ArticlePDF AvailableLiterature Review

Neurovascular coupling in the normal brain and in hypertension, stroke, and Alzheimer disease

Authors:

Abstract and Figures

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.
Content may be subject to copyright.
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
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 de-
pendent 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, endothe-
lial 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 hyperten-
sion, Alzheimer disease, and ischemic stroke. Consequently, CBF is no longer
matched to the metabolic requirements of the tissue. This cerebrovascular dysregu-
lation 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.
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 homeo-
static unbalance may contribute to brain dysfunction. This
review provides a brief summary of neurovascular coupling in
the normal state and in diseases including hypertension, ische-
mic 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 (astro-
cytes, microglia, oligodendrocytes), and vascular cells (endo-
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 neu-
rovascular 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 sep-
arated from the brain by the Virchow-Robin space, which
contains cerebrospinal fluid. On the outer side of the Virchow-
Robin space, astrocytes give rise to the glia limitans mem-
brane. As the arterioles penetrate deeper into the brain, the
Virchow-Robin space disappears and the vascular basement
Address for reprint requests and other correspondence: C. Iadecola, Division
of Neurobiology, 411 East 69th St., Rm. KB410, New York, NY 10021
(e-mail: coi2001@med.cornell.edu).
J Appl Physiol 100: 328 –335, 2006;
doi:10.1152/japplphysiol.00966.2005.
8750-7587/06 $8.00 Copyright
©
2006 the American Physiological Society http://www. jap.org328
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 astro-
cytic 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. Endo-
thelial 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 hyper-
polarizing factor, and endothelin (19). Pericytes have contrac-
tile properties and may modulate capillary diameter (31).
Perivascular astrocytes surround most of the capillary ablumi-
nal 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).
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 neuro-
vascular coupling (Table 1). These include ions, metabolic
by-products, vasoactive neurotransmitters, and vasoactive fac-
tors released in response to neurotransmitters.
Vasoactive ions. K
and H
are generated by the extracel
-
lular 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
muscle cells (20, 54), leading to their hyperpolarization and
subsequent relaxation. During sustained activation, ATP de-
pletion could lead to opening of ATP-sensitive K
channels
(K
ATP
) on vessels. Therefore, K
ATP
channels have been impli
-
cated in the mechanisms of neurovascular coupling (54). Fur-
thermore, K
ATP
could also participate in neurovascular cou
-
pling 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 Ca
2
may produce vasodilation (28).
Vasoactive factors related to energy metabolism. The brain
has little energy reserve and requires a continuous supply of
glucose and O
2
through CBF. A sudden increase in the demand
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.
Table 1. Factors implicated in neurovascular coupling
Agent References
Vasoactive ions
K
45
H
45
Ca
2
28
Metabolic factors
Lactate, CO
2
71, 73
Hypoxia 73
Adenosine 72
Vasoactive neurotransmitters
Dopamine 43
GABA 22
Acetylcholine 74
Vasoactive intestinal peptide 83
Other
NO 61
COX-2 products 55
P450 products 67
CO 51
NO, nitric oxide; COX-2, cyclooxygenase-2; CO, carbon monoxide.
Invited Review
329NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
for energy during synaptic activity could result in a relative
lack of O
2
and glucose, which may be a factor in triggering the
hemodynamic response (3). However, the reduction in brain O
2
concentration at the site of activation is small and transient and
cannot account for sustained increases in flow (2). Further-
more, the CBF response to activation is not altered by hypo-
glycemia or hypoxia, suggesting that lack of glucose or O
2
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 cerebel-
lum (47) and cerebral cortex (41). Lactate produced during
brain activation could also be an important mediator of func-
tional hyperemia by increasing H
concentration and produc
-
ing 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 hippocam-
pal 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 O
2
(“dip”) observed at the onset of
activation with optical imaging, functional MRI, and O
2
elec
-
trodes (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 neurotrans-
mitters. It has long been proposed that vasoactive neurotrans-
mitters released during neural activity contribute to the vaso-
dilation (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 acti-
vation 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. Vaso-
active factors can also be generated by the intracellular signal-
ing induced by activation of neurotransmitter receptors. For
example, activation of glutamate receptors produces vaso-
dilation and increases blood flow. In neocortex and hip-
pocampus, 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-methylisoxazole-
propionic acid (CAMPA) and/or NMDA receptor blockers (1,
33, 62). Because glutamate is not vasoactive in isolated cere-
bral arteries (18), the vasodilation is mediated by vasoactive
factors whose synthesis is triggered by the changes in intracel-
lular Ca
2
associated with glutamate receptor activation. The
increase in Ca
2
activates Ca
2
-dependent enzymes that pro
-
duce potent vasodilators. One such enzyme is the neuronal
isoform of NO synthase (nNOS), which produces the vasodi-
lator 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 inhib-
itors (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 ob-
served also in nNOS null mice (86, 87). These findings suggest
that, in cerebellum, NO plays an obligatory role in the mech-
anisms of the vasodilation.
The increase in intracellular Ca
2
evoked by glutamate also
activates phospholipase A2, leading to production of arachi-
donic acid. Arachidonic acid is then metabolized by the cyclo-
oxygenase (COX) pathway, producing vasodilatatory prosta-
glandins (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 metabo-
lites 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 in-
crease in CBF produced by activation (see Koehler RC, Ge-
bremedhin 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
Invited Review
330 NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
astrocytes, leading to increases in intracellular Ca
2
, COX
activation, and local vasodilation (90). A critical role for Ca
2
fluxes from neurons to astrocytes and arterioles is also sup-
ported by another study in brain slices (23). Arterioles at rest
exhibit periodic contractions and relaxations, called vasomo-
tion, a phenomenon resulting from Ca
2
oscillations in smooth
muscle cells (23). It was found that electrical stimulation of
brain slices induces Ca
2
waves in astrocytes, which then
propagate to arterioles and inhibit vasomotion (23). The inhi-
bition 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 Ca
2
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 be-
cause they provide evidence that activity-induced Ca
2
tran
-
sients in astrocytes are linked to changes in cerebrovascular
tone.
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 opti-
mize 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 vasodi-
lation of Duling and Berne (75). Indeed, retrograde vasodila-
tion in response to ATP has been observed in isolated rat
cerebral arterioles (12), whereas upstream vasodilation has
been demonstrated in cerebellar cortex arterioles during acti-
vation 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 re-
sponse 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 up-
stream 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 vasodi-
lator substances from the glia limitans adjacent to upstream
pial arteries plays a role in remote vasodilation because flush-
ing 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 prop-
agation 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, peri-
cytes, 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 hyper-
tension, 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 struc-
tural alterations facilitate vascular occlusions and compromise
cerebral perfusion. In addition, hypertension impairs the func-
tion of cerebral blood vessels (17, 27). Hypertension impairs
endothelium-dependent relaxation (19) and alters cerebrovas-
cular 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 somato-
sensory 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
Invited Review
331NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
does not reproduce the effects of ANG II on functional hyper-
emia. Although these observations suggest that the cerebrovas-
cular effect of short-term administration of ANG II is indepen-
dent 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 (Ta-
ble 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 indi-
viduals (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 coex-
isting 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 de-
mentia and is characterized by deposition of amyloid -peptide
(A) in the neuropil (neuritic plaques) and blood vessels
(amyloid angiopathy), and by accumulation of hyperphospho-
rylated 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 dys-
function 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, re-
sponses to vasoconstrictors are exaggerated (35, 59). Func-
tional hyperemia is impaired and cerebrovascular autoregula-
tion 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 exacer-
bate the dementia (31). Furthermore, insufficient CBF may
alter A trafficking across the blood-brain barrier (89). There-
fore, 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 de-
mentia.
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 hypoper-
fusion). After ischemia, the cerebral circulation is in a state of
vasoparalysis (30). After ischemic stroke in patients, the reac-
Table 2. Alterations in functional hyperemia in Alzheimer disease, cerebrovascular diseases,
and hypertension: selected human studies
Pathology Activation Task
Method for CBF
Detection Effect on Cortical Hemodynamic Change Reference
Alzheimer disease and APOE genotype
Alzheimer disease Visual stimulus PET 2 Middle temporal 50
Alzheimer disease Verbal task PET and NIRS 2 Parietal, frontal 29
Alzheimer disease Verbal task Xenon-133 2 Dorsolateral 82
APOEε4 carriers Visual naming and letter task fMRI 2 Prefrontal 77
APOEε4 carriers Memory tasks fMRI 2 Mid and posterior inferior temporal 4
Cerebrovascular diseases
Ischemic stroke with full recovery Hand grip fMRI 2 Left hippocampus 42
Lacunar stroke Finger tapping fMRI 2 Somatosensory cortex, supplementary
motor area
69
Extra- and intracranial stenosis Hand grip fMRI 2 Somatosensory cortex, supplementary
motor area
26
Common carotid or vertebral
artery occlusion
Finger tapping fMRI 2 Supplementary motor area 70
Internal carotid or middle cerebral
artery occlusion
Bimanual motor tasks PET No change in somatosensory cortex 36
Essential hypertension Memory task PET 2 Posterior parietal 37
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
332 NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
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 reduc-
tion 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.
Vascular Oxidative Stress: a Final Common Pathway to
Cerebrovascular Dysfunction
There is accumulating evidence that vascular oxidative
stress leads to profound alterations in cerebrovascular regula-
tion (16). Hypertension, AD, and cerebral ischemia are asso-
ciated 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 ob-
served 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 scav-
enging 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 ob-
served in hypertension and AD. Indeed, ROS scavengers
ameliorate the disturbance in CBF produced by ischemia-
reperfusion (30, 78), but it is not known whether ROS scav-
engers 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 hyperten-
sion and AD, whereas mice lacking the catalytic subunit of the
enzyme (gp91
phox
) are protected from the deleterious cerebro
-
vascular effects of hypertension or A (39, 64). Mice lacking
gp91
phox
have reduced brain damage after middle cerebral
artery occlusion (79). Therefore, NADPH oxidase-derived
ROS could also play a role in postischemic cerebrovascular
dysregulation.
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 eluci-
dated in sufficient detail. New and powerful tools are needed to
probe into the functional relationship among these cells with-
out 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 dysregula-
tion contribute to the overall brain dysfunction. Vascular dys-
regulation may prove to be a valuable therapeutic target in a
wide variety of brain diseases.
ACKNOWLEDGMENTS
We thank Dr. Carrie Drake for comments.
GRANTS
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, Pericak-
Vance 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 dy-
namics 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 Phar-
macol 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 vaso-
motor 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 hyperten-
sion. 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.
Invited Review
333NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
18. Faraci FM and Breese KR. Nitric oxide mediates vasodilation in re-
sponse 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 cyclooxy-
genases. 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 Ca
2
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: Cerebro-
vascular Diseases, edited by Ginsberg M and Bogousslavsky J. Cam-
bridge, UK: Blackwell, 1998, p. 319 –332.
31. Iadecola C. Neurovascular regulation in the normal brain and in Alzhei-
mer’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.
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 Neu-
rosci 1: 286–289, 1998.
44. Krizanac-Bengez L, Mayberg MR, and Janigro D. The cerebral vas-
culature 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 potas-
sium 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 somato-
sensory 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. Neuro-
report 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 chan-
nels may participate in the coupling of neuronal activity and cerebrovas-
cular tone. Am J Physiol Heart Circ Physiol 278: H878 –H885, 2000.
55. Niwa K, Araki E, Morham SG, Ross ME, and Iadecola C. Cyclooxy-
genase-2 contributes to functional hyperemia in whisker-barrel cortex.
J Neurosci 20: 763–770, 2000.
56. Niwa K, Carlson GA, and Iadecola C. Exogenous A1– 40 reproduces
cerebrovascular alterations resulting from amyloid precursor protein over-
expression 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. A1– 40-related reduction in
functional hyperemia in mouse neocortex during somatosensory activa-
tion. 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 activity-
dependent increases of neuronal activity and blood flow in rat somatosen-
sory 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.
Invited Review
334 NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
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. Al-
tered 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. Microcircu-
lation 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.
78. Traystman RJ, Kirsch JR, and Koehler RC. Oxygen radical mecha-
nisms 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 activity-
induced 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. Neuro-
vascular 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.
Invited Review
335NEUROVASCULAR COUPLING
J Appl Physiol VOL 100 JANUARY 2006 www.jap.org
... A similar statement holds for the right PFC when the EEG gamma power served as a node of the network. This observation is in good agreement with the neurophysiological source of oscillation in the endogenic band where the contraction and relaxation of the endothelial layer of the vessel wall are a key component of the blood-brain barrier (BBB), which regulates the hemodynamics, metabolism, and neural activity in the human brain 12,66 . In other words, the oxygen demand-supply balance between cerebral hemodynamics versus metabolic activity is mainly controlled in this frequency band. ...
... Novel construction of frequency-specific neurophysiological networks and potential applications In this study, we investigated and constructed cerebral neurophysiological networks in three ISO frequency bands based on their physiological representations. The slowest endogenic band is associated with relaxationcontraction oscillation in the endothelium mediated by released potent vasoactive factors, such as nitric oxide (NO), prostacyclin, free radicals, endothelin, and endothelium-derived hyperpolarizing factor 12,66 . The neurogenic band corresponds to dilation-contraction cycles of vessel walls affected by neurotransmitters and vasoactive ions from neurons 67 . ...
Article
Full-text available
Cerebral infra-slow oscillation (ISO) is a source of vasomotion in endogenic (E; 0.005–0.02 Hz), neurogenic (N; 0.02–0.04 Hz), and myogenic (M; 0.04–0.2 Hz) frequency bands. In this study, we quantified changes in prefrontal concentrations of oxygenated hemoglobin (Δ[HbO]) and redox-state cytochrome c oxidase (Δ[CCO]) as hemodynamic and metabolic activity metrics, and electroencephalogram (EEG) powers as electrophysiological activity, using concurrent measurements of 2-channel broadband near-infrared spectroscopy and EEG on the forehead of 22 healthy participants at rest. After preprocessing, the multi-modality signals were analyzed using generalized partial directed coherence to construct unilateral neurophysiological networks among the three neurophysiological metrics (with simplified symbols of HbO, CCO, and EEG) in each E/N/M frequency band. The links in these networks represent neurovascular, neurometabolic, and metabolicvascular coupling (NVC, NMC, and MVC). The results illustrate that the demand for oxygen by neuronal activity and metabolism (EEG and CCO) drives the hemodynamic supply (HbO) in all E/N/M bands in the resting prefrontal cortex. Furthermore, to investigate the effect of transcranial photobiomodulation (tPBM), we performed a sham-controlled study by delivering an 800-nm laser beam to the left and right prefrontal cortex of the same participants. After performing the same data processing and statistical analysis, we obtained novel and important findings: tPBM delivered on either side of the prefrontal cortex triggered the alteration or reversal of directed network couplings among the three neurophysiological entities (i.e., HbO, CCO, and EEG frequency-specific powers) in the physiological network in the E and N bands, demonstrating that during the post-tPBM period, both metabolism and hemodynamic supply drive electrophysiological activity in directed network coupling of the prefrontal cortex (PFC). Overall, this study revealed that tPBM facilitates significant modulation of the directionality of neurophysiological networks in electrophysiological, metabolic, and hemodynamic activities.
... the brain's glial and neural networks (i.e., the hemo-neural hypothesis) (Moore and Cao, 2008). More specifically, an exercisemediated increase in CBF is proposed to improve neurovascular coupling and cerebral perfusion (for review see, Girouard and Iadecola, 2006) that enhance the efficiency of local neural circuits via vessel diameter changes specific to pial and penetrating arterioles (i.e., mechanical change) and improved temperature regulation (i.e., cooling) (for extensive review see, Nippert et al., 2018). Passive exercise entails movement of the limbs via an external mechanical force and is a technique known to increase CBF independent of volitional muscle activation or increased cardiorespiratory demands. ...
Article
Full-text available
Introduction: Passive exercise involves limb movement via an external force and is an intervention providing an immediate postexercise executive function (EF) benefit. It is, however, unknown whether EF is improved simultaneous with passive exercise—a salient question given the advent of passive (and active) exercise workstations designed to enhance productivity and wellbeing for individuals engaged in sedentary occupations. Methods: Here, participants (N = 23) completed separate 20-min conditions involving active (i.e., via volitional muscle activation) and passive (i.e., via mechanically driven cycle ergometer) cycle ergometry and a non-exercise control condition. EF was assessed prior to (i.e., preintervention), simultaneous with, and immediately after (post-intervention) each condition via the antipointing task. Antipointing involves a goal-directed limb movement mirror-symmetrical to a target and is an ideal tool for the current investigation given that the task is mediated via EF inhibitory control networks that show response-dependent changes following a single bout of exercise. Results and discussion: Results showed that passive exercise produced a simultaneous and post-intervention reduction in antipointing reaction time (RT), whereas active exercise selectively produced a post-intervention—but not simultaneous—RT reduction. Thus, passive and active exercise elicited a postexercise EF benefit; however, only passive exercise produced a simultaneous benefit. That passive—but not active—exercise produced a simultaneous benefit may reflect that the intervention provides the necessary physiological or psychological changes to elicit improved EF efficiency without the associated dual-task cost(s) of volitional muscle activity.
... Commensurate with the cytokines and multi-omic ( Figure 3B) associations, cerebral perfusion, and metabolism via uncoupling analysis revealed a sexually dimorphic dysregulation with age and genotype ( Figure 7,B). Importantly, the addition of a HFD in LOAD1 mice, resulted in Type 1 uncoupling (i.e., reductions in glycolysis and compensatory hyperemia), consistent with a cytokine (i.e., TNFα, IL-1β, IL-6, and/or IL-12) 26-29 driven down regulation of insulin receptors, 27 which have been shown to result in a reduction in neuronal glucose uptake via GLUT transporters. [28][29][30] Importantly, these data closely parallel the Type 2 diabetic phenotype 31,32 with reactive hyperemia via activation of eNOS. ...
Article
Full-text available
INTRODUCTION MODEL‐AD (Model Organism Development and Evaluation for Late‐Onset Alzheimer's Disease) is creating and distributing novel mouse models with humanized, clinically relevant genetic risk factors to capture the trajectory and progression of late‐onset Alzheimer's disease (LOAD) more accurately. METHODS We created the LOAD2 model by combining apolipoprotein E4 (APOE4), Trem2*R47H, and humanized amyloid‐beta (Aβ). Mice were subjected to a control diet or a high‐fat/high‐sugar diet (LOAD2+HFD). We assessed disease‐relevant outcome measures in plasma and brain including neuroinflammation, Aβ, neurodegeneration, neuroimaging, and multi‐omics. RESULTS By 18 months, LOAD2+HFD mice exhibited sex‐specific neuron loss, elevated insoluble brain Aβ42, increased plasma neurofilament light chain (NfL), and altered gene/protein expression related to lipid metabolism and synaptic function. Imaging showed reductions in brain volume and neurovascular uncoupling. Deficits in acquiring touchscreen‐based cognitive tasks were observed. DISCUSSION The comprehensive characterization of LOAD2+HFD mice reveals that this model is important for preclinical studies seeking to understand disease trajectory and progression of LOAD prior to or independent of amyloid plaques and tau tangles. Highlights By 18 months, unlike control mice (e.g., LOAD2 mice fed a control diet, CD), LOAD2+HFD mice presented subtle but significant loss of neurons in the cortex, elevated levels of insoluble Ab42 in the brain, and increased plasma neurofilament light chain (NfL). Transcriptomics and proteomics showed changes in gene/proteins relating to a variety of disease‐relevant processes including lipid metabolism and synaptic function. In vivo imaging revealed an age‐dependent reduction in brain region volume (MRI) and neurovascular uncoupling (PET/CT). LOAD2+HFD mice also demonstrated deficits in acquisition of touchscreen‐based cognitive tasks.
... Changes in electrical neuronal activity (as detected by EEG) and haemodynamic changes (as measured by brain perfusion SPECT) are tightly linked by neurovascular coupling: increased neuronal activity is associated with increased local energy demand and therefore triggers an increase of local blood flow to provide additional oxygen and glucose [4,[38][39][40][41]. Neurovascular coupling is active not only during normal somatosensory processing, but also in pathological states including epilepsy, although it is probably not sufficient to fully match the far supranormal local energy demands during a seizure [42,43]. However, neurovascular coupling does not imply that the temporal scales of changes in electrical neuronal activity and the resulting changes in cerebral perfusion are the same. ...
Article
Full-text available
Background The aim of this study was to assess the impact of the post-injection electrical seizure duration on the identification of the seizure onset zone (SOZ) in ictal brain perfusion SPECT in presurgical evaluation of drug-resistant epilepsy. Methods 176 ictal SPECT performed with 99mTc-HMPAO (n = 140) or -ECD (n = 36) were included retrospectively. Visual interpretation of the SPECT images (together with individual MRI and statistical hyperperfusion maps) with respect to lateralization (right, left, none) and localization (temporal, frontal, parietal, occipital) of the SOZ was performed by 3 independent readers. Between-readers agreement was characterized by Fleiss’ κ. An ictal SPECT was considered "lateralizing" if all readers agreed on right or left hemisphere. It was considered "localizing" if it was lateralizing and all readers agreed on the same lobe within the same hemisphere. The impact of injection latency and post-injection seizure duration on the proportion of lateralizing/localizing SPECT was tested by ANOVA with dichotomized (by the median) injection latency and post-injection seizure duration as between-subjects factors. Results Median [interquartile range] (full range) of injection latency and post-injection seizure duration were 30 [24, 40] (3–120) s and 50 [27, 70] (-20–660) s, respectively. Fleiss’ κ for lateralization of the SOZ was largest for the combination of early (< 30 s) injection and long (> 50 s) post-injection seizure duration (κ = 0.894, all other combinations κ = 0.659–0.734). Regarding Fleiss’ κ for localization of the SOZ in the 141 (80.1%) lateralizing SPECT, it was largest for early injection and short post-injection seizure duration (κ = 0.575, all other combinations κ = 0.329–0.368). The proportion of lateralizing SPECT was lower with short compared to long post-injection seizure duration (estimated marginal means 74.3% versus 86.3%, p = 0.047). The effect was mainly driven by cases with very short post-injection seizure duration ≤ 10 s (53.8% lateralizing). Injection latency in the considered range had no significant impact on the proportion of lateralizing SPECT (p = 0.390). The proportion of localizing SPECT among the lateralizing cases did not depend on injection latency or post-injection seizure duration (p ≥ 0.603). Conclusions Short post-injection seizure duration is associated with a lower proportion of lateralizing cases in ictal brain perfusion SPECT.
... In addition to their vital roles in neuroinflammation, these cells are also crucial components of the neuro-glio-vascular unit (NGVU), which act in concert to locally regulate microvascular blood flow to satisfy the brain's dynamic, heterogeneous energy demands [10]. Disruptions to the NGVU account for many of AD's other preclinical hallmarks, including reductions to cerebral perfusion, impaired glucose and oxygen metabolism, and neurovascular decoupling [11][12][13][14][15]. ...
Article
Full-text available
Background Neuroinflammation, impaired metabolism, and hypoperfusion are fundamental pathological hallmarks of early Alzheimer’s disease (AD). Numerous studies have asserted a close association between neuroinflammation and disrupted cerebral energetics. During AD progression and other neurodegenerative disorders, a persistent state of chronic neuroinflammation reportedly exacerbates cytotoxicity and potentiates neuronal death. Here, we assessed the impact of a neuroinflammatory challenge on metabolic demand and microvascular hemodynamics in the somatosensory cortex of an AD mouse model. Methods We utilized in vivo 2-photon microscopy and the phosphorescent oxygen sensor Oxyphor 2P to measure partial pressure of oxygen (pO2) and capillary red blood cell flux in cortical microvessels of awake mice. Intravascular pO2 and capillary RBC flux measurements were performed in 8-month-old APPswe/PS1dE9 mice and wildtype littermates on days 0, 7, and 14 of a 14-day period of lipopolysaccharide-induced neuroinflammation. Results Before the induced inflammatory challenge, AD mice demonstrated reduced metabolic demand but similar capillary red blood cell flux as their wild type counterparts. Neuroinflammation provoked significant reductions in cerebral intravascular oxygen levels and elevated oxygen extraction in both animal groups, without significantly altering red blood cell flux in capillaries. Conclusions This study provides evidence that neuroinflammation alters cerebral oxygen demand at the early stages of AD without substantially altering vascular oxygen supply. The results will guide our understanding of neuroinflammation’s influence on neuroimaging biomarkers for early AD diagnosis.
Article
Visualization of human brain activity is crucial for understanding normal and aberrant brain function. Currently available neural activity recording methods are highly invasive, have low sensitivity, and cannot be conducted outside of an operating room. Functional ultrasound imaging (fUSI) is an emerging technique that offers sensitive, large-scale, high-resolution neural imaging; however, fUSI cannot be performed through the adult human skull. Here, we used a polymeric skull replacement material to create an acoustic window compatible with fUSI to monitor adult human brain activity in a single individual. Using an in vitro cerebrovascular phantom to mimic brain vasculature and an in vivo rodent cranial defect model, first, we evaluated the fUSI signal intensity and signal-to-noise ratio through polymethyl methacrylate (PMMA) cranial implants of different thicknesses or a titanium mesh implant. We found that rat brain neural activity could be recorded with high sensitivity through a PMMA implant using a dedicated fUSI pulse sequence. We then designed a custom ultrasound-transparent cranial window implant for an adult patient undergoing reconstructive skull surgery after traumatic brain injury. We showed that fUSI could record brain activity in an awake human outside of the operating room. In a video game “connect the dots” task, we demonstrated mapping and decoding of task-modulated cortical activity in this individual. In a guitar-strumming task, we mapped additional task-specific cortical responses. Our proof-of-principle study shows that fUSI can be used as a high-resolution (200 μm) functional imaging modality for measuring adult human brain activity through an acoustically transparent cranial window.
Article
BACKGROUND Alzheimer's disease (AD) is the most common cause of dementia worldwide, with apolipoprotein Eε4 (APOEε4) being the strongest genetic risk factor. Current clinical diagnostic imaging focuses on amyloid and tau; however, new methods are needed for earlier detection. METHODS PET imaging was used to assess metabolism‐perfusion in both sexes of aging C57BL/6J, and hAPOE mice, and were verified by transcriptomics, and immunopathology. RESULTS All hAPOE strains showed AD phenotype progression by 8 months, with females exhibiting the regional changes, which correlated with GO‐term enrichments for glucose metabolism, perfusion, and immunity. Uncoupling analysis revealed APOEε4/ε4 exhibited significant Type‐1 uncoupling (↓ glucose uptake, ↑ perfusion) at 8 and 12 months, while APOEε3/ε4 demonstrated Type‐2 uncoupling (↑ glucose uptake, ↓ perfusion), while immunopathology confirmed cell specific contributions. DISCUSSION This work highlights APOEε4 status in AD progression manifests as neurovascular uncoupling driven by immunological activation, and may serve as an early diagnostic biomarker. Highlights We developed a novel analytical method to analyze PET imaging of ¹⁸ F‐FDG and ⁶⁴ Cu‐PTSM data in both sexes of aging C57BL/6J, and hAPOEε3/ε3, hAPOEε4/ε4, and hAPOEε3/ε4 mice to assess metabolism‐perfusion profiles termed neurovascular uncoupling. This analysis revealed APOEε4/ε4 exhibited significant Type‐1 uncoupling (decreased glucose uptake, increased perfusion) at 8 and 12 months, while APOEε3/ε4 demonstrated significant Type‐2 uncoupling (increased glucose uptake, decreased perfusion) by 8 months which aligns with immunopathology and transcriptomic signatures. This work highlights that there may be different mechanisms underlying age related changes in APOEε4/ε4 compared with APOEε3/ε4. We predict that these changes may be driven by immunological activation and response, and may serve as an early diagnostic biomarker.
Article
Full-text available
Myalgic Encephalomyelitis/Chronic Fatigue Syndrome (ME/CFS) with orthostatic intolerance (OI) is characterized by neuro-cognitive deficits perhaps related to upright hypocapnia and loss of cerebral autoregulation (CA). We performed N-back neurocognition testing and calculated the phase synchronization index (PhSI) between Arterial Pressure (AP) and cerebral blood velocity (CB V ) as a time-dependent measurement of cerebral autoregulation in 11 control (mean age=24.1 years) and 15 ME/CFS patients (mean age=21.8 years). All ME/CFS patients had postural tachycardia syndrome (POTS). A 10-minute 60⁰ head-up tilt (HUT) significantly increased heart rate (109.4 ± 3.9 vs. 77.2 ± 1.6 beats/min, P <0.05) and respiratory rate (20.9 ± 1.7 vs. 14.2 ± 1.2 breaths/min, P < 0.05) and decreased end-tidal CO 2 (ETCO 2 ; 33.9 ± 1.1 vs. 42.8 ± 1.2 Torr, P < 0.05) in ME/CFS vs. control. In ME/CFS, HUT significantly decreased CB V compared to control (-22.5% vs -8.7%, p<0.005). To mitigate the orthostatic CB V reduction, we administered supplemental CO 2 , phenylephrine and acetazolamide and performed N-back testing supine and during HUT. Only phenylephrine corrected the orthostatic decrease in neurocognition by reverting % correct n=4 N-back during HUT in ME/CFS similar to control (ME/CFS=38.5±5.5 vs. ME/CFS+PE= 65.6±5.7 vs. Control 56.9±7.5). HUT in ME/CFS resulted in increased PhSI values indicating decreased CA. While CO 2 and Acetazolamide had no effect on PhSI in ME/CFS, PE caused a significant reduction in PhSI (ME/CFS=0.80±0.03 vs ME/CFS+PE= 0.69±0.04, p< 0.05) and improved cerebral autoregulation. Thus, PE improved neurocognitive function in ME/CFS patients, perhaps related to improved neurovascular coupling, cerebral autoregulation and maintenance of CB V .
Article
Full-text available
The prostanoid-synthesizing enzyme cyclooxygenase-2 (COX-2) is expressed in selected cerebral cortical neurons and is involved in synaptic signaling. We sought to determine whether COX-2 participates in the increase in cerebral blood flow produced by synaptic activity in the somatosensory cortex. In anesthetized mice, the vibrissae were stimulated mechanically, and cerebral blood flow was recorded in the contralateral somatosensory cortex by a laser-Doppler probe. We found that the COX-2 inhibitor NS-398 attenuates the increase in somatosensory cortex blood flow produced by vibrissal stimulation. Furthermore, the flow response was impaired in mice lacking the COX-2 gene, whereas the associated increase in whisker-barrel cortex glucose use was not affected. The increases in cerebral blood flow produced by hypercapnia, acetylcholine, or bradykinin were not attenuated by NS-398, nor did they differ between wild-type and COX-2 null mice. The findings provide evidence for a previously unrecognized role of COX-2 in the mechanisms coupling synaptic activity to neocortical blood flow and provide an insight into one of the functions of constitutive COX-2 in the CNS.
Article
Since adenosine is present in normal brain tissue and cerebrosipinal fluid and since it dilates the pial vessels, it is possible that adenosine, in addition to H-+, is also a mediator of the metabolic regulation of cerebral blood flow. Evidence supporting this hypothesis was obtained under various experimental conditions characterized by achange in brain oxygen supply. The brain was frozen in situ by means of a small bonerongeur precooled in liquid N2 and the tissue was processed for adenosine determination (nmol/g of tissue). Electrical stimulation of the cortex at 0, 15, 30, and 45 Hz yielded adenosine levels of 5.4 plus or minus 0.7, 10.5 plus or minus 1.7, 13.0 plusor minus 1.2, and 9.0 plus or minus 2.1 nmol/g. Arterial pressures of 87, 60, and 40mmHg gave adenosine levels of 7.5 plus or minus 0.76, 13 plus or minus 2.6, and 26.6plus or minus 3.3, respectively. Ventilation with 29.7, 20, 10.7, and5.5% O2 significantly increased the adenosine levels to 9.4 plus or minus 3.0, 6.4 plus or minus 1.2, 30.0 plus or minus 9.3, and 63.3 plus or minus 18.2 nmol/g, respectively. Hyperventilation significantly increased adenosine form 6.7 plus or minus 1.0 to 11.8 plus or minus 1.4 nmol/g. This increased adenosine level was reduced by additionof CO2 to the ventilating gas mixture. Lactate, the main H-+ donor, pyruvate, and cAMP changed in a fashion parallel to adenosine. However, cAMP showedonly a small increase in adenosine. These findings are in accordance with the concept that adenosine and H-+ may act synergistally to regulate cerebral blood flow and that endogenous adenosine may exert a small effect on cAMP formation.
Article
We investigated the role of nitric oxide (NO)/cGMP in the coupling of neuronal activation to regional cerebral blood flow (rCBF) in α-chloralose-anesthetized rats. Whisker deflection (60 s) increased rCBF by 18 ± 3%. NO synthase (NOS) inhibition by Nω-nitro-l-arginine (l-NNA; topically) reduced the rCBF response to 9 ± 4% and resting rCBF to 80 ± 8%. NO donors [ S-nitroso- N-acetylpenicillamine (SNAP; 50 μM), 3-morpholinosydnonimine (10 μM)] or 8-bromoguanosine 3',5'-cyclic-monophosphate (8-BrcGMP; 100 μM)] restored resting rCBF andl-NNA-induced attenuation of the whisker response in the presence ofl-NNA, whereas the NO-independent vasodilator papaverine (1 mM) had no effect on the whisker response. Basal cGMP levels were decreased to 35% byl-NNA and restored to 65% of control by subsequent SNAP superfusion. Inhibition of neuronal NOS by 7-nitroindazole (7-NI; 40 mg/kg ip) or soluble guanylyl cyclase by 1 H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one (ODQ; 100 μM) significantly reduced resting rCBF to 86 ± 8 and 92 ± 10% and whisker rCBF response to 7 ± 4 and 12 ± 3%, respectively. ODQ reduced tissue cGMP to 54%. 8-BrcGMP restored the whisker response in the presence of 7-NI or ODQ. We conclude that NO, produced by neuronal NOS, is a modulator in the coupling of neuronal activation and rCBF in rat somatosensory cortex and that this effect is mainly mediated by cGMP.l-NNA-induced vasomotion was significantly reduced during increased neuronal activity and after restoration of basal NO levels, but not after restoration of cGMP.
Book
This volume documents the proceedings of a symposium on "The Pulmonary Circulation" held at the Ettore Majorana Center for Scientific Culture, in Erice, Sicily, between 16th July and 21st July 1979. This was attended by about 200 participants drawn from Europe as a whole, but the majority were from Southern Europe. The discussion was recorded either in English or Italian and the tapes were reduced to a verbatim typescript by the Ente Nazionale Interpeti Congresso. The verbatim typescript has been edited using a few guiding principles as foliows: Titles and honorifics have been eliminated unless the 1. statement is addressed to a specific person. The style of the speakers in the discussion has been 2. preserved as far as possible and not reduced to a strictly grammatical format. 3. Where references to illustrations (e.g., on the blackboard) are made, the comments have been left unaltered and many are understandable. Removing them detracted from the sense. 4. The air of informality in the proceedings has been preserved so far as possible. 5. The responsibility for the discussion rests solely with the editors, and no contributor has had the opportunity of correcting whae he said. No manuscript was received from one participant, but the 6. discussion of his presentation has been included since it contains some points of substance. G. Cumming and G. Bonsignore v CONTENTS Functional Morphology of the Pulmonary 1 Circulation. ........ .
Article
The haemodynamic responses to neural activity that underlie the blood-oxygen-level-dependent (BOLD) signal used in functional magnetic resonance imaging (fMRI) of the brain are often assumed to be driven by energy use, particularly in presynaptic terminals or glia. However, recent work has suggested that most brain energy is used to power postsynaptic currents and action potentials rather than presynaptic or glial activity and, furthermore, that haemodynamic responses are driven by neurotransmitter-related signalling and not directly by the local energy needs of the brain. A firm understanding of the BOLD response will require investigation to be focussed on the neural signalling mechanisms controlling blood flow rather than on the locus of energy use.
Article
Although it is well established that neural activation increases local cerebral blood flow, the microvascular adjustments responsible for the flow increase have not been elucidated. We used the parallel fiber (PF) system of the cerebellar cortex to investigate the local and remote microvascular changes evoked by increased neural activity. The PF were stimulated (25-150 μA; 30 Hz) in halothane-anesthetized rats equipped with a cranial window. Arteriolar diameter was measured using a videomicroscopy system and blood flow was monitored by laser-Doppler flowmetry. The field potentials evoked by PF stimulation were recorded using microelectrodes. PF stimulation increased the diameter of local arterioles (+26±1%) in the activated folium (n=10; p<0.01). The vasodilation was greatest in smaller arterioles (16.5±0.8 μm), was graded with the intensity of stimulation and was less marked than the vasodilation produced by hypercapnia in comparably sized vessels (+58±5%; pCO2: 50-60 mmHg; n=8). The increases in vascular diameter were associated with increases in flow in the activated area (+55±4%; n=5). PF stimulation increased vascular diameter (+10±0.5%; p<0.01; n=10) also in larger arterioles (30-40 μm) whose higher order branches supplied the activated area. These vessels were located in a different folium, 2-3 mm away from the site of stimulation. No field potentials were evoked by PF stimulation in the area where these upstream vessels resided. The data suggest that increased synaptic activity in the PF system produces a "local" hemodynamic response mediated by synaptic release of vasoactive agents and a "remote" response which is propagated to upstream arterioles from vessels residing in the activated folium. These propagated vascular responses are important in the coordination of segmental vascular resistance that is required to increase flow effectively during functional brain hyperemia.
Article
We tested the hypothesis that conduction of vasomotor responses occurs in cannulated and isolated rat cerebral penetrating arterioles. Both at the site of stimulation (local) and 500-650 μm distant from it, we observed the diameter responses and time courses thereof to pressure-ejected vasoactive stimuli. ATP locally caused an initial constriction (response onset at 0.3 s, average diameter 85% of control at 450-ms pulse with a maximum at 1.6 s after stimulation) followed by a secondary dilation (111% at 7 s). Conducted vasodilation of 111% was observed over a distance of 520 μm. Prostaglandin F 2α (PGF 2α) constricted the vessels locally (80%) and caused conducted vasodilation (110%). For both ATP and PGF 2α the local constriction occurred simultaneously to the conducted vasodilation. Adenosine dilated the vessels (123%) but produced only inconsistent conducted vasodilation. Hydrogen ions initially constricted the vessels (88%) and then dilated them to 113%. Thus, although ATP and PGF 2α are strong promoters of conduction, adenosine and hydrogen ions are not. Paradoxically, ATP and PGF 2α caused conducted vasodilation even though the initial local response was a vasoconstriction, indicating that in cerebral arterioles conduction may be mediated through endothelial cell mechanisms rather than through smooth muscle cell communication.