Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Spreading depression (SD) is a transient wave of near-complete neuronal and glial depolarization associated with massive transmembrane ionic and water shifts. It is evolutionarily conserved in the central nervous systems of a wide variety of species from locust to human. The depolarization spreads slowly at a rate of only millimeters per minute by way of grey matter contiguity, irrespective of functional or vascular divisions, and lasts up to a minute in otherwise normal tissue. As such, SD is a radically different breed of electrophysiological activity compared with everyday neural activity, such as action potentials and synaptic transmission. Seventy years after its discovery by Leão, the mechanisms of SD and its profound metabolic and hemodynamic effects are still debated. What we did learn of consequence, however, is that SD plays a central role in the pathophysiology of a number of diseases including migraine, ischemic stroke, intracranial hemorrhage, and traumatic brain injury. An intriguing overlap among them is that they are all neurovascular disorders. Therefore, the interplay between neurons and vascular elements is critical for our understanding of the impact of this homeostatic breakdown in patients. The challenges of translating experimental data into human pathophysiology notwithstanding, this review provides a detailed account of bidirectional interactions between brain parenchyma and the cerebral vasculature during SD and puts this in the context of neurovascular diseases.

Free full text 


Logo of physrevLink to Publisher's site
Physiol Rev. 2015 Jul; 95(3): 953–993.
Published online 2015 Jul 1. https://doi.org/10.1152/physrev.00027.2014
PMCID: PMC4491545
PMID: 26133935

Spreading Depression, Spreading Depolarizations, and the Cerebral Vasculature

Abstract

Spreading depression (SD) is a transient wave of near-complete neuronal and glial depolarization associated with massive transmembrane ionic and water shifts. It is evolutionarily conserved in the central nervous systems of a wide variety of species from locust to human. The depolarization spreads slowly at a rate of only millimeters per minute by way of grey matter contiguity, irrespective of functional or vascular divisions, and lasts up to a minute in otherwise normal tissue. As such, SD is a radically different breed of electrophysiological activity compared with everyday neural activity, such as action potentials and synaptic transmission. Seventy years after its discovery by Leão, the mechanisms of SD and its profound metabolic and hemodynamic effects are still debated. What we did learn of consequence, however, is that SD plays a central role in the pathophysiology of a number of diseases including migraine, ischemic stroke, intracranial hemorrhage, and traumatic brain injury. An intriguing overlap among them is that they are all neurovascular disorders. Therefore, the interplay between neurons and vascular elements is critical for our understanding of the impact of this homeostatic breakdown in patients. The challenges of translating experimental data into human pathophysiology notwithstanding, this review provides a detailed account of bidirectional interactions between brain parenchyma and the cerebral vasculature during SD and puts this in the context of neurovascular diseases.

I. INTRODUCTION

In 1944, Aristides A. P. Leão reported a peculiar electrophysiological observation that he serendipitously discovered while studying seizure activity in rabbit cortex (Figure 1A). He observed that a sufficiently intense but not necessarily injurious focal electrical or mechanical stimulus triggered a wave of electrocorticographic silence that slowly propagated across the cortex (270). This spreading depression (SD) of cortical activity, which has since been reproduced in a wide variety of species from mammals (Figure 1B), reptiles, and birds to locusts and cephalopods, required grey matter contiguity, as preexisting lesions or cuts prevented the wave propagation. The cerebrovascular response to SD was conspicuous enough for Leão to dedicate the succeeding paper in the same issue to describe the profound pial vascular changes accompanying the SD wave (269). He observed arterial dilation accompanied by a red shift (i.e., arterialization) in the color of venous blood suggesting luxury oxygenation and perfusion. Following these two landmark papers, research on SD gained steady momentum, fueled by technological advances as well as the recognition of its role in a number of human diseases (259).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390001.jpg

Electrophysiological, ionic, and optical changes during SD. A: although possibly observed by others in one form or another as early as 1930s (99, 274), Leão (270) was the first to recognize, systematically investigate, and describe in detail the hallmarks of SD. In his 1944 paper, he demonstrated the suppression of electrocorticogram (ECoG) activity that spreads in contiguous grey matter in a relatively concentric manner at an approximate rate of 3 mm/min using serially placed surface electrodes on rabbit cortex as shown in this figure. Recovery follows the same order within 10 min. [Modified from Leão (270).] B: typical electrophysiological tracings showing ECoG depression spatiotemporally coincident with a large negative extracellular direct coupled (DC) potential shift that lasts up to a minute when recorded by two serial intracortical microelectrodes in the rat as shown in the inset. [Modified from Ayata (15) with permission.] C: sample ECoG and DC potential tracings from human brain showing SDs in the setting of traumatic brain injury recorded by subdural electrode strips. ECoG traces are bipolar between adjacent channels on the strip. DC traces are relative to an extracranial reference. Note the sequential involvement of electrode pairs by SD. (Courtesy of Dr. Jed Hartings.) D: representative time-lapse images of intrinsic optical signal changes taken every 6.4 s from rat cortex show relative reflectance (upper two rows) and rate of change in reflectance (lower two rows) during pinprick-induced SD expanding in a concentric fashion regardless of functional or vascular territories. [Modified from Chen et al. (56) with permission.] E: the transmembrane ionic shifts during SD lead to a large drop in [Na+]o, [Ca2+]o, [Cl]o, and pH, and a surge in [K+]o simultaneously with the extracellular DC potential shift (V), as shown by ion-selective microelectrodes (147, 339, 477). [Modified from Hansen and Zeuthen (180) with permission.] F: time-lapse images taken every 2 s show retinal SD changing the optical properties of isolated chick retinal cup, seen as a concentric pallor spreading from the point of induction (arrowhead). [Modified from Hanke and de Lima (173) with permission.]

For many years, SD was perceived as a curiosity, an experimental phenomenon in animals. Over the past decade the perception changed. Today, ample evidence from advanced imaging technologies and landmark studies by the Cooperative Study on Brain Injury Depolarizations (COSBID; www.cosbid.org) indicate that SD does occur in human brain in ischemic or hemorrhagic stroke, and traumatic injury (Figure 1C), and perhaps in other less common conditions such as transient global amnesia (79, 93, 115, 183, 451). There is strong albeit indirect evidence suggesting that SD is the electrophysiological substrate of migraine aura as well (13). Moreover, data also indicate that SD is not an innocent bystander or an epiphenomenon, but does indeed worsen injury outcomes by vascular and metabolic mechanisms. As such, SD is now a diagnostic and therapeutic target in neurovascular diseases. Numerous studies have explored the bidirectional interactions between parenchymal depolarization during SD and the cerebral vasculature. However, to put the metabolic and vascular effects of SD in proper context, we will start with a brief overview of its electrophysiological properties and the underlying transmembrane ionic and water fluxes.

II. NEUROPHYSIOLOGICAL FEATURES OF SPREADING DEPRESSION

SD is a slowly propagating wave of intense but transient regional depolarization of most neurons and glia (indeed perhaps all cells), lasting up to a minute or more in otherwise normal tissue (40, 63, 455). Because complete membrane depolarization precludes action potentials or synaptic transmission, SD is accompanied by suppression of all spontaneous or evoked electrical activity in that region (Figure 1B), thus prompting Leão to coin the term depression (268-271). In terms of membrane potential, however, the term depression is somewhat of a misnomer as the underlying electrophysiological process is a near-complete and prolonged depolarization (i.e., strong excitation), a fact that should be kept in mind when interpreting its metabolic and vascular concomitants.

As the name implies, the depolarization spreads at a rate of only a few millimeters in a minute, often centrifugally from the point of origin by way of grey matter contiguity, regardless of functional divisions or vascular territories (Figure 1D), although large cerebral arteries can sometimes act as barriers (137). Interruption of grey matter by preexisting lesions blocks the spread, and high astrocyte density slows the propagation (137). By the same token, SD cannot penetrate white matter, and therefore, intervening white matter bundles form barriers to SD propagation. Contiguous spread in grey matter is independent of action potentials or physiological synaptic transmission and, therefore, of long-range axonal connections. Instead, it is dependent on depolarization-induced massive transmembrane ionic and water shifts that locally flood neighboring tissues with depolarizing ions and neurotransmitters that trigger the same depolarization cycle in adjacent cells. This slow, regenerative propagation of the depolarization is primarily driven by neurons and passively followed by astrocytes, since astrocytic [Ca2+]i waves temporally lag behind neuronal changes (59, 367).

SD is triggered when a sufficiently strong stimulus simultaneously depolarizes a minimum critical volume of brain tissue estimated to be ~1 mm3 in rodent cortex, in vivo (295), although the critical volume is smaller in cortical slices in vitro (471). The depolarizing stimulus overloads the extracellular K+ ([K+]o) clearance mechanisms causing [K+]o to exceed a critical threshold concentration of ~12 mM (162, 180, 190, 280, 295). These thresholds can vary in different species and brain regions depending on neuronal and excitatory synaptic density, and with age, among other factors (43, 294). The inciting event causes a sudden drop in membrane resistance via opening of nonselective large-conductance cation channels, the presence and identity of which are yet incompletely understood. As a result, both intracellular and extracellular ions move along their transmembrane concentration gradients (Figure 1E). Massive K+ efflux raises [K+]o from ~3 mM at resting state (219, 378, 507) to over ~30–50 mM, and sometimes as high as 80 mM, in an all-or-none fashion, in most species, tissues, and model systems (39, 179, 180, 232, 280, 342, 507, 518). This large K+ efflux is reciprocated by Na+ and Cl influx that pulls water, causing cell swelling (134, 180, 341, 495, 497499, 518). Production of intracellular osmolutes presumably contributes to intracellular water accumulation and cell swelling. Extracellular space shrinks by more than 50%, which accentuates extracellular ion concentrations (178, 308, 370, 518). Depolarization also triggers Ca2+ influx and a more than 10-fold drop in [Ca2+]o (146, 180, 340), which, along with Na+ and water influx, leads to release of many if not all neurotransmitters and neuromodulators within the depolarized tissue. Extracellular glutamate, aspartate, glycine, GABA, and taurine concentrations increase during SD (117, 232, 326), and similar increases have been shown for adenosine, catecholamine, and ascorbate levels (277, 278, 325, 391, 465). The massive rise in [K+]o to levels sufficient to depolarize neighboring cells is the critical factor mediating the contiguous spread of the wave (i.e., reaction-diffusion model of SD propagation; Figure 2) (161, 162, 180). Elevated extracellular levels of the strongly depolarizing excitatory amino acid glutamate further fuels SD and facilitates its propagation by activating NMDA receptors (289, 324).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390002.jpg

Reaction-diffusion model of SD propagation and regenerative processes. Basic mechanisms believed to be responsible for the self-regenerating propagation of SD within contiguous grey matter and the mechanisms of recovery.

The massive redistribution of ions, water, and neurotransmitters is self-limited. A number of mechanisms, including the Na+-K+-ATPase, intracellular buffering of [Ca2+]i, reuptake and metabolism or spatial buffering by the astrocytic network, and quite likely vascular clearance, all help restore the homeostasis usually within a minute. The process is in part energy-dependent and strongly stimulates O2 and glucose consumption. Therefore, in severely energy-compromised tissue, restoration of homeostasis is understandably delayed, with deleterious consequences on tissue viability.

Neurons are like batteries. They store a substantial charge across their polarized cell membranes. In contrast to the highly regulated and brief currents comprising action potentials and synaptic transmission, SD is analogous to shorting out the battery, discharging the entire membrane potential of all cells in a narrow (2–3 mm) strip of brain tissue simultaneously. SD may indeed generate heat in the tissue (250, 251, 473, 489), although changes in blood flow complicate the measurements and interpretation (1, 188). When recorded extracellularly, this massive discharge during SD creates a signature potential shift that can reach up to −30 mV or more in amplitude. Owing to its slow onset and resolution (many seconds), this potential shift is often not detected using traditional high-pass AC amplifiers, and requires unfiltered direct-coupled (DC) amplification; because of this, it has classically been referred to as the DC shift.

These properties distinguish SD from normal electrophysiological activity in the brain. For example, an action potential lasts milliseconds, generates an extracellular potential shift that is microvolts in amplitude, and can propagate meters in 1 s. In contrast, SD lasts up to a minute, generates an extracellular potential shift on the order of millivolts, and propagates only a few millimeters in a minute. It is easily evoked in vertebrates (e.g., mammalians, birds, reptiles, amphibians, fish) as well as invertebrates (e.g., locust, cockroach, cephalopods) (43, 266, 290, 302, 341, 388, 391, 393, 399, 530, 531). In human brain, SD was first detected electrophysiologically upon stereotactic injection of concentrated KCl in caudate and hippocampus in neurosurgical patients (448), and then spontaneously in cortex in a brain-injured patient (302). In vertebrates, cortex, striatum, thalamus, cerebellum, brain stem, spinal cord, and retina (Figure 1F) can all sustain SD, albeit with varying degrees of susceptibility. Because in vitro preparations (e.g., retinal cup, brain slices) can readily undergo SD, perfused vasculature is not required for SD generation, propagation, or recovery.

Experimentally, SD can be evoked when a sufficiently strong stimulus raises [K+]o above a threshold of ~12 mM in a minimum critical volume of grey matter, as noted earlier. This can be achieved chemically by direct application of depolarizing substances, such as concentrated KCl, glutamate receptor agonists particularly of NMDA subtype (496), and Ca2+ or Na+ channel openers (e.g., BAY K 8644, aconitine); electrically by direct cathodal stimulation of tissue; and by mechanical stimulation (e.g., pinprick, focused ultrasound or laser) (230, 381, 407). Other chemical stimuli can also precipitate SD, such as Na+-K+-ATPase inhibitors (e.g., ouabain) (21, 76, 292, 449), membrane-impermeable Cl substitutes (293), and metabolic toxins (e.g., cyanide, azide, 2,4-dinitrophenol, fluoride, fluoroacetate, iodoacetate) (41, 49, 145). Indeed, hypoxia and ischemia are potent inducers of an SD-like event termed anoxic depolarization (AD) (441). Moreover, spontaneous and repetitive peri-infarct or peri-injury depolarizations (PIDs) that are often indistinguishable from SD occur in focal ischemic brain. Such injury-induced spreading depolarization waves have been detected in human ischemic or hemorrhagic stroke and trauma (79, 93, 115, 183). It should be noted that in the context of stroke and brain injury, usage of the term SD has become problematic, as it requires depression of spontaneous electrophysiological (i.e., electrocorticographic) activity. Because the latter is often already absent in peri-infarct or injured brain, its suppression cannot define or be relied upon to detect injury depolarizations (IDs). The arguably more accurate and definitely more useful term spreading depolarization has aptly been coined as an alternative to describe both (83). For historical reasons, and because the distinction between spreading depressions and spreading depolarizations is less clear in normal brain, we refer to both as SD in this article.

A large number of physiological (e.g., pH, temperature, fasting, hormones), pharmacological (e.g., Ca2+ channel blockers, K+ channel openers, σ-receptor inhibitors, nitric oxide synthase inhibitors) and genetic (e.g., mutations in Cav2.1 P/Q type Ca2+ channel, Notch3 receptor) modulators of susceptibility to and electrophysiological properties of SD have been described (37, 103, 105, 369, 482), while other modulators of neuronal activity and synaptic transmission have been ineffective (e.g., GABAA receptor agonists, voltage-gated Na+ channel blockers, AMPA/KA subtype of glutamate receptor antagonists), as reviewed in detail elsewhere (15).

Despite the relatively constant electrophysiological features of SD across a wide range of tissues and species, its vascular effects appear to be highly variable, depending on the species (e.g., rats versus mice), brain region, vascular bed (e.g., large arteries versus precapillary arterioles and capillaries), and experimental conditions (e.g., anesthesia, systemic physiology). However, to put the hemodynamic response in proper context, we shall first review the metabolic consequences of SD in normal brain.

III. METABOLIC IMPACT OF SPREADING DEPRESSION

As early as 1950s, it was recognized that SD precipitates marked metabolic changes in brain tissue (Figure 3). First and foremost, ATP consumption is stimulated, presumably as a result of activation of Na+-K+-ATPase and other ATP-dependent pumps to restore ionic gradients across neuronal and glial membranes, and to recycle and replenish the neurotransmitters after their release during the depolarization. Although some studies have failed to detect a significant change in ATP levels during SD (42, 43, 226, 237, 238, 262), others have convincingly shown a decrease by as much as 50% (143, 321, 416), along with increased ADP and AMP (262). Adenosine levels are also elevated and correlate with the metabolic burden and mismatch imposed on the tissue by SD (278). The onset of these changes approximately coincides with the DC shift onset (143, 321). Recovery starts at peak DC negativity and prior to hyperemia and is complete within 1–2 min in otherwise normal tissue (321). The ATP/ADP systems are linked through a rapid equilibration in the creatine phosphokinase reaction (110). As a result, creatine phosphate decreases by up to 80% during the peak DC negativity and recovers within 1–2 min (143, 238, 262, 412, 416). Despite these changes in high-energy phosphates, the total energy charge does not appear to be diminished during and after SD (262). These results suggest that while a great turnover takes place, otherwise healthy brain tissue can cope with the challenge.

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390003.jpg

Metabolic impact of SD. A summary diagram showing changes in metabolic indexes as a function of time after SD onset. Blue scale indicates the magnitude of change (darker is bigger change). Some degree of uncertainty exists in the literature in terms of the precise timing and magnitude of metabolic changes. Therefore, we included in the diagram only those indices with reproducible changes reported in the literature. Notably, temporal changes in Po2, CMRO2, and NADH have been complex and highly variable among studies, as described in detail in the text, possibly related to differences such as measurement technique, species, experimental conditions, and the hemodynamic response. The latter is an important determinant of supply-demand mismatch and supply limitation.

The restoration of ionic gradients after SD is a challenge for the local energy homeostasis, and the rise in metabolism is among the largest observed in the living brain. Changes in high-energy phosphates are accompanied by a rapid stimulation of glucose consumption. Increased cerebral metabolic rate of glucose (CMRGlu) is detectable as early as the DC shift onset, reaches a peak at repolarization, and returns to normal levels after about 15 min (150, 258). Multiple consecutive SDs can indeed double or triple CMRGlu in broad areas of cortex (163, 226, 320, 332, 433). As a result, tissue glucose decreases by 30–60% and lactate increases by more than twofold at or within a minute of peak depolarization (68, 69, 143, 150, 226, 237, 238, 262, 321, 329, 394, 411, 416). Cumulative changes were detected upon repeated SDs using microdialysis (186). Tissue pH decreases by up to 0.5 units depending on the recording technique and the tissue and species under study (69, 143, 231, 329, 380, 411, 461, 487). In fact, tissue acidification during SD (pH ~6.80–6.90) approaches that observed during ischemia (pH ~6.75). Studies with sufficient temporal resolution also show a preceding alkaline shift of 0.05–0.16 pH units lasting a few seconds at the onset of depolarization (231, 329, 411). It is worth noting that intracellular pH also drops in neurons, but not in astrocytes where it increases (57, 536), and may transiently interfere with neuronal glycolytic phosphorylation due to the pH sensitivity of enzymes such as phosphofructokinase (337, 485).

As with high-energy phosphates, intracellular pH recovers within a minute or two (143). Tissue glucose concentration recovers after ~5 min, followed by a modest overshoot, while extracellular lactate and pH recover in ~10 min (237, 262, 329, 411), although recovery may be incomplete for more than 30 min as shown by microdialysis in one study (186). Glucose recovery is in part accomplished by a rapid increase in blood-brain glucose transfer, which starts during the depolarization prior to the onset of hyperemia without an apparent change in permeability for glucose (150). Indeed, a single SD decreases tissue glycogen by up to 50%, which is detectable within 2 min, and lasts 10–90 min (236, 237, 262, 416). To put in context, this degree of glycogen drop is larger than that seen after seizures, moderate hypoxia, or days of fasting and is exceeded only by severe anoxia with depolarization (5, 167). Glycogen breakdown is a result of the twofold increase in glycogen phosphorylase activity (235), stimulated by the two- to threefold increase in cAMP levels (234), both of which appear to peak shortly after DC shift recovery and last more than 10 min (235). The activation of glycogenolysis is consistent with intracellular alkalinization of astrocytes after SD (57), which stimulates the glycogen phosphorylase kinase via a rise in cAMP (196). The sensor may be the soluble adenylyl cyclase, which is highly expressed in astrocytes and becomes activated in response to HCO3 entry via the electrogenic Na+/HCO3 cotransporter. Activated soluble adenylyl cyclase increases intracellular cAMP levels, causing glycogen breakdown, enhanced glycolysis, and the release of lactate into the extracellular space. This process is recruited over a broad physiological range of extracellular K+ and also during aglycemic episodes, helping to maintain synaptic function (58). Although neurons may also contain glycogen (403), astrocytic glycogenolysis is likely to be the predominant source. Recovery of glycogen lags behind all other metabolites including creatine phosphate, glucose, and lactate. Unlike other metabolites, multiple repetitive SDs do have an additive effect on the magnitude of glycogen reduction (236, 237).

It is well established that SD stimulates oxidative phosphorylation and O2 demand (43, 98, 131, 138, 248, 273, 285, 290, 305, 325, 371, 396, 406, 469, 500, 516), with the exception of isolated retina preparations (501). Stimulation-induced rises in cerebral metabolic rate of O2 (CMRO2) are controlled by the ATP turnover, which depends on the energy used to fuel the Na+-K+-ATPase to reestablish ionic gradients after SD (110, 111). Therefore, SD-induced rises in CMRO2 support the increased Na+-K+-ATPase activity (42, 68, 69, 75, 143, 146, 258, 262, 298, 320, 364, 433, 440). The huge rise in cytosolic Ca2+ after SD depolarizes neuronal mitochondria and may trigger a rise in CMRO2 via the Ca2+ uniporter in the inner mitochondrial membrane (536). In most studies, CMRO2 increases by up to twofold coincident with the hyperemia (131, 371, 406). However, there are also data suggesting that during the depolarization (i.e., preceding the hyperemia) there may be a transient impediment to mitochondrial O2 utilization (285, 371, 532). Spectroscopy has detected a reduced rather than oxidized state of cytochromes c and aa3 (cytochrome c oxidase, complex IV) during the depolarization, despite apparently sufficient O2 availability (464, 520, 521, 523, 528). This could be explained by a number of mechanisms including diffusional barrier for O2, compartmentalization of energy metabolism that relies on glycolytic rather than oxidative phosphorylation, or disruption of mitochondrial respiration during SD because of mitochondrial depolarization and swelling which may alter the proton-motive potential (10, 133, 138, 203, 233, 285). As an exception, the profoundly vasoconstrictive response in the mouse brain severely limits the O2 supply, and by this way CMRO2 is paradoxically reduced, as discussed in detail later (17, 38, 54, 432, 532).

In rats, CMRO2 remains elevated in the post-SD period (10–20% above resting state) and gradually recovers over a period of 2 h (131, 371). This protracted increase in CMRO2 may be explained by oxidation of lactate converted to pyruvate and resynthesis of glycogen from glucose for the first 10–15 min. But this mechanism cannot account for the rise in CMRO2 from 15–60 min after the SD, when brain glucose and glucose usage, lactate, and glycogen have returned to normal (258, 262). This is reminiscent of the small rise in the cerebral oxygen/glucose uptake and consumption ratio during recovery from sensory stimulation that was not explained by delayed consumption of lactate or glucose (286). Glucose must traverse a series of intermediate metabolic pools before it is used directly for energy metabolism, and the size of the intermediate pools may enlarge during SD. The subsequent increase in CMRO2 may serve to decrease the size of these pools to resting levels (286) or to increase anaplerotic formation of glutamate (417) during SD and subsequent glutamate oxidation (192). In addition, the prolonged CMRO2 rise may relate to “oxygen debt,” known from studies of muscle metabolism, which is a persistent rise in oxygen use after exercise that relates to both oxidation of lactate and to replenishment of intermediate metabolic pools (288). Regardless of its mechanisms, protracted elevation in CMRO2 is clinically relevant because of its potential contribution to secondary brain damage in acutely injured human brain (176), where SD occurs in a large proportion of patients (79, 93, 115, 116).

IV. CEREBRAL HEMODYNAMIC EFFECTS OF SPREADING DEPRESSION

A. Historical Perspective

Leão (269) was struck by a conspicuous hyperemic response accompanying the SD wave that nearly doubled the pial arterial diameters, occasionally followed by a longer lasting constriction in rabbits. The dilation peaked within a couple of minutes and lasted a few minutes. In a subsequent study, Leão (271) spectroscopically showed increased oxyhemoglobin concentrations in pial veins during SD, supporting a hyperemic response. Several years later, Van Harreveld and Stamm (500) used tissue volume change as a surrogate marker for the vascular response in rabbits and reached a different conclusion; they detected cortical volume decrease during the DC shift, interpreted as vasoconstriction, sometimes followed by a slower but longer lasting volume increase suggesting dilation. Using a similar approach, Marshall et al. detected only vasodilation in cats (290), yet the same group had earlier reported no change in vessel diameters during SD in rabbits and monkeys (291). Both studies may have been confounded by tissue volume changes during SD that are independent of perfusion as has been shown in isolated retina (51).

Subsequent studies using the thermocouple approach, where heat transfer by blood flow is monitored, also yielded mixed results. One study showed a hyperemic response in most cats and rabbits, but an initial reduction in blood flow was observed during the DC shift in a few rabbits (498). Another study in cats and rabbits found a highly variable response among animals and brain regions where “blood flow might increase, decrease, or show a biphasic change,” although the response remained relatively uniform within a given region upon repeated SDs (487). The same technique in rats mainly showed hyperemia that started more than a minute after the DC shift onset, without any evidence of hypoperfusion during or after the DC shift (44). It was also associated with increased cortical hemoglobin content as a surrogate measure for cerebral blood volume (CBV) in the same study, detected 3 min after the DC shift and resolved after 5 min. Hyperemia was later confirmed in cats using a thermocouple (290).

Arguably the most telling data among early studies were obtained by flash cortical freezing approach, which minimized trauma by monitoring devices and enabled visualization of both pial and penetrating vessels in histological sections cut parallel to the propagation path of SD in transit across the cortex. All rabbits studied in this manner showed a narrow band of vasoconstriction that spatiotemporally coincided with the DC shift, followed by a much broader area of vasodilation (498). Interestingly, none of the cats studied the same way displayed vasoconstriction even when their gyrencephalic cortex was appropriately sectioned to allow assessment of a sufficiently long path of SD propagation, hinting for species differences (498).

In the next two decades, cortical surface spectroscopy of changes in tissue hemoglobin content as a surrogate for CBV once again showed marked heterogeneity during SD; monophasic CBV reductions or increases during the DC shift, biphasic or triphasic responses, and even no change, have all been reported (280, 297, 300, 301, 307, 398). Therefore, heterogeneity in the observed hemodynamic responses persisted despite methodological advances over time, suggesting a complex response pattern composed of multiple components with independent mediators and modulators. First, we review these components in order of their dominance in the hemodynamic response.

B. Prototypical Cerebral Hemodynamic Response to SD

In most species (e.g., rats, rabbits, cats) and under normal physiological conditions, SD triggers a massive peak hyperemia that dominates the vasomotor response (Figure 4; component II) (2, 17, 20, 22, 66, 86, 90, 94, 95, 97, 98, 113, 114, 118, 119, 124, 130, 131, 150, 152155, 163, 179, 186, 200, 207, 220, 221, 226, 242, 248, 257, 260, 263, 305, 320, 321, 343, 352, 354, 371, 373, 376, 382, 384386, 389, 392, 394, 406, 410, 422, 424426, 443, 444, 456, 457, 460, 461, 469, 489491, 520, 521, 523, 530, 534, 535). Electrophysiological recordings and voltage-sensitive dye imaging coupled to blood flow measurements show that the hyperemia starts 15–20 s after the depolarization onset, and peaks only after full repolarization, usually 1–2 min after DC shift onset (17, 352). The magnitude of hyperemia has been highly variable, between 30–250% cerebral blood flow (CBF) increase, although exceptions certainly exist ranging between no detectable hyperemia (257) particularly in awake animals (95, 96), and up to fivefold CBF increase (327, 426, 427). No doubt such variability is multifactorial, including the species, experimental conditions, and measurement techniques (see below for a detailed discussion). Direct pial artery diameter measurements showed a corresponding dilation (25–120%) in most studies, which appeared to be more prominent in smaller caliber arterioles than larger ones (16, 38, 46, 59, 64, 65, 314, 315, 407, 419422, 439, 460, 461, 512, 513). Multispectral reflectance, near-infrared spectroscopy, light transmission, and intravascular fluorescence imaging studies also showed hyperemia during SD in rats and cats (Figure 5A; component II) (18, 81, 163, 195, 227, 351, 407, 474, 528). Hyperemia was further confirmed by magnetic resonance imaging (MRI), where CBV increased by ~30% for 1–2 min with a latency of 15–20 s after the apparent diffusion coefficient (ADC) changes during the depolarization (71). There is good evidence that peak hyperemia associated with SD leads to functional recruitment of red blood cells into the microcirculation (i.e., reduction in the percentage of capillaries free of erythrocytes; Figure 5B; component II) (287, 343). In contrast to arteries and arterioles, venous caliber changes are often inconspicuous, delayed, or absent (38, 59, 419, 460, 461).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390004.jpg

CBF response to SD. Sample tracings show a range of CBF responses to SD in the rat (A–G), cat (H), and mouse (I) cortex obtained by laser Doppler (A–H) or speckle flowmetry (I) from different laboratories, selected to demonstrate the variable presence, timing, and magnitude of the four vasomotor components (I–IV) shaping the CBF response as described in text. A–C: CBF responses to single or consecutive cortical SDs and their temporal relationship to the DC shift recorded using intracortical microelectrodes under isoflurane anesthesia. [Modified from Ayata et al. (17) and Sukhotinsky et al. (456).] D: CBF responses to two consecutive cortical SDs 15–20 min apart, and their temporal relationship to the depolarization as imaged using a voltage-sensitive dye under halothane anesthesia. [Modified from Farkas et al. (124).] E and F: CBF responses to cortical SD under halothane anesthesia. [Modified from Fabricius and Lauritzen (119) and Farkas et al. (123).] G: evolution of the CBF response to cortical SDs triggered consecutively every 30 min under halothane anesthesia. Note diminished initial hypoperfusion (component I), augmented hyperemia (component II), decreased latency to component III merging into the hyperemia, and absence of a change in post-SD oligemia (component IV). [Modified from Fabricius et al. (114).] H: CBF response to cortical SD in a cat under urethane/α-chloralose anesthesia. Note the prominent vasomotion at resting state and its disappearance after the SD. [Modified from Piper et al. (376).] I: CBF responses to two consecutive cortical SDs triggered 15 min apart in a mouse under isoflurane anesthesia. Although the pronounced hypoperfusion (component I) dominates the response, all other components are also readily identifiable. The second SD is superimposed on severe post-SD oligemia (component IV) after the first SD, and does not show component I, closely resembling the response to SD in other species. [Modified from Yuzawa et al. (532).]

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390005.jpg

CBV, capillary flow, and pial artery diameter responses to SD. A: time-lapse subtraction images of transilluminated cortical light transmission show a spreading concentric wave of an initial increase in light transmittance (i.e., CBV reduction, I) followed by a decrease in light transmittance (i.e., CBV increase, II) images in a rat (left panel) and a cat (right panel) under urethane/α-chloralose anesthesia. Time after SD induction is indicated on the top left of each image. Arrow: the site of KCl injection. [Modified from Tomita et al. (481).] B: capillary blood flow dynamics during SD studied by line scans using in vivo multi-photon microscopy (top panel) through a closed cranial window in a representative isoflurane-anesthetized rat (postnatal day 15) show blood cell flow arrest (middle panel) during the DC shift (bottom panel) that corresponds to component I. Following repolarization, there was a rebound increase in blood flow (component II). [Modified from Chuquet et al. (59).] C: combined electrophysiological recording and multimodal optical imaging in a representative isoflurane-anesthetized mouse show changes in total hemoglobin ([Hbtot]), pial artery diameter (PAD), hemoglobin O2 saturation (SatO2; %), and optical intrinsic signal (OIS) in relation to the DC shift (field potential; mV). Different vasomotor components (I–IV) are labeled as described in the text as well as in Figure 3. Note the severe vasoconstriction and hemoglobin desaturation during the DC shift (I), and long-lasting and severe post-SD oligemia (IV). [Hbtot], PAD, and OIS are normalized to pre-CSD baseline. [Modified from Chang et al. (54).]

The hyperemia usually lasts up to 3 min, and often gives way to a prolonged post-SD oligemia; CBF remains 10–40% below baseline for an hour or longer (Figure 4; component IV) (17, 90, 9497, 114, 119, 124, 130, 131, 152154, 171, 226, 242, 248, 255, 263, 343, 352, 371, 373, 376, 394, 410, 424, 425, 491, 502, 534, 535). Once again, corresponding constrictions have been observed in pial arteries (407, 421, 512), and reduced CBV by MRI, near-infrared spectroscopy, multispectral reflectance, and intravascular dye imaging (18, 71, 195, 227, 351, 407).

Many studies with sufficient spatiotemporal resolution have also observed an initial hypoperfusion coincident with the DC shift (Figure 4; component I) (17, 86, 90, 114, 119, 124, 128, 131, 150, 152, 154, 163, 179, 186, 257, 260, 343, 352, 376, 385, 389, 394, 406, 424, 443, 456, 460, 474, 490, 491). When present, this transient vasoconstrictive tone usually decreases CBF by 5–30% for 5–30 s, and can either precede the hyperemia leading to an early dip in CBF, or be superimposed on the rising slope giving the response a notched or jittery upstroke. Corresponding changes in vessel diameter have been detected by intravital microscopy of pial arteries (16, 46, 54, 59, 359, 419, 420). Moreover, several surface reflectance and light transmission studies have detected reduced CBV preceding the hyperemia (Figure 5A; component I), or notched or jittery upstroke, in multiple species including rats and cats (98, 163, 227, 299, 303, 305, 306, 359, 407, 442, 444, 474, 479, 481). Indeed, dramatic capillary red blood cell slowing, flow arrest, and even transient disappearance of red blood cells from capillary segments have all been observed (Figure 5B; component I), lasting as long as a minute during SD, without any evidence for focal or diffuse capillary narrowing or constriction, compression or collapse, or occlusion (59, 479, 480, 490, 491, 532). The flow reduction in parenchymal arterioles and capillaries may precede the vasoconstriction in larger pial arteries (359). Altogether, these data indicate that variable degree of vasoconstriction and hypoperfusion occur during the DC shift. Although the principle site(s) at which the cerebrovascular resistance increases during tissue depolarization is unclear, the phenomenon has critical relevance in the pathophysiology of stroke and other brain injury states (discussed in detail later).

Another inconsistently observed component of the vasomotor response is a second smaller (10–50%) but longer lasting (4–8 min) late hyperemia (or dilation) peaking 3–5 min after the peak hyperemia, forming a characteristic hump (Figure 4, B and D–H; component III) (114, 119, 124, 152, 154, 200, 352, 376, 385, 394, 406, 419, 422, 520, 521). In addition, a mild hyperemia (~10%) (120) and a much delayed oligemia (~30%) (360) have been shown to follow at 1–2 h and 3 days later, respectively, although the latter was in response to numerous SDs triggered consecutively over a couple of hours. Lastly, an early rise or fall in CBF preceded the DC shift sometimes by up to a minute in some studies (17, 456), possibly vascular conduction of the vasomotor signal ahead of the approaching wave front (38).

Despite the comparable hemodynamic responses in rats, rabbits, and cats, SD may evoke a different response in mice, where a CBF decrease of 60–70% may accompany the SD wavefront (Figure 4I; component I) (17, 432, 492, 532). This initial hypoperfusion starts near simultaneously with the DC shift and reaches a nadir ~20 s later. The duration of hypoperfusion is closely coupled to that of the DC shift, suggesting a causal link. During the repolarization, CBF transiently normalizes to a peak at ~1 min after DC shift onset, temporally corresponding to the peak hyperemia in other species (Figure 4I; component II). A smaller late hyperemia starts at ~2 min after the SD onset (Figure 4I; component III), similar to that observed in other species. This is followed by a decline in CBF of 60% for up to an hour corresponding to post-SD oligemia (Figure 4I; component IV). This multiphasic vasomotor response in mice, with a profound constriction coupled to the DC shift, has been confirmed using direct pial arterial diameter and tissue hemoglobin concentration (i.e., CBV) measurements (Figure 5C) (38, 50, 54, 492, 532), although not all studies agree (469, 476).

A hyperemic response to SD has been reported in barbiturate-anesthetized primates using PET imaging, as the species most proximate to human (530). Although the electrophysiological properties of SD in primate cortex were comparable to other species, SD rarely propagated more than a few centimeters, suggesting reduced SD susceptibility compared with other gyrencephalic species such as the cat. In this model, SD evoked a focal and rather small (30–50%) peak hyperemia that was most prominent at the KCl application site compared with contralateral cortex, and there was no post-SD oligemia, raising the question whether the closely related human brain can mount as profound a hyperemic response as observed in anesthetized rats, cats, and rabbits (e.g., 100–200% CBF increase). Nevertheless, hyperemic responses have been detected in human brain during migraine aura and after brain injury, as summarized below in detail (169, 194, 519).

In summary, data so far suggest that the hemodynamic response to SD is composed of multiple opposing vasomotor influences at different stages of the SD wave (Figure 6), variations in the presence, amplitude, and timing of which account for much of the heterogeneity of the hemodynamic response. The prototypical CBF response has at least four distinct vasomotor components: I) an initial hypoperfusion or constriction coincident with the DC shift; II) a peak hyperemia or dilation that emerges during repolarization and reaches a maximum after complete DC shift recovery; III) a smaller late hyperemia or dilation that develops 3–5 min after SD; and IV) a post-SD oligemia or constriction lasting an hour or more. Each component is an aggregate response originating from one or more vascular compartments. Throughout this review, we will use these four terms uniformly when referring to hemodynamic components I–IV to minimize confusion. Undoubtedly oversimplified, the framework nevertheless serves as a starting point to better understand the sources of heterogeneity of the hemodynamic responses in the literature, and to dissect the dilator and constrictor mechanisms, mediators and modulators (197).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390006.jpg

A conceptual framework of vasomotor components shaping the CBF response to SD. Four distinct vasomotor components can often be identified in response to SD. Local depolarization (i.e., DC shift) coincides with a vasoconstrictive tone that creates an initial hypoperfusion (component I). This component has been rather capricious; in some cases it is large and conspicuous, in others it is small to only create a brief dip or a notchy pattern superimposed on the upstroke of hyperemia, or it may be absent. The second component (II) is a vasodilator tone that gradually builds after the depolarization onset and gives rise to a peak hyperemia that reaches a peak after full repolarization. This is often the most conspicuous component of the vasomotor response to SD. The vasodilator component subsides within minutes, giving way to a prolonged vasoconstrictive tone leading to post-SD oligemia for an hour or more (component IV). More often than not, a second smaller CBF rise (late hyperemia; component III) is interposed between the peak hyperemia and post-SD oligemia. [Modified from Ayata (14).]

C. Potential Sources of Heterogeneity in the CBF Response to SD

A number of factors may account for the heterogeneity in the hemodynamic response among studies. For example, species and strain differences may account for some of the variation. Most notably, as described above, the hemodynamic response in mice radically differs from most other species, and consists of a pronounced initial hypoperfusion, and small or no hyperemia above baseline (17, 38, 54, 532). There can be minor differences among different strains of the same species as well (17), which may become relevant when examining transgenic models. Sex difference has not been observed so far (532). In the majority of species studied, however, peak hyperemia dominates the response, and therefore, most heterogeneity stems from other factors.

The response can also differ among brain regions and arterial territories (532). More importantly, there is evidence suggesting that different vascular segments and compartments (e.g., pial arteries, penetrating arterioles, capillaries, veins) may respond differently to SD. Pial arteries may dilate, when penetrating vessels constrict (376, 512), and smaller caliber pial arterioles may dilate more than the larger ones (419). The principle was nicely demonstrated by multiphoton microscopy in young rats where the magnitude of initial constriction and subsequent dilation were both twofold higher in small pial arteries compared with large ones, and smaller penetrating arterioles responded almost exclusively with vasoconstriction (59).

Differential responses in vascular compartments and segments have implications for the techniques used to measure the hemodynamic response. Some techniques sample a particular vascular segment or compartment, adding to the observed heterogeneity of the hemodynamic response. For example, intravital microscopy of pial artery diameter responses can differ from those recorded by laser Doppler flowmetry, which samples CBF in ~1 mm3 of cortex under the probe. Indeed, post-SD constriction of pial arterioles is smaller and lasts shorter than post-SD oligemia measured by laser Doppler flowmetry (LDF) or indicator fractionation in otherwise similar experimental models (421). Further supporting the notion, depth-resolved LDF has shown larger peak hyperemic responses in deeper cortical layers (~120–130% increase at a depth of >250 μm) compared with superficial layers (~80–90% increase within the first 250 μm) (113, 343). Diffuse optical coherence tomography showed minimal or no CBF change at even greater depths within the rat cortex (~2 mm) and underlying white matter (~3 mm) when CBF was nearly doubled at a depth of ~1 mm (535). Moreover, some techniques may be more sensitive to speed of flow, while others to bulk flow. It is also possible that some measurements (e.g., LDF through intact skull or dura) may be contaminated by changes in meningeal flow after SD, even if its contribution is small (30). and lastly, the spatiotemporal resolution of each technique is different (e.g., μm with multiphoton imaging, mm with LDF), and small and brief vasomotor components may be lost because of spatial or temporal averaging.

Resting CBF on which SD is superimposed is another important source of heterogeneity. SD-induced peak hyperemia (component II) is often smaller, and the initial hypoperfusion (component I) is more prominent, when resting CBF and vessel calibers are elevated (e.g., acetazolamide, inhalational anesthesia, dimethyl sulfoxide) than when they are reduced (e.g., barbiturate or propofol anesthesia, indomethacin, hypocapnia, or post-SD oligemia after a preceding SD) (71, 95, 257, 394, 460). Although anesthetics can alter SD susceptibility (74, 191, 200, 240, 241, 375, 404, 443, 466), they do not appear to have a consistent direct effect on the hemodynamic response (443); both absent and potent peak hyperemic responses have been reported in unanesthetized animals in different studies (95, 424, 443). Nevertheless, anesthesia can indirectly influence the hemodynamic response by changing resting CBF (128) or systemic blood pressure (see below).

Arguably one of the overlooked confounders of the hemodynamic response to SD is inadvertent triggering of one or more SDs during surgical preparation (54, 478). Although repeated SDs are highly reproducible electrophysiologically, the hemodynamic response to the first SD occurring in naive tissue is decidedly different than the response to subsequent SDs occurring less than 30–45 min apart, superimposed on post-SD oligemia caused by the preceding SD. During subsequent SDs, initial hypoperfusion (component I) is diminished, peak hyperemia (component II) tends to get larger, late hyperemia (component III) gradually increases in magnitude and merges with the peak hyperemia that precedes it, and post-SD oligemia (component IV) is only minimally augmented (Figure 4, C, ,D,D, and andGG) (22, 71, 114, 124, 152, 163, 227, 352, 406, 427, 443, 474). In the mouse, subsequent SDs triggered within 30 min are superimposed on a severe post-SD oligemia (40% of baseline CBF) (17, 532). Therefore, the initial hypoperfusion coupled to the DC shift is minimal or absent, and instead of a transient normalization of CBF, there is an ostensibly monophasic peak hyperemia that represents 100% increase relative to the post-SD oligemic baseline (Figure 4I). After this apparent hyperemic phase, CBF returns to approximately the same post-SD oligemia level (17, 532). Similarly, pial arterial diameter measurements have shown diminished vasoconstrictive responses with each SD triggered in rapid succession (54). In all species, if SDs are repeated more than 1 h apart, the response resembles that in the naive cortex. Therefore, failure to detect and recognize the occurrence of SD during experimental preparation may be a confounder. For example, CBF measurements using methods involving electrode insertion into the tissue (e.g., hydrogen clearance, microdialysis) have yielded abnormally low CBF values. A careful analysis of possible explanations has eliminated tissue injury as a cause, and instead identified SD triggered during electrode insertion as the culprit (24, 478, 502). Indeed, craniotomy is a frequent cause of SD, especially in susceptible species and under favorable experimental conditions (e.g., barbiturate anesthesia), and can profoundly alter the hemodynamic responses to SD (54). Because the initial hypoperfusion virtually disappears after the first SD, it is the most frequently missed vasomotor component of the CBF response (469). The absence of post-SD oligemia and presence of abnormal vascular reactivity (see below) in some studies may also be due to a preceding undetected SD (248).

Systemic physiological parameters (e.g., arterial blood pressure, Po2) can dramatically change the shape and the magnitude of the vasomotor response to SD. Systemic hypotension in rats, even when it is relatively mild within the autoregulatory range, brings out and augments the initial hypoperfusion (component I) during the DC shift, and diminishes the peak hyperemia (component II), while prolonging both components; systemic hypoxia has a similar effect, albeit milder (Figure 7A) (123, 198, 445, 456, 457). Systemic hypercapnia has also been reported to diminish the peak hyperemia, possibly because SD was superimposed on an elevated resting CBF (523). Systemic hyperglycemia appeared to augment both the peak hyperemia and the post-SD oligemia (515, 523). Therefore, it is critical to monitor all systemic physiological parameters during SD experiments, in vivo. Lastly, aging has been reported to diminish the magnitude of all vasomotor components in rats (2 versus 8 mo old) (123), and to a lesser extent the initial hypoperfusion in mice (19 mo old) (432).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390007.jpg

Systemic physiological and pharmacological modulation of CBF response to SD. A: hypotension (blood pressure ~40–45 mmHg) transforms the large predominantly hyperemic response to SD (left) into a biphasic or triphasic response with predominantly hypoperfusion during the DC shift (middle); resting CBF was only mildly reduced (~10%). The addition of hypoxia (Po2 ~55 mmHg) on to hypotension did not further alter the response (right). Note also the prolongation of DC shift (top tracings) during hypotension. CBF was recorded by LDF in isoflurane-anesthetized rats. [Modified from Sukhotinsky et al. (456).] B: systemic administration of nitric oxide synthase (NOS) inhibitor N-nitro-l-arginine methyl ester (l-NAME) also transforms the response by bringing out an initial hypoperfusion (component I), as shown by [14C]-iodoantipyrine autoradiography. [Modified from Duckrow (94).] C: systemic NOS inhibitor N-nitro-l-arginine (l-NNA) exerts a similar effect in rats recorded by laser Doppler flowmetry. Note the appearance of an initial hypoperfusion (component I) and diminished hyperemia (component II). Top two tracings show two consecutive cortical SDs under control conditions. The effect is reversed by systemic l-arginine. [Modified from Fabricius et al. (114).]

V. MECHANISMS, MEDIATORS, and MODULATORS OF THE CBF RESPONSE TO SPREADING DEPRESSION

The mediators and permissive modulators of the hemodynamic response to SD have been subject to much investigation. Because SD markedly stimulates ATP, glucose, and O2 consumption, the hemodynamic response, particularly the peak hyperemia, was thought to reflect metabolic coupling. The fact that hyperemia starts during the DC shift when tissue glucose reaches a nadir (321), and that its magnitude is inversely related to resting CBF (257), further supported the notion that it is a reactive hyperperfusion state such as after a period of ischemia. However, providing a surplus of oxygen and glucose during SD did not diminish the CBF response (276, 523), nor did oxygen or glucose shortage augment it (457), casting doubt over this simplistic view.

Indeed, there are ample data suggesting that, to the contrary, SD causes neurovascular uncoupling (54, 371), and that the hemodynamic response is not titrated to match the increased metabolic demand. The principle (i.e., lack of substrate sensing) also applies to neurovascular coupling during more physiological levels of activation (54, 132, 133). The CBF response to SD is likely a result of direct cell-cell contact or via diffusible mediators within the neurovascular unit. Neurovascular unit is a conceptual ensemble consisting of many cell types, including astrocyte end feet, parenchymal neuronal projections, perivascular nerves, pericytes, smooth muscle cells, and the endothelium. Hence, a multitude of potential vasoactive substances are released during SD (160, 441), many of which have been implicated in shaping the hemodynamic response (Figure 8; Table 1). The mechanisms likely differ among vascular compartments (e.g., pial versus parenchymal arterioles and capillaries), further adding to the complexity.

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390008.jpg

Propagation of SD across the dense cerebral vascular network and the neurovascular unit. SD invading the tissue profoundly impacts all cells and constituents of the cerebrovascular unit. The intense pandepolarization causes massive extracellular ionic, neurotransmitter, and metabolic changes. The cerebrovascular response is the result of numerous complex processes and interactions within the unit, mediated by key cell types and molecular signals, some of which are listed. The large number of mediators and modulators simultaneously released from multiple different cell types has made it extremely difficult to dissect the role of individual factors in the hemodynamic response, further complicated by the heterogeneity of the normal response among studies. Therefore, we have refrained from attempting a contrived representation of the vasomotor actions of each potential mediator or modulator, and instead summarized the available data in Table 1. EDHF, endothelium-derived hyperpolarizing factor; PACAP, pituitary adenylate cyclase activating peptide; NE, norepinephrine; NPY, neuropeptide Y; ACh, acetylcholine; SP, substance P; NK-A, neurokinin A.

Table 1.

Physiological and pharmacological modulators of the cerebral hemodynamic response to spreading depression

Hemodynamic ComponentAugmented ByAttenuated By
Initial hypoperfusion/constriction (component I)Elevated [K+]oPreceding SD
Reduced NO availability
Voltage-gated Na+ channel inhibition
High resting CBF
Hypotension
Peak hyperemia/dilation (component II)5-HT1D/1B activationCGRP inhibition
KATP inhibitionNO inhibition
Voltage-gated Na+ channel inhibitionMuscarinic inhibition
Adenosine A1 receptor inhibition5-HT2A/2C inhibition
Cyclooxygenase inhibitionSensory perivascular nerve inhibition
Lipoxygenase inhibitionParasympathetic perivascular nerve inhibition
mPTP inhibitionHypotension
Low resting CBF
Hyperglycemia
Post-SD oligemia/constriction (component IV)Voltage-gated Na+ channel inhibition20-HETE inhibition
HyperglycemiaCyclooxygenase inhibition
L-type Ca2+ channel inhibition
NO activation
Preceding SD
mPTP inhibition
Calcineurin inhibition

SD, spreading depression; [K+]o, extracellular K+ concentration; CBF, cerebral blood flow; NO, nitric oxide; CGRP, calcitonin gene-related peptide; mPTP, mitochondrial permeability transition pore; 5-HT, serotonin.

A. Cellular Constituents of the Neurovascular Unit and the Vasomotor Response to SD

1. Astrocytes

Although all cell types within the neurovascular unit are likely to contribute in a temporally coordinated manner, astrocytes are critically positioned to regulate the cerebrovascular tone in response to events taking place in brain parenchyma. Astrocytes can trigger vasodilation or constriction in a context-dependent manner (159, 275, 318, 328, 468, 537). During normal (i.e., functional) neuronal activation, metabotropic glutamate receptor activation induces Ca2+ elevations in astrocytic end feet. Although these inositol 1,4,5-trisphosphate receptor (IP3R)-generated Ca2+ rises are too slow to explain alone the relatively rapid increase in CBF that accompanies normal neuronal activation (328, 537), prolonged neuronal and glial depolarization during SD can certainly recruit astrocytic mechanisms of vascular regulation. Astrocytic mediation of neurovascular coupling is generally slower than direct neuronal mechanisms, but its relevance is likely to be amplified during the sustained depolarization of SD (52).

Astrocytic Ca2+ elevations stimulate arachidonic acid production by phospholipase A2, which can then be converted to dilator epoxyeicosatrienoic acids (EETs) by the epoxygenase CYP42C11 (6), constrictor 20-HETE by CYP4A in vascular smooth muscle (328), or dilator prostaglandin E2 by cyclooxygenase-1, to relay the vasomotor signal to the vessel wall. The magnitude of end feet Ca2+ transients, background nitric oxide synthase (NOS) activity, tissue O2 availability, and extracellular lactate and adenosine levels may all play a role in determining the polarity of the vasomotor effect (159, 275, 328, 468, 537). In immature rats, fast astrocytic Ca2+ elevations during SD trigger vasoconstriction of intracortical arterioles, in vivo (59), instead of the anticipated dilation (159, 537). The data suggest that the astrocytic control of vasomotor response is context dependent and that there may be subcellular compartmentalization of astrocytic Ca2+ elevations with specific downstream targets and mediators.

Astrocytes are also critical for perivascular K+ signaling as a complementary mechanism of neurovascular coupling. Astrocytic Ca2+ waves triggered by neuronal activation open Ca2+-sensitive large conductance K+ (BK) channels clustered on the end feet, which elevates perivascular [K+]o to hyperpolarize vascular smooth muscle via inward rectifying K+ (Kir) channels (129, 377). Although the contribution of end feet BK channels to SD-induced vasomotor response is not known, the mechanism can depolarize vascular smooth muscle if [K+]o rises to sufficiently high levels (>20 mM), providing bipolar modulation (149).

Lastly, swelling of astrocytic end feet during SD has been hypothesized as a potential mechanism to compress the microvasculature and restrict perfusion, which may explain the capillary flow arrest during SD (8, 479, 480). Capillary narrowing can have a dramatic effect on cerebrovascular resistance. However, the phenomenon has not been observed in studies with sufficient spatiotemporal resolution (469, 532, 536), and further studies are needed to address this issue.

2. Parenchymal neurons

Neurons have direct vasomotor effects. Both pyramidal and interneurons regulate vasomotor tone and contribute to neurovascular coupling upon synaptic glutamate release and raised neuronal [Ca2+]i. For example, somatosensory stimulation induces hyperemia in part via activation of cortical pyramidal cell cyclooxygenase-2 (272, 345). In addition, cortical GABAergic interneurons receive input from subcortical cholinergic, cathecolaminergic, and serotonergic efferents and make direct contacts with local microvasculature to induce dilation by vasoactive intestinal peptide and nitric oxide, and constriction by somatostatin and neuropeptide Y (53, 108, 225). Activation of noradrenergic input from locus ceruleus on to the astrocytic end feet around the blood vessels reduces resting CBF (62, 72, 365, 379). However, data on direct neuronal control of vasomotor responses during SD are sparse. For example, inhibition of controlled neuronal vesicular exocytotic neurotransmitter release by tetrodotoxin or tetanus toxin did not alter the vasomotor response to SD (59). In contrast, topical application of tetrodotoxin at concentrations that abolished the electrocorticogram reduced resting CBF and augmented both the initial hypoperfusion and the post-SD oligemia in another study (114).

3. Perivascular nerves

Pial (i.e., extraparenchymal) arteries are richly innervated by perivascular nerves originating from superior cervical ganglion (sympathetics containing norepinephrine and neuropeptide Y), sphenopalatine, otic and internal carotid ganglia (parasympathetics containing acetylcholine, vasoactive intestinal polypeptide, and NOS), and trigeminal ganglion (sensory nerves containing calcitonin gene-related peptide, substance P, neurokinin A, and pituitary adenylate-cyclase activating polypeptide) (172). As with parenchymal neurons discussed above, perivascular nerves are in contact with the arteries and capable of directly regulating the vasomotor tone. Sympathetic nerve activation constricts the pial arteries, particularly large caliber vessels, while parasympathetic nerves are dilators (29, 102, 462). Sympathectomy reportedly has no effect on the peak hyperemic response to SD in the rat either acutely or chronically, although the spatiotemporal resolution of the methods may not have been sufficient to resolve early or late constrictive components in this study (44). In contrast, both sensory nerve transection and parasympathectomy diminished the peak hyperemic response to SD by 20–30% without altering the hypercapnic hyperemia or the DC shift (389). Importantly, a chronic lesion was required, suggesting that distal degeneration of perivascular nerve fibers is essential, and that local depolarization of perivascular nerve endings triggers the release of vasoactive substances, rather than a nerve conduction-dependent effect via efferent activity from the ganglia. Indeed, both elevated [K+]o and glutamate can potently trigger the release of vasoactive substances from perivascular nerve endings (312). Consistent with these data, acute systemic administration of capsaicin or direct trigeminal ganglion stimulation to transiently deplete vasoactive peptides in sensory nerve endings also transiently diminished the peak hyperemia by 20–30% (22).

4. Endothelium

Endothelium does not appear to directly mediate or modulate the vasomotor response to SD. For example, inhibition of endothelium-dependent hyperpolarizing factor production by cytochrome P-450 epoxygenase inhibitor miconazole, direct endothelial disruption by intravenous phorbol 12,13-dibutyrate, or endothelial dysfunction induced by insulin resistance did not alter the hyperemia or the DC shift during SD (427). Of note, endothelial cells do change their morphology (e.g., swelling, contraction) acutely in response to local oxygenation changes (212, 213); however, such changes have not been observed or reported during SD, in part due to insufficient spatial resolution of the imaging techniques.

5. Pericytes

Pericytes are critically positioned to regulate capillary blood flow, and pericyte contraction in response to oxidative stress reportedly contributes to the no-reflow phenomenon after focal ischemia (527). One might therefore speculate that the initial hypoperfusion during the depolarization (component I) and the post-SD oligemia (component IV) may be produced by pericyte constriction. However, high spatiotemporal resolution microvascular imaging has thus far failed to detect any focal constriction of capillaries during SD, in vivo (469, 480, 532). Moreover, pial artery constriction observed during and after SD (512) cannot be explained through a pericyte mechanism. Nevertheless, these modified vascular smooth muscle cells may contribute to CBF regulation under physiological conditions. Recent data suggest that neuronal activity and the neurotransmitter glutamate evoke the release of messengers that dilate capillaries by actively relaxing pericytes (170), although not all studies support a role for pericytes in functional hyperemia (127). Future work is expected to shed light on pericyte signaling to regulate capillary diameter during SD and the possible importance of pericytes for local blood flow control.

B. Vasoactive Substances Released During SD

Numerous vasoactive substances are released during SD, largely inferred from studies on the effect of blockers of synthesis, release, and interaction with vascular receptors. In the following section, we review the data that mainly stem from pharmacological interventions, i.e., without directly monitoring the activity of a specific pathway targeted by the drug. This weakness, inherent to in vivo dissection of physiological mechanisms, is difficult to circumvent and the matter is complex because SD evokes a rise in almost any mediator or modulator that has been measured so far. These substances differ in a number of important aspects. Some are released from stores, while others require de novo synthesis. Some are released in the vicinity of the vasculature, while the levels of others rise dramatically in the parenchyma and may diffuse or be siphoned to the vasculature. Some directly mediate a response, while others play a permissive role. and some have effects that last seconds, while others trigger changes that can last hours. The complexity is brought about by the intensity, duration, and multicellular nature of the pandepolarization and ion fluxes during SD. Concentrations of vasoactive substances often are exceedingly small and difficult to directly measure. Parallel slice experiments are most commonly required because the vasoactive compounds, especially the lipophilic ones, disappear within the brain's largest sink, the blood flow (131). Therefore, our understanding of vascular mediators and modulators is based on indirect evidence, mainly from pharmacological or ablation studies.

1. Ions and channels

During SD [K+]o rises ~10-fold from 3–4 mM to ~30–60 mM (Figure 1), and all cells within the neurovascular unit are likely exposed, including vascular smooth muscle. [K+]o elevations up to 20 mM hyperpolarize smooth muscle membrane via activation of Kir channels, in particular Kir2.1, and reduce [Ca2+]i and muscle tone (157, 224, 246, 309). The dilation response to low elevations of [K+]o appears to be specific for cerebral vessels (157, 402). When [K+]o exceeds 20 mM, direct smooth muscle membrane depolarization, [Ca2+]i rise, and constriction ensue, reaching near maximum above 60 mM (157, 310, 402). Of note, cerebral arteries are more sensitive to the constrictive effects of elevated [K+]e compared with peripheral vasculature (e.g., mesenteric artery), and perhaps more importantly, human pial arterioles are more sensitive to [K+]o than either rabbit or rat pial arteries (e.g., EC50 and ECmax 18 and 45 mM in human, versus 37 and 109 mM in the rat, respectively), suggesting that the constrictive effect of [K+]o elevation may be stronger in human brain (402). Similar species differences in the cerebrovascular constrictive response to high [K+]o have also been reported between rats and mice, where isolated mouse basilar arteries were more sensitive to K+-induced constriction compared with rats (17).

Taking into account the vasoactive properties of K+, [K+]o changes during SD are expected to initially increase CBF, and subsequently to constrict the resistance vessels when [K+]o exceeds ~20 mM. Because the [K+]o surge during SD is tightly coupled to the extracellular DC potential shift (Figure 1E), one can use the DC shift as a surrogate marker to temporally correlate [K+]o changes to CBF. As such, the initial hypoperfusion coincident with the DC shift may reflect the constrictive influence of [K+]o on the cerebral vasculature. The fact that not all animals and SDs show initial hypoperfusion during the DC shift may be because spatial K+ buffering prevents the perivascular [K+]o reaching more than 20 mM, or because concurrent release of dilator substances counteracts the vasoconstrictive influence. There appears to be competition between dilators and constrictors during the DC shift; K+ is the leading constrictor known released during the DC shift, and pharmacological inhibition of opposing dilators such as nitric oxide (NO) can bring out the constrictive effect of K+. Peak hyperemia starts during the recovery of [K+]o and peaks after complete restoration of [K+]o. Vasomotor changes closely resembling the response to SD have indeed been reproduced by focally applying high-K+ pulse with a micropipette in cats (359), although it is not clear whether this is a direct effect of K+ or the result of an SD-like cascade of events triggered by the pulse injection.

An important consideration is how elevated [K+]o during SD gets into contact with the vasculature to exert its vasomotor effects. Astrocytic end feet completely encase parenchymal microvasculature. Astrocytes also play a critical role in restoring the [K+]o during SD by uptake and spatial buffering, and may release K+ into the perivascular space through large conductance Ca2+-activated K+ (BK) channels which are capable of carrying large K+ fluxes. However, inhibition of BK channels by charybdotoxin did not have an effect on the CBF response to SD (426). Microinjections of KCl into the cortex can raise [K+]o comparable to levels evoked by stimulation, but do not achieve the same magnitude of CBF increases that develop in response to stimulation (48). The data suggest that mediators other than [K+]o are also important for activity-dependent rises in CBF (47). Passive diffusion is a possibility in superficial vasculature within the Virchow-Robin space and the interstitial fluid. However, parenchymal [K+]o changes during SD may be unlikely to have significant influence on larger pial vasculature, because pial perivascular [K+]o reportedly does not increase significantly during SD, and topical superfusion of up to 12 mM K+ does not reproduce the hemodynamic changes during SD such as the large hyperemia (420). Alternatively, K+-induced local vasomotor changes can be conducted undiminished for more than a millimeter away via endothelial direct coupling, and therefore, it may be hypothesized that parenchymal [K+]o changes trigger upstream pial arterial diameter changes during SD (201). In support of this, conducted vasodilation has been observed during SD (38).

Effects of K+ channel modulators on the CBF response to SD have also been studied. Topical application of ATP-dependent K+ channel (KATP) inhibitor glibenclamide augmented the peak hyperemia by ~30% in two studies, without altering the resting CBF or the DC shift in rats (200, 426). KATP channels are activated and hyperpolarize the cells when ATP/ADP levels drop, such as during hypoxia or hypoglycemia, and during SD, and are expected to increase CBF. Therefore, KATP inhibition is expected to promote depolarization, and although the target cell type (i.e., neurons versus vasculature) responsible for glibenclamide effect is unclear, it is unlikely to be the smooth muscle. In contrast, inhibition of mitochondrial KATP by 5-hydroxydecanoate did not alter SD-induced CBF changes in rats (200).

There are much less data on voltage-gated Na+ and Ca2+ channels. In one study, voltage-gated Na+ channel inhibitor tetrodotoxin, which does not suppress SD, reduced resting CBF, augmented initial hypoperfusion, increased the peak hyperemia relative to the reduced baseline, and augmented post-SD oligemia (114). and lastly, L-type Ca2+ channel inhibitors lomerizine and flunarizine have been reported to dose-dependently attenuate post-SD oligemia when administered systemically (425). The effect, however, may be nonspecific via direct smooth muscle relaxation, as these drugs also elevate resting CBF and inhibit pharmacological vasoconstriction.

Cerebral arteries and arterioles are responsive to perivascular pH (245). Acid pH dilates and alkaline pH constricts pial arterioles (246, 310, 509). During SD there is a brief extracellular alkaline transient (~0.1 pH unit lasting a few seconds) at the onset of depolarization, and this is followed by a larger and longer-lasting extracellular acidification (~0.3–0.4 pH unit lasting a few minutes); the latter is indeed capable of dilating the pial arteries by ~20%, and increase CBF by ~100%. Although these pH shifts could contribute to the initial hypoperfusion and the subsequent hyperemia evoked by SD (114), CBF changes showed no consistent relationship to pH changes in one earlier study (487). Superfusion with artificial CSF, which abolished the hypercapnic vasodilation presumably by buffering the pH changes, did not diminish the pial dilator response to SD (420, 421). Although these data argue against a direct role of pH in the pial arterial response to SD, a contribution at the parenchymal arteriolar level cannot be excluded. Another plausible mechanism by which high levels of external lactate (68) may contribute to the hyperemia independent of pH is inhibition of transporter-mediated uptake of vasodilator prostaglandin E2 from the extracellular space (159, 371).

2. NO

Owing to its rapid and potent vasodilator effect, NO has arguably been the most intensely investigated molecule as a potential mediator of the CBF response to SD. During physiological neurovascular coupling, neuronal NOS-derived NO acts as a permissive factor in the cortex (i.e., blocking effects of NOS inhibition are reversed by NO donors), and as a mediator in the cerebellum (3, 525). Glutamate release, NMDA receptor activation, and massive Ca2+ influx during SD are predicted to stimulate neuronal NOS, which is attached to the NMDA receptor via the scaffolding protein, postsynaptic density protein 95 kDa (408). NMDA receptor activation stimulates NO release (121, 317, 347). Stimulation of NO synthesis and release during SD is indirectly supported by a drop in l-arginine levels and a rise in cGMP, the downstream effector (117, 143). Direct polarographic measurements during SD in rats and cats have indeed been reported to detect a sharp two- to threefold increase in tissue NO levels coincident with the DC shift, followed by a sharp decrease that coincided with the hyperemia; NO levels then decreased to a post-SD plateau lower than baseline for at least 10 min (382, 383, 385, 387). Importantly, peak NO increases were gradually reduced upon repeated SDs, unlike the gradually augmented hyperemic response to recurrent SDs (385). Unfortunately, the results showing NO release have not been reproduced by other groups, and it is still uncertain to what extent tissue NO rises during SD. In addition to cortical interneurons, perivascular nerves, and the endothelium may contribute NO during SD (112, 204, 205, 348).

Inhibitors NOS have been tested to dissect how NO mediates or modulates the CBF response to SD, with highly conflicting results. Given the permissive role of NO for vascular reactions in the cerebral cortex, it is not surprising that the effect of NOS inhibition during SD is ambiguous. In some studies, systemic NOS inhibitor N-nitro-l-arginine methyl ester (l-NAME) completely abolished the CBF response to SD in cats, without altering the single unit burst firing that was used to monitor SD occurrence, but the lack of proper control experiments make it difficult to assess the data (154, 385). As expected, l-NAME increased systemic blood pressure, reduced resting NO levels, and significantly diminished the biphasic NO response to SD (385). Although l-NAME has antimuscarinic effects that may be confounding, N-nitro-l-arginine (l-NA), which is a more specific NOS inhibitor, also reversibly suppressed SD-induced peak hyperemia by up to 40–70% when applied topically or systemically in rabbits (65, 66, 314, 315). Moreover, 7-nitroindazole (7-NI), a relatively specific inhibitor of neuronal NOS in vivo, also reduced the peak hyperemia to SD by 35%, without a significant change in resting CBF or systemic blood pressure, implicating a neuronal origin for NO (66). Lastly, NOS inhibition also attenuated NMDA-induced dilations (121), but the dilations could in part be due to SD in that model (16). Several other studies failed to detect any effect of NOS inhibition on the CBF response to SD. For example, the same group that found potent inhibition of hyperemia by l-NAME in cats failed to show an effect in rats using an otherwise similar experimental model, despite complete inhibition of the increase in NO levels during SD (382). Other rat studies also failed to detect any inhibition of peak hyperemia after systemic or topical l-NA or l-NAME; peak hyperemia to repeated SDs was in fact even larger (118, 427, 428, 520, 534). NOS inhibition was verified by direct enzyme assays as well as by absent endothelium-dependent vasodilation and attenuation of hypercapnia-induced rises in CBF in a subset of these studies, and others that have used similar doses and administration routes observed similar lack of effect on CSD-induced increases in CBF. Therefore, insufficient enzyme inhibition or systemic versus topical administration do not appear to explain the discrepant results. Besides the obvious possibility of experimental pitfalls and caveats, one potential source of variation is the fact that NOS inhibition can reduce resting CBF by up to 10–50%, and the magnitude of peak hyperemia critically depends on whether preinhibition or pre-SD CBF is taken as the denominator. This, however, does not explain all variability in reported effects. There is more consistent data suggesting that NO counteracts the vasoconstrictive tone that develops during SD. NOS inhibition augmented the initial hypoperfusion and diminished the peak hyperemia without significantly altering the other vasomotor components of CBF response described earlier (Figure 7, B and andCC) (86, 94, 114, 445). This effect was subsequently confirmed in mice (17). Altogether the data suggest that NOS inhibition augments the initial hypoperfusion, which would be consistent with the sharp increase in tissue NO levels during the DC shift.

Indeed, resting NO levels may modulate the cerebrovascular response by augmenting the effect of other vasodilators, and diminishing the effect of vasoconstrictors released during SD (87, 322). For example, NOS inhibition, or NO scavenging using hemoglobin, severely augmented the vasoconstrictive response to elevated [K+]o (35 mM) to the extent that spontaneous and prolonged SDs developed that caused severe vasoconstriction which reduced the perfusion level to <20% of baseline for up to an hour (86). SDs induced during topical l-NA with elevated [K+]o (35 mM) evoked a severe and 5- to 10-fold prolonged hypoperfusion coupled to a similarly prolonged DC shift; peak hyperemia was also diminished and prolonged (90). These changes were partially reversed by NO donor S-nitroso-N-acetylpenicillamine (SNAP), Ca2+ channel blocker nimodipine, and to a lesser extent direct vasodilator papaverine (86, 90). Similarly, in the absence of NO, extracellular ionic changes that cerebral blood vessels are exposed to during SD caused vasoconstriction in isolated middle cerebral arteries, in vitro (518). NO may also hinder vasoconstriction by inhibiting cytochrome P (CYP)-450 ω-hydroxylase activity, which catalyzes the conversion of arachidonic acid (AA) to 20-hydroxyeicosatetraenoic acid (20-HETE) (328, 395, 459), as described below.

These data clearly suggest that ambient NO modulates vasoconstriction during SD. In support of a permissive role for NO in SD, application of the NO donor SNAP or cell-permeable cGMP analog 8-bromo-cGMP, to mimic steady resting levels of NO availability, restored the dilations in response to elevated [K+]o, after NOS inhibition had transformed the response into constriction in isolated cerebral arteries, in vitro (156). Lastly, topical or systemic pretreatment with l-arginine, the substrate for NO synthesis, ameliorated the post-SD oligemia in rats, and enhanced the late hyperemia (component III) that followed the peak hyperemia (Figure 7C) (114). These observations point to a functional downregulation of the NO-cGMP pathway following SD. The central role for NO in the protracted vascular changes following SD is further corroborated by the observation that topical application of NO donors or 8-Br-cGMP restored resting CBF during post-SD oligemia to values measured before SD (410). In contrast, systemic glyceryl trinitrate infusion did not alter the CBF response to SD, despite apparently enhanced NO availability at resting state and increased production during SD (386).

3. Arachidonic acid metabolites

Prostanoids are abundantly produced by astrocytes and possibly other cells types in the neurovascular unit, such as pyramidal neurons that express cyclooxygenase-2 (331). SD is a well-documented trigger for rapid arachidonic acid production, presumably due to activation of phospholipase A2 via elevated [Ca2+]i through NMDA receptors (262). Perivascular prostaglandin E2, I2, and F levels increase after a single SD by 30–90%; the increase is even larger after multiple consecutive SDs (50–300%), which also increases thromboxane B2 (421, 422). Importantly, vasoconstrictor prostaglandin F shows the largest increase. Although the time course of prostanoid release is not known, it is likely to be slower than the rapid hyperemic components of the hemodynamic response. Indeed, although systemic indomethacin can completely abolish the prostanoid release during SD, it does not diminish the peak hyperemia at doses that achieve complete inhibition of cerebrovascular cyclooxygenase, making this component unlikely to be mediated by prostanoids (257, 426428). To the contrary, indomethacin markedly augmented the pial arteriolar dilation and the peak hyperemia to SD in several studies, not fully explained by the reduction in resting diameters and CBF induced by indomethacin (95, 315, 419, 422). In addition to the augmented peak hyperemia, indomethacin abolished post-SD vasoconstriction (421).

More recently, lipoxygenase products have been investigated, because of their vasoactive properties, and because their CSF levels appear to be linked to delayed cerebral ischemia and poor outcomes after aneurysmal SAH (67, 222) where injury depolarizations akin to SD occur spontaneously (93). Within the neurovascular unit, arachidonic acid released from astrocytes by Ca2+-activated phospholipase A2 is converted to its main vasoconstrictor metabolite 20-HETE by a cytochrome P-450 enzyme (CYP4A) in vascular smooth muscle (328). Indeed, 20-HETE levels doubled after SD in brain slices, peaking within 20 min and lasting up to 2 h, congruent with the time course of post-SD oligemia (131). The prolonged 20-HETE elevation is interesting in light of the much shorter [Ca2+]i and arachidonic acid rise and phospholipase A2 activation during SD, implicating an alternative mechanism such as binding of the phospholipid derivative 20-HETE to an intracellular store (e.g., fatty acid binding protein) during SD-induced synthesis (517), followed by slow but sustained release after SD. When tested in vivo in rats, HET0016, a specific inhibitor of 20-HETE synthesis (222, 323), blocked the increase in 20-HETE levels and significantly ameliorated post-SD oligemia by more than half, without altering the resting CBF, Po2, and CMRO2, or changes in these parameters during the SD (131). These data suggested that an astrocyte-dependent mechanism might cause the post-SD oligemia. Importantly, 20-HETE levels increase by 10-fold in CSF after subarachnoid hemorrhage in rats (222), and by 6-fold in rat cerebral arteries during hypertension (144), underscoring the broad physiological impact of this vasoconstrictor metabolite.

Altogether these data suggest that both cycloxygenase and lipoxygenase products of the arachidonic acid metabolism derived from astrocytes exert a vasoconstrictive influence during and after SD. In support of this, blocking arachidonic acid production by an inhibitor of phospholipase A2 markedly diminishes astrocyte-mediated constriction of parenchymal arterioles during the depolarization, as shown by in vivo two-photon microscopy (59). Lastly, a direct interaction between NO and arachidonic acid metabolites in cerebrovascular regulation has been proposed (395, 459). In one study where such an interaction was sought during SD, NOS inhibition diminished the peak hyperemic response, and contrary to the expectations, this was prevented by indomethacin (315). The importance of these findings, however, is not clear.

4. Peptide neurotransmitters

Calcitonin gene-related peptide (CGRP) is a potent, endothelium-independent vasodilator that is present in perivascular trigeminal sensory nerve endings in experimental animals, and more sparsely in humans (102, 151, 281). Cortical CGRP release has been detected during exposure to elevated [K+]o (26–40 mM) both in vivo in anesthetized rats and in vitro in rat brain slices (26 mM) (100, 483). Topical CGRP antagonist (CGRP8–37) reversibly diminished SD-induced pial artery vasodilation by ~40%, albeit at relatively high concentrations, without altering resting diameter in rats, rabbits, and cats (64, 389, 513). When administered after chronic sensory denervation, CGRP8–37 was ineffective, confirming its specificity (389). Importantly, CGRP antagonists appeared to inhibit SD-induced intrinsic optical signals in one study (483), while others have reported no change in propagation speed, in vivo (64, 513). Therefore, whether CGRP modulates SD remains to be studied using standard electrophysiological methods.

Endothelin is another vasoactive peptide that is primarily released from the endothelium with potent constrictive as well as neuronal effects. Endothelin receptors are amply expressed by vascular smooth muscle, endothelium, neurons, and glia. Although endothelin is capable of triggering SD, in vivo, it fails to do so in brain slices, in vitro (85), suggesting a perfusion-dependent effect likely vasoconstrictive ischemia. Blockade of endothelin receptors did not alter the cerebrovascular response to SD, arguing against a role for this mediator in the initial hypoperfusion and post-SD oligemia components of the CBF response to SD (153).

5. Small molecule neurotransmitters

Acetylcholine dilates cerebral vessels via endothelium-dependent and -independent mechanisms largely via muscarinic receptors when applied topically (247, 313, 316, 414). In one study, muscarinic receptor antagonist atropine diminished SD-induced peak hyperemia by ~40%; atropine inhibition was largely prevented after chronic trigeminal and parasympathetic denervation (389). In another study, atropine was reported not to change the CBF response to SD (95). Serotonin (5-HT) is an indoleamine neurotransmitter with complex vasomotor properties, including endothelium-dependent vasodilation via 5-HT2c receptors (12). Central serotoninergic nerves also project onto pial and parenchymal arteries and may play a role in the cerebrovascular response to SD (101). The 5-HT2A/2C antagonist ritanserin reportedly decreased SD-induced peak hyperemia by ~40%, but also prolonged the hyperemia duration as well as the DC shift, without significantly altering the hypercapnic hyperemia in rats (155). In contrast, 5-HT1D/1B agonist zolmitriptan augmented the peak hyperemia amplitude albeit at very high doses, and did not affect the DC shift. Adenosine, a purine metabolite of ATP, markedly increases in response to hypoxia and metabolic activation, as well as during SD (217, 277, 400). Although adenosine is a vasodilator when applied topically (510), adenosine A1 receptor inhibitor 8-cyclopentyl-1,3-dipropylxanthine (DPCPX) has been reported to augment rather than inhibit SD-induced peak hyperemia (200). However, changes in baseline flow may in part be responsible for the apparent effect. Much less is known on the involvement of norepinephrine and epinephrine in the cerebrovascular response to SD. Perivascular application of these biogenic amines causes dose-dependent pial constriction via α-adrenergic receptors (244, 511), and in one report, the β-adrenergic receptor blocker propranolol did not alter the CBF response to SD (95). Because glutamate receptor inhibitors suppress SD, it has been difficult to assess its role in the cerebrovascular response (261). It is well known that glutamate receptor activation leads to vasodilation (126). This could be mediated by parenchymal depolarization and neuronal NOS activation via NMDA receptors (121, 122, 317), rather than being a direct vasomotor effect (181), although activation of non-NMDA (i.e., AMPA) subtype of glutamate receptors may be involved as well. Indeed, AMPA receptor activation is the main mechanism in neurovascular coupling during normal physiological activity, NMDA receptors are only involved during strong stimulation (199). Future studies will need to examine if vascular responses during SD involve glutamate receptors other than NMDA receptors. Lastly, it should be remembered that glutamate agonists can trigger SD (266), which may in part contribute to the CBF responses (16).

6. Free radicals

Free radical production by glial cells has been demonstrated during the post-SD period (164, 165, 503), although not all studies agree (86). A few studies have tested free radical scavenger antioxidants (e.g., tirilazad, superoxide dismutase, oxypurinol) on pial artery diameter and CBF responses to SD and did not observe an effect (96, 152, 314), although post-SD oligemia was reportedly prevented by tirilazad in one study (171). It should be noted that under artificial conditions creating a spreading ischemia pattern (i.e., reduced NO availability and elevated [K+]o), SD was associated with significant free radical production, presumably a result of tissue hypoxia and ischemia (86). However, its significance for the hemodynamic response is unclear.

7. Mitochondrial permeability transition pore

SD has many similarities to cerebral ischemia: during SD, Ca2+ accumulates in neurons and astrocytes (23, 59), free fatty acids are released (262), and neuronal mitochondria depolarize (536). However, as noted above, there are divergent views on oxygen free radical formation during SD, which is a hallmark of brain damage (86), and recurrent waves of SD do not induce irreversible neuronal injury in otherwise normal rat brain (334). Nevertheless, Ca2+ accumulation and free fatty acid release all favor formation of the mitochondrial permeability transition pore (mPTP), an important mechanism of mitochondrial dysfunction and apoptosis (135). Blockade of mPTP prior to, during, or following episodes of brain ischemia can rescue mitochondrial function, O2 metabolism, and tissue integrity, suggesting that these signaling pathways contribute importantly to impairment of function in models of acute brain diseases (135, 458). Whether mPTP modulates the generic features of SD and its impact on cerebral vasculature has recently been tested (373). The mPTP blockers cyclosporin A, and its more specific derivative NIM811 (514), did not affect the electrophysiological or metabolic features of SD, but augmented the peak hyperemia and ameliorated post-SD oligemia (373). Neither resting CBF nor thromboxane-induced vasoconstriction was altered, ruling out a nonspecific dilator effect. The findings suggest that a transient opening of mPTP may occur in a small population of highly Ca2+-sensitive mitochondria, most likely in interneurons (189) with slow recovery of function. This mechanism may explain part of the vascular impairment following SD (373).

C. Vascular Propagation of Vasomotor Responses

Intercellular gap junctions (e.g., connexin 40) directly couple endothelial and smooth muscle cells to allow electrotonic and chemical spread of vasomotor signals along the vessel length as a mechanism to induce adaptive changes upstream to the activated segment (201, 338, 484, 526). Given the fact that parenchymal substances released during SD are much diluted by the time they reach the pial vasculature (e.g., [K+]o ~4–5 mM) (420), and that superfusion with artificial cerebrospinal fluid, which inhibits hypercapnic pial vasodilation, fails to alter the response to SD (420), it is likely that pial vasomotor responses observed during SD are at least in part conducted upstream from the penetrating arterioles, or that the main mechanisms occur at a depth that is not affected by changing the superfusion fluid. Indeed, vascular conduction of vasomotor responses, mostly dilator, has been suggested during SD in both mice and rats (38). The conducted arteriolar response was ~500 μm and 10 s ahead of the depolarization wave front, independent of the baseline vessel diameter or the direction of SD propagation in relation to the blood flow direction, and never skipped across segments or from one artery to another, arguing against endothelial shear stress or conduction via perivascular nerves as propagation mechanism.

VI. IMPAIRED CEREBROVASCULAR REACTIVITY AFTER SPREADING DEPRESSION

SD not only has direct vasomotor effects, but it also influences physiological cerebrovascular responses. In the post-SD period vascular responses to most if not all physiological and pharmacological stimuli are markedly blunted. For example, vasoreactivity to perivascular pipette application of acidic or basic solutions (pH 6.87–7.57), varying concentrations of K+ (0–6 mM), adenosine, and bradykinin, was greatly diminished for at least 75 min after an SD when studied by intravital microscopy in rats and cats (512). Because solutions were applied locally onto the pial arterioles, diminished reactivity was likely due to changes within the vessels, rather than the parenchyma. Indeed, isolated vessels harvested after they were exposed to an SD in situ showed much diminished reactivity to pH and [K+]o changes (415). Even the response to the direct vasodilator papaverine appeared to be diminished after an SD in one study (130), although responses to acetylcholine, CGRP, NMDA, KATP channel opener aprikalim, and adenylate cyclase activator forskolin were reportedly preserved after SD in rabbits (46, 80), suggesting that the vasomotor paralysis may not apply to all species or mediators.

A single SD impairs both resting and stimulus-evoked neurovascular coupling. In the wake of a cortical SD in the rat, both CMRGlu and CMRO2 are preserved or increased when CBF is reduced by 30% or more, indicating flow-metabolism uncoupling at resting state (258, 373). Similar findings have been reported after a single SD in the cat using PET imaging (242), and using optical imaging in the mouse (54), where CMRO2 becomes O2 supply-limited in part due to the pronounced post-SD oligemia (532). Interestingly, neurovascular uncoupling has also been reported in subcortical structures (e.g., hippocampus) after a cortical SD, although it is not clear whether this was due to subcortical propagation of the cortical SD (327). In addition to the resting state supply-demand mismatch, cortical activation-induced hyperemic responses to transcallosal or somatosensory stimulation are also diminished after an SD, without a corresponding change in the electrophysiological or CMRO2 response, suggesting impaired neurovascular coupling with intact neurometabolic coupling (166, 371). A persistent reduction in the activity of interneurons containing vasoactive intestinal polypeptide (VIP) and NO may be a potential mechanism for the uncoupling (108, 239, 371). More recently, inhibition of mitochondrial permeability transition pore formation and calcineurin activation by cyclosporin, FK506, and NIM811 restored neurovascular coupling after SD, without altering the cortical electrical activity, resting CMRO2, or the CMRO2 increase during and after SD (373). A possible mechanism of action of cyclosporin and FK506 may be enhanced phosphorylation of neuronal NOS, which is required for normal neurovascular coupling (363), especially since data suggest a functional downregulation of NO/cGMP pathway after an SD (410). In contrast, inhibition of 20-HETE synthesis by HET0016 did not restore neurovascular coupling after an SD despite the fact that post-SD oligemia was ameliorated, suggesting that different mechanisms control resting CBF and activation-induced hyperemia during post-SD oligemia (131).

Hypercapnic hyperemia is also abolished or markedly diminished in the wake of an SD, starting as early as 15 min, and lasting anywhere between 20 min to 12 h in different studies and species (46, 114, 119, 130, 152, 248, 255, 376, 410, 533), although not all studies agree (80). Pretreatment with NOS substrate l-arginine, NO donors, and cell-permeable cGMP analog 8-Br-cGMP, but not papaverine, significantly restores hypercapnic hyperemia after SD (114, 410), suggesting a permissive role for NO. The data on blood flow autoregulation during blood pressure fluctuations are mixed. Both upper and lower limit of autoregulation were preserved, when sufficient time was allowed after the blood pressure drop (i.e., slow component) (255, 376). In contrast, the fast component of autoregulation that takes place within a minute after blood pressure reduction was completely abolished within 15 min after an SD, and took more than an hour to recover (130). Similarly, hypoxic hyperemia was diminished by 66% at 20–25 min after a single SD in one study (119), and unchanged in another (248). The hyperemic response to substantia innominata (basal forebrain) stimulation was almost completely abolished for up to 2 h (248). Lastly, spontaneous vasomotion (0.1 Hz) was also abolished immediately after SD, unrelated to systemic physiology (Figure 4H) (38, 166, 376).

Of note, diminished neurovascular coupling and hypercapnic hyperemia with relatively preserved autoregulation after an SD mimics the changes observed after controlled cortical impact, and focal or global ischemia in experimental animals, and traumatic brain injury and migraine attacks in humans (60, 109, 169, 265, 335, 413, 533). Although blunted physiological responses temporally overlap with post-SD oligemia, their mechanisms likely differ because they recover at different rates and show differential pharmacological sensitivity. Regardless of the mechanisms, blunted physiological and pharmacological responses after an SD have implications for experimental studies where invasive procedures (e.g., cranial window) can inadvertently trigger an SD during the preparation that may go unnoticed and confound subsequent testing of such responses, as previously pointed out by others (512, 532). In summary, post-SD period is characterized by disrupted vascular reactivity to all substances and conditions tested so far.

VII. BLOOD-BRAIN BARRIER AFTER SPREADING DEPRESSION

SD has also been shown to disrupt blood-brain barrier (BBB). A comprehensive examination of this phenomenon revealed that cortical SD increased MMP-9 gene expression within 1 h (168). MMP-9 (but not MMP-2) activity increased by more than 10-fold starting within 6 h, peaking at 24 h, and normalizing after 2 days. Enhanced MMP-9 activity was localized around blood vessels, detectable as soon as 15–30 min after an SD when examined in tissue sections. MMP-9 activation was associated with loss of basement membrane and tight junction proteins, and disruption of BBB. Indeed, SD caused plasma protein leakage into perivascular tissues detected as early as 3 h after SD and the BBB remained open for at least a day. Brain water content was also increased with a similar time course. Importantly, these changes were prevented by an MMP inhibitor and absent in MMP-9 knockout mice (168). In another study, 2 h of repetitive cortical SD induction increased the number of pinocytic vesicles in endothelial cells in the rat, which may be another mechanism contributing to BBB disruption after SD (529). However, not all studies agree with BBB disruption after SD. For example, an earlier study has failed to detect any plasma protein extravasation from pial arteries 4 h after SD (73), and more recently, plasma protein extravasation after 1 h of SD induction by topical KCl appeared to be a direct effect of KCl rather than SD itself as extravasation rapidly diminished as a function of distance to KCl application site, and was absent more than 5 mm away (356).

VIII. METABOLIC SUPPLY-DEMand MISMATCH DURING SPREADING DEPRESSION

As reviewed in detail above, SD markedly stimulates energy consumption and glycolytic and oxidative metabolic rates. Whether this creates a critical supply-demand mismatch and metabolic stress has long been debated. To that effect, SD induces hypoxia inducible factor 1α expression in rat cortex (504). Although exposure to numerous repetitive SDs does not cause acute injury in otherwise healthy brain with normal perfusion and oxygenation at baseline (334), long-term or cumulative effects are certainly possible. Two surrogates that have been studied as markers of metabolic stress are tissue O2 availability and mitochondrial redox state. The data altogether suggest that SD indeed causes O2 supply-demand mismatch in brain tissue, but the magnitude of mismatch may be small in tissue with normal perfusion at resting state and a hyperemic response to SD, and large-enough to create a hypoxic state in tissue with hypoperfusion at resting state and in the absence of a hyperemic response to SD.

A. O2 Availability

Most but not all studies show a decrease in O2 availability (i.e., Po2 or mixed arteriovenous hemoglobin O2 saturation) during the DC shift, sometimes preceded by a brief increase (20, 54, 131, 185, 248, 251, 285, 290, 304, 371, 406, 469, 486, 490, 491, 500). Reduced supply and/or increased demand may limit tissue O2 availability during SD (131, 371, 469, 532). While reduced O2 availability can be explained by vasoconstriction in some species and experimental models (54, 227, 406, 500, 532), similar Po2 dips were also observed in the absence of vasoconstriction and even in the presence of hyperemia, and were accompanied by increased tissue Pco2, thus implicating increased O2 use (131, 248, 371, 469). Upon repolarization, when the hemodynamic response becomes predominantly hyperemic, O2 availability may increase, which could suggest luxury oxygenation (18, 20, 89, 98, 248, 271, 406, 500, 520, 521, 523), even though some studies have reported tissue hypoxia (131, 371, 469). By luxury oxygenation or perfusion we mean an overabundant cerebral perfusion relative to the metabolic needs of the brain tissue (254).

The observation of “pink veins” (i.e., veins with a high hemoglobin O2 saturation) (269) poses an apparent paradox with low tissue Po2 during SD. One theoretical mechanism by which this could happen is increased capillary transit time heterogeneity, where the linear velocity of red blood cells in a subset of capillaries is too fast for O2 to be extracted efficiently (i.e., decreased O2 extraction fraction), leading to relative tissue hypoxia (215). The possibility remains to be tested.

When peak hyperemia is absent, Po2 increase is also absent (304, 532). In the post-SD period, tissue O2 availability remains significantly reduced for up to 2 h as a result of the post-SD oligemia that leads to a reduction in tissue Po2 (18, 54, 131, 185, 195, 371, 532). Despite this reduction in supply, CMRO2 is increased after an SD for up to 2 h. In summary, O2 delivery and availability can be reduced during the DC shift because of reduced delivery (initial hypoperfusion) and/or increased capillary transit time heterogeneity, and after CSD because of a reduced delivery (post-SD oligemia). In the mouse, severe initial hypoperfusion during the DC shift and subsequent post-SD oligemia are associated with marked increases in O2 extraction fraction, and reductions in O2 saturation as well as the CMRO2 measured by minimally invasive optical imaging tools (54, 532).

B. Mitochondrial Redox State

SD generates a rise in both aerobic and anaerobic metabolism, but it is less clear whether this results in a shortage of ATP in otherwise normal cortex and a change in the redox state (469). One way to address this is by monitoring the mitochondrial redox state. Monitoring NADH and FAD fluorescence has been commonly used to measure changes in mitochondrial redox state. The origin of NADH is largely cytosolic, while FAD fluorescence is exclusively mitochondrial (36, 249), the vast majority of which resides in neurons (216), and reflects the balance between the “reduced” and “oxidized” forms of these cofactors (488). Stimulation of oxidative phosphorylation, or elevating ADP levels, induces a shift from NADH (i.e., reduced) to NAD+ (i.e., oxidized), and from FADH2 to FAD, whereas stimulation of glycolysis and the tricarboxylic acid cycle shifts the balance in the opposite direction (488). Therefore, NADH and FAD fluorescence intensities change, respectively, with glycolytic and oxidative metabolic rates.

The reported patterns of NADH response to SD have been variable. While many studies detected a monophasic decrease in NADH fluorescence (i.e., oxidation) (252, 280, 296, 297, 299, 303, 305, 307, 397, 398, 444, 464), a monophasic increase, or biphasic responses with an initial decrease and a subsequent increase or vice versa, have all been reported in various species (50, 81, 185, 187, 304, 390, 442, 443, 532). A longer lasting NADH increase followed in the wake of SD in some studies, coinciding with the post-SD oligemia (50, 185, 532). Regardless of the direction of change, the NADH redox shifts were fairly large in magnitude, approaching the decrease induced by hyperbaric hyperoxia or mitochondrial uncouplers, or the increase (e.g., 2-fold) induced by anoxia or inhibitors of mitochondrial respiratory chain, in vivo (187, 296, 300, 301, 311, 390). Because ouabain significantly slows down the NADH oxidation rate during SD, stimulation of Na+-K+-ATPase is likely to be a significant contributor to the increase in energy consumption (252).

Variable NADH response patterns are a result of superimposed glycolytic and oxidative responses with variable magnitudes and time course. Upon depolarization, rapid glycolytic surge tends to increase NADH production, but with the onset of hyperemia during and after repolarization, oxidative phosphorylation increases NADH consumption. If the rate of glycolytic NADH production exceeds the rate of NADH consumption by the electron transport chain, the net result would be NADH accumulation. The direction of change during different phases is context sensitive, i.e., determined by the balance between O2 supply and demand, which is a function of the resting blood flow and metabolic supply-demand balance, as well as the blood flow response and the supply-demand reserve capacity, both of which can vary greatly under different experimental conditions (e.g., in vivo versus in vitro, species, presence or absence and type or depth of anesthesia, arterial Po2 and blood pressure, presence of ischemia). In the absence of O2 supply limitation, metabolic stimulation during SD decreases NADH (i.e., oxidation). If O2 supply is limited, NADH increase (i.e., reduction) becomes prominent (306, 397, 409, 445, 488, 532). Inhibition of NOS has a similar effect (445). In support of this, the vast majority of SDs that spontaneously occur in the partially ischemic peri-infarct penumbra tissue (i.e., PIDs) are associated with NADH increase and are transformed into NADH decrease only when the PID propagates into adjacent nonischemic tissue where O2 is not supply-limited (193, 452, 454).

It should be noted that NADH fluorescence measurements can be confounded by changes in tissue hemoglobin concentration (i.e., cerebral blood volume) and light scattering properties during the depolarization (i.e., transient cell swelling) (472), and both the magnitude and the direction of change in above-mentioned studies should be interpreted with caution, despite the standard correction by subtracting the reflected light intensity from the fluorescent signal. Two-photon microscopy, which is less prone to such confounders, has revealed a spatially heterogeneous NADH response during SD in mice (469). In the vicinity of capillaries, where O2 is presumably more available, NADH displayed a marked decrease (~20%), whereas farther away from the capillaries it increased (~25%). Following repolarization, all regions showed uniform and lasting NADH increases. In a more recent study using two-photon microscopy, NADH responses were again multiphasic and spatially heterogeneous, but did not show a relationship to individual capillaries (532). Nevertheless, the NADH response differed between the capillary bed and the tissue in the vicinity of a penetrating arteriole. Within the capillary bed NADH mainly increased during the depolarization and remained elevated throughout the imaging, whereas in the vicinity of an arteriole NADH was multiphasic but remained close to baseline, suggesting higher O2 availability from the nearby vessel. Altogether, the data from these two studies suggest that NADH changes during SD are spatially related to the local O2 availability, which appears to be dependent on the distance to arterioles in normal cortex, and distance to capillaries in hypoperfused cortex (405).

IX. ROLE OF VASCULATURE IN SPREADING DEPRESSION RECOVERY

Although it is believed that the mechanism of SD recovery is an energy-dependent cellular restoration of transmembrane ionic gradients by uptake and possibly spatial buffering, and that severe energy compromise delays and even prevents SD recovery and repolarization (147), there are a number of observations that are incongruent with this mechanism as well. First, reduced cerebral perfusion pressure by systemic hypotension or arterial stenosis markedly prolongs the DC shift (e.g., more than doubled at a blood pressure of ~40 mmHg) even when resting perfusion is maintained near normal levels (i.e., >90% of baseline) by autoregulation (105, 456, 457). In contrast, mild systemic hypoxia (Po2 ~45 mmHg) did not significantly prolong the DC shift as long as the arterial pressure and cerebral perfusion were maintained within a normal range (456, 457). Indeed, there was no correlation between tissue Po2 and the duration of the DC shift, and excess tissue O2 or glucose availability by systemic hyperoxia (arterial Po2 ~400 mmHg) or hyperglycemia (blood glucose ~400 mg/dl) did not restore prolonged DC shift durations to normal during hypotension (198, 457). Similarly, SD recovery in isolated retinal preparation is independent of O2 availability, albeit with substantially elevated glucose levels (501). Moreover, pharmacological interventions that diminish the hyperemic response to SD also prolong the DC shift (155, 353), and the durations of peri-infarct SDs are prolonged in hypoperfused penumbra despite relatively mild ischemia (202, 229). These data suggest that tissue perfusion per se may be critical for the restoration of ionic homeostasis. Because the recovery of SD is critically dependent on restoration of [K+]o, there may be mechanisms of K+ clearance that are not energy dependent (140142, 174). Astrocytic spatial buffering through intercellular junctions can be an efficient mechanism for clearance of small rises in [K+]o (i.e., <10 mM) during normal physiological activity that occur within small volumes of tissue (e.g., less than a few hundred microns) at any given point in time. However, the ability of this mechanism to spatially redistribute the [K+]o among astrocytic network is likely overwhelmed when a large strip of brain tissue (i.e., ~3 mm wide) is simultaneously exposed to over 40 mM of [K+]o during an SD. An alternative route of K+ dissipation may be the capillary network, which densely penetrates the brain tissue with an average intercapillary distance of ~50 μm. Therefore, the capillary network is the largest sink within the brain tissue that can contribute to the vascular clearance of [K+]o. Indeed, astrocyte end-feet tightly ensheath the capillaries and are ideally positioned to clear parenchymal [K+]o to the vasculature (129, 336, 366). Although the cerebrovascular endothelium is highly impermeable to K+ on the luminal side (70, 177), active or facilitated transport systems for K+ do exist on the abluminal endothelial membranes and get recruited to pump K+ into the capillaries upon elevated [K+]o (27, 28, 32, 107, 158). The mechanism becomes physiologically relevant at [K+]o that is reached during SD based on capillary K+ permeability of 3 × 10−7 cm/s and capillary area of 200 cm2/g (175, 177). In this way, a drop in capillary perfusion or an increase in capillary transit time heterogeneity, such as during hypotension, may impact cerebrovascular [K+]o clearance even if it does not create energy shortage. This may explain the transformation of CBF response from hyperemia to a biphasic response during hypotension (as discussed above), as well as in ischemic penumbra (see below). Reduced vascular [K+]o clearance together with an augmented vasoconstrictive tone and reduced intraluminal pressure to keep the vessels open may also explain the prolonged DC shift durations during hypotension in the absence of energy deficit (456, 457). In summary, direct vascular regulation of extracellular ionic milieu (i.e., vascular clearance of [K+]o) has implications for not only SD recovery and brain injury states, but also for migraine triggers (182). Lastly, it is worth noting that the perivascular space, suggested as a potential route of cerebral waste clearance through bulk CSF flow (210, 211), may also contribute to [K+]o clearance during SD. However, bulk CSF flow is limited to the perivascular rather than interstitial space (463), and passive diffusion is too slow to explain the dissipation of [K+]o during SD by this route (140, 142). Although SD-induced vasomotor and cell volume changes undoubtedly impact the perivascular space, the relevance of this phenomenon for SD recovery is unclear.

X. SPREADING DEPRESSION UNDER PATHOLOGICAL CIRCUMSTANCES

The complex cerebrovascular response to SD is further influenced by the physiological state of the tissue. Pathological states, such as reduced perfusion pressure, ischemia, and subarachnoid hemorrhage, all transform the hemodynamic response to SD from hyperemia into hypoperfusion (i.e., inversion) by augmenting the constrictive component I.

A. Focal Cerebral Ischemia

PIDs akin to SD occur with high frequency in focal cerebral ischemia in all species studied to date (34, 104, 332, 333). They arise in an ostensibly spontaneous fashion in the periphery of the infarct (333), in tissue that is functionally and metabolically compromised but not yet irreversibly damaged, the so-called ischemic “penumbra” (11). Recent data show that PIDs are indeed triggered by episodic drops in metabolic supply (e.g., hypoxic or hypotensive transients) or increases in metabolic demand (e.g., functional activation) that cause supply-demand mismatch transients in susceptible peri-infarct hot zones (506). Consequently, both hyperglycemia and hyperoxia can suppress PIDs (430, 454). Reduced tissue perfusion delays the recovery of DC shifts and [K+]o elevations during PIDs (35). Therefore, PIDs are prolonged in ischemic penumbra. However, when they propagate into normally perfused tissue, PIDs become indistinguishable from SDs in amplitude, duration, and propagation speed (115, 202, 319). Infarct expansion correlates with the number of PIDs (55, 282, 319, 467) or with their cumulative duration (77). SDs artificially evoked in the ischemic hemisphere accelerate infarct growth proportional to their number (19, 45), and suppression of PIDs can ameliorate injury (148, 206, 332, 431). The basis for this relationship may lie in the imbalance between metabolic work load and delivery of O2 and glucose, i.e., the CBF.

In focal ischemic penumbra, PIDs predominantly cause hypoperfusion secondary to vasoconstriction in all species studied to date, including mice, rats, and cats, as well as humans. In contrast, when PIDs propagate into the nonischemic tissue, they are associated with a peak hyperemia indistinguishable from SD. The deeper into the ischemic territory, the stronger the hypoperfusion response, and the more severe the loss of perfusion (26, 61, 218, 243, 284, 431, 450, 470). In mildly ischemic regions, the response is often biphasic with a conspicuous initial hypoperfusion (component I) followed by peak hyperemia (component II) (Figure 9). In more severely ischemic regions, the initial hypoperfusion becomes more severe, and peak hyperemia disappears. Therefore, the response becomes a monophasic hypoperfusion. Indeed, the loss of ability to mount a hyperemic response during the early stages of ischemia predicted tissue deterioration (128). The onset of hypoperfusion follows the onset of depolarization rather than precedes it, confirming that it is the effect of depolarization but not the cause (450). For the same reason, the steal phenomenon that has been proposed as one mechanism by which PIDs can worsen ischemic tissue perfusion (374) is unlikely to be a direct contributor (431). Adjacent to the ischemic core the hypoperfusion may become permanent if the depolarization fails to recover (197, 330, 431, 450), marking the stepwise expansion of ischemic core into the penumbra. Thus the vascular signature depends on the properties of the underlying tissue microenvironment. As such, PIDs are associated with more severe O2 supply-demand mismatch in ischemic penumbra compared with SD in normal tissue (20, 452, 522), and the oligemic response contributes to ischemic injury rather than being an epiphenomenon (330, 450). Pharmacological (e.g., NMDA receptor antagonist MK-801) or physiological (normobaric hyperoxia) suppression of PIDs improves tissue perfusion and affords neuroprotection in ischemic brain in mice (429, 431). Interestingly, the same has not been observed with MK-801 in rats possibly due to a lower PID frequency in rats than in mice, and systemic hypotension with MK-801 as a potential confounder (206, 362).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390009.jpg

CBF response to PIDs in focal ischemic brain. A hypothetical diagram showing the transformation of the CBF response (top graphs) to peri-infarct SDs (PIDs) recorded in tissue with increasing severity of ischemia as shown on the CBF profile across focal ischemic tissue (bottom graph). In nonischemic tissue, SD evokes a predominantly hyperemic response (A), whereas in mildly ischemic penumbra (B), a biphasic response is observed. In moderately ischemic penumbra (C), the response is mainly a monophasic hypoperfusion. In the severely ischemic core-penumbra junction (D), both the DC shift and the hypoperfusion may not recover completely after a PID. Horizontal bars represent the DC shift during PIDs.

The duration of hypoperfusion correlates directly with the duration of PID. Because more severe perfusion deficits in penumbra are associated with longer lasting depolarizations, their vasoconstrictive effect is also prolonged, which slows down and often prevents repolarization, creating a vicious cycle in penumbra. Indeed, when a PID fails to recover in critically hypoperfused tissue (i.e., when a portion of penumbra is incorporated into the core), its vasoconstrictive effect also becomes permanent. Since anoxic depolarization (AD) is a permanent (or at least a very long-lasting) SD-like state, it is associated with permanent loss of perfusion in focal ischemic core (431). This is also observed in global ischemia (175, 434). The mechanism may be relevant for the no-reflow phenomenon (7), possibly linked to the vasoconstrictive effect of elevated [K+]o that invariably accompanies SD, PID, and AD (7, 508). Pericytes may contribute (170). The permanent vasoconstrictive effect of AD may also explain the close correlation between MRI indexes of reduced cerebral blood volume and final tissue infarction in acute stroke in humans (446).

The exact mechanisms that transform the predominantly hyperemic CBF response into hypoperfusion in ischemic brain are unknown. By definition, focal arterial occlusion decreases the perfusion pressure in the territory supplied by the occluded artery. Intravascular perfusion pressure is a major determinant of the hemodynamic response to SD. As noted earlier, reduced perfusion pressure such as in systemic hypotension or proximal arterial stenosis or occlusion augments or unmasks the initial hypoperfusion (i.e., component I of the hemodynamic response) and diminishes the magnitude and prolongs the duration of subsequent dilator component (component II). Accordingly, hypotension transforms the CBF response in rats from monophasic hyperemia into a biphasic response (456, 457). Carotid artery occlusion also transforms the CBF response in the same manner (25) even when the resting CBF is maintained within normal range by autoregulatory mechanisms (207, 444, 445).

It is unclear, however, whether reduced luminal perfusion pressure in ischemic penumbra can alone explain the hypoperfusion response observed during PIDs, or whether tissue hypoxia or ischemia also independently transforms the hemodynamic response. In support of the latter, the response to SD also becomes vasoconstrictive after pial vein occlusion in rats, where perfusion pressure is not expected to decrease (361). In fact, CBF response to SD remains transformed and diminished at the infarct edge even 1 day after transient focal cerebral ischemia followed by successful reperfusion in rats, once again suggesting that perfusion pressure is not the only determinant of the transformation of response (428). Therefore, mechanisms other than simple reduction in intraluminal perfusion pressure appear to play a role as well. One can speculate that NOS activity may be diminished in ischemic tissue in part due to reduced O2 availability as a substrate for NO synthesis. This is supported by data showing that pharmacological NOS inhibition transforms the CBF response similar to that in ischemic tissue (17, 114). Alternatively, maximum autoregulatory vasodilation may be approached in ischemic penumbra, limiting any further dilation during SD and unmasking the vasoconstrictive component. Lastly, [K+]o plays a significant role in determining the morphology of the vascular response during injury depolarizations (246, 310). Resting [K+]o is often elevated in ischemic and otherwise injured brain tissue, which augments vasoconstriction during PIDs, as we have learned in the setting of subarachnoid hemorrhage.

B. Subarachnoid Hemorrhage

Although spontaneous SDs reminiscent of PIDs have not been reported in animal models of subarachnoid hemorrhage, conditions mimicking subarachnoid hemorrhage do transform the peak hyperemic response to induced SDs (86). For example, topical application of hemoglobin, presumed to act as an NO scavenger, transforms the response from a monophasic peak hyperemia into a biphasic response with an initial hypoperfusion followed by diminished hyperemia. When topical hemoglobin is combined with elevated resting [K+]o (20 mM), the hemodynamic response to SD transforms into a severe monophasic hypoperfusion (80% CBF decrease). The combination of topical hemoglobin and 35 mM resting [K+]o indeed triggers spontaneous SDs that are accompanied by severe reduction of CBF to ~15–20% of baseline, the so-called spreading ischemia that lasts anywhere from 10 min to an hour. The condition resembles the vicious cycle during PIDs in ischemic core and penumbra, i.e., initial hypoperfusion (component I) severe enough to interfere with repolarization (90). Interestingly, combination of topical hemoglobin or NOS inhibition with either low glucose solution (0.8 mM) or endothelin-1 (ET-1, 100 nM) also triggered spreading ischemia (84, 368); however, ETA or ETB receptor inhibitors did not prevent spreading ischemia triggered by NOS inhibition combined with 35 mM resting [K+]o (368). The ionic milieu during spreading ischemia, when reproduced in vitro, also constricts isolated MCA when NOS is inhibited (518). Spreading ischemia was proposed as a model for migrainous infarction, or in the pathophysiology of mitochondrial encephalopathy, lactic acidosis, and stroke (MELAS). Despite the severe transformation, the hemodynamic response to SD can be normalized by restoring [K+]o and washout of hemoglobin, by nimodipine coapplication (86), and partially by adenosine coapplication (92). Indeed, normalization of the microvascular hemodynamic response to SD by nimodipine may be one mechanism to explain its clinical benefit in preventing the delayed ischemic neurological deficits after subarachnoid hemorrhage, since large artery vasospasm was not reversed (475).

Subarachnoid blood has also been shown to directly transform functional neurovascular coupling in brain slices via increased amplitude of spontaneous astrocytic Ca2+ oscillations, augmenting the K+ efflux through BK channels both at rest and during neural activation (228). As a result, [K+]o approaches and exceeds the threshold at which its normal dilatory effect via Kir channels is transformed into a direct constrictive effect on vascular smooth muscle, effectively inverting the response. Whether the same mechanism occurs in vivo or transforms the hemodynamic response to SD after subarachnoid hemorrhage has not been tested.

C. Traumatic Brain Injury

Investigation of SD in the setting of trauma has been limited. Spontaneous SDs occurring after traumatic cortical injury were uncommon in one study and occurred mostly during the first hour of injury (505). The CBF response to such spontaneous SDs showed a prominent initial hypoperfusion during the depolarization and a diminished peak hyperemia. Interestingly, artificially (KCl) induced SDs after injury did not worsen the outcome in this model, setting a contrast with focal cerebral ischemia and suggesting compromised perfusion as a critical factor in SD-induced damage in injured brain.

XI. SPREADING DEPRESSION and CEREBRAL VASCULATURE IN HUMAN BRAIN

Although there is ample evidence indicating the occurrence of SD in human brain, data on the hemodynamic response to SD are limited to SD occurring under pathological conditions such as migraine and brain injury, which may not fully represent normal brain. Moreover, it is highly probable that recordings in subjects with brain injury have missed the first SD occurring in naive tissue, the hemodynamic response to which may significantly differ from the response to subsequent SDs. These potential caveats notwithstanding, a number of studies have attempted to characterize the CBF response to SD in human.

A. Migraine

SD is the pathophysiological substrate of migraine aura and a trigger for headache (13). Aura may be any neurological disturbance that appears shortly before or during the development of a migraine headache, and may have different features depending of the brain regions involved in SD (253, 256). Hemodynamic studies during migraine attacks have revealed a reduction in CBF, termed spreading oligemia, that spatiotemporally closely followed the development of neurological symptoms (263, 265, 267, 346, 350, 357, 436438). Spreading oligemia most commonly propagates from the occipital lobe into the parietal and temporal lobes at a rate of 2–3 mm/min, the maximal CBF decrease being above the ischemic level (i.e., 20–40% reduction). The original findings of a spreading oligemia have been confirmed with SPECT, PET, and functional MRI techniques (169, 264, 524). A delayed hyperemia (~20%) has also been reported following oligemia during headache resolution (~5 h after attack onset) in a small cohort of migraineurs (9), reminiscent of the delayed hyperemia following the post-SD oligemia reported in experimental animals (120). However, a preceding hyperemia during the aura stage, corresponding to the peak hyperemia (component II) in experimental animals, has been detected in only a small subset of migraineurs (169, 357). This may be because the hyperemia is more prominent in relative hypoperfusion states usually not present in humans (257), or because it is very brief and occupies a narrow strip of tissue (Figure 10) and may have been missed by the poor spatiotemporal resolution of clinically available noninvasive techniques, or partial volume averaging. It is of course possible that hyperemia may not be a prominent feature of the CBF response to SD in human brain, although studies from injured human brain suggest otherwise (see below). In regions showing oligemia during an attack, CO2 reactivity and functional hyperemia are attenuated (169), and autoregulation during blood pressure changes is preserved (4, 265, 267, 435), as the vascular signature of SD in humans, as it is in animals (255, 376).

An external file that holds a picture, illustration, etc.
Object name is z9j0031527390010.jpg

The wave-width of various physiological signals that serve as footprints of SD. The duration of each signal (min) and the cortical width (mm) that undergoes the specific changes at any given time are shown based on the known propagation rate of SD in normal cortex. Most electrophysiological and neuroimaging signatures occupy a strip of cortical tissue not more than several millimeters in width, a major factor that limits noninvasive clinical detection of SD.

Oligemia occurs during familial hemiplegic migraine attacks as well (214), often associated with more severe hypoperfusion approaching ischemic levels (136). However, the spatiotemporal evolution of oligemia did not show the typical spread pattern observed in classical migraine with aura. It was rather chaotic, appeared and disappeared faster and unpredictably, and was associated with superimposed large amplitude fluctuations in CBF sometimes showing hyperemia (136, 208). Recently, longitudinal imaging revealed a biphasic pattern with multifocal oligemia within hours of attack onset, and hyperemia after 18 h (209). Although an intrinsic vasomotor instability was suggested to explain these observations, enhanced SD susceptibility and propagation speed and the propensity for reentrant SD waves in this disease may also be responsible (103, 104, 106, 493, 494).

The obvious difficulty of capturing patients during an attack notwithstanding, angiographic vasospasm has not yet been demonstrated during oligemia. Its absence may implicate increased resistance at the pial or parenchymal arteriolar level (437, 512). Because pial vasculature represents an anastomosing network, it is difficult, based on local changes in pial arteriolar diameters, to produce persistent changes in CBF downstream from the brain surface by local pial vasoconstriction (423). In contrast, a slight constriction of penetrating arterioles embedded in cortical tissue through which the SD propagates can explain the reduction in CBF (344). Interestingly, elevated plasma endothelin-1 levels have been reported during the first few hours of migraine attacks with or without aura that may provide a mechanism for increased parenchymal cerebrovascular resistance (125, 139).

B. Brain Injury

For several decades, it was speculated whether PIDs occur in human ischemic stroke (176), and anecdotal evidence has appeared (302). The hypothesis was systematically tested using subdural electrodes in patients with malignant MCA infarcts, a condition that occurs in ~10% of patients with supratentorial infarcts (33) that necessitates hemicraniectomy because of risk for herniation (358). Numerous PIDs were detected over the course of 7 days after stroke in most if not all stroke patients, and more than 90% of the total of depolarization events occurred in temporal clusters (78). As in experimental animals, PIDs increase the infarct size in human brain, demonstrated in a patient with malignant stroke who showed lesion progression by sequential MRI exactly in the zone experiencing around a hundred PIDs over a period of 5 days (330). Intraoperative visualization of the human cerebral cortex in the peri-infarct region of malignant hemispheric stroke showed multiple PIDs with a mixed pattern of CBF responses. More than 70% of events were coupled to hyperemia, and a smaller subset showed biphasic CBF responses or a monophasic hypoperfusion (519). Altogether the types of CBF responses coupled to PIDs in this cohort of malignant hemispheric infarct were consistent with experimental data (283, 330, 431, 450), suggesting that PIDs can be associated with both protective and unfavorable hemodynamic responses.

A similar pattern of CBF changes ranging from monophasic transient peak hyperemia to monophasic permanent hypoperfusion were observed coupled to SDs after subarachnoid hemorrhage as well (88). Although arterial vasospasm is a much-feared complication of aneurysmal subarachnoid hemorrhage, whether arterial vasospasm alone can explain the delayed ischemic neurological deficits days after the initial bleeding has been unclear. Patients with aneurysmal subarachnoid hemorrhage have a high incidence of SD events, and there is a clear correlation between SD occurrence and development of delayed ischemic neurological deficits (93). Indeed, small cortical infarcts are more common than large territorial infarcts corresponding to major cerebral arteries after aneurysmal subarachnoid hemorrhage, suggesting that SD events might contribute to delayed ischemia (91). In one study of 13 patients with aneurysmal subarachnoid hemorrhage, SDs were observed in 12 patients, associated with an increase, no change, or a decrease in CBF, with corresponding changes in tissue Po2 (88). Consistently, progressively hypoxic responses developed during clusters of SDs in patients with delayed ischemic neurological deficits, which were explained both by changes in local CBF due to microcirculatory dysfunction and to augmented metabolism (31). SDs also occur with high frequency in patients with traumatic brain injury and intracerebral hemorrhage (115, 184, 451). Recent systematic investigation of human traumatic brain injury revealed heterogeneous hemodynamic responses to SD ranging from monophasic hyperemia to monophasic hypoperfusion, mimicking the response patterns in peri-infarct tissue as well as after subarachnoid hemorrhage (194), as has been noted in earlier small studies (302).

In summary, SDs have been shown to worsen tissue perfusion in injured human brain, much like in experimental animals, as a potential mechanism leading to detrimental outcomes in stroke, intracranial hemorrhage, and traumatic brain injury (355). This is not surprising because human cerebral arteries are more sensitive to the vasoconstrictive effects of elevated [K+]o compared with rat and rabbit arteries (402). Altogether the data suggest a bidirectional relationship leading to a vicious cycle in injured brain: injury radically transforms the neurovascular response to SD, and in return the transformed neurovascular response exacerbates the tissue injury by both decreasing CBF and metabolic supply and by increasing the energy demand (364). Therefore, SD should be considered a therapeutic target in injured brain (453).

XII. PRIMARY VASCULAR EVENTS TRIGGERING SPREADING DEPRESSION

Data over the past few years have strongly suggested a reverse direction of causality as well, where primary vascular events can lead to SD occurrence. For example, transient vascular microembolic events caused by air microbubbles, microspheres, or cholesterol microcrystals can all trigger SD without causing permanent brain injury (349). The mechanism involves brief ischemia severe enough to induce anoxic depolarization and has been implicated as a trigger for attacks of migraine with aura in patients with patent foramen ovale, particularly in the setting of agitated saline test during diagnostic echocardiography (418, 447). The theory has been challenged by the failure of patent foramen ovale closure in improving migraine in a pilot randomized trial (82), and lack of an association between patent foramen ovale and migraine in an older population (401). Another interesting possibility is the release of one or more factors from the vasculature to trigger or predispose to SD either by directly acting on neurons and glia, or by causing severe constriction and transient ischemia as above. Indeed, endothelin is a potent vasoconstrictor peptide released primarily by the endothelial cells, and when administered exogenously, it is capable of triggering SD, in vivo but not in vitro, likely via a constrictive or ischemic mechanism through ETA receptors (85, 223). It remains to be determined whether endothelial irritation can induce endothelin release as a potential vascular trigger of SD. Lastly, the clinical association between several genetic vasculopathies and migraine with aura (e.g., CADASIL, RVCL, or mutations in COL4A1) has raised the possibility of a primary vascular determinant of SD susceptibility or induction. However, at least in CADASIL, the phenotype appears to involve parenchymal hyperexcitability (105).

XIII. ON THE EVOLUTIONARY ORIGINS OF VASOCONSTRICTIVE RESPONSE TO SPREADING DEPRESSION

Teleologically, vasoconstriction to PIDs is a highly paradoxical response in ischemic brain as it leads to further loss of perfusion. One has to remember, however, that ischemic stroke is a disease of modern age, an exceedingly short time span in evolutionary history. This is because ischemic stroke is largely a disease of elderly and is relatively rare in young adults; although it is among the major causes of death and disability today, this was not the case a mere millennium ago. Back then and until the beginning of 20th century, human life was much shorter (average global life expectancy <30 yr), and the major form of intracranial pathology that humans were exposed to was presumably traumatic brain injury and hemorrhage. In fact, pediatric strokes are more often hemorrhagic (279). Even today, spontaneous strokes are quite rare in the majority of mammals and phylogenetically lower animals that have been studied in sufficient detail, suggesting that our brains have never been evolutionarily challenged by ischemic strokes to force the natural selection of resistant genotypes. As indicated above, PIDs are in many ways analogous to SDs observed after intracerebral or subarachnoid hemorrhage and traumatic brain injury, and in case of the latter it makes a lot more sense for cerebral vasculature to constrict and minimize hemorrhage, rather than dilate to improve brain perfusion which would be more appropriate in ischemic stroke. Our brain simply does not know the difference between a PID in ischemic brain and an SD in traumatized brain. Because traumatic injury had more evolutionary influence, brain tissue assumes the latter and reacts as such when the tissue homeostasis is disturbed in some way (e.g., ischemia). Even in the normal brain (i.e., migraine), the invariable component of the cerebrovascular response to SD is post-SD oligemia that lasts up to an hour and is observed across species; the peak hyperemia during SD, which has traditionally attracted more attention, are highly variable in direction as well as magnitude, and last only a couple of minutes. It appears from these observations that vasoconstriction is the more biologically and clinically relevant and deleterious response to intense brain depolarizations.

XIV. Summary and Conclusions

SD represents a massive pandepolarization of brain tissue with profound vascular and metabolic consequences. It is an evolutionarily conserved intrinsic ability of neuronal membranes to undergo such depolarizations upon intense excitation or metabolic failure. We here provide a detailed overview of the complexity of cerebrovascular regulation during and after SD. The multitude of vasoactive substances simultaneously released from neurons and astrocytes makes it one of the most difficult vasomotor responses to dissect physiologically and pharmacologically. We have, therefore, refrained from contrived and reductionist simplifications of the vasomotor actions of each potential mediator or modulator, and instead provided a detailed summary of our knowledge base for the reader to draw upon. The image we would like to impress upon the reader is one of chaos, where countless cells and signaling pathways are simultaneously recruited and perturbed. Our job is then to seek targets that can be utilized to better understand the neurovascular processes and alter the course of events during SD. The topic is not only of theoretical but also clinical relevance since SD occurs in patients with migraine and acutely injured cortex, with severe outcome implications in the latter. The goals of ongoing neurovascular research in SD are to suppress the depolarization events, to facilitate substrate supply at a time when the tissue depolarizes, and to prevent the post-SD oligemia that may relate to vasospasm after SAH. The search for vasoconstrictor mechanisms is ongoing and of great importance for our understanding of the neurovascular function and the approach to treat neurovascular disorders. Our knowledge of the loss of vascular reactivity after SD and the persistent impairment in neurovascular coupling is also incomplete and of similar importance. New research tools, such as optical imaging, will be essential and are likely to yield critical data on the contribution of SD to neurovascular disorders, and facilitate new concepts and drugs to mitigate the vascular consequences of SD.

GRANTS

C. Ayata was supported by the National Institutes of Health Grants NS061505;, NS055104;, and NS35611; Fondation Leducq; The Heitman Foundation; and The Ellison Foundation. M. Lauritzen was supported by the Lundbeck Foundation, the NOVO-Nordisk Foundation, the Nordea Foundation for the Center for Healthy Aging, Foundation Leducq, and the Danish Medical Research Council.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

ACKNOWLEDGMENTS

We thank Dr. Jed Hartings, Dr. Qingming Luo, Dr. Wolfgang Hanke, and Dr. Eszter Farkas for kindly providing figures.

Address for reprint requests and other correspondence: C. Ayata, 149 13th St., Rm. 6403, Charlestown, MA 02129 (e-mail: gro.srentrap@atayac) or M. Lauritzen, Dept. of Neuroscience and Pharmacology, Panum Institute, Univ. of Copenhagen, Blegdamsvej 3, 2200-DK Copenhagen, Denmark (e-mail: kd.uk.dnus@tirualm).

REFERENCES

1. Abrams RM, Stolwijk JA, Hammel HT, Graichen H. Brain temperature and brain blood flow in unanesthetized rats. Life Sci 4: 2399–2410, 1965. [Abstract] [Google Scholar]
2. Akerman S, Holland PR, Goadsby PJ. Mechanically-induced cortical spreading depression associated regional cerebral blood flow changes are blocked by Na+ ion channel blockade. Brain Res 1229: 27–36, 2008. [Abstract] [Google Scholar]
3. Akgoren N, Fabricius M, Lauritzen M. Importance of nitric oxide for local increases of blood flow in rat cerebellar cortex during electrical stimulation. Proc Natl Acad Sci USA 91: 5903–5907, 1994. [Europe PMC free article] [Abstract] [Google Scholar]
4. Akin A, Bilensoy D, Emir UE, Gulsoy M, Candansayar S, Bolay H. Cerebrovascular dynamics in patients with migraine: near-infrared spectroscopy study. Neurosci Lett 400: 86–91, 2006. [Abstract] [Google Scholar]
5. Albaum HG, Noell WK, Chinn HI. Chemical changes in rabbit brain during anoxia. Am J Physiol 174: 408–412, 1953. [Abstract] [Google Scholar]
6. Alkayed NJ, Birks EK, Narayanan J, Petrie KA, Kohler-Cabot AE, Harder DR. Role of P-450 arachidonic acid epoxygenase in the response of cerebral blood flow to glutamate in rats. Stroke 28: 1066–1072, 1997. [Abstract] [Google Scholar]
7. Ames A 3rd, Wright RL, Kowada M, Thurston JM, Majno G. Cerebral ischemia. II. The no-reflow phenomenon. Am J Pathol 52: 437–453, 1968. [Europe PMC free article] [Abstract] [Google Scholar]
8. anderova M, Kubinova S, Mazel T, Chvatal A, Eliasson C, Pekny M, Sykova E. Effect of elevated K+, hypotonic stress, and cortical spreading depression on astrocyte swelling in GFAP-deficient mice. Glia 35: 189–203, 2001. [Abstract] [Google Scholar]
9. andersen AR, Friberg L, Olsen TS, Olesen J. Delayed hyperemia following hypoperfusion in classic migraine. Single photon emission computed tomographic demonstration. Arch Neurol 45: 154–159, 1988. [Abstract] [Google Scholar]
10. andersen BJ, Marmarou A. Functional compartmentalization of energy production in neural tissue. Brain Res 585: 190–195, 1992. [Abstract] [Google Scholar]
11. Astrup J, Siesjo BK, Symon L. Thresholds in cerebral ischemia: the ischemic penumbra. Stroke 12: 723–725, 1981. [Abstract] [Google Scholar]
12. Auer LM, Leber K, Sayama I. Effect of serotonin and its antagonist ketanserin on pial vessels. J Cereb Blood Flow Metab 5: 517–522, 1985. [Abstract] [Google Scholar]
13. Ayata C. Cortical spreading depression triggers migraine attack: pro. Headache 50: 725–730, 2010. [Abstract] [Google Scholar]
14. Ayata C. Spreading depression and neurovascular coupling. Stroke 44: S87–S89, 2013. [Abstract] [Google Scholar]
15. Ayata C. Spreading depression: from serendipity to targeted therapy in migraine prophylaxis. Cephalalgia 29: 1095–1114, 2009. [Abstract] [Google Scholar]
16. Ayata C, Moskowitz MA. Cortical spreading depression confounds concentration-dependent pial arteriolar dilation during N-methyl-d-aspartate superfusion. Am J Physiol Heart Circ Physiol 290: H1837–H1841, 2006. [Abstract] [Google Scholar]
17. Ayata C, Shin HK, Salomone S, Ozdemir-Gursoy Y, Boas DA, Dunn AK, Moskowitz MA. Pronounced hypoperfusion during spreading depression in mouse cortex. J Cereb Blood Flow Metab 24: 1172–1182, 2004. [Abstract] [Google Scholar]
18. Ba AM, Guiou M, Pouratian N, Muthialu A, Rex DE, Cannestra AF, Chen JW, Toga AW. Multiwavelength optical intrinsic signal imaging of cortical spreading depression. J Neurophysiol 88: 2726–2735, 2002. [Abstract] [Google Scholar]
19. Back T, Ginsberg MD, Dietrich WD, Watson BD. Induction of spreading depression in the ischemic hemisphere following experimental middle cerebral artery occlusion: effect on infarct morphology. J Cereb Blood Flow Metab 16: 202–213, 1996. [Abstract] [Google Scholar]
20. Back T, Kohno K, Hossmann KA. Cortical negative DC deflections following middle cerebral artery occlusion and KCl-induced spreading depression: effect on blood flow, tissue oxygenation, and electroencephalogram. J Cereb Blood Flow Metab 14: 12–19, 1994. [Abstract] [Google Scholar]
21. Balestrino M, Young J, Aitken P. Block of (Na+,K+)ATPase with ouabain induces spreading depression-like depolarization in hippocampal slices. Brain Res 838: 37–44, 1999. [Abstract] [Google Scholar]
22. Bari F, Paprika D, Jancso G, Domoki F. Capsaicin-sensitive mechanisms are involved in cortical spreading depression-associated cerebral blood flow changes in rats. Neurosci Lett 292: 17–20, 2000. [Abstract] [Google Scholar]
23. Basarsky TA, Duffy SN, andrew RD, MacVicar BA. Imaging spreading depression and associated intracellular calcium waves in brain slices. J Neurosci 18: 7189–7199, 1998. [Abstract] [Google Scholar]
24. Benveniste H, Drejer J, Schousboe A, Diemer NH. Regional cerebral glucose phosphorylation and blood flow after insertion of a microdialysis fiber through the dorsal hippocampus in the rat. J Neurochem 49: 729–734, 1987. [Abstract] [Google Scholar]
25. Bere Z, Obrenovitch TP, Bari F, Farkas E. Ischemia-induced depolarizations and associated hemodynamic responses in incomplete global forebrain ischemia in rats. Neuroscience 260: 217–226, 2014. [Abstract] [Google Scholar]
26. Bere Z, Obrenovitch TP, Kozak G, Bari F, Farkas E. Imaging reveals the focal area of spreading depolarizations and a variety of hemodynamic responses in a rat microembolic stroke model. J Cereb Blood Flow Metab 34: 1695–1705, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
27. Betz AL, Firth JA, Goldstein GW. Polarity of the blood-brain barrier: distribution of enzymes between the luminal and antiluminal membranes of brain capillary endothelial cells. Brain Res 192: 17–28, 1980. [Abstract] [Google Scholar]
28. Betz AL, Goldstein GW. The basis for active transport at the blood-brain barrier. Adv Exp Med Biol 131: 5–16, 1980. [Abstract] [Google Scholar]
29. Bevan JA, Duckworth J, Laher I, Oriowo MA, McPherson GA, Bevan RD. Sympathetic control of cerebral arteries: specialization in receptor type, reserve, affinity, and distribution. FASEB J 1: 193–198, 1987. [Abstract] [Google Scholar]
30. Bolay H, Reuter U, Dunn AK, Huang Z, Boas DA, Moskowitz MA. Intrinsic brain activity triggers trigeminal meningeal afferents in a migraine model. Nat Med 8: 136–142, 2002. [Abstract] [Google Scholar]
31. Bosche B, Graf R, Ernestus RI, Dohmen C, Reithmeier T, Brinker G, Strong AJ, Dreier JP, Woitzik J. Recurrent spreading depolarizations after subarachnoid hemorrhage decreases oxygen availability in human cerebral cortex. Ann Neurol 67: 607–617, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
32. Bradbury MW, Segal MB, Wilson J. Transport of potassium at the blood-brain barrier. J Physiol 221: 617–632, 1972. [Abstract] [Google Scholar]
33. Brandt T, Grau AJ, Hacke W. Severe stroke. Bailliere's Clin Neurol 5: 515–541, 1996. [Abstract] [Google Scholar]
34. Branston NM, Strong AJ, Symon L. Extracellular potassium activity, evoked potential and tissue blood flow. Relationships during progressive ischaemia in baboon cerebral cortex. J Neurol Sci 32: 305–321, 1977. [Abstract] [Google Scholar]
35. Branston NM, Strong AJ, Symon L. Kinetics of resolution of transient increases in extracellular potassium activity: relationships to regional blood flow in primate cerebral cortex. Neurol Res 4: 1–19, 1982. [Abstract] [Google Scholar]
36. Brennan AM, Connor JA, Shuttleworth CW. NAD(P)H fluorescence transients after synaptic activity in brain slices: predominant role of mitochondrial function. J Cereb Blood Flow Metab 26: 1389–1406, 2006. [Abstract] [Google Scholar]
37. Brennan KC, Bates EA, Shapiro RE, Zyuzin J, Hallows WC, Huang Y, Lee HY, Jones CR, Fu YH, Charles AC, Ptacek LJ. Casein kinase idelta mutations in familial migraine and advanced sleep phase. Sci Transl Med 5: 183ra156, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
38. Brennan KC, Beltran-Parrazal L, Lopez-Valdes HE, Theriot J, Toga AW, Charles AC. Distinct vascular conduction with cortical spreading depression. J Neurophysiol 97: 4143–4151, 2007. [Abstract] [Google Scholar]
39. Brinley FJ Jr, Kandel ER, Marshall WH. Potassium outflux from rabbit cortex during spreading depression. J Neurophysiol 23: 246–256, 1960. [Abstract] [Google Scholar]
40. Brozek G. Changes in the membrane potential of cortical cells during spreading depression. Physiol Bohemoslov 15: 98–103, 1966. [Abstract] [Google Scholar]
41. Bures J. Some metabolic aspects of Leao's spreading depression. J Neurochem 1: 153–158, 1956. [Abstract] [Google Scholar]
42. Bures J, Buresova O, Krivanek J. The meaning and significance of Leao's spreading depression. Ann Acad Bras Cienc 56: 385–400, 1984. [Abstract] [Google Scholar]
43. Bures J, Buresova O, Krivanek J. The Mechanism and Applications of Leao's Spreading Depression of Electroencephalographic Activity. New York: Academic, 1974. [Google Scholar]
44. Buresova O. Changes in cerebral circulation in rats during spreading EEG depression. Physiol Bohemoslov 6: 1–11, 1957. [Google Scholar]
45. Busch E, Gyngell ML, Eis M, Hoehn-Berlage M, Hossmann KA. Potassium-induced cortical spreading depressions during focal cerebral ischemia in rats: contribution to lesion growth assessed by diffusion-weighted NMR and biochemical imaging. J Cereb Blood Flow Metab 16: 1090–1099, 1996. [Abstract] [Google Scholar]
46. Busija DW, Meng W. Retention of cerebrovascular dilation after cortical spreading depression in anesthetized rabbits. Stroke 24: 1740–1745, 1993. [Abstract] [Google Scholar]
47. Caesar K, Akgoren N, Mathiesen C, Lauritzen M. Modification of activity-dependent increases in cerebellar blood flow by extracellular potassium in anaesthetized rats. J Physiol 520: 281–292, 1999. [Abstract] [Google Scholar]
48. Caesar K, Hashemi P, Douhou A, Bonvento G, Boutelle MG, Walls AB, Lauritzen M. Glutamate receptor-dependent increments in lactate, glucose and oxygen metabolism evoked in rat cerebellum in vivo. J Physiol 586: 1337–1349, 2008. [Abstract] [Google Scholar]
49. Canals S, Larrosa B, Pintor J, Mena MA, Herreras O. Metabolic challenge to glia activates an adenosine-mediated safety mechanism that promotes neuronal survival by delaying the onset of spreading depression waves. J Cereb Blood Flow Metab 28: 1835–1844, 2008. [Abstract] [Google Scholar]
50. Carlson AP, Carter RE, Shuttleworth CW. Vascular, electrophysiological, and metabolic consequences of cortical spreading depression in a mouse model of simulated neurosurgical conditions. Neurol Res 34: 223–231, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
51. Castro Gde O, Martins-Ferreira H. Deformations and thickness variations accompanying spreading depression in the retina. J Neurophysiol 33: 891–900, 1970. [Abstract] [Google Scholar]
52. Cauli B, Hamel E. Revisiting the role of neurons in neurovascular coupling. Front Neuroenerget 2: 9, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
53. Cauli B, Tong XK, Rancillac A, Serluca N, Lambolez B, Rossier J, Hamel E. Cortical GABA interneurons in neurovascular coupling: relays for subcortical vasoactive pathways. J Neurosci 24: 8940–8949, 2004. [Abstract] [Google Scholar]
54. Chang JC, Shook LL, Biag J, Nguyen EN, Toga AW, Charles AC, Brennan KC. Biphasic direct current shift, haemoglobin desaturation and neurovascular uncoupling in cortical spreading depression. Brain 133: 996–1012, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
55. Chen Q, Chopp M, Bodzin G, Chen H. Temperature modulation of cerebral depolarization during focal cerebral ischemia in rats: correlation with ischemic injury. J Cereb Blood Flow Metab 13: 389–394, 1993. [Abstract] [Google Scholar]
56. Chen S, Feng Z, Li P, Jacques SL, Zeng S, Luo Q. In vivo optical reflectance imaging of spreading depression waves in rat brain with and without focal cerebral ischemia. J Biomed Opt 11: 34002, 2006. [Abstract] [Google Scholar]
57. Chesler M, Kraig RP. Intracellular pH of astrocytes increases rapidly with cortical stimulation. Am J Physiol Regul Integr Comp Physiol 253: R666–R670, 1987. [Europe PMC free article] [Abstract] [Google Scholar]
58. Choi HB, Gordon GR, Zhou N, Tai C, Rungta RL, Martinez J, Milner TA, Ryu JK, McLarnon JG, Tresguerres M, Levin LR, Buck J, MacVicar BA. Metabolic communication between astrocytes and neurons via bicarbonate-responsive soluble adenylyl cyclase. Neuron 75: 1094–1104, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
59. Chuquet J, Hollender L, Nimchinsky EA. High-resolution in vivo imaging of the neurovascular unit during spreading depression. J Neurosci 27: 4036–4044, 2007. [Abstract] [Google Scholar]
60. Cipolla MJ, Curry AB. Middle cerebral artery function after stroke: the threshold duration of reperfusion for myogenic activity. Stroke 33: 2094–2099, 2002. [Abstract] [Google Scholar]
61. Clark D, Institoris A, Kozak G, Bere Z, Tuor U, Farkas E, Bari F. Impact of aging on spreading depolarizations induced by focal brain ischemia in rats. Neurobiol Aging 35: 2803–2811, 2014. [Abstract] [Google Scholar]
62. Cohen Z, Molinatti G, Hamel E. Astroglial and vascular interactions of noradrenaline terminals in the rat cerebral cortex. J Cereb Blood Flow Metab 17: 894–904, 1997. [Abstract] [Google Scholar]
63. Collewijn H, Harreveld AV. Membrane potential of cerebral cortical cells during spreading depression and asphyxia. Exp Neurol 15: 425–436, 1966. [Abstract] [Google Scholar]
64. Colonna DM, Meng W, Deal DD, Busija DW. Calcitonin gene-related peptide promotes cerebrovascular dilation during cortical spreading depression in rabbits. Am J Physiol Heart Circ Physiol 266: H1095–H1102, 1994. [Abstract] [Google Scholar]
65. Colonna DM, Meng W, Deal DD, Busija DW. Nitric oxide promotes arteriolar dilation during cortical spreading depression in rabbits. Stroke 25: 2463–2470, 1994. [Abstract] [Google Scholar]
66. Colonna DM, Meng W, Deal DD, Gowda M, Busija DW. Neuronal NO promotes cerebral cortical hyperemia during cortical spreading depression in rabbits. Am J Physiol Heart Circ Physiol 272: H1315–H1322, 1997. [Abstract] [Google Scholar]
67. Crago EA, Thampatty BP, Sherwood PR, Kuo CW, Bender C, Balzer J, Horowitz M, Poloyac SM. Cerebrospinal fluid 20-HETE is associated with delayed cerebral ischemia and poor outcomes after aneurysmal subarachnoid hemorrhage. Stroke 42: 1872–1877, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
68. Cruz NF, Adachi K, Dienel GA. Rapid efflux of lactate from cerebral cortex during K+-induced spreading cortical depression. J Cereb Blood Flow Metab 19: 380–392, 1999. [Abstract] [Google Scholar]
69. Csiba L, Paschen W, Mies G. Regional changes in tissue pH and glucose content during cortical spreading depression in rat brain. Brain Res 336: 167–170, 1985. [Abstract] [Google Scholar]
70. Davson H, Welch K. The permeation of several materials into the fluids of the rabbit's brain. J Physiol 218: 337–351, 1971. [Abstract] [Google Scholar]
71. De Crespigny A, Rother J, van Bruggen N, Beaulieu C, Moseley ME. Magnetic resonance imaging assessment of cerebral hemodynamics during spreading depression in rats. J Cereb Blood Flow Metab 18: 1008–1017, 1998. [Abstract] [Google Scholar]
72. De la Torre JC. Evidence for central innervation of intracerebral blood vessels: local cerebral blood flow measurements and histofluorescence analysis by the sucrose-phosphate-glyoxylic acid (SPG) method. Neuroscience 1: 455–457, 1976. [Abstract] [Google Scholar]
73. De-Araujo-Pinheiro RS, Martins-Ferreira H. Spreading depression and the blood-brain barrier. Braz J Med Biol Res 17: 265–270, 1984. [Abstract] [Google Scholar]
74. Dhir A, Lossin C, Rogawski MA. Propofol hemisuccinate suppresses cortical spreading depression. Neurosci Lett 514: 67–70, 2012. [Abstract] [Google Scholar]
75. Dienel GA, Liu K, Cruz NF. Local uptake of 14C-labeled acetate and butyrate in rat brain in vivo during spreading cortical depression. J Neurosci Res 66: 812–820, 2001. [Abstract] [Google Scholar]
76. Dietz RM, Weiss JH, Shuttleworth CW. Zn2+ influx is critical for some forms of spreading depression in brain slices. J Neurosci 28: 8014–8024, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
77. Dijkhuizen RM, Beekwilder JP, van der Worp HB, Berkelbach van der Sprenkel JW, Tulleken KA, Nicolay K. Correlation between tissue depolarizations and damage in focal ischemic rat brain. Brain Res 840: 194–205, 1999. [Abstract] [Google Scholar]
78. Dohmen C, Sakowitz OW, Fabricius M, Bosche B, Reithmeier T, Ernestus RI, Brinker G, Dreier JP, Woitzik J, Strong AJ, Graf R. Spreading depolarizations occur in human ischemic stroke with high incidence. Ann Neurol 63: 720–728, 2008. [Abstract] [Google Scholar]
79. Dohmen C, Sakowitz OW, Fabricius M, Bosche B, Reithmeier T, Ernestus RI, Brinker G, Dreier JP, Woitzik J, Strong AJ, Graf R., Co-Operative Study of Brain Injury. Spreading depolarizations occur in human ischemic stroke with high incidence. Ann Neurol 63: 720–728, 2008. [Abstract] [Google Scholar]
80. Domoki F, Veltkamp R, Bari F, Louis TM, Busija DW. Cerebrovascular reactivity remains intact after cortical depolarization in newborn piglets. Pediatr Res 45: 834–837, 1999. [Abstract] [Google Scholar]
81. Dora E, Gyulai L, Kovach AG. Determinants of brain activation-induced cortical NAD/NADH responses in vivo. Brain Res 299: 61–72, 1984. [Abstract] [Google Scholar]
82. Dowson A, Mullen MJ, Peatfield R, Muir K, Khan AA, Wells C, Lipscombe SL, Rees T, De Giovanni JV, Morrison WL, Hildick-Smith D, Elrington G, Hillis WS, Malik IS, Rickards A. Migraine Intervention With STARFlex Technology (MIST) trial: a prospective, multicenter, double-blind, sham-controlled trial to evaluate the effectiveness of patent foramen ovale closure with STARFlex septal repair implant to resolve refractory migraine headache. Circulation 117: 1397–1404, 2008. [Abstract] [Google Scholar]
83. Dreier JP. The role of spreading depression, spreading depolarization and spreading ischemia in neurological disease. Nat Med 17: 439–447, 2011. [Abstract] [Google Scholar]
84. Dreier JP, Ebert N, Priller J, Megow D, Lindauer U, Klee R, Reuter U, Imai Y, Einhaupl KM, Victorov I, Dirnagl U. Products of hemolysis in the subarachnoid space inducing spreading ischemia in the cortex and focal necrosis in rats: a model for delayed ischemic neurological deficits after subarachnoid hemorrhage? J Neurosurg 93: 658–666, 2000. [Abstract] [Google Scholar]
85. Dreier JP, Kleeberg J, Petzold G, Priller J, Windmuller O, Orzechowski HD, Lindauer U, Heinemann U, Einhaupl KM, Dirnagl U. Endothelin-1 potently induces Leao's cortical spreading depression in vivo in the rat: a model for an endothelial trigger of migrainous aura? Brain 125: 102–112, 2002. [Abstract] [Google Scholar]
86. Dreier JP, Korner K, Ebert N, Gorner A, Rubin I, Back T, Lindauer U, Wolf T, Villringer A, Einhaupl KM, Lauritzen M, Dirnagl U. Nitric oxide scavenging by hemoglobin or nitric oxide synthase inhibition by N-nitro-l-arginine induces cortical spreading ischemia when K+ is increased in the subarachnoid space. J Cereb Blood Flow Metab 18: 978–990, 1998. [Abstract] [Google Scholar]
87. Dreier JP, Korner K, Gorner A, Lindauer U, Weih M, Villringer A, Dirnagl U. Nitric oxide modulates the CBF response to increased extracellular potassium. J Cereb Blood Flow Metab 15: 914–919, 1995. [Abstract] [Google Scholar]
88. Dreier JP, Major S, Manning A, Woitzik J, Drenckhahn C, Steinbrink J, Tolias C, Oliveira-Ferreira AI, Fabricius M, Hartings JA, Vajkoczy P, Lauritzen M, Dirnagl U, Bohner G, Strong AJ. Cortical spreading ischaemia is a novel process involved in ischaemic damage in patients with aneurysmal subarachnoid haemorrhage. Brain 132: 1866–1881, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
89. Dreier JP, Major S, Manning A, Woitzik J, Drenckhahn C, Steinbrink J, Tolias C, Oliveira-Ferreira AI, Fabricius M, Hartings JA, Vajkoczy P, Lauritzen M, Dirnagl U, Bohner G, Strong AJ., CS Group. Cortical spreading ischaemia is a novel process involved in ischaemic damage in patients with aneurysmal subarachnoid haemorrhage. Brain 132: 1866–1881, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
90. Dreier JP, Petzold G, Tille K, Lindauer U, Arnold G, Heinemann U, Einhaupl KM, Dirnagl U. Ischaemia triggered by spreading neuronal activation is inhibited by vasodilators in rats. J Physiol 531: 515–526, 2001. [Abstract] [Google Scholar]
91. Dreier JP, Sakowitz OW, Harder A, Zimmer C, Dirnagl U, Valdueza JM, Unterberg AW. Focal laminar cortical MR signal abnormalities after subarachnoid hemorrhage. Ann Neurol 52: 825–829, 2002. [Abstract] [Google Scholar]
92. Dreier JP, Tille K, Dirnagl U. Partial antagonistic effect of adenosine on inverse coupling between spreading neuronal activation and cerebral blood flow in rats. Neurocrit Care 1: 85–94, 2004. [Abstract] [Google Scholar]
93. Dreier JP, Woitzik J, Fabricius M, Bhatia R, Major S, Drenckhahn C, Lehmann TN, Sarrafzadeh A, Willumsen L, Hartings JA, Sakowitz OW, Seemann JH, Thieme A, Lauritzen M, Strong AJ. Delayed ischaemic neurological deficits after subarachnoid haemorrhage are associated with clusters of spreading depolarizations. Brain 129: 3224–3237, 2006. [Abstract] [Google Scholar]
94. Duckrow RB. A brief hypoperfusion precedes spreading depression if nitric oxide synthesis is inhibited. Brain Res 618: 190–195, 1993. [Abstract] [Google Scholar]
95. Duckrow RB. Regional cerebral blood flow during spreading cortical depression in conscious rats. J Cereb Blood Flow Metab 11: 150–154, 1991. [Abstract] [Google Scholar]
96. Duckrow RB, Beard DC. Cortical hypoperfusion following spreading depression is not altered by tirilazad mesylate (U-74006F) in awake rats. Brain Res 572: 296–299, 1992. [Abstract] [Google Scholar]
97. Dunn AK, Bolay H, Moskowitz MA, Boas DA. Dynamic imaging of cerebral blood flow using laser speckle. J Cereb Blood Flow Metab 21: 195–201, 2001. [Abstract] [Google Scholar]
98. Dunn AK, Devor A, Bolay H, andermann ML, Moskowitz MA, Dale AM, Boas DA. Simultaneous imaging of total cerebral hemoglobin concentration, oxygenation, and blood flow during functional activation. Opt Lett 28: 28–30, 2003. [Abstract] [Google Scholar]
99. Dusser de Barenne JG, McCullogh WS. Factors for facilitation and extinction in the central nervous system. J Neurophysiol 319–355, 1939. [Google Scholar]
100. Ebersberger A, Schaible HG, Averbeck B, Richter F. Is there a correlation between spreading depression, neurogenic inflammation, and nociception that might cause migraine headache? Ann Neurol 49: 7–13, 2001. [Abstract] [Google Scholar]
101. Edvinsson L, Degueurce A, Duverger D, MacKenzie ET, Scatton B. Central serotonergic nerves project to the pial vessels of the brain. Nature 306: 55–57, 1983. [Abstract] [Google Scholar]
102. Edvinsson L, Jansen I, Cunha e Sa M, Gulbenkian S. Demonstration of neuropeptide containing nerves and vasomotor responses to perivascular peptides in human cerebral arteries. Cephalalgia 14: 88–96, 1994. [Abstract] [Google Scholar]
103. Eikermann-Haerter K, Dilekoz E, Kudo C, Savitz SI, Waeber C, Baum MJ, Ferrari MD, van den Maagdenberg AM, Moskowitz MA, Ayata C. Genetic and hormonal factors modulate spreading depression and transient hemiparesis in mouse models of familial hemiplegic migraine type 1. J Clin Invest 119: 99–109, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
104. Eikermann-Haerter K, Lee JH, Yuzawa I, Liu CH, Zhou Z, Shin HK, Zheng Y, Qin T, Kurth T, Waeber C, Ferrari MD, van den Maagdenberg AM, Moskowitz MA, Ayata C. Migraine mutations increase stroke vulnerability by facilitating ischemic depolarizations. Circulation 125: 335–345, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
105. Eikermann-Haerter K, Yuzawa I, Dilekoz E, Joutel A, Moskowitz MA, Ayata C. Cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy syndrome mutations increase susceptibility to spreading depression. Ann Neurol 69: 413–418, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
106. Eikermann-Haerter K, Yuzawa I, Qin T, Wang Y, Baek K, Kim YR, Hoffmann U, Dilekoz E, Waeber C, Ferrari MD, van den Maagdenberg AM, Moskowitz MA, Ayata C. Enhanced subcortical spreading depression in familial hemiplegic migraine type 1 mutant mice. J Neurosci 31: 5755–5763, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
107. Eisenberg HM, Suddith RL. Cerebral vessels have the capacity to transport sodium and potassium. Science 206: 1083–1085, 1979. [Abstract] [Google Scholar]
108. Enager P, Piilgaard H, Offenhauser N, Kocharyan A, Fernandes P, Hamel E, Lauritzen M. Pathway-specific variations in neurovascular and neurometabolic coupling in rat primary somatosensory cortex. J Cereb Blood Flow Metab 29: 976–986, 2009. [Abstract] [Google Scholar]
109. Enevoldsen EM, Jensen FT. Autoregulation and CO2 responses of cerebral blood flow in patients with acute severe head injury. J Neurosurg 48: 689–703, 1978. [Abstract] [Google Scholar]
110. Erecinska M, Silver IA. ATP and brain function. J Cereb Blood Flow Metab 9: 2–19, 1989. [Abstract] [Google Scholar]
111. Erecinska M, Silver IA. Ions and energy in mammalian brain. Prog Neurobiol 43: 37–71, 1994. [Abstract] [Google Scholar]
112. Estrada C, Mengual E, Gonzalez C. Local NADPH-diaphorase neurons innervate pial arteries and lie close or project to intracerebral blood vessels: a possible role for nitric oxide in the regulation of cerebral blood flow. J Cereb Blood Flow Metab 13: 978–984, 1993. [Abstract] [Google Scholar]
113. Fabricius M, Akgoren N, Dirnagl U, Lauritzen M. Laminar analysis of cerebral blood flow in cortex of rats by laser-Doppler flowmetry: a pilot study. J Cereb Blood Flow Metab 17: 1326–1336, 1997. [Abstract] [Google Scholar]
114. Fabricius M, Akgoren N, Lauritzen M. Arginine-nitric oxide pathway and cerebrovascular regulation in cortical spreading depression. Am J Physiol Heart Circ Physiol 269: H23–H29, 1995. [Abstract] [Google Scholar]
115. Fabricius M, Fuhr S, Bhatia R, Boutelle M, Hashemi P, Strong AJ, Lauritzen M. Cortical spreading depression and peri-infarct depolarization in acutely injured human cerebral cortex. Brain 129: 778–790, 2006. [Abstract] [Google Scholar]
116. Fabricius M, Fuhr S, Willumsen L, Dreier JP, Bhatia R, Boutelle MG, Hartings JA, Bullock R, Strong AJ, Lauritzen M. Association of seizures with cortical spreading depression and peri-infarct depolarisations in the acutely injured human brain. Clin Neurophysiol 119: 1973–1984, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
117. Fabricius M, Jensen LH, Lauritzen M. Microdialysis of interstitial amino acids during spreading depression and anoxic depolarization in rat neocortex. Brain Res 612: 61–69, 1993. [Abstract] [Google Scholar]
118. Fabricius M, Lauritzen M. Examination of the role of nitric oxide for the hypercapnic rise of cerebral blood flow in rats. Am J Physiol Heart Circ Physiol 266: H1457–H1464, 1994. [Abstract] [Google Scholar]
119. Fabricius M, Lauritzen M. Laser-Doppler evaluation of rat brain microcirculation: comparison with the [14C]-iodoantipyrine method suggests discordance during cerebral blood flow increases. J Cereb Blood Flow Metab 16: 156–161, 1996. [Abstract] [Google Scholar]
120. Fabricius M, Lauritzen M. Transient hyperemia succeeds oligemia in the wake of cortical spreading depression. Brain Res 602: 350–353, 1993. [Abstract] [Google Scholar]
121. Faraci FM, Breese KR. Nitric oxide mediates vasodilatation in response to activation of N-methyl-d-aspartate receptors in brain. Circ Res 72: 476–480, 1993. [Abstract] [Google Scholar]
122. Faraci FM, Brian JE Jr. 7-Nitroindazole inhibits brain nitric oxide synthase and cerebral vasodilatation in response to N-methyl-d-aspartate. Stroke 26: 2172–2176, 1995. [Abstract] [Google Scholar]
123. Farkas E, Obrenovitch TP, Institoris A, Bari F. Effects of early aging and cerebral hypoperfusion on spreading depression in rats. Neurobiol Aging 32: 1707–1715, 2011. [Abstract] [Google Scholar]
124. Farkas E, Pratt R, Sengpiel F, Obrenovitch TP. Direct, live imaging of cortical spreading depression and anoxic depolarisation using a fluorescent, voltage-sensitive dye. J Cereb Blood Flow Metab 28: 251–262, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
125. Farkkila M, Palo J, Saijonmaa O, Fyhrquist F. Raised plasma endothelin during acute migraine attack. Cephalalgia 12: 383–384, 1992. [Abstract] [Google Scholar]
126. Fergus A, Lee KS. Regulation of cerebral microvessels by glutamatergic mechanisms. Brain Res 754: 35–45, 1997. [Abstract] [Google Scholar]
127. Fernandez-Klett F, Offenhauser N, Dirnagl U, Priller J, Lindauer U. Pericytes in capillaries are contractile in vivo, but arterioles mediate functional hyperemia in the mouse brain. Proc Natl Acad Sci USA 107: 22290–22295, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
128. Feuerstein D, Takagaki M, Gramer M, Manning A, Endepols H, Vollmar S, Yoshimine T, Strong AJ, Graf R, Backes H. Detecting tissue deterioration after brain injury: regional blood flow level versus capacity to raise blood flow. J Cereb Blood Flow Metab 34: 1117–1127, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
129. Filosa JA, Bonev AD, Straub SV, Meredith AL, Wilkerson MK, Aldrich RW, Nelson MT. Local potassium signaling couples neuronal activity to vasodilation in the brain. Nat Neurosci 9: 1397–1403, 2006. [Abstract] [Google Scholar]
130. Florence G, Bonvento G, Charbonne R, Seylaz J. Spreading depression reversibly impairs autoregulation of cortical blood flow. Am J Physiol Regul Integr Comp Physiol 266: R1136–R1140, 1994. [Abstract] [Google Scholar]
131. Fordsmann JC, Ko RW, Choi HB, Thomsen K, Witgen BM, Mathiesen C, Lonstrup M, Piilgaard H, MacVicar BA, Lauritzen M. Increased 20-HETE synthesis explains reduced cerebral blood flow but not impaired neurovascular coupling after cortical spreading depression in rat cerebral cortex. J Neurosci 33: 2562–2570, 2013. [Abstract] [Google Scholar]
132. Fox PT, Raichle ME. Focal physiological uncoupling of cerebral blood flow and oxidative metabolism during somatosensory stimulation in human subjects. Proc Natl Acad Sci USA 83: 1140–1144, 1986. [Europe PMC free article] [Abstract] [Google Scholar]
133. Fox PT, Raichle ME, Mintun MA, Dence C. Nonoxidative glucose consumption during focal physiologic neural activity. Science 241: 462–464, 1988. [Abstract] [Google Scholar]
134. Freygang WH Jr, Landau WM. Some relations between resistivity and electrical activity in the cerebral cortex of the cat. J Cell Physiol 45: 377–392, 1955. [Abstract] [Google Scholar]
135. Friberg H, Wieloch T. Mitochondrial permeability transition in acute neurodegeneration. Biochimie 84: 241–250, 2002. [Abstract] [Google Scholar]
136. Friberg L, Olsen TS, Roland PE, Lassen NA. Focal ischaemia caused by instability of cerebrovascular tone during attacks of hemiplegic migraine. A regional cerebral blood flow study. Brain 110: 917–934, 1987. [Abstract] [Google Scholar]
137. Fujita S, Mizoguchi N, Aoki R, Cui Y, Koshikawa N, Kobayashi M. Cytoarchitecture-Dependent decrease in propagation velocity of cortical spreading depression in the rat insular cortex revealed by optical imaging. Cereb Cortex. In press. [Abstract] [Google Scholar]
138. Galeffi F, Somjen GG, Foster KA, Turner DA. Simultaneous monitoring of tissue Po2 and NADH fluorescence during synaptic stimulation and spreading depression reveals a transient dissociation between oxygen utilization and mitochondrial redox state in rat hippocampal slices. J Cereb Blood Flow Metab 31: 626–639, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
139. Gallai V, Sarchielli P, Firenze C, Trequattrini A, Paciaroni M, Usai F, Palumbo R. Endothelin 1 in migraine and tension-type headache. Acta Neurol Scand 89: 47–55, 1994. [Abstract] [Google Scholar]
140. Gardner-Medwin AR. Analysis of potassium dynamics in mammalian brain tissue. J Physiol 335: 393–426, 1983. [Abstract] [Google Scholar]
141. Gardner-Medwin AR. Possible roles of vertebrate neuroglia in potassium dynamics, spreading depression and migraine. J Exp Biol 95: 111–127, 1981. [Abstract] [Google Scholar]
142. Gardner-Medwin AR. A study of the mechanisms by which potassium moves through brain tissue in the rat. J Physiol 335: 353–374, 1983. [Abstract] [Google Scholar]
143. Gault LM, Lin CW, LaManna JC, Lust WD. Changes in energy metabolites, cGMP and intracellular pH during cortical spreading depression. Brain Res 641: 176–180, 1994. [Abstract] [Google Scholar]
144. Gebremedhin D, Lange AR, Lowry TF, Taheri MR, Birks EK, Hudetz AG, Narayanan J, Falck JR, Okamoto H, Roman RJ, Nithipatikom K, Campbell WB, Harder DR. Production of 20-HETE and its role in autoregulation of cerebral blood flow. Circ Res 87: 60–65, 2000. [Abstract] [Google Scholar]
145. Gerich FJ, Hepp S, Probst I, Muller M. Mitochondrial inhibition prior to oxygen-withdrawal facilitates the occurrence of hypoxia-induced spreading depression in rat hippocampal slices. J Neurophysiol 96: 492–504, 2006. [Abstract] [Google Scholar]
146. Gido G, Katsura K, Kristian T, Siesjo BK. Influence of plasma glucose concentration on rat brain extracellular calcium transients during spreading depression. J Cereb Blood Flow Metab 13: 179–182, 1993. [Abstract] [Google Scholar]
147. Gido G, Kristian T, Siesjo BK. Induced spreading depressions in energy-compromised neocortical tissue: calcium transients and histopathological correlates. Neurobiol Dis 1: 31–41, 1994. [Abstract] [Google Scholar]
148. Gill R, andine P, Hillered L, Persson L, Hagberg H. The effect of MK-801 on cortical spreading depression in the penumbral zone following focal ischaemia in the rat. J Cereb Blood Flow Metab 12: 371–379, 1992. [Abstract] [Google Scholar]
149. Girouard H, Bonev AD, Hannah RM, Meredith A, Aldrich RW, Nelson MT. Astrocytic endfoot Ca2+ and BK channels determine both arteriolar dilation and constriction. Proc Natl Acad Sci USA 107: 3811–3816, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
150. Gjedde A, Hansen AJ, Quistorff B. Blood-brain glucose transfer in spreading depression. J Neurochem 37: 807–812, 1981. [Abstract] [Google Scholar]
151. Goadsby PJ. Inhibition of calcitonin gene-related peptide by h-CGRP(8–37) antagonizes the cerebral dilator response from nasociliary nerve stimulation in the cat. Neurosci Lett 151: 13–16, 1993. [Abstract] [Google Scholar]
152. Goadsby PJ. The oligemic phase of cortical spreading depression is not blocked by tirilazad mesylate (U-74006F). Brain Res 588: 140–143, 1992. [Abstract] [Google Scholar]
153. Goadsby PJ, Adner M, Edvinsson L. Characterization of endothelin receptors in the cerebral vasculature and their lack of effect on spreading depression. J Cereb Blood Flow Metab 16: 698–704, 1996. [Abstract] [Google Scholar]
154. Goadsby PJ, Kaube H, Hoskin KL. Nitric oxide synthesis couples cerebral blood flow and metabolism. Brain Res 595: 167–170, 1992. [Abstract] [Google Scholar]
155. Gold L, Back T, Arnold G, Dreier J, Einhaupl KM, Reuter U, Dirnagl U. Cortical spreading depression-associated hyperemia in rats: involvement of serotonin. Brain Res 783: 188–193, 1998. [Abstract] [Google Scholar]
156. Golding EM, Steenberg ML, Johnson TD, Bryan RM. Nitric oxide in the potassium-induced response of the rat middle cerebral artery: a possible permissive role. Brain Res 889: 98–104, 2001. [Abstract] [Google Scholar]
157. Golding EM, Steenberg ML, Johnson TD, Bryan RM Jr. The effects of potassium on the rat middle cerebral artery. Brain Res 880: 159–166, 2000. [Abstract] [Google Scholar]
158. Goldstein GW. Relation of potassium transport to oxidative metabolism in isolated brain capillaries. J Physiol 286: 185–195, 1979. [Abstract] [Google Scholar]
159. Gordon GR, Choi HB, Rungta RL, Ellis-Davies GC, MacVicar BA. Brain metabolism dictates the polarity of astrocyte control over arterioles. Nature 456: 745–749, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
160. Gorji A. Spreading depression: a review of the clinical relevance. Brain Res 38: 33–60, 2001. [Abstract] [Google Scholar]
161. Grafstein B. Locus of propagation of spreading cortical depression. J Neurophysiol 19: 308–316, 1956. [Abstract] [Google Scholar]
162. Grafstein B. Mechanism of spreading cortical depression. J Neurophysiol 19: 154–171, 1956. [Abstract] [Google Scholar]
163. Gramer M, Feuerstein D, Steimers A, Takagaki M, Kumagai T, Sue M, Vollmar S, Kohl-Bareis M, Backes H, Graf R. Device for simultaneous positron emission tomography, laser speckle imaging and RGB reflectometry: validation and application to cortical spreading depression and brain ischemia in rats. Neuroimage 94: 250–262, 2014. [Abstract] [Google Scholar]
164. Grinberg YY, Dibbern ME, Levasseur VA, Kraig RP. Insulin-like growth factor-1 abrogates microglial oxidative stress and TNF-alpha responses to spreading depression. J Neurochem 126: 662–672, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
165. Grinberg YY, van Drongelen W, Kraig RP. Insulin-like growth factor-1 lowers spreading depression susceptibility and reduces oxidative stress. J Neurochem 122: 221–229, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
166. Guiou M, Sheth S, Nemoto M, Walker M, Pouratian N, Ba A, Toga AW. Cortical spreading depression produces long-term disruption of activity-related changes in cerebral blood volume and neurovascular coupling. J Biomed Opt 10: 11004, 2005. [Abstract] [Google Scholar]
167. Gurdjian ES, Webster JE, Stone WE. Cerebral constituents in relation to blood gases. Am J Physiol 156: 149–157, 1949. [Abstract] [Google Scholar]
168. Gursoy-Ozdemir Y, Qiu J, Matsuoka N, Bolay H, Bermpohl D, Jin H, Wang X, Rosenberg GA, Lo EH, Moskowitz MA. Cortical spreading depression activates and upregulates MMP-9. J Clin Invest 113: 1447–1455, 2004. [Europe PMC free article] [Abstract] [Google Scholar]
169. Hadjikhani N, Sanchez Del Rio M, Wu O, Schwartz D, Bakker D, Fischl B, Kwong KK, Cutrer FM, Rosen BR, Tootell RB, Sorensen AG, Moskowitz MA. Mechanisms of migraine aura revealed by functional MRI in human visual cortex. Proc Natl Acad Sci USA 98: 4687–4692, 2001. [Europe PMC free article] [Abstract] [Google Scholar]
170. Hall CN, Reynell C, Gesslein B, Hamilton NB, Mishra A, Sutherland BA, O'Farrell FM, Buchan AM, Lauritzen M, Attwell D. Capillary pericytes regulate cerebral blood flow in health and disease. Nature 508: 55–60, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
171. Hall ED, Smith SL. The 21-aminosteroid antioxidant tirilazad mesylate, U-74006F, blocks cortical hypoperfusion following spreading depression. Brain Res 553: 243–248, 1991. [Abstract] [Google Scholar]
172. Hamel E. Perivascular nerves and the regulation of cerebrovascular tone. J Appl Physiol 100: 1059–1064, 2006. [Abstract] [Google Scholar]
173. Hanke W, de Lima VM. Central nervous tissue: an excitable medium. A study using the retinal spreading depression as a tool. Philos Transact A Math Phys Eng Sci 366: 359–368, 2008. [Abstract] [Google Scholar]
174. Hansen AJ. Effect of anoxia on ion distribution in the brain. Physiol Rev 65: 101–148, 1985. [Abstract] [Google Scholar]
175. Hansen AJ, Gjedde A, Siemkowicz E. Extracellular potassium and blood flow in the post-ischemic rat brain. Pflügers Arch 389: 1–7, 1980. [Abstract] [Google Scholar]
176. Hansen AJ, Lauritzen M. The role of spreading depression in acute brain disorders. Ann Acad Bras Cienc 56: 457–479, 1984. [Abstract] [Google Scholar]
177. Hansen AJ, Lund-andersen H, Crone C. K+-permeability of the blood-brain barrier, investigated by aid of a K+-sensitive microelectrode. Acta Physiol Scand 101: 438–445, 1977. [Abstract] [Google Scholar]
178. Hansen AJ, Olsen CE. Brain extracellular space during spreading depression and ischemia. Acta Physiol Scand 108: 355–365, 1980. [Abstract] [Google Scholar]
179. Hansen AJ, Quistorff B, Gjedde A. Relationship between local changes in cortical blood flow and extracellular K+ during spreading depression. Acta Physiol Scand 109: 1–6, 1980. [Abstract] [Google Scholar]
180. Hansen AJ, Zeuthen T. Extracellular ion concentrations during spreading depression and ischemia in the rat brain cortex. Acta Physiol Scand 113: 437–445, 1981. [Abstract] [Google Scholar]
181. Hardebo JE, Wieloch T, Kahrstrom J. Excitatory amino acids and cerebrovascular tone. Acta Physiol Scand 136: 483–485, 1989. [Abstract] [Google Scholar]
182. Harrington MG, Fonteh AN, Arakaki X, Cowan RP, Ecke LE, Foster H, Huhmer AF, Biringer RG. Capillary endothelial Na+,K+,ATPase transporter homeostasis and a new theory for migraine pathophysiology. Headache 50: 459–478, 2010. [Abstract] [Google Scholar]
183. Hartings JA, Strong AJ, Fabricius M, Manning A, Bhatia R, Dreier JP, Mazzeo AT, Tortella FC, Bullock MR. Spreading depolarizations and late secondary insults after traumatic brain injury. J Neurotrauma 26: 1857–1866, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
184. Hartings JA, Vidgeon S, Strong AJ, Zacko C, Vagal A, andaluz N, Ridder T, Stanger R, Fabricius M, Mathern B, Pahl C, Tolias CM, Bullock MR., Co-Operative Studies on Brain Injury. Surgical management of traumatic brain injury: a comparative-effectiveness study of 2 centers. J Neurosurg 120: 434–446, 2014. [Abstract] [Google Scholar]
185. Haselgrove JC, Bashford CL, Barlow CH, Quistorff B, Chance B, Mayevsky A. Time resolved 3-dimensional recording of redox ratio during spreading depression in gerbil brain. Brain Res 506: 109–114, 1990. [Abstract] [Google Scholar]
186. Hashemi P, Bhatia R, Nakamura H, Dreier JP, Graf R, Strong AJ, Boutelle MG. Persisting depletion of brain glucose following cortical spreading depression, despite apparent hyperaemia: evidence for risk of an adverse effect of Leao's spreading depression. J Cereb Blood Flow Metab 29: 166–175, 2009. [Abstract] [Google Scholar]
187. Hashimoto M, Takeda Y, Sato T, Kawahara H, Nagano O, Hirakawa M. Dynamic changes of NADH fluorescence images and NADH content during spreading depression in the cerebral cortex of gerbils. Brain Res 872: 294–300, 2000. [Abstract] [Google Scholar]
188. Hayward JN, Baker MA. A comparative study of the role of the cerebral arterial blood in the regulation of brain temperature in five mammals. Brain Res 16: 417–440, 1969. [Abstract] [Google Scholar]
189. Hazelton JL, Petrasheuskaya M, Fiskum G, Kristian T. Cyclophilin D is expressed predominantly in mitochondria of gamma-aminobutyric acidergic interneurons. J Neurosci Res 87: 1250–1259, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
190. Heinemann U, Lux HD. Ceiling of stimulus induced rises in extracellular potassium concentration in the cerebral cortex of cat. Brain Res 120: 231–249, 1977. [Abstract] [Google Scholar]
191. Hertle DN, Dreier JP, Woitzik J, Hartings JA, Bullock R, Okonkwo DO, Shutter LA, Vidgeon S, Strong AJ, Kowoll C, Dohmen C, Diedler J, Veltkamp R, Bruckner T, Unterberg AW, Sakowitz OW. Effect of analgesics and sedatives on the occurrence of spreading depolarizations accompanying acute brain injury. Brain 135: 2390–2398, 2012. [Abstract] [Google Scholar]
192. Hertz L, Fillenz M. Does the ‘mystery of the extra glucose’ during CNS activation reflect glutamate synthesis? Neurochem Int 34: 71–75, 1999. [Abstract] [Google Scholar]
193. Higuchi T, Takeda Y, Hashimoto M, Nagano O, Hirakawa M. Dynamic changes in cortical NADH fluorescence and direct current potential in rat focal ischemia: relationship between propagation of recurrent depolarization and growth of the ischemic core. J Cereb Blood Flow Metab 22: 71–79, 2002. [Abstract] [Google Scholar]
194. Hinzman JM, andaluz N, Shutter LA, Okonkwo DO, Pahl C, Strong AJ, Dreier JP, Hartings JA. Inverse neurovascular coupling to cortical spreading depolarizations in severe brain trauma. Brain 137: 2960–2972, 2014. [Abstract] [Google Scholar]
195. Hirabayashi H, Kurita D, Takizawa S, Shinohara Y. Phosphate-related energy compounds are not exhausted in chronically hypoperfused rat brain cortex after cortical spreading depression. J Stroke Cerebrovasc Dis 13: 271–279, 2004. [Abstract] [Google Scholar]
196. Hof PR, Pascale E, Magistretti PJ. K+ at concentrations reached in the extracellular space during neuronal activity promotes a Ca2+-dependent glycogen hydrolysis in mouse cerebral cortex. J Neurosci 8: 1922–1928, 1988. [Abstract] [Google Scholar]
197. Hoffmann U, Ayata C. Neurovascular coupling during spreading depolarizations. Acta Neurochir Suppl 115: 161–165, 2013. [Abstract] [Google Scholar]
198. Hoffmann U, Sukhotinsky I, Atalay YB, Eikermann-Haerter K, Ayata C. Increased glucose availability does not restore prolonged spreading depression durations in hypotensive rats without brain injury. Exp Neurol 238: 130–132, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
199. Hoffmeyer HW, Enager P, Thomsen KJ, Lauritzen MJ. Nonlinear neurovascular coupling in rat sensory cortex by activation of transcallosal fibers. J Cereb Blood Flow Metab 27: 575–587, 2007. [Abstract] [Google Scholar]
200. Horiguchi T, Kis B, Rajapakse N, Shimizu K, Busija DW. Cortical spreading depression (CSD)-induced tolerance to transient focal cerebral ischemia in halothane anesthetized rats is affected by anesthetic level but not ATP-sensitive potassium channels. Brain Res 1062: 127–133, 2005. [Abstract] [Google Scholar]
201. Horiuchi T, Dietrich HH, Hongo K, Dacey RG Jr. Mechanism of extracellular K+-induced local and conducted responses in cerebral penetrating arterioles. Stroke 33: 2692–2699, 2002. [Abstract] [Google Scholar]
202. Hossmann KA. Periinfarct depolarizations. Cerebrovasc Brain Metab Rev 8: 195–208, 1996. [Abstract] [Google Scholar]
203. Hotez L, Dailey JW, Geelhoed GW, Gainer JL. The role of oxygen diffusivity in biochemical reactions. Experientia 33: 1424–1425, 1977. [Abstract] [Google Scholar]
204. Iadecola C, Beitz AJ, Renno W, Xu X, Mayer B, Zhang F. Nitric oxide synthase-containing neural processes on large cerebral arteries and cerebral microvessels. Brain Res 606: 148–155, 1993. [Abstract] [Google Scholar]
205. Iadecola C, Pelligrino DA, Moskowitz MA, Lassen NA. Nitric oxide synthase inhibition and cerebrovascular regulation. J Cereb Blood Flow Metab 14: 175–192, 1994. [Abstract] [Google Scholar]
206. Iijima T, Mies G, Hossmann KA. Repeated negative DC deflections in rat cortex following middle cerebral artery occlusion are abolished by MK-801: effect on volume of ischemic injury. J Cereb Blood Flow Metab 12: 727–733, 1992. [Abstract] [Google Scholar]
207. Iijima T, Shimase C, Iwao Y, Sankawa H. Relationships between glutamate release, blood flow and spreading depression: real-time monitoring using an electroenzymatic dialysis electrode. Neurosci Res 32: 201–207, 1998. [Abstract] [Google Scholar]
208. Iizuka T, Takahashi Y, Sato M, Yonekura J, Miyakawa S, Endo M, Hamada J, Kan S, Mochizuki H, Momose Y, Tsuji S, Sakai F. Neurovascular changes in prolonged migraine aura in FHM with a novel ATP1A2 gene mutation. J Neurol Neurosurg Psychiatry 83: 205–212, 2012. [Abstract] [Google Scholar]
209. Iizuka T, Tominaga N, Kaneko J, Sato M, Akutsu T, Hamada J, Sakai F, Nishiyama K. Biphasic neurovascular changes in prolonged migraine aura in familial hemiplegic migraine type 2. J Neurol Neurosurg Psychiatry 86: 344–353, 2015. [Abstract] [Google Scholar]
210. Iliff JJ, Wang M, Liao Y, Plogg BA, Peng W, Gundersen GA, Benveniste H, Vates GE, Deane R, Goldman SA, Nagelhus EA, Nedergaard M. A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid beta. Sci Transl Med 4: 147ra111, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
211. Iliff JJ, Wang M, Zeppenfeld DM, Venkataraman A, Plog BA, Liao Y, Deane R, Nedergaard M. Cerebral arterial pulsation drives paravascular CSF-interstitial fluid exchange in the murine brain. J Neurosci 33: 18190–18199, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
212. Inoue K, Tomita M, Fukuuchi Y, Tanahashi N, Kobari M, Takao M, Takeda H, Yokoyama M. Dynamic observation of oxygenation-induced contraction of and transient fiber-network formation-disassembly in cultured human brain microvascular endothelial cells. J Cereb Blood Flow Metab 23: 821–828, 2003. [Abstract] [Google Scholar]
213. Itoh Y, Suzuki N. Control of brain capillary blood flow. J Cereb Blood Flow Metab 32: 1167–1176, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
214. Jensen TS, Voldby B, de Fine Olivarius B, Jensen FT. Cerebral hemodynamics in familial hemiplegic migraine. Cephalalgia 1: 121–125, 1981. [Abstract] [Google Scholar]
215. Jespersen SN, Ostergaard L. The roles of cerebral blood flow, capillary transit time heterogeneity, and oxygen tension in brain oxygenation and metabolism. J Cereb Blood Flow Metab 32: 264–277, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
216. Jobsis FF, Keizer JH, LaManna JC, Rosenthal M. Reflectance spectrophotometry of cytochrome aa3 in vivo. J Appl Physiol 43: 858–872, 1977. [Abstract] [Google Scholar]
217. Kaku T, Hada J, Hayashi Y. Endogenous adenosine exerts inhibitory effects upon the development of spreading depression and glutamate release induced by microdialysis with high K+ in rat hippocampus. Brain Res 658: 39–48, 1994. [Abstract] [Google Scholar]
218. Kao YC, Li W, Lai HY, Oyarzabal EA, Lin W, Shih YY. Dynamic perfusion and diffusion MRI of cortical spreading depolarization in photothrombotic ischemia. Neurobiol Dis 71: 131–139, 2014. [Abstract] [Google Scholar]
219. Katzman R. Maintenance of a constant brain extracellular potassium. Federation Proc 35: 1244–1247, 1976. [Abstract] [Google Scholar]
220. Kaube H, Goadsby PJ. Anti-migraine compounds fail to modulate the propagation of cortical spreading depression in the cat. Eur Neurol 34: 30–35, 1994. [Abstract] [Google Scholar]
221. Kaube H, Knight YE, Storer RJ, Hoskin KL, May A, Goadsby PJ. Vasodilator agents and supracollicular transection fail to inhibit cortical spreading depression in the cat. Cephalalgia 19: 592–597, 1999. [Abstract] [Google Scholar]
222. Kehl F, Cambj-Sapunar L, Maier KG, Miyata N, Kametani S, Okamoto H, Hudetz AG, Schulte ML, Zagorac D, Harder DR, Roman RJ. 20-HETE contributes to the acute fall in cerebral blood flow after subarachnoid hemorrhage in the rat. Am J Physiol Heart Circ Physiol 282: H1556–H1565, 2002. [Abstract] [Google Scholar]
223. Kleeberg J, Petzold GC, Major S, Dirnagl U, Dreier JP. ET-1 induces cortical spreading depression via activation of the ETA receptor/phospholipase C pathway in vivo. Am J Physiol Heart Circ Physiol 286: H1339–H1346, 2004. [Abstract] [Google Scholar]
224. Knott HJ, Zimmermann PA, Nelson MT. Extracellular K+-induced hyperpolarizations and dilatations of rat coronary and cerebral arteries involve inward rectifier K+ channels. J Physiol 492: 419–430, 1996. [Abstract] [Google Scholar]
225. Kocharyan A, Fernandes P, Tong XK, Vaucher E, Hamel E. Specific subtypes of cortical GABA interneurons contribute to the neurovascular coupling response to basal forebrain stimulation. J Cereb Blood Flow Metab 28: 221–231, 2008. [Abstract] [Google Scholar]
226. Kocher M. Metabolic and hemodynamic activation of postischemic rat brain by cortical spreading depression. J Cereb Blood Flow Metab 10: 564–571, 1990. [Abstract] [Google Scholar]
227. Kohl M, Lindauer U, Dirnagl U, Villringer A. Separation of changes in light scattering and chromophore concentrations during cortical spreading depression in rats. Opt Lett 23: 555–557, 1998. [Abstract] [Google Scholar]
228. Koide M, Bonev AD, Nelson MT, Wellman GC. Inversion of neurovascular coupling by subarachnoid blood depends on large-conductance Ca2+-activated K+ (BK) channels. Proc Natl Acad Sci USA 109: E1387–1395, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
229. Koroleva VI, Bures J. The use of spreading depression waves for acute and long-term monitoring of the penumbra zone of focal ischemic damage in rats. Proc Natl Acad Sci USA 93: 3710–3714, 1996. [Europe PMC free article] [Abstract] [Google Scholar]
230. Koroleva VI, Vykhodtseva NI, Elagin VA. Spreading depression in the cortex and subcortical structures of the brain of the rat induced by exposure to focused ultrasound. Neirofiziologiia 18: 55–61, 1986. [Abstract] [Google Scholar]
231. Kraig RP, Ferreira-Filho CR, Nicholson C. Alkaline and acid transients in cerebellar microenvironment. J Neurophysiol 49: 831–850, 1983. [Abstract] [Google Scholar]
232. Kraig RP, Nicholson C. Extracellular ionic variations during spreading depression. Neuroscience 3: 1045–1059, 1978. [Abstract] [Google Scholar]
233. Kreisman NR, Sick TJ, LaManna JC, Rosenthal M. Local tissue oxygen tension-cytochrome a,a3 redox relationships in rat cerebral cortex in vivo. Brain Res 218: 161–174, 1981. [Abstract] [Google Scholar]
234. Krivanek J. Adenosine 3’,5’-monophosphate in rat cerebral cortex: effect of potassium ions in vivo (cortical spreading depression). J Neurochem 26: 413–415, 1976. [Abstract] [Google Scholar]
235. Krivanek J. Brain cyclic adenosine 3’,5’-monophosphate during depolarization of the cerebral cortical cells in vivo. Brain Res 120: 493–505, 1977. [Abstract] [Google Scholar]
236. Krivanek J. Changes of brain glycogen in the spreading EEG-depression of Leao. J Neurochem 2: 337–343, 1958. [Abstract] [Google Scholar]
237. Krivanek J. Concerning the dynamics of the metabolic changes accompanying cortical spreading depression. Physiol Bohemoslov 11: 383–391, 1962. [Abstract] [Google Scholar]
238. Krivanek J. Some metabolic changes accompanying Leao's spreading cortical depression in the rat. J Neurochem 6: 183–189, 1961. [Abstract] [Google Scholar]
239. Kruger H, Luhmann HJ, Heinemann U. Repetitive spreading depression causes selective suppression of GABAergic function. Neuroreport 7: 2733–2736, 1996. [Abstract] [Google Scholar]
240. Kudo C, Nozari A, Moskowitz MA, Ayata C. The impact of anesthetics and hyperoxia on cortical spreading depression. Exp Neurol 212: 201–206, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
241. Kudo C, Toyama M, Boku A, Hanamoto H, Morimoto Y, Sugimura M, Niwa H. Anesthetic effects on susceptibility to cortical spreading depression. Neuropharmacology 67: 32–36, 2013. [Abstract] [Google Scholar]
242. Kuge Y, Hasegawa Y, Yokota C, Minematsu K, Hashimoto N, Miyake Y, Yamaguchi T. Effects of single and repetitive spreading depression on cerebral blood flow and glucose metabolism in cats: a PET study. J Neurol Sci 176: 114–123, 2000. [Abstract] [Google Scholar]
243. Kumagai T, Walberer M, Nakamura H, Endepols H, Sue M, Vollmar S, Adib S, Mies G, Yoshimine T, Schroeter M, Graf R. Distinct spatiotemporal patterns of spreading depolarizations during early infarct evolution: evidence from real-time imaging. J Cereb Blood Flow Metab 31: 580–592, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
244. Kuschinsky W, Wahl M. Alpha-receptor stimulation by endogenous and exogenous norepinephrine and blockade by phentolamine in pial arteries of cats. Circ Res 37: 168–174, 1975. [Abstract] [Google Scholar]
245. Kuschinsky W, Wahl M. Local chemical and neurogenic regulation of cerebral vascular resistance. Physiol Rev 58: 656–689, 1978. [Abstract] [Google Scholar]
246. Kuschinsky W, Wahl M, Bosse O, Thurau K. Perivascular potassium and pH as determinants of local pial arterial diameter in cats. A microapplication study. Circ Res 31: 240–247, 1972. [Abstract] [Google Scholar]
247. Kuschinsky W, Wahl M, Neiss A. Evidence for cholinergic dilatatory receptors in pial arteries of cats. A microapplication study. Pflügers Arch 347: 199–208, 1974. [Abstract] [Google Scholar]
248. Lacombe P, Sercombe R, Correze JL, Springhetti V, Seylaz J. Spreading depression induces prolonged reduction of cortical blood flow reactivity in the rat. Exp Neurol 117: 278–286, 1992. [Abstract] [Google Scholar]
249. Lakowicz JR, Szmacinski H, Nowaczyk K, Johnson ML. Fluorescence lifetime imaging of free and protein-bound NADH. Proc Natl Acad Sci USA 89: 1271–1275, 1992. [Europe PMC free article] [Abstract] [Google Scholar]
250. LaManna JC, McCracken KA, Patil M, Prohaska O. Brain tissue temperature: activation-induced changes determined with a new multisensor probe. Adv Exp Med Biol 222: 383–389, 1988. [Abstract] [Google Scholar]
251. LaManna JC, McCracken KA, Patil M, Prohaska OJ. Stimulus-activated changes in brain tissue temperature in the anesthetized rat. Metab Brain Dis 4: 225–237, 1989. [Abstract] [Google Scholar]
252. LaManna JC, Rosenthal M. Effect of ouabain and phenobarbital on oxidative metabolic activity associated with spreading cortical depression in cats. Brain Res 88: 145–149, 1975. [Abstract] [Google Scholar]
253. Lashley KS. Patterns of cerebral integration indicated by the scotomas of migraine. Arch Neurol Psychiatry 46: 331–339, 1941. [Google Scholar]
254. Lassen NA. The luxury-perfusion syndrome and its possible relation to acute metabolic acidosis localised within the brain. Lancet 2: 1113–1115, 1966. [Abstract] [Google Scholar]
255. Lauritzen M. Long-lasting reduction of cortical blood flow of the brain after spreading depression with preserved autoregulation and impaired CO2 response. J Cereb Blood Flow Metab 4: 546–554, 1984. [Abstract] [Google Scholar]
256. Lauritzen M. Pathophysiology of the migraine aura. The spreading depression theory. Brain 117: 199–210, 1994. [Abstract] [Google Scholar]
257. Lauritzen M. Regional cerebral blood flow during cortical spreading depression in rat brain: increased reactive hyperperfusion in low-flow states. Acta Neurol Scand 75: 1–8, 1987. [Abstract] [Google Scholar]
258. Lauritzen M, Diemer NH. Uncoupling of cerebral blood flow and metabolism after single episode of cortical spreading depression in the rat brain. Brain Res 370: 405–408, 1986. [Abstract] [Google Scholar]
259. Lauritzen M, Dreier JP, Fabricius M, Hartings JA, Graf R, Strong AJ. Clinical relevance of cortical spreading depression in neurological disorders: migraine, malignant stroke, subarachnoid and intracranial hemorrhage, and traumatic brain injury. J Cereb Blood Flow Metab 31: 17–35, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
260. Lauritzen M, Fabricius M. Real time laser-Doppler perfusion imaging of cortical spreading depression in rat neocortex. Neuroreport 6: 1271–1273, 1995. [Abstract] [Google Scholar]
261. Lauritzen M, Hansen AJ. The effect of glutamate receptor blockade on anoxic depolarization and cortical spreading depression. J Cereb Blood Flow Metab 12: 223–229, 1992. [Abstract] [Google Scholar]
262. Lauritzen M, Hansen AJ, Kronborg D, Wieloch T. Cortical spreading depression is associated with arachidonic acid accumulation and preservation of energy charge. J Cereb Blood Flow Metab 10: 115–122, 1990. [Abstract] [Google Scholar]
263. Lauritzen M, Jorgensen MB, Diemer NH, Gjedde A, Hansen AJ. Persistent oligemia of rat cerebral cortex in the wake of spreading depression. Ann Neurol 12: 469–474, 1982. [Abstract] [Google Scholar]
264. Lauritzen M, Olesen J. Regional cerebral blood flow during migraine attacks by Xenon-133 inhalation and emission tomography. Brain 107: 447–461, 1984. [Abstract] [Google Scholar]
265. Lauritzen M, Olsen TS, Lassen NA, Paulson OB. Regulation of regional cerebral blood flow during and between migraine attacks. Ann Neurol 14: 569–572, 1983. [Abstract] [Google Scholar]
266. Lauritzen M, Rice ME, Okada Y, Nicholson C. Quisqualate, kainate and NMDA can initiate spreading depression in the turtle cerebellum. Brain Res 475: 317–327, 1988. [Abstract] [Google Scholar]
267. Lauritzen M, Skyhoj Olsen T, Lassen NA, Paulson OB. Changes in regional cerebral blood flow during the course of classic migraine attacks. Ann Neurol 13: 633–641, 1983. [Abstract] [Google Scholar]
268. Leao AAP. Further observations on spreading depression of activity in the cerebral cortex. J Neurophysiol 10: 409–414, 1947. [Abstract] [Google Scholar]
269. Leao AAP. Pial circulation and spreading depression of activity in the cerebral cortex. J Neurophysiol 7: 391–396, 1944. [Google Scholar]
270. Leao AAP. Spreading depression of activity in cerebral cortex. J Neurophysiol 7: 359–390, 1944. [Google Scholar]
271. Leao AAP, Morison RS. Propagation of spreading cortical depression. J Neurophysiol 8: 33–45, 1945. [Google Scholar]
272. Lecrux C, Toussay X, Kocharyan A, Fernandes P, Neupane S, Levesque M, Plaisier F, Shmuel A, Cauli B, Hamel E. Pyramidal neurons are “neurogenic hubs” in the neurovascular coupling response to whisker stimulation. J Neurosci 31: 9836–9847, 2011. [Abstract] [Google Scholar]
273. Lehmenkuhler A, Speckmann EJ, Caspers H. Cortical spreading depression in relation to potassium activity, oxygen tension, local flow and carbon dioxide tension. In: Ion and Enzyme Electrodes in Biology and Medicine, edited by Kessler M., Clark L.C., Lübbers D.W., Silver I.A., and Simon W. München, Germany: Urban and Scwerzenberg, 1976, p. 311–315. [Google Scholar]
274. Libet B, Gerard RW. Control of the potential rhythm of the isolated frog brain. J Neurophysiol 153–169, 1939. [Google Scholar]
275. Lind BL, Brazhe AR, Jessen SB, Tan FC, Lauritzen MJ. Rapid stimulus-evoked astrocyte Ca2+ elevations and hemodynamic responses in mouse somatosensory cortex in vivo. Proc Natl Acad Sci USA 110: E4678–4687, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
276. Lindauer U, Leithner C, Kaasch H, Rohrer B, Foddis M, Fuchtemeier M, Offenhauser N, Steinbrink J, Royl G, Kohl-Bareis M, Dirnagl U. Neurovascular coupling in rat brain operates independent of hemoglobin deoxygenation. J Cereb Blood Flow Metab 30: 757–768, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
277. Lindquist BE, Shuttleworth CW. Adenosine receptor activation is responsible for prolonged depression of synaptic transmission after spreading depolarization in brain slices. Neuroscience 223: 365–376, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
278. Lindquist BE, Shuttleworth CW. Spreading depolarization-induced adenosine accumulation reflects metabolic status in vitro and in vivo. J Cereb Blood Flow Metab 34: 1779–1790, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
279. Lo W, Stephens J, Fernandez S. Pediatric stroke in the United States and the impact of risk factors. J Child Neurol 24: 194–203, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
280. Lothman E, Lamanna J, Cordingley G, Rosenthal M, Somjen G. Responses of electrical potential, potassium levels, and oxidative metabolic activity of the cerebral neocortex of cats. Brain Res 88: 15–36, 1975. [Abstract] [Google Scholar]
281. Louis TM, Meng W, Bari F, Errico RA, Busija DW. Ischemia reduces CGRP-induced cerebral vascular dilation in piglets. Stroke 27: 134–139, 1996. [Abstract] [Google Scholar]
282. Luckl J, Dreier JP, Szabados T, Wiesenthal D, Bari F, Greenberg JH. Peri-infarct flow transients predict outcome in rat focal brain ischemia. Neuroscience 226: 197–207, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
283. Luckl J, Zhou C, Durduran T, Yodh AG, Greenberg JH. Characterization of periinfarct flow transients with laser speckle and Doppler after middle cerebral artery occlusion in the rat. J Neurosci Res 87: 1219–1229, 2009. [Abstract] [Google Scholar]
284. Luckl J, Zhou C, Durduran T, Yodh AG, Greenberg JH. Characterization of periinfarct flow transients with laser speckle and Doppler after middle cerebral artery occlusion in the rat. J Neurosci Res 87: 1219–1229, 2009. [Abstract] [Google Scholar]
285. Lukyanova LD, Bures J. Changes in pO2 due to spreading depression in the cortex and nucleus caudatus of the rat. Physiol Bohemoslov 16: 449–455, 1967. [Abstract] [Google Scholar]
286. Madsen PL, Cruz NF, Sokoloff L, Dienel GA. Cerebral oxygen/glucose ratio is low during sensory stimulation and rises above normal during recovery: excess glucose consumption during stimulation is not accounted for by lactate efflux from or accumulation in brain tissue. J Cereb Blood Flow Metab 19: 393–400, 1999. [Abstract] [Google Scholar]
287. Malonek D, Dirnagl U, Lindauer U, Yamada K, Kanno I, Grinvald A. Vascular imprints of neuronal activity: relationships between the dynamics of cortical blood flow, oxygenation, and volume changes following sensory stimulation. Proc Natl Acad Sci USA 94: 14826–14831, 1997. [Europe PMC free article] [Abstract] [Google Scholar]
288. Margaria R, Edwards HT, Dill DB. The possible mechanisms of contracting and paying the oxygen debt and the role of lactic acid in muscular contraction. Am J Physiol 106: 689–715, 1933. [Google Scholar]
289. Marrannes R, Willems R, De Prins E, Wauquier A. Evidence for a role of the N-methyl-d-aspartate (NMDA) receptor in cortical spreading depression in the rat. Brain Res 457: 226–240, 1988. [Abstract] [Google Scholar]
290. Marshall WH. Spreading cortical depression of Leao. Physiol Rev 39: 239–279, 1959. [Abstract] [Google Scholar]
291. Marshall WH, Essig CF, Dubroff SJ. Relation of temperature of cerebral cortex to spreading depression of Leao. J Neurophysiol 14: 153–166, 1951. [Abstract] [Google Scholar]
292. Martins-Ferreira H, De Oliveira Castro G, Struchiner CJ, Rodrigues PS. Circling spreading depression in isolated chick retina. J Neurophysiol 37: 773–784, 1974. [Abstract] [Google Scholar]
293. Martins-Ferreira H, Nedergaard M, Nicholson C. Perspectives on spreading depression. Brain Res 32: 215–234, 2000. [Abstract] [Google Scholar]
294. Maslarova A, Alam M, Reiffurth C, Lapilover E, Gorji A, Dreier JP. Chronically epileptic human and rat neocortex display a similar resistance against spreading depolarization in vitro. Stroke 42: 2917–2922, 2011. [Abstract] [Google Scholar]
295. Matsuura T, Bures J. The minimum volume of depolarized neural tissue required for triggering cortical spreading depression in rat. Exp Brain Res 12: 238–249, 1971. [Abstract] [Google Scholar]
296. Mayevsky A. Brain energy metabolism of the conscious rat exposed to various physiological and pathological situations. Brain Res 113: 327–338, 1976. [Abstract] [Google Scholar]
297. Mayevsky A. Pyridine nucleotide oxidation-reduction state of the cerebral cortex in the awake gerbil. J Neurosci Res 3: 369–374, 1978. [Abstract] [Google Scholar]
298. Mayevsky A, Bar-Sagie D. The interrelation between CBF, energy metabolism and ECoG in a new awake brain model. Adv Exp Med Biol 94: 761–767, 1977. [Abstract] [Google Scholar]
299. Mayevsky A, Chance B. Intracellular oxidation-reduction state measured in situ by a multichannel fiber-optic surface fluorometer. Science 217: 537–540, 1982. [Abstract] [Google Scholar]
300. Mayevsky A, Chance B. Metabolic responses of the awake cerebral cortex to anoxia hypoxia spreading depression and epileptiform activity. Brain Res 98: 149–165, 1975. [Abstract] [Google Scholar]
301. Mayevsky A, Chance B. Repetitive patterns of metabolic changes during cortical spreading depression of the awake rat. Brain Res 65: 529–533, 1974. [Abstract] [Google Scholar]
302. Mayevsky A, Doron A, Manor T, Meilin S, Zarchin N, Ouaknine GE. Cortical spreading depression recorded from the human brain using a multiparametric monitoring system. Brain Res 740: 268–274, 1996. [Abstract] [Google Scholar]
303. Mayevsky A, Duckrow RB, Yoles E, Zarchin N, Kaushausky D. Brain mitochondrial redox state, tissue haemodynamic and extracellular ion responses to four-vessel occlusion and spreading depression in the rat. Neurol Res 12: 243–248, 1990. [Abstract] [Google Scholar]
304. Mayevsky A, Lebourdais S, Chance B. The interrelation between brain PO2 and NADH oxidation-reduction state in the gerbil. J Neurosci Res 5: 173–182, 1980. [Abstract] [Google Scholar]
305. Mayevsky A, Weiss HR. Cerebral blood flow and oxygen consumption in cortical spreading depression. J Cereb Blood Flow Metab 11: 829–836, 1991. [Abstract] [Google Scholar]
306. Mayevsky A, Zarchin N, Friedli CM. Factors affecting the oxygen balance in the awake cerebral cortex exposed to spreading depression. Brain Res 236: 93–105, 1982. [Abstract] [Google Scholar]
307. Mayevsky A, Zeuthen T, Chance B. Measurements of extracellular potassium, ECoG and pyridine nucleotide levels during cortical spreading depression in rats. Brain Res 76: 347–349, 1974. [Abstract] [Google Scholar]
308. Mazel T, Richter F, Vargova L, Sykova E. Changes in extracellular space volume and geometry induced by cortical spreading depression in immature and adult rats. Physiol Res 51 Suppl 1: S85–93, 2002. [Abstract] [Google Scholar]
309. McCarron JG, Halpern W. Potassium dilates rat cerebral arteries by two independent mechanisms. Am J Physiol Heart Circ Physiol 259: H902–H908, 1990. [Abstract] [Google Scholar]
310. McCulloch J, Edvinsson L, Watt P. Comparison of the effects of potassium and pH on the calibre of cerebral veins and arteries. Pflügers Arch 393: 95–98, 1982. [Abstract] [Google Scholar]
311. Meirovithz E, Sonn J, Mayevsky A. Effect of hyperbaric oxygenation on brain hemodynamics, hemoglobin oxygenation and mitochondrial NADH. Brain Res Rev 54: 294–304, 2007. [Abstract] [Google Scholar]
312. Mejia JA, Pernow J, von Holst H, Rudehill A, Lundberg JM. Effects of neuropeptide Y, calcitonin gene-related peptide, substance P, and capsaicin on cerebral arteries in man and animals. J Neurosurg 69: 913–918, 1988. [Abstract] [Google Scholar]
313. Meng W, Ayata C, Waeber C, Huang PL, Moskowitz MA. Neuronal NOS-cGMP-dependent ACh-induced relaxation in pial arterioles of endothelial NOS knockout mice. Am J Physiol Heart Circ Physiol 274: H411–H415, 1998. [Abstract] [Google Scholar]
314. Meng W, Busija DW. Oxygen radicals do not play a role in arteriolar dilation during cortical spreading depression. J Cereb Blood Flow Metab 16: 175–179, 1996. [Abstract] [Google Scholar]
315. Meng W, Colonna DM, Tobin JR, Busija DW. Nitric oxide and prostaglandins interact to mediate arteriolar dilation during cortical spreading depression. Am J Physiol Heart Circ Physiol 269: H176–H181, 1995. [Abstract] [Google Scholar]
316. Meng W, Ma J, Ayata C, Hara H, Huang PL, Fishman MC, Moskowitz MA. ACh dilates pial arterioles in endothelial and neuronal NOS knockout mice by NO-dependent mechanisms. Am J Physiol Heart Circ Physiol 271: H1145–H1150, 1996. [Abstract] [Google Scholar]
317. Meng W, Tobin JR, Busija DW. Glutamate-induced cerebral vasodilation is mediated by nitric oxide through N-methyl-d-aspartate receptors. Stroke 26: 857–863, 1995. [Abstract] [Google Scholar]
318. Metea MR, Newman EA. Glial cells dilate and constrict blood vessels: a mechanism of neurovascular coupling. J Neurosci 26: 2862–2870, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
319. Mies G, Iijima T, Hossmann KA. Correlation between peri-infarct DC shifts and ischaemic neuronal damage in rat. Neuroreport 4: 709–711, 1993. [Abstract] [Google Scholar]
320. Mies G, Niebuhr I, Hossmann KA. Simultaneous measurement of blood flow and glucose metabolism by autoradiographic techniques. Stroke 12: 581–588, 1981. [Abstract] [Google Scholar]
321. Mies G, Paschen W. Regional changes of blood flow, glucose, and ATP content determined on brain sections during a single passage of spreading depression in rat brain cortex. Exp Neurol 84: 249–258, 1984. [Abstract] [Google Scholar]
322. Minato H, Hashizume M, Masuda Y, Hosoki K. Modulation of extraluminally induced vasoconstrictions by endothelium-derived nitric oxide in the canine basilar artery. Arzneimittelforschung 45: 675–678, 1995. [Abstract] [Google Scholar]
323. Miyata N, Taniguchi K, Seki T, Ishimoto T, Sato-Watanabe M, Yasuda Y, Doi M, Kametani S, Tomishima Y, Ueki T, Sato M, Kameo K. HET0016, a potent and selective inhibitor of 20-HETE synthesizing enzyme. Br J Pharmacol 133: 325–329, 2001. [Europe PMC free article] [Abstract] [Google Scholar]
324. Mody I, Lambert JD, Heinemann U. Low extracellular magnesium induces epileptiform activity and spreading depression in rat hippocampal slices. J Neurophysiol 57: 869–888, 1987. [Abstract] [Google Scholar]
325. Moghaddam B, Schenk JO, Stewart WB, Hansen AJ. Temporal relationship between neurotransmitter release and ion flux during spreading depression and anoxia. Can J Physiol Pharmacol 65: 1105–1110, 1987. [Abstract] [Google Scholar]
326. Molchanova S, Koobi P, Oja SS, Saransaari P. Interstitial concentrations of amino acids in the rat striatum during global forebrain ischemia and potassium-evoked spreading depression. Neurochem Res 29: 1519–1527, 2004. [Abstract] [Google Scholar]
327. Mraovitch S, Calando Y, Goadsby PJ, Seylaz J. Subcortical cerebral blood flow and metabolic changes elicited by cortical spreading depression in rat. Cephalalgia 12: 137–141, 1992. [Abstract] [Google Scholar]
328. Mulligan SJ, MacVicar BA. Calcium transients in astrocyte endfeet cause cerebrovascular constrictions. Nature 431: 195–199, 2004. [Abstract] [Google Scholar]
329. Mutch WA, Hansen AJ. Extracellular pH changes during spreading depression and cerebral ischemia: mechanisms of brain pH regulation. J Cereb Blood Flow Metab 4: 17–27, 1984. [Abstract] [Google Scholar]
330. Nakamura H, Strong AJ, Dohmen C, Sakowitz OW, Vollmar S, Sue M, Kracht L, Hashemi P, Bhatia R, Yoshimine T, Dreier JP, Dunn AK, Graf R. Spreading depolarizations cycle around and enlarge focal ischaemic brain lesions. Brain 133: 1994–2006, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
331. Nam MJ, Thore C, Busija D. Effects of protein kinase C activation on prostaglandin production and cyclooxygenase mRNA levels in ovine astroglia. Prostaglandins 51: 203–213, 1996. [Abstract] [Google Scholar]
332. Nedergaard M, Astrup J. Infarct rim: effect of hyperglycemia on direct current potential and [14C]2-deoxyglucose phosphorylation. J Cereb Blood Flow Metab 6: 607–615, 1986. [Abstract] [Google Scholar]
333. Nedergaard M, Hansen AJ. Characterization of cortical depolarizations evoked in focal cerebral ischemia. J Cereb Blood Flow Metab 13: 568–574, 1993. [Abstract] [Google Scholar]
334. Nedergaard M, Hansen AJ. Spreading depression is not associated with neuronal injury in the normal brain. Brain Res 449: 395–398, 1988. [Abstract] [Google Scholar]
335. Nemoto EM, Snyder JV, Carroll RG, Morita H. Global ischemia in dogs: cerebrovascular CO2 reactivity and autoregulation. Stroke 6: 425–431, 1975. [Abstract] [Google Scholar]
336. Newman EA. High potassium conductance in astrocyte endfeet. Science 233: 453–454, 1986. [Europe PMC free article] [Abstract] [Google Scholar]
337. Newman GC, Hospod FE, Schissel SL. Ischemic brain slice glucose utilization: effects of slice thickness, acidosis, and K+. J Cereb Blood Flow Metab 11: 398–406, 1991. [Abstract] [Google Scholar]
338. Ngai AC, Nguyen TS, Meno JR, Britz GW. Postischemic augmentation of conducted dilation in cerebral arterioles. Stroke 38: 124–130, 2007. [Abstract] [Google Scholar]
339. Nicholson C. Comparative neurophysiology of spreading depression in the cerebellum. An Acad Bras Cienc 56: 481–494, 1984. [Abstract] [Google Scholar]
340. Nicholson C, Bruggencate GT, Steinberg R, Stockle H. Calcium modulation in brain extracellular microenvironment demonstrated with ion-selective micropipette. Proc Natl Acad Sci USA 74: 1287–1290, 1977. [Europe PMC free article] [Abstract] [Google Scholar]
341. Nicholson C, Kraig RP. Chloride and potassium changes measured during spreading depression in catfish cerebellum. Brain Res 96: 384–389, 1975. [Abstract] [Google Scholar]
342. Nicholson C, ten Bruggencate G, Stockle H, Steinberg R. Calcium and potassium changes in extracellular microenvironment of cat cerebellar cortex. J Neurophysiol 41: 1026–1039, 1978. [Abstract] [Google Scholar]
343. Nielsen AN, Fabricius M, Lauritzen M. Scanning laser-Doppler flowmetry of rat cerebral circulation during cortical spreading depression. J Vasc Res 37: 513–522, 2000. [Abstract] [Google Scholar]
344. Nishimura N, Schaffer CB, Friedman B, Lyden PD, Kleinfeld D. Penetrating arterioles are a bottleneck in the perfusion of neocortex. Proc Natl Acad Sci USA 104: 365–370, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
345. Niwa K, Araki E, Morham SG, Ross ME, Iadecola C. Cyclooxygenase-2 contributes to functional hyperemia in whisker-barrel cortex. J Neurosci 20: 763–770, 2000. [Abstract] [Google Scholar]
346. Norris JW, Hachinski VC, Cooper PW. Changes in cerebral blood flow during a migraine attack. Br Med J 3: 676–677, 1975. [Europe PMC free article] [Abstract] [Google Scholar]
347. Northington FJ, Tobin JR, Koehler RC, Traystman RJ. In vivo production of nitric oxide correlates with NMDA-induced cerebral hyperemia in newborn sheep. Am J Physiol Heart Circ Physiol 269: H215–H221, 1995. [Abstract] [Google Scholar]
348. Nozaki K, Moskowitz MA, Maynard KI, Koketsu N, Dawson TM, Bredt DS, Snyder SH. Possible origins and distribution of immunoreactive nitric oxide synthase-containing nerve fibers in cerebral arteries. J Cereb Blood Flow Metab 13: 70–79, 1993. [Abstract] [Google Scholar]
349. Nozari A, Dilekoz E, Sukhotinsky I, Stein T, Eikermann-Haerter K, Liu C, Wang Y, Frosch MP, Waeber C, Ayata C, Moskowitz MA. Microemboli may link spreading depression, migraine aura, and patent foramen ovale. Ann Neurol 67: 221–229, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
350. O'Brien MD. Cerebral-cortex-perfusion rates in migraine. Lancet 1: 1036, 1967. [Abstract] [Google Scholar]
351. O'Farrell AM, Rex DE, Muthialu A, Pouratian N, Wong GK, Cannestra AF, Chen JW, Toga AW. Characterization of optical intrinsic signals and blood volume during cortical spreading depression. Neuroreport 11: 2121–2125, 2000. [Abstract] [Google Scholar]
352. Obrenovitch TP, Chen S, Farkas E. Simultaneous, live imaging of cortical spreading depression and associated cerebral blood flow changes, by combining voltage-sensitive dye and laser speckle contrast methods. Neuroimage 45: 68–74, 2009. [Abstract] [Google Scholar]
353. Obrenovitch TP, Urenjak J, Wang M. Nitric oxide formation during cortical spreading depression is critical for rapid subsequent recovery of ionic homeostasis. J Cereb Blood Flow Metab 22: 680–688, 2002. [Abstract] [Google Scholar]
354. Obrig H, Hirth C, Junge-Hulsing JG, Doge C, Wenzel R, Wolf T, Dirnagl U, Villringer A. Length of resting period between stimulation cycles modulates hemodynamic response to a motor stimulus. Adv Exp Med Biol 411: 471–480, 1997. [Abstract] [Google Scholar]
355. Offenhauser N, Windmuller O, Strong AJ, Fuhr S, Dreier JP. The gamut of blood flow responses coupled to spreading depolarization in rat and human brain: from hyperemia to prolonged ischemia. Acta Neurochir Suppl 110: 119–124, 2011. [Abstract] [Google Scholar]
356. Olah G, Heredi J, Menyhart A, Czinege Z, Nagy D, Fuzik J, Kocsis K, Knapp L, Krucso E, Gellert L, Kis Z, Farkas T, Fulop F, Pardutz A, Tajti J, Vecsei L, Toldi J. Unexpected effects of peripherally administered kynurenic acid on cortical spreading depression and related blood-brain barrier permeability. Drug Design Dev Ther 7: 981–987, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
357. Olesen J, Larsen B, Lauritzen M. Focal hyperemia followed by spreading oligemia and impaired activation of rCBF in classic migraine. Ann Neurol 9: 344–352, 1981. [Abstract] [Google Scholar]
358. Orakcioglu B, Becker K, Sakowitz OW, Herweh C, Kohrmann M, Huttner HB, Steiner T, Unterberg A, Schellinger PD. MRI of the perihemorrhagic zone in a rat ICH model: effect of hematoma evacuation. Neurocrit Care 8: 448–455, 2008. [Abstract] [Google Scholar]
359. Osada T, Tomita M, Suzuki N. Spindle-shaped constriction and propagated dilation of arterioles during cortical spreading depression. Neuroreport 17: 1365–1368, 2006. [Abstract] [Google Scholar]
360. Otori T, Greenberg JH, Welsh FA. Cortical spreading depression causes a long-lasting decrease in cerebral blood flow and induces tolerance to permanent focal ischemia in rat brain. J Cereb Blood Flow Metab 23: 43–50, 2003. [Abstract] [Google Scholar]
361. Otsuka H, Ueda K, Heimann A, Kempski O. Effects of cortical spreading depression on cortical blood flow, impedance, DC potential, and infarct size in a rat venous infarct model. Exp Neurol 162: 201–214, 2000. [Abstract] [Google Scholar]
362. Park CK, Nehls DG, Teasdale GM, McCulloch J. Effect of the NMDA antagonist MK-801 on local cerebral blood flow in focal cerebral ischaemia in the rat. J Cereb Blood Flow Metab 9: 617–622, 1989. [Abstract] [Google Scholar]
363. Park L, Gallo EF, Anrather J, Wang G, Norris EH, Paul J, Strickland S, Iadecola C. Key role of tissue plasminogen activator in neurovascular coupling. Proc Natl Acad Sci USA 105: 1073–1078, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
364. Parkin M, Hopwood S, Jones DA, Hashemi P, Landolt H, Fabricius M, Lauritzen M, Boutelle MG, Strong AJ. Dynamic changes in brain glucose and lactate in pericontusional areas of the human cerebral cortex, monitored with rapid sampling on-line microdialysis: relationship with depolarisation-like events. J Cereb Blood Flow Metab 25: 402–413, 2005. [Abstract] [Google Scholar]
365. Paspalas CD, Papadopoulos GC. Ultrastructural relationships between noradrenergic nerve fibers and non-neuronal elements in the rat cerebral cortex. Glia 17: 133–146, 1996. [Abstract] [Google Scholar]
366. Paulson OB, Newman EA. Does the release of potassium from astrocyte endfeet regulate cerebral blood flow? Science 237: 896–898, 1987. [Europe PMC free article] [Abstract] [Google Scholar]
367. Peters O, Schipke CG, Hashimoto Y, Kettenmann H. Different mechanisms promote astrocyte Ca2+ waves and spreading depression in the mouse neocortex. J Neurosci 23: 9888–9896, 2003. [Abstract] [Google Scholar]
368. Petzold GC, Einhaupl KM, Dirnagl U, Dreier JP. Ischemia triggered by spreading neuronal activation is induced by endothelin-1 and hemoglobin in the subarachnoid space. Ann Neurol 54: 591–598, 2003. [Abstract] [Google Scholar]
369. Petzold GC, Haack S, von Bohlen Und Halbach O, Priller J, Lehmann TN, Heinemann U, Dirnagl U, Dreier JP. Nitric oxide modulates spreading depolarization threshold in the human and rodent cortex. Stroke 39: 1292–1299, 2008. [Abstract] [Google Scholar]
370. Phillips JM, Nicholson C. Anion permeability in spreading depression investigated with ion-sensitive microelectrodes. Brain Res 173: 567–571, 1979. [Abstract] [Google Scholar]
371. Piilgaard H, Lauritzen M. Persistent increase in oxygen consumption and impaired neurovascular coupling after spreading depression in rat neocortex. J Cereb Blood Flow Metab 29: 1517–1527, 2009. [Abstract] [Google Scholar]
373. Piilgaard H, Witgen BM, Rasmussen P, Lauritzen M. Cyclosporine A, FK506, and NIM811 ameliorate prolonged CBF reduction and impaired neurovascular coupling after cortical spreading depression. J Cereb Blood Flow Metab 31: 1588–1598, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
374. Pinard E, Nallet H, MacKenzie ET, Seylaz J, Roussel S. Penumbral microcirculatory changes associated with peri-infarct depolarizations in the rat. Stroke 33: 606–612, 2002. [Abstract] [Google Scholar]
375. Piper RD, Lambert GA. Inhalational anesthetics inhibit spreading depression: relevance to migraine. Cephalalgia 16: 87–92, 1996. [Abstract] [Google Scholar]
376. Piper RD, Lambert GA, Duckworth JW. Cortical blood flow changes during spreading depression in cats. Am J Physiol Heart Circ Physiol 261: H96–H102, 1991. [Abstract] [Google Scholar]
377. Price DL, Ludwig JW, Mi H, Schwarz TL, Ellisman MH. Distribution of rSlo Ca2+-activated K+ channels in rat astrocyte perivascular endfeet. Brain Res 956: 183–193, 2002. [Abstract] [Google Scholar]
378. Prince DA, Lux HD, Neher E. Measurement of extracellular potassium activity in cat cortex. Brain Res 50: 489–495, 1973. [Abstract] [Google Scholar]
379. Raichle ME, Hartman BK, Eichling JO, Sharpe LG. Central noradrenergic regulation of cerebral blood flow and vascular permeability. Proc Natl Acad Sci USA 72: 3726–3730, 1975. [Europe PMC free article] [Abstract] [Google Scholar]
380. Rapoport SI, Marshall WH. Measurement of cortical pH in spreading cortical depression. Am J Physiol 206: 1177–1180, 1964. [Abstract] [Google Scholar]
381. Raymond SB, Skoch J, Hynynen K, Bacskai BJ. Multiphoton imaging of ultrasound/Optison mediated cerebrovascular effects in vivo. J Cereb Blood Flow Metab 27: 393–403, 2007. [Abstract] [Google Scholar]
382. Read SJ, Parsons AA. Nitric oxide does not mediate cerebral blood flow changes during cortical spreading depression in the anaesthetised rat. Neurosci Lett 250: 115–118, 1998. [Abstract] [Google Scholar]
383. Read SJ, Parsons AA. Sumatriptan modifies cortical free radical release during cortical spreading depression. A novel antimigraine action for sumatriptan? Brain Res 870: 44–53, 2000. [Abstract] [Google Scholar]
384. Read SJ, Smith MI, Benham CD, Hunter AJ, Parsons AA. Furosemide inhibits regenerative cortical spreading depression in anaesthetized cats. Cephalalgia 17: 826–832, 1997. [Abstract] [Google Scholar]
385. Read SJ, Smith MI, Hunter AJ, Parsons AA. The dynamics of nitric oxide release measured directly and in real time following repeated waves of cortical spreading depression in the anaesthetised cat. Neurosci Lett 232: 127–130, 1997. [Abstract] [Google Scholar]
386. Read SJ, Smith MI, Hunter AJ, Parsons AA. Enhanced nitric oxide release during cortical spreading depression following infusion of glyceryl trinitrate in the anaesthetized cat. Cephalalgia 17: 159–165, 1997. [Abstract] [Google Scholar]
387. Read SJ, Smith MI, Hunter AJ, Upton N, Parsons AA. SB-220453, a potential novel antimigraine agent, inhibits nitric oxide release following induction of cortical spreading depression in the anaesthetized cat. Cephalalgia 20: 92–99, 2000. [Abstract] [Google Scholar]
388. Rebert CS. Spreading depression in squirrel monkey lissencephalic cortex. Physiol Behav 5: 239–241, 1970. [Abstract] [Google Scholar]
389. Reuter U, Weber JR, Gold L, Arnold G, Wolf T, Dreier J, Lindauer U, Dirnagl U. Perivascular nerves contribute to cortical spreading depression-associated hyperemia in rats. Am J Physiol Heart Circ Physiol 274: H1979–H1987, 1998. [Abstract] [Google Scholar]
390. Rex A, Pfeifer L, Fink F, Fink H. Cortical NADH during pharmacological manipulations of the respiratory chain and spreading depression in vivo. J Neurosci Res 57: 359–370, 1999. [Abstract] [Google Scholar]
391. Rice ME, Nicholson C. Interstitial ascorbate in turtle brain is modulated by release and extracellular volume change. J Neurochem 49: 1096–1104, 1987. [Abstract] [Google Scholar]
392. Richter F, Bauer R, Lehmenkuhler A, Schaible HG. Spreading depression in the brainstem of the adult rat: electrophysiological parameters and influences on regional brainstem blood flow. J Cereb Blood Flow Metab 28: 984–994, 2008. [Abstract] [Google Scholar]
393. Rodgers CI, Armstrong GA, Shoemaker KL, Labrie JD, Moyes CD, Robertson RM. Stress preconditioning of spreading depression in the locust CNS. PLoS One 2: e1366, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
394. Rogers ML, Feuerstein D, Leong CL, Takagaki M, Niu X, Graf R, Boutelle MG. Continuous online microdialysis using microfluidic sensors: dynamic neurometabolic changes during spreading depolarization. ACS Chem Neurosci 4: 799–807, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
395. Roman RJ. P-450 metabolites of arachidonic acid in the control of cardiovascular function. Physiol Rev 82: 131–185, 2002. [Abstract] [Google Scholar]
396. Rosenthal M, LaManna JC. Oxidative metabolism and electrophysiological activity in intact central nervous system. In: Oxygen and Physiological Function, edited by Jobsis F. F., editor. Dallas, TX: Professional Information Library, 1977, p. 515–530. [Google Scholar]
397. Rosenthal M, Martel DL. Ischemia-induced alterations in oxidative “recovery” metabolism after spreading cortical depression in situ. Exp Neurol 63: 367–378, 1979. [Abstract] [Google Scholar]
398. Rosenthal M, Somjen G. Spreading depression, sustained potential shifts, and metabolic activity of cerebral cortex of cats. J Neurophysiol 36: 739–749, 1973. [Abstract] [Google Scholar]
399. Rounds HD. KC1-induced “spreading depression” in the cockroach. J Insect Physiol 13: 869–872, 1967. [Abstract] [Google Scholar]
400. Rubio R, Berne RM, Bockman EL, Curnish RR. Relationship between adenosine concentration and oxygen supply in rat brain. Am J Physiol 228: 1896–1902, 1975. [Abstract] [Google Scholar]
401. Rundek T, Elkind MS, Di Tullio MR, Carrera E, Jin Z, Sacco RL, Homma S. Patent Foramen Ovale and Migraine. A Cross-Sectional Study From the Northern Manhattan Study (NOMAS). Circulation 118: 1419–1424, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
402. Ryman T, Brandt L, andersson KE, Mellergard P. Regional and species differences in vascular reactivity to extracellular potassium. Acta Physiol Scand 136: 151–159, 1989. [Abstract] [Google Scholar]
403. Saez I, Duran J, Sinadinos C, Beltran A, Yanes O, Tevy MF, Martinez-Pons C, Milan M, Guinovart JJ. Neurons have an active glycogen metabolism that contributes to tolerance to hypoxia. J Cereb Blood Flow Metab 34: 945–955, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
404. Saito R, Graf R, Hubel K, Taguchi J, Rosner G, Fujita T, Heiss WD. Halothane, but not alpha-chloralose, blocks potassium-evoked cortical spreading depression in cats. Brain Res 699: 109–115, 1995. [Abstract] [Google Scholar]
405. Sakadzic S, Roussakis E, Yaseen MA, Mandeville ET, Srinivasan VJ, Arai K, Ruvinskaya S, Devor A, Lo EH, Vinogradov SA, Boas DA. Two-photon high-resolution measurement of partial pressure of oxygen in cerebral vasculature and tissue. Nature Methods 7: 755–759, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
406. Sakadzic S, Yuan S, Dilekoz E, Ruvinskaya S, Vinogradov SA, Ayata C, Boas DA. Simultaneous imaging of cerebral partial pressure of oxygen and blood flow during functional activation and cortical spreading depression. Appl Opt 48: D169–177, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
407. Sato S, Kawauchi S, Okuda W, Nishidate I, Nawashiro H, Tsumatori G. Real-time optical diagnosis of the rat brain exposed to a laser-induced shock wave: observation of spreading depolarization, vasoconstriction and hypoxemia-oligemia. PLoS One 9: e82891, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
408. Sattler R, Xiong Z, Lu WY, Hafner M, MacDonald JF, Tymianski M. Specific coupling of NMDA receptor activation to nitric oxide neurotoxicity by PSD-95 protein. Science 284: 1845–1848, 1999. [Abstract] [Google Scholar]
409. Schechter M, Sonn J, Mayevsky A. Brain oxygen balance under various experimental pathophysiologycal conditions. Adv Exp Med Biol 645: 293–299, 2009. [Abstract] [Google Scholar]
410. Scheckenbach KE, Dreier JP, Dirnagl U, Lindauer U. Impaired cerebrovascular reactivity after cortical spreading depression in rats: restoration by nitric oxide or cGMP. Exp Neurol 202: 449–455, 2006. [Abstract] [Google Scholar]
411. Scheller D, Kolb J, Tegtmeier F. Lactate and pH change in close correlation in the extracellular space of the rat brain during cortical spreading depression. Neurosci Lett 135: 83–86, 1992. [Abstract] [Google Scholar]
412. Scheller D, Kolb J, Tegtmeier F, Lehmenkuhler A. Extracellular changes of inorganic phosphate are different during spreading depression and global cerebral ischemia of rats. Neurosci Lett 141: 269–272, 1992. [Abstract] [Google Scholar]
413. Schmitz B, Bottiger BW, Hossmann KA. Functional activation of cerebral blood flow after cardiac arrest in rat. J Cereb Blood Flow Metab 17: 1202–1209, 1997. [Abstract] [Google Scholar]
414. Scremin OU, Rovere AA, Raynald AC, Giardini A. Cholinergic control of blood flow in the cerebral cortex of the rat. Stroke 4: 233–239, 1973. [Abstract] [Google Scholar]
415. Seitz I, Dirnagl U, Lindauer U. Impaired vascular reactivity of isolated rat middle cerebral artery after cortical spreading depression in vivo. J Cereb Blood Flow Metab 24: 526–530, 2004. [Abstract] [Google Scholar]
416. Selman WR, Lust WD, Pundik S, Zhou Y, Ratcheson RA. Compromised metabolic recovery following spontaneous spreading depression in the penumbra. Brain Res 999: 167–174, 2004. [Abstract] [Google Scholar]
417. Serres S, Raffard G, Franconi JM, Merle M. Close coupling between astrocytic and neuronal metabolisms to fulfill anaplerotic and energy needs in the rat brain. J Cereb Blood Flow Metab 28: 712–724, 2008. [Abstract] [Google Scholar]
418. Sevgi EB, Erdener SE, Demirci M, Topcuoglu MA, Dalkara T. Paradoxical air microembolism induces cerebral bioelectrical abnormalities and occasionally headache in patent foramen ovale patients with migraine. JAMA 1: e001735, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
419. Shibata M, Leffler CW, Busija DW. Cerebral hemodynamics during cortical spreading depression in rabbits. Brain Res 530: 267–274, 1990. [Abstract] [Google Scholar]
420. Shibata M, Leffler CW, Busija DW. Evidence against parenchymal metabolites directly promoting pial arteriolar dilation during cortical spreading depression in rabbits. Brain Res Bull 26: 753–758, 1991. [Abstract] [Google Scholar]
421. Shibata M, Leffler CW, Busija DW. Pial arteriolar constriction following cortical spreading depression is mediated by prostanoids. Brain Res 572: 190–197, 1992. [Abstract] [Google Scholar]
422. Shibata M, Leffler CW, Busija DW. Prostanoids attenuate pial arteriolar dilation induced by cortical spreading depression in rabbits. Am J Physiol Regul Integr Comp Physiol 261: R828–R834, 1991. [Abstract] [Google Scholar]
423. Shih AY, Friedman B, Drew PJ, Tsai PS, Lyden PD, Kleinfeld D. Active dilation of penetrating arterioles restores red blood cell flux to penumbral neocortex after focal stroke. J Cereb Blood Flow Metab 29: 738–751, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
424. Shimazawa M, Hara H. An experimental model of migraine with aura: cortical hypoperfusion following spreading depression in the awake and freely moving rat. Clin Exp Pharmacol Physiol 23: 890–892, 1996. [Abstract] [Google Scholar]
425. Shimazawa M, Hara H, Watano T, Sukamoto T. Effects of Ca2+ channel blockers on cortical hypoperfusion and expression of c-Fos-like immunoreactivity after cortical spreading depression in rats. Br J Pharmacol 115: 1359–1368, 1995. [Europe PMC free article] [Abstract] [Google Scholar]
426. Shimizu K, Bari F, Busija DW. Glibenclamide enhances cortical spreading depression-associated hyperemia in the rat. Neuroreport 11: 2103–2106, 2000. [Abstract] [Google Scholar]
427. Shimizu K, Miller AW, Erdos B, Bari F, Busija DW. Role of endothelium in hyperemia during cortical spreading depression (CSD) in the rat. Brain Res 928: 40–49, 2002. [Abstract] [Google Scholar]
428. Shimizu K, Veltkamp R, Busija DW. Characteristics of induced spreading depression after transient focal ischemia in the rat. Brain Res 861: 316–324, 2000. [Abstract] [Google Scholar]
429. Shin HK, Dunn AK, Jones PB, Boas DA, Lo EH, Moskowitz MA, Ayata C. Normobaric hyperoxia improves cerebral blood flow and oxygenation, and inhibits peri-infarct depolarizations in experimental focal ischaemia. Brain 130: 1631–1642, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
430. Shin HK, Dunn AK, Jones PB, Boas DA, Lo EH, Moskowitz MA, Ayata C. Normobaric hyperoxia improves cerebral blood flow and oxygenation, and inhibits peri-infarct depolarizations in experimental focal ischaemia. Brain 130: 1631–1642, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
431. Shin HK, Dunn AK, Jones PB, Boas DA, Moskowitz MA, Ayata C. Vasoconstrictive neurovascular coupling during focal ischemic depolarizations. J Cereb Blood Flow Metab 26: 1018–1030, 2006. [Abstract] [Google Scholar]
432. Shin HK, Jones PB, Garcia-Alloza M, Borrelli L, Greenberg SM, Bacskai BJ, Frosch MP, Hyman BT, Moskowitz MA, Ayata C. Age-dependent cerebrovascular dysfunction in a transgenic mouse model of cerebral amyloid angiopathy. Brain 130: 2310–2319, 2007. [Abstract] [Google Scholar]
433. Shinohara M, Dollinger B, Brown G, Rapoport S, Sokoloff L. Cerebral glucose utilization: local changes during and after recovery from spreading cortical depression. Science 203: 188–190, 1979. [Abstract] [Google Scholar]
434. Siemkowicz E. Cerebrovascular resistance in ischemia. Pflügers Arch 388: 243–247, 1980. [Abstract] [Google Scholar]
435. Simard D, Paulson OB. Cerebral vasomotor paralysis during migraine attack. Arch Neurol 29: 207–209, 1973. [Abstract] [Google Scholar]
436. Skinhoj E. Hemodynamic studies within the brain during migraine. Arch Neurol 29: 95–98, 1973. [Abstract] [Google Scholar]
437. Skinhoj E, Paulson OB. Regional blood flow in internal carotid distribution during migraine attack. Br Med J 3: 569–570, 1969. [Europe PMC free article] [Abstract] [Google Scholar]
438. Skyhoj Olsen T, Friberg L, Lassen NA. Ischemia may be the primary cause of the neurologic deficits in classic migraine. Arch Neurol 44: 156–161, 1987. [Abstract] [Google Scholar]
439. Smith MI, Read SJ, Chan WN, Thompson M, Hunter AJ, Upton N, Parsons AA. Repetitive cortical spreading depression in a gyrencephalic feline brain: inhibition by the novel benzoylamino-benzopyran SB-220453. Cephalalgia 20: 546–553, 2000. [Abstract] [Google Scholar]
440. Sokoloff L. Relationships among local functional activity, energy metabolism, and blood flow in the central nervous system. Federation Proc 40: 2311–2316, 1981. [Abstract] [Google Scholar]
441. Somjen GG. Mechanisms of spreading depression and hypoxic spreading depression-like depolarization. Physiol Rev 81: 1065–1096, 2001. [Abstract] [Google Scholar]
442. Sonn J, Mayevsky A. The effect of ethanol on metabolic, hemodynamic and electrical responses to cortical spreading depression. Brain Res 908: 174–186, 2001. [Abstract] [Google Scholar]
443. Sonn J, Mayevsky A. Effects of anesthesia on the responses to cortical spreading depression in the rat brain in vivo. Neurol Res 28: 206–219, 2006. [Abstract] [Google Scholar]
444. Sonn J, Mayevsky A. Effects of brain oxygenation on metabolic, hemodynamic, ionic and electrical responses to spreading depression in the rat. Brain Res 882: 212–216, 2000. [Abstract] [Google Scholar]
445. Sonn J, Mayevsky A. Responses to cortical spreading depression under oxygen deficiency. Open Neurol J 6: 6–17, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
446. Sorensen AG, Copen WA, Ostergaard L, Buonanno FS, Gonzalez RG, Rordorf G, Rosen BR, Schwamm LH, Weisskoff RM, Koroshetz WJ. Hyperacute stroke: simultaneous measurement of relative cerebral blood volume, relative cerebral blood flow, and mean tissue transit time. Radiology 210: 519–527, 1999. [Abstract] [Google Scholar]
447. Sorensen SG, Aguilar H, Reichert W, Goldstein M, Thoen D, Thomas H, Muhlestein JB. Transient neurological symptoms during contrast transcranial Doppler right-to-left shunt testing in patients with cryptogenic neurological disease. J Interv Cardiol 23: 284–290, 2010. [Abstract] [Google Scholar]
448. Sramka M, Brozek G, Bures J, Nadvornik P. Functional ablation by spreading depression: possible use in human stereotactic neurosurgery. Appl Neurophysiol 40: 48–61, 1977. [Abstract] [Google Scholar]
449. Stach R, Smialowski A, Dulska E. The effect of ouabain on the bioelectrical activity of extrapyramidal cortex and caudate nucleus in the rabbit brain. Neurosci Lett 5: 351–355, 1977. [Abstract] [Google Scholar]
450. Strong AJ, anderson PJ, Watts HR, Virley DJ, Lloyd A, Irving EA, Nagafuji T, Ninomiya M, Nakamura H, Dunn AK, Graf R. Peri-infarct depolarizations lead to loss of perfusion in ischaemic gyrencephalic cerebral cortex. Brain 130: 995–1008, 2007. [Abstract] [Google Scholar]
451. Strong AJ, Fabricius M, Boutelle MG, Hibbins SJ, Hopwood SE, Jones R, Parkin MC, Lauritzen M. Spreading and synchronous depressions of cortical activity in acutely injured human brain. Stroke 33: 2738–2743, 2002. [Abstract] [Google Scholar]
452. Strong AJ, Harland SP, Meldrum BS, Whittington DJ. The use of in vivo fluorescence image sequences to indicate the occurrence and propagation of transient focal depolarizations in cerebral ischemia. J Cereb Blood Flow Metab 16: 367–377, 1996. [Abstract] [Google Scholar]
453. Strong AJ, Hartings JA, Dreier JP. Cortical spreading depression: an adverse but treatable factor in intensive care? Curr Opin Crit Care 13: 126–133, 2007. [Abstract] [Google Scholar]
454. Strong AJ, Smith SE, Whittington DJ, Meldrum BS, Parsons AA, Krupinski J, Hunter AJ, Patel S, Robertson C. Factors influencing the frequency of fluorescence transients as markers of peri-infarct depolarizations in focal cerebral ischemia. Stroke 31: 214–222, 2000. [Abstract] [Google Scholar]
455. Sugaya E, Takato M, Noda Y. Neuronal and glial activity during spreading depression in cerebral cortex of cat. J Neurophysiol 38: 822–841, 1975. [Abstract] [Google Scholar]
456. Sukhotinsky I, Dilekoz E, Moskowitz MA, Ayata C. Hypoxia and hypotension transform the blood flow response to cortical spreading depression from hyperemia into hypoperfusion in the rat. J Cereb Blood Flow Metab 28: 1369–1376, 2008. [Abstract] [Google Scholar]
457. Sukhotinsky I, Yaseen MA, Sakadzic S, Ruvinskaya S, Sims JR, Boas DA, Moskowitz MA, Ayata C. Perfusion pressure-dependent recovery of cortical spreading depression is independent of tissue oxygenation over a wide physiologic range. J Cereb Blood Flow Metab 30: 1168–1177, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
458. Sullivan PG, Rabchevsky AG, Waldmeier PC, Springer JE. Mitochondrial permeability transition in CNS trauma: cause or effect of neuronal cell death? J Neurosci Res 79: 231–239, 2005. [Abstract] [Google Scholar]
459. Sun CW, Falck JR, Okamoto H, Harder DR, Roman RJ. Role of cGMP versus 20-HETE in the vasodilator response to nitric oxide in rat cerebral arteries. Am J Physiol Heart Circ Physiol 279: H339–H350, 2000. [Abstract] [Google Scholar]
460. Sun X, Li P, Luo W, Chen S, Feng N, Wang J, Luo Q. Investigating the effects of dimethylsulfoxide on hemodynamics during cortical spreading depression by combining laser speckle imaging with optical intrinsic signal imaging. Lasers Surg Med 42: 649–655, 2010. [Abstract] [Google Scholar]
461. Sun X, Wang Y, Chen S, Luo W, Li P, Luo Q. Simultaneous monitoring of intracellular pH changes and hemodynamic response during cortical spreading depression by fluorescence-corrected multimodal optical imaging. Neuroimage 57: 873–884, 2011. [Abstract] [Google Scholar]
462. Suzuki N, Hardebo JE, Kahrstrom J, Owman C. Selective electrical stimulation of postganglionic cerebrovascular parasympathetic nerve fibers originating from the sphenopalatine ganglion enhances cortical blood flow in the rat. J Cereb Blood Flow Metab 10: 383–391, 1990. [Abstract] [Google Scholar]
463. Sykova E, Nicholson C. Diffusion in brain extracellular space. Physiol Rev 88: 1277–1340, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
464. Sylvia AL, Rosenthal M. Effects of age on brain oxidative metabolism in vivo. Brain Res 165: 235–248, 1979. [Abstract] [Google Scholar]
465. Szerb JC. Glutamate release and spreading depression in the fascia dentata in response to microdialysis with high K+: role of glia. Brain Res 542: 259–265, 1991. [Abstract] [Google Scholar]
466. Takagaki M, Feuerstein D, Kumagai T, Gramer M, Yoshimine T, Graf R. Isoflurane suppresses cortical spreading depolarizations compared with propofol–implications for sedation of neurocritical care patients. Exp Neurol 252: 12–17, 2014. [Abstract] [Google Scholar]
467. Takano K, Latour LL, Formato JE, Carano RA, Helmer KG, Hasegawa Y, Sotak CH, Fisher M. The role of spreading depression in focal ischemia evaluated by diffusion mapping. Ann Neurol 39: 308–318, 1996. [Abstract] [Google Scholar]
468. Takano T, Tian GF, Peng W, Lou N, Libionka W, Han X, Nedergaard M. Astrocyte-mediated control of cerebral blood flow. Nat Neurosci 9: 260–267, 2006. [Abstract] [Google Scholar]
469. Takano T, Tian GF, Peng W, Lou N, Lovatt D, Hansen AJ, Kasischke KA, Nedergaard M. Cortical spreading depression causes and coincides with tissue hypoxia. Nat Neurosci 10: 754–762, 2007. [Abstract] [Google Scholar]
470. Takeda Y, Zhao L, Jacewicz M, Pulsinelli WA, Nowak TS Jr. Metabolic and perfusion responses to recurrent peri-infarct depolarization during focal ischemia in the Spontaneously Hypertensive Rat: dominant contribution of sporadic CBF decrements to infarct expansion. J Cereb Blood Flow Metab 31: 1863–1873, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
471. Tang YT, Mendez JM, Theriot JJ, Sawant PM, Lopez-Valdes HE, Ju YS, Brennan KC. Minimum conditions for the induction of cortical spreading depression in brain slices. J Neurophysiol 112: 2572–2579, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
472. Tao L, Masri D, Hrabetova S, Nicholson C. Light scattering in rat neocortical slices differs during spreading depression and ischemia. Brain Res 952: 290–300, 2002. [Abstract] [Google Scholar]
473. Tasaki I, Byrne PM. Demonstration of heat production associated with spreading depression in the amphibian retina. Biochem Biophys Res Commun 174: 293–297, 1991. [Abstract] [Google Scholar]
474. Teplov V, Shatillo A, Nippolainen E, Grohn O, Giniatullin R, Kamshilin AA. Fast vascular component of cortical spreading depression revealed in rats by blood pulsation imaging. J Biomed Opt 19: 046011, 2014. [Abstract] [Google Scholar]
475. Tettenborn D, Dycka J. Prevention and treatment of delayed ischemic dysfunction in patients with aneurysmal subarachnoid hemorrhage. Stroke 21: IV85–89, 1990. [Abstract] [Google Scholar]
476. Thrane AS, Takano T, Rangroo Thrane V, Wang F, Peng W, Ottersen OP, Nedergaard M, Nagelhus EA. In vivo NADH fluorescence imaging indicates effect of aquaporin-4 deletion on oxygen microdistribution in cortical spreading depression. J Cereb Blood Flow Metab 33: 996–999, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
477. Tobiasz C, Nicholson C. Tetrodotoxin resistant propagation and extracellular sodium changes during spreading depression in rat cerebellum. Brain Res 241: 329–333, 1982. [Abstract] [Google Scholar]
478. Tomida S, Wagner HG, Klatzo I, Nowak TS Jr. Effect of acute electrode placement on regional CBF in the gerbil: a comparison of blood flow measured by hydrogen clearance, [3H]nicotine, and [14C]iodoantipyrine techniques. J Cereb Blood Flow Metab 9: 79–86, 1989. [Abstract] [Google Scholar]
479. Tomita M, Schiszler I, Tomita Y, Tanahashi N, Takeda H, Osada T, Suzuki N. Initial oligemia with capillary flow stop followed by hyperemia during K+-induced cortical spreading depression in rats. J Cereb Blood Flow Metab 25: 742–747, 2005. [Abstract] [Google Scholar]
480. Tomita M, Tomita Y, Unekawa M, Toriumi H, Suzuki N. Oscillating neuro-capillary coupling during cortical spreading depression as observed by tracking of FITC-labeled RBCs in single capillaries. Neuroimage 56: 1001–1010, 2011. [Abstract] [Google Scholar]
481. Tomita Y, Tomita M, Schiszler I, Amano T, Tanahashi N, Kobari M, Takeda H, Ohtomo M, Fukuuchi Y. Repetitive concentric wave-ring spread of oligemia/hyperemia in the sensorimotor cortex accompanying K+-induced spreading depression in rats and cats. Neurosci Lett 322: 157–160, 2002. [Abstract] [Google Scholar]
482. Tong CK, Chesler M. Modulation of spreading depression by changes in extracellular pH. J Neurophysiol 84: 2449–2457, 2000. [Abstract] [Google Scholar]
483. Tozzi A, de Iure A, Di Filippo M, Costa C, Caproni S, Pisani A, Bonsi P, Picconi B, Cupini LM, Materazzi S, Geppetti P, Sarchielli P, Calabresi P. Critical role of calcitonin gene-related peptide receptors in cortical spreading depression. Proc Natl Acad Sci USA 109: 18985–18990, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
484. Tran CH, Vigmond EJ, Goldman D, Plane F, Welsh DG. Electrical communication in branching arterial networks. Am J Physiol Heart Circ Physiol 303: H680–H692, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
485. Trivedi B, Danforth WH. Effect of pH on the kinetics of frog muscle phosphofructokinase. J Biol Chem 241: 4110–4112, 1966. [Abstract] [Google Scholar]
486. Tsacopoulos M, Lehmenkuhler A. A double-barrelled Pt-microelectrode for simultaneous measurement of PO2 and bioelectrical activity in excitable tissues. Experientia 33: 1337–1338, 1977. [Abstract] [Google Scholar]
487. Tschirgi RD, Inanaga K, Taylor JL, Walker RM, Sonnenschein RR. Changes in cortical pH and blood flow accompanying spreading cortical depression and convulsion. Am J Physiol 190: 557–562, 1957. [Abstract] [Google Scholar]
488. Turner DA, Foster KA, Galeffi F, Somjen GG. Differences in O2 availability resolve the apparent discrepancies in metabolic intrinsic optical signals in vivo and in vitro. Trends Neurosci 30: 390–398, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
489. Ueda M, Watanabe N, Ushikubo Y, Tsuzuki T, Aoki K, Yamazaki Y, Samejima H. Changes in regional cortical temperature and cerebral blood flow after cortical spreading depression. Neurol Med Chir 37: 441–446, 1997. [Abstract] [Google Scholar]
490. Unekawa M, Tomita M, Tomita Y, Toriumi H, Suzuki N. Sustained decrease and remarkable increase in red blood cell velocity in intraparenchymal capillaries associated with potassium-induced cortical spreading depression. Microcirculation 19: 166–174, 2012. [Abstract] [Google Scholar]
491. Unekawa M, Tomita Y, Toriumi H, Masamoto K, Kanno I, Suzuki N. Potassium-induced cortical spreading depression bilaterally suppresses the electroencephalogram but only ipsilaterally affects red blood cell velocity in intraparenchymal capillaries. J Neurosci Res 91: 578–584, 2013. [Abstract] [Google Scholar]
492. Unekawa M, Tomita Y, Toriumi H, Osada T, Masamoto K, Kawaguchi H, Itoh Y, Kanno I, Suzuki N. Hyperperfusion counteracted by transient rapid vasoconstriction followed by long-lasting oligemia induced by cortical spreading depression in anesthetized mice. J Cereb Blood Flow Metab. In press. [Europe PMC free article] [Abstract] [Google Scholar]
493. Van den Maagdenberg AM, Pietrobon D, Pizzorusso T, Kaja S, Broos LA, Cesetti T, van de Ven RC, Tottene A, van der Kaa J, Plomp JJ, Frants RR, Ferrari MD. A Cacna1a knockin migraine mouse model with increased susceptibility to cortical spreading depression. Neuron 41: 701–710, 2004. [Abstract] [Google Scholar]
494. Van den Maagdenberg AM, Pizzorusso T, Kaja S, Terpolilli N, Shapovalova M, Hoebeek FE, Barrett CF, Gherardini L, van de Ven RC, Todorov B, Broos LA, Tottene A, Gao Z, Fodor M, De Zeeuw CI, Frants RR, Plesnila N, Plomp JJ, Pietrobon D, Ferrari MD. High cortical spreading depression susceptibility and migraine-associated symptoms in Ca(v)2.1 S218L mice. Ann Neurol 67: 85–98, 2010. [Abstract] [Google Scholar]
495. Van Harreveld A. Changes in the diameter of apical dendrites during spreading depression. Am J Physiol 192: 457–463, 1958. [Abstract] [Google Scholar]
496. Van Harreveld A. Compounds in brain extracts causing spreading depression of cerebral cortical activity and contraction of crustacean muscle. J Neurochem 3: 300–315, 1959. [Abstract] [Google Scholar]
497. Van Harreveld A, Khattab FI. Changes in cortical extracellular space during spreading depression investigated with the electron microscope. J Neurophysiol 30: 911–929, 1967. [Abstract] [Google Scholar]
498. Van Harreveld A, Ochs S. Electrical and vascular concomitants of spreading depression. Am J Physiol 189: 159–166, 1957. [Abstract] [Google Scholar]
499. Van Harreveld A, Schade JP. Chloride movements in cerebral cortex after circulatory arrest and during spreading depression. J Cell Comp Physiol 54: 65–84, 1959. [Abstract] [Google Scholar]
500. Van Harreveld A, Stamm JS. Vascular concomitants of spreading cortical depression. J Neurophysiol 15: 487–496, 1952. [Abstract] [Google Scholar]
501. Vercesi A, Martins-Ferreira H. Oxygen and glucose requirements in chick retinal spreading depression. An Acad Bras Cienc 55: 309–316, 1983. [Abstract] [Google Scholar]
502. Verhaegen MJ, Todd MM, Warner DS, James B, Weeks JB. The role of electrode size on the incidence of spreading depression and on cortical cerebral blood flow as measured by H2 clearance. J Cereb Blood Flow Metab 12: 230–237, 1992. [Abstract] [Google Scholar]
503. Viggiano A, Viggiano E, Valentino I, Monda M, Viggiano A, De Luca B. Cortical spreading depression affects reactive oxygen species production. Brain Res 1368: 11–18, 2011. [Abstract] [Google Scholar]
504. Viggiano E, Ferrara D, Izzo G, Viggiano A, Minucci S, Monda M, De Luca B. Cortical spreading depression induces the expression of iNOS, HIF-1alpha, and LDH-A. Neuroscience 153: 182–188, 2008. [Abstract] [Google Scholar]
505. Von Baumgarten L, Trabold R, Thal S, Back T, Plesnila N. Role of cortical spreading depressions for secondary brain damage after traumatic brain injury in mice. J Cereb Blood Flow Metab 28: 1353–1360, 2008. [Abstract] [Google Scholar]
506. Von Bornstädt D, Houben T, Seidel J, Zheng Y, Dilekoz E, Qin T, Sandow D, Kura S, Eikermann-Haerter K, Endres M, Boas DA, Moskowitz MA, Lo EH, Dreier JP, Woitzik J, Sakadžić S, Ayata C. Supply-demand mismatch transients in susceptible peri-infarct hot zones explain the origin of spreading injury depolarizations. Neuron. In press. [Europe PMC free article] [Abstract] [Google Scholar]
507. Vyskocil F, Kritz N, Bures J. Potassium-selective microelectrodes used for measuring the extracellular brain potassium during spreading depression and anoxic depolarization in rats. Brain Res 39: 255–259, 1972. [Abstract] [Google Scholar]
508. Wade JG, Amtorp O, Sorensen SC. No-flow state following cerebral ischemia. Role of increase in potassium concentration in brain interstitial fluid. Arch Neurol 32: 381–384, 1975. [Abstract] [Google Scholar]
509. Wahl M, Deetjen P, Thurau K, Ingvar DH, Lassen NA. Micropuncture evaluation of the importance of perivascular pH for the arteriolar diameter on the brain surface. Pflügers Arch 316: 152–163, 1970. [Abstract] [Google Scholar]
510. Wahl M, Kuschinsky W. The dilatatory action of adenosine on pial arteries of cats and its inhibition by theophylline. Pflügers Arch 362: 55–59, 1976. [Abstract] [Google Scholar]
511. Wahl M, Kuschinsky W, Bosse O, Olesen J, Lassen NA, Ingvar DH, Michaelis J, Thurau K. Effect of 1-norepinephrine on the diameter of pial arterioles and arteries in the cat. Circ Res 31: 248–256, 1972. [Abstract] [Google Scholar]
512. Wahl M, Lauritzen M, Schilling L. Change of cerebrovascular reactivity after cortical spreading depression in cats and rats. Brain Res 411: 72–80, 1987. [Abstract] [Google Scholar]
513. Wahl M, Schilling L, Parsons AA, Kaumann A. Involvement of calcitonin gene-related peptide (CGRP) and nitric oxide (NO) in the pial artery dilatation elicited by cortical spreading depression. Brain Res 637: 204–210, 1994. [Abstract] [Google Scholar]
514. Waldmeier PC, Feldtrauer JJ, Qian T, Lemasters JJ. Inhibition of the mitochondrial permeability transition by the nonimmunosuppressive cyclosporin derivative NIM811. Mol Pharmacol 62: 22–29, 2002. [Abstract] [Google Scholar]
515. Wang Z, Luo W, Li P, Qiu J, Luo Q. Acute hyperglycemia compromises cerebral blood flow following cortical spreading depression in rats monitored by laser speckle imaging. J Biomed Opt 13: 064023, 2008. [Abstract] [Google Scholar]
516. Weber B, Burger C, Wyss MT, von Schulthess GK, Scheffold F, Buck A. Optical imaging of the spatiotemporal dynamics of cerebral blood flow and oxidative metabolism in the rat barrel cortex. Eur J Neurosci 20: 2664–2670, 2004. [Abstract] [Google Scholar]
517. Widstrom RL, Norris AW, Spector AA. Binding of cytochrome P450 monooxygenase and lipoxygenase pathway products by heart fatty acid-binding protein. Biochemistry 40: 1070–1076, 2001. [Abstract] [Google Scholar]
518. Windmuller O, Lindauer U, Foddis M, Einhaupl KM, Dirnagl U, Heinemann U, Dreier JP. Ion changes in spreading ischaemia induce rat middle cerebral artery constriction in the absence of NO. Brain 128: 2042–2051, 2005. [Abstract] [Google Scholar]
519. Woitzik J, Hecht N, Pinczolits A, Sandow N, Major S, Winkler MK, Weber-Carstens S, Dohmen C, Graf R, Strong AJ, Dreier JP, Vajkoczy P., CS Group. Propagation of cortical spreading depolarization in the human cortex after malignant stroke. Neurology 80: 1095–1102, 2013. [Abstract] [Google Scholar]
520. Wolf T, Lindauer U, Obrig H, Dreier J, Back T, Villringer A, Dirnagl U. Systemic nitric oxide synthase inhibition does not affect brain oxygenation during cortical spreading depression in rats: a noninvasive near-infrared spectroscopy and laser-Doppler flowmetry study. J Cereb Blood Flow Metab 16: 1100–1107, 1996. [Abstract] [Google Scholar]
521. Wolf T, Lindauer U, Obrig H, Villringer A, Dirnagl U. Extra- and intracellular oxygen supply during cortical spreading depression in the rat. Adv Exp Med Biol 388: 299–304, 1996. [Abstract] [Google Scholar]
522. Wolf T, Lindauer U, Reuter U, Back T, Villringer A, Einhaupl K, Dirnagl U. Noninvasive near infrared spectroscopy monitoring of regional cerebral blood oxygenation changes during peri-infarct depolarizations in focal cerebral ischemia in the rat. J Cereb Blood Flow Metab 17: 950–954, 1997. [Abstract] [Google Scholar]
523. Wolf T, Lindauer U, Villringer A, Dirnagl U. Excessive oxygen or glucose supply does not alter the blood flow response to somatosensory stimulation or spreading depression in rats. Brain Res 761: 290–299, 1997. [Abstract] [Google Scholar]
524. Woods RP, Iacoboni M, Mazziotta JC. Brief report: bilateral spreading cerebral hypoperfusion during spontaneous migraine headache. N Engl J Med 331: 1689–1692, 1994. [Abstract] [Google Scholar]
525. Yang G, Chen G, Ebner TJ, Iadecola C. Nitric oxide is the predominant mediator of cerebellar hyperemia during somatosensory activation in rats. Am J Physiol Regul Integr Comp Physiol 277: R1760–R1770, 1999. [Abstract] [Google Scholar]
526. Yashiro Y, Duling BR. Integrated Ca2+ signaling between smooth muscle and endothelium of resistance vessels. Circ Res 87: 1048–1054, 2000. [Abstract] [Google Scholar]
527. Yemisci M, Gursoy-Ozdemir Y, Vural A, Can A, Topalkara K, Dalkara T. Pericyte contraction induced by oxidative-nitrative stress impairs capillary reflow despite successful opening of an occluded cerebral artery. Nat Med 15: 1031–1037, 2009. [Abstract] [Google Scholar]
528. Yin C, Zhou F, Wang Y, Luo W, Luo Q, Li P. Simultaneous detection of hemodynamics, mitochondrial metabolism and light scattering changes during cortical spreading depression in rats based on multi-spectral optical imaging. Neuroimage 76: 70–80, 2013. [Abstract] [Google Scholar]
529. Yisarakun W, Supornsilpchai W, Chantong C, Srikiatkhachorn A, Maneesri-le Grand S. Chronic paracetamol treatment increases alterations in cerebral vessels in cortical spreading depression model. Microvasc Res 94: 36–46, 2014. [Abstract] [Google Scholar]
530. Yokota C, Kuge Y, Hasegawa Y, Tagaya M, Abumiya T, Ejima N, Tamaki N, Yamaguchi T, Minematsu K. Unique profile of spreading depression in a primate model. J Cereb Blood Flow Metab 22: 835–842, 2002. [Abstract] [Google Scholar]
531. Young W. Spreading depression in elasmobranch cerebellum. Brain Res 199: 113–126, 1980. [Abstract] [Google Scholar]
532. Yuzawa I, Sakadzic S, Srinivasan VJ, Shin HK, Eikermann-Haerter K, Boas DA, Ayata C. Cortical spreading depression impairs oxygen delivery and metabolism in mice. J Cereb Blood Flow Metab 32: 376–386, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
533. Zhang F, Sprague SM, Farrokhi F, Henry MN, Son MG, Vollmer DG. Reversal of attenuation of cerebrovascular reactivity to hypercapnia by a nitric oxide donor after controlled cortical impact in a rat model of traumatic brain injury. J Neurosurg 97: 963–969, 2002. [Abstract] [Google Scholar]
534. Zhang ZG, Chopp M, Maynard KI, Moskowitz MA. Cerebral blood flow changes during cortical spreading depression are not altered by inhibition of nitric oxide synthesis. J Cereb Blood Flow Metab 14: 939–943, 1994. [Abstract] [Google Scholar]
535. Zhou C, Yu G, Furuya D, Greenberg J, Yodh A, Durduran T. Diffuse optical correlation tomography of cerebral blood flow during cortical spreading depression in rat brain. Optics Express 14: 1125–1144, 2006. [Abstract] [Google Scholar]
536. Zhou N, Gordon GR, Feighan D, MacVicar BA. Transient swelling, acidification, and mitochondrial depolarization occurs in neurons but not astrocytes during spreading depression. Cereb Cortex 20: 2614–2624, 2010. [Abstract] [Google Scholar]
537. Zonta M, Angulo MC, Gobbo S, Rosengarten B, Hossmann KA, Pozzan T, Carmignoto G. Neuron-to-astrocyte signaling is central to the dynamic control of brain microcirculation. Nat Neurosci 6: 43–50, 2003. [Abstract] [Google Scholar]

Articles from Physiological Reviews are provided here courtesy of American Physiological Society

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/4233489
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/4233489

Article citations


Go to all (259) article citations

Other citations

Funding 


Funders who supported this work.

HHS | NIH | National Institute of Neurological Disorders and Stroke (1)

HHS | NIH | National Institute of Neurological Disorders and Stroke (NINDS) (1)

NINDS NIH HHS (5)