Skip to main content
Intended for healthcare professionals
Free access
Review article
First published November 2003

Energy Substrates for Neurons during Neural Activity: A Critical Review of the Astrocyte-Neuron Lactate Shuttle Hypothesis

Abstract

Glucose had long been thought to fuel oxidative metabolism in active neurons until the recently proposed astrocyte-neuron lactate shuttle hypothesis (ANLSH) challenged this view. According to the ANLSH, activity-induced uptake of glucose takes place predominantly in astrocytes, which metabolize glucose anaerobically. Lactate produced from anaerobic glycolysis in astrocytes is then released from astrocytes and provides the primary metabolic fuel for neurons. The conventional hypothesis asserts that glucose is the primary substrate for both neurons and astrocytes during neural activity and that lactate produced during activity is removed mainly after neural activity. The conventional hypothesis does not assign any particular fraction of glucose metabolism to the aerobic or anaerobic pathways. In this review, the authors discuss the theoretical background and critically review the experimental evidence regarding these two hypotheses. The authors conclude that the experimental evidence for the ANLSH is weak, and that existing evidence and theoretical considerations support the conventional hypothesis.
Until recently, glucose had stood unchallenged as the principal metabolic substrate of the mature brain (for example, see McIlwain and Bachelard, 1985; Sokoloff, 1989). The primacy of glucose, though, has been questioned recently by some investigators who view lactate as a substrate that active neurons prefer over glucose (Magistretti, 1999; Magistretti et al., 1999). The brain can consume lactate as a substrate, as has been demonstrated by studies showing that the brain uses lactate during hypoglycemia or during periods of elevated blood lactate (Ide et al., 2000; Nemoto et al., 1974). Also, some in vitro studies have shown that lactate supports neural activity in brain tissues in the absence of glucose (e.g., Izumi et al., 1997; Schurr et al., 1988). However, because lactate does not pass through the blood-brain barrier nearly as well as glucose (e.g., McIlwain and Bachelard, 1985), lactate cannot serve the brain as a blood-borne substrate the way glucose does. In the mid 1990s, an astrocyte-neuron lactate shuttle hypothesis (ANLSH) was proposed that assigned a major metabolic role to brain-derived lactate (Pellerin and Magistretti, 1994; Tsacopoulos and Magistretti, 1996). According to this hypothesis, lactate is produced in an activity-dependent and glutamate-mediated manner by astrocytes and is then transferred to and used by active neurons (Pellerin et al., 1998a). In greater detail, the ANLSH postulates that neuronal activation increases the extracellular concentration of brain glutamate, which astrocytes take up via high affinity, Na+-dependent transporters. Increases in glutamate and Na+ in astrocytes activate glutamine synthetase and Na+-K+ ATPase, respectively. Activation of Na+-K+ ATPase spurs astrocytic ATP consumption. This leads to activation of anaerobic glycolysis1 in astrocytes, and the resulting lactate is transported out of astrocytes and into neurons, where it fuels the activity-related energy needs of neurons (for reviews, see Bouzier-Sore et al., 2002; Magistretti, 1999, 2000; Magistretti and Pellerin, 1999; Magistretti et al., 1999). As currently expressed, the ANLSH applies only to glutamatergic neurons and does not address how activity-related energy demands are met in nonglutametergic neurons.
Early support for the ANLSH came from a study by Pellerin and Magistretti (1994) that showed that glutamate increased 2-deoxy-D-[1,2-3H]glucose uptake and lactate release in cultured astrocytes. Later, Magistretti and colleagues (Bittar et al., 1996) proposed that the distribution of lactate dehydrogenase (LDH) isoforms in brain cells might favor lactate oxidation in neurons. Also, results from several other studies were used as evidence that lactate may be preferred over glucose as an energy substrate for the brain (e.g., Hu and Wilson, 1997; Larrabee, 1995; Poitry-Yamate et al., 1995; Schurr et al., 1999). In 1998, Shulman and coworkers (Sibson et al., 1998) found a 1:1 stoichiometry between oxidative glucose metabolism and glutamate cycling. Magistretti and coworkers (Magistretti, 1999; Magistretti and Pellerin, 1999; Magistretti et al., 1999) used this finding to suggest that glucose is used predominantly, if not exclusively, by astrocytes and that neurons use lactate released from astrocytes as their primary substrate during activity. In 2002, Magistretti and colleagues (Bouzier-Sore et al., 2002) reviewed the evidence supporting the ANLSH and proposed again that astrocytes provide lactate as an energy substrate for neurons, especially during periods of enhanced synaptic activity.
The conventional view of glucose metabolism (Fig. 1) and the ANLSH (Fig. 2) differ most noticeably in (1) the sites and modes of glucose use, (2) the sites of lactate production and use, and (3) the timing of lactate use. The conventional hypothesis contends that activity-induced energy demand is met predominantly by metabolizing glucose oxidatively (e.g., Chih et al., 2001b; Dienel and Hertz, 2001; McIlwain and Bachelard, 1985; Sokoloff, 1989). In the conventional view, increased Na+-K+ ATPase activity in brain cells resulting from neural activation causes increased ATP use, which activates the glycolytic pathway and increases oxidative metabolism in both neurons and astrocytes. Production of lactate in the brain during neural activity is considered to be a result of glycolytic activity transiently exceeding the rate of oxidative metabolism (e.g., Prichard et al., 1991). In this view, a buildup of lactate is viewed as potentially detrimental (e.g., Prichard et al., 1991; Siesjö, 1982); therefore, lactate must be either removed via the circulation or consumed by inactive brain cells after neural activity (e.g., Chih et al., 2001a, 2001b). The conventional hypothesis also maintains that both neurons and astrocytes can produce and use lactate (Dringen et al., 1993a; Chih et al., 2001b; Dienel and Hertz, 2001). However, lactate is used only when the prevailing glycolytic rate is low.
FIG. 1. Schematic illustration of glucose metabolism during neural activity in neurons and astrocytes according to the conventional hypothesis. According to this hypothesis, glucose use increases in both neurons and astrocytes during activity. Heightened Na+-K+ ATPase activity in both cell types during neural activity increases ATP consumption and stimulates glucose use by activating glycolytic enzymes. Glucose is transported into neurons and astrocytes via the glucose transporters GLUT3 and GLUT1, respectively. Glucose is metabolized to pyruvate (hereafter designated as glycolysis) to produce 2 ATPs in either cell type. Pyruvate is either converted to lactate (anaerobic glycolysis) by LDH or taken up by mitochondria, where it enters the TCA cycle and is metabolized via oxidative metabolism. Lactate produced by both neurons and astrocytes is transported to the extracellular space via MCT. Increased energy demands resulting from neural activity are met by ATP generated from both glycolysis and oxidative metabolism. The conventional hypothesis does not ascribe any particular fraction of glucose metabolism to either the aerobic or anaerobic pathways. O.P., oxidative phosphorylation; Glc, glucose; Pyr, pyruvate; LDH, lactate dehydrogenase; Lac, lactate; MCT, monocarboxylate transporters.
FIG. 2. Schematic illustration of glucose metabolism during neural activity in neurons and astrocytes, according to the ANLSH. Activation of neurons causes release of glutamate, which is then taken up by astrocytes. Cotransport of Na+ with glutamate into astrocytes (not shown) activates astrocytic Na+-K+ ATPase, which in turn stimulates ATP consumption and activates glycolytic enzymes in astrocytes. According to the ANLSH, activity-induced glucose uptake occurs predominantly, if not exclusively, in astrocytes, where glucose is metabolized anaerobically to lactate. The two ATPs produced by anaerobic glycolysis are used for glutamine synthesis and Na+-K+ ATPase activity. Lactate produced by astrocytes is released to the extracellular space and taken up by neurons to fuel neuronal oxidative metabolism. One stipulation of the ANLSH is that glutamate cycling, which includes uptake of glutamate and conversion of glutamate to glutamine, is fueled entirely by astrocytic anaerobic glycolysis, and is the driving force for glucose use during neural activity. The ANLSH also contends that active neurons prefer to use lactate released from astrocytes rather than ambient glucose as an energy substrate. X's in metabolic pathways emphasize that, for the ANLSH, increased neural activity does not activate neuronal glycolysis, and that activity-induced increases in astrocytic energy metabolism are nonoxidative. ANLSH, astrocyte-neuron lactate shuttle hypothesis; Glu, glutamate; Gln, glutamine; O.P., oxidative phosphorylation; Glc, glucose; Pyr, pyruvate; LDH, lactate dehydrogenase; Lac, lactate; MCT, monocarboxylate transporters.
In contrast, the ANLSH proposes that activity-induced increases in glucose uptake occur mainly in astrocytes and that astrocytes consume glucose anaerobically to produce lactate (Magistretti, 1999; Magistretti et al., 1999). The ANLSH postulates that neural activity increases anaerobic glycolysis in astrocytes without enhancing the neuronal glycolytic pathway (Magistretti et al., 1999) and that, during activity, astrocytes release lactate destined to fuel oxidative metabolism in active neurons. Thus, with regard to lactate use, the two hypotheses not only differ in the timing of lactate use and in the cell types using lactate, but also in the rationale for why lactate is produced and used in the brain.
The features of both hypotheses deserve a deeper examination, given the fundamental importance of glucose use to our understanding of brain energy metabolism and the widespread acceptance of the ANLSH by neuroscientists. In this review, we discuss the theoretical background and critically review the experimental evidence regarding the conventional hypothesis and the ANLSH. We find little convincing experimental support for the ANLSH and that existing evidence and theoretical considerations support the conventional hypothesis that brain cells, including neurons, consume glucose directly to meet their activity-related energy needs. In Part I we explain, from the perspectives of enzyme kinetics, thermodynamics, and substrate availability, that glucose is probably the primary substrate used by neurons during activity. In Part II we critically examine current experimental evidence to evaluate the validity of the two hypotheses. In Part III we summarize the role of lactate as an energy substrate for the brain.

PART I: COMPARISON OF GLUCOSE AND LACTATE AS ENERGY SUBSTRATES FOR NEURONS DURING ACTIVITY

Availability of glucose and lactate to neurons during activity

The contention of the ANLSH that neural activity does not stimulate the neuronal glycolytic pathway is valid if glucose is not available to neurons during activity or if neurons have little capacity to use glucose. However, although neuronal processes are not directly coupled to the bloodstream as astrocyte processes are, much indirect evidence suggests that glucose is readily available to and easily used by neurons. First, synaptic membranes have glucose transporters in abundance (Leino et al., 1997; McCall et al., 1994). Second, neurons have GLUT3 as their dominant glucose transporter. GLUT3 transports glucose seven times faster than GLUT1, which is the dominant glucose transporter in astrocytes (Vannucci et al., 1997). Third, glucose, which is evenly distributed between the intra- and extracellular compartments of the brain (Pfeuffer et al., 2000), is present at extracellular concentrations that are well above the Km for hexokinase (0.04 mM) (Lowry and Passonneau, 1964). This is true whether the brain is at rest or stimulated in vivo (see data of Fray et al., 1996; Hu and Wilson, 1997). For example, resting extracellular glucose concentrations measured directly with glucose-sensitive microelectrodes (Hu and Wilson, 1997; Silver and Erecinska, 1994) are 2.4 to 2.6 mM. Fourth, neurons have high activities of glycolytic enzymes (Cimino et al., 1998; Lai et al., 1999; Kao-Jen and Wilson, 1980; Wilson, 1972). Finally, a recent in vivo study (Wittendorp-Rechenmann et al., 2001) showed that 14C-labeled 2-deoxyglucose (2-DG) was taken up approximately equally by both astrocytes (42% to 58%) and neurons (54% to 62%). Many in vitro studies also have shown that both neurons and astrocytes can use glucose as a substrate (Itoh et al., 2003; Peng et al., 1994; Vicario et al., 1993; Walz and Mukerji, 1988) and that neurons may increase drastically their glucose use during periods of heightened energy demand (Peng et al., 1994; Walz and Mukerji, 1988). Thus the existing evidence suggests that neurons use glucose directly during neural activity in situ.
Neurons and astrocytes also possess the cellular machinery necessary for using lactate. For example, both cell types have LDH (Bittar et al., 1996). In addition, both neurons and astrocytes have monocarboxylate transporters (MCTs) (Broer et al., 1997; Pierre et al., 2002). However, as discussed below, the lack of a mechanism to activate LDH and slow increases in lactate levels during activity hinder the use of lactate during activity.

Responses of glycolytic enzymes and LDH to increases in energy demand during neural activity

Lactate must compete with glucose for use if glucose is readily available to neurons. Because energy demands fluctuate rapidly between active and inactive states in neurons, the key question regarding substrate use is whether glucose or lactate metabolism offers a quicker response to sudden increases in energy demand. Glucose metabolism via the glycolytic pathway appears well adapted to meet increased cellular energy demands because it has the important characteristic of being tightly coupled to energy demand (Clarke et al., 1989). For example, during periods of low energy demand, the glycolytic rate is low because key glycolytic enzymes, including hexokinase and phosphofructokinase, are inhibited by high ATP levels (Clarke et al., 1989). For brain hexokinase, this inhibition is 97% for resting conditions (Clarke et al., 1989). This basal inhibition gives neurons the potential to respond quickly to increased energy demands. When neurons are active, increased intracellular Na+ and extracellular K+ stimulate neuronal ATP use via Na+-K+ ATPase. The resulting increases in adenine nucleotide and phosphate levels activate the glycolytic pathway and oxidative metabolism at several key control points. The arrangement of these control points allows increased rates of glucose use without increases in glucose concentrations (Peng et al., 1994; Walz and Mukerji, 1988).
In contrast to the responsiveness of glycolytic enzymes to energy status, increased energy demand does not activate LDH. Thus, increased rates of lactate dehydrogenation require either increases in lactate levels or decreases in pyruvate levels. Given that pyruvate levels increase during elevated neural activity (Ferrendelli and McDougal, 1971; Goldberg et al., 1966), extracellular lactate levels must first go up for neuronal lactate use to increase. This requirement has been demonstrated in an in vitro study of cultured cerebellar neurons (Peng et al., 1994). In this study, substrate concentrations were held constant, and energy demand was increased by raising the extracellular K+ concentration. The rate of lactate oxidation in the cultured neurons did not increase significantly with increased K+. In contrast, both glucose and pyruvate oxidation increased significantly with elevated K+, despite unchanged ambient glucose and pyruvate levels. Thus, rapid increases in extracellular lactate must occur if lactate is to fuel the increased neuronal oxidative metabolism seen with elevated neural activity.
However, increases in extracellular lactate in situ lag behind neural activity (Fellows et al., 1993; Fray et al., 1996; Hu and Wilson, 1997) and are sometimes not seen until after activity has ended (Hu and Wilson, 1997). For example, extracellular lactate does not increase in the rat brain for the first 2.5 minutes of a 5 minute tail pinch stimulation (Fellows et al., 1993). By the end of the 5 minute stimulation, lactate levels had increased 40% to 50% and did not peak (55% to 80%) until approximately 5 minutes later. When extracellular lactate changes were examined in the hippocampus with lactate-sensitive electrodes (Hu and Wilson, 1997), lactate levels began increasing only 10 to 12 seconds after 5 seconds of electrical stimulation had ended. Lactate peaked (40% to 100% above control) after another 50 to 60 seconds.
The increases in extracellular lactate noted in the preceding studies are far too slow to meet cellular metabolic needs resulting from neural activity. For example, the half-life for the clearance of extracellular K+ in hippocampal slices after they have been intensely activated electrically is only approximately 5 seconds (Roberts, 1993; Roberts and Feng, 1996). Also, as noted from hemoglobin desaturation, oxygen use during neural activity in situ increases rapidly within the first 1 to 3 seconds of activity (Malonek and Grinvald, 1996; Malonek et al., 1997). Because it rises slowly, extracellular lactate probably cannot fuel these rapid responses. Thus, the available physiological evidence tends not to support lactate's use as an energy substrate early during neural activity.
Unlike lactate, glucose use in neurons can rise rapidly because key glycolytic enzymes are highly sensitive to declines in cellular energy status, and rate changes only require the diffusion of small metabolites within the cell. Many studies have shown that neurons can increase their glycolytic rates in response to increased energy demands (e.g., Peng et al., 1994; Sokoloff, 1999; Walz and Mukerji, 1988). Measurements of oxygen-glucose metabolism in the human visual cortex also suggest that, during visual stimulation, oxidative metabolism and glycolysis in neurons increase concurrently (Gjedde and Marrett, 2001; Gjedde et al., 2002). Therefore, the available evidence suggests that neural activity should activate the glycolytic pathway in both neurons and astrocytes and not in astrocytes alone, as the ANLSH proposes (Magistretti, 1999; Magistretti et al., 1999).

Suppression of lactate oxidation by increased glycolysis

The above analysis suggests that the faster activation of the glycolytic pathway makes glucose more likely than lactate to be the substrate used during the initial stages of neural activity. An additional question is whether lactate can become an important energy source for neurons during prolonged activation when lactate levels increase. Successful use of lactate during prolonged activation would come at the expense of glucose because both lactate dehydrogenation and glycolysis need NAD+ (Fig. 3). Also, the cotransport of lactate and H+ (Schneider et al., 1993) into neurons would cause acidification, which may inhibit glycolysis.
FIG. 3. Schematic illustration of mechanisms for maintaining the redox balance when lactate is the energy substrate. NADH generated from lactate dehydrogenation can only be reoxidized to NAD+ via (1) the G3P shuttle and (2) the malate shuttle. In areas where mitochondria are rare, such as in dendritic spines (see text), NADH generated by LDH must diffuse to areas where mitochondria are present. When energy demand increases, no mechanism is available to increase the activity of the G3P shuttle if lactate is the energy substrate. Also, NADH generated by LDH must compete with NADH generated via the action of glyceraldehyde-3-phosphate dehydrogenase for access to the two NADH shuttles. G3P, glycerol-3-phosphate; LDH, lactate dehydrogenase; Asp, aspartate; E.T.C., electron transport chain; FBP, fructose-1,6-bisphosphate; Mal, malate; OAA, oxaloacetate; GAD-3P, glyceraldehyde-3-phosphate; 1,3-BPG, 1,3-biphosphoglycerate; DHAP, dihydroxyacetone phosphate.
The key issue for brain cells using lactate is the thermodynamic feasibility for conversion of lactate to pyruvate. In the initial stages of neural activation, as glycolytic rates increase, pyruvate (Ferrendelli and McDougal, 1971; Goldberg et al., 1966) and H+ (Chesler and Kaila, 1992; Hochachka and Mommsen, 1983) levels go up, and the cytosolic NAD+/NADH ratio decreases (Clarke et al., 1989). These changes drive the LDH-catalyzed reaction toward lactate production rather than lactate use. Also, changes in extracellular lactate during activity may fall within a range of only 0% to 135% (see Fellows et al., 1993; Fray et al., 1996; Hu and Wilson, 1997) above its resting levels (0.6 to 1.37 mM) (Harada et al., 1992; Prichard et al., 1991) during activity. Thus, even if lactate increases during neural activity, whether it rises to levels that can create conditions thermodynamically favorable for driving the LDH-catalyzed reaction in neurons toward lactate oxidation remains unclear. Even if conditions were favorable for lactate oxidation in neurons during activity, LDH must still compete with glyceraldehyde-3-phosphate (P) dehydrogenase (GAPDH) for NAD+ (Fig. 3). GAPDH, like LDH, is present at high concentrations in brain cells (McIlwain and Bachelard, 1985). The increased glyceraldehyde-3-P levels that occur with elevated glycolytic activity increase the activity of GAPDH (Williamson, 1965). Because the conversion of glucose to pyruvate, which produces two net ATPs, is more favorable energetically than the conversion of lactate to pyruvate, it is doubtful that lactate use will occur at the expense of glucose use during periods of increased energy demand.
As discussed in Part II, little unequivocal experimental evidence exists for active brain areas using lactate during neural activity. Also, if lactate levels are high enough to inhibit glycolysis, then glycolysis in both neurons and astrocytes would be inhibited, leaving open the possibility that both cell types would use lactate. Indeed, both neurons and astrocytes can use lactate as an energy substrate (Peng et al., 1994; Vicario et al., 1993). In fact, the high sensitivity of LDH-1, the dominant LDH isoform in neurons (Bittar et al., 1996), to pyruvate product inhibition (Stambaugh and Post, 1966; also see Part I, The LDH isoform distribution) may help suppress lactate use in neurons when glycolytic rates increase and pyruvate levels go up. Thus, if lactate is being used during activity, it may be more likely that astrocytes are using it because astrocytes have both LDH-1 and LDH-5 and because LDH-5 has a lower sensitivity to pyruvate product inhibition than LDH-1.
As indicated in Part II, little unambiguous experimental evidence exists for the shuttling of lactate from astrocytes to neurons during neural activity. The observation that lactate levels diminish only after neural activity in most in vivo studies supports the idea that lactate is used in the brain mainly after neural activity instead of during activity, as postulated by the ANLSH (Bouzier-Sore et al., 2002; Pellerin et al., 1998a).

Thermodynamics of glucose oxidation and lactate oxidation

ANLSH supporters also have proposed that lactate oxidation is favored over glucose oxidation in neurons because ATP consumption at the hexokinase (HK) and phosphofructokinase (PFK) steps of glycolysis imposes an energetic disadvantage (Magistretti, 1999; Schurr and Rigor, 1998). However, the reactions catalyzed by HK and PFK are thermodynamically very favorable and are essentially irreversible under most physiological conditions (Lehninger et al., 1993). In contrast, as seen in the preceding section, the thermodynamics of conversion of lactate to pyruvate may not favor the occurrence of this reaction during neural activity. Also, declining ATP levels greatly activate both HK and PFK (Clarke et al., 1989) but not LDH. In addition, conversion of glucose to pyruvate produces two net ATPs, whereas conversion of lactate to pyruvate produces no ATP. These factors favor glucose use during times of increased ATP consumption. Furthermore, because HK and PFK have Km levels for ATP (0.13 mM for HK and 0.025 mM for PFK) (Lowry and Passonneau, 1964) that are well below resting ATP levels (approximately 3 mM) (McIlwain and Bachelard, 1985), fluxes through these two reactions are affected only if ATP drops to extremely low levels, such as under pathological conditions. Thus, from a thermodynamic standpoint, neurons gain an energetic advantage by oxidizing glucose instead of lactate once increased energy demand has activated key glycolytic enzymes. In fact, many cell types usually respond to increased energy demands by activating glycolysis, including cultured neurons (Peng et al., 1994; Walz and Mukerji, 1988) and astrocytes (Pellerin and Magistretti, 1994).

Glycolysis as an essential energy source for localized areas of brain cells

Glycolysis may generate ATP locally in both neurons and astrocytes for ion pumping and other processes (Wu et al., 1997). Glycolytically derived ATP may be particularly important for synapses. This is especially true for dendritic spines, which rarely have mitochondria (Sorra and Harris, 2000; Wu et al., 1997). In the hippocampus, mitochondria are found only in the large, complex spines found in subfield CA3 (Chicurel and Harris, 1992). Sometimes, mitochondria are absent presynaptically. For example, in the CA3 to CA1 projection in the hippocampus, mitochondria occur with low frequency in axonal varicosities and are in fact absent in 50% of synaptic boutons (Shepherd and Harris, 1998). Glycolytic enzymes have been clearly identified in dendritic spines, where they are bound to postsynaptic densities (PSDs) (Wu et al., 1997). PSDs contain not only neurotransmitter receptors and ion channels but also protein kinases (Wu et al., 1997). These protein kinases consume ATP generated in PSDs via glycolysis (Wu et al., 1997). The presence of glycolytic enzymes, lactate dehydrogenase (Wu et al., 1997), and the monocarboxylate transporter (Bergersen et al., 2001) in PSDs, and the rarity of mitochondria in dendritic spines, suggests that at least some glucose may be used anaerobically and that the lactate produced via anaerobic glycolysis in PSDs may be released via the monocarboxylate transporter in active neurons. The rarity of mitochondria in dendritic spines and in synaptic boutons in some regions of the brain may help explain at least partially the increased lactate production during heightened neural activity in such brain areas, even when there is no shortage of oxygen (e.g., Fox et al., 1988). Similarly, the peripheral processes of astrocytes may require glycolysis for energy generation because they are essentially bereft of mitochondria, which may be too large to fit into these processes (Dienel and Cruz, 2003).
Moreover, energy generated from glycolysis has been linked to ion transport via Na+-K+ ATPase. Studies of erythrocytes (Proverbio and Hoffman, 1977), smooth muscle (Paul et al., 1979), and cardiac muscle (Weiss and Hiltbrand, 1985) have provided compelling evidence that glycolytically generated ATP is linked to Na+-K+ ATPase. Although no direct evidence for a similar link has been identified yet in the brain, both glycolytic enzymes and Na+-K+ ATPase are bound at high specific activity within synaptosomal membranes (Knull, 1978; Lim et al., 1983). Also, several studies have shown that ion transport is less robust when glucose is replaced either by lactate or by pyruvate (Lipton and Robacker, 1983; Roberts, 1993; Silver et al., 1997). The association of glycolytic enzymes with Na+-K+ ATPase in synaptic membranes (Knull, 1978; Lim et al., 1983) suggests that glucose, rather than lactate, is used as the substrate to provide energy for the Na+-K+ pump in active neurons.

Maintenance of the cytoplasmic oxidation/reduction balance

A factor that argues against the use of lactate as a substrate by neurons is the cytoplasmic oxidation/reduction (redox) balance. With glucose as the substrate, the redox balance in neurons can be maintained by NADH shuttles and by converting pyruvate to lactate, which regenerates NAD+ (Fig. 4). In contrast, when lactate is the energy substrate, the redox balance is maintained by NADH shuttles alone (Fig. 3). The fact that lactate is produced during neural activity suggests that activation of NADH shuttles in the brain may lag behind increases in energy demand. Also, as mentioned in Part I, Glycolysis as an essential energy source for localized areas of brain cells, mitochondria are extremely rare in dendritic spines (Sorra and Harris, 2000; Wu et al., 1997) and infrequent in the synaptic boutons of some brain areas, such as the hippocampal CA3 to CA1 projection (Shepherd and Harris, 1998). In those synaptic elements containing no mitochondria, reliance on NADH shuttles to maintain the redox balance may not be possible. This may constitute another disadvantage for metabolizing lactate to meet increased cellular energy demands during neural activation.
FIG. 4. Schematic illustration of cellular mechanisms used to maintain the redox balance when glucose is the energy substrate. NADH generated by the conversion of GAD-3P to 1,3-BPG (catalyzed by glyceraldehyde-3-phosphate dehydrogenase) in the glycolytic pathway can be reoxidized to NAD+ via (1) LDH, which converts pyruvate to lactate, (2) the G3P shuttle, which transfers reducing equivalents from cytosolic NADH to FAD inside mitochondria, or (3) the malate shuttle, which transfers reducing equivalents from cytosolic NADH to NAD+ inside mitochondria. In areas where mitochondria are rare, such as dendritic spines (see text), lactate formation may be the dominant mechanism for maintaining the redox balance because diffusion of NADH to where mitochondria are may be slow. During increased glycolytic activity, the activity of the G3P shuttle may be accelerated by increased levels of DHAP, which is also a glycolytic intermediate. GAD-3P, glyceraldehyde-3-phosphate; 1,3-BPG, 1,3-bisphosphoglycerate; G3P, glycerol-3-phosphate; DHAP, dihydroxyacetone phosphate; Asp, aspartate; E.T.C., electron transport chain; FBP, fructose-1,6-bisphosphate; Mal, malate; OAA, oxaloacetate.
Another advantage of glucose oxidation is that accelerated activity in at least one of the NADH shuttles can occur via activation of the glycolytic pathway. NADH generated from glycolysis via GAPDH can be reoxidized by both glycerol-3-P dehydrogenase (glycerol-3-P shuttle) and malate dehydrogenase (malate-aspartate shuttle) (Cheeseman and Clark, 1988; McKenna et al., 1993) (Fig. 4). Each NADH shuttle system depends upon the presence of carrier metabolites (Berry et al., 1992). The glycolytic intermediate dihydroxyacetone-3-P is the carrier for the glycerol-3-P shuttle. Dihydroxyacetone-3-P levels rise as the glycolytic rate increases (Williamson, 1965), so the rate of NAD+ regeneration via the glycerol-3 phosphate shuttle can increase (Berry et al., 1992) to maintain a high glycolytic rate. In contrast, no mechanism is available to increase dihydroxyactone-3-P levels with lactate use; therefore, NAD+ regeneration may be slower when lactate, instead of glucose, is used as the substrate.
Also, LDH may not have as good an access to the glycerol-3-P shuttle as GAPDH does. For example, in hepatocytes, NADH produced by lactate dehydrogenation failed to reach the glycerol-3-P shuttle, possibly because the enzymes for glycolysis are segregated from the enzymes for gluconeogenesis (Berry et al., 1992). Thus, NADH shuttles may limit lactate oxidation, as suggested in an important study by Hertz and colleagues (Peng et al., 1994). In this study, cultured cerebellar granule cells exposed to high K+ concentrations doubled their rates of glucose and pyruvate oxidation but experienced no statistically significant changes in lactate oxidation (Peng et al., 1994). The finding that pyruvate oxidation, but not lactate oxidation, increased with high energy demand points out that lactate oxidation is limited at the step where lactate is converted to pyruvate, possibly because of a limited supply of NAD+. Thus, neurons face difficulties in increasing lactate oxidation during heightened energy demand even when glucose is absent. This may explain why in some studies lactate is less efficient than glucose in supporting ion pumping activity in brain tissue (Lipton and Robacker, 1983; Roberts, 1993; Silver et al., 1997). The findings of Hertz and colleagues serve as strong evidence against lactate competing with glucose as a substrate during elevations in energy demand.

The LDH isoform distribution

The distribution of LDH isoforms in neurons and astrocytes often has been used to support the ANLSH. LDH-1 is the main LDH isoform in neurons, whereas astrocytes have both LDH-1 and LDH-5 (Bittar et al., 1996; Tsacopoulos and Magistretti, 1996). ANLSH supporters suggest that the dominant presence of LDH-1 in neurons means that neurons are sites of lactate use, whereas astrocytes are sites of lactate production (Bittar et al., 1996; Pellerin et al., 1998a), because (1) the higher sensitivity of LDH-1 to pyruvate inhibition drives the reaction catalyzed by LDH-1 toward pyruvate production (Bittar et al., 1996), and (2) the lower Km of LDH-1 for lactate (Bittar et al., 1996) makes LDH-1 better suited than LDH-5 for lactate oxidation. However, the free energy of a biochemical reaction determines its direction, not the enzyme catalyzing it (Lehninger et al., 1993), so the dominant presence of LDH-1 in neurons does not influence the direction of the LDH reaction. Also, the lower Vmax of LDH-1 can offset the effects of its lower Km when the enzyme activity of LDH-1 is compared with that of LDH-5 (Nitisewojo and Hultin, 1976). Of greater importance is that LDH-1 is more sensitive to pyruvate product inhibition (Ki = 0.18 mM for LDH-1 and 0.28 mM for LDH-5) (Stambaugh and Post, 1966). Because pyruvate levels are between 0.1 to 0.2 mM in brain tissue (McIlwain and Bachelard, 1985), the higher sensitivity of LDH-1 to pyruvate product inhibition implies that LDH-1 has greater difficulty oxidizing lactate than LDH-5 in most physiological situations, particularly when pyruvate levels increase. Thus, LDH-1 may help to inhibit lactate oxidation in neurons during neural activity when the glycolytic pathway is activated and pyruvate levels increase.
The prevalence of LDH-1 in neurons also may help neurons oxidize glucose fully. This is because of the higher sensitivity of LDH-1 to pyruvate substrate inhibition and lactate product inhibition, which helps direct glycolytically derived pyruvate to the TCA cycle instead of toward lactate production (Cahn et al., 1962). The fact that less lactate is produced at rest in cultured neurons than in cultured astrocytes (Schousboe et al., 1997; Walz and Mukerji, 1988) supports this conclusion. LDH-1 is also the dominant isoform in other highly aerobic cells, such as cardiac muscle cells, where glucose is often fully oxidized, while LDH-5 is the dominant isoform in cells where anaerobic glycolysis (or glycogenolysis) is a major mode of energy production, such as in skeletal muscle cells. Therefore, the presence of both LDH-1 and LDH-5 in astrocytes means that astrocytes may metabolize glucose either aerobically or anaerobically.
The prevailing glycolytic rate probably influences lactate use more than the LDH isoform. For example, in skeletal muscle in vivo, where LDH-5 is the predominant isoform, lactate produced and released from active muscle, where the rate of glycogenolysis is high, is used by inactive muscle, where the rate of glycogenolysis is low (Bangsbo et al., 1995). Thus, the dominant presence of LDH-1 in neurons does not support the claim of the ANLSH that neurons are the primary users of lactate.

Distribution of monocarboxylate transporters in neurons and astrocytes

The distribution of monocarboxylate transporters in the brain is heterogeneous because neurons have mainly monocarboxylate transporter 2 (MCT2), whereas astrocytes have MCT1, MCT2, and MCT4 (Bergersen et al., 2001; Gerhart et al., 1998; Pellerin et al., 1998b). This heterogeneous distribution has been cited as additional evidence for the ANLSH (Bergersen et al., 2001; Broer et al., 1997; Pellerin et al., 1998b; Pierre et al., 2002). Of the different MCT isoforms, MCT2 has the highest affinity for lactate (Bergersen et al., 2001). Based on this, Bergersen et al. (2001) concluded that MCT2 allows movement of lactate from astrocytes into neurons. However, the gradient for lactate across brain cell membranes determines the direction of lactate transport, not the type of MCT present. For example, when glycolysis is activated, lactate efflux from cultured neurons can equal that of astrocytes (Walz and Mukerji, 1988). Also, the association of MCT, LDH, and glycolytic enzymes with PSDs in dendritic spines (Bergersen et al., 2001; Wu et al., 1997), where mitochondria are rare, suggests that lactate is being produced and transported out of dendritic spines during heightened synaptic activity (also see Part I, Glycolysis as an essential energy source for localized areas of brain cells). In addition, the activity of the MCT is determined not only by its Km, but also by its Vmax. Thus, a higher affinity of MCT2 for lactate does not necessarily imply a higher rate of lactate uptake. Indeed, the initial rates of lactate uptake are similar in cultured astrocytes and neurons for lactate concentrations of 1 to 5 mM (Dienel and Hertz, 2001; Dringen et al., 1993c; Tildon et al., 1993). Therefore, it seems unlikely that differences in MCT isoforms significantly influence lactate transport in either neurons or astrocytes.

SUMMARY

The characteristics of glycolytic enzymes and the kinetics of metabolite changes suggest that glucose is the primary energy substrate in both neurons and astrocytes during neural activation. Factors arguing against active neurons using lactate as an energy substrate include (1) the lack of a mechanism to activate LDH during increased energy demand, (2) the slow increase of brain lactate levels during neural activity, (3) the rapid activation of the glycolytic pathway by increased energy demand, and (4) the likelihood that activation of glycolysis may suppress lactate oxidation. The observed decline of lactate after physiological brain stimulation suggests that use or removal of lactate occurs mainly after neural activity.

PART II: EXPERIMENTAL EVIDENCE REGARDING GLUCOSE AND LACTATE USE

Sites of glucose uptake

1:1 coupling of oxidative glucose consumption to glutamate cycling

A study by Shulman and coworkers (Sibson et al., 1998) regarding the coupling between oxidative glucose use and glutamate cycling is thought to provide key support for the ANLSH (Magistretti and Pellerin, 1999; Magistretti et al., 1999). These investigators used 13C-NMR to follow the fate of 13C-glucose injected into the rat brain in vivo when different levels of neural activity were simulated with various levels of anesthesia. They concluded that a 1:1 stoichiometry existed between increases in oxidative glucose consumption and increases in glutamate cycling in astrocytes (measured as increases in glutamine synthesis) as neural activity increased. ANLSH supporters have suggested that this 1:1 ratio means that glutamate cycling in astrocytes is the principal driving force behind brain glucose metabolism.
The support this study gives the ANLSH is debatable for two major reasons. First, the calculated rate of glutamate cycling may not be applicable in situ. A critical assumption of the Sibson et al. (1998) study was that the rate of glutamine synthesis via the anaplerotic pathway is unchanged when the rate of total glutamine synthesis goes up with increased neural activity. This assumption has no obvious foundation and may seriously underestimate the stoichiometry between oxidative glucose consumption and glutamate cycling. For example, if the rate of glutamine synthesis via the anaplerotic pathway during normal activity had been 30% of total glutamine synthesis, as others have observed (Kunnecke et al., 1993; Lapidot and Haber, 2000), then the stoichiometry between oxidative glucose consumption and glutamate cycling would be roughly 1.5:1 rather than 1:1 (estimated from Sibson et al., 1998). Thus, the stoichiometry between oxidative glucose metabolism and glutamate cycling may vary greatly depending on the assumptions used. Additional assumptions made in the Sibson et al. study also may bring into question how applicable the results of this study are in situ (also see Gruetter, in press; Roberts and Chih, in press).
Second, even if the data did support a 1:1 stoichiometry between oxidative glucose metabolism and glutamate cycling, other models besides the ANLSH, including the conventional hypothesis, may explain the data. Because the sites of glucose uptake and modes of glucose consumption in astrocytes and neurons were not determined in the Sibson et al. study, the calculated amount of glucose consumed by neurons and astrocytes could vary greatly depending on the percentage of glucose metabolized aerobically in neurons and astrocytes. If activity-induced increases in astrocyte glucose metabolism were partially aerobic instead of entirely anaerobic as assumed by the ANLSH, glutamate cycling would require much less glucose consumption in astrocytes. This is due to the large difference in efficiency between aerobic and anaerobic glucose metabolism (34 to 36 ATP versus 2 ATP). For example, if as little as 6% of glucose metabolism in astrocytes had been aerobic, the total glucose needed for the measured glutamate cycling in astrocytes would be only 50% of measured glucose consumption. This means neurons could have consumed the other 50% of glucose aerobically. Likewise, if 20% of astrocytic glucose metabolism had been aerobic, then neurons would have accounted for 75% of total glucose consumption in the Sibson et al. study. Thus, for the results of Sibson et al. to suggest that glucose is predominantly consumed in astrocytes, the assumption of the ANLSH that glutamate cycling in astrocytes is supported entirely by anaerobic glucose metabolism needs verification.
However, this assumption is controversial based on existing evidence. Although uptake of exogenous glutamate is associated with lactate production in cultured astrocytes (Pellerin and Magistretti, 1994) and in the brain in vivo (Demestre et al., 1997), the percent of glutamate uptake in astrocytes fueled by anaerobic glycolysis is unknown. Many studies have shown that glutamate uptake in astrocytes can be sustained at least partially by oxidative metabolism (e.g., Hertz et al., 1998; Swanson, 1992). In fact, Hertz et al. (1998) concluded that energy for glutamate uptake in cultured astroglia was largely provided by oxidative metabolism of accumulated glutamate. Also, astrocyte cultures increase their oxygen consumption when exposed to 100 μM glutamate (Eriksson et al., 1995), which suggests that glutamate cycling in astrocytes is at least partially supported by oxidative metabolism. In addition, studies of mammalian astrocytes have shown no preferential use of energy produced by either glycolysis or oxidative phosphorylation for the support of Na+-K+ pump activity (Silver and Erecinska, 1997). Moreover, Rothman and coworkers have recently determined that the astroglial TCA cycle accounts for 14% of brain oxidative metabolism (Lebon et al., 2002). Other researchers have estimated a higher percentage of glucose metabolism via the glial TCA cycle based on the relative activity of pyruvate carboxylase and pyruvate dehydrogenase (Hassel et al., 1995; Qu et al., 2000). Thus, glutamate probably activates both anaerobic glycolysis and oxidative metabolism in astrocytes. Because aerobic metabolism is more efficient than anaerobic metabolism, the glucose consumed by astrocytes should be much less than the observed total glucose consumption in the Sibson et al. study. Therefore, this calculated 1:1 stoichiometry should not be used to draw conclusions regarding the relative contributions of neurons and astrocytes to glucose consumption in the brain during neural activity.

Uptake of 2-deoxyglucose

Activity-induced 2-DG uptake has been found to occur primarily in the neuropil, a region enriched in axon terminals, dendrites, and synapses (see Kadekaro et al., 1985). Because the membranes of synaptic terminals contain a high density of the glucose transporter GLUT3 (Leino et al., 1997; McCall et al., 1994), the pattern of 2-DG uptake suggests that active neurons do take up glucose. A recent in vivo study (Wittendorp-Rechenmann et al., 2001), in which a microautoradiographic imaging procedure was used also showed that 54% to 62% of 14C-2-DG was picked up by microtubule-associated protein 2 (MAP2)-positive cells (neurons), compared with 42% to 58% by glial fibrillary acidic protein (GFAP)-positive cells (astrocytes). Thus, these findings call into question the idea of the ANLSH that activity-induced glucose uptake occurs predominantly in astrocytes.

Consumption of glucose in cultured neurons and astrocytes

Many studies have shown that both neurons and astrocytes use glucose as a substrate (Peng et al., 1994; Walz and Mukerji, 1988). The fact that more glucose is consumed by cultured astrocytes than by cultured neurons (Lopes-Cardozo et al., 1986; Waagepetersen et al., 2000) has been cited as evidence that glucose is preferentially consumed in astrocytes (e.g., Magistretti, 1999; Magistretti et al., 1999). However, neurons tend to oxidize glucose fully instead of metabolizing glucose to lactate. This is seen in a higher rate of glucose oxidation (Itoh et al., 2003; Vicario et al., 1993) and in a smaller production of lactate by cultured neurons compared with cultured astrocytes (Waagepetersen et al., 2000; Walz and Mukerji, 1988). The dominant presence of LDH-1 in neurons may help direct pyruvate generated from the glycolytic pathway toward the TCA cycle instead of toward lactate production (also see Part I, The LDH isoform distribution). Thus, cultured neurons may need less glucose than cultured astrocytes because they may rely less than astrocytes do on anaerobic glycolysis for energy production. In vivo, the actual consumption of glucose by neurons and astrocytes would depend on the relative masses and energy requirements of these brain cells.

Sites of lactate release

Lactate release from cultured neurons and astrocytes

As noted in the previous section, under conditions considered normoxic for cell cultures, both cultured neurons and astrocytes generate lactate from glucose, but neurons typically produce less lactate then astrocytes (Itoh et al., 2003; Walz and Mukerji, 1988). Consistent with this finding is the observation that neurons typically have a higher rate of glucose oxidation than astrocytes (Itoh et al., 2003; Vicario et al., 1993). These observations suggest that glucose is more fully metabolized in neurons than in astrocytes (also see Part II, Sites of glucose uptake, Consumption of glucose in cultured neurons and astrocytes). However, observations of lower lactate production in cultured neurons at rest do not mean neurons have a smaller glycolytic capacity than astrocytes. This is because, when oxidative metabolism is blocked, neurons can increase lactate production as much as astrocytes (Walz and Mukerji, 1988).

Lactate release during neural activity in vivo

Two microdialysis studies (Demestre et al., 1997; Fray et al., 1996) have been cited by ANLSH proponents as providing evidence for lactate release specifically from astrocytes during stimulated neural activity (Bouzier-Sore et al., 2002). These studies showed that the competitive glutamate uptake blockers threo-hydroxyaspartate (THA) and L-trans-pyrrolidine-2,4-dicarboxylate (tPDC) blocked stimulation-induced increases in lactate (Demestre et al., 1997; Fray et al., 1996). From these results, it was concluded that stimulated release of lactate is dependent on the uptake of glutamate (Demestre et al., 1997). However, the uptake blockers themselves elevated lactate before sensory stimulation as much as sensory stimulation did in the absence of uptake inhibitors. Thus, the lack of an effect of sensory stimulation on lactate release may have been because the glycolytic rate was maximal or near maximal before stimulation. Also, the two glutamate uptake blockers used are not specific for glial glutamate transporters and would inhibit glutamate uptake into either glia or neurons (e.g., Danbolt, 2001), given that glutamate is taken up by pre- and postsynaptic neurons as well as by astrocytes (Conti and Weinberg, 1999; Danbolt, 2001). Thus, the results from the microdialysis studies do not establish that lactate was produced specifically by glia, as has been suggested by some investigators (e.g., Bouzier-Sore et al., 2002). As in cultured neurons, neurons in vivo may produce less lactate during neural activity because they may oxidize glucose more fully (see above).

Sites of lactate use

Comparison of lactate use in neuronal and astrocyte cultures

The ability of both cultured neurons and astrocytes to use lactate has been well established (Itoh et al., 2003; McKenna et al., 1993; Peng et al., 1994; Vicario et al., 1993). Some studies have shown that neurons and astrocytes have similar rates of lactate oxidation (Peng et al., 1994; Vicario et al., 1993), whereas others have reported that rates of lactate oxidation are higher in neurons or isolated synaptic terminals than in astrocytes (Itoh et al., 2003; McKenna et al., 1993). Neurons and astrocytes may use lactate via other metabolic pathways besides oxidation. Tabernero et al. (1993) reported that the rate of lipogenesis from lactate is approximately twofold greater in astrocytes than in neurons. Dringen et al. (1993b) showed that cultured astrocytes can incorporate lactate into glycogen. Thus, both astrocytes and neurons can use lactate.
The higher rates of lactate oxidation in neurons compared with astrocytes observed in some studies (Itoh et al., 2003; McKenna et al., 1993) are probably due to generally higher rates of oxidative metabolism in neurons because the rates of glucose oxidation observed in neurons are also higher than those in astrocytes (Itoh et al., 2003; Vicario et al., 1993). It is interesting to note that both cultured neurons and astrocytes oxidize lactate much faster than they do glucose. In astrocytes, rates of lactate oxidation are approximately 4 to11 times greater than rates of glucose oxidation, whereas in neurons, rates of lactate oxidation are approximately 3 to 10 times greater than rates of glucose oxidation (Itoh et al., 2003; McKenna et al., 1993; Vicario et al., 1993). This may be caused in part by the fact that some glucose is metabolized anaerobically to lactate to provide energy in cultured brain cells. Also, as mentioned previously (see Part I, Responses of glycolytic enzymes and LDH to increases in energy demand during neural activity), the glycolytic pathway is controlled mainly by energy demand (Sokoloff, 1989), not by glucose concentration, and so is greatly suppressed at rest. In contrast, LDH catalyzes a reaction near equilibrium at rest and can proceed rapidly when the ratio of its substrates and products levels are favorable. Other substrates, such as glutamine and 3-hydroxybutyrate, were also oxidized at a much higher rate than glucose in isolated synaptic terminals and astrocytes (McKenna et al., 1993).

Sites of lactate use in vivo

Two 13C-NMR spectroscopy studies of metabolic substrate turnover in the brain in vivo (Bouzier et al., 2000; Hassel and Bråthe, 2000) have been cited as providing convincing evidence that lactate is specifically a neuronal substrate in vivo (Bouzier-Sore et al., 2002). In these studies, a similar labeling of the glutamine C2 and C3 positions was found after infusion of rats with 3−13C-lactate. Bouzier et al. (2000) suggested that the similar labeling of the glutamine C2 and C3 positions meant that 3−13C-lactate was metabolized in a compartment devoid of pyruvate carboxylase (PC), specifically neurons. However, whereas a greater labeling of glutamine C2 than glutamine C3 may indicate higher glial PC activity (Bouzier et al., 2000; Merle et al., 2002), it does not necessarily follow that similar labeling of glutamine C2 and C3 means that astrocytes were not metabolizing 3−13C-lactate. For example, 3−13C-pyruvate derived from 3−13C-lactate could have been metabolized in astrocytes via the pyruvate dehydrogenase (PDH)-mediated pathway. Zwingman et al. (2000) found little difference in the labeling of the glutamine C2 and C3 positions in cultured astrocytes after they were incubated in 3−13C-alanine, which, like 3−13C-lactate, leads to the production of 3−13C-pyruvate. Martin et al. (1993) reported that labeling of the glutamine C2 position was 1.4 times greater than that of the glutamine C3 position in astrocyte cultures exposed to 1-13C-glucose, which is also converted metabolically to 3−13C-pyruvate. The inconsistency of the glutamine C2/C3 labeling ratio in astrocytes between different studies indicates that the labeling pattern from 3−13C-pyruvate may not reflect the site(s) of substrate use.
Moreover, the glutamine C2/C3 ratio can be affected by other factors such as time (Hassel et al., 1995; Merle et al., 2002) and anesthesia (Shank et al., 1993). Several investigators have reported on the effects of time (e.g., Hassel et al., 1995; Merle et al., 2002; Shank et al., 1993). For example, Hassel et al. (1995) found that the glutamine C2/C3 ratio in the mouse brain decreased from 1.96 at 15 minutes to 1.12 at 30 minutes after they injected 1-13C-glucose intravenously into conscious mice. Similarly, Shank et al. (1993) reported a decline in the glutamine C2/C3 ratio of the rat forebrain from around 1.8 at 30 minutes to around 1.2 at approximately 45 minutes (estimated from Shank et al., 1993) after they administered 1-13C-glucose intraperitoneally to conscious rats. The glutamine C2/C3 ratios obtained at the later time points in these studies were similar to the ratios found by Hassel and Bråthe (2000) (1.3 at 15 minutes) and by Bouzier et al. (2000) (1.1 at 20 minutes), all of whom used 3−13C-lactate as the labeling substrate.
The time dependence of glutamine C2/C3 labeling may be caused in part by the relative contributions of neuronal and astrocytic pools of glutamate to glutamine synthesis changing with time (Merle et al., 2002). It could also be caused by the equilibration of oxaloacetate with fumarate, which scrambles the label between the glutamine C2 and C3 positions (Hassel et al., 1995). It has been suggested that the glutamine C2/C3 ratio may be most accurate in the early phase after the infusion of labeled substrates (Merle et al., 2002; Shank et al., 1993). However, accurate determination of the amino acid enrichment in the initial phase is hindered by many technical difficulties, including low signal-to-noise ratios in 13C-NMR spectra (Merle et al., 2002). These factors further increase the uncertainty of predicting the sites of substrate use based on the labeling of the glutamine C2 and C3 positions.
In addition, results from a study by Sonnewald and coworkers (Qu et al., 2000) conflict with the studies of Hassel and Bråthe (2000) and Bouzier et al. (2000). In the Qu et al. study (2000), the relative contributions of PC and PDH to glutamine labeling in the rat brain were approximately 0.31 after rats were injected with [U-13C]lactate, and approximately 0.41 after rats were injected with [U-13C]glucose. Their results suggested that lactate, like glucose, is used by astrocytes in vivo.

Evidence for an astrocyte-neuron lactate shuttle during normal neural activity

Substrate use in an isolated retinal preparation

Major support for the existence of an astrocyte-neuron lactate shuttle has been drawn from a study by Poitry-Yamate et al. (1995). They concluded that photoreceptors used lactate released from retinal astrocytes (Müller cells) even in the presence of glucose. The principal finding was that the radioactivity of bath 14C-lactate released from Müller cells was reduced 70% when photoreceptors were included in the cell suspension. However, as discussed in detail elsewhere (Chih et al., 2001a, 2001b; Roberts and Chih, in press), most of the decrease in 14C-lactate radioactivity was due to dilution caused by adding photoreceptors to Müller cell cultures. Thus, it is difficult to conclude from the Poitry-Yamate et al. study whether a shuttle exists between Müller cells and photoreceptors.

Exchange of lactate between astrocytes and neurons in coculture systems

Two other studies of astrocyte-neuron cocultures (Waagepetersen et al., 2000; Zwingmann et al., 2000) have been used as direct evidence for the existence of an astrocyte-neuron lactate shuttle (Bouzier-Sore et al., 2002). In one study, monocultures of neurons and astrocytes, and astrocyte-neuron cocultures, were incubated in 1 mM [U-13C]lactate and 2.5 mM glucose (Waagepetersen et al., 2000). After 4 hours of incubation, unlabeled lactate formed in all cultures. The consumption of [U-13C]lactate in each group was calculated based on the difference between bath [U-13C]lactate at the beginning and end of incubation. Neurons in monoculture apparently consumed 0.7 μmol/mg protein of [U-13C]lactate in 4 h, whereas consumption of [U-13C]lactate in both astrocyte cultures and astrocyte-neuron cocultures was listed as “nonquantifiable”. Waagepetersen et al. postulated that because less [U-13C]lactate was consumed by neurons in cocultures than by the neurons in monocultures, the neurons in the cocultures must have used the unlabeled lactate produced by astrocytes. However, when estimated from the data in Table 1 of Waagepetersen et al. (2000), the calculated averaged value of bath [U-13C]lactate in both astrocyte cultures and astrocyte-neuron cocultures apparently increased from the beginning to the end of the incubation period. This calculated increase in [U-13C]lactate would reflect experimental error and may not be equated to zero (nonquantifiable) [U-13C]lactate consumption. For statistical comparisons of the presumed consumption of [U-13C]lactate among the three groups, the mean values and variations of [U-13C]lactate concentration in each culture must be evaluated. Also, the proportional content of neurons in the cocultures was not determined. Without the preceding information, comparisons of [U-13C]lactate consumption between neurons in monoculture and coculture are not possible. Moreover, even if bath [U-13C]lactate decreased less in cocultures than it did in neuronal monocultures, the result does not necessarily mean that neurons used the unlabeled lactate released from astrocytes instead of [U-13C]lactate. Other plausible explanations for the results exist, such as increased glucose consumption by neurons in the coculture. Thus, the data from the Waagepetersen et al. study provide little evidence that a lactate shuttle exists between astrocytes and neurons.
Also in the Waagepetersen et al. study, the formation of unlabeled lactate exceeds the calculated decrease in bath [U-13C]lactate in neuronal monocultures by a factor of two. A net formation of lactate suggests that the thermodynamics of the LDH reaction in neurons favors lactate instead of pyruvate production for the given experimental conditions. Thus, it is unclear whether the calculated decrease in bath [U-13C]lactate in the media really represents neuronal use of [U-13C]lactate. Even if it does, the calculated [U-13C]lactate consumption in neuronal monocultures only represents approximately 38% of energy production from glucose use, taking into account the facts that glucose produces two pyruvates and lactate produces only one pyruvate and that approximately 47% of glucose was metabolized anaerobically in this study, which does not support the idea of the ANLSH that neurons prefer lactate as a substrate over glucose.
In another study, monocultures of neurons and astrocytes, and astrocyte-neuron cocultures, were incubated in 1.5 mM [3−13C]alanine and 2.5 mM glucose (Zwingmann et al., 2000). In this study, astrocytes released more lactate than neurons, and lactate produced from alanine was higher in astrocyte-neuron cocultures than in astrocyte cultures. The authors of this study speculated that a lactate-alanine shuttle between astrocytes and neurons existed, but no evidence was presented showing that neurons consumed lactate released by astrocytes. Thus, the existence of an astrocyte-neuron lactate shuttle cannot be substantiated based on the results of either the Zwingmann et al. study or the Waagepetersen et al. study.

Breakdown of glycogen in astrocytes as a source of lactate for neighboring cells

Two studies (Dringen et al., 1993a; Wender et al., 2000) are thought by some investigators (e.g., Bouzier-Sore et al., 2002) to provide evidence that lactate produced by astrocytes as a result of glycogen breakdown is transferred to neurons. In the first study, cultured astrocytes deprived of glucose depleted their glycogen stores and released lactate instead of glucose (Dringen et al., 1993a). However, in the absence of glucose, astrocytes may have to metabolize glucose produced through glycogen breakdown to meet their own energy needs. Also, glucose derived from glycogen may be metabolized anaerobically to lactate in cultured astrocytes because cultured astrocytes typically produce large amounts of lactate even under normoxic conditions (Schousboe et al., 1997; Waagepetersen et al., 2000; Walz and Mukerji, 1988). The breakdown of glycogen to lactate in the absence of glucose does not imply that a similar breakdown of glycogen to lactate will occur in situ when glucose is present. In fact, Fillenz and colleagues (Forsyth et al., 1996; Fray et al., 1996) found evidence that astrocytes may export glucose derived from glycogen degradation to the extracellular space of the brain in vivo when the brain is stimulated. Whether glucose is released by astrocytes and used by active neurons during normal neural activity in vivo is as yet a question needing further exploration.
In a second study, Wender et al. (2000) showed that the lactate transport blockers α-cyano-4-hydroxycinnamate (4-CIN) and p-chloromercurobenzosulfonate (p CMBS) reduced the time needed for failure of axonal function during glucose deprivation and decreased the recovery of axonal function after glucose replenishment. The study implied that when glucose is unavailable, such as during hypoglycemia, active neurons may use lactate derived from glycogen degradation in astrocytes as a fuel. However, the results can only be applied to pathologic conditions when glucose is scarce. As mentioned earlier, lactate must compete with, instead of supplement, glucose as a substrate for active neurons under normal physiological conditions (see Part I, Suppression of lactate oxidation by increased glycolysis). The results of Wender et al. do not imply that, when glucose is readily available, lactate released by astrocytes will be shuttled to and used by active neurons during normal neural activity, as proposed by the ANLSH. Also, the results of this study do not conflict with the conventional view of glucose metabolism, which contends that glucose is the predominant substrate during normal neural activity in the mature brain but that alternative substrates such as lactate may be used when glucose is unavailable.

Distributions of carbonic anhydrase and the sodium-bicarbonate transporter

The distributions of carbonic anhydrase (CA) and the Na+-HCO3 transporter (NBC) also have been considered in some reviews (e.g., Ames, 2000; Bouzier-Sore et al., 2002) as reinforcing lactate shuttling between astrocytes and neurons. In these reviews, CA is said to be present only in astrocytes, and the NBC is said to be found only in astrocytes. However, CA and the NBC occur in both neurons and astrocytes. Indeed, CA has been found in neurons and nerve fibers, particularly those associated with sensory functions in different regions of the brain (Ridderståle and Wistrand, 1998; Willoughby and Schwiening, 2002). The NBC also has been found in rat brain neurons (Schmitt et al., 2000). Thus, speculation that the distribution of CA and of the NBC plays a role in promoting lactate shuttling in the brain is not supported by the existing evidence.

Comparison of glucose and lactate as substrates in the brain and brain tissue

Changes in extracellular glucose and lactate levels in vivo

Several studies have investigated changes in extracellular glucose or lactate levels during and after neural activation in the brain in vivo (Fellows and Boutelle, 1993; Fellows et al., 1993; Fillenz and Lowry, 1998; Fray et al., 1996; Hu and Wilson, 1997). In one study, glucose and lactate use was inferred from changes in extracellular lactate (Hu and Wilson, 1997). However, changes in extracellular glucose or lactate levels are complex because they represent not simply metabolite use alone but also a balance between supply and use. For example, increases in extracellular glucose resulting from neural activation may reflect increases in the glucose supplied from the bloodstream. This is because (1) the plasma glucose concentration (approximately 7 to 8 mM) (Hu and Wilson, 1997; Silver and Erecinska, 1994) is higher than glucose levels in brain tissue, and (2) cerebral blood flow goes up sharply (50% to 100%) in stimulated regions (Fellows and Boutelle, 1993). As a result, extracellular glucose during induced neural activity has been shown to increase in some studies (Fellows and Boutelle, 1993) and decrease in other studies (Fillenz and Lowry, 1998; Fray et al., 1996; Hu and Wilson, 1997). Thus, drawing conclusions regarding glucose or lactate use based on changes in extracellular glucose or lactate levels is difficult.
In the study by Hu and Wilson (1997) mentioned above, transient changes in extracellular lactate and glucose levels measured with lactate- and glucose-sensitive microsensors were used as evidence for glucose and lactate use. As has been discussed in detail elsewhere (Chih et al., 2001a, 2001b; Roberts and Chih, in press), the results of this study were inconclusive because of several reasons, including a lack of statistical analysis.

Changes in brain lactate levels in vivo, as measured with magnetic resonance spectroscopy

In vivo magnetic resonance spectroscopy has been used in several studies to measure changes in lactate levels during stimulated neural activity (e.g., Chhina et al., 2001; Prichard et al., 1991; Ueki et al., 1988). Whereas one study found that intense visual stimulation did not produce significant changes in lactate levels (Chhina et al., 2001), other studies have shown that brain lactate levels were elevated after physiological stimulation (Prichard et al., 1991; Ueki et al., 1988). In a study by Prichard et al. (1991), lactate transiently increased 0.3 to 0.9 mM during the first 3 minutes of stimulation, followed by a gradual decline (approximately 20% to 60% in 10 minutes) during the stimulation period and a continuous decline after termination of stimulation. It is unclear whether this slow decline in lactate levels represented lactate use in brain cells. If it did represent this, then the rate of lactate use was probably declining as brain lactate levels went down because the rate of lactate dehydrogenation depends primarily on lactate levels (see Part I, Responses of glycolytic enzymes and LDH to increases in energy demand during neural activity). Also, the type (neurons or astrocytes) and status (active or inactive) of brain cells contributing to this decline remain unidentified.

Inhibition of glucose consumption by lactate in brain cells in vitro

One study frequently used to support the ANLSH shows that, in unstimulated sympathetic ganglia isolated from chicken embryos, lactate carbons can displace 50% to 70% of the glucose carbons used for CO2 production (Larrabee, 1995). It should be noted that the concentration of lactate used in this study (5 mM) was higher than the extracellular lactate concentration estimated in situ (0.6 to 1.37 mM) (see Part I, Suppression of lactate oxidation by increased glycolysis) and that the cell types using lactate were not identified. The results from this study have been taken as evidence that brain tissue prefers lactate to glucose as an energy substrate (Magistretti et al., 1999). However, as mentioned earlier, the glycolytic rate is tightly coupled to energy demand and so is greatly depressed at rest (section Part I, Responses of glycolytic enzymes and LDH to increases in energy demand during neural activity). In contrast, the rate of lactate dehydrogenation is mainly determined by the concentrations of its substrates and products (Part I, Responses of glycolytic enzymes and LDH to increases in energy demand during neural activity). Thus, when extracellular lactate levels rise through experimental manipulation, lactate oxidation will increase and may suppress glucose use in unstimulated neural tissue. For example, increasing lactate levels inhibit glucose oxidation in both neuronal and astrocyte cultures (Itoh et al., 2003) and reduce 2-DG uptake in vivo (Bliss and Sapolsky, 2001). High lactate levels (10 mM) also suppress lipogenesis from glucose in cultured neurons and astrocytes (Tabernero et al., 1996). The removal of lactate in the brain after neural activity (Fellows et al., 1993; Fray et al., 1996; Hu and Wilson, 1997) may be a protective mechanism allowing the brain to avoid the potentially detrimental effects of a lactate buildup on both neuronal and glial glycolysis.
Inhibition of glycolysis by high lactate levels also occurs during resting conditions in other tissues, such as red skeletal muscle fibers. When unstimulated rat soleus muscles were incubated with 8 mM lactate, lactate metabolism accounted for 70% of the oxygen consumption, and both glucose use and glycerol use were suppressed (Pearce and Connett, 1980). This result does not mean that lactate is the preferred substrate for red muscle fibers during exercise. Likewise, the results from unstimulated nerve tissue in Larabee's study do not necessarily apply to active nerve tissue in vivo.

Comparison of the abilities of glucose and lactate to support neural function in brain slices

The ability of glucose and lactate to support neural activity has been studied in several brain slice preparations. Some studies have shown that lactate supports basal neural activity in vitro as well as glucose (Izumi et al., 1997; Schurr et al., 1988). However, other studies have reported that lactate cannot support neural activity as well as glucose (Cox and Bachelard, 1988; Kanatani et al., 1995; Takata and Okada, 1995; Wada et al., 1997). In two of these studies, population spikes maintained in brain slices exposed to glucose-containing artificial cerebrospinal fluid (ACSF) disappeared in slices provided only with lactate (Kanatani et al., 1995; Wada et al., 1997). In addition, lactate is less able than glucose to aid the recovery of ion homeostasis in brain tissue that is nonpathologically activated (Roberts, 1993): When CA1 pyramidal cells of rat hippocampal slices are activated transsynaptically at high frequency (40 Hz), extracellular K+ recovers more slowly in ACSF containing lactate than in ACSF containing glucose (Roberts, 1993). Thus, although lactate may be used to support activity in brain tissue in the absence of glucose, the issue remains whether lactate can successfully compete with glucose when energy demand increases.
A series of studies in hippocampal slices by Schurr and coworkers (1997, 1999) has been considered as strong support for the ANLSH (Magistretti et al., 1999) because these studies showed that 4-CIN, which inhibits the cell membrane monocarboxylate transporter, prevented the recovery of synaptic transmission even in the presence of adequate glucose. However, 4-CIN blocks not only lactate transport across the plasmalemma but also pyruvate entry into isolated mitochondria (Cox et al., 1985; Halestrap and Denton, 1974). Several studies have shown that 4-CIN by itself compromises normal oxidative metabolism and blocks neural activity (e.g., Amos and Richards, 1996; Chih et al., 2001a; McKenna et al., 2001). The effect of 4-CIN on oxidative metabolism, rather than inhibition of the lactate (monocarboxylate) transporter, could account for the lack of recovery seen in the studies of Schurr and coworkers.

Summary

The above analysis shows that current experimental evidence provides little support for the ANLSH because both neurons and astrocytes may take up glucose, release lactate, and use lactate. Also, the existence of an astrocyte-neuron lactate shuttle has yet to be demonstrated under normal physiological conditions. Although lactate may support basic brain function in the absence of glucose, lactate may be unable to compete successfully with glucose during neural activity.

PART III: ROLE OF LACTATE AS AN ENERGY SUBSTRATE

Brain tissue produces small amounts of lactate even when it is at rest and well oxygenated (Clarke et al., 1989). The reason for increased lactate production during neuronal activity, as deduced from measurements of oxygen and glucose consumption (Fox and Raichle, 1986; Fox et al., 1988; Raichle, 1991; Sappeymarinier et al., 1992) and as demonstrated by increases in extracellular lactate described previously (Fray et al., 1996; Hu and Wilson, 1997; Prichard et al., 1991) is not entirely clear. Lactate production may occur in those areas of neurons and astrocytes where mitochondria are absent and glycolytic enzymes are present, such as cerebrocortical dendritic spine heads (Wu et al., 1997) or distal astrocytic processes (Dienel and Cruz, 2003) (also see Part I, Glycolysis as an essential energy source for localized areas of brain cells). Lactate production may be seen when glycolytic activity transiently exceeds ongoing oxidative metabolism, when NADH shuttles are not activated as quickly as needed to meet increased energy demands (Clarke et al., 1989), or when cellular supplies of oxygen are temporarily limited. An analogous situation is also found in red muscle fibers, where lactate production occurs during moderate exercise although these muscle fibers are fully oxygenated (Connett et al., 1985). Whatever the basis is for lactate production, increases in lactate in the brain during neural activity are moderate and slow.
Buildups of high levels of lactate must be avoided because of its potential inhibitory effects on both neuronal and glial glycolysis. Because the energy demands of brain cells fluctuate rapidly as cells switch between their active and inactive states, the brain needs to remove excess lactate at rest so that the glycolytic pathway in both neurons and astrocytes can be easily activated when brain cells become active (also see Part I, Responses of glycolytic enzymes and LDH to increases in energy demand during neural activity). Several studies have demonstrated that high lactate levels (5 to 10 mM) may significantly suppress glucose use in both neurons and astrocytes at rest (Itoh et al., 2003; Larrabee, 1995; Tabernero et al., 1996). Similar inhibition of glycolysis by high lactate levels (8 mM) occurs in red muscle fibers (Pearce and Connett, 1980) (also see Part II, Comparison of glucose and lactate as substrates in the brain and brain tissue, Inhibition of glucose consumption by lactate in brain cells in vitro). Rapid removal of lactate has already been demonstrated in skeletal muscle, where rapid fluctuations in energy demand and glycolytic rates often takes place. For example, lactate produced by active skeletal muscle fibers is shuttled to and used by inactive skeletal muscle fibers during exercise (Bangsbo et al., 1995).
Because of the relatively slow transport of lactate across the blood-brain barrier (McIlwain and Bachelard, 1985), some lactate produced during and after neural activity probably is used by inactive brain cells. Several possible pathways exist for the metabolism of lactate built up during neural activity. Neurons and astrocytes may use lactate for oxidative metabolism following activity, when energy demands are reduced and glycolytic rates are downregulated. Astrocytes may also use lactate for glycogen synthesis (Dringen et al., 1993b). Furthermore, lactate may be used by those cells with little glycolytic capacity, such as Purkinje neurons, microglia, and oligodendrocytes (Dringen et al., 1993a). In addition, removal of lactate from the brain via the bloodstream may occur when lactate levels are elevated, as during acute ammonia challenge (Hawkins et al., 1973), spreading cortical depression (Dienel and Cruz, 2003), and generalized sensory stimulation (Dienel and Cruz, 2003). The rates of lactate removal via the bloodstream for these stimuli are 15%, 22%, and 6%, respectively, of glucose influx (Dienel and Cruz, 2003; Dienel and Hertz, 2001). The observed decline in lactate after neural activity elicited by physiologically relevant stimuli in most in vivo studies (e.g., Fellows et al., 1993; Fillenz and Lowry, 1998; Fray et al., 1996; Hu and Wilson, 1997) supports the notion that use or removal of lactate occurs mainly after neural activity.

CONCLUSION

The major difference between the conventional hypothesis and the ANLSH concerns which substrate (glucose or lactate) is used by neurons during neuronal activation. The analysis in Part I has provided evidence that, from a theoretical perspective, glucose is probably the major substrate for active neurons. The analysis in Part II suggests that current experimental evidence provides little unequivocal support for the ANLSH. Most importantly, no convincing evidence has yet been offered to affirm the declaration of the ANLSH that shuttling of lactate from astrocytes to active neurons occurs during normal (physiological) neural activity.
The analyses in this review do not rule out the possibility that some brain cells use lactate to meet their energetic needs. Brain cells are heterogeneous and show great structural and biochemical variability. This includes variability in the cellular levels of glycolytic enzymes and in the oxidative capacity of brain cells. Brain cells with low levels of glycolytic enzymes, such as oligodendrocytes, may preferentially metabolize lactate or other substrates. Lactate may also be used by active brain cells when glucose is temporarily unavailable. The conventional hypothesis contends that glucose is the primary substrate but does not exclude the possibility that lactate can be used as an alternative substrate under certain conditions, such as hypoglycemia. However, the analyses in this review point out (1) that the kinetic characteristics of metabolic enzymes and the observed changes in metabolites lend little support to the basic concept of the ANLSH that lactate is produced by astrocytes and used by neurons in an activity-dependent manner, (2) the often overlooked uncertainty in the evidence used to support the ANLSH, and (3) that the existing evidence and theoretical considerations provide a great deal of credibility to the conventional hypothesis.

Acknowledgments

The authors thank Dr. Peter Lipton, Department of Physiology, University of Wisconsin School of Medicine, for his invaluable advice and suggestions that aided greatly with the preparation of this manuscript. The authors also thank Dr. Gerald Dienel, Department of Neurology and Physiology & Biophysics, University of Arkansas for Medical Sciences, for the illuminating discussions regarding the manuscript.

Footnote

1 In this paper, Glycolysis refers to the conversion of glucose to pyruvate under aerobic conditions, and anaerobic glycolysis refers to the conversion of glucose to lactate.

References

Ames A III (2000) CNS energy metabolism as related to function. Brain Res Rev 34:42–68
Amos BJ, Richards CD (1996) Intrinsic hydrogen ion buffering in rat CNS neurones maintained in culture. Exp Physiol 81:261–271
Bangsbo J, Aagaard T, Olsen M, Kiens B (1995) Lactate and H+ uptake in inactive muscles during intense exercise in man. J Physiol (Lond) 488:219–229
Bergersen L, Warhaug O, Helm J, Thomas M, Laake P, Davies AJ, Wilson MC, Halestrap AP, Ottersen OP (2001) A novel postsynaptic density protein: The monocarboxylate transporter MCT2 is co-localized with glutamate receptors in postsynaptic densities of parallel fiber-Purkinje cell synapses. Exp Brain Res 136:523–534
Berry MN, Phillips JW, Grivell AR (1992) Interactions between mitochondria and cytoplasm in isolated hepatocytes. Curr Top Cell Regul 33:309–328
Bittar PG, Charnay Y, Pellerin L, Bouras C, Magistretti PJ (1996) Selective distribution of lactate dehydrogenase isoenzymes in neurons and astrocytes of human brain. J Cereb Blood Flow Metab 16:1079–1089
Bliss TM, Sapolsky RM (2001) Interactions among glucose, lactate and adenosine regulate energy substrate utilization in hippocampal cultures. Brain Res 899:134–141
Bouzier A-K, Thiaudiere E, Biran M, Rouland R, Canioni P, Merle M (2000) The metabolism of [3-13C]lactate in the rat brain is specific of a pyruvate carboxylase-deprived compartment. J Neurochem 75:480–486
Bouzier-Sore A-K, Merle M, Magistretti PJ, Pellerin L (2002) Feeding active neurons: (re)emergence of a nursing role for astrocytes. J Physiol (Paris) 96:273–282
Broer S, Rahman B, Pellegri G, Pellerin L, Martin JL, Verleysdonk S, Hamprecht B, Magistretti PJ (1997) Comparison of lactate transport in astroglial cells and monocarboxylate 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
Cahn RD, Kaplan NO, Levine L, Zwilling LE (1962) Nature and development of lactic dehydrogenases. Science 136:962–969
Cheeseman AJ, Clark JB (1988) Influence of the malate-aspartate shuttle on oxidative metabolism in synaptosomes. J Neurochem 50:1559–1565
Chesler M, Kaila K (1992) Modulation of pH by neuronal activity. Trends Neurosci 15:396–402
Chhina N, Kuestermann E, Halliday J, Simpson LJ, MacDonald IA, Bachelard HS, Morris PG (2001) Measurement of human tricarboxylic acid cycle rates during visual activation by 13C magnetic resonance spectroscopy. J Neurosci Res 66:737–746
Chicurel ME, Harris KM (1992) Three-dimensional analysis of the structure and composition of CA3 branched dendritic spines and their synaptic relationships with mossy fiber boutons in the rat hippocampus. J Comp Neurol 325:169–182
Chih CP, He J, Sly TS, Roberts EL Jr (2001a) Comparison of glucose and lactate as substrates during NMDA-induced activation of hippocampal slices. Brain Res 893:143–154
Chih CP, Lipton P, Roberts EL Jr (2001b) Do active cerebral neurons really use lactate rather than glucose? Trends Neurosci 24:573–578
Cimino M, Balduini W, Marini P, Cattabeni F, Court JA, Bianchi M, Magnani M (1998) Expression of hexokinase mRNA in human hippocampus. Molec Brain Res 53:297–300
Clarke DD, Lajtha AL, Maker HS (1989) Intermediary Metabolism. In: Basic Neurochemistry, fourth edn New York, Raven Press, pp. 542–550
Connett RJ, Gayeski TE, Honig CR (1985) Energy sources in fully aerobic rest-work transitions: A new role for glycolysis. Am J Physiol 248(6 Pt 2):H922–H929
Conti F, Weinberg RJ (1999) Shaping excitation at glutamatergic synapses. Trends Neurosci 22:451–458
Cox DDWG, Drower J, Bachelard HS (1985) Effects of metabolic inhibitors on evoked activity and the energy state of hippocampal slices superfused in vitro. Exp Brain Res 57:464–470
Cox DW, 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
Danbolt NC (2001) Glutamate uptake. Prog Neurobiol 65:1–105
Demestre M, Boutelle M, Fillenz M (1997) Stimulated release of lactate in freely moving rats is dependent on the uptake of glutamate. J Physiol (Lond) 499:825–832
Dienel GA, Cruz NF (2003) Neighborly interactions of metabolicallyactivated astrocytes in vivo. Neurochem Int 1334:1–16
Dienel GA, Hertz L (2001) Glucose and lactate metabolism during brain activation. J Neurosci Res 66:824–838
Dringen R, Gebhardt R, Hamprecht B (1993a) Glycogen in astrocytes: Possible function as lactate supply for neighboring cells. Brain Res 623:208–214
Dringen R, Schmoll D, Cesar M, Hamprecht B (1993b) Incorporation of radioactivity from [14C]lactate into the glycogen of cultured mouse astroglial cells: Evidence for gluconeogenesis in brain cells. Biol Chem 374:343–347
Dringen R, Wiesinger H, Hamprecht B (1993c) Uptake of L-lactate by cultured rat brain neurons. Neurosci Lett 163:5–7
Eriksson G, Peterson A, Iverfeldt K, Walum E (1995) Sodiumdependent glutamate uptake as an activator of oxidative metabolism in primary astrocyte cultures from newborn rat. Glia 15:152–156
Fellows LK, Boutelle MG (1993) Rapid changes in extracellular glucose levels and blood flow in the striatum of the freely moving rat. Brain Res 604:225–231
Fellows LK, Boutelle MG, Fillenz M (1993) Physiological stimulation increases nonoxidative glucose metabolism in the brain of the freely moving rat. J Neurochem 60:1258–1263
Ferrendelli JA, McDougal DB Jr (1971) The effect of audiogenic seizures on regional CNS energy reserves, glycolysis and citric acid cycle flux. J Neurochem 18:1207–1220
Fillenz M, Lowry JP (1998) Studies of the source of glucose in the extracellular compartment of the rat brain. Dev Neurosci 20:365–368
Forsyth R, Fray A, Boutelle M, Fillenz M, Middleditch C, Burchell A (1996) A role for astrocytes in glucose delivery to neurons. Dev Neurosci 18:360–370
Fox PT, Raichle ME (1986) Focal physiological uncoupling of cerebral blood flow and oxidative metabolism during somatosensory stimulation in human subjects. Proc Nat Acad Sci U S A 83:1140–1144
Fox PT, Raichle ME, Mintun MA, Dence C (1988) Nonoxidative glucose consumption during focal physiologic neural activity. Science 241:462–464
Fray AE, Forsyth RJ, Boutelle MG, Fillenz M (1996) The mechanisms controlling physiologically stimulated changes in rat brain glucose and lactate: A microdialysis study. J Physiol (Lond) 496:49–57
Gerhart DZ, Enerson BE, Zhdankina OY (1998) Expression of the monocarboxylate transporter MCT2 by rat brain glia. Glia 22:272–281
Gjedde A, Marrett S (2001) Glycolysis in neurons, not astrocytes, delays oxidative metabolism of human visual cortex during sustained checkerboard stimulation in vivo. J Cereb Blood Flow Metab 21:1384–1392
Gjedde A, Marrett S, Vafaee M (2002) Oxidative and nonoxidative metabolism of excited neurons and astrocytes. J Cereb Blood Flow Metab 22:1–14
Goldberg ND, Passonneau JV, Lowry OH (1966) Effects of changes in brain metabolism on the levels of citric acid cycle intermediates. J Biol Chem 241:3997–4003
Gruetter R (in press) Principles of the measurement of neuro-glial metabolism using in vivo13C NMR spectroscopy. In: Advances in molecular and cell biology: Neuronal cells in the nervous system: function and dysfunction (Hertz L, ed), Elsevier
Halestrap AP, Denton RM (1974) Specific inhibition of pyruvate transport in rat liver mitochondria and human erythrocytes by cyano-4-hydroxycinnamate. Biochem J 138:313–316
Harada M, Okuda C, Sawa T, Murakami T (1992) Cerebral extracellular glucose and lactate concentrations during and after moderate hypoxia in glucose-infused and saline-infused rats. Anesthesiology 77:728–734
Hassel B, Bråthe A (2000) Cerebral metabolism of lactate in vivo: Evidence for neuronal pyruvate carboxylation. J Cereb Blood Flow Metab 20:327–336
Hassel B, Sonnewald U, Fonnum F (1995) Glial-neuronal interactions as studied by cerebral metabolism of [2–13C]acetate and [1–13C]glucose: an ex vivo13C NMR spectroscopic study. J Neurochem 64:2773–2782
Hawkins RA, Miller AL, Nielsen RC, Veech RL (1973) The acute action of ammonia on rat brain metabolism in vivo. Biochem J 134:1001–1008
Hertz L, Swanson RA, Newman GC, Marrif H, Juurlink BHJ, Peng LA (1998) Can experimental conditions explain the discrepancy over glutamate stimulation of aerobic glycolysis. Dev Neurosci 20:339–347
Hochachka PW, Mommsen TP (1983) Protons and anaerobiosis. Science 219:1391–1397
Hu YB, Wilson GS (1997) A temporary local energy pool coupled to neuronal activity: Fluctuations of extracellular lactate levels in rat brain monitored with rapid-response enzyme-based sensor. J Neurochem 69:1484–1490
Ide K, Schmalruch IK, Quistorff B, Horn A, Secher NH (2000) Lactate, glucose and O2 uptake in human brain during recovery from maximal exercise. J Physiol (Lond) 522:159–164
Itoh Y, Esaki T, Shimoji K, Cook M, Law MJ, Kaufman E, Sokoloff L (2003) Dichloroacetate effects on glucose and lactate oxidation by neurons and astroglia in vitro and on glucose utilization by brain in vivo. Proc Nat Acad Sci U S A 100:4879–4884
Izumi Y, Benz AM, Katsuki H, Zorumski CF (1997) Endogenous monocarboxylates sustain hippocampal synaptic function and morphological integrity during energy deprivation. J Neurosci 17:9448–9457
Kadekaro M, Crane AM, Sokoloff L (1985) Differential effects of electrical stimulation of sciatic nerve on metabolic activity in spinal cord and dorsal root ganglion in the rat. Proc Nat Acad Sci U S A 82:6010–6013
Kanatani T, Mizuno K, Okada Y (1995) Effects of glycolytic metabolites on preservation of high energy phosphate level and synaptic transmission in the granule cells of guinea pig hippocampal slices. Experientia 51:213–216
Kao-Jen J, Wilson JE (1980) Localization of hexokinase in neural tissue: Electron microscopic studies of rat cerebellar cortex. J Neurochem 35:667–678
Knull HR (1978) Association of glycolytic enzymes with particulate fractions from nerve endings. Biochim Biophys Acta 522:1–9
Kunnecke B, Cerdan S, Seelig J (1993) Cerebral metabolism of [1,2-13C] glucose and [U-13C] 3-hydroxybutyrate in rat brain as detected by 13C NMR spectroscopy. NMR Biomed 6:264–277
Lai JC, Behar KL, Liang BB, Hertz L (1999) Hexokinase in astrocytes: Kinetic and regulatory properties. Metab Brain Dis 14:125–133
Lapidot A, Haber S (2000) Effect of acute insulin-induced hypoglycemia on fetal versus adult brain fuel utilization, assessed by 13C MRS isotopomer analysis of [U-13C]glucose metabolites. Dev Neurosci 22:444–455
Larrabee MG (1995) Lactate metabolism and its effects on glucose metabolism in an excised neural tissue. J Neurochem 64:1734–1741
Lebon V, Petersen KF, Cline GW, Shen J, Mason GF, Dufour S, Behar KL, Shulman GI, Rothman DL (2002) Astroglial contribution to brain energy metabolism in humans revealed by 13C nuclear magnetic resonance spectroscopy: elucidation of the dominant pathway for neurotransmitter glutamate repletion and measurement of astrocytic oxidative metabolism. J Neurosci 22:1523–1531
Lehninger AL, Nelson DL, Cox MM (1993) Principles of Biochemistry, 2nd ed., New York, Worth Publishers
Leino RL, Gerhart DZ, Vanbueren AM, McCall AL, Drewes LR (1997) Ultrastructural localization of GLUT 1 and GLUT 3 glucose transporters in rat brain. J Neurosci Res 49:617–626
Lim L, Hall C, Leung T, Mahadevan L, Whatley S (1983) Neuronespecific enolase and creatine phosphokinase are protein components of rat brain synaptic plasma membrane. J Neurochem 41:1177–1182
Lipton P, Robacker K (1983) Glycolysis and brain function: [K+]o stimulation of protein synthesis and K+ uptake require glycolysis. Fed Proc 42:2875–2880
Lopes-Cardozo M, Larsson OM, Schousboe A (1986) Acetoacetate and glucose as lipid precursors and energy substrates in primary cultures of astrocytes and neurons from mouse cerebral cortex. J Neurochem 46:773–778
Lowry OH, Passonneau JV (1964) The relationships between substrates and enzymes of glycolysis in brain. J Biol Chem 239:31–42
Magistretti PJ (1999) Brain energy metabolism. In: Fundamental neuroscience (Zigmond MJ, Bloom FE, Landis SC, Roberts JL, Squire LR, eds) New York, Academic Press, pp. 389–413
Magistretti PJ (2000) Cellular bases of functional brain imaging: Insights from neuron-glia metabolic coupling. Brain Res 886:108–112
Magistretti PJ, Pellerin L (1999) Cellular mechanisms of brain energy metabolism and their relevance to functional brain imaging. Philos Trans R Soc Lond B Biol Sci 354:1155–1163
Magistretti PJ, Pellerin L, Rothman DL, Shulman RG (1999) Energy on demand. Science 283:496–497
Malonek D, Grinvald A (1996) Interactions between electrical activity and cortical microcirculation revealed by imaging spectroscopy: Implications for functional brain mapping. Science 272:551–554
Malonek D, Dirnagl U, Lindauer U, Yamada K, Kanno I, Grinvald A (1997) Vascular imprints of neuronal activity: Relationships between the dynamics of cortical blood flow, oxygenation, and volume changes following sensory stimulation. Proc Nat Acad Sci U S A 94:14826–14831
Martin M, Portais JC, Labouesse J, Canioni P, Merle M (1993) [1–13C]Glucose metabolism in rat cerebellar granule cells and astrocytes in primary culture. Evaluation of flux parameters by 13C and 1H-NMR spectroscopy. Eur J Biochem 217:617–625
McCall AL, Van Bueren AM, Moholt-Siebert M, Cherry NJ, Woodward WR (1994) Immunohistochemical localization of the neuronspecific glucose transporter (GLUT3) to neuropil in adult rat brain. Brain Res 659:292–297
McIlwain H, Bachelard HS (1985) Biochemistry and the central nervous system, 5th edn, New York, Churchill Livingstone
McKenna MC, Tildon JT, Stevenson JH, Boatright R, Huang S (1993) Regulation of energy metabolism in synaptic terminals and cultured rat brain astrocytes: Differences revealed using aminooxyacetate. Dev Neurosci 15:320–329
McKenna MC, Hopkins IB, Carey A (2001) α-cyano-4-hydroxycinnamate decreases both glucose and lactate metabolism in neurons and astrocytes: Implications for lactate as an energy substrate for neurons. J Neurosci Res 66:747–754
Merle M, Bouzier-Sore AK, Canioni P (2002) Time-dependence of the contribution of pyruvate carboxylase versus pyruvate dehydrogenase to rat brain glutamine labeling from [1–13C]glucose metabolism. J Neurochem 82:47–57
Nemoto EM, Hoff JT, Severinghaus JW (1974) Lactate uptake and metabolism by brain during hyperlactatemia and hypoglycemia. Stroke 5:48–53
Nitisewojo P, Hultin HO (1976) A comparison of some kinetic properties of soluble and bound lactate dehydrogenase isoenzymes at different temperatures. Eur J Biochem 67:87–94
Paul RJ, Bauer W, Pease W (1979) Vascular smooth muscle: Aerobic glycolysis linked to sodium and potassium transport proceses. Science 206:1414–1416
Pearce FJ, Connett RJ (1980) Effect of lactate and palmitate on substrate utilization of isolated rat soleus. Am J Physiol 238:C149–C159
Pellerin L, Magistretti PJ (1994) Glutamate uptake into astrocytes stimulates aerobic glycolysis: A mechanism coupling neuronal activity to glucose utilization. Proc Nat Acad Sci U S A 91:10625–10629
Pellerin L, Pellegri G, Bittar PG, Charnay Y, Bouras C, Martin JL, Stella N, Magistretti PJ (1998a) Evidence supporting the existence of an activity-dependent astrocyte-neuron lactate shuttle. Dev Neurosci 20:291–299
Pellerin L, Pellegri G, Martin JL, Magistretti PJ (1998b) Expression of monocarboxylate transporter mRNAs in mouse brain: Support for a distinct role of lactate as an energy substrate for the neonatal vs. Adult brain. Proc Nat Acad Sci U S A 95:3990–3995
Peng L, Zhang XH, Hertz L (1994) High extracellular potassium concentrations stimulate oxidative metabolism in a glutamatergic neuronal culture and glycolysis in cultured astrocytes but have no stimulatory effect in a GABAergic neuronal culture. Brain Res 663:168–172
Pfeuffer J, Tkac I, Gruetter R (2000) Extracellular-intracellular distribution of glucose and lactate in the rat brain assessed noninvasively by diffusion-weighted 1H nuclear magnetic resonance spectroscopy in vivo. J Cereb Blood Flow Metab 20:736–746
Pierre K, Magistretti PJ, Pellerin L (2002) MCT2 is a major neuronal monocarboxylate transporter in the adult mouse brain. J Cereb Blood Flow Metab 22:586–595
Poitry-Yamate CL, Poitry S, Tsacopoulos M (1995) Lactate released by Muller glial cells is metabolized by photoreceptors from mammalian retina. J Neurosci 15:5179–5191
Prichard J, Rothman D, Novotny E, Petroff O, Kuwabara T, Avison M, Howseman A, Hanstock C, Shulman R (1991) Lactate rise detected by 1H NMR in human visual cortex during physiologic stimulation. Proc Nat Acad Sci U S A 88:5829–5831
Proverbio F, Hoffman JF (1977) Membrane compartmentalized ATP and its preferential use by Na,K-ATPase of human red cell ghosts. J Gen Physiol 69:605–632
Qu H, Haberg A, Haraldseth O, Unsgard G, Sonnewald U (2000) C-13 MR spectroscopy study of lactate as substrate for rat brain. Dev Neurosci 22:429
Raichle ME (1991) The metabolic requirements of functional activity in the human brain: A positron emission tomography study. Adv Exp Med Biol 291:1–4
Ridderståle Y, Wistrand PJ (1998) Carbonic anhydrase isoforms in the mammalian nervous system. In: pH and Brain Function (Kaila K, Ransom BR, eds) New York, Wiley-Liss, Inc., pp. 21–43
Roberts EL Jr (1993) Glycolysis and recovery of potassium ion homeostasis and synaptic transmission in hippocampal slices after anoxia or stimulated potassium release. Brain Res 620:251–258
Roberts EL Jr, Chih CP (in press) A role for lactate released from astrocytes in energy production during neural activity? In: Advances in molecular and cell biology: Neuronal cells in the nervous system: function and dysfunction (Hertz L, ed), Elsevier
Roberts EL Jr, Feng ZC (1996) Influence of age on the clearance of K+ from the extracellular space of rat hippocampal slices. Brain Res 708:16–20
Sappeymarinier D, Calabrese G, Fein G, Hugg JW, Biggins C, Weiner MW (1992) Effect of photic stimulation on human visual cortex lactate and phosphates using 1H and 31P magnetic resonance spectroscopy. J Cereb Blood Flow Metab 12:584–592
Schmitt BM, Berger UV, Douglas RM, Bevensee MO, Hediger MA, Haddad GG, Boron WF (2000) Na/HCO3 cotransporters in rat brain: Expression in glia, neurons, and choroid plexus. J Neurosci 20:6839–6848
Schneider U, Poole RC, Halestrap AP, Grafe P (1993) Lactate-proton co-transport and its contribution to interstitial acidification during hypoxia in isolated rat spinal roots. Neuroscience 53:1153–1162
Schousboe A, Westergaard N, Waagepetersen HS, Larsson OM, Bakken IJ, Sonnewald U (1997) Trafficking between glia and neurons of TCA cycle intermediates and related metabolites. Glia 21:99–105
Schurr A, Rigor BM (1998) Brain anaerobic lactate production - a suicide note or a survival kit. Dev Neurosci 20:348–357
Schurr A, West CA, Rigor BM (1988) Lactate-supported synaptic function in the rat hippocampal slice preparation. Science 240:1326–1328
Schurr A, Payne RS, Miller JJ, Rigor BM (1997) Brain lactate is an obligatory aerobic energy substrate for functional recovery after hypoxia: Further in vitro validation. J Neurochem 69:423–426
Schurr A, Miller JJ, Payne RS, Rigor BM (1999) An increase in lactate output by brain tissue serves to meet the energy needs of glutamate-activated neurons. J Neurosci 19:34–39
Shank RP, Leo GC, Zielke HR (1993) Cerebral metabolic compartmentation as revealed by nuclear magnetic resonance analysis of D-[1–13C]glucose metabolism. J Neurochem 61:315–323
Shepherd GMG, Harris KM (1998) Three-dimensional structure and composition of CA3-CA1 axons in rat hippocampal slices: Implications for presynaptic connectivity and compartmentalization. J Neurosci 18:8300–8310
Sibson NR, Dhankhar A, Mason GF, Rothman DL, Behar KL, Shulman RG (1998) Stoichiometric coupling of brain glucose metabolism and glutamatergic neuronal activity. Proc Nat Acad Sci U S A 95:316–321
Siesjö BK (1982) Lactic acidosis in the brain: Occurrence, triggering mechanisms and pathophysiological importance. Ciba Found Symp 87:77–100
Silver IA, Erecinska M (1994) Extracellular glucose concentration in mammalian brain: Continuous monitoring of changes during increased neuronal activity and upon limitation in oxygen supply in normo-, hypo-, and hyperglycemic animals. J Neurosci 14:5068–5076
Silver IA, Erecinska M (1997) Energetic demands of the NA+/K+ ATPase in mammalian astrocytes. Glia 21:35–45
Silver IA, Deas J, Erecinska M (1997) Ion homeostasis in brain cells: Differences in intracellular ion responses to energy limitation between cultured neurons and glial cells. Neuroscience 78:589–601
Sokoloff L (1989) Circulation and energy metabolism of the brain. In: Basic Neurochemistry (Siegel G, Agranoff B, Albers RW, Molinoff P, eds) New York, Raven Press, pp. 565–590
Sokoloff L (1999) Energetics of functional activation in neural tissues. Neurochem Res 24:321–329
Sorra KE, Harris KM (2000) Overview on the structure, composition, function, development, and plasticity of hippocampal dendritic spines. Hippocampus 10:501–511
Stambaugh R, Post D (1966) Substrate and product inhibition of rabbit muscle lactic dehydrogenase heart (H4) and muscle (M4) isozymes. J Biol Chem 241(7):1462–1467
Swanson RA (1992) Astrocyte glutamate uptake during chemical hypoxia in vitro. Neurosci Lett 147:143–146
Tabernero A, Bolanos JP, Medina JM (1993) Lipogenesis from lactate in rat neurons and astrocytes in primary culture. Biochem J 294:635–638
Tabernero A, Vicario C, Medina JM (1996) Lactate spares glucose as a metabolic fuel in neurons and astrocytes from primary culture. Neurosci Res 26:369–376
Takata T, Okada Y (1995) Effects of deprivation of oxygen or glucose on the neural activity in the guinea pig hippocampal slice - intracellular recording study of pyramidal neurons. Brain Res 683:109–116
Tildon JT, McKenna MC, Stevenson J, Couto R (1993) Transport of L-lactate by cultured rat brain astrocytes. Neurochem Res 18:177–184
Tsacopoulos M, Magistretti PJ (1996) Metabolic coupling between glia and neurons. J Neurosci 16:877–885
Ueki M, Linn F, Hossmann KA (1988) Functional activation of cerebral blood flow and metabolism before and after global ischemia of rat brain. J Cereb Blood Flow Metab 8:486–494
Vannucci SJ, Maher F, Simpson IA (1997) Glucose transporter proteins in brain: Delivery of glucose to neurons and glia. Glia 21:2–21
Vicario C, Tabernero A, Medina JM (1993) Regulation of lactate metabolism by albumin in rat neurons and astrocytes from primary culture. Ped Res 34:709–715
Waagepetersen HS, Sonnewald U, Larsson OM, Schousboe A (2000) A possible role of alanine for ammonia transfer between astrocytes and glutamatergic neurons. J Neurochem 75:471–479
Wada H, Okada Y, Nabetani M, Nakamura H (1997) The effects of lactate and β-hydroxybutyrate on the energy metabolism and neural activity of hippocampal slices from adult and immature rat. Dev Brain Res 101:1–7
Walz W, Mukerji S (1988) Lactate release from cultured astrocytes and neurons: A comparison. Glia 1:366–370
Weiss J, Hiltbrand B (1985) Functional compartmentation of glycolytic versus oxidative metabolism in isolated rabbit heart. J Clin Invest 75:436–447
Wender R, Brown AM, Fern R, Swanson RA, Farrell K, Ransom BR (2000) Astrocytic glycogen influences axon function and survival during glucose derivation in central white matter. J Neurosci 20:6804–6810
Williamson JR (1965) Metabolic control in the perfused rat heart. In: Control of energy metabolism (Chance B, Estabrook RW, Williamson JR, eds) New York, Academic Press, pp. 333–355
Willoughby D, Schwiening CJ (2002) Electrically evoked dendritic pH transients in rat cerebellar Purkinje cells. J Physiol (Lond) 544:487–499
Wilson JE (1972) The localization of latent brain hexokinase on synaptosomal mitochondria. Arch Biochem Biophys 150:96–104
Wittendorp-Rechenmann E, Lam CD, Lasbennes F, Nehlig A (2001) First in vivo demonstrateion of the uptake of [14C]2-deoxyglucose by astrocytes and neurons: a microautoradiographic study. J Cereb Blood Flow Metab 21, Suppl 1:S321
Wu K, Aoki C, Elste A, Rogalski-Wilk AA, Siekevitz P (1997) The synthesis of ATP by glycolytic enzymes in the postsynaptic density and the effect of endogenously generated nitric oxide. Proc Nat Acad Sci U S A 94:13273–13278
Zwingmann C, Richter-Landsberg C, Brand A, Leibfritz D (2000) NMR spectroscopic study on the metabolic fate of [3–13C]alanine in astrocytes, neurons, and cocultures: Implications for glia-neuron interactions in neurotransmitter metabolism. Glia 32:286–303

Cite article

Cite article

Cite article

OR

Download to reference manager

If you have citation software installed, you can download article citation data to the citation manager of your choice

Share options

Share

Share this article

Share with email
EMAIL ARTICLE LINK
Share on social media

Share access to this article

Sharing links are not relevant where the article is open access and not available if you do not have a subscription.

For more information view the Sage Journals article sharing page.

Information, rights and permissions

Information

Published In

Article first published: November 2003
Issue published: November 2003

Keywords

  1. Brain
  2. Energy metabolism
  3. Glucose
  4. Glycolysis
  5. Lactate
  6. Oxidative phosphorylation

Rights and permissions

© 2003 The International Society for Cerebral Blood Flow and Metabolism.
Request permissions for this article.
PubMed: 14600433

Authors

Affiliations

Ching-Ping Chih
Geriatric Research, Education, and Clinical Center, and Research Office, Miami VA Medical Center, Miami, Florida
Eugene L Roberts, Jr.
Geriatric Research, Education, and Clinical Center, and Research Office, Miami VA Medical Center, Miami, Florida
Department of Neurology, University of Miami School of Medicine, Miami, Florida

Notes

Address correspondence and reprint requests to Eugene L. Roberts, Jr., Department of Neurology, University of Miami School of Medicine, P.O. Box 016960, Miami, FL 33101 U.S.A.; e-mail: [email protected].

Metrics and citations

Metrics

Journals metrics

This article was published in Journal of Cerebral Blood Flow & Metabolism.

VIEW ALL JOURNAL METRICS

Article usage*

Total views and downloads: 3675

*Article usage tracking started in December 2016


Altmetric

See the impact this article is making through the number of times it’s been read, and the Altmetric Score.
Learn more about the Altmetric Scores



Articles citing this one

Receive email alerts when this article is cited

Web of Science: 238 view articles Opens in new tab

Crossref: 0

  1. Lactate as a supplemental fuel for synaptic transmission and neuronal ...
    Go to citation Crossref Google Scholar
  2. A tribute to Leif Hertz: The historical context of his pioneering stud...
    Go to citation Crossref Google Scholar
  3. In vivo calibration of genetically encoded metabolite biosensors must ...
    Go to citation Crossref Google Scholar
  4. Functional brain network controllability dysfunction in Alzheimer's di...
    Go to citation Crossref Google Scholar
  5. ATP-mediated increase in H + efflux from r...
    Go to citation Crossref Google Scholar
  6. Resting state and activated brain glutamate–glutamine, brain lactate, ...
    Go to citation Crossref Google Scholar
  7. A monocarboxylate transporter-dependent mechanism confers resistance t...
    Go to citation Crossref Google Scholar
  8. A functional account of stimulation-based aerobic glycolysis and its r...
    Go to citation Crossref Google Scholar
  9. Impact of Opioids on Cellular Metabolism: Implications for Metabolic P...
    Go to citation Crossref Google Scholar
  10. Monocarboxylate transporter-dependent mechanism is involved in the ada...
    Go to citation Crossref Google Scholar
  11. Aging and memory are altered by genetically manipulating lactate dehyd...
    Go to citation Crossref Google Scholar
  12. State of the Science on Brain Insulin Resistance and Cognitive Decline...
    Go to citation Crossref Google Scholar
  13. From antioxidant to neuromodulator: The role of ascorbate in the manag...
    Go to citation Crossref Google Scholar
  14. Differential role of neuronal glucose and PFKFB3 ...
    Go to citation Crossref Google Scholar
  15. Multimodality Monitoring and Goal-Directed Therapy for the Treatment o...
    Go to citation Crossref Google Scholar
  16. In Brief
    Go to citation Crossref Google Scholar
  17. Brain energy failure in dementia syndromes: Opportunities and challeng...
    Go to citation Crossref Google Scholar
  18. Increase of Lactate Concentration During Spreading Depression
    Go to citation Crossref Google Scholar
  19. Artificial neurovascular network (ANVN) to study the accuracy vs. effi...
    Go to citation Crossref Google Scholar
  20. Overview of age‐related changes in psychomotor and cognitive functions...
    Go to citation Crossref Google Scholar
  21. Dihydrocapsaicin effectively mitigates cerebral ischemia-induced patho...
    Go to citation Crossref Google Scholar
  22. Alzheimer's disease alters oligodendrocytic glycolytic and ketolytic g...
    Go to citation Crossref Google Scholar
  23. Possible mechanisms of HIV neuro-infection in alcohol use: Interplay o...
    Go to citation Crossref Google Scholar
  24. Participation of L-Lactate and Its Receptor HCAR1/GPR81 in Neurovisual...
    Go to citation Crossref Google Scholar
  25. Hypothalamic Astrocytes as a Specialized and Responsive Cell Populatio...
    Go to citation Crossref Google Scholar
  26. Effect of Diet as a Factor of Exposome on Brain Function
    Go to citation Crossref Google Scholar
  27. Exploring the crosstalk of glycolysis and mitochondrial metabolism in ...
    Go to citation Crossref Google Scholar
  28. Integrating hippocampal metabolomics and network pharmacology decipher...
    Go to citation Crossref Google Scholar
  29. Saikosaponin D Rescues Deficits in Sexual Behavior and Ameliorates Neu...
    Go to citation Crossref Google Scholar
  30. Energy Metabolism | Brain Energy Metabolism
    Go to citation Crossref Google Scholar
  31. Disfunções neurológicas e declínio cognitivo
    Go to citation Crossref Google Scholar
  32. Universal Glia to Neurone Lactate Transfer in the Nervous System: Phys...
    Go to citation Crossref Google Scholar
  33. Information Processing in Affective Disorders: Did an Ancient Peptide ...
    Go to citation Crossref Google Scholar
  34. Evidence of the Role of Omega-3 Polyunsaturated Fatty Acids in Brain G...
    Go to citation Crossref Google Scholar
  35. Computational singular perturbation analysis of brain lactate metaboli...
    Go to citation Crossref Google Scholar
  36. Astrocytes: Heterogeneous and Dynamic Phenotypes in Neurodegeneration ...
    Go to citation Crossref Google ScholarPub Med
  37. Consequences of a Metabolic Glucose-Depletion on the Survival and the ...
    Go to citation Crossref Google Scholar
  38. Brain energetics plays a key role in the coordination of electrophysio...
    Go to citation Crossref Google Scholar
  39. The energetic brain – A review from students to students
    Go to citation Crossref Google Scholar
  40. Neuron-specific deficits of bioenergetic processes in the dorsolateral...
    Go to citation Crossref Google Scholar
  41. The involvement of astrocytes in early‐life adversity induced programm...
    Go to citation Crossref Google Scholar
  42. Energetic substrate availability regulates synchronous activity in an ...
    Go to citation Crossref Google Scholar
  43. Mechanisms for the maintenance and regulation of axonal energy supply
    Go to citation Crossref Google Scholar
  44. A SLC16A1 Mutation in an Infant With Ketoacidosis and Neuroimaging Ass...
    Go to citation Crossref Google Scholar
  45. In Vivo NMR Spectroscopy – Static Aspects
    Go to citation Crossref Google Scholar
  46. Metabolism and metabolomics of opiates: A long way of forensic implica...
    Go to citation Crossref Google Scholar
  47. Development of a Model to Test Whether Glycogenolysis Can Support Astr...
    Go to citation Crossref Google Scholar
  48. Tissue-specific kinase expression and activity regulate flux through t...
    Go to citation Crossref Google Scholar
  49. Brain Glucose Metabolism: Integration of Energetics with Function
    Go to citation Crossref Google Scholar
  50. Astrocyte‐neuron lactate shuttle sensitizes nociceptive transmission i...
    Go to citation Crossref Google Scholar
  51. Glycolysis Paradigm Shift Dictates a Reevaluation of Glucose and Oxyge...
    Go to citation Crossref Google Scholar
  52. One-pot enzymatic synthesis of l-[3-11C]lactate for pharmacokinetic an...
    Go to citation Crossref Google Scholar
  53. Peroxidase-mimicking PtNP-coated, 3D-printed multi-well plate for rapi...
    Go to citation Crossref Google Scholar
  54. A Comparison of Oxidative Lactate Metabolism in Traumatically Injured ...
    Go to citation Crossref Google Scholar
  55. PDGF/PDGFR axis in the neural systems
    Go to citation Crossref Google Scholar
  56. Astrocyte glycogen and lactate: New insights into learning and memory ...
    Go to citation Crossref Google Scholar
  57. Defects in Bioenergetic Coupling in Schizophrenia
    Go to citation Crossref Google Scholar
  58. Lactate metabolism: historical context, prior misinterpretations, and ...
    Go to citation Crossref Google Scholar
  59. Neurochemical changes in unilateral cerebral hemisphere during the sub...
    Go to citation Crossref Google Scholar
  60. 3.15 Neuronal Energy Production
    Go to citation Crossref Google Scholar
  61. High-fat diet reduces the hippocampal content level of lactate which i...
    Go to citation Crossref Google Scholar
  62. Physiology of Astroglia
    Go to citation Crossref Google Scholar
  63. Metabolic Changes Associated with a Rat Model of Diabetic Depression D...
    Go to citation Crossref Google Scholar
  64. Lanthanum damages learning and memory and suppresses astrocyte–neuron ...
    Go to citation Crossref Google Scholar
  65. Lack of appropriate stoichiometry: Strong evidence against an energeti...
    Go to citation Crossref Google Scholar
  66. Biology of obesity and weight regain
    Go to citation Crossref Google Scholar
  67. Laser Acupuncture at GV20 Improves Brain Damage and Oxidative Stress i...
    Go to citation Crossref Google Scholar
  68. Glucose metabolism in nerve terminals
    Go to citation Crossref Google Scholar
  69. Astroglia as a cellular target for neuroprotection and treatment of ne...
    Go to citation Crossref Google Scholar
  70. Aspects on the Physiological and Biochemical Foundations of Neurocriti...
    Go to citation Crossref Google Scholar
  71. Key Metabolic Enzymes Underlying Astrocytic Upregulation of GABAergic ...
    Go to citation Crossref Google Scholar
  72. Monocarboxylate Transporter 1 in the Medial Prefrontal Cortex Developm...
    Go to citation Crossref Google Scholar
  73. Emerging Roles for Glycogen in the CNS
    Go to citation Crossref Google Scholar
  74. Before, after, & after-after
    Go to citation Crossref Google Scholar
  75. Lactate Shuttles in Neuroenergetics—Homeostasis, Allostasis and Beyond
    Go to citation Crossref Google Scholar
  76. Effects of Systemic Metabolic Fuels on Glucose and Lactate Levels in t...
    Go to citation Crossref Google Scholar
  77. How AMPK and PKA Interplay to Regulate Mitochondrial Function and Surv...
    Go to citation Crossref Google Scholar
  78. Uncertainty quantification in flux balance analysis of spatially lumpe...
    Go to citation Crossref Google Scholar
  79. PINK1 expression increases during brain development and stem cell diff...
    Go to citation Crossref Google Scholar
  80. Hippocampus-based contextual memory alters the morphological character...
    Go to citation Crossref Google Scholar
  81. Microdialysate concentration changes do not provide sufficient informa...
    Go to citation Crossref Google ScholarPub Med
  82. The Effects of Capillary Transit Time Heterogeneity (CTH) on the Cereb...
    Go to citation Crossref Google Scholar
  83. Metabolic reprogramming by the pyruvate dehydrogenase kinase–lactic ac...
    Go to citation Crossref Google Scholar
  84. Biosynthesis of glycerol phosphate is associated with long-term potent...
    Go to citation Crossref Google Scholar
  85. Enzyme-Immobilized 3D-Printed Reactors for Online Monitoring of Rat Br...
    Go to citation Crossref Google Scholar
  86. Exercise Counteracts Aging-Related Memory Impairment: A Potential Role...
    Go to citation Crossref Google Scholar
  87. Cerebral Metabolic Profiling of Hypothermic Circulatory Arrest with an...
    Go to citation Crossref Google Scholar
  88. The Role of Lactate-Mediated Metabolic Coupling between Astrocytes and...
    Go to citation Crossref Google Scholar
  89. Pyruvate Dehydrogenase Kinase-mediated Glycolytic Metabolic Shift in t...
    Go to citation Crossref Google Scholar
  90. Glutamate Metabolism
    Go to citation Crossref Google Scholar
  91. Functional magnetic resonance imaging
    Go to citation Crossref Google Scholar
  92. Effect of subconjunctival glucose on retinal ganglion cell survival in...
    Go to citation Crossref Google Scholar
  93. Neuromodulation accompanying focused ultrasound-induced blood-brain ba...
    Go to citation Crossref Google Scholar
  94. Old Things New View: Ascorbic Acid Protects the Brain in Neurodegenera...
    Go to citation Crossref Google Scholar
  95. Direct neuronal glucose uptake heralds activity-dependent increases in...
    Go to citation Crossref Google Scholar
  96. The relationship between iron dyshomeostasis and amyloidogenesis in Al...
    Go to citation Crossref Google Scholar
  97. Electro-acupuncture up-regulates astrocytic MCT1 expression to improve...
    Go to citation Crossref Google Scholar
  98. A 1H-NMR Based Study on Hemolymph Metabolomics in Eri Silkworm after O...
    Go to citation Crossref Google Scholar
  99. The restless brain: how intrinsic activity organizes brain function
    Go to citation Crossref Google Scholar
  100. MicroRNA-7 Promotes Glycolysis to Protect against 1-Methyl-4-phenylpyr...
    Go to citation Crossref Google Scholar
  101. Cancer‐like metabolism of the mammalian retina
    Go to citation Crossref Google Scholar
  102. Central Presynaptic Terminals Are Enriched in ATP but the Majority Lac...
    Go to citation Crossref Google Scholar
  103. The metabolic response to excitotoxicity – lessons from single-cell im...
    Go to citation Crossref Google Scholar
  104. Distinct Functional States of Astrocytes During Sleep and Wakefulness:...
    Go to citation Crossref Google Scholar
  105. Methylglyoxal, the dark side of glycolysis
    Go to citation Crossref Google Scholar
  106. Contributions of glycogen to astrocytic energetics during brain activa...
    Go to citation Crossref Google Scholar
  107. Brain aerobic glycolysis functions and Alzheimer’s disease
    Go to citation Crossref Google Scholar
  108. Posterior Cingulate Lactate as a Metabolic Biomarker in Amnestic Mild ...
    Go to citation Crossref Google Scholar
  109. Energy substrates that fuel fast neuronal network oscillations
    Go to citation Crossref Google Scholar
  110. Cellular Origin and Regulation of D-and L-Serine in in Vitro and in Vi...
    Go to citation Crossref Google ScholarPub Med
  111. Ex vivo 1H nuclear magnetic resonance spectroscopy reveals systematic ...
    Go to citation Crossref Google Scholar
  112. Cerebral glycolysis: a century of persistent misunderstanding and misc...
    Go to citation Crossref Google Scholar
  113. Direct evidence for activity-dependent glucose phosphorylation in neur...
    Go to citation Crossref Google Scholar
  114. Astrogliosis Is a Possible Player in Preventing Delayed Neuronal Death
    Go to citation Crossref Google Scholar
  115. Mechanisms of Metabonomic for a Gateway Drug: Nicotine Priming Enhance...
    Go to citation Crossref Google Scholar
  116. Energy Metabolism in the Brain
    Go to citation Crossref Google Scholar
  117. Fatty Acids in Energy Metabolism of the Central Nervous System
    Go to citation Crossref Google Scholar
  118. Brain Ischemia in Patients with Intracranial Hemorrhage: Pathophysiolo...
    Go to citation Crossref Google ScholarPub Med
  119. Stroke Neuroprotection: Targeting Mitochondria
    Go to citation Crossref Google Scholar
  120. Regulation and dysregulation of astrocyte activation and implications ...
    Go to citation Crossref Google Scholar
  121. Glucose metabolism disorder is a risk factor in ethanol exposure induc...
    Go to citation Crossref Google Scholar
  122. A Mathematical Model of the Metabolic and Perfusion Effects on Cortica...
    Go to citation Crossref Google Scholar
  123. Modeling of Brain Metabolism and Pyruvate Compartmentation Using 13C N...
    Go to citation Crossref Google ScholarPub Med
  124. Function of the master energy regulator adenosine monophosphate‐activa...
    Go to citation Crossref Google Scholar
  125. Subcellular Biochemical Investigation of Purkinje Neurons Using Synchr...
    Go to citation Crossref Google Scholar
  126. Pink1 deficiency attenuates astrocyte proliferation through mitochondr...
    Go to citation Crossref Google Scholar
  127. Reevaluating Metabolism in Alzheimer's Disease from the Perspective of...
    Go to citation Crossref Google Scholar
  128. Metabolic Changes Detected by Ex Vivo High Resolution 1H NMR Spectrosc...
    Go to citation Crossref Google Scholar
  129. Adenosine and Energy Metabolism—Relationship to Brain Bioenergetics
    Go to citation Crossref Google Scholar
  130. Carbohydrate Metabolism in the Central Nervous System
    Go to citation Crossref Google Scholar
  131. Energy Metabolism | Carbohydrate Metabolism in the Central Nervous Sys...
    Go to citation Crossref Google Scholar
  132. The Metabolism of Neurons and Astrocytes Through Mathematical Models
    Go to citation Crossref Google Scholar
  133. Functional magnetic resonance imaging
    Go to citation Crossref Google Scholar
  134. A mutation in Drosophila Aldolase Causes T...
    Go to citation Crossref Google Scholar
  135. Ménage à Trois: The Role of Neurotransmitters in the Energy Metabolism...
    Go to citation Crossref Google ScholarPub Med
  136. Fueling and Imaging Brain Activation
    Go to citation Crossref Google ScholarPub Med
  137. Brain Lactate Metabolism: The Discoveries and the Controversies
    Go to citation Crossref Google ScholarPub Med
  138. 1H NMR-based metabonomic analysis of brain in rats of morphine depende...
    Go to citation Crossref Google Scholar
  139. Physical and Functional Interaction of NCX1 and EAAC1 Transporters Lea...
    Go to citation Crossref Google Scholar
  140. Astroglial Pentose Phosphate Pathway Rates in Response to High-Glucose...
    Go to citation Crossref Google ScholarPub Med
  141. Glucose Metabolism During Neural Activation
    Go to citation Crossref Google Scholar
  142. Brain Metabolic Compartmentalization, Metabolism Modeling, and Cerebra...
    Go to citation Crossref Google Scholar
  143. Energy Metabolism of the Brain
    Go to citation Crossref Google Scholar
  144. Effects of fluctuating glucose concentrations on oxidative metabolism ...
    Go to citation Crossref Google Scholar
  145. Astrogliopathy as a loss of astroglial protective function against gly...
    Go to citation Crossref Google Scholar
  146. Biosensors Based On Prussian Blue Modified Carbon Fibers Electrodes fo...
    Go to citation Crossref Google Scholar
  147. Ascorbic acid-dependent GLUT3 inhibition is a critical step for switch...
    Go to citation Crossref Google Scholar
  148. Brain Energy Metabolism: Focus on Astrocyte-Neuron Metabolic Cooperati...
    Go to citation Crossref Google Scholar
  149. Gliovascular and cytokine interactions modulate brain endothelial barr...
    Go to citation Crossref Google Scholar
  150. Sympathetic influence on cerebral blood flow and metabolism during exe...
    Go to citation Crossref Google Scholar
  151. Multi-dimensional MR spectroscopy: towards a better understanding of h...
    Go to citation Crossref Google Scholar
  152. Cellular and Metabolic Origins of Flavoprotein Autofluorescence in the...
    Go to citation Crossref Google Scholar
  153. Astrocytes and energy metabolism
    Go to citation Crossref Google Scholar
  154. Why does the brain (not) have glycogen?
    Go to citation Crossref Google Scholar
  155. Dynamic activation model for a glutamatergic neurovascular unit
    Go to citation Crossref Google Scholar
  156. Comparison of Metabolic Profiling of Cyanidin-3- O ...
    Go to citation Crossref Google Scholar
  157. Simultaneous Monitoring of Tissue Po2 and NADH Fluorescence During Syn...
    Go to citation Crossref Google ScholarPub Med
  158. Gamma oscillations in the hippocampus require high complex I gene expr...
    Go to citation Crossref Google Scholar
  159. The role(s) of astrocytes and astrocyte activity in neurometabolism, n...
    Go to citation Crossref Google Scholar
  160. Neurons and Neuronal Stem Cells Survive in Glucose-Free Lactate and in...
    Go to citation Crossref Google Scholar
  161. Targeting Intermediary Metabolism in the Hypothalamus as a Mechanism t...
    Go to citation Crossref Google Scholar
  162. Two views of brain function
    Go to citation Crossref Google Scholar
  163. Myelination and the trophic support of long axons
    Go to citation Crossref Google Scholar
  164. Changes in Glucose Uptake Rather than Lactate Shuttle Take Center Stag...
    Go to citation Crossref Google ScholarPub Med
  165. Stoffwechselnotfälle
    Go to citation Crossref Google Scholar
  166. Pioglitazone enhances pyruvate and lactate oxidation in cultured neuro...
    Go to citation Crossref Google Scholar
  167. Astrocyte–neuron interactions in neurological disorders
    Go to citation Crossref Google Scholar
  168. Astrocytes are poised for lactate trafficking and release from activat...
    Go to citation Crossref Google Scholar
  169. A metabolic switch in brain: glucose and lactate metabolism modulation...
    Go to citation Crossref Google Scholar
  170. Astrocytes as the Glucose Shunt for Glutamatergic Neurons at High Acti...
    Go to citation Crossref Google Scholar
  171. The in vivo neuron‐to‐astrocyte lactate sh...
    Go to citation Crossref Google Scholar
  172. Metabolic and Hemodynamic Events after Changes in Neuronal Activity: C...
    Go to citation Crossref Google ScholarPub Med
  173. Role of NMDA receptor upon [14C]acetate uptake into intact rat brain
    Go to citation Crossref Google Scholar
  174. The Potential Role of Glucose Transport Changes in Hot Flash Physiolog...
    Go to citation Crossref Google ScholarPub Med
  175. Glucose administration prior to a divided attention task improves trac...
    Go to citation Crossref Google Scholar
  176. Glycogen Metabolism in CNS White Matter
    Go to citation Crossref Google Scholar
  177. Astrocyte and Neuron Intone Through Glutamate
    Go to citation Crossref Google Scholar
  178. Imaging Brain Activation
    Go to citation Crossref Google Scholar
  179. Ascorbic acid participates in a general mechanism for concerted glucos...
    Go to citation Crossref Google Scholar
  180. Lactate production and neurotransmitters; evidence from microdialysis ...
    Go to citation Crossref Google Scholar
  181. Modulation of astrocytic metabolic phenotype by proinflammatory cytoki...
    Go to citation Crossref Google Scholar
  182. Uridine Protects Cortical Neurons from Glucose Deprivation-Induced Dea...
    Go to citation Crossref Google Scholar
  183. A glia–neuron alanine/ammonium shuttle is central to energy metabolism...
    Go to citation Crossref Google Scholar
  184. Coupling between neuronal activity and microcirculation: Implications ...
    Go to citation Crossref Google Scholar
  185. Effect of Alternate Energy Substrates on Mammalian Brain Metabolism Du...
    Go to citation Crossref Google Scholar
  186. Cerebral Metabolism
    Go to citation Crossref Google Scholar
  187. Norepinephrine activity, as measured by MHPG, is associated with menop...
    Go to citation Crossref Google Scholar
  188. Metabolic compartmentalization in the human cortex and hippocampus: ev...
    Go to citation Crossref Google Scholar
  189. Lactate uptake contributes to the NAD(P)H biphasic response and tissue...
    Go to citation Crossref Google Scholar
  190. Timing of potential and metabolic brain energy
    Go to citation Crossref Google Scholar
  191. Supply and Demand in Cerebral Energy Metabolism: The Role of Nutrient ...
    Go to citation Crossref Google ScholarPub Med
  192. Role of glutamate transporters in the modulation of stress-induced lac...
    Go to citation Crossref Google Scholar
  193. Caudal hindbrain lactate infusion alters glucokinase, SUR1, and neuron...
    Go to citation Crossref Google Scholar
  194. d-Lactate inhibition of memory in a single trial discrimination avoida...
    Go to citation Crossref Google Scholar
  195. Compartmentalization of brain energy metabolism between glia and neuro...
    Go to citation Crossref Google Scholar
  196. Cellular pathways of energy metabolism in the brain: Is glucose used b...
    Go to citation Crossref Google Scholar
  197. Activity‐dependent regulation of energy metabolism by astrocytes: An u...
    Go to citation Crossref Google Scholar
  198. Intracellular ascorbic acid inhibits transport of glucose by neurons, ...
    Go to citation Crossref Google Scholar
  199. A glycogen phosphorylase inhibitor selectively enhances local rates of...
    Go to citation Crossref Google Scholar
  200. The Effect of Dietary Intake on Hot Flashes in Menopausal Women
    Go to citation Crossref Google Scholar
  201. Kinetic Parameters and Lactate Dehydrogenase Isozyme Activities Suppor...
    Go to citation Crossref Google Scholar
  202. Is lactate food for neurons? Comparison of monocarboxylate transporter...
    Go to citation Crossref Google Scholar
  203. Mitochondria and neuronal activity
    Go to citation Crossref Google Scholar
  204. ABCC8 and ABCC9: ABC transporters that regulate K+ channels
    Go to citation Crossref Google Scholar
  205. 8.3 Modeling of Electron Transport: Implications to Mitochondrial Dise...
    Go to citation Crossref Google Scholar
  206. 2.1 The Support of Energy Metabolism in the Central Nervous System wit...
    Go to citation Crossref Google Scholar
  207. Metabolic Biopsy of the Brain
    Go to citation Crossref Google Scholar
  208. Glucose Oxidase-immobilized Glass Disks for Imaging of d-Glucose in Ac...
    Go to citation Crossref Google Scholar
  209. Automated no-carrier-added synthesis of [1-11C]-labeled d- and l-enant...
    Go to citation Crossref Google Scholar
  210. Glucose is Necessary to Maintain Neurotransmitter Homeostasis during S...
    Go to citation Crossref Google ScholarPub Med
  211. Elevated lactate suppresses neuronal firing in vivo and inhibits gluco...
    Go to citation Crossref Google Scholar
  212. BRAIN WORK AND BRAIN IMAGING
    Go to citation Crossref Google Scholar
  213. Neuronal–Glial Glucose Oxidation and Glutamatergic–GABAergic Function
    Go to citation Crossref Google ScholarPub Med
  214. Fuelling Cerebral Activity in Exercising Man
    Go to citation Crossref Google ScholarPub Med
  215. Cognitive dysfunction and emotional–behavioural changes in MS: The pot...
    Go to citation Crossref Google Scholar
  216. Functional Brain Imaging
    Go to citation Crossref Google Scholar
  217. Oxidative Metabolism in Cultured Rat Astroglia: Effects of Reducing th...
    Go to citation Crossref Google ScholarPub Med
  218. Lactate: The Ultimate Cerebral Oxidative Energy Substrate?
    Go to citation Crossref Google ScholarPub Med
  219. Chapter 5.4 Online glucose and lactate monitoring during physiological...
    Go to citation Crossref Google Scholar
  220. Interaction between Astrocytes and Neurons Studied using a Mathematica...
    Go to citation Crossref Google ScholarPub Med
  221. Glucose transport to the brain: A systems model
    Go to citation Crossref Google Scholar
  222. The role of lactate in brain metabolism
    Go to citation Crossref Google Scholar
  223. The Astrocyte—Neuron Lactate Shuttle: A Debated but still Valuable Hyp...
    Go to citation Crossref Google ScholarPub Med
  224. Transfer of glycogen‐derived lactate from astrocytes to axons via spec...
    Go to citation Crossref Google Scholar
  225. The central role of astrocytes in neurometabolic coupling: A decade's ...
    Go to citation Crossref Google Scholar
  226. Theoretical support for the astrocyte-neuron lactate shuttle hypothesi...
    Go to citation Crossref Google Scholar
  227. Understanding of cerebral energy metabolism by dynamic living brain sl...
    Go to citation Crossref Google Scholar
  228. Astrocytic contributions to bioenergetics of cerebral ischemia
    Go to citation Crossref Google Scholar
  229. Energy substrate requirements for survival of rat retinal cells in cul...
    Go to citation Crossref Google Scholar
  230. Neurobarrier Coupling in the Brain: A Partner of Neurovascular and Neu...
    Go to citation Crossref Google ScholarPub Med
  231. In vivo neurochemical monitoring and the study of behaviour
    Go to citation Crossref Google Scholar
  232. Differential Effects of Glucose and Lactate on Glucosensing Neurons in...
    Go to citation Crossref Google Scholar
  233. Ex vivo NMR study of lactate metabolism in rat brain under various dep...
    Go to citation Crossref Google Scholar
  234. Lactate transport and transporters: General principles and functional ...
    Go to citation Crossref Google Scholar
  235. On-line monitoring of striatum glucose and lactate in the endothelin-1...
    Go to citation Crossref Google Scholar
  236. The Astrocyte-Neuron Lactate Shuttle: A Challenge of a Challenge
    Go to citation Crossref Google ScholarPub Med
  237. The Role of Hypotheses in Current Research, Illustrated by Hypotheses ...
    Go to citation Crossref Google ScholarPub Med
  238. Ex Vivo Analysis of Lactate and Glucose Metabolism in the Rat Brain un...
    Go to citation Crossref Google Scholar
  239. Selective Uptake of [14C]2-Deoxyglucose by Neurons and Astrocytes: Hig...
    Go to citation Crossref Google ScholarPub Med
  240. Lactate muscles its way into consciousness: fueling brain activation
    Go to citation Crossref Google Scholar
  241. Whole‐brain glutamate metabolism evaluated by steady‐state kinetics us...
    Go to citation Crossref Google Scholar
  242. Nutrition during brain activation: does cell-to-cell lactate shuttling...
    Go to citation Crossref Google Scholar
  243. Brain glycogen re‐awakened
    Go to citation Crossref Google Scholar
  244. Cerebral glucose metabolism in diabetes mellitus
    Go to citation Crossref Google Scholar
  245. Hyperoxia in head injury
    Go to citation Crossref Google Scholar
  246. Glial-neuronal interactions and brain energy metabolism
    Go to citation Crossref Google Scholar
  247. Food for Thought: Challenging the Dogmas
    Go to citation Crossref Google ScholarPub Med
  248. Astrocyte Metabolism and Astrocyte-Neuron Interaction
    Go to citation Crossref Google Scholar
  249. Neuroprotection by N-methyl-D-aspartate antagonists
    Go to citation Crossref Google Scholar
  250. Mitochondrial Disease
    Go to citation Crossref Google Scholar
  251. Stoffwechselnotfälle
    Go to citation Crossref Google Scholar

Figures and tables

Figures & Media

Tables

View Options

View options

PDF/ePub

View PDF/ePub

Get access

Access options

If you have access to journal content via a personal subscription, university, library, employer or society, select from the options below:

ISCBFM members can access this journal content using society membership credentials.

ISCBFM members can access this journal content using society membership credentials.


Alternatively, view purchase options below:

Purchase 24 hour online access to view and download content.

Access journal content via a DeepDyve subscription or find out more about this option.