Download Full PDF - American Journal of Physiology

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Neuropharmacology wikipedia , lookup

Brain damage wikipedia , lookup

Neurodegeneration wikipedia , lookup

Cerebral palsy wikipedia , lookup

Cortical stimulation mapping wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

History of neuroimaging wikipedia , lookup

Transcranial Doppler wikipedia , lookup

Transcript
J Appl Physiol 100: 328 –335, 2006;
doi:10.1152/japplphysiol.00966.2005.
Invited Review
HIGHLIGHTED TOPIC
Regulation of the Cerebral Circulation
Neurovascular coupling in the normal brain and in hypertension,
stroke, and Alzheimer disease
Helene Girouard and Costantino Iadecola
Division of Neurobiology, Department of Neurology and Neuroscience, Weill Medical College
of Cornell University, New York, New York
cerebral blood flow; astrocytes; NADPH oxidase; free radicals
A LARGE BODY OF EVIDENCE INDICATES that neural activity is
closely related to cerebral blood flow (CBF). The close spatial
and temporal relationship between neural activity and CBF,
termed neurovascular coupling, is at the basis of modern
neuroimaging techniques that utilize the cerebrovascular
changes induced by activation to map regional changes in
function in the behaving human brain. However, in several
brain pathologies, the interaction between neural activity and
cerebral blood vessels is disrupted, and the resulting homeostatic unbalance may contribute to brain dysfunction. This
review provides a brief summary of neurovascular coupling in
the normal state and in diseases including hypertension, ischemic stroke, and Alzheimer disease (AD).
NEUROVASCULAR COUPLING IN THE NORMAL STATE
Cerebral blood vessels have many unique structural and
functional characteristics that differentiate them from vessels
in other organs. Perhaps the most distinctive feature of cerebral
blood vessels is their close interaction with neurons and glia. A
growing body of evidence indicates that neurons, glia (astrocytes, microglia, oligodendrocytes), and vascular cells (endoAddress for reprint requests and other correspondence: C. Iadecola, Division
of Neurobiology, 411 East 69th St., Rm. KB410, New York, NY 10021
(e-mail: [email protected]).
328
thelium, smooth muscle cells or pericytes, adventitial cells) are
closely related developmentally, structurally, and functionally.
The term “neurovascular unit” was introduced to highlight the
intimate functional relationships between these cells and their
coordinated pattern of reaction to injury (see Ref. 31 for
references). The increase in CBF produced by brain activity, or
functional hyperemia, is an example of the close interaction
between neurons, glia, and vascular cells. In the next sections,
we will describe the anatomical and functional bases of neurovascular coupling in the normal brain.
Cerebral Blood Vessels
Large cerebral arteries arising from the circle of Willis
branch out into smaller pial arteries and arterioles that travel on
the surface of the brain across the subarachnoid space. Pial
arteries give rise to penetrating arteries and arterioles that enter
into the substance of the brain. These arteries consist of an
endothelial cell layer, a smooth muscle cell layer, and an outer
layer, termed adventitia, containing collagen, fibroblasts, and
perivascular nerves (68) (Fig. 1). Penetrating vessels are separated from the brain by the Virchow-Robin space, which
contains cerebrospinal fluid. On the outer side of the VirchowRobin space, astrocytes give rise to the glia limitans membrane. As the arterioles penetrate deeper into the brain, the
Virchow-Robin space disappears and the vascular basement
8750-7587/06 $8.00 Copyright © 2006 the American Physiological Society
http://www. jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
Girouard, Helene, and Costantino Iadecola. Neurovascular coupling in the
normal brain and in hypertension, stroke, and Alzheimer disease. J Appl Physiol 100:
328 –335, 2006; doi: 10.1152/japplphysiol.00966.2005.—The brain is critically dependent on a continuous supply of blood to function. Therefore, the cerebral
vasculature is endowed with neurovascular control mechanisms that assure that the
blood supply of the brain is commensurate to the energy needs of its cellular
constituents. The regulation of cerebral blood flow (CBF) during brain activity
involves the coordinated interaction of neurons, glia, and vascular cells. Thus,
whereas neurons and glia generate the signals initiating the vasodilation, endothelial cells, pericytes, and smooth muscle cells act in concert to transduce these
signals into carefully orchestrated vascular changes that lead to CBF increases
focused to the activated area and temporally linked to the period of activation.
Neurovascular coupling is disrupted in pathological conditions, such as hypertension, Alzheimer disease, and ischemic stroke. Consequently, CBF is no longer
matched to the metabolic requirements of the tissue. This cerebrovascular dysregulation is mediated in large part by the deleterious action of reactive oxygen species
on cerebral blood vessels. A major source of cerebral vascular radicals in models
of hypertension and Alzheimer disease is the enzyme NADPH oxidase. These
findings, collectively, highlight the importance of neurovascular coupling to the
health of the normal brain and suggest a therapeutic target for improving brain
function in pathologies associated with cerebrovascular dysfunction.
Invited Review
329
NEUROVASCULAR COUPLING
membrane enters into direct contact with the astrocytic end
feet. Arterioles become progressively smaller, lose the smooth
muscle cell layer, and become cerebral capillaries. The density
of brain capillaries within the brain is regionally heterogeneous
and varies according to regional blood flow and regional
metabolic demands (81). Capillaries consist of endothelial
cells, pericytes, and the capillary basal lamina on which astrocytic feet are attached (68) (Fig. 1). Brain endothelial cells are
unique in that they are not fenestrated and are sealed by tight
junctions, features that underlie the blood-brain barrier. Endothelial cells play an important role in the regulation of vascular
tone by releasing potent vasoactive factors, such as nitric oxide
(NO), free radicals, prostacyclin, endothelium-derived hyperpolarizing factor, and endothelin (19). Pericytes have contractile properties and may modulate capillary diameter (31).
Perivascular astrocytes surround most of the capillary abluminal surface with their end feet. There is close association
between cerebral arteries, arterioles, and capillaries with nerves
originating from central and peripheral sources (Fig. 1). These
relationships are addressed elsewhere in this series (see
Ref. 25a).
muscle cells (20, 54), leading to their hyperpolarization and
subsequent relaxation. During sustained activation, ATP depletion could lead to opening of ATP-sensitive K⫹ channels
(KATP) on vessels. Therefore, KATP channels have been implicated in the mechanisms of neurovascular coupling (54). Furthermore, KATP could also participate in neurovascular coupling by mediating the vasodilation produced by agents that
increase cAMP, such as adenosine or prostacyclin (20). The
vasodilatatory effect of increased concentrations of H⫹ is also
mediated, at least in part, by the opening of K⫹ channels (20).
It has also been suggested that activity-induced reductions in
extracellular Ca2⫹ may produce vasodilation (28).
Vasoactive factors related to energy metabolism. The brain
has little energy reserve and requires a continuous supply of
glucose and O2 through CBF. A sudden increase in the demand
Table 1. Factors implicated in neurovascular coupling
Agent
Vasoactive ions
K⫹
H⫹
Ca2⫹
Metabolic factors
Lactate, CO2
Hypoxia
Adenosine
Vasoactive neurotransmitters
Dopamine
GABA
Acetylcholine
Vasoactive intestinal peptide
Other
NO
COX-2 products
P450 products
CO
Mechanisms of Neurovascular Coupling: The Quest for
the Mediators
The mechanisms underlying neurovascular coupling have
been the subject of enquiry for more than a century (31), and
numerous vasoactive factors have been implicated in neurovascular coupling (Table 1). These include ions, metabolic
by-products, vasoactive neurotransmitters, and vasoactive factors released in response to neurotransmitters.
Vasoactive ions. K⫹ and H⫹ are generated by the extracellular ionic currents induced by action potentials and synaptic
transmission. Elevations in extracellular K⫹ up to 8 –10 mM
cause dilation of arterioles both in vitro and in vivo (45, 54).
This effect is mediated by the opening of K⫹ channels, mainly
of the inward rectifier type, on the membrane of arterial smooth
J Appl Physiol • VOL
References
45
45
28
71, 73
73
72
43
22
74
83
61
55
67
51
NO, nitric oxide; COX-2, cyclooxygenase-2; CO, carbon monoxide.
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
Fig. 1. Relationship of cerebrovascular cells with neurons, glia, and perivascular nerves. Pial arteries and arterioles are innervated by nerve fibers arising from
cranial autonomic ganglia. Smaller cerebral arterioles (⬍100 ␮m) come in contact with nerve terminals arising from local interneurons and from central pathways
originating from distant sites in the brain stem or basal forebrain. These neurovascular associations often terminate on astrocytic end feet lining the abluminal
vascular surface. Pericytes, contractile cells embedded in the capillary wall, are closely associated with astrocytic end feet and endothelial cells. The term
“neurovascular unit” has been coined to define the close structural and functional relationships between brain cells and vascular cells. In diseases states, such
as ischemic stroke, Alzheimer disease, and hypertension, the function of the neurovascular unit is profoundly disrupted resulting in alterations in cerebrovascular
reactivity that compromise brain function.
Invited Review
330
NEUROVASCULAR COUPLING
J Appl Physiol • VOL
33, 62). Because glutamate is not vasoactive in isolated cerebral arteries (18), the vasodilation is mediated by vasoactive
factors whose synthesis is triggered by the changes in intracellular Ca2⫹ associated with glutamate receptor activation. The
increase in Ca2⫹ activates Ca2⫹-dependent enzymes that produce potent vasodilators. One such enzyme is the neuronal
isoform of NO synthase (nNOS), which produces the vasodilator NO. The vasodilation produced by topical application of
NMDA or glutamate is reduced by nNOS inhibitors (18, 85).
There is ample evidence that NO contributes to the increase in
CBF produced by functional activity. Thus the increase in CBF
in the somatosensory cortex induced by sensory stimulation is
associated with NO release and is attenuated by nNOS inhibitors (5, 48, 61). However, in cerebral cortex, the effect of NO
synthase inhibition on functional hyperemia can be reversed by
application of exogenous NO. This finding suggests that the
presence of NO is required for the vasodilation, but that NO is
not the ultimate mediator of smooth muscle relaxation (48).
NO is also involved in the increase in CBF induced by
activation of the cerebellar cortex (1, 33). However, at variance
with the cerebral cortex, in cerebellum the attenuation of the
CBF response cannot be reversed by NO donors and is observed also in nNOS null mice (86, 87). These findings suggest
that, in cerebellum, NO plays an obligatory role in the mechanisms of the vasodilation.
The increase in intracellular Ca2⫹ evoked by glutamate also
activates phospholipase A2, leading to production of arachidonic acid. Arachidonic acid is then metabolized by the cyclooxygenase (COX) pathway, producing vasodilatatory prostaglandins (25). Although there are several isoforms of COX
(COX-1, -2, and -3) (25), COX-2 is the main isoform involved
in functional hyperemia. COX-2 is present in axon terminals
and dendritic processes separated from penetrating arterioles
and capillaries by glial processes (80). Functionally, the CBF
increase evoked by somatosensory stimulation is attenuated by
COX-2 inhibitors or in COX-2 null mice, whereas COX-1 does
not participate in the response (55, 57). The COX-2 metabolites responsible for the vasodilation are likely to involve
vasodilatatory prostaglandins. Other arachidonic acid products
that are involved in functional hyperemia include metabolites
of the p450 pathway, such as epoxyeicosatrienoic acids (67).
Furthermore, carbon monoxide has also been proposed to
contribute to functional hyperemia (51). Epoxyeicosatrienoic
acids and carbon monoxide are addressed elsewhere in this
series (see Leffler CW, Parfenova H, Jaggar JH, and Wang R,
unpublished review and Koehler RC, Gebremedhin D, and
Harder DH, unpublished review).
Neurovascular Coupling: Role of Astrocytes
Astrocytes are uniquely positioned to contribute to the increase in CBF produced by activation (see Koehler RC, Gebremedhin D, and Harder DH, unpublished review). Paulson
and Newman (65) provided theoretical evidence that astrocytes
could participate in functional hyperemia by shunting K⫹ ions,
which are vasoactive, from the synapses to the astrocytic end
feet surrounding blood vessels. However, this theory remains
to be proven experimentally. More direct evidence for a role of
astrocytes has been provided by Zonta et al. (90) in brain slices.
These investigators showed that glutamate released by neural
activity activates metabotropic glutamate receptors in cortical
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
for energy during synaptic activity could result in a relative
lack of O2 and glucose, which may be a factor in triggering the
hemodynamic response (3). However, the reduction in brain O2
concentration at the site of activation is small and transient and
cannot account for sustained increases in flow (2). Furthermore, the CBF response to activation is not altered by hypoglycemia or hypoxia, suggesting that lack of glucose or O2 is
not the primary factor triggering vasodilation (3). On the other
hand, adenosine, a potent vasodilator produced during ATP
catabolism, is involved in neurovascular coupling in cerebellum (47) and cerebral cortex (41). Lactate produced during
brain activation could also be an important mediator of functional hyperemia by increasing H⫹ concentration and producing vasodilation (3). Pellerin and colleagues (66) hypothesized
that astrocytes metabolize glucose through glycolysis, leading
to lactate production. Lactate, in turn, is taken up by neurons
and used as fuel for ATP synthesis. Recent data in hippocampal slices support this theory. Using two-photon confocal
microscopy of NADH fluorescence, Kasischke et al. (38) have
investigated the spatial and temporal characteristics of energy
metabolism in neurons and astrocytes during activation. They
provided evidence of an early NADH decrease in neurons,
reflecting oxidative metabolism, followed by a NADH increase
in astrocytes, reflecting glycolytic metabolism (38). These
observations are consistent with the hypothesis that neuronal
oxidative metabolism precedes glial glycolysis and fit well
with the reduction in tissue O2 (“dip”) observed at the onset of
activation with optical imaging, functional MRI, and O2 electrodes (2). However, the data of Kasischke et al. do not shed
light on the role of lactate in the vasodilation produced by
neural activation. Rather, the lactate rise is small and transient,
and it cannot account in full for the increase in flow produced
by neural activity (3).
Central pathways, interneurons, and vasoactive neurotransmitters. It has long been proposed that vasoactive neurotransmitters released during neural activity contribute to the vasodilation (Table 1). These neurotransmitters could be released
from neurovascular projections that terminate close to blood
vessels originating from local interneurons or from distant
nuclei and modulate CBF (Fig. 1). However, the contribution
of the vascular innervation to functional hyperemia has not
been firmly established. The role of interneurons has recently
been addressed by Cauli et al. (8), who, using forebrain slices,
have been able to show that activation of interneurons with
neurovascular contacts produces vasodilation. Furthermore, in
cerebellum the CBF response evoked by somatosensory activation is decreased in cyclin D2-null mice, which lack stellate
interneurons in the cerebellar molecular layer (84). These data
implicate specific classes of interneurons in the regulation of
the cerebral microcirculation during neural activity (see also
Ref. 25a).
Other vasoactive factors released by neural activity. Vasoactive factors can also be generated by the intracellular signaling induced by activation of neurotransmitter receptors. For
example, activation of glutamate receptors produces vasodilation and increases blood flow. In neocortex and hippocampus, exogenous glutamate or N-methyl-D-aspartate
(NMDA) dilates pial arterioles and/or cerebral microvessels
(18, 49). Functional hyperemia in cerebral and cerebellar
cortex is inhibited by DL-␣-amino-3-hydroxy-5-methylisoxazolepropionic acid (CAMPA) and/or NMDA receptor blockers (1,
Invited Review
NEUROVASCULAR COUPLING
Neurovascular Coupling: Local vs. Remote Vasodilation
In brain, flow is controlled by pial arteries, which are the
major site of vascular resistance (27). Consequently, to optimize perfusion, the dilation of intracerebral vessels at the site
of activation must be associated with dilation of upstream pial
arteries (32). Additional vascular adjustments are needed to
increase flow in the activated area, but not in inactive regions
vascularized by other branches of the same pial artery. The
mechanisms responsible for this complex and coordinated
chain of events have not been completely elucidated, but
several factors have emerged. The dilation initiated by active
neurons may be propagated retrogradely to pial arterioles
through an intrinsic mechanism such as the retrograde vasodilation of Duling and Berne (75). Indeed, retrograde vasodilation in response to ATP has been observed in isolated rat
cerebral arterioles (12), whereas upstream vasodilation has
been demonstrated in cerebellar cortex arterioles during activation of the parallel fibers (34). In systemic vessels, the
cellular mechanisms by which the vasodilation is transmitted
upstream involves intercellular conduction of signals between
endothelial cells and/or vascular smooth muscle cells via gap
junctions (75). The increase in shear stress on the endothelium
of the feeding vessels could contribute to propagate the response further upstream (flow-mediated vasodilation) (32), but
there is limited evidence supporting this possibility in the
neocortical microcirculation. Less clear are the mechanisms by
which the vasodilation is restricted only to the arterial branches
supplying the activated areas. One possibility is that the upstream dilatation leads to increased transmural pressure in
nondilated downstream branches supplying nonactivated areas
inducing a myogenic response that constricts the vessels and
maintains CBF constant. It is unlikely that release of vasodiJ Appl Physiol • VOL
lator substances from the glia limitans adjacent to upstream
pial arteries plays a role in remote vasodilation because flushing the brain surface with artificial cerebrospinal fluid does not
attenuate the pial vasodilation induced by activation (53).
Irrespective of the cellular mechanisms of the retrograde propagation of the vasodilatatory signals, vascular cells play a key
role in the expression of functional hyperemia by coordinating
a complex hemodynamic response that, in the end, results in
focused and timely increase in CBF to the activated area.
Mechanisms of Neurovascular Coupling: an Integrated View
For many decades, investigators believed that functional
hyperemia was the result of the action of a single vasoactive
agent reaching local blood vessels by simple diffusion and
producing the vasodilation (32). However, as discussed in the
previous section, this view is no longer tenable. Active neurons
and glia release a multitude of vasoactive factors that act in
concert to increase CBF (31). Cerebral endothelial cells, pericytes, and smooth muscle cells are the target of these signals
and transduce them into coordinated vascular adjustments that
ultimately lead to an increase in CBF. Therefore, the increase
in flow evoked by brain activity is mediated by the concerted
action of multiple mediators that originate from different cells
and act at different levels of the cerebral vasculature.
NEUROVASCULAR COUPLING IN DISEASE
The relationship between neural activity and CBF is altered
in several pathologies. These alterations perturb the delivery of
substrates to active brain cells and impair the removal of
potentially deleterious by-products of cerebral metabolism.
The ensuing disruption of the cerebral microenvironment is
likely to contribute to brain dysfunction. In this section, we will
focus on the cerebrovascular dysfunction that occurs in hypertension, AD, and ischemic stroke (Table 2).
Hypertension
Hypertension exerts deleterious actions on the brain and its
circulation. Hypertension alters the structure of cerebral blood
vessels by producing vascular hypertrophy and remodeling and
by promoting atherosclerosis in large cerebral arteries and
lipohyalinosis in penetrating arterioles (11, 17). These structural alterations facilitate vascular occlusions and compromise
cerebral perfusion. In addition, hypertension impairs the function of cerebral blood vessels (17, 27). Hypertension impairs
endothelium-dependent relaxation (19) and alters cerebrovascular autoregulation (27), defined as the ability of the cerebral
circulation to maintain relatively constant CBF in the face of
changes in arterial pressure within a certain range. Recent
evidence suggests that hypertension also alters neurovascular
coupling. Administration of ANG II to mice increases arterial
pressure (20 –30 mmHg) and attenuates the increase in somatosensory cortex CBF produced by whisker stimulation (⫺65%),
without reducing resting CBF (40). The effects of ANG II on
neurovascular coupling are blocked by losartan, indicating that
they are mediated by AT1 receptors (40). The attenuation in
functional hyperemia is observed even if the increase in arterial
pressure is prevented by removal of a small amount of arterial
blood, or if the pressor effect of ANG II is avoided by applying
the peptide directly to the cerebral cortex (40). Furthermore,
elevation of arterial pressure by phenylephrine administration
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
astrocytes, leading to increases in intracellular Ca2⫹, COX
activation, and local vasodilation (90). A critical role for Ca2⫹
fluxes from neurons to astrocytes and arterioles is also supported by another study in brain slices (23). Arterioles at rest
exhibit periodic contractions and relaxations, called vasomotion, a phenomenon resulting from Ca2⫹ oscillations in smooth
muscle cells (23). It was found that electrical stimulation of
brain slices induces Ca2⫹ waves in astrocytes, which then
propagate to arterioles and inhibit vasomotion (23). The inhibition of vasomotion would allow the arteriole to be more
relaxed and, presumably, increase flow. In contrast, Mulligan
and MacVicar (52) demonstrated that release of caged Ca2⫹ in
hippocampal astrocytes produces vasoconstriction, a response
mediated by the cytochrome p450 metabolites 20-HETEs. The
discrepancy between the studies of Zonta et al. and Mulligan
and MacVicar is perhaps due to differences in the experimental
conditions. For example, Zonta et al. preconstricted the vessels
with a NO synthase inhibitor. A major limitation of studies in
brain slices is the lack of cerebral perfusion. Changes in vessel
diameter under these conditions are difficult to assess because
the vessels lack the intrinsic smooth muscle tone provided by
intravascular pressure (myogenic tone) and are not subjected to
shear stress, which has profound effects on endothelial cell
function (44). Nevertheless, these studies are noteworthy because they provide evidence that activity-induced Ca2⫹ transients in astrocytes are linked to changes in cerebrovascular
tone.
331
Invited Review
332
NEUROVASCULAR COUPLING
Table 2. Alterations in functional hyperemia in Alzheimer disease, cerebrovascular diseases,
and hypertension: selected human studies
Pathology
Method for CBF
Detection
Activation Task
Effect on Cortical Hemodynamic Change
Reference
Middle temporal
Parietal, frontal
Dorsolateral
Prefrontal
Mid and posterior inferior temporal
50
29
82
77
4
42
69
Alzheimer disease and APOE␧ genotype
PET
PET and NIRS
Xenon-133
fMRI
fMRI
2
2
2
2
2
Alzheimer disease
Alzheimer disease
Alzheimer disease
APOEε4 carriers
APOEε4 carriers
Visual stimulus
Verbal task
Verbal task
Visual naming and letter task
Memory tasks
Ischemic stroke with full recovery
Lacunar stroke
Hand grip
Finger tapping
fMRI
fMRI
Extra- and intracranial stenosis
Hand grip
fMRI
Common carotid or vertebral
artery occlusion
Internal carotid or middle cerebral
artery occlusion
Essential hypertension
Finger tapping
fMRI
2 Left hippocampus
2 Somatosensory cortex, supplementary
motor area
2 Somatosensory cortex, supplementary
motor area
2 Supplementary motor area
Bimanual motor tasks
PET
No change in somatosensory cortex
36
Memory task
PET
2 Posterior parietal
37
Cerebrovascular diseases
26
70
does not reproduce the effects of ANG II on functional hyperemia. Although these observations suggest that the cerebrovascular effect of short-term administration of ANG II is independent of the increase in arterial pressure, they do not rule out
that in chronic models of hypertension, mediated by either
ANG II or other factors, the elevation of arterial pressure plays
a role in the cerebrovascular pathology. There is evidence that
hypertension alters functional hyperemia also in humans (Table 2). The increase in CBF in posterior parietal and thalamic
areas produced by cognitive tasks is reduced in patients with
chronic untreated hypertension relative to normotensive individuals (37). The attenuated CBF response was associated with
a lower cognitive performance (37). Reduction in resting CBF
in hypertensive subjects have also been reported (24), but not
by all studies (37, 46). Although confounding effects of coexisting brain pathologies cannot be ruled out, these findings
provide evidence that hypertension impairs neurovascular
function also in humans.
Alzheimer Disease
Alzheimer disease (AD) is the most common form of dementia and is characterized by deposition of amyloid ␤-peptide
(A␤) in the neuropil (neuritic plaques) and blood vessels
(amyloid angiopathy), and by accumulation of hyperphosphorylated neurofilament in neurons (neurofibrillary tangles) (76).
Cerebrovascular structure is profoundly altered in AD (21).
Cerebral microvessels are reduced in number, endothelial cells
are flattened, and smooth muscle cells undergo degeneration
(21). Cerebrovascular function is also altered in AD (31).
Resting CBF is reduced and the increase in CBF produced by
activation is attenuated (Table 2). The cerebrovascular dysfunction often precedes the onset of cognitive impairment,
suggesting a role in the mechanisms of the dementia (31).
Mouse models of AD, in which mutated amyloid precursor
protein (APP) is overexpressed to increase A␤ levels, have a
profound dysregulation of the cerebral circulation (31).
Whereas endothelium-dependent responses are attenuated, reJ Appl Physiol • VOL
sponses to vasoconstrictors are exaggerated (35, 59). Functional hyperemia is impaired and cerebrovascular autoregulation is nearly abolished (58, 60). These cerebrovascular effects
are present in the absence of plaques or vascular amyloid (58,
63). The cerebrovascular dysfunction observed in APP mice
can be reproduced in normal mice by topical superfusion of
A␤1– 40 on the neocortex (56, 63). Furthermore, alterations in
vascular reactivity can also be observed in isolated vessels of
normal mice exposed to A␤1– 40 (9, 59). In APP mice, the
deleterious vascular effects of A␤ worsen cerebral ischemia
and enhance the ensuing brain damage (88).
How do these cerebrovascular alterations contribute to the
brain dysfunction? The reduced cerebral perfusion can promote
ischemic lesions, which act synergistically with A␤ to exacerbate the dementia (31). Furthermore, insufficient CBF may
alter A␤ trafficking across the blood-brain barrier (89). Therefore, reduced CBF may slow down A␤ clearance and promote
its accumulation in the brain. In addition, the CBF reduction
may attenuate cerebral protein synthesis, which is essential for
normal cognition (31). Thus the structural and functional
alterations of the microvasculature in AD could contribute to
the mechanisms of the brain dysfunction underlying the dementia.
Ischemic Stroke
Focal or global cerebral ischemia exerts profound effects on
the normal regulation of the cerebral circulation (30). In focal
cerebral ischemia, the flow reduction resulting from the arterial
occlusion is greatest in the center of the ischemic territory
(ischemic core) and less pronounced at the periphery (ischemic
penumbra) (30). In global ischemia, the perfusion of the entire
brain is interrupted, usually because of cardiac arrest. In both
focal and global ischemia, when perfusion is reestablished
there is a transient increase in flow (postischemic hyperemia)
followed by a period of reduced flow (postischemic hypoperfusion). After ischemia, the cerebral circulation is in a state of
vasoparalysis (30). After ischemic stroke in patients, the reac-
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
CBF, cerebral blood flow; PET, positron emission tomography; NIRS: near-infrared spectroscopy; fMRI, functional magnetic resonance imaging; 2,
attenuation of the CBF increase.
Invited Review
NEUROVASCULAR COUPLING
333
tivity of the cerebral circulation to vasomotor stimuli is altered,
autoregulation is impaired, and the increase in CBF produced
by functional activation is decreased (30). Studies in which
indexes of neural activity were measured have suggested that
the reduction in the CBF response to activation is secondary to
a reduction of the neural activity driving the hemodynamic
response (6). Thus, in a rat global ischemia model, the reduction in the CBF response to cortical electrical stimulation is
associated with a reduction in the amplitude of the oxidative
response (oxidation of cytochrome a) evoked by activation
(14). Similarly, the increase in cerebral glucose utilization
evoked by stimulation of the rat whiskers after transient global
ischemia is severely depressed 1 day after ischemia (13).
Therefore, after focal and global ischemia, both CBF and
metabolic responses are depressed even in intact regions.
probe into the functional relationship among these cells without disruption of their natural working environment. Recently
introduced in vivo imaging technologies, such as two-photon
confocal microscopy, may provide the opportunity to shed
light on these complex relationships. In disease states, vascular
dysregulation may act synergistically with other pathologies to
aggravate the intensity of the insult. It would be important to
gain insight into these synergistic interactions to determine the
extent to which the deleterious effects of vascular dysregulation contribute to the overall brain dysfunction. Vascular dysregulation may prove to be a valuable therapeutic target in a
wide variety of brain diseases.
Vascular Oxidative Stress: a Final Common Pathway to
Cerebrovascular Dysfunction
GRANTS
FUTURE DIRECTIONS
As discussed above, the mechanisms of the regulation of
CBF during brain activity involve the coordinated interaction
of neurons, glia, and vascular cells. However, the molecular
nature of these highly complex processes has not been elucidated in sufficient detail. New and powerful tools are needed to
J Appl Physiol • VOL
We thank Dr. Carrie Drake for comments.
This research was supported by National Institutes of Health Grants
HL-18974 and NS-37853.
REFERENCES
1. Akgoren N, Fabricius M, and 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.
2. Ances BM. Coupling of changes in cerebral blood flow with neural
activity: what must initially dip must come back up. J Cereb Blood Flow
Metab 24: 1– 6, 2004.
3. Attwell D and Iadecola C. The neural basis of functional brain imaging
signals. Trends Neurosci 25: 621– 625, 2002.
4. Bookheimer SY, Strojwas MH, Cohen MS, Saunders AM, PericakVance MA, Mazziotta JC, and Small GW. Patterns of brain activation
in people at risk for Alzheimer’s disease. N Engl J Med 343: 450 – 456,
2000.
5. Buerk DG, Ances BM, Greenberg JH, and Detre JA. Temporal dynamics of brain tissue nitric oxide during functional forepaw stimulation
in rats. Neuroimage 18: 1–9, 2003.
6. Bundo M, Inao S, Nakamura A, Kato T, Ito K, Tadokoro M, Kabeya
R, Sugimoto T, Kajita Y, and Yoshida J. Changes of neural activity
correlate with the severity of cortical ischemia in patients with unilateral
major cerebral artery occlusion. Stroke 33: 61– 66, 2002.
7. Cai H, Griendling KK, and Harrison DG. The vascular NAD(P)H
oxidases as therapeutic targets in cardiovascular diseases. Trends Pharmacol Sci 24: 471– 478, 2003.
8. Cauli B, Tong XK, Rancillac A, Serluca N, Lambolez B, Rossier J, and
Hamel E. Cortical GABA interneurons in neurovascular coupling: relays
for subcortical vasoactive pathways. J Neurosci 24: 8940 – 8949, 2004.
9. Crawford F, Suo Z, Fang C, and Mullan M. Characteristics of the in
vitro vasoactivity of beta-amyloid peptides. Exp Neurol 150: 159 –168,
1998.
10. De Champlain J, Wu R, Girouard H, Karas M, El Midaoui A,
Laplante MA, and Wu L. Oxidative stress in hypertension. Clin Exp
Hypertens 26: 593– 601, 2004.
11. Dickinson CJ. Why are strokes related to hypertension? Classic studies
and hypotheses revisited. J Hypertens 19: 1515–1521, 2001.
12. Dietrich HH, Kajita Y, and Dacey RG Jr. Local and conducted vasomotor responses in isolated rat cerebral arterioles. Am J Physiol Heart Circ
Physiol 271: H1109 –H1116, 1996.
13. Dietrich WD, Ginsberg MD, and Busto R. Effect of transient cerebral
ischemia on metabolic activation of a somatosensory circuit. J Cereb
Blood Flow Metab 6: 405– 413, 1986.
14. Duckrow RB, LaManna JS, and Rosenthal M. Disparate recovery of
resting and stimulated oxidative metabolism following transient ischemia.
Stroke 12: 677– 686, 1981.
15. Escobales N and Crespo MJ. Oxidative-nitrosative stress in hypertension. Curr Vasc Pharmacol 3: 231–246, 2005.
16. Faraci FM. Oxidative stress: the curse that underlies cerebral vascular
dysfunction? Stroke 36: 186 –188, 2005.
17. Faraci FM, Baumbach GL, and Heistad DD. Cerebral circulation:
humoral regulation and effects of chronic hypertension. J Am Soc Nephrol
1: 53–57, 1990.
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
There is accumulating evidence that vascular oxidative
stress leads to profound alterations in cerebrovascular regulation (16). Hypertension, AD, and cerebral ischemia are associated with evidence of oxidative stress in cerebral blood
vessels (10, 31, 78). Thus it is likely that vascular oxidative
stress is responsible for the cerebrovascular alterations observed in these conditions. There is ample evidence that ANG
II-induced experimental hypertension, as well as hypertension
in humans, is associated with vascular oxidative stress (15).
Furthermore, studies in APP mice have shown that there is free
radical production in cerebral vessels at a time when there is no
evidence of oxidative stress in the brain parenchyma (63).
Antioxidants attenuate the cerebrovascular dysfunction in
models of ANG II-induced hypertension and in APP mice,
whereas transgenic mice overexpressing the free radical scavenging enzyme SOD are protected from the dysregulation (35,
39, 63). Because an increase in reactive oxygen species (ROS)
generation has been demonstrated in cerebral ischemia (78), it
is conceivable that the mechanisms responsible for an impaired
functional hyperemia after ischemia are similar to those observed in hypertension and AD. Indeed, ROS scavengers
ameliorate the disturbance in CBF produced by ischemiareperfusion (30, 78), but it is not known whether ROS scavengers improve the alteration in neurovascular coupling.
ROS are produced by several enzymatic systems (78), but
recent findings have identified NADPH oxidase as a major
source of ROS at the vascular level (7). Inhibition of NADPH
oxidase attenuates the ROS production in models of hypertension and AD, whereas mice lacking the catalytic subunit of the
enzyme (gp91phox) are protected from the deleterious cerebrovascular effects of hypertension or A␤ (39, 64). Mice lacking
gp91phox have reduced brain damage after middle cerebral
artery occlusion (79). Therefore, NADPH oxidase-derived
ROS could also play a role in postischemic cerebrovascular
dysregulation.
ACKNOWLEDGMENTS
Invited Review
334
NEUROVASCULAR COUPLING
J Appl Physiol • VOL
41. Ko KR, Ngai AC, and Winn HR. Role of adenosine in regulation of
regional cerebral blood flow in sensory cortex. Am J Physiol Heart Circ
Physiol 259: H1703–H1708, 1990.
42. Krainik A, Hund-Georgiadis M, Zysset S, and von Cramon DY.
Regional impairment of cerebrovascular reactivity and BOLD signal in
adults after stroke. Stroke 36: 1146 –1152, 2005.
43. Krimer LS, Muly EC 3rd, Williams GV, and Goldman-Rakic PS.
Dopaminergic regulation of cerebral cortical microcirculation. Nat Neurosci 1: 286 –289, 1998.
44. Krizanac-Bengez L, Mayberg MR, and Janigro D. The cerebral vasculature as a therapeutic target for neurological disorders and the role of
shear stress in vascular homeostatis and pathophysiology. Neurol Res 26:
846 – 853, 2004.
45. Kuschinsky W, Wahl M, Bosse O, and Thurau K. Perivascular potassium and pH as determinants of local pial arterial diameter in cats. A
microapplication study. Circ Res 31: 240 –247, 1972.
46. Lassen NA. Cerebral blood flow and oxygen consumption in man. Physiol
Rev 39: 183–238, 1959.
47. Li J and Iadecola C. Nitric oxide and adenosine mediate vasodilation
during functional activation in cerebellar cortex. Neuropharmacology 33:
1453–1461, 1994.
48. Lindauer U, Megow D, Matsuda H, and Dirnagl U. Nitric oxide: a
modulator, but not a mediator, of neurovascular coupling in rat somatosensory cortex. Am J Physiol Heart Circ Physiol 277: H799 –H811, 1999.
49. Lovick TA, Brown LA, and Key BJ. Neurovascular relationships in
hippocampal slices: physiological and anatomical studies of mechanisms
underlying flow-metabolism coupling in intraparenchymal microvessels.
Neuroscience 92: 47– 60, 1999.
50. Mentis MJ, Horwitz B, Grady CL, Alexander GE, VanMeter JW,
Maisog JM, Pietrini P, Schapiro MB, and Rapoport SI. Visual cortical
dysfunction in Alzheimer’s disease evaluated with a temporally graded
“stress test” during PET. Am J Psychiatry 153: 32– 40, 1996.
51. Montecot C, Seylaz J, and Pinard E. Carbon monoxide regulates
cerebral blood flow in epileptic seizures but not in hypercapnia. Neuroreport 9: 2341–2346, 1998.
52. Mulligan SJ and MacVicar BA. Calcium transients in astrocyte endfeet
cause cerebrovascular constrictions. Nature 431: 195–199, 2004.
53. Ngai AC and Winn HR. Pial arteriole dilation during somatosensory
stimulation is not mediated by an increase in CSF metabolites. Am J
Physiol Heart Circ Physiol 282: H902–H907, 2002.
54. Nguyen TS, Winn HR, and Janigro D. ATP-sensitive potassium channels may participate in the coupling of neuronal activity and cerebrovascular tone. Am J Physiol Heart Circ Physiol 278: H878 –H885, 2000.
55. Niwa K, Araki E, Morham SG, Ross ME, and Iadecola C. Cyclooxygenase-2 contributes to functional hyperemia in whisker-barrel cortex.
J Neurosci 20: 763–770, 2000.
56. Niwa K, Carlson GA, and Iadecola C. Exogenous A␤1– 40 reproduces
cerebrovascular alterations resulting from amyloid precursor protein overexpression in mice. J Cereb Blood Flow Metab 20: 1659 –1668, 2000.
57. Niwa K, Haensel C, Ross ME, and Iadecola C. Cyclooxygenase-1
participates in selected vasodilator responses of the cerebral circulation.
Circ Res 88: 600 – 608, 2001.
58. Niwa K, Kazama K, Younkin L, Younkin SG, Carlson GA, and
Iadecola C. Cerebrovascular autoregulation is profoundly impaired in
mice overexpressing amyloid precursor protein. Am J Physiol Heart Circ
Physiol 283: H315–H323, 2002.
59. Niwa K, Porter VA, Kazama K, Cornfield D, Carlson GA, and
Iadecola C. A␤-peptides enhance vasoconstriction in cerebral circulation.
Am J Physiol Heart Circ Physiol 281: H2417–H2424, 2001.
60. Niwa K, Younkin L, Ebeling C, Turner SK, Westaway D, Younkin S,
Ashe KH, Carlson GA, and Iadecola C. A␤1– 40-related reduction in
functional hyperemia in mouse neocortex during somatosensory activation. Proc Natl Acad Sci USA 97: 9735–9740, 2000.
61. Northington FJ, Matherne GP, and Berne RM. Competitive inhibition
of nitric oxide synthase prevents the cortical hyperemia associated with
peripheral nerve stimulation. Proc Natl Acad Sci USA 89: 6649 – 6652,
1992.
62. Norup Nielsen A and Lauritzen M. Coupling and uncoupling of activitydependent increases of neuronal activity and blood flow in rat somatosensory cortex. J Physiol 533: 773–785, 2001.
63. Park L, Anrather J, Forster C, Kazama K, Carlson GA, and Iadecola
C. Abeta-induced vascular oxidative stress and attenuation of functional
hyperemia in mouse somatosensory cortex. J Cereb Blood Flow Metab 24:
334 –342, 2004.
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
18. Faraci FM and Breese KR. Nitric oxide mediates vasodilation in response to activation of N-methyl-D-aspartate receptors in brain. Circ Res
72: 476 – 480, 1993.
19. Faraci FM and Heistad DD. Regulation of the cerebral circulation: role
of endothelium and potassium channels. Physiol Rev 78: 53–97, 1998.
20. Faraci FM and Sobey CG. Role of potassium channels in regulation of
cerebral vascular tone. J Cereb Blood Flow Metab 18: 1047–1063, 1998.
21. Farkas E and Luiten PG. Cerebral microvascular pathology in aging and
Alzheimer’s disease. Prog Neurobiol 64: 575– 611, 2001.
22. Fergus A and Lee KS. GABAergic regulation of cerebral microvascular
tone in the rat. J Cereb Blood Flow Metab 17: 992–1003, 1997.
23. Filosa JA, Bonev AD, and Nelson MT. Calcium dynamics in cortical
astrocytes and arterioles during neurovascular coupling. Circ Res 95:
e73– e81, 2004.
24. Fujishima M, Ibayashi S, Fujii K, and Mori S. Cerebral blood flow and
brain function in hypertension. Hypertens Res 18: 111–117, 1995.
25. Garavito RM and Mulichak AM. The structure of mammalian cyclooxygenases. Annu Rev Biophys Biomol Struct 32: 183–206, 2003.
25a.Hamel E. Perivascular nerves and the regulation of cerebrovascular tone.
J Appl Physiol. In press.
26. Hamzei F, Knab R, Weiller C, and Rother J. The influence of extra- and
intracranial artery disease on the BOLD signal in FMRI. Neuroimage 20:
1393–1399, 2003.
27. Heistad DD and Kontos HA. Cerebral circulation. In: Handbook of
Physiology. The Cardiovascular System. Circulation and Organ Blood
Flow. Bethesda, MD: Am. Physiol. Soc., 1983, sect. 2, vol. III, pt. 1, chapt.
5, p. 137–182.
28. Heuser D. The significance of cortical extracellular H⫹, K⫹ and Ca2⫹
activities for regulation of local cerebral blood flow under conditions of
enhanced neuronal activity. In: Cerebral Vascular Smooth Muscle and Its
Control, edited by Elliot K and O’Connor M. New York: Elsevier, 1978,
p. 339 –353.
29. Hock C, Villringer K, Muller-Spahn F, Wenzel R, Heekeren H,
Schuh-Hofer S, Hofmann M, Minoshima S, Schwaiger M, Dirnagl U,
and Villringer A. Decrease in parietal cerebral hemoglobin oxygenation
during performance of a verbal fluency task in patients with Alzheimer’s
disease monitored by means of near-infrared spectroscopy (NIRS)—
correlation with simultaneous rCBF-PET measurements. Brain Res 755:
293–303, 1997.
30. Iadecola C. Cerebral circulatory dysregulation in ischemia. In: Cerebrovascular Diseases, edited by Ginsberg M and Bogousslavsky J. Cambridge, UK: Blackwell, 1998, p. 319 –332.
31. Iadecola C. Neurovascular regulation in the normal brain and in Alzheimer’s disease. Nat Rev Neurosci 5: 347–360, 2004.
32. Iadecola C. Regulation of the cerebral microcirculation during neural
activity: is nitric oxide the missing link? Trends Neurosci 16: 206 –214,
1993.
33. Iadecola C, Li J, Ebner TJ, and Xu X. Nitric oxide contributes to
functional hyperemia in cerebellar cortex. Am J Physiol Regul Integr
Comp Physiol 268: R1153–R1162, 1995.
34. Iadecola C, Yang G, Ebner TJ, and Chen G. Local and propagated
vascular responses evoked by focal synaptic activity in cerebellar cortex.
J Neurophysiol 78: 651– 659, 1997.
35. Iadecola C, Zhang F, Niwa K, Eckman C, Turner SK, Fischer E,
Younkin S, Borchelt DR, Hsiao KK, and Carlson GA. SOD1 rescues
cerebral endothelial dysfunction in mice overexpressing amyloid precursor
protein. Nat Neurosci 2: 157–161, 1999.
36. Inao S, Tadokoro M, Nishino M, Mizutani N, Terada K, Bundo M,
Kuchiwaki H, and Yoshida J. Neural activation of the brain with
hemodynamic insufficiency. J Cereb Blood Flow Metab 18: 960 –967,
1998.
37. Jennings JR, Muldoon MF, Ryan C, Price JC, Greer P, Sutton-Tyrrell
K, van der Veen FM, and Meltzer CC. Reduced cerebral blood flow
response and compensation among patients with untreated hypertension.
Neurology 64: 1358 –1365, 2005.
38. Kasischke KA, Vishwasrao HD, Fisher PJ, Zipfel WR, and Webb
WW. Neural activity triggers neuronal oxidative metabolism followed by
astrocytic glycolysis. Science 305: 99 –103, 2004.
39. Kazama K, Anrather J, Zhou P, Girouard H, Frys K, Milner TA, and
Iadecola C. Angiotensin II impairs neurovascular coupling in neocortex
through NADPH oxidase-derived radicals. Circ Res 95: 1019 –1026, 2004.
40. Kazama K, Wang G, Frys K, Anrather J, and Iadecola C. Angiotensin
II attenuates functional hyperemia in the mouse somatosensory cortex.
Am J Physiol Heart Circ Physiol 285: H1890 –H1899, 2003.
Invited Review
NEUROVASCULAR COUPLING
J Appl Physiol • VOL
78. Traystman RJ, Kirsch JR, and Koehler RC. Oxygen radical mechanisms of brain injury following ischemia and reperfusion. J Appl Physiol
71: 1185–1195, 1991.
79. Walder CE, Green SP, Darbonne WC, Mathias J, Rae J, Dinauer MC,
Curnutte JT, and Thomas GR. Ischemic stroke injury is reduced in mice
lacking a functional NADPH oxidase. Stroke 28: 2252–2258, 1997.
80. Wang H, Hitron IM, Iadecola C, and Pickel VM. Synaptic and vascular
associations of neurons containing cyclooxygenase-2 and nitric oxide
synthase in rat somatosensory cortex. Cereb Cortex 15: 1250 –1260, 2005.
81. Ward NL and Lamanna JC. The neurovascular unit and its growth
factors: coordinated response in the vascular and nervous systems. Neurol
Res 26: 870 – 883, 2004.
82. Warkentin S and Passant U. Functional imaging of the frontal lobes in
organic dementia. Regional cerebral blood flow findings in normals, in
patients with frontotemporal dementia and in patients with Alzheimer’s
disease, performing a word fluency test. Dement Geriatr Cogn Disord 8:
105–109, 1997.
83. Yaksh TL, Wang JY, and Go VL. Cortical vasodilatation produced by
vasoactive intestinal polypeptide (VIP) and by physiological stimuli in the
cat. J Cereb Blood Flow Metab 7: 315–326, 1987.
84. Yang G, Huard JM, Beitz AJ, Ross ME, and Iadecola C. Stellate
neurons mediate functional hyperemia in the cerebellar molecular layer.
J Neurosci 20: 6968 – 6973, 2000.
85. Yang G and Iadecola C. Glutamate microinjections in cerebellar cortex
reproduce cerebral vascular effects of parallel fiber stimulation. Am J
Physiol Regul Integr Comp Physiol 271: R1568 –R1575, 1996.
86. Yang G and Iadecola C. Obligatory role of NO in glutamate-dependent
hyperemia evoked from cerebellar parallel fibers. Am J Physiol Regul
Integr Comp Physiol 272: R1155–R1161, 1997.
87. Yang G, Zhang Y, Ross ME, and Iadecola C. Attenuation of activityinduced increases in cerebellar blood flow in mice lacking neuronal nitric
oxide synthase. Am J Physiol Heart Circ Physiol 285: H298 –H304, 2003.
88. Zhang F, Eckman C, Younkin S, Hsiao KK, and Iadecola C. Increased
susceptibility to ischemic brain damage in transgenic mice overexpressing
the amyloid precursor protein. J Neurosci 17: 7655–7661, 1997.
89. Zlokovic BV, Deane R, Sallstrom J, Chow N, and Miano JM. Neurovascular pathways and Alzheimer amyloid beta-peptide. Brain Pathol 15:
78 – 83, 2005.
90. Zonta M, Angulo MC, Gobbo S, Rosengarten B, Hossmann KA,
Pozzan T, and Carmignoto G. Neuron-to-astrocyte signaling is central to
the dynamic control of brain microcirculation. Nat Neurosci 6: 43–50,
2003.
100 • JANUARY 2006 •
www.jap.org
Downloaded from http://jap.physiology.org/ by 10.220.33.3 on May 5, 2017
64. Park L, Anrather J, Zhou P, Frys K, Pitstick R, Younkin S, Carlson
GA, and Iadecola C. NADPH oxidase-derived reactive oxygen species
mediate the cerebrovascular dysfunction induced by the amyloid beta
peptide. J Neurosci 25: 1769 –1777, 2005.
65. Paulson OB and Newman EA. Does the release of potassium from
astrocyte endfeet regulate cerebral blood flow? Science 237: 896 – 898,
1987.
66. Pellerin L, Pellegri G, Bittar PG, Charnay Y, Bouras C, Martin JL,
Stella N, and Magistretti PJ. Evidence supporting the existence of an
activity-dependent astrocyte-neuron lactate shuttle. Dev Neurosci 20:
291–299, 1998.
67. Peng X, Carhuapoma JR, Bhardwaj A, Alkayed NJ, Falck JR, Harder
DR, Traystman RJ, and Koehler RC. Suppression of cortical functional
hyperemia to vibrissal stimulation in the rat by epoxygenase inhibitors.
Am J Physiol Heart Circ Physiol 283: H2029 –H2037, 2002.
68. Peters APS and Webster HD. The Fine Structure of the Nervous System.
New York: Oxford University Press, 1991.
69. Pineiro R, Pendlebury S, Johansen-Berg H, and Matthews PM. Altered hemodynamic responses in patients after subcortical stroke measured
by functional MRI. Stroke 33: 103–109, 2002.
70. Rother J, Knab R, Hamzei F, Fiehler J, Reichenbach JR, Buchel C,
and Weiller C. Negative dip in BOLD fMRI is caused by blood flow—
oxygen consumption uncoupling in humans. Neuroimage 15: 98 –102,
2002.
71. Roy C and Sherrington C. On the regulation of the blood-supply of the
brain. J Physiol 11: 85–108, 1890.
72. Rubio R, Berne RM, Bockman EL, and Curnish RR. Relationship
between adenosine concentration and oxygen supply in rat brain. Am J
Physiol 228: 1896 –1902, 1975.
73. Schmidt C. The Cerebral Circulation in Health and Disease. Springfield,
IL: Thomas, 1950.
74. Scremin OU, Rovere AA, Raynald AC, and Giardini A. Cholinergic
control of blood flow in the cerebral cortex of the rat. Stroke 4: 232–239,
1973.
75. Segal SS. Regulation of blood flow in the microcirculation. Microcirculation 12: 33– 45, 2005.
76. Selkoe DJ and Schenk D. Alzheimer’s disease: molecular understanding
predicts amyloid-based therapeutics. Annu Rev Pharmacol Toxicol 43:
545–584, 2003.
77. Smith CD, Andersen AH, Kryscio RJ, Schmitt FA, Kindy MS,
Blonder LX, and Avison MJ. Altered brain activation in cognitively
intact individuals at high risk for Alzheimer’s disease. Neurology 53:
1391–1396, 1999.
335