The role of the cerebral capillaries in acute ischemic stroke: the extended penumbra model

Leif Østergaard, Sune Nørhøj Jespersen, Kim Mouridsen, Irene Klærke Mikkelsen, Kristjana Ýr Jonsdottír, Anna Tietze, Jakob Udby Blicher, Rasmus Aamand, Niels Hjort, Nina Kerting Iversen, Changsi Cai, Kristina Dupont Hougaard, Claus Z Simonsen, Paul Von Weitzel-Mudersbach, Boris Modrau, Kartheeban Nagenthiraja, Lars Riisgaard Ribe, Mikkel Bo Hansen, Susanne Lise Bekke, Martin Gervais Dahlman, Josep Puig, Salvador Pedraza, Joaquín Serena, Tae-Hee Cho, Susanne Siemonsen, Götz Thomalla, Jens Fiehler, Norbert Nighoghossian, Grethe Andersen, Leif Østergaard, Sune Nørhøj Jespersen, Kim Mouridsen, Irene Klærke Mikkelsen, Kristjana Ýr Jonsdottír, Anna Tietze, Jakob Udby Blicher, Rasmus Aamand, Niels Hjort, Nina Kerting Iversen, Changsi Cai, Kristina Dupont Hougaard, Claus Z Simonsen, Paul Von Weitzel-Mudersbach, Boris Modrau, Kartheeban Nagenthiraja, Lars Riisgaard Ribe, Mikkel Bo Hansen, Susanne Lise Bekke, Martin Gervais Dahlman, Josep Puig, Salvador Pedraza, Joaquín Serena, Tae-Hee Cho, Susanne Siemonsen, Götz Thomalla, Jens Fiehler, Norbert Nighoghossian, Grethe Andersen

Abstract

The pathophysiology of cerebral ischemia is traditionally understood in relation to reductions in cerebral blood flow (CBF). However, a recent reanalysis of the flow-diffusion equation shows that increased capillary transit time heterogeneity (CTTH) can reduce the oxygen extraction efficacy in brain tissue for a given CBF. Changes in capillary morphology are typical of conditions predisposing to stroke and of experimental ischemia. Changes in capillary flow patterns have been observed by direct microscopy in animal models of ischemia and by indirect methods in humans stroke, but their metabolic significance remain unclear. We modeled the effects of progressive increases in CTTH on the way in which brain tissue can secure sufficient oxygen to meet its metabolic needs. Our analysis predicts that as CTTH increases, CBF responses to functional activation and to vasodilators must be suppressed to maintain sufficient tissue oxygenation. Reductions in CBF, increases in CTTH, and combinations thereof can seemingly trigger a critical lack of oxygen in brain tissue, and the restoration of capillary perfusion patterns therefore appears to be crucial for the restoration of the tissue oxygenation after ischemic episodes. In this review, we discuss the possible implications of these findings for the prevention, diagnosis, and treatment of acute stroke.

Figures

Figure 1
Figure 1
Vascular changes in conditions that predispose to stroke. The figure illustrates the three levels at which the vascular system is affected in conditions predisposing to stroke. The schematic drawing is modified from a figure by Lanza and Crea. At the level of conduit vessels (A), blood flow can be restricted by atherosclerotic lesions (red arrow). At the level of resistance vessels (B), small vessel disease causes thickening of the vessel wall (red arrows). More peripherally, the capillary wall undergoes morphological changes, such as thickened basement membrane, pericapillary fibrosis, and pericyte loss in aging and a number of conditions predisposing to stroke (A, right)—see also Table 1. Images from Farkas et al. Ischemia can induce irreversible constrictions of capillary pericytes (B, right) and thereby disturb capillary flow patterns—from Yemisci et al.
Figure 2
Figure 2
Classical Bohr–Kety–Crone–Renkin (BKCR) flow-diffusion relation for oxygen. The classical BKCR curve shows the maximum amount of oxygen that can diffuse from a single capillary into tissue, for a given flow CMRO2max. The curve shape predicts three important properties of CMRO2max parallel-coupled capillaries: first, the curve slope decreases toward high perfusion values, making vasodilation increasingly inefficient as a means of improving tissue oxygenation toward high perfusion rates. This property is reflected at the macroscopic level in that CBF increases are generally several-fold larger than the corresponding increase in oxygen utilization. The resulting blood ‘hyperoxygenation' is thought to explain the blood oxygen level-dependent (BOLD) contrast mechanism. Second, if erythrocyte flows differ among capillary paths (case B) instead of being equal (case A), the net tissue oxygen availability declines. This is seen by using the BKCR curve to determine the net tissue oxygen availability resulting from the individual flows in case (B). The resulting net tissue oxygen availability is the weighted average of the oxygen availabilities for the two flows, labeled B in the plot. Note that the resulting tissue oxygen availability will always be less than that of the homogenous case, labeled A. Conversely, homogenization of capillary flows during hyperemia has the opposite effect, and serves to compensate for the first property. Third, if erythrocyte flows are hindered (rather than continuously redistributed) along single capillary paths (as indicated by slow-passing white blood cell (WBC) and rugged capillary walls) upstream vasodilation is likely to amplify the redistribution losses, as erythrocytes are forced through other branches at very high speeds, with negligible net oxygenation gains. CBF, cerebral blood flow; CMRO2max, cerebral metabolic rate of oxygen.
Figure 3
Figure 3
Changes in cerebral blood flow (CBF) and tissue oxygen tension that must accompany increasing levels of capillary dysfunction to maintain tissue oxygen availability before stroke. The figure displays the adaptations of CBF (second panel from the top) and PtO2 (third panel from the top) that are necessary to maintain tissue oxygen availability as capillary transit time heterogeneity (CTTH) levels (top panel) gradually increase, both during rest (upper graph in each panel) and during functional activation (lower graph in each panel). These changes are hypothesized to occur in relation to conditions that predispose to stroke—cf. Table 1—years before symptoms develop. The most important hemodynamic change occurs toward the end of Stage 1 when the increase in tissue oxygen availability during hyperemia is no longer sufficient to meet the metabolic needs of the tissue. Therefore, vasodilation during episodes of increased metabolic demands must be attenuated. This transition is hypothesized to mark the onset of neurovascular dysfunction, and of increased oxidative damage due to the accompanying release of reactive oxygen species (ROS). Later, vasodilation induced during rest must also be attenuated to maintain tissue oxygen availability above the needs of resting tissue. This is hypothesized to mark the onset of detectable reductions in cerebrovascular reserve capacity. As CTTH continues to increase, oxygen availability can be secured by attenuated CBF responses and more efficient blood-tissue concentration gradients—see text. This results in increasing maximum oxygen extraction fraction (OEFmax) values. The gradual reduction in metabolic reserve capacity is indicated in the lower panel: As CTTH increases, net oxygen extraction is increasingly limited by CBF as the oxygen extraction fraction approaches unity. In the stroke prone state, minor reductions in CBF or increases in CTTH are therefore predicted to result in neurologic symptoms as maximum cerebral metabolic rate of oxygen (CMRO2max) approaches the actual, metabolic needs of the tissue. HIF-1, hypoxia-inducible transcription factor 1.
Figure 4
Figure 4
Effects of mean transit time (MTT), capillary transit time heterogeneity (CTTH), and oxygen tension on oxygen extraction. (A, B) The x axis position is indicated by the value of both MTT (bottom x axis) and cerebral blood flow (CBF) (top x axis). (A) Contour plot of maximum oxygen extraction fraction (OEFmax) as a function of these parameters and the CTTH at a tissue oxygen tension of PtO2=26 mm Hg. The value of OEFmax is indicated by a color at the corresponding location in the (MTT, CTTH) plane. The OEFmax value corresponding to this color or location in the contour plot is most easily derived from the OEFmax values indicated on the two nearest solid black lines, which display OEFmax iso-contours. Note that the oxygen extraction efficacy always increases with increasing MTT and with decreasing CTTH. The image inserts show schematic capillary beds with homogenous capillary transit times (lower left insert) and with CTTH typical of resting brain (upper right insert). The arrow indicates the changes in MTT and CTTH that occur during functional activation: The reduction in CTTH that occurs in parallel with the arteriolar dilation partly maintains the oxygen extraction efficacy during hyperemia. The maximum supported cerebral metabolic rate of oxygen (CMRO2max) is shown in units of mL/100 mL per minute in (B). Recall that this figure is derived from (A) by multiplying its OEFmax values by Ca and CBF, assuming that the capillary transit times of erythrocytes are inversely proportional to CBV. Note that CMRO2max, and thereby tissue oxygen availability, always increases with decreasing flow heterogeneity. The physiologic CBF response to functional activation tends to reduce OEFmax in normal tissue, both due to the shorter MTT and due to a higher PtO2. Hemodynamic states above the yellow line in (B) are unique in that increases in their CBF (reductions in MTT) will lead to states of lower tissue oxygen availability: these states are referred to as having malignant CTTH. The horizontal arrow indicates the metabolic consequence of increased CBF without concomitant reductions in CTTH: in the absence of any homogenization of capillary flows, vasodilation would not lead to a net increase in the availability of oxygen in tissue. The origin of this paradox effect is illustrated by the insert in the upper right of (A). The oxygenation of blood along capillary paths in this insert was modeled according to in vivo erythrocyte velocity recordings in resting rat brain, and the single-capillary Bohr–Kety–Crone–Renkin (BKCR) model, and then indicated as a color code, ranging from fully oxygenated (left) to levels well below that of mixed venular blood (right) along some capillary paths—cf. also Figure 3 in Jespersen and Østergaard. Note that, along some capillary paths, flow velocities are so high that little oxygen is extracted from blood. In normal brain, hyperemia is accompanied by a reduction in CTTH (lower left insert in A). This homogenization decreases the number of capillary paths with high flow and limited oxygen extraction, causing net oxygen availability to increase as a function of CBF. If this heterogeneity persist during hyperemia, for example due to changes in capillary morphology or blood rheology, the effective shunting of oxygenated blood can become critical. As discussed in the text, vasodilation may in fact be attenuated by hypoxia-sensitive mechanisms as hemodynamic states approach this critical limit. Paradoxically, the neurovascular dysfunction and the reduced CVRC observed in conditions predisposing to stroke may therefore reflect adaptations to prevent critical hypoxia in cases of high CTTH. (C) This panel shows how tissue oxygen metabolism can be secured when CBF increases no longer lead to increased tissue oxygen availability. The figure shows net oxygenation as a function of tissue oxygen tension and CTTH for fixed MTT (CBF=60 mL/100 mL per minute; MTT 1.4 seconds) to illustrate how tissue metabolism, by reducing tissue oxygen tension, can facilitate net oxygen extraction in cases where capillary dysfunction cause critical levels of ‘shunting'—provided that CBF responses are attenuated. Note that a decrease in oxygen tension of 5 mm Hg can support a CMRO2 increase of 20%. This corresponds roughly to the additional energy requirements of neuronal firing.
Figure 5
Figure 5
Metabolic thresholds. The green iso-contour surface corresponds to the metabolic rate of contralateral tissue in patients with focal ischemia. The red plane marks the boundary, left of which vasodilation reduces tissue oxygen availability (malignant capillary transit time heterogeneity (CTTH)). The maximum value CTTH can attain at a PtO2 of 25 mm Hg if oxygen availability is to remain above that of resting tissue is indicated by the label A. As CTTH increases further, a critical limit is reached as PtO2 approaches zero—label B. At this stage, the metabolic needs of tissue cannot be supported unless mean transit time (MTT) is prolonged to a threshold of ∼4 seconds, corresponding to cerebral blood flow (CBF)=21 mL/100 mL per minute. If MTT increases instead, owing to limitations in blood supply, then the threshold corresponding to resting contralateral tissue is reached at MTT values of roughly 6.3 seconds, or CBF=13 mL/100 mL per minute, provided CTTH is negligible or moderate (label C). If, however, CTTH increases as a result of the reduction in CBF or because of per-ischemic changes in capillary patency (cf. Figure 1), then tissue oxygen availability cannot be maintained above the metabolic needs of resting tissue for CBF values below 21 mL/100 mL per minute. The blue arrows indicate how tissue hypoperfusion and capillary dysfunction causes MTT and CTTH to change, and tissue oxygen availability to approach the metabolic requirements of resting brain tissue (the green iso-contour).
Figure 6
Figure 6
Digital phantom simulation of the inherent bias of the Tmax parameter by the values of mean transit time (MTT) and capillary transit time heterogeneity (CTTH). (A) Forty-nine squares with ‘true' maximum oxygen extraction fraction (OEFmax) values corresponding to combinations of seven MTT values and seven CTTH values, cf. Figure 4A, are shown. Each square consists of 14-by-14 voxels, with associated concentration time curves characteristic of the tissue bolus passage which would result from the vascular retention caused by the corresponding values of MTT and CTTH. Random noise characteristic of magnetic resonance perfusion raw data was then added at each noise free, simulated data point, and the tissue curve analyzed by a customized algorithm, which determine OEFmax based on CTTH and MTT. (B) This panel shows OEFmax values estimated by this algorithm. They are seen to correspond well with the true values in panel A. We also determined the corresponding Tmax value using two common singular value decomposition based deconvolution techniques., (C, D) Maps of Tmax derived from the delay-corrected (C) and delay un-corrected (D) singular value decomposition algorithms. Note that the Tmax maps contain an incidental bias that mimics OEFmax.

References

    1. Moustafa RS, Baron JC.Perfusion Thresholds in Cerebral IschemiaIn: Donnan GA, Baron JC, Davis SM, Sharp FR, (eds).. The Ischemic Penumbra Informa Healthcare USA, Inc.: New York, USA; 200731–36.
    1. Astrup J, Symon L, Branston NM, Lassen NA. Cortical evoked potential and extracellular K+ and H+ at critical levels of brain ischemia. Stroke. 1977;8:51–57.
    1. Astrup J, Siesjo BK, Symon L. Thresholds in cerebral ischemia—the ischemic penumbra. Stroke. 1981;12:723–725.
    1. Donnan GA, Baron JC, Davis SM, Sharp FR.The ischemic penumbra: overview, definition, and criteriaIn: Donnan GA, Baron JC, Davis SM, Sharp FR, (eds).. The Ischemic Penumbra Informa Healthcare USA, Inc.: New York; 20077–20.
    1. Albers GW, Thijs VN, Wechsler L, Kemp S, Schlaug G, Skalabrin E, et al. Magnetic resonance imaging profiles predict clinical response to early reperfusion: the diffusion and perfusion imaging evaluation for understanding stroke evolution (DEFUSE) study. Ann Neurol. 2006;60:508–517.
    1. Davis SM, Donnan GA, Parsons MW, Levi C, Butcher KS, Peeters A, et al. Effects of alteplase beyond 3 hour after stroke in the Echoplanar Imaging Thrombolytic Evaluation Trial (EPITHET): a placebo-controlled randomised trial. Lancet Neurol. 2008;7:299–309.
    1. Nagakane Y, Christensen S, Brekenfeld C, Ma H, Churilov L, Parsons MW, et al. EPITHET: positive result after reanalysis using baseline diffusion-weighted imaging/perfusion-weighted imaging co-registration. Stroke. 2011;42:59–64.
    1. Mishra NK, Albers GW, Davis SM, Donnan GA, Furlan AJ, Hacke W, et al. Mismatch-based delayed thrombolysis: a meta-analysis. Stroke. 2010;41:e25–e33.
    1. Amarenco P, Bogousslavsky J, Caplan LR, Donnan GA, Hennerici MG. New approach to stroke subtyping: the A-S-C-O (phenotypic) classification of stroke. Cerebrovasc Dis. 2009;27:502–508.
    1. Moskowitz MA, Lo EH, Iadecola C. The science of stroke: mechanisms in search of treatments. Neuron. 2010;67:181–198.
    1. Iadecola C, Davisson RL. Hypertension and cerebrovascular dysfunction. Cell Metab. 2008;7:476–484.
    1. Dawson VL, Dawson TM. Nitric oxide neurotoxicity. J Chem Neuroanat. 1996;10:179–190.
    1. Pacher P, Beckman JS, Liaudet L. Nitric oxide and peroxynitrite in health and disease. Physiol Rev. 2007;87:315–424.
    1. Shah K, Qureshi SU, Johnson M, Parikh N, Schulz PE, Kunik ME. Does use of antihypertensive drugs affect the incidence or progression of dementia? A systematic review. Am J Geriatr Pharmacother. 2009;7:250–261.
    1. Schulz E, Jansen T, Wenzel P, Daiber A, Munzel T. Nitric oxide, tetrahydrobiopterin, oxidative stress, and endothelial dysfunction in hypertension. Antioxid Redox Signal. 2008;10:1115–1126.
    1. Jackson SP. Arterial thrombosis-insidious, unpredictable and deadly. Nat Med. 2011;17:1423–1436.
    1. Gursoy-Ozdemir Y, Yemisci M, Dalkara T. Microvascular protection is essential for successful neuroprotection in stroke. J Neurochem. 2012;123 (Suppl 2:2–11.
    1. Yemisci M, Gursoy-Ozdemir Y, Vural A, Can A, Topalkara K, Dalkara T. Pericyte contraction induced by oxidative-nitrative stress impairs capillary reflow despite successful opening of an occluded cerebral artery. Nat Med. 2009;15:1031–1037.
    1. Dalkara T, Arsava EM. Can restoring incomplete microcirculatory reperfusion improve stroke outcome after thrombolysis. J Cereb Blood Flow Metab. 2012;32:2091–2099.
    1. Kuschinsky W, Paulson OB. Capillary circulation in the brain. Cerebrovasc Brain Metab Rev. 1992;4:261–286.
    1. Tomita Y, Tomita M, Schiszler I, Amano T, Tanahashi N, Kobari M, et al. Moment analysis of microflow histogram in focal ischemic lesion to evaluate microvascular derangement after small pial arterial occlusion in rats. J Cereb Blood Flow Metab. 2002;22:663–669.
    1. Hudetz AG, Feher G, Weigle CG, Knuese DE, Kampine JP. Video microscopy of cerebrocortical capillary flow: response to hypotension and intracranial hypertension. Am J Physiol. 1995;268:H2202–H2210.
    1. Østergaard L, Sorensen AG, Chesler DA, Weisskoff RM, Koroshetz WJ, Wu O, et al. Combined diffusion-weighted and perfusion-weighted flow heterogeneity magnetic resonance imaging in acute stroke. Stroke. 2000;31:1097–1103.
    1. Simonsen CZ, Rohl L, Vestergaard-Poulsen P, Gyldensted C, Andersen G, Østergaard L. Final infarct size after acute stroke: prediction with flow heterogeneity. Radiology. 2002;225:269–275.
    1. Perkio J, Soinne L, Østergaard L, Helenius J, Kangasmaki A, Martinkauppi S, et al. Abnormal intravoxel cerebral blood flow heterogeneity in human ischemic stroke determined by dynamic susceptibility contrast magnetic resonance imaging. Stroke. 2005;36:44–49.
    1. Jespersen SN, Østergaard L. The roles of cerebral blood flow, capillary transit time heterogeneity and oxygen tension in brain oxygenation and metabolism. J Cereb Blood Flow Metab. 2012;32:264–277.
    1. Renkin EM. B. W. Zweifach Award lecture. Regulation of the microcirculation. Microvasc Res. 1985;30:251–263.
    1. Pawlik G, Rackl A, Bing RJ. Quantitative capillary topography and blood flow in the cerebral cortex of cats: an in vivo microscopic study. Brain Res. 1981;208:35–58.
    1. Villringer A, Them A, Lindauer U, Einhaupl K, Dirnagl U. Capillary perfusion of the rat brain cortex. An in vivo confocal microscopy study. Circ Res. 1994;75:55–62.
    1. Kleinfeld D, Mitra PP, Helmchen F, Denk W. Fluctuations and stimulus-induced changes in blood flow observed in individual capillaries in layers 2 through 4 of rat neocortex. Proc Natl Acad Sci USA. 1998;95:15741–15746.
    1. Mazzoni MC, Schmid-Schonbein GW. Mechanisms and consequences of cell activation in the microcirculation. Cardiovasc Res. 1996;32:709–719.
    1. Lipowsky HH, Gao L, Lescanic A. Shedding of the endothelial glycocalyx in arterioles, capillaries and venules and its effect on capillary hemodynamics during inflammation. Am J Physiol Heart Circ Physiol. 2011;30:H2235–H2245.
    1. King RB, Raymond GM, Bassingthwaighte JB. Modeling blood flow heterogeneity. Ann Biomed Eng. 1996;24:352–372.
    1. Stewart GN. Researches on the circulation time in organs and on the influences which affect it. Parts I-III. J Physiol. 1894;15:1–89.
    1. Hayashi T, Watabe H, Kudomi N, Kim KM, Enmi J, Hayashida K, et al. A theoretical model of oxygen delivery and metabolism for physiologic interpretation of quantitative cerebral blood flow and metabolic rate of oxygen. J Cereb Blood Flow Metab. 2003;23:1314–1323.
    1. Mintun MA, Lundstrom BN, Snyder AZ, Vlassenko AG, Shulman GL, Raichle ME. Blood flow and oxygen delivery to human brain during functional activity: theoretical modeling and experimental data. Proc Natl Acad Sci USA. 2001;98:6859–6864.
    1. Crone C. The permeability of capillaries in various organs as determined by use of the ‘Indicator Diffusion' Method. Acta Physiol Scand. 1963;58:292–305.
    1. Arfken GB, Weber HJ. Mathematical Methods for Physicists, Sixth Edition: A Comprehensive Guide. Elsevier Academic Press: Boston, USA; 2005. p. 1182.
    1. Stefanovic B, Hutchinson E, Yakovleva V, Schram V, Russell JT, Belluscio L, et al. Functional reactivity of cerebral capillaries. J Cereb Blood Flow Metab. 2008;28:961–972.
    1. Schulte ML, Wood JD, Hudetz AG. Cortical electrical stimulation alters erythrocyte perfusion pattern in the cerebral capillary network of the rat. Brain Res. 2003;963:81–92.
    1. Hudetz AG, Biswal BB, Feher G, Kampine JP. Effects of hypoxia and hypercapnia on capillary flow velocity in the rat cerebral cortex. Microvasc Res. 1997;54:35–42.
    1. Vorstrup S, Henriksen L, Paulson OB. Effect of acetazolamide on cerebral blood flow and cerebral metabolic rate for oxygen. J Clin Invest. 1984;74:1634–1639.
    1. Powers WJ. Cerebral hemodynamics in ischemic cerebrovascular disease. Ann Neurol. 1991;29:231–240.
    1. Maeda H, Matsumoto M, Handa N, Hougaku H, Ogawa S, Itoh T, et al. Reactivity of cerebral blood flow to carbon dioxide in hypertensive patients: evaluation by the transcranial Doppler method. J Hypertens. 1994;12:191–197.
    1. Lavi S, Gaitini D, Milloul V, Jacob G. Impaired cerebral CO2 vasoreactivity: association with endothelial dysfunction. Am J Physiol Heart Circ Physiol. 2006;291:H1856–H1861.
    1. Sette G, Baron JC, Mazoyer B, Levasseur M, Pappata S, Crouzel C. Local brain haemodynamics and oxygen metabolism in cerebrovascular disease. Positron emission tomography. Brain. 1989;112 (Pt 4:931–951.
    1. Smith EE, Schneider JA, Wardlaw JM, Greenberg SM. Cerebral microinfarcts: the invisible lesions. Lancet Neurol. 2012;11:272–282.
    1. Smith WS, Lev MH, English JD, Camargo EC, Chou M, Johnston SC, et al. Significance of large vessel intracranial occlusion causing acute ischemic stroke and TIA. Stroke. 2009;40:3834–3840.
    1. Purroy F, Begue R, Quilez A, Pinol-Ripoll G, Sanahuja J, Brieva L, et al. The California, ABCD, and unified ABCD2 risk scores and the presence of acute ischemic lesions on diffusion-weighted imaging in TIA patients. Stroke. 2009;40:2229–2232.
    1. Elkind MS. Why now? Moving from stroke risk factors to stroke triggers. Curr Opin Neurol. 2007;20:51–57.
    1. Macko RF, Ameriso SF, Barndt R, Clough W, Weiner JM, Fisher M. Precipitants of brain infarction. Roles of preceding infection/inflammation and recent psychological stress. Stroke. 1996;27:1999–2004.
    1. Grau AJ, Buggle F, Becher H, Zimmermann E, Spiel M, Fent T, et al. Recent bacterial and viral infection is a risk factor for cerebrovascular ischemia: clinical and biochemical studies. Neurology. 1998;50:196–203.
    1. Zurru MC, Alonzo C, Brescacin L, Romano M, Camera LA, Waisman G, et al. Recent respiratory infection predicts atherothrombotic stroke: case-control study in a Buenos Aires healthcare system. Stroke. 2009;40:1986–1990.
    1. Sheth T, Nair C, Muller J, Yusuf S. Increased winter mortality from acute myocardial infarction and stroke: the effect of age. J Am Coll Cardiol. 1999;33:1916–1919.
    1. Woodhouse PR, Khaw KT, Plummer M, Foley A, Meade TW. Seasonal variations of plasma fibrinogen and factor VII activity in the elderly: winter infections and death from cardiovascular disease. Lancet. 1994;343:435–439.
    1. Lavallee P, Perchaud V, Gautier-Bertrand M, Grabli D, Amarenco P. Association between influenza vaccination and reduced risk of brain infarction. Stroke. 2002;33:513–518.
    1. Nichol KL, Nordin J, Mullooly J, Lask R, Fillbrandt K, Iwane M. Influenza vaccination and reduction in hospitalizations for cardiac disease and stroke among the elderly. N Engl J Med. 2003;348:1322–1332.
    1. Grau AJ, Fischer B, Barth C, Ling P, Lichy C, Buggle F. Influenza vaccination is associated with a reduced risk of stroke. Stroke. 2005;36:1501–1506.
    1. Elkind MS. Inflammatory mechanisms of stroke. Stroke. 2010;41:S3–S8.
    1. Peppiatt CM, Howarth C, Mobbs P, Attwell D. Bidirectional control of CNS capillary diameter by pericytes. Nature. 2006;443:700–704.
    1. Kwon I, Kim EH, del Zoppo GJ, Heo JH. Ultrastructural and temporal changes of the microvascular basement membrane and astrocyte interface following focal cerebral ischemia. J Neurosci Res. 2009;87:668–676.
    1. Rowat A, Graham C, Dennis M. Dehydration in hospital-admitted stroke patients: detection, frequency, and association. Stroke. 2012;43:857–859.
    1. Yang GY, Betz AL. Reperfusion-induced injury to the blood-brain barrier after middle cerebral artery occlusion in rats. Stroke. 1994;25:1658–1664.
    1. Aronowski J, Strong R, Grotta JC. Reperfusion injury: demonstration of brain damage produced by reperfusion after transient focal ischemia in rats. J Cereb Blood Flow Metab. 1997;17:1048–1056.
    1. Dirnagl U, Iadecola C, Moskowitz MA. Pathobiology of ischaemic stroke: an integrated view. Trends Neurosci. 1999;22:391–397.
    1. Lassen NA. The luxury-perfusion syndrome and its possible relation to acute metabolic acidosis localised within the brain. Lancet. 1966;2:1113–1115.
    1. Kidwell CS, Saver JL, Mattiello J, Starkman S, Vinuela F, Duckwiler G, et al. Diffusion-perfusion MRI characterization of post-recanalization hyperperfusion in humans. Neurology. 2001;57:2015–2021.
    1. Zhao H. Ischemic postconditioning as a novel avenue to protect against brain injury after stroke. J Cereb Blood Flow Metab. 2009;29:873–885.
    1. Girouard H, Iadecola C. Neurovascular coupling in the normal brain and in hypertension, stroke, and Alzheimer disease. J Appl Physiol. 2006;100:328–335.
    1. Kazama K, Anrather J, Zhou P, Girouard H, Frys K, Milner TA, et al. Angiotensin II impairs neurovascular coupling in neocortex through NADPH oxidase-derived radicals. Circ Res. 2004;95:1019–1026.
    1. Girouard H, Park L, Anrather J, Zhou P, Iadecola C. Cerebrovascular nitrosative stress mediates neurovascular and endothelial dysfunction induced by angiotensin II. Arterioscler Thromb Vasc Biol. 2007;27:303–309.
    1. Capone C, Faraco G, Park L, Cao X, Davisson RL, Iadecola C. The cerebrovascular dysfunction induced by slow pressor doses of angiotensin II precedes the development of hypertension. Am J Physiol Heart Circ Physiol. 2011;300:H397–H407.
    1. Matsugi T, Chen Q, Anderson DR. Contractile responses of cultured bovine retinal pericytes to angiotensin II. Arch Ophthalmol. 1997;115:1281–1285.
    1. Kawamura H, Kobayashi M, Li Q, Yamanishi S, Katsumura K, Minami M, et al. Effects of angiotensin II on the pericyte-containing microvasculature of the rat retina. J Physiol. 2004;561:671–683.
    1. Yuan G, Khan SA, Luo W, Nanduri J, Semenza GL, Prabhakar NR. Hypoxia-inducible factor 1 mediates increased expression of NADPH oxidase-2 in response to intermittent hypoxia. J Cell Physiol. 2011;226:2925–2933.
    1. Sorce S, Krause KH. NOX enzymes in the central nervous system: from signaling to disease. Antioxid Redox Signal. 2009;11:2481–2504.
    1. Diomedi M, Placidi F, Cupini LM, Bernardi G, Silvestrini M. Cerebral hemodynamic changes in sleep apnea syndrome and effect of continuous positive airway pressure treatment. Neurology. 1998;51:1051–1056.
    1. Marin JM, Carrizo SJ, Vicente E, Agusti AG. Long-term cardiovascular outcomes in men with obstructive sleep apnoea-hypopnoea with or without treatment with continuous positive airway pressure: an observational study. Lancet. 2005;365:1046–1053.
    1. Haefliger IO, Zschauer A, Anderson DR. Relaxation of retinal pericyte contractile tone through the nitric oxide-cyclic guanosine monophosphate pathway. Invest Ophthalmol Vis Sci. 1994;35:991–997.
    1. Haefliger IO, Anderson DR. Oxygen modulation of guanylate cyclase-mediated retinal pericyte relaxations with 3-morpholino-sydnonimine and atrial natriuretic peptide. Invest Ophthalmol Vis Sci. 1997;38:1563–1568.
    1. Attwell D, Buchan AM, Charpak S, Lauritzen M, Macvicar BA, Newman EA. Glial and neuronal control of brain blood flow. Nature. 2010;468:232–243.
    1. Jensen FB. The dual roles of red blood cells in tissue oxygen delivery: oxygen carriers and regulators of local blood flow. J Exp Biol. 2009;212:3387–3393.
    1. Eltzschig HK, Carmeliet P. Hypoxia and inflammation. N Engl J Med. 2011;364:656–665.
    1. van Faassen EE, Bahrami S, Feelisch M, Hogg N, Kelm M, Kim-Shapiro DB, et al. Nitrite as regulator of hypoxic signaling in mammalian physiology. Med Res Rev. 2009;29:683–741.
    1. Jung KH, Chu K, Ko SY, Lee ST, Sinn DI, Park DK, et al. Early intravenous infusion of sodium nitrite protects brain against in vivo ischemia-reperfusion injury. Stroke. 2006;37:2744–2750.
    1. Terpolilli NA, Moskowitz MA, Plesnila N. Nitric oxide: considerations for the treatment of ischemic stroke. J Cereb Blood Flow Metab. 2012;32:1332–1346.
    1. Dirnagl U, Becker K, Meisel A. Preconditioning and tolerance against cerebral ischaemia: from experimental strategies to clinical use. Lancet Neurol. 2009;8:398–412.
    1. Powers WJ, Clarke WR, Grubb RL, Videen TO, Adams HP, Derdeyn CP, et al. Extracranial-intracranial bypass surgery for stroke prevention in hemodynamic cerebral ischemia: the Carotid Occlusion Surgery Study randomized trial. JAMA. 2011;306:1983–1992.
    1. Fiebach JB, Al-Rawi Y, Wintermark M, Furlan AJ, Rowley HA, Lindsten A, et al. Vascular occlusion enables selecting acute ischemic stroke patients for treatment with desmoteplase. Stroke. 2012;43:1561–1566.
    1. De Silva DA, Churilov L, Olivot JM, Christensen S, Lansberg MG, Mlynash M, et al. Greater effect of stroke thrombolysis in the presence of arterial obstruction. Ann Neurol. 2011;70:601–605.
    1. Mouridsen K, Friston K, Hjort N, Gyldensted L, Østergaard L, Kiebel S. Bayesian estimation of cerebral perfusion using a physiological model of microvasculature. Neuroimage. 2006;33:570–579.
    1. Mouridsen K, Østergaard L, Christensen S, Jespersen SN. Proceedings of the International Society for Magnetic Resonance in Medicines 19th Annual Meeting and Exhibition. Montréal: Canada; 2011. Reliable estimation of capillary transit time distributions at voxel-level using DSC-MRI; p. 3915.
    1. Butcher KS, Parsons M, MacGregor L, Barber PA, Chalk J, Bladin C, et al. Refining the perfusion-diffusion mismatch hypothesis. Stroke. 2005;36:1153–1159.
    1. Christensen S, Mouridsen K, Wu O, Hjort N, Karstoft H, Thomalla G, et al. Comparison of 10 perfusion MRI parameters in 97 sub-6-hour stroke patients using voxel-based receiver operating characteristics analysis. Stroke. 2009;40:2055–2061.
    1. Takasawa M, Jones PS, Guadagno JV, Christensen S, Fryer TD, Harding S, et al. How reliable is perfusion MR in acute stroke? Validation and determination of the penumbra threshold against quantitative PET. Stroke. 2008;39:870–877.
    1. Olivot JM, Mlynash M, Thijs VN, Kemp S, Lansberg MG, Wechsler L, et al. Optimal Tmax threshold for predicting penumbral tissue in acute stroke. Stroke. 2009;40:469–475.
    1. Calamante F, Christensen S, Desmond PM, Østergaard L, Davis SM, Connelly A. The physiological significance of the time-to-maximum (Tmax) parameter in perfusion MRI. Stroke. 2010;41:1169–1174.
    1. Østergaard L, Aamand R, Gutierrez-Jimenez E, Ho Y-L, Blicher JU, Madsen SM, et al. The capillary dysfunction hypothesis of Alzheimer's disease. Neurobiol Aging. 2013;34:1018–1031.
    1. Elkins JS, Knopman DS, Yaffe K, Johnston SC. Cognitive function predicts first-time stroke and heart disease. Neurology. 2005;64:1750–1755.
    1. Wang Q, Capistrant BD, Ehntholt A, Glymour MM. Long-term rate of change in memory functioning before and after stroke onset. Stroke. 2012;43:2561–2566.
    1. Bell MA, Ball MJ. Morphometric comparison of hippocampal microvasculature in ageing and demented people: diameters and densities. Acta Neuropathol. 1981;53:299–318.
    1. Kalaria RN. Cerebral vessels in ageing and Alzheimer's disease. Pharmacol Ther. 1996;72:193–214.
    1. Tagami M, Nara Y, Kubota A, Fujino H, Yamori Y. Ultrastructural changes in cerebral pericytes and astrocytes of stroke-prone spontaneously hypertensive rats. Stroke. 1990;21:1064–1071.
    1. Junker U, Jaggi C, Bestetti G, Rossi GL. Basement membrane of hypothalamus and cortex capillaries from normotensive and spontaneously hypertensive rats with streptozotocin-induced diabetes. Acta Neuropathol. 1985;65:202–208.
    1. McCuskey PA, McCuskey RS. In vivo and electron microscopic study of the development of cerebral diabetic microangiography. Microcirc Endothelium Lymphatics. 1984;1:221–244.
    1. Johnson PC, Brendel K, Meezan E. Thickened cerebral cortical capillary basement membranes in diabetics. Arch Pathol Lab Med. 1982;106:214–217.
    1. Reske-Nielsen E, Lundbæk K, Rafaelsen OJ. Pathological changes in the central and peripheral nervous system of young long-term diabetics. I. Diabetic encephalopathy. Diabetologia. 1965;1:233–241.
    1. Mayhan WG, Sharpe GM. Nicotine impairs histamine-induced increases in macromolecular efflux: role of oxygen radicals. J Appl Physiol. 1998;84:1589–1595.
    1. Karwacka H. Ultrastructural and biochemical studies of the brain and other organs in rats after chronic ethanol administration. I. Electronmicroscopic investigations of the morphologic elements of the blood-brain barrier in the rat after ethanol intoxication. Exp Pathol (Jena) 1980;18:118–126.
    1. Dalkara T, Gursoy-Ozdemir Y, Yemisci M. Brain microvascular pericytes in health and disease. Acta Neuropathol. 2011;122:1–9.
    1. Lanza GA, Crea F. Primary coronary microvascular dysfunction: clinical presentation, pathophysiology, and management. Circulation. 2010;121:2317–2325.
    1. Farkas E, de Vos RA, Donka G, Jansen Steur EN, Mihaly A, Luiten PG. Age-related microvascular degeneration in the human cerebral periventricular white matter. Acta Neuropathol. 2006;111:150–157.
    1. Fox PT, Raichle ME. Focal physiological uncoupling of cerebral blood flow and oxidative metabolism during somatosensory stimulation in human subjects. Proc Natl Acad Sci USA. 1986;83:1140–1144.
    1. Raichle ME, Mintun MA. Brain work and brain imaging. Annu Rev Neurosci. 2006;29:449–476.
    1. Ndubuizu O, LaManna JC. Brain tissue oxygen concentration measurements. Antioxid Redox Signal. 2007;9:1207–1219.
    1. Wu O, Østergaard L, Weisskoff RM, Benner T, Rosen BR, Sorensen AG. Tracer arrival timing-insensitive technique for estimating flow in MR perfusion-weighted imaging using singular value decomposition with a block-circulant deconvolution matrix. Magn Reson Med. 2003;50:164–174.
    1. Østergaard L, Weisskoff RM, Chesler DA, Gyldensted C, Rosen BR. High resolution measurement of cerebral blood flow using intravascular tracer bolus passages. Part I: Mathematical approach and statistical analysis. Magn Reson Med. 1996;36:715–725.

Source: PubMed

3
S'abonner