CHAPTER 343 Cerebral Blood Flow and Metabolism and Cerebral Ischemia
Cerebral Metabolism
Overview
Although many aspects of human cerebral metabolism are common to the metabolism of other tissues and organs in the body, there are a few fundamental differences. First, the brain is an unusual organ in having the highest energy requirement by mass. Even though it constitutes less than 2% of body weight, the adult brain receives 25% of cardiac output at rest and uses 20% of the total energy produced by the body.1 In children, the figures are even more impressive, with up to 50% of the energy consumption of the body being accounted for by the brain. Much of this energy allocation is devoted to activities connected to neural signaling, most of all (>50%) to the work of adenosine triphosphate (ATP)-driven ion pumps, particularly sodium-potassium adenosine triphosphatase (Na+,K+-ATPase), which maintains and restores the transmembrane Na+/K+ gradients that are repeatedly diminished by the propagation of action potentials and synaptic transmission.2,3 Other signaling-related costs pertain to neurotransmitter synthesis and reuptake from the synaptic cleft and axoplasmic transport. The remainder of energy expenditure is due to so-called housekeeping activities, such as the synthesis of molecules for general cellular purposes.4 These energy demands necessitate that the brain have reliable mechanisms to adequately protect its supply of oxygen and glucose from blood and ensure that it is tightly matched with demand (i.e., the level of neural activity). Second, the metabolism of the brain is distinguished by the singular contribution of astrocytes. Third, the brain possesses a BBB. Fourth, the brain exhibits highly developed metabolic compartmentalization, which refers to the fact that astrocytes and neuronal cells are so metabolically specialized that certain substrates and synthesized products are restricted to a particular cell type or “compartment” even though they may be required for the function of another. As detailed later, this feature necessitates close interaction and trafficking of molecules between cells of different type.5–8 More is said on this interdependence later.
Cerebral Metabolic Rate
The global cerebral metabolic rate (CMR) is conventionally expressed in terms of the consumption of glucose (CMRGlc) or oxygen (CMRO2), which respectively measures 25 to 30 µmol/100 g per minute and 130 to 180 µmol/100 g per minute in a resting human adult.9,10 CMRO2 is considered a function of mitochondrial activity and can be calculated from CBF and the arteriovenous oxygen content difference. Notably, CMRGlc is considerably higher during the first few years of life because of rapid brain growth and myelination, with a gradual reduction to adult levels by the end of the second decade.11 Good correlation between CBF, CMRO2, and CMRGlc is seen during rest, with the ratio between CMRO2 and CMRGlc being maintained at around 6 : 1. However, with neural activation, this relationship is altered.
The metabolism of the brain exhibits considerable variation on multiple levels: by region, activation state, cell type, and subcellular location. Regional variance is reflected in the local metabolic rates for oxygen and glucose (LCMRO2 and LCMRGlc), as well as in the levels of some metabolic enzymes. On the whole, a wider and much higher range of LCMRGlc is exhibited by cerebral gray matter than by white matter, with the highest values recorded in physiologically more active areas such as the auditory cortex, inferior colliculus, and somatosensory cortex.12,13 Consistent with this disparity, gray matter exceeds white matter in the activity of the key enzyme of mitochondrial energy metabolism, cytochrome oxidase.14,15 At the cellular level, neuronal energy needs are estimated to greatly exceed that of glial cells, even though neurons are significantly outnumbered by glia.2 The higher metabolic activity of neurons relative to glial cells is suggested by their higher mitochondrial density and expression of cytochrome oxidase,16 as well as by their heightened sensitivity to hypoxic or hypoglycemic injury. Within the neuron itself, there is further evidence of heterogeneity of metabolic activity, with dendrites and synaptic terminals having higher cytochrome oxidase activity than cell bodies and axons.16
Energy Capture and Transfer
ATP is the principal energy currency of all living cells, including neurons and glial cells. The energy of catabolic processes is captured in the two high-energy phosphate bonds of ATP, largely through the process of mitochondrial oxidative phosphorylation (Fig. 343-1). Only minor contributions derive from substrate-level phosphorylation in the glycolytic pathway and the tricarboxylic acid (TCA; also known as the citric acid or Kreb’s) cycle. In contrast, ATP utilization occurs mostly in the cytosol through cleavage of the terminal orthophosphate group of ATP by hydrolytic enzymes, collectively referred to as ATP hydrolases or “ATPases” for short, to liberate adenosine diphosphate (ADP). ADP can then undergo similar breakdown to release adenosine monophosphate (AMP). ATP can also be directly hydrolyzed to AMP and inorganic pyrophosphate (PPi). All these hydrolytic reactions are energy producing, or exergonic, and are coupled to many energy-requiring, or endergonic, reactions that could not otherwise proceed because of unfavorable thermodynamics. The high energy demand in the brain results in fast cycling between ATP, ADP, and inorganic phosphate (Pi), which requires fast transport between the mitochondria and cytosol. When ATP utilization is rapid or ATP generation is inhibited, the brain is assisted in the maintenance of ATP supply by the following buffer systems:
Reaction 1, catalyzed by creatine kinase, yields ATP from phosphocreatine (PCr), a storage form of high-energy phosphate. PCr also functions as a shuttle for the transfer of high-energy phosphate from mitochondria to the cytosol. The importance of PCr in energy homeostasis is underscored by the fact that the total creatine pool (as Cr and PCr) in the brain is at least 3 times larger than the adenosine nucleotide pool (AMP, ADP, and ATP). Reaction 2, catalyzed by adenylate kinase, re-creates ATP from ADP. Although other nucleoside triphosphates, such as guanosine triphosphate (GTP), cytidine triphosphate, and uridine triphosphate (UTP), serve functions similar to those of ATP, their regeneration from the corresponding nucleoside diphosphates occurs at the expense of ATP.17 Therefore, directly or indirectly, ATP drives all endergonic reactions of the cell.
Choice of Metabolic Substrates
Although the brain harnesses energy from a variety of substrates, it relies predominantly on glucose. Indeed, measurement of cerebral arteriovenous levels of a range of substrates and their metabolic products in mature humans under normal steady-state conditions has established that glucose is the only energy substrate that is taken up by endothelial cells of the BBB in more than trivial amounts.18 The virtual parity of the respiratory quotient under normal conditions demonstrates the overwhelming but not complete consumption of this glucose by oxidation. The brain also obtains some energy from the metabolism of other substrates such as amino acids and endogenously produced carbohydrates and lactate. Recent studies have even suggested that glial-derived lactate may actually be the preferred fuel of neurons,19,20 but on a whole-organ basis, no single metabolic process apart from glucose oxidation has the yield to support the intense activity of the brain for more than very brief periods. Therefore, with the brain possessing just low stores of glucose,8 moderate to severe hypoglycemia can result in rapid deterioration in the level of consciousness and essential cerebral functions.21 A notable exception occurs in fasting individuals and nursing babies, in whom the imposition of hypoglycemia and high fat metabolism can lead to the dominant use of ketone bodies by the brain for cellular fuel. In this regard, it is interesting that the rate of transport of ketone bodies across the BBB is the limiting step in terms of their cerebral metabolism.22,23 However, even though the oxidation of ketone bodies may provide up to 75% of the total cerebral energy supply,24 it is unable to serve as a complete replacement for the oxidation of glucose.
Metabolism of Glucose
Glucose and its metabolites occupy key positions at the intersection point of a number of essential catabolic and anabolic pathways. As discussed earlier, glucose is the main substrate for energy production, which occurs via glycolytic and TCA cycle metabolism (see Fig. 343-1). In addition, phosphorylated glucose can be condensed into glycogen to serve as the main energy reserve in the brain. Furthermore, glucose enters the synthetic pathways of three key neurotransmitters: glutamate, γ-aminobutyric acid (GABA), and acetylcholine, as well as that of a range of amino acids via TCA cycle intermediates. Finally, metabolism of glucose via the pentose phosphate pathway provides both ribose 5-phosphate for the synthesis of nucleotides and the reducing molecule nicotinamide adenine dinucleotide phosphate (NADPH) for lipid synthesis and antioxidant defense.
Glucose and Oxygen Delivery
Glucose is an extremely hydrophilic molecule, and its delivery to brain cells involves facilitated transport, which is tightly regulated in a cell- and region-specific manner by the glucose transporter proteins (GLUTs) of the solute carrier family 2 transport protein group.25,26 Many of the 13 known GLUT isoforms have been identified in brain but, in most cases, have not been recognized to have any defined role in glucose transport.27–30 Cerebral glucose uptake is predominantly mediated by GLUT1 and GLUT3. GLUT1 exists in two molecular weights of 45 and 55 kD as a result of differing extents of glycosylation. Glucose enters the brain through the 55-kD GLUT1 transporters of microvascular endothelial cells, which are variably distributed between the luminal and abluminal cell membranes, as well as a sizable intracellular pool, probably as a mechanism to modulate cerebral glucose uptake according to metabolic demands.26,31 The 45-kD GLUT1 is usually localized to glial cells, the choroid plexus, and the ependyma26 and has just limited expression in neurons, except in response to stress from conditions such as hypoxia or hypoglycemia.32,33 Neurons express the higher affinity, higher capacity GLUT3 transporter.34
At the average rate of global CBF (55 mL/100 g per minute), the oxygen extraction fraction (i.e., the proportion of oxygen extracted by the brain relative to the amount delivered to it in arterial blood) is approximately 0.5 (or 50%), as opposed to the glucose extraction fraction of about 0.1 (10%).10 Thus, the supply of glucose is typically far in excess of requirements, which leaves plenty of reserve for periods of high metabolic demand.
Energy Production from Glucose
Glycolysis is the main pathway for glucose metabolism (see Fig. 343-1) and takes place in the cytosol with the net formation of 2 equivalents of ATP and 2 equivalents of pyruvate from 1 equivalent of glucose:
Depending on the redox state of the cell, the resultant pyruvate can take one of two paths. Under normal aerobic conditions, pyruvate and the reduced form of nicotinamide adenine dinucleotide (NADH) are taken up by mitochondria, where their oxidation by the TCA cycle and respiratory chain provide for the vast bulk of ATP production (see Fig. 343-1). In the absence of oxygen, reoxidation of NADH through the respiratory chain is blocked and must instead occur by the reductive conversion of pyruvate to lactate by lactate dehydrogenase (LDH), or else glycolysis cannot continue.
Pyruvate enters the mitochondrial matrix to undergo stepwise oxidation in the TCA cycle to carbon dioxide and water, with much of the resulting release of energy made available to the mitochondrial respiratory chain as reducing equivalents. Usually, the pyruvate is converted to acetyl coenzyme A (acetyl CoA) by pyruvate dehydrogenase (see Fig. 343-1).
Acetyl CoA then enters the TCA cycle, where the first step involves its condensation with oxaloacetate to yield citrate. In the next few steps, citrate is isomerized and oxidatively decarboxylated to yield α-ketoglutarate, which in turn is oxidatively decarboxylated to form succinate. The remaining steps involve the regeneration of oxaloacetate from succinate (see Fig. 343-1). With each round of the cycle, two carbon atoms enter as acetyl CoA and two carbon atoms leave as CO2. Because an acetyl group is more reduced than CO2, the completed cycle involves four oxidation-reduction reactions that give rise to three molecules of NADH and one of reduced flavin adenine dinucleotide (FADH2). In addition, conversion of succinyl CoA to succinate involves the cleavage of an energy-rich thioester bond, which is coupled to the phosphorylation of guanosine diphosphate to form GTP. If the GTP is not used for protein synthesis or signal transduction, its γ-phosphate group can readily be transferred to ADP to form ATP.
The NADH and FADH2 from each of the preceding stages in the oxidation of glucose are used by mitochondria for the generation of ATP by oxidative phosphorylation (see Fig. 343-1). The driving force of oxidative phosphorylation is the electron transfer potential of NADH or FADH2 relative to O2. Electrons donated by NADH are passed sequentially down a “respiratory” chain of three large protein complexes embedded in the inner mitochondrial membrane, from NADH-ubiquinone (Q) reductase (complex I) to cytochrome reductase (complex III) to cytochrome-c oxidase (complex IV). Transfer of electrons between complexes I and III and between complexes III and IV is accomplished by reduced ubiquinone (QH2) and cytochrome c, respectively. In contrast to NADH, QH2 is the entry point for electrons from FADH2. At complex IV (cytochrome-c oxidase), the electrons are consumed in a reaction with oxygen, the final electron sink, to form water. This electron flow leads to translocation of H+ from the mitochondrial matrix to the intermembrane space, and the resultant concentration gradient creates a proton motive force or transmembrane electrochemical potential (Δψm) to drive an inner membrane–bound ATP synthase (complex V). In this fashion, oxidation of NADH and FADH2 theoretically drives the formation of 3 and 2 ATP molecules, respectively. On tallying up the energy yield from each stage of the metabolism of glucose, it can be seen that a molecule of glucose that undergoes complete oxidation may theoretically yield up to 38 molecules of ATP, the majority of which are contributed by oxidative phosphorylation.
In each of the stages of generation of ATP from glucose, energy supply is coupled to energy demand by enzymes whose activities are subject to feedback regulation by one or more downstream ionic or molecular species. The main site of regulation of glycolysis is at the level of phosphofructokinase (PFK), an enzyme that catalyzes the ATP-dependent addition of a phosphate group to fructose 6-phosphate early in the pathway (see Fig. 343-1). Increases in levels of ATP (a fundamental end product of glucose metabolism), citrate (the first molecule generated in the Krebs cycle), and H+ (which is produced in numerous downstream reactions) indicate a relatively robust energy supply and exert a negative feedback effect on PFK, which dampens further glycolysis. Conversely, increases in AMP, cyclic AMP (cAMP), ADP, K+, NH4+, and Pi tend to accompany a rundown of cellular energy and thereby stimulate the activity of PFK. At the stage of acetyl CoA synthesis, the pyruvate dehydrogenase complex is directly stimulated by increasing levels of its substrate pyruvate and inhibited by its end products acetyl CoA and NADH (see Fig. 343-1). It is also responsive to energy status, as shown by the [NADH]/[NAD+], [acetyl CoA]/[CoA] and [ATP]/[ADP] ratios. Increases in these ratios result in the phosphorylation and deactivation of pyruvate dehydrogenase. The rate of the TCA cycle is immediately dependent on the availability of the oxidized form of nicotinamide adenine dinucleotide (NAD+), which in turn is dependent on the availability of ADP and hence ultimately on the rate of utilization of ATP. The enzymes of the TCA cycle are also individually regulated. For example, the dehydrogenases of the TCA cycle are activated by Ca2+, which increases in concentration during the work of muscle contraction. The most important factor regulating the overall rate of oxidative phosphorylation is the level of ADP. As ADP levels rise (reflecting higher consumption or inadequate production of ATP), oxidative phosphorylation is stimulated; that is, the respiratory chain is activated by a need for ATP synthesis. Taken together, these complex and multiple regulatory mechanisms reflect the precise control of glucose metabolism in response to the prevailing cellular conditions.
Other Metabolic Fates of Glucose
Amino Acid and Neurotransmitter Synthesis
Glutamate can be converted back to α-ketoglutarate by oxidative deamination to undergo further oxidation in the TCA cycle for the purpose of energy generation35 or redirected toward the synthesis of other amino acids, GABA, fatty acids, or glutathione.
Glycerol Synthesis
New evidence indicates that glucose undergoes conversion to free glycerol within the brain. In theory, this could occur through reduction of the glycolytic intermediate dihydroxyacetone phosphate to glycerol 3-phosphate by glycerol-3-phosphate dehydrogenase, followed by dephosphorylation to glycerol by a phosphatase. The natural operation of this pathway is suggested by the presence of the necessary enzymes in the brain. Glycerol 3-phosphate has been demonstrated in astrocytes and neurons,36 as well as in the cytosol of oligodendrocytes.37 More importantly, it has recently been reported that 13C-labeled glycerol accumulates in the media of cultured astrocytes and cerebellar granule cells and in the brains of rats within 5 minutes after treatment with [U-13C] glucose.38
Storage as Glycogen
The brain converts a limited amount of glucose into glycogen to form its principal energy reserve. Although there is typically 3 to 4 times more glycogen than free glucose in the brain, it still amounts to no more than approximately 4 µmol/g, and were it to serve as the sole fuel source, it would be consumed completely in no more than a few minutes. Instead, more contemporary evidence suggests that it exists as a metabolic buffer system and is metabolized slowly, so complete turnover of brain glycogen stores normally takes 3 to 5 days.39
Glycogenesis requires the action of glycogen synthase on glucose subunits that have first undergone phosphorylation by ATP. Although both astrocytes and neurons possess the necessary enzymes, synthesis of glycogen is normally confined to astrocytes. Storage of glycogen is also almost exclusive to astrocytes and serves to augment neuronal energy requirements during periods of intense activity and pathologic shortages of glucose. Glycogen undergoes glycogenolysis by the enzyme glycogen phosphorylase to form lactate under stimulation by neurotransmitters such as norepinephrine, vasoactive intestinal polypeptide (VIP), histamine, serotonin, and certain metabolic by-products of neuronal activity such as K+ and adenosine.40 It has been hypothesized that this lactate is transferred from astrocytes to neurons for use as an energy substrate.41,42 In contrast to the metabolism of glucose, release of glucose equivalents from glycogen does not require prior “priming” with ATP.
Metabolic Contributions of Brain Structural Elements
Astrocytes
A growing body of research has resulted in a revision of the traditional view of astrocytes as a passive supporting act to neurons. The proliferation and development of astrocytes are indeed driven by trophic cues associated with neuronal activity, but the growth and survival of neurons are in turn dependent on astrocytes.43 In terms of metabolism, astrocytes form a highly active compartment that is separate from that of neurons but with which it interacts in a dynamic and essential fashion. This interaction occurs via the narrow extracellular space (ECS) and involves the shuttling of ionic and molecular metabolites and neurotransmitters such as lactate and glutamate. Despite being essentially nonexcitable, astrocytes have also been implicated in the integration of neuronal input, modulation of synaptic activity, and even long-range signaling. Neurotransmitter-evoked elevations in astrocytic calcium can trigger the release of chemical transmitters that can cause sustained modulatory actions on neighboring neurons. Astrocytes are also involved in brain water homeostasis and induction of the BBB (see review by Bennaroch5).
Before considering the influence of astrocytes on cerebral metabolism, it is relevant to consider their structural relationship to each other and to other cells in the brain. In the mammalian brain, neurons make up no more than 50% of the cerebral cortical volume and in most regions are outnumbered 10 : 1 by astrocytes.44 Each astrocyte typically defines a nonoverlapping three-dimensional domain and is polarized such that one or more astrocytic processes contact a capillary while many more are entwined within the neuropil and engage with hundreds to thousands of synapses. Where astrocytic processes abut capillaries, they are specialized into structures known as end-feet. These end-feet are so numerous that almost the entire surface of the capillaries is covered.45 The potential significance of this interposition of astrocytes between neurons and the capillary (endothelium) has been recognized since the late 19th century.44 It puts astrocytes in a special position to both sense neuronal signaling and capture glucose directly from the capillary, thereby permitting them to govern excitation-metabolism coupling. Another noteworthy feature of astrocytes is their ample connections with each other and the ECS through gap junctions and hemichannels composed mainly of connexin.43,46,47
As alluded to before, astrocytes play a leading role in the flux of glucose into neurons for energy. The glucose taken up by astrocytes may have one of two primary fates: it may be converted to lactate via astrocytic glycolysis, or it may be converted via glycogenesis to the glucose storage polymer glycogen. By contrast, in adult neurons, aerobic glycolysis results in the formation of pyruvate, not lactate, and glycogen metabolism and storage normally do not occur. That glycogen, the storage form of glucose, is located almost entirely in astrocytes is one indication of the dominant position of astrocytes in the metabolic processing of glucose. However, it should be noted that glycogenolysis in astrocytes is dictated by specific neurotransmitters and neuromodulators, thus giving neurons tight rein over this energy store.48 As elaborated on earlier, astrocytic glycolysis is also stimulated by neuronal activation, which leads to the production of lactate. It has been proposed that this lactate is secreted into the ECS and is taken up and oxidized by activated neurons as their main energy source. Notably, 1 molecule of lactate entering the TCA cycle to undergo oxidation can yield 17 ATP molecules under normoxic conditions, which is about half the energy production of aerobic glucose metabolism. Although controversial, the existence of astrocyte-neuron lactate shuttling is based on evidence. For instance, immunohistochemistry has identified a selective distribution of LDH isoforms, with neurons predominantly expressing LDH1, the form that is enriched in lactate-consuming tissues, and astrocytes expressing LDH5, the form that is enriched in lactate-producing tissues.6 Localization of the lactate monocarboxylate transporters MCT-1 and MCT-4 to astrocytes and MCT-2 to neurons demonstrates the capacity for lactate exchange between these cells.49 When primary neuronal cultures are in the presence of lactate and glucose, they preferentially consume lactate as their oxidative substrate.
The fidelity and safety of glutamate-mediated neurotransmission are dependent on very efficient uptake and modification of glutamate by astrocytes (Fig. 343-2). They keep the extracellular glutamate concentration very low by the rapid Na+-dependent GLT1 (predominant in the cortex and hippocampus) and GLAST (predominant in the cerebellum) cotransporter-mediated removal of the neurotransmitter from the synapse, thereby enabling quick termination of glutamate signaling before excitotoxic neuronal injury can occur.50 Astrocytes can handle this influx of glutamate in several different ways. They possess the necessary aminotransferases to transfer the α-amino group of glutamate to oxaloacetate or pyruvate to yield α-ketoglutarate and either aspartate or alanine. The resultant α-ketoglutarate can be oxidized for energy metabolism in the TCA cycle. Indeed, cultured astrocytes have been reported to use glutamate as an energy substrate even in the presence of glucose.51 Glutamate can also be converted into α-ketoglutarate by glutamate dehydrogenase–catalyzed oxidative deamination. However, the most common means by which astrocytes inactivate glutamate is by conversion to glutamine through the attachment of ammonium ions (see Fig. 343-2). This last reaction is endergonic and requires glutamine synthase, an enzyme limited to astrocytes, and serves the additional important function of removing ammonia.52
Being electrically “inert,” the glutamine can then be safely released to neurons for recycling into glutamate by neuronal mitochondrial glutaminase. This cycling between glutamine and glutamate is commonly referred to as the glutamine-glutamate shuttle (see Fig. 343-2). The sodium that is cotransported into astrocytes during the uptake of glutamate (glutamate/Na+ ratio, 1 : 3) stimulates the activity of Na+,K+-ATPase, which depletes intracellular ATP and in turn stimulates PFK, the principal rate-limiting enzyme of glycolysis.53 The resulting net production by glycolysis of 2 molecules each of ATP and lactate per molecule of glucose would in theory supply enough energy for astrocytic uptake (1 ATP for the glutamate transporter) and conversion (1 ATP to drive amidation) of 1 molecule of glutamate to glutamine, thus leaving lactate in surplus of requirements.54 Herein lies one possible explanation for the well-documented early rise in lactate in cerebral tissue during cortical activation, as occurs in seizures.55,56 More importantly, this hypothesized stimulation of glycolysis by glutamate provides a mechanism for the observed coupling of excitation and metabolism in the brain. Cortical activation causes rapid glutamatergic signaling, which increases the glutamine/glutamate flux and therefore drives the increased astrocytic consumption of glucose by glycolysis. This logic is supported by evidence of coupling of the glutamate/glutamine shuttle to glucose energetics in the cerebral cortex in vivo.57,58
Astrocytes are capable of long-range signaling and buffering functions through the use of their gap junctions to form a large intercellular network, or syncytium. For example, increases in the extracellular K+ concentration secondary to neuronal activity leads to the entry of K+ into astrocytes through strong inwardly rectifying K+ (Kir) channels.59,60 The resulting local depolarization is propagated electrotonically through the glial cell network via the gap junctions, which leads to the efflux of K+ at distant cell processes that are not experiencing the elevated K+ concentration. The high density of K+ channels on astrocytic end-feet allows K+ to be deposited in the perivascular space, where it can be recycled when neural activity ceases. This so-called spatial buffering of K+ is important because even modest efflux of K+ from neurons can lead to considerable changes in the concentration of K+ in the ECS, with potentially detrimental effects on maintenance of neuronal membrane potential, activation and inactivation of voltage-gated channels, synaptic transmission, and electrogenic transport of neurotransmitters. The gap junctions of astrocytes also contribute to the propagation of intercellular Ca2+ waves, probably by enhancing the release of ATP, as well as by providing an intercellular pathway. Astrocyte membrane depolarization by glutamate causes the mobilization of Ca2+ from the endoplasmic reticulum via the inositol triphosphate (IP3) generated by the activation of metabotropic glutamate receptor 5.61 It has previously been thought that IP3, Ca2+, or both are propagated across gap junctions to create a “Ca2+ wave.” More contemporary evidence suggests that the Ca2+ wave is propagated through the paracrine actions of ATP released from astrocytes on purinergic receptors.62 Release of glutamate from astrocytes has been associated with this calcium wave. Taken together, it is clear that an increase in neuronal depolarization is coupled to an increase in astrocytic depolarization, which in turn influences local metabolic and electrical activity via K+, H+, and Ca2+ ions.
Finally, yet another distinct role of astrocytes in cerebral metabolism is regulation of the development and function of the endothelial cells of the BBB. Astrocytes have important influences on the BBB, which plays a crucial role in cerebral metabolism (see later). First, contact of astrocytic foot processes with endothelial cells of the BBB upregulates the formation of tight junctions in the latter, mainly by inducing the production of occludin, an integral protein of tight junctions.63,64 Second, astrocytes can also induce the expression in endothelial cells of membrane transporters and specialized enzymes such as γ-glutamyl transpeptidase.65,66 Astrocytes submitted to hypoglycemic conditions may release factors that increase glucose uptake through the BBB.67
Blood-Brain Barrier
The BBB is a protective structure formed by capillaries to restrict the exchange of solutes between the brain and blood and thus to ensure a chemically controlled intracerebral milieu for optimal cellular performance. Morphologically, the BBB is composed of a monolayer of specialized capillary endothelial cells surrounded by a thin basal lamina and closely invested by the foot processes of astrocytes and cells known as pericytes. Unlike endothelial cells elsewhere in the body, those of the BBB are characterized by specialized regions of circumferential intercellular contact known as tight junctions or zonula occludens, minimal pinocytotic activity, and the virtual absence of fenestrations, all of which combine to necessitate that molecules passing across the BBB take a transcellular rather than a paracellular route. Consequently, free diffusion through an intact BBB is limited to lipid-soluble substances such as CO2, O2, ethanol, and lipophilic drugs or to very small polar molecules with a radius of less than 0.8 nm.68 Passage of even the smallest charged molecules (i.e., inorganic ions) is severely restricted, so transendothelial electrical resistance, which is typically 2 to 20 Ω•cm2, can be increased to greater than 1000 Ω•cm2.69 Hence, ions are transported by channels or by active means, such as via Na+,K+-ATPase.70 Transport of ions (particularly Na+) is coupled to the obligatory flow of water via osmotic forces, and both the luminal Na+ transporter and abluminal Na+,K+-ATPase are implicated in the secretion of extracellular fluid by brain capillaries.68,70 There are specific transport systems for the transcellular traffic of small lipophilic nutrient and amino acid molecules across the BBB. For example, glucose and neutral amino acids are transferred by the GLUT1 and L-amino acid transporters, respectively. Large hydrophilic proteins such as insulin, transferrin, and insulin-like growth factor may be taken up by a saturable receptor-mediated transcytosis mechanism, whereas others, particularly polycationic proteins, may cross the BBB via a nonspecific, non–receptor-mediated process referred to as adsorptive transcytosis. The BBB also has active mechanisms for excluding potentially harmful substances from the brain. Physiologically expressed carriers such as P-glycoprotein actively transport lipophilic molecules across the BBB and out of the brain. In addition, a metabolic barrier is provided by a combination of intracellular and extracellular enzymes: ectoenzymes such as peptidases and nucleotidases for the degradation of peptides and ATP, respectively, and intracellular enzymes such as monoamine oxidase and cytochrome P-450 for the inactivation of neuroactive and toxic compounds.71 Finally many, if not most of the features of the BBB are dynamic. For instance, during hypoglycemia, upregulation of GLUT1 and MCT transporter expression allows transport of glucose and ketone bodies into the brain to be increased.72
Cerebral Blood Flow
Hemodynamics
Here, r refers to vessel radius, ΔP to pressure gradient, η to the coefficient of fluid viscosity, and L to vessel length.73 In practical terms, L and η can usually be regarded as effectively constant, and the implication of the Hagen-Poiseuille law is that blood flow not only varies proportionally with the pressure gradient but also with the fourth power of the vessel radius. This provides an explanation for the clinical observation of the large change in blood flow that can occur with only small changes in vessel diameter. However, it only approximates real life because contrary to the key assumptions behind the Hagen-Poiseuille law, normal blood flow is not continuous but pulsatile, and blood vessels are not rigid and branchless tubes. In addition, if the rate of flow is continuously increased, there comes a point when resistance to flow increases sharply and the flow ceases to be laminar, instead forming a turbulent pattern. The situation in the brain is made even more complex by the operation of cerebrovascular autoregulation (discussed in detail later), the possible presence of arterial stenosis, and the diameter and extent of arterial collaterals.74 The intracranial arteries of the circle of Willis represent the naturally existing site of collateral blood flow for the cerebral circulation. It should be noted, however, that when a major cerebral artery undergoes gradual occlusion, the extracranial arteries can also provide important collateral supply to the cerebral circulation via the ophthalmic, meningeal, and leptomeningeal arteries.
Under normal physiologic conditions, blood flow is regulated in the brain through changes in vascular resistance. By combining Equations 10 and 14, the significant relationship between vessel radius r and CVR is revealed:
Indeed, the resistance of the cerebral circulation is subject to dynamic changes in the contractile state of vascular smooth muscle (VSM), most of all at the level of the penetrating precapillary arterioles; that is, the principal cerebral resistance vessels are those that arise perpendicularly from the pial arteries on the brain surface before penetrating the parenchyma. However, up to 50% of total CVR arises from smaller pial arteries (150 to 200 µm in diameter) and arteries of the circle of Willis.73,75
Hemorheology
The composition and environment of blood confer on itself complex and anomalous viscous properties.76,77 Importantly, by being composed of a concentrated suspension of cells within proteinaceous plasma, blood is a particulate fluid. Furthermore, many of these cells are capable of altering their shape and forming physical interactions with each other or the glycocalyx on the endothelial wall.78,79 From the foregoing, it is intuitive that blood viscosity is not only a function of plasma viscosity but also depends on the concentration of cells, one measure of the latter being the hematocrit. Less obvious is that such factors as the deformability and aggregability of erythrocytes and the adherence or nonadherence of leukocytes to the endothelium80–82 can lead to marked deviations from Newton’s law. However, in blood vessels, where the internal diameter is very large in comparison to the size of the cells, blood of normal hematocrit approximates Newtonian behavior reasonably well.77
The non-Newtonian behavior of blood is best demonstrated in its passage through the microcirculation. In large arteries, where the shear rate is low, calculations using Poiseuille’s equation yield an “apparent” viscosity that is much higher than expected. This has been explained by the tendency of erythrocytes to form clumps or rouleaux at low shear rates, thereby increasing resistance to flow. With the higher shear rates typically found in the microvasculature, the apparent viscosity falls because any cell aggregates are dispersed into single cells that stretch and align themselves with the axial and fastest moving laminae of the bloodstream, thus leaving a slower, cell-depleted zone of plasma along the vessel margin. This marginal zone is thought to progressively dilute the hematocrit as the caliber of the blood vessel is reduced and may be accentuated by plasma skimming or cell screening at branching points. As a result, when blood flows through progressively diminishing arterioles or capillaries with a diameter of 300 µm or smaller, there is a linear fall in apparent viscosity known as the Fahraeus-Lindqvist effect. Eventually, however, when blood reaches capillaries with a diameter that is less than that of an erythrocyte (6 to 8 µm), a steep rise in blood viscosity and an inversion phenomenon (i.e., reversal of the Fahraeus-Lindqvist effect) take place.83,84 The anomaly represented by the Fahraeus-Lindqvist effect is of critical importance in counteracting the adverse influence of blood vessel geometry (number, length, and diameter) on resistance to microcirculatory blood flow.
Relationship between Cerebral Blood Flow and Intracranial Pressure
CBF and ICP are related by the Monro-Kellie doctrine, the modern version of which states that because the intracranial space is fixed and its principal components, the brain parenchyma, blood, and CSF, are nearly incompressible, any change in the volume of one component must lead to a reciprocal change in volume of the other components or ICP must rise. Cerebral blood volume (CBV) takes up a significant proportion of intracranial volume. At any point in time, changes in CBV are determined by changes in arterial inflow relative to venous drainage and are therefore reliant on CBF. If intracranial compliance is low, a rise in CBF can increase ICP. However, as ICP rises, there must be a compensatory rise in MAP or CPP will fall (see Equations 11 and 12), with a deleterious effect on CBF.
Regulation of Cerebral Blood Flow
Unlike other organs, regulation of blood flow in the brain is distinguished by the influence of astrocytes and neurons. Extracerebral blood vessels receive a rich “extrinsic” supply of perivascular fibers from the parasympathetic (mainly the sphenopalatine, otic, and internal carotid) and sympathetic (mainly from the superior cervical) ganglia, as well as the sensory roots of the trigeminal ganglia. On entering the brain parenchyma, cerebral arteries lose this ganglionic nerve supply and instead acquire “intrinsic” innervation from parenchymal neurons. The best characterized intrinsic neural pathways that project to cortical blood vessels are those from the nucleus basalis, locus caeruleus, and raphe nucleus. With electrical or chemical stimulation of these areas, increases or decreases in cortical CBF occur. Anatomic studies have shown that neurons in these areas send projection fibers to cortical blood vessels, as well as to astrocytes. In fact, noradrenergic afferents from the locus caeruleus target mainly perivascular astrocytes. Changes in perivascular astrocytic [Ca2+]i secondary to noradrenaline cause vasoconstriction of the adjacent arterioles.85 Hence, arterial tone is influenced by astrocytes, as well as by neurons.
Major Mediators
Nitric Oxide
Nitric oxide (NO) is a ubiquitous cellular messenger that was referred to as endothelium-derived relaxing factor (EDRF) on its discovery and remains known as one of the principal mediators of vasodilation. With regard to this vasodilator function, NO is the mediator for the action of mechanical stimuli and a large variety of agonists on the endothelium. The latter include acetylcholine, bradykinin, oxytocin, histamine, endothelin-1 (ET-1 via ETB receptors; see later), and prostaglandin F2α (PGF2α). Apart from vascular actions, NO signaling has been suggested or established in a wide variety of other physiologic functions, including neurotransmission, behavioral inhibition, prevention of platelet and neutrophil aggregation, promotion of gastrointestinal motility, and the immune response. Moreover, alterations in the NO system are implicated in the pathogenesis of many diseases as diverse as Alzheimer’s disease, cerebral vasospasm, pyloric stenosis, and nephrogenic diabetes insipidus.86–88 In terms of its vascular actions, which involve extensive crosstalk with a variety of other vasoactive systems, it is clear that the NO signaling system is a major mediator and possibly part of a “final common pathway” of vascular modulation.
Being a short-lived and freely diffusible gas, NO cannot be stored and must be synthesized close to its site of action by nitric oxide synthase (NOS) from L-arginine. There are actually three isoforms of NOS, each encoded by distinct and highly conserved genes: neuronal/type I (nNOS), inducible/type II (iNOS), and endothelial/type III (eNOS). The endothelial and neuronal isoforms are constitutively expressed in cerebral blood vessels, the latter occurring in the nerve plexus in the outer adventitial layer.89–91 Recently, constitutive expression of a posttranslationally modified variant nNOS isoform, nNOS-α or mtNOS, has been identified in the mitochondria (mt) of brain and other organs.92 By contrast, expression of iNOS is induced in cells, including macrophages, endothelial cells, and neurons, under specific conditions such as inflammation, trauma, and infection86 and appears to have no role in normal cerebrovascular function. The enzyme is active when dimerized. The catalytic reaction requires molecular oxygen and NADPH in addition to the cofactors of FAD, flavin mononucleotide, iron protoporphyrin IX, and tetrahydrobiopterin (BH4). Once formed, NO diffuses within or between cells, thus permitting it to function as both an autocrine and paracrine messenger. In its most common redox state, NO possesses an unpaired electron, which renders it highly reactive, particularly with metalloproteins containing iron or thiol groups. Indeed, much of NO signaling relies on avid binding of NO to the heme moiety of soluble guanylate cyclase, thereby leading to activation of the enzyme, followed by increased levels of cyclic guanosine monophosphate (cGMP). In VSM cells, cGMP in turn activates protein kinase G, which causes relaxation by opening K+ channels or decreasing the sensitivity of the contractile machinery to Ca2+.
NO has an important role in the regulation of cerebrovascular tone under normal, basal conditions, as shown by the constrictive effects of nonspecific NOS inhibitors on resting cerebral arteries both in vitro and in vivo, accompanied in the latter case by a reduction in CBF.93–96 The individual blockade of eNOS and nNOS causes a decrease in vascular tone, thus suggesting a tonic influence of both isoforms.97–99 This effect is confirmed in the former case by the phenotype of modest systemic hypertension in mice with targeted deletion of eNOS.100 This may not merely be due to a simple loss of vasodilatory influence inasmuch as eNOS-null animals demonstrate cerebral arteriolar hypertrophy that has been attributed to the lost inhibition of VSM proliferation,101 thus echoing previous hypotheses that eNOS-derived NO is a negative regulator of vascular remodeling.102,103 In contrast, tonic input is supplied to forebrain arteries by the nitroxidergic parasympathetic postganglionic fibers arising from the pterygopalatine ganglion.104 This is evidenced by ipsilateral cerebral vasoconstriction on denervation or ablation of this ganglion or its preganglionic afferents in dogs and monkeys.105–107
By and large, evidence suggests that the NO synthesized by nNOS is one of the mediators in the coupling of CBF to synaptic activity. Selective nNOS inhibitors such as 7-nitroindazole have been shown in animal models to inhibit increases in CBF induced by neuronal activity.108–111 Curiously, nNOS-null mice have an intact hyperemic response to hypercapnia112 and neural activation108,113,114 in the cerebral cortex but not in the cerebellum.115 However, unlike their wild-type brethren, the cerebral cortical hyperemia in mutant nNOS mice is not altered by pharmacologic NO inhibition, which provides a clue that normalcy can be maintained, at least in the supratentorial compartment, by the compensatory activity of NO-independent pathways. In light of the evolutionary conservation of NOS genes, such redundancy is noteworthy.
Regulation of NOS isoforms in the context of the cerebral circulation is extremely complex and occurs at multiple levels.116 In general, transcriptional regulation governs cell- and tissue-specific expression of all isoforms.117 In addition, eNOS transcription can be upregulated by shear stress on endothelial cells as a result of the presence in the promoter region of a shear stress responsive element.118,119 The occurrence of posttranscriptional control mechanisms, such as alterations in messenger RNA splicing and stability, is well described.120 Apart from the availability of substrate and cofactors, the activity of the enzyme itself is regulated by a variety of protein-protein interactions, as well as by phosphorylation. Binding of calmodulin is essential for the electron transfer function of all NOS isoforms,121 but eNOS and nNOS do so in a manner that is reversible and dependent on [Ca2+]i, whereas iNOS does so avidly and in an essentially Ca2+-independent manner.122 Ca2+-calmodulin has the additional important function of competing with caveolins in the allosteric modulation of eNOS and nNOS activity. Another apparent regulator of eNOS, with which it colocalizes, is the 90-kD heat shock protein (Hsp90).123 Better known as a molecular chaperone, Hsp90 enhances the activation of eNOS, possibly by facilitating the competitive displacement of caveolin-1 from eNOS by Ca2+-calmodulin.124,125 As eNOS is regulated by phosphorylation on serine and, in specific situations, on tyrosine and threonine residues,126,127