competing lactate with

pyruvate:

1. 2 mM sodium [3-13C]pyruvate

2. 2 mM sodium [3-13C]pyruvate plus 0.74 mM sodium

lactate

3. 2 mM sodium pyruvate plus 0.74 mM sodium

[3-13C]lactate

To determine the effect of modulating levels of NAD?

or nicotinamide on acetate metabolism:

1. 5 mM D-glucose and 5 mM sodium [2-13C]acetate

2. 5 mM D-glucose and 5 mM sodium [2-13C]acetate plus

250 lM NAD?

3. 5 mM D-glucose and 5 mM sodium [2-13C]acetate plus

250 lM nicotinamide.

All experiments with lactate and pyruvate as substrates

were incubated for 60 min, while the experiment with

glucose was incubated for 90 min.

To measure citrate in the 13C spectrum, 4 mM EDTA was

added to the samples to chelate paramagnetic species and

sharpen the resonance [15] and 13C NMR spectra of these

samples were reacquired. This was not successful with the

Neurochem Res (2012) 37:2541–2553 2543

123

experiment with lactate and pyruvate as the samples had

deteriorated due to repeated removal from the NMR tubes

and signal to noise was poor. Instead the citrate concentration

was estimated from the broadened metal-bound resonance.

The number of samples was N = 4 in all cases.

Preparation of Samples and NMR Analysis

On completion of the incubation period, slices were removed

from the incubation buffer by rapid filtration and extracted in

methanol/chloroform according to the method of Le Belle

[31]. Extracts were lyophilized, and the pellet retained for

protein estimation by the Lowry technique. Lyophilized

supernatants were stored at -20 �C until required for NMR

analysis. Samples were resuspended in 0.60 mL D2O con-

taining 2 mM sodium [13C]formate as an internal intensity

and chemical shift reference (13C d 171.8). Fully relaxed 1H

and 1H[13C-decoupled] spectra (total cycle of 30 s, compris-

ing 24 s relaxation delay, 4 s water suppression and *1 s

acquisition time), WURST-40 [32] with a 112-step phase

cycle [33], decoupling during acquisition) were obtained at

800.15 MHz on a Bruker Avance-III 800 MHz spectrometer

equipped with a TCI cryoprobe probe, followed by 13C

[1H-decoupled] spectra (800 transients, cycle of 4 s com-

prising 2 s relaxation delay and *2 s acquisition time, con-

tinuous WALTZ-65 decoupling, 185180 data points).

Assignments were made as described previously [34].13C [1H-Decoupled] spectra were Fourier transformed

using 1 Hz exponential line-broadening and peak areas

were determined by integration using standard Bruker

software (TOPSPIN, Version 1.3, or 2.5) following base-

line correction. Peak areas were adjusted for nuclear

Overhauser effect, saturation and natural abundance effects

and quantified by reference to the area of the internal

standard resonance of [13C]formate. Glu C3 was not

quantified due to resonance overlap with pyruvate signal.

Metabolite pool sizes (lactate, alanine, GABA, glutamate,

glutamine and aspartate) were determined by integration of

resonances in fully relaxed 800 MHz 1H[13C-decoupled]

spectra using [13C]formate as the internal intensity refer-

ence. The amount of acetate labeled with 13C at C2 was

also determined from the 13C satellite at d = 1.776 ppm in

the fully-relaxed 1H NMR spectrum.

Experimental data (N = 4) are given as means (standard

deviation). Statistical analysis was done using ANOVA for

comparing experimental interventions with control (N = 4,

followed, only where statistical significance was indicated by

Scheffe F test, by a nonparametric (Mann–Whitney U) test

(Statview Student)). Significance was assumed at a = 0.05.

Principal Components Analysis

In order to discover what metabolic factors influence the

presence of labeled acetate in the cortical tissue slice we

constructed a statistical model using data from previously

published experiments [35, 36], but also including integrals

from the CH313C satellites of acetate C2. Fifteen variables

from 163 experiments were used as input for a principal

components model using Simca P? (v11.5, Umetrics,

Umea, Sweden). Each dataset for a particular manipulation

was imported as the relative change from the average value

obtained from the control group for that particular experi-

ment. Data were univariance scaled to standardize variance

between the high and low concentration metabolites [37],

to ensure that the 13C labelling and steady state pool size

concentrations equally contributed to the model.

Table 1 Substrate affinities for the various monocarboxylate

transporters

Transporter Substrates KM (mM)

MCT1 Acetate 1.6 ± 0.4 [68]

D, L-b-hydroxybutyrate 12.5 ± 0.26 [69]

Lactate 5.6 [70]

D-lactate [60 [71]

16 ± 6 [68]

L-lactate 4.38 ± 0.74 [71]

3.5 ± 0.4 [28]

5 ± 1 [68]

Pyruvate 1.01 ± 0.06 [28]

2.09 ± 0.37 [71]

MCT2 Acetate 2.58 ± 0.8 Fig. 2

D,L-b-Hydroxybutyrate 1.2 ± 0.2 [27]

L-lactate 0.74 ± 0.07 [27]

1.0 ± 0.2 [72]

Pyruvate 0.08 ± 0.01 [27, 71]

0.025 [73]

MCT4 Acetate Non-saturable Fig. 2

L-b-Hydroxybutyrate 824 ± 64 [71]

D-b-Hydroxybutyrate 130 ± 9.6 [71]

Lactate 17–34 [72]

D-lactate 519 ± 35 [71]

L-lactate 28 ± 4 [71]

33.7 ± 4.9 [74]

Pyruvate 153 ± 6 [71]

SMCT1 Acetate 1.6 ± 0.1 [75]

2.46 ± 0.89 [76]

L-b-Hydroxybutyrate 2.33 ± 0.17 [40]

D-b-Hydroxybutyrate 1.44 ± 0.12 [40]

L-lactate 0.184 ± 0.008 [40]

0.235 ± 0.24 [76]

D-lactate 1.088 ± 0.068 [40]

0.742 ± 0.33 [76]

Pyruvate 0.387 ± 0.043 [40]

SMCT2 L-lactate 16.9 ± 3.7 [77]

2544 Neurochem Res (2012) 37:2541–2553

123

Results

Acetate Uptake by MCT2 and MCT4

Acetate was found to be a good substrate for MCT2 with a

Michaelis–Menten constant comparable to that of MCT1

(Fig. 2a). Acetate is a very poor substrate for MCT4 (Fig. 2b).

The potential use of acetate by any cell type in the brain

will depend on the availability of transport proteins and

metabolic enzymes that can convert acetate to acetyl-CoA.

In situ hybridization data from the Allen Brain Atlas

(http://mouse.brain-map.org) suggest that expression of the

mitochondrial acetyl-CoA synthetase 2 (AceCS2) in the

brain is extremely low. The cytosolic AceCS1, by contrast,

is found in significant amounts in all brain areas. In terms

of transport, the brain expresses MCT1, MCT2, MCT4 and

SMCT1. It appears that MCT2 and SMCT1 are largely

neuronal transporters, while MCT1 and MCT4 are

expressed in astrocytes. Apart from MCT4, all of these are

capable of transporting acetate with similar KM-values

(Fig. 2, Table 1). There is an argument that although the

KM for acetate is low enough, capacity is limited by low

levels of MCT2 expression in neurons. However, it is

evident from the residual 13C acetate signal that acetate

was accumulated by at least one brain compartment to

levels higher than in the medium. If this accumulation were

limited to just the astrocytic pool which is thought to me-

tabolise acetate (estimated at around 30 % of brain vol-

ume) this would mean that the concentration of acetate in

this compartment would be in the order of [30 mM.

Thus, it can be reasonably expected that acetate could

penetrate both astrocytes and neurons, but that this may

involve transporters other than MCTs.

Acetate Synthesis and Presence in NMR Spectra

Acetate is reliably present in 1H NMR spectra of cortical

tissue slices, appearing as a singlet resonance which is co-

resonant with the bCH2 of GABA at 1.8 ppm (Fig. 3). It is

difficult to know whether this acetate is endogenous or

whether it might arise due to a putative breakdown of N-

acetylaspartate in the extraction process. There is a doublet

(1JCH = 127.1 ± 0.03 Hz) which is collapsed by broad-

band 13C decoupling of the 1H spectrum, indicating that a

portion of the acetate present is also labeled with 13C at C2

(i.e. in excess of natural abundance at 1.1 %). This doublet

is present in 1H NMR spectra of brain slice extracts fol-

lowing incubation with [3-13C]pyruvate, [3-13C]lactate and

[1-13C]glucose and indicates that acetate labeled at C2 is

synthesized from these substrates (Fig. 3).

Acetate Compartmentation—Pyruvate and Lactate

The effects of adding 0.74 mM L-lactate to 2 mM pyruvate

are shown in Fig. 4. The amount of label incorporated into

acetate C2 was reduced by addition of unlabelled lactate

(control acetate C2 = 1.04 (0.21) cf. 0.76 (0.12) lmol/h/

100 mg protein.). When the label is transferred to lactate

by use of 0.74 mM [3-13C]lactate and 2 mM pyruvate label

was also incorporated into acetate C2. Acetate is therefore

also made in a compartment accessible to 0.74 mM lactate.

Scrutiny of the predicted vs actual labeling of the other

isotopomers revealed that the labeling of Gln C4 from

[3-13C]pyruvate was not significantly different in the

A

B

Fig. 2 Transport of acetate by monocarboxylate transporters.

Oocytes were injected with cRNA encoding MCT2 (20 ng) or

MCT4 (10 ng). After 3–5 days, uptake of [14C]acetate was measured

over a period of 20 min at the indicated concentrations. Each datapoint represents the mean ± SD uptake activity of 8–10 oocytes

Neurochem Res (2012) 37:2541–2553 2545

123

presence (0.11 (0.03) lmol/h/100 mg protein) or absence

of lactate (0.11 (0.02) lmol/h/100 mg protein). It was not

possible to detect labeling of Gln C4 from 0.74 mM

[3-13C]lactate. This suggests that the astrocyte compart-

ment where glutamine is labeled is not accessible to lactate

under these conditions. Labelling of citrate was not sig-

nificantly affected by addition of lactate, and only a small

amount of citrate was labeled from [3-13C]lactate.

There was also little significant change in total metab-

olite pool sizes due to addition of lactate (Fig. 4b).

Acetate Compartmentation—Pyruvate and Acetate

Incubation of 2 mM [3-13C]pyruvate with 5 mM unlabelled

acetate resulted generally in increased labelling from

[3-13C]pyruvate into all isotopomers measured apart from Gln

C4 and Lac C3, which were unchanged (Fig. 5). Labelling of

acetate C2 increased significantly (0.72 (0.1) cf (1.16 (0.14)

lmol/h/100 mg protein as did labeling of citrate (1.47 (0.38)

cf 2.77 (0.5) lmol/h/100 mg protein) as well as Glu C2 and

GABA C2. When the label source was exchanged, incorpo-

ration into measured isotopomers dropped significantly. In

comparison with the control (2 mM [3-13C]pyruvate and no

acetate) experiment, little label was incorporated. Gln C4 was

the most highly labeled isotopomer and was labeled to

*30 % of the levels of the control experiment. Interestingly,

citrate was not labeled from acetate at all; it was not detected in

any spectra from the experiment with 2 mM pyruvate and

5 mM [2-13C]acetate, showing that citrate was not made from

acetyl-CoA which had been synthesized from acetate under

these circumstances.

The total pool size of glutamate and aspartate were sig-

nificantly increased by addition of [2-13C]acetate (Fig. 5b).

Acetate Compartmentation—Lactate and Acetate

Incubation of 0.74 mM [3-13C]L-lactate with 2 mM glu-

cose and unlabelled acetate (5 mM) had no significant

effect on the incorporation of label from [3-13C]L-lactate

into isotopomers of Glu, GABA or Asp (Fig. 6a). Incor-

poration of label into Gln C4 and Ala C3 was significantly

lower and there was no significant effect of acetate on label

incorporation into citrate. When the label source was

exchanged to [2-13C]acetate, label incorporation dropped

significantly. Label was incorporated above natural abun-

dance in Gln C4, Glu C4 and Ala C3.

Addition of acetate, either labeled or unlabelled, resul-

ted in a significant decrease in the total pool sizes of glu-

tamate, lactate, GABA, aspartate and alanine but had no

significant effect on the total pool size of glutamine.

Acetate Metabolism—Effect of NAD?

and Nicotinamide

When slices were incubated with 5 mM D-glucose and 5 mM

[2-13C]acetate, label was incorporated into Glu C4, GABA

C2 and Gln C4 to levels above natural abundance (Fig. 7).

Gln C4 was labeled more from 5 mM [2-13C]acetate in the

Fig. 3 Sections of fully-relaxed 600 MHz 1H NMR spectra of brain

cortical tissue slice extracts following metabolism of added 13C-

labelled substrates showing 13C-labelled acetate. Spectra in the toprow are [13C]-decoupled, while those in the bottom row are not. Left-most spectra are extracts following one hour of incubation with 2 mM

[3-13C]pyruvate [24], the middle spectra are extracts following

30 min of incubation with 10 mM [1-13C]glucose under depolarizing

(40 mM K?) conditions [23] and the right hand spectra are extracts

following 1 h of incubation with 8 mM [3-13C]lactate [16]. Fully

relaxed 1H and 1H[13C-decoupled] spectra (total cycle of 30 s,

comprising 24 s relaxation delay, 4 s water suppression and *2 s

acquisition time), WURST-40 [32] with a 112-step phase cycle [33],

decoupling during acquisition) were obtained at 600.13 MHz on a

Bruker DRX-600 spectrometer with a 5 mm dual 1H/13C probe

2546 Neurochem Res (2012) 37:2541–2553

123

presence of 5 mM glucose (11.7 %) than in the presence of

2 mM pyruvate (8.8 %) or 0.74 mM lactate and 2 mM

glucose (3.9 %).

Addition of 250 lM NAD? resulted in a significant

decrease in labelling from [2-13C]acetate into Gln C4 and

citrate. There was also a significant decrease in the total

pool size of aspartate and glutamine (Fig. 7b).

Addition of 250 lM nicotinamide resulted in significant

decreases in label incorporation into Glu C2 and C4 and

Gln C4. Total lactate was significantly increased (Fig. 7).

Acetate Production in the Presence of 4-

hydroxycyanocinnamate

The inhibitor 4-hydroxycyanocinnamate (4-CIN) is a broad

spectrum inhibitor of monocarboxylate transporters at

7 mM [38] and also of the mitochondria pyruvate carrier

[39]. It therefore has a significant effect on mitochondrial

metabolism. The effect of this inhibitor on production of

[2-13C]acetate from [1-13C]glucose is shown in Fig. 8.

Labelling of [2-13C]acetate was significantly decreased by

4CIN but the degree of reduction was relatively minor

compared to the reduction of label into mitochondrially

produced metabolites such as glutamate. Labeling of

metabolites produced in the cytosol, such as aspartate C2

and C3, was increased.

Pattern Recognition

Principal components analysis of the data from 163

experiments [35, 36] generated a two component model

accounting for 69 % of the variance in the data (45 and

24 %, respectively) with a Q2 of 60 %. Q2 is a measure of

the goodness of fit of the model. The two principal com-

ponents of the model are shown in Fig. 9, along with the

variables in a ‘‘Bi-plot’’. Inspection of this plot shows that

acetate C2 lies closest to Lactate C3 and is at the opposite

side of the plot to variables such as Glu C2 and C4, Asp C2

and C4 and Gln C4.

A

B

Fig. 4 Effects of competing 0.74 mM lactate with 2 mM pyruvate in

brain cortical tissue slices for 60 min. a Incorporation of 13C. b Total

metabolite pool sizes. Clear bars 2 mM [3-13C]pyruvate (control);

hatched bars 2 mM [3-13C]pyruvate plus 0.74 mM sodium L-lactate;

spotted bars 2 mM sodium pyruvate plus 0.74 mM [3-13C]- L-lactate.

* significantly different to control (P \ 0.05); # significantly different

to 2 mM [3-13C]pyruvate plus 0.74 mM sodium L-lactate

A

B

Fig. 5 Effects of competing 5 mM acetate with 2 mM pyruvate in

brain cortical tissue slices for 60 min. a Incorporation of 13C. b Total

metabolite pool sizes. Clear bars 2 mM [3-13C]pyruvate (control);

hatched bars 2 mM [3-13C]pyruvate plus 5.0 mM sodium acetate;

spotted bars 2 mM sodium pyruvate plus 5.0 mM [2-13C]- acetate.

* Significantly different to control (P \ 0.05); # significantly different

to 2 mM [3-13C]pyruvate plus 5.0 mM sodium acetate

Neurochem Res (2012) 37:2541–2553 2547

123

Discussion

Uptake of Acetate

Acetate proved to be a poor substrate for the low affinity

lactate transporter MCT4 with a KM [ 2 M, but a rela-

tively good substrate for the high-affinity (mostly) neuronal

transporter MCT2 (KM = 2.5 mM; Table 1, Fig. 2). This

value compares well with the reported KM for MCT1 of

1.6 mM (Table 1). It is therefore unlikely that failure of the

monocarboxylate transporter to take up acetate into neu-

rons is an adequate explanation for neuronal failure to use

acetate as a substrate with any efficacy [21]. Moreover,

neurons express the sodium dependent monocarboxylate

transporter (SMCT1, [40]), which has an affinity for acetate

comparable to that of the monocarboxylate transporters

MCT1 and MCT2 (Table 1). This transporter may there-

fore also be involved in accumulating acetate. SMCT1 is a

neuronal transporter but we cannot rule out involvement of

SMCT2. The affinity of SMCT2 for acetate as far as we can

determine has not yet been established, nor has the brain

cellular location of this transporter, although the transporter

is described as a low affinity transporter and has been

reported in Muller cells in the retina [20]. When slices are

incubated with 5 mM [2-13C]acetate, acetate is accumu-

lated to high levels, giving a calculated concentration of

8.85 (0.99) lmol/h/100 mg protein. However, only 9 % of

this acetate was actually metabolized into Krebs cycle

intermediates or their metabolites under the conditions

used (i.e. in competition with 2 mM pyruvate). It has

previously been reported that acetate is taken up into brain

after cisternal injection at high specific radioactivity to a

greater extent than lactate, pyruvate or glucose [41] and

also rapidly (within 2 min) incorporated into the lipid

fraction. The fact that acetate is accumulated further

A

B

Fig. 6 Effects of competing 5 mM acetate with 0.74 mM lactate in

brain cortical tissue slices for 60 min. a Incorporation of 13C. b Total

metabolite pool sizes. Clear bars 0.74 mM [3-13C]lactate (control);

hatched bars 0.74 mM [3-13C]lactate plus 5.0 mM sodium acetate;

spotted bars 2 mM sodium lactate plus 5.0 mM [2-13C]-acetate.

* Significantly different to control (P \ 0.05); # significantly different

to 0.74 mM [3-13C]lactate plus 5.0 mM sodium acetate. Slices were

also incubated with 2 mM D-glucose to provide adequate respiratory

support

A

B

Fig. 7 Effects of addition of NAD? and nicotinamide to incorpo-

ration of label supplied from 5 mM [2-13C]acetate. Slices were

incubated with 5 mM D-glucose and 5 mM [2-13C]acetate (clear bars;

control); 250 lM NAD? (hatched bars) or 250 lM nicotinamide

(spotted bars)

2548 Neurochem Res (2012) 37:2541–2553

123

supports a role of SMCT1 in brain acetate transport,

because MCTs only equilibrate intracellular and extracel-

lular pools of monocarboxylates in the absence of sub-

stantial pH gradients [38].

Other workers have shown uptake of acetate by neurons

[8, 13] and also by synaptosomes [42] in contrast to the

conclusions of Waniewski and Martin [21]. Here, we have

shown that the neuronal MCT2 is capable of transporting

acetate, plus provided corroborating evidence that brain

accumulates acetate, suggesting a role for active uptake in

acetate transport.

Production of Acetate

It is apparent from the data that acetate is synthesized in

non-trivial amounts from a range of different substrates

supplied to brain tissue cortical slices. While it might be

expected that acetate be synthesized from 2 mM pyruvate,

since this relatively high substrate concentration is likely to

cause accumulation of acetyl-CoA [43], acetate is also

synthesized from added lactate and also from glucose.

Pyruvate is the most efficacious substrate, producing the

most labeled acetate (mean value *0.35 lmol (h mmol

substrate 100 mg protein)-1, but lactate (used at 5 mM

concentration [23]) also produces non-trivial amounts

(*0.16 lmol (h mmol substrate 100 mg protein)-1,

exceeded when lower amounts of lactate (0.74 mM) are

used *0.54 lmol (h mmol substrate 100 mg protein)-1.

Glucose is also a good source when adjusted for the

amount of label (*0.22 lmol/(90 min mmol substrate

100 mg protein)-1; data taken from the control experiment

with 4-CIN using 5 mM glucose).

Acetate is made by hydrolysis of acety-lCoA by the

ubiquitous enzyme acetylCoA hydrolase [44] an enzyme

which may be product-stimulated by acetate [45, 46].

Acetyl-CoA can come from citrate via the ATP-citrate

lyase an enzyme which is expressed in both neurons and

glia [47].

Free acetate has been detected in brain by NMR in areas

of ischemia and ischemic penumbra [48] but not in normal

brain tissue in vivo. Levels of free acetate have been

measured by enzymatic assay at 1.51 ± 0.34 lmol/g tissue

in mouse brain, significantly higher than levels in liver

[49]. Acetate is reported as a common metabolite across a

Fig. 8 Effects of 4-hydroxycyanocinnamate on incorporation of 13C

in brain cortical tissue slices incubated with [1-13C]glucose for

90 min. Clear bars, control (N = 4); hatched bars, 7 mM 4-CIN

(N = 4). Values shown are means, while error bars show standard

deviations. * Significantly different to control

Fig. 9 Bi-plot showing principal components and variable clustering.

The figure shows the two principal components generated by the

model plus the individual variables and allows us to interpret the

observations in terms of the variables. Data from 163 experiments

(observations) [35, 36] was used as input, with 15 variables as shown

on the plot. Data were univariance scaled to standardize variance

between the high and low concentration metabolites [37], to ensure

that the 13C labelling and steady state pool size concentrations equally

contributed to the model, and were expressed as the relative change

from the average value obtained from the control group for that

particular experiment. Observations are shown as grey squares,

variables as black stars with labels

Neurochem Res (2012) 37:2541–2553 2549

123

range of primary cell cultures, including in neurons,

astrocytes and oligodendrocytes [50].

Acetate production was not significantly impacted by

inhibition of the mitochondrial pyruvate carrier with the

broad spectrum inhibitor 4-CIN (Fig. 8). This is in keeping

with the cytosolic location of acetyl-CoA hydrolase [46].

Inspection of Fig. 9 shows that acetate C2 is on the

opposite side of the principal components plot to variables

such as GluC2, C4, Asp C2 and C3 and Gln C4 indicating

that production of acetate occurs when Krebs cycle flux is

low. Acetate C2 is located near to lactate C3, indicating

that acetate production relates to pyruvate clearance. One

could go a step further and infer that acetate C2 production

relates to clearance of acetyl-CoA; i.e. high levels of

acetyl-CoA due to lowered flux in the Krebs cycle corre-

spond to high rates of acetate C2 production.

Metabolism of Acetate

It is apparent from the data that acetate is used to make

glutamine. This likely occurs in the so-called ‘‘small

compartment’’ and has been the subject of many previous

reports [15, 51–54]. Acetate has previously been reported

to label citrate. Citrate is problematic to study by NMR due

to its role in chelation of paramagnetic ions which results in

broadening of the NMR spectrum. It is possible to reduce

the linewidth of the two citrate CH2 carbons in the 13C

NMR spectrum through addition of the chelating agent

EDTA [55]. Here, we found that addition of lactate to

2 mM pyruvate did not significantly alter incorporation of

label into citrate, while addition of unlabelled acetate

almost doubled incorporation of label from [3-13C]pyru-

vate into citrate. By contrast, very little labelling of citrate

was detected when label was supplied as 5 mM

[2-13C]acetate (less than 4 % of that labelled by pyruvate).

Very little label was incorporated into citrate when

0.74 mM [3-13C]lactate was supplied as labelled substrate

and this incorporation was not altered by addition of ace-

tate. Incorporation of label into citrate from [2-13C]acetate,

although minimal, was reduced further by addition of

NAD? and nicotinamide.

Citrate has been shown to be synthesised in both neu-

rons and astrocytes [56], with astrocytes exporting a much

higher fraction of citrate than neurons [57]. Waagepetersen

has described two separate mitochondrial compartments in

astrocytes that metabolise acetate; one that produces glu-

tamine and one that preferentially produces citrate [58].

This is in agreement with earlier work showing that the

citrate, which enters the small metabolic pool that also

labels glutamine, does not fully interact with glutamine

labelling from acetate [41], and with work that has sug-

gested further subdivision of the ‘‘small’’ glutamate pool

[59]. Here, it is apparent that acetate is influencing the rate

of synthesis of citrate, since incorporation of label into

citrate from [3-13C]pyruvate doubles when unlabelled

5 mM acetate is present. It is also apparent that the added

acetate does not participate fully in this reaction, as very

little label is incorporated from added 5 mM [2-13C]ace-

tate. This is consistent with the subdivision of the small

glutamate pool proposed previously [41, 59]. There was no

significant increase in anaplerotic activity (as measured by

the difference between label incorporation into Asp C3 vs.

C2) in the presence of added acetate [0.23 (0.07) vs. 0.23

(0.01) lmol/h/100 mg protein] suggesting that the synthe-

sis of citrate was not a net cataplerotic process.

We would suggest that, rather than seeing a reduction in

citrate synthesis, we are seeing an alteration in the break-

down of citrate. Citrate is in an equilibrium with acetyl-

CoA, and with acetate [60]. Therefore, in situations where

acetate is plentiful (and by inference, acetyl-CoA) the

activity of ATP-citrate lyase is reduced. This may explain

why adding acetate appears to increase the amount of cit-

rate that is labeled; it does this by inhibiting the breakdown

of citrate.

Alteration of Acetate Metabolism by NAD?

and Nicotinamide

Modulation of enzyme activity by acetylation is emerging

as a major metabolic control mechanism [61]. The major

acetate metabolising enzymes, acetyl-CoA synthetases 1

and 2 are but one set of enzymes whose activity is con-

trolled by reversible acetylation. De-acetylaction of these

enzymes is under the control of the silent information

regulator enzymes, the NAD?-dependent deacetylases

known as sirtuins [62]. Nicotinamide is a potent non-

competitive inhibitor of SIRT1 [63]while NAD? is a sub-

strate although the effect of altering NAD? on SIRT

activity is not well understood[64]. Extracellular NAD?

has been shown to enter cells and influence metabolic

activity [65, 66]. Here, we conducted preliminary experi-

ments to determine whether or not nicotinamide or NAD?

could influence acetate metabolism.

Both molecules decreased the amount of acetate being

used as a substrate for glutamine synthesis along with a

reduction in label incorporation into citrate, which we have

postulated is due to increased breakdown of citrate by

ATP-citrate lyase. This points to a lower availability of

acetyl-CoA.

This is important information to consider when inter-

preting data acquired using the acetate/glucose model of

metabolism. The original papers describing the model all

contained a warning that using this approach from a strictly

compartmentalised viewpoint had limitations [55, 67]. The

possibilities opened up by considering changes in metab-

olism of acetate caused by cellular energy and acetylation

2550 Neurochem Res (2012) 37:2541–2553

123

status offer a potentially more interesting and useful

interpretation of data obtained using this model.

Summary

We have shown that acetate is transported by the neuronal

MCT2, that acetate is actively accumulated by tissue slices

but not catabolised to the same degree as other mono-

carboxylates such as pyruvate or lactate. A likely expla-

nation for the lack of enthusiasm for catabolism of acetate

resides in the control of metabolism via acetylation of

enzymes, in particular acetylation of acetyl-CoA synthetase

[5]. Our data suggest that levels of acetyl-CoA are strictly

controlled and that acetate and citrate are key metabolites

in this control.

AcetylCo-A is a central metabolite in many key reac-

tions (see Fig. 1) and plainly we have much further to go to

reach a full understanding of how these reactions are reg-

ulated in the brain.

Acknowledgments The authors are grateful to Dr Ann Kwan of the

School of Molecular Biosciences at the University of Sydney for

expert technical assistance. This work was supported by the Austra-

lian National Health and Medical Research Council (Grant #568767

to CR and Fellowship to CR #630516).

References

1. Gonzalez SV, Nguyen NHT, Rise F, Hassel B (2005) Brain

metabolism of exogenous pyruvate. J Neurochem 95:284–293

2. Cox DWG, Bachelard HS (1988) Partial attenuation of dentate

granule cell evoked activity by the alternative substrates, lactate

and pyruvate: evidence for a postsynaptic action. Exp Brain Res

69:368–372

3. Tarasenko AS, Linetska MV, Storchak LG, Himmelreich NH

(2006) Effectiveness of extracellular lactate/pyruvate for sus-

taining synaptic vesicle proton gradient generation and vesicular

accumulation of GABA. J Neurochem 99:787–796

4. Berl S, Frigyesi TL (1969) Turnover of glutamate glutamine

aspartate and GABA labeled with 1–14C acetate in caudate

nucleus thalamus and motor cortex (cat). Brain Res 12:444–455

5. Starai VJ, Escalante-Semerena JC (2004) Acetyl-coenzyme A

synthetase (AMP forming). Cell Mol Life Sci 61:2020–2030

6. Dadamo AF, Yatsu FM (1966) Acetate metabolism in nervous

system. N-acetyl-L-aspartic acid and biosynthesis of brain lipids.

J Neurochem 13:961

7. Soliman ML, Rosenberger TA (2011) Acetate supplementation

increases brain histone acetylation and inhibits histone deace-

tylase activity and expression. Mol Cell Biochem 352:173–180

8. Carroll PT (1997) Evidence to suggest that extracellular acetate is

accumulated by rat hippocampal cholinergic nerve terminals for

acetylcholine formation and release. Brain Res 753:47–55

9. Ariyannur PS, Moffett JR, Madhavarao CN, Arun P, Vishnu N,

Jacobowitz DM, Hallows WC, Denu JM, Namboodiri AMA

(2010) Nuclear-cytoplasmic localization of acetyl coenzyme A

synthetase-1 in the rat brain. J Comp Neurol 518:2952–2977

10. Tucek S (1967) Subcellular distribution of acetyl-CoA synthe-

tase, ATP citrate lyase, citrate synthase, choline acetyltransferase,

fumarate hydratase and lactate dehydrogenase in mammalian

brain tissue. J Neurochem 14:531–545

11. Hirschey MD, Shimazu T, Capra JA, Pollard KS, Verdin E

(2011) SIRT1 and SIRT3 deacetylate homologous substrates:

AceCS1,2 and HMGCS1,2. Aging-US 3:635–642

12. Ramadori G, Lee CE, Bookout AL, Lee S, Williams KW,

Anderson J, Elmquist JK, Coppari R (2008) Brain SIRT1: ana-

tomical distribution and regulation by energy availability.

J Neurosci 28:9989–9996

13. Brand A, Richter-Landsberg C, Leibfritz D (1997) Metabolism of

acetate in rat brain neurons, astrocytes and cocultures: metabolic

interactions between neurons and glia cells, monitored by NMR

spectroscopy. Cell Mol Biol 43:645–657

14. Chapa F, Cruz F, Garcia-Martin ML, Garcia-Espinosa MA,

Cerdan S (1999) Metabolism of (1–13C) glucose and (2–13C,

2–2H3) acetate in the neuronal and glial compartments of the

adult rat brain as detected by 13C, 2H NMR spectroscopy.

Pergamon-Elsevier Science Ltd, Wierzba

15. Badar-Goffer R, Bachelard H, Morris P (1990) Cerebral metab-

olism of acetate and glucose studied by 13C NMR spectroscopy.

Biochem J 266:133–139

16. Rae C, Hare N, Bubb WA, McEwan SR, Broer A, McQuillan JA,

Balcar VJ, Conigrave AD, Broer S (2003) Inhibition of glutamine

transport depletes glutamate and GABA neurotransmitter pools:

further evidence for metabolic compartmentation. J Neurochem

85:503–514

17. Bergersen LH, Magistretti PJ, Pellerin L (2005) Selective post-

synaptic co-localisation of MCT2 with AMPA receptor GluR2/3

subunits at excitatory synapses exhibiting AMPA receptor traf-

ficking. Cereb Cortex 15:361–370

18. Pierre K, Pellerin L (2005) Monocarboxylate transporters in the

central nervous system: distribution, regulation and function.

J Neurochem 94:1–14

19. Rafiki A, Boulland JL, Halestrap AP, Ottersen OP, Bergersen LH

(2003) Highly differential expression of the monocarboxylate

transporters MCT2 and MCT4 in the developing rat brain. Neu-

roscience 122:677–688

20. Martin PM, Dun Y, Mysona B, Ananth S, Roon P, Smith SB,

Ganapathy V (2007) Expression of the sodium-coupled mono-

carboxylate transporters SMCT1 (SLC5A8) and SMCT2

(SLC5A12) in retina. Invest Ophthalmol Vis Sci 48:3356–3363

21. Waniewski R, Martin D (1998) Preferential utilization of acetate

by astrocytes is attributable to transport. J Neurosci 18:5225–

5233

22. Berl S, Nunez R, Colon AD, Clarke DD (1983) Acetylaction of

synaptosomal protein—inhibition by veratridine. J Neurochem

40:176–183

23. Rae C, Nasrallah F, Broer S (2009) Metabolic effects of blocking

lactate transport in brain cortical tissue slices using an inhibitor

specific to MCT1 and MCT2. Neurochem Res 34:1783–1791

24. Nasrallah F, Griffin JL, Balcar VJ, Rae C (2007) Understanding

your inhibitions. Modulation of brain cortical metabolism by

GABA-B receptors. J Cereb Blood Flow Metab 27:1510–1520

25. Broer S, Broer A, Hansen JT, Bubb WA, Balcar VJ, Nasrallah

FA, Garner B, Rae C (2007) Alanine metabolism, transport and

cycling in the brain. J Neurochem 102:1758–1770

26. Broer S (2010) Xenopus laevis oocytes. Methods Mol Biol 637:

295–310

27. Broer S, Broer A, Schneider HP, Stegen C, Halestrap AP, Deit-

mer JW (1999) Characterization of the high-affinity monocarb-

oxylate transporter MCT2 in Xenopus laevis oocytes. Biochem J

341:529–535

28. Broer S, Schneider HP, Broer A, Rahman B, Hamprecht B,

Deitmer JW (1998) Characterization of the monocarboxylate

transporter 1 expressed in Xenopus laevis oocytes by changes in

cytosolic pH. Biochem J 333:167–174

Neurochem Res (2012) 37:2541–2553 2551

123

29. Tosco M, Orsenigo MN, Gastaldi G, Faelli A (2000) An endoge-

nous monocarboxylate transport in Xenopus laevis oocytes. Am J

Physiol-Regul Integr Comp Physiol 278:R1190–R1195

30. McIlwain H, Bachelard H (1985) Biochemistry and the central

nervous system. Churchill Livingstone, Edinburgh, pp 7–29

31. Le Belle JE, Harris NG, Williams SR, Bhakoo KK (2002) A

comparison of cell and tissue extraction techniques using high-

resolution 1H NMR spectroscopy. NMR Biomed 15:37–44

32. Kupce E, Freeman R (1995) Adiabatic pulses for wideband

inversion and broadband decoupling. J Mag Reson A 115:

273–276

33. Skinner TE, Bendall MR (1997) A phase-cycling algorithm for

reducing sidebands in adiabatic decoupling. J Magn Reson 124:

474–478

34. Rae C, Lawrance ML, Dias LS, Provis T, Bubb WA, Balcar VJ

(2000) Strategies for studies of potentially neurotoxic mecha-

nisms involving deficient transport of L-glutamate: antisense

knockout in rat brain in vivo and changes in the neurotransmitter

metabolism following inhibition of glutamate transport in guinea

pigs brain slices. Brain Res Bull 53:373–381

35. Nasrallah FA, Balcar VJ, Rae CD (2011) Activity dependent

GABA release controls brain cortical tissue slice metabolism.

J Neurosci Res 89:1935–1945

36. Nasrallah F, Griffin JL, Balcar VJ, Rae CD (2009) Understanding

your inhibitions. Effects of GABA and GABAA receptors on

brain cortical metabolism. J Neurochem 108:57–71

37. Wold S, Antti H, Lindgren F, Ohman J (1998) Orthogonal signal

correction of near-infrared spectra. Chemom Intell Lab Syst

44:175–185

38. Halestrap AP, Price NT (1999) The proton-linked monocarb-

oxylate transporter (MCT) family: structure, function and regu-

lation. Biochem J 343:281–299

39. Halestrap AP, Denton RM (1974) Specific inhibition of pyruvate

transport in rat liver mitochondria and human erythrocytes by

a-cyano-4-hydroxycinnamate. Biochem J 138:313–316

40. Martin PM, Gopal E, Ananth S, Zhuang L, Itagaki S, Prasad BM,

Smith SB, Prasad PD, Ganapathy V (2006) Identify of

SMCT1(SLC5A8) as a neuron-specific Na?-coupled transporter

for active uptake of L-lactate and ketone bodies in the brain.

J Neurochem 98:279–288

41. Tucek S, Cheng SC (1974) Provenance of acetyl group of ace-

tylcholine and compartmentation of acetyl-CoA and Krebs cycle

intermediates in brain in vivo. J Neurochem 22:893–914

42. Berl S, Clarke DD, Colon A, Nunez R (1983) H-3-Labelled acetate

incorporation into protein of synaptosomes—effect of Na? chan-

nel toxins. Hoppe Seylers Z Physiol Chem 364:1098–1099

43. Nakamura R, Cheng SC, Naruse H (1970) A study on precursors

of acetyl moiety of acetylcholine in brain slices—observations on

compartmentalization of acetyl-coenzyme A pool Biochem J

118:443

44. Kirkby B, Roman N, Kobe B, Kellie S, Forwood JK (2010)

Functional and structural properties of mammalian acyl-coen-

zyme A thioesterases. Prog Lipid Res 49:366–377

45. Robinson JB, Mahan DE, Koeppe RE (1976) Studies on rat brain

acyl-coenzyme-A hydrolase (short chain). Biochem Biophys Res

Commun 71:959–965

46. Matsuda T, Yonehara N, Yoshida H (1976) Metabolism of pre-

cursors of acetylcholine synthesis in central nervous system—

some properties of acetyl-CoA hydrolase in cytoplasmic soluble

fraction from rat brain. Folia Pharmacol Jap 72:P63

47. Beigneux AP, Kosinski C, Gavino B, Horton JD, Skarnes WC,

Young SG (2004) ATP-citrate lyase deficiency in the mouse.

J Biol Chem 279:9557–9564

48. Alf MF, Lei H, Berthet C, Hirt L, Gruetter R, Mlynarik V (2012)

High-resolution spatial mapping of changes in the neurochemical

profile after focal ischemia in mice. NMR Biomed 25:247–254

49. Madhavarao CN, Arun P, Moffett JR, Szucs S, Surendran S,

Matalon R, Garbern J, Hristova D, Johnson A, Jiang W, Nam-

boodiri MAA (2005) Defective N-acetylaspartate catabolism

reduces brain acetate levels and myelin lipid synthesis in Cana-

van’s disease. Proc Natl Acad Sci USA 102:5221–5226

50. Urenjak J, Williams SR, Gadian DG, Noble M (1993) Proton

nuclear magnetic resonance spectroscopy unambiguously identi-

fies different neural types. J Neurosci 13:981–989

51. Berl S, Clarke DD, Nicklas WJ (1970) Compartmentation of

citric acid cycle metabolism in brain—effect of aminooxyacetic

acid, ouabain and Ca2? on labelling of glutamate, glutamine,

aspartate and GABA by 1-C-14acetate, U-C-14glutamate and

U-C-14 aspartate. J Neurochem 17:999–1007

52. Berl S, Lajtha A, Waelsch H (1961) Amino acid and protein

metabolism -VI Cerebral compartments of glutamic acid metab-

olism. J Neurochem 7:186–197

53. Sonnewald U, Westergaard N, Hassel B, Muller TB, Unsgard G,

Fonnum F, Hertz L, Schousboe A, Petersen SB (1993) NMR

spectroscopy studies of 13C acetate and 13C glucose metabolism

in neocortical astrocytes: evidence for mitochondrial heteroge-

neity. Dev Neurosci 15:351–358

54. Wyss MT, Magistretti PJ, Buck A, Weber B (2011) Labeled

acetate as a marker of astrocytic metabolism. J Cereb Blood Flow

Metab 31:1668–1674

55. Badar-Goffer R, Bachelard H, Morris P (1992) Neuronal-glial

metabolism under depolarising conditions. Biochem J 282:225–

230

56. Westergaard N, Sonnewald U, Unsgard G, Peng L, Hertz L,

Schousboe A (1994) Uptake, release and metabolism of citrate in

neurons and astrocytes in primary cultures. J Neurochem 62:

1727–1733

57. Sonnewald U, Westergaard N, Krane J, Unsgard G, Petersen SB,

Schousboe A (1991) First direct demonstration of preferential

release of citrate from astrocytes using C-13 NMR spectroscopy

of cultured neurons and astrocytes. Neurosci Lett 128:235–239

58. Waagepetersen HS, Sonnewald U, Larsson OM, Schousboe A

(2001) Multiple compartments with different metabolic charac-

teristics are involved in biosynthesis of intracellular and released

glutamine and citrate in astrocytes. Glia 35:246–252

59. Cheng SC, Nakamura R (1972) Metabolism related to tricar-

boxylic acid cycle in rat brain slices—observations on CO2 fix-

ation and metabolic compartmentation. Brain Res 38:355

60. Wellen KE, Hatzivassiliou G, Sachdeva UM, Bui TV, Cross JR,

Thompson CB (2009) ATP-citrate lyase links cellular metabolism

to histone acetylation. Science 324:1076–1080

61. Shin JY, Zhang D, Chen D (2011) Reversible acetylation of

metabolic enzymes celebration: SIRT2 and p300 join the party.

Mol Cell 43:3–5

62. Hallows WC, Lee S, Denu JM (2006) Sirtuins deacetylate and

activate mammalian acetyl-CoA synthetases. Proc Natl Acad Sci

USA 103:10230–10235

63. Bitterman KJ, Anderson RM, Cohen HY, Latorre-Esteves M,

Sinclair DA (2002) Inhibition of silencing and accelerated aging

by nicotinamide, a putative negative regulator of yeast Sir2 and

human SIRT1. J Biol Chem 277:45099–45107

64. Canto C, Auwerx J (2012) Targeting Sirtuin 1 to improve

metabolism: all you need is NAD(?)? Pharmacol Rev 64:

166–187

65. Billington RA, Travelli C, Ercolano E, Galli U, Roman CB,

Grolla AA, Canonico PL, Condorelli F, Genazzani AA (2008)

Characterization of NAD uptake in mammalian cells. J Biol

Chem 283:6367–6374

66. Pittelli M, Felici R, Pitozzi V, Giovannelli L, Bigagli E, Cialdai

F, Romano G, Moroni F, Chiarugi A (2011) Pharmacological

effects of exogenous NAD on mitochondrial bioenergetics, DNA

repair, and apoptosis. Mol Pharmacol 80:1136–1146

2552 Neurochem Res (2012) 37:2541–2553

123

67. Clarke DD (1999) Acetate and fluoroacetate as possible markers

for glial metabolism: an historical perspective. J Neurochem

72:S2

68. Galic S, Schneider HP, Broer A, Deitmer JW, Broer S (2003) The

loop between helix 4 and helix 5 in the monocarboxylate trans-

porter MCT1 is important for substrate selection and protein

stability. Biochem J 376:413–422

69. Carpenter L, Halestrap AP (1994) The kinetics, substrate and

inhibitor specificity of the lactate transporter of ehrlich-lettre

tumor cells studied with the intracellular pH indicator BCECF.

Biochem J 304:751–760

70. Broer S, Rahman B, Pellegri G, Pellerin L, Martin JL, Verleys-

donk S, Hamprecht B, Magistretti PJ (1997) Comparison of

lactate transport in astroglial cells and transporter 1 (MCT 1)

expressing Xenopus laevis oocytes. Expression of two different

monocarboxylate transporters in astroglial cells and neurons.

J Biol Chem 272:30096–30102

71. Fox JEM, Meredith D, Halestrap AP (2000) Characterisation of

human monocarboxylate transporter 4 substantiates its role in

lactic acid efflux from skeletal muscle. J Physiol Lond 529:

285–293

72. Ovens MJ, Manoharan C, Wilson MC, Murray CM, Halestrap AP

(2010) The inhibition of monocarboxylate transporter 2 (MCT2)

by AR-C155858 is modulated by the associated ancillary protein.

Biochem J 431:217–225

73. Lin RY, Vera JC, Chaganti RSK, Golde DW (1998) Human

monocarboxylate transporter 2 (MCT2) is a high affinity pyruvate

transporter. J Biol Chem 273:28959–28965

74. Dimmer KS, Friedrich B, Lang F, Deitmer JW, Broer S (2000)

The low-affinity monocarboxylate transporter MCT4 is adapted

to the export of lactate in highly glycolytic cells. Biochem J

350:219–227

75. Babu E, Ramachandran S, CoothanKandaswamy V, Elangovan S,

Prasad PD, Ganapathy V, Thangaraju M (2011) Role of SLC5A8,

a plasma membrane transporter and a tumor suppressor, in the

antitumor activity of dichloroacetate. Oncogene 30:4026–4037

76. Miyauchi S, Gopal E, Fei YJ, Ganapathy V (2004) Functional

identification of SLC5A8, a tumor suppressor down-regulated in

colon cancer, as a Na?-coupled transporter for short-chain fatty

acids. J Biol Chem 279:13293–13296

77. Gopal E, Umapathy NS, Martin PM, Ananth S, Gnana-Prakasam

JP, Becker H, Wagner CA, Ganapathy V, Prasad PD (2007)

Cloning and functional characterization of human SMCT2

(SLC5A12) and expression pattern of the transporter in kidney.

Biochim Biophys Acta 1768:2690–2697

Neurochem Res (2012) 37:2541–2553 2553

123