Skip to main content

Treatment of Acute Ischaemic Stroke

  • Chapter
  • First Online:
  • 3353 Accesses

Part of the book series: Contemporary Medical Imaging ((CMI,volume 1))

Abstract

The elements of successful treatment for acute stroke are: (1) Rapid evaluation and decision making, (2) Careful patient selection, and (3) Rapid and effective pharmacologic or mechanical thrombolysis. Approach any potential candidate for endovascular therapy with an eye towards first administering IV t-PA, or getting the patient on the angio suite table as soon as possible. The overall strategy consists of the following steps: make the correct diagnosis and focus on the patient examination.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

References

  1. The NINDS t-PA Stroke Study Group. Generalized efficacy of t-PA for acute stroke. Subgroup analysis of the NINDS t-PA Stroke Trial. Stroke. 1997;28:2119–25.

    Google Scholar 

  2. Marler JR, Tilley BC, Lu M, et al. Early stroke treatment associated with better outcome: the NINDS rt-PA stroke study. Neurology. 2000;55:1649–55.

    Article  PubMed  CAS  Google Scholar 

  3. Steiner T, Bluhmki E, Kaste M, et al. The ECASS 3-hour cohort. Secondary analysis of ECASS data by time stratification. ECASS Study Group. European Cooperative Acute Stroke Study. Cerebrovasc Dis. 1998;8:198–203.

    Article  PubMed  CAS  Google Scholar 

  4. Hacke W, Donnan G, Fieschi C, et al. Association of outcome with early stroke treatment: pooled analysis of ATLANTIS, ECASS, and NINDS rt-PA stroke trials. Lancet. 2004;363:768–74.

    Article  PubMed  Google Scholar 

  5. Moser DK, Kimble LP, Alberts MJ, et al. Reducing delay in seeking treatment by patients with acute coronary syndrome and stroke: a scientific statement from the American Heart Association Council on Cardiovascular Nursing and Stroke Council. Circulation. 2006;114:168–82.

    Article  PubMed  Google Scholar 

  6. Sagar G, Riley P, Vohrah A. Is admission chest radiography of any clinical value in acute stroke patients? Clin Radiol. 1996;51:499–502.

    Article  PubMed  CAS  Google Scholar 

  7. Adams Jr HP, Adams RJ, Brott T, et al. Guidelines for the early management of patients with ischemic stroke: a scientific statement from the Stroke Council of the American Stroke Association. Stroke. 2003;34:1056–83.

    Article  PubMed  Google Scholar 

  8. American College of Radiology. Manual on contrast Media. Version 7. Reston, VA; 2010.

    Google Scholar 

  9. Hopyan JJ, Gladstone DJ, Mallia G, et al. Renal safety of CT angiography and perfusion imaging in the emergency evaluation of acute stroke. AJNR Am J Neuroradiol. 2008;29:1826–30.

    Article  PubMed  CAS  Google Scholar 

  10. Lima FO, Lev MH, Levy RA, et al. Functional contrast-enhanced CT for evaluation of acute ischemic stroke does not increase the risk of contrast-induced nephropathy. AJNR Am J Neuroradiol. 2010;31:817–21.

    Article  PubMed  CAS  Google Scholar 

  11. del Zoppo GJ, Higashida RT, Furlan AJ, Pessin MS, Rowley HA, Gent M. PROACT: a phase II randomized trial of recombinant pro-urokinase by direct arterial delivery in acute middle cerebral artery stroke. PROACT Investigators. Prolyse in Acute Cerebral Thromboembolism. Stroke. 1998;29:4–11.

    Article  PubMed  Google Scholar 

  12. Furlan A, Higashida R, Wechsler L, et al. Intra-arterial prourokinase for acute ischemic stroke. The PROACT II study: a randomized controlled trial. Prolyse in Acute Cerebral Thromboembolism. JAMA. 1999;282:2003–11.

    Article  PubMed  CAS  Google Scholar 

  13. Eckert B, Kucinski T, Neumaier-Probst E, Fiehler J, Rother J, Zeumer H. Local intra-arterial fibrinolysis in acute hemispheric stroke: effect of occlusion type and fibrinolytic agent on recanalization success and neurological outcome. Cerebrovasc Dis. 2003;15:258–63.

    Article  PubMed  CAS  Google Scholar 

  14. Davydov L, Cheng JW. Tenecteplase: a review. Clin Ther. 2001;23:982–97; discussion 1.

    Article  PubMed  CAS  Google Scholar 

  15. Hoffmeister HM, Szabo S, Kastner C, et al. Thrombolytic therapy in acute myocardial infarction: comparison of procoagulant effects of streptokinase and alteplase regimens with focus on the kallikrein system and plasmin. Circulation. 1998;98:2527–33.

    Article  PubMed  CAS  Google Scholar 

  16. Haley Jr EC, Lyden PD, Johnston KC, Hemmen TM, the TNKiSI. A pilot dose-escalation safety study of tenecteplase in acute ischemic stroke. Stroke. 2005;36:607–12.

    Article  PubMed  CAS  Google Scholar 

  17. Haley EC, Thompson JLP, Grotta JC, et al. Phase IIB/III trial of tenecteplase in acute ischemic stroke. Stroke. 2010;41:707–11.

    Article  PubMed  CAS  Google Scholar 

  18. Liberatore GT, Samson A, Bladin C, Schleuning W-D, Medcalf RL. Vampire Bat salivary plasminogen activator (desmoteplase): a unique fibrinolytic enzyme that does not promote neurodegeneration. Stroke. 2003;34:537–43.

    Article  PubMed  CAS  Google Scholar 

  19. Hacke W, Albers G, Al-Rawi Y, et al. The Desmoteplase in Acute Ischemic Stroke Trial (DIAS): a phase II MRI-based 9-hour window acute stroke thrombolysis trial with intravenous desmoteplase. Stroke. 2005;36:66–73.

    Article  PubMed  CAS  Google Scholar 

  20. Furlan AJ, Eyding D, Albers GW, et al. Dose Escalation of Desmoteplase for Acute Ischemic Stroke (DEDAS): evidence of safety and efficacy 3 to 9 hours after stroke onset. Stroke. 2006;37:1227–31.

    Article  PubMed  CAS  Google Scholar 

  21. Ewart MR, Hatton MW, Basford JM, Dodgson KS. The proteolytic action of Arvin on human fibrinogen. Biochem J. 1970;118:603–9.

    PubMed  CAS  Google Scholar 

  22. Ehrly AM. Influence of Arwin on the flow properties of blood. Biorheology. 1973;10:453–6.

    PubMed  CAS  Google Scholar 

  23. Tsivgoulis G, Alexandrov AV, Chang J, et al. Safety and outcomes of intravenous thrombolysis in stroke mimics: a 6-year, single-care center study and a pooled analysis of reported series. Stroke. 2011;42:1771–4.

    Article  PubMed  Google Scholar 

  24. Winkler DT, Fluri F, Fuhr P, et al. Thrombolysis in stroke mimics: frequency, clinical characteristics, and outcome. Stroke. 2009;40:1522–5.

    Article  PubMed  Google Scholar 

  25. Latchaw RE, Alberts MJ, Lev MH, et al. Recommendations for imaging of acute ischemic stroke: a scientific statement from the American Heart Association. Stroke. 2009;40:3646–78.

    Article  PubMed  Google Scholar 

  26. Becker RC, Hochman JS, Cannon CP, et al. Fatal cardiac rupture among patients treated with thrombolytic agents and adjunctive thrombin antagonists: observations from the Thrombolysis and Thrombin Inhibition in Myocardial Infarction 9 Study. J Am Coll Cardiol. 1999;33:479–87.

    Article  PubMed  Google Scholar 

  27. Diedler JMD, Ahmed NMDP, Sykora MMD, et al. Safety of intravenous thrombolysis for acute ischemic stroke in patients receiving antiplatelet therapy at stroke onset. Stroke. 2010;41:288–94.

    Article  PubMed  CAS  Google Scholar 

  28. Riedel CH, Zimmermann P, Jensen-Kondering U, Stingele R, Deuschl G, Jansen O. The importance of size: successful recanalization by intravenous thrombolysis in acute anterior stroke depends on thrombus length. Stroke. 2011;42:1775–7.

    Article  PubMed  Google Scholar 

  29. Brinjikji W, Rabinstein AA, Kallmes DF, Cloft HJ. Patient outcomes with endovascular embolectomy therapy for acute ischemic stroke: a study of the national inpatient sample: 2006 to 2008. Stroke. 2011;42:1648–52.

    Article  PubMed  Google Scholar 

  30. Brandt T, von Kummer R, Muller-Kuppers M, Hacke W. Thrombolytic therapy of acute basilar artery occlusion. Variables affecting recanalization and outcome. Stroke. 1996;27:875–81.

    Article  PubMed  CAS  Google Scholar 

  31. Kirton A, Wong JH, Mah J, et al. Successful endovascular therapy for acute basilar thrombosis in an adolescent. Pediatrics. 2003;112:e248–51.

    Article  PubMed  Google Scholar 

  32. Adams HP, del Zoppo GJ, von Kummer R. Management of stroke: a practical guide for the prevention, evaluations and treatment of acute stroke. 2nd ed. Caddo: Professional Communications, Inc.; 2002.

    Google Scholar 

  33. Agarwal P, Kumar S, Hariharan S, et al. Hyperdense middle cerebral artery sign: can it be used to select intra-arterial versus intravenous thrombolysis in acute ischemic stroke? Cerebrovasc Dis. 2004;17:182–90.

    Article  PubMed  Google Scholar 

  34. Demchuk AM, Tanne D, Hill MD, et al. Predictors of good outcome after intravenous tPA for acute ischemic stroke. Neurology. 2001;57:474–80.

    Article  PubMed  CAS  Google Scholar 

  35. The National Institute of Neurological Disorders and Stroke rt-PA Stroke Study Group. Tissue plasminogen activator for acute ischemic stroke. The National Institute of Neurological Disorders and Stroke rt-PA Stroke Study Group. N Engl J Med. 1995;333:1581–7.

    Google Scholar 

  36. Chesebro JH, Knatterud G, Roberts R, et al. Thrombolysis in Myocardial Infarction (TIMI) Trial, phase I: a comparison between intravenous tissue plasminogen activator and intravenous streptokinase. Clinical findings through hospital discharge. Circulation. 1987;76:142–54.

    Article  PubMed  CAS  Google Scholar 

  37. Alexandrov AV, Molina CA, Grotta JC, et al. Ultrasound-enhanced systemic thrombolysis for acute ischemic stroke. N Engl J Med. 2004;351:2170–8.

    Article  PubMed  CAS  Google Scholar 

  38. Qureshi AI. New grading system for angiographic evaluation of arterial occlusions and recanalization response to intra-arterial thrombolysis in acute ischemic stroke. Neurosurgery. 2002;50:1405–15.

    PubMed  Google Scholar 

  39. Francis CW, Blinc A, Lee S, Cox C. Ultrasound accelerates transport of recombinant tissue plasminogen activator into clots. Ultrasound Med Biol. 1995;21:419–24.

    Article  PubMed  CAS  Google Scholar 

  40. Daffertshofer M, Gass A, Ringleb P, et al. Transcranial low-frequency ultrasound-mediated thrombolysis in brain ischemia: increased risk of hemorrhage with combined ultrasound and tissue plasminogen activator: results of a phase II clinical trial. Stroke. 2005;36:1441–6.

    Article  PubMed  Google Scholar 

  41. Viguier A, Petit R, Rigal M, Cintas P, Larrue V. Continuous monitoring of middle cerebral artery recanalization with transcranial color-coded sonography and Levovist. J Thromb Thrombolysis. 2005;19:55–9.

    Article  PubMed  Google Scholar 

  42. Molina CA, Ribo M, Rubiera M, et al. Microbubble administration accelerates clot lysis during continuous 2-MHz ultrasound monitoring in stroke patients treated with intravenous tissue plasminogen activator. Stroke. 2006;37:425–9.

    Article  PubMed  CAS  Google Scholar 

  43. Alexandrov AV, Mikulik R, Ribo M, et al. A pilot randomized clinical safety study of sonothrombolysis augmentation with ultrasound-activated perflutren-lipid microspheres for acute ischemic stroke. Stroke. 2008;39:1464–9.

    Article  PubMed  CAS  Google Scholar 

  44. Molina CA, Barreto AD, Tsivgoulis G, et al. Transcranial ultrasound in clinical sonothrombolysis (TUCSON) trial. Ann Neurol. 2009;66:28–38.

    Article  PubMed  CAS  Google Scholar 

  45. Eggers J, Koch B, Meyer K, Konig I, Seidel G. Effect of ultrasound on thrombolysis of middle cerebral artery occlusion. Ann Neurol. 2003;53:797–800.

    Article  PubMed  Google Scholar 

  46. Eggers J, König IR, Koch B, Händler G, Seidel G. Sonothrombolysis with transcranial color-coded sonography and recombinant tissue-type plasminogen activator in acute middle cerebral artery main stem occlusion: results from a randomized study. Stroke. 2008;39:1470–5.

    Article  PubMed  CAS  Google Scholar 

  47. Perren F, Loulidi J, Poglia D, Landis T, Sztajzel R. Microbubble potentiated transcranial duplex ultrasound enhances IV thrombolysis in acute stroke. J Thromb Thrombolysis. 2008;25:219–23.

    Article  PubMed  Google Scholar 

  48. Tsivgoulis GMD, Eggers JMD, Ribo MMD, et al. Safety and efficacy of ultrasound-enhanced thrombolysis: a comprehensive review and meta-analysis of randomized and nonrandomized studies. Stroke. 2010;41:280–7.

    Article  PubMed  Google Scholar 

  49. Butcher KMDP, Christensen SP, Parsons MPF, et al. Postthrombolysis blood pressure elevation is associated with hemorrhagic transformation. Stroke. 2010;41:72–7.

    Article  PubMed  Google Scholar 

  50. Graham GD. Tissue plasminogen activator for acute ischemic stroke in clinical practice: a meta-analysis of safety data. Stroke. 2003;34:2847–50.

    Article  PubMed  CAS  Google Scholar 

  51. Hill MD, Lye T, Moss H, et al. Hemi-orolingual angioedema and ACE inhibition after alteplase treatment of stroke. Neurology. 2003;60:1525–7.

    Article  PubMed  CAS  Google Scholar 

  52. Hill MD, Buchan AM. Thrombolysis for acute ischemic stroke: results of the Canadian Alteplase for Stroke Effectiveness Study. CMAJ. 2005;172:1307–12.

    Article  PubMed  Google Scholar 

  53. Engelter ST, Fluri F, Buitrago-Tellez C, et al. Life-threatening orolingual angioedema during thrombolysis in acute ischemic stroke. J Neurol. 2005;252:1167–70.

    Article  PubMed  CAS  Google Scholar 

  54. Abou-Chebl A, Lin R, Hussain MS, et al. Conscious sedation versus general anesthesia during endovascular therapy for acute anterior circulation stroke. Stroke. 2010;41:1175–9.

    Article  PubMed  CAS  Google Scholar 

  55. Hemmer LB, Zeeni C, Gupta DK. Generalizations about general anesthesia: the unsubstantiated condemnation of general anesthesia for patients undergoing intra-arterial therapy for anterior circulation stroke. Stroke. 2010;41:e573.

    Article  PubMed  Google Scholar 

  56. Kumpe DA. Thrombolysis of acute stroke syndromes. In: Kandarpa K, Aruny JE, editors. Handbook of interventional radiologic procedures. Philadelphia: Lippincott Williams & Wilkins; 2002. p. 47–62.

    Google Scholar 

  57. Rubiera M, Cava L, Tsivgoulis G, et al. Diagnostic criteria and yield of real-time transcranial Doppler monitoring of intra-arterial reperfusion procedures. Stroke. 2010;41:695–9.

    Article  PubMed  Google Scholar 

  58. Khatri P, Broderick JP, Khoury JC, Carrozzella JA, Tomsick TA. Microcatheter contrast injections during intra-arterial thrombolysis may increase intracranial hemorrhage risk. Stroke. 2008;39:3283–7.

    Article  PubMed  Google Scholar 

  59. Chopko BW, Kerber C, Wong W, Georgy B. Transcatheter snare removal of acute middle cerebral artery thromboembolism: technical case report. Neurosurgery. 2000;46:1529–31.

    Article  PubMed  CAS  Google Scholar 

  60. Kerber CW, Barr JD, Berger RM, Chopko BW. Snare retrieval of intracranial thrombus in patients with acute stroke. J Vasc Interv Radiol. 2002;13:1269–74.

    Article  PubMed  Google Scholar 

  61. Fourie P, Duncan IC. Microsnare-assisted mechanical removal of intraprocedural distal middle cerebral arterial thromboembolism. AJNR Am J Neuroradiol. 2003;24:630–2.

    PubMed  Google Scholar 

  62. Wikholm G. Transarterial embolectomy in acute stroke. AJNR Am J Neuroradiol. 2003;24:892–4.

    PubMed  Google Scholar 

  63. Nesbit GM, Luh G, Tien R, Barnwell SL. New and future endovascular treatment strategies for acute ischemic stroke. J Vasc Interv Radiol. 2004;15:103S–10.

    Article  Google Scholar 

  64. Kerber CW, Wanke I, Bernard Jr J, Woo HH, Liu MW, Nelson PK. Rapid intracranial clot removal with a new device: the alligator retriever. AJNR Am J Neuroradiol. 2007;28:860–3.

    PubMed  CAS  Google Scholar 

  65. Lutsep HL, Clark WM, Nesbit GM, Kuether TA, Barnwell SL. Intraarterial suction thrombectomy in acute stroke. AJNR Am J Neuroradiol. 2002;23:783–6.

    PubMed  Google Scholar 

  66. Chapot R, Houdart E, Rogopoulos A, Mounayer C, Saint-Maurice JP, Merland JJ. Thromboaspiration in the basilar artery: report of two cases. AJNR Am J Neuroradiol. 2002;23:282–4.

    PubMed  Google Scholar 

  67. Nedeltchev K, Remonda L, Do DD, et al. Acute stenting and thromboaspiration in basilar artery occlusions due to embolism from the dominating vertebral artery. Neuroradiology. 2004;46:686–91.

    Article  PubMed  CAS  Google Scholar 

  68. Cross 3rd DT, Moran CJ, Akins PT, Angtuaco EE, Derdeyn CP, Diringer MN. Collateral circulation and outcome after basilar artery thrombolysis. AJNR Am J Neuroradiol. 1998;19:1557–63.

    PubMed  Google Scholar 

  69. Nakayama T, Tanaka K, Kaneko M, Yokoyama T, Uemura K. Thrombolysis and angioplasty for acute occlusion of intracranial vertebrobasilar arteries. Report of three cases. J Neurosurg. 1998;88:919–22.

    Article  PubMed  CAS  Google Scholar 

  70. Mori T, Kazita K, Mima T, Mori K. Balloon angioplasty for embolic total occlusion of the middle cerebral artery and ipsilateral carotid stenting in an acute stroke stage. AJNR Am J Neuroradiol. 1999;20:1462–4.

    PubMed  CAS  Google Scholar 

  71. Ringer AJ, Qureshi AI, Fessler RD, Guterman LR, Hopkins LN. Angioplasty of intracranial occlusion resistant to thrombolysis in acute ischemic stroke. Neurosurgery. 2001;48:1282–90.

    PubMed  CAS  Google Scholar 

  72. Shi ZS, Liebeskind DS, Loh Y, et al. Predictors of subarachnoid hemorrhage in acute ischemic stroke with endovascular therapy. Stroke. 2010;41:2775–81.

    Article  PubMed  Google Scholar 

  73. Gupta R, Vora NA, Horowitz MB, et al. Multimodal reperfusion therapy for acute ischemic stroke: factors predicting vessel recanalization. Stroke. 2006;37:986–90.

    Article  PubMed  Google Scholar 

  74. Levy EI, Mehta R, Gupta R, et al. Self-expanding stents for recanalization of acute cerebrovascular occlusions. AJNR Am J Neuroradiol. 2007;28:816–22.

    PubMed  CAS  Google Scholar 

  75. Castano CMDP, Dorado LMD, Guerrero CMD, et al. Mechanical thrombectomy with the solitaire AB device in large artery occlusions of the anterior circulation: a pilot study. Stroke. 2010;41:1836–40.

    Article  PubMed  Google Scholar 

  76. Smith WS, Sung G, Starkman S, et al. Safety and efficacy of mechanical embolectomy in acute ischemic stroke: results of the MERCI trial. Stroke. 2005;36:1432–8.

    Article  PubMed  Google Scholar 

  77. Smith WS. Safety of mechanical thrombectomy and intravenous tissue plasminogen activator in acute ischemic stroke. Results of the multi Mechanical Embolus Removal in Cerebral Ischemia (MERCI) trial, part I. AJNR Am J Neuroradiol. 2006;27:1177–82.

    PubMed  CAS  Google Scholar 

  78. Layton KF, White JB, Cloft HJ, Kallmes DF. Use of the Perclose ProGlide device with the 9 French Merci retrieval system. Neuroradiology. 2006;48:324–6.

    Article  PubMed  Google Scholar 

  79. Spiotta AM, Hussain MS, Sivapatham T, et al. The versatile distal access catheter: the Cleveland Clinic experience. Neurosurgery. 2011;68:1677–86; discussion 86.

    Article  PubMed  Google Scholar 

  80. Hui FK, Hussain MS, Spiotta A, et al. Merci retrievers as access adjuncts for reperfusion catheters: the grappling hook technique. Neurosurgery. 2012;70(2):456–60.

    Article  PubMed  Google Scholar 

  81. Henkes H, Flesser A, Brew S, et al. A novel microcatheter-delivered, highly-flexible and fully-retrievable stent, specifically designed for intracranial use. Technical note. Interv Neuroradiol. 2003;9:391–3.

    PubMed  CAS  Google Scholar 

  82. Roth CM, Papanagiotou PM, Behnke SM, et al. Stent-assisted mechanical recanalization for treatment of acute intracerebral artery occlusions. Stroke. 2010;41:2559–67.

    Article  PubMed  CAS  Google Scholar 

  83. SolitareTM FR With the Intention For Thrombectomy (SWIFT) study. 2011. http://clinicaltrials.gov/ct2/show/NCT01054560. Accessed 11 Aug 2011.

  84. Mahon BR, Nesbit GM, Barnwell SL, et al. North American clinical experience with the EKOS MicroLysUS infusion catheter for the treatment of embolic stroke. AJNR Am J Neuroradiol. 2003;24:534–8.

    PubMed  Google Scholar 

  85. IMS II Trial Investigators. The Interventional Management of Stroke (IMS) II Study. Stroke. 2007;38:2127–35.

    Article  Google Scholar 

  86. Opatowsky MJ, Morris PP, Regan JD, Mewborne JD, Wilson JA. Rapid thrombectomy of superior sagittal sinus and transverse sinus thrombosis with a rheolytic catheter device. AJNR Am J Neuroradiol. 1999;20:414–7.

    PubMed  CAS  Google Scholar 

  87. Bellon RJ, Putman CM, Budzik RF, Pergolizzi RS, Reinking GF, Norbash AM. Rheolytic thrombectomy of the occluded internal carotid artery in the setting of acute ischemic stroke. AJNR Am J Neuroradiol. 2001;22:526–30.

    PubMed  CAS  Google Scholar 

  88. Molina CA, Saver JL. Extending reperfusion therapy for acute ischemic stroke: emerging pharmacological, mechanical, and imaging strategies. Stroke. 2005;36:2311–20.

    Article  PubMed  Google Scholar 

  89. Henkes H, Reinartz J, Lowens S, et al. A device for fast mechanical clot retrieval from intracranial arteries (Phenox clot retriever). Neurocrit Care. 2006;5:134–40.

    Article  PubMed  Google Scholar 

  90. Liebig T, Reinartz J, Hannes R, Miloslavski E, Henkes H. Comparative in vitro study of five mechanical embolectomy systems: effectiveness of clot removal and risk of distal embolization. Neuroradiology. 2008;50:43–52.

    Article  PubMed  Google Scholar 

  91. Brekenfeld C, Schroth G, El-Koussy M, et al. Mechanical thromboembolectomy for acute ischemic stroke: comparison of the catch thrombectomy device and the Merci Retriever in vivo. Stroke. 2008;39:1213–9.

    Article  PubMed  Google Scholar 

  92. Lylyk P, Vila JF, Miranda C, Ferrario A, Romero R, Cohen JE. Partial aortic obstruction improves cerebral perfusion and clinical symptoms in patients with symptomatic vasospasm. Neurol Res. 2005;27 Suppl 1:S129–35.

    Article  PubMed  Google Scholar 

  93. Alnaami I, Saqqur M, Chow M. A novel treatment of distal cerebral vasospasm. A case report. Interv Neuroradiol. 2009;15:417–20.

    PubMed  CAS  Google Scholar 

  94. Shuaib A, Bornstein NM, Diener HC, et al. Partial aortic occlusion for cerebral perfusion augmentation: safety and efficacy of NeuroFlo in Acute Ischemic Stroke trial. Stroke. 2011;42:1680–90.

    Article  PubMed  Google Scholar 

  95. Kase CS, Furlan AJ, Wechsler LR, et al. Cerebral hemorrhage after intra-arterial thrombolysis for ischemic stroke: the PROACT II trial. Neurology. 2001;57:1603–10.

    Article  PubMed  CAS  Google Scholar 

  96. Lisboa RC, Jovanovic BD, Alberts MJ. Analysis of the safety and efficacy of intra-arterial thrombolytic therapy in ischemic stroke. Stroke. 2002;33:2866–71.

    Article  PubMed  CAS  Google Scholar 

  97. Khatri P, Hill MD, Palesch YY, et al. Methodology of the interventional management of stroke III trial. Int J Stroke. 2008;3:130–7.

    Article  PubMed  Google Scholar 

  98. Lewandowski CA, Frankel M, Tomsick TA, et al. Combined intravenous and intra-arterial r-TPA versus intra-arterial therapy of acute ischemic stroke: Emergency Management of Stroke (EMS) Bridging Trial. Stroke. 1999;30:2598–605.

    Article  PubMed  CAS  Google Scholar 

  99. The IMSSI. Combined intravenous and intra-arterial recanalization for acute ischemic stroke: the interventional management of stroke study. Stroke. 2004;35:904–11.

    Article  Google Scholar 

  100. Mazighi M, Serfaty JM, Labreuche J, et al. Comparison of intravenous alteplase with a combined intravenous-endovascular approach in patients with stroke and confirmed arterial occlusion (RECANALISE study): a prospective cohort study. Lancet Neurol. 2009;8:802–9.

    Article  PubMed  CAS  Google Scholar 

  101. Rubiera M, Ribo M, Pagola J, et al. Bridging intravenous-intra-arterial rescue strategy increases recanalization and the likelihood of a good outcome in nonresponder intravenous tissue plasminogen activator-treated patients: a case–control study. Stroke. 2011;42:993–7.

    Article  PubMed  CAS  Google Scholar 

  102. Eckert B, Koch C, Thomalla G, et al. Aggressive therapy with intravenous abciximab and intra-arterial rtPA and additional PTA/stenting improves clinical outcome in acute vertebrobasilar occlusion: combined local fibrinolysis and intravenous abciximab in acute vertebrobasilar stroke treatment (FAST): results of a multicenter study. Stroke. 2005;36:1160–5.

    Article  PubMed  CAS  Google Scholar 

  103. Nagel S, Schellinger PD, Hartmann M, et al. Therapy of acute basilar artery occlusion: intraarterial thrombolysis alone vs bridging therapy. Stroke. 2009;40:140–6.

    Article  PubMed  Google Scholar 

  104. Hill MD, Barber PA, Takahashi J, Demchuk AM, Feasby TE, Buchan AM. Anaphylactoid reactions and angioedema during alteplase treatment of acute ischemic stroke. CMAJ. 2000;162:1281–4.

    PubMed  CAS  Google Scholar 

  105. Brandt T. Diagnosis and thrombolytic therapy of acute basilar artery occlusion: a review. Clin Exp Hypertens. 2002;24:611–22.

    Article  PubMed  Google Scholar 

  106. Brandt T, Knauth M, Wildermuth S, et al. CT angiography and Doppler sonography for emergency assessment in acute basilar artery ischemia. Stroke. 1999;30:606–12.

    Article  PubMed  CAS  Google Scholar 

  107. Schellinger PD, Hacke W. Intra-arterial thrombolysis is the treatment of choice for basilar thrombosis: pro. Stroke. 2006;37:2436–7.

    Article  PubMed  Google Scholar 

  108. Ford GA. Intra-arterial thrombolysis is the treatment of choice for basilar thrombosis: con. Stroke. 2006;37:2438–9.

    Article  PubMed  Google Scholar 

  109. Macleod MR, Davis SM, Mitchell PJ, et al. Results of a multicentre, randomised controlled trial of intra-arterial urokinase in the treatment of acute posterior circulation ischaemic stroke. Cerebrovasc Dis. 2005;20:12–7.

    Article  PubMed  CAS  Google Scholar 

  110. Schonewille WJ, Wijman CA, Michel P, et al. Treatment and outcomes of acute basilar artery occlusion in the Basilar Artery International Cooperation Study (BASICS): a prospective registry study. Lancet Neurol. 2009;8:724–30.

    Article  PubMed  Google Scholar 

  111. Davis SM, Donnan GA. Basilar artery thrombosis: recanalization is the key. Stroke. 2006;37:2440.

    Article  PubMed  Google Scholar 

  112. Hacke W, Zeumer H, Ferbert A, Bruckmann H, del Zoppo GJ. Intra-arterial thrombolytic therapy improves outcome in patients with acute vertebrobasilar occlusive disease. Stroke. 1988;19:1216–22.

    Article  PubMed  CAS  Google Scholar 

  113. Wang H, Fraser K, Wang D, Alvernia J, Lanzino G. Successful intra-arterial basilar artery thrombolysis in a patient with bilateral vertebral artery occlusion: technical case report. Neurosurgery. 2005;57:E398; discussion E.

    Article  PubMed  Google Scholar 

  114. Lindsberg PJ, Mattle HP. Therapy of basilar artery occlusion: a systematic analysis comparing intra-arterial and intravenous thrombolysis. Stroke. 2006;37:922–8.

    Article  PubMed  Google Scholar 

  115. Nesbit GM, Clark WM, O’Neill OR, Barnwell SL. Intracranial intraarterial thrombolysis facilitated by microcatheter navigation through an occluded cervical internal carotid artery. J Neurosurg. 1996;84:387–92.

    Article  PubMed  CAS  Google Scholar 

  116. Hui FK, Hussain MS, Elgabaly MH, Sivapatham T, Katzan IL, Spiotta AM. Embolic protection devices and the Penumbra 054 catheter: utility in tandem occlusions in acute ischemic stroke. J Neurointerv Surg. 2011;3:50–3.

    Article  PubMed  Google Scholar 

  117. Steinhubl SR, Talley JD, Braden GA, et al. Point-of-care measured platelet inhibition correlates with a reduced risk of an adverse cardiac event after percutaneous coronary intervention: results of the GOLD (AU-Assessing Ultegra) multicenter study. Circulation. 2001;103:2572–8.

    Article  PubMed  CAS  Google Scholar 

  118. Quinn MJ, Plow EF, Topol EJ. Platelet glycoprotein IIb/IIIa inhibitors: recognition of a two-edged sword? Circulation. 2002;106:379–85.

    Article  PubMed  CAS  Google Scholar 

  119. Kleinman N. Assessing platelet function in clinical trials. In: Quinn M, Fitzgerald D, editors. Platelet function assessment, diagnosis, and treatment. Totowa: Humana Press; 2005. p. 369–84.

    Google Scholar 

  120. Fisher CM, Ojemann RG, Roberson GH. Spontaneous dissection of cervico-cerebral arteries. Can J Neurol Sci. 1978;5:9–19.

    PubMed  CAS  Google Scholar 

  121. Fitzsimmons BFM, Becske T, Nelson PK. Rapid stent-supported revascularization in acute ischemic stroke. AJNR Am J Neuroradiol. 2006;27:1132–4.

    PubMed  Google Scholar 

  122. Kessler IM, Mounayer C, Piotin M, Spelle L, Vanzin JR, Moret J. The use of balloon-expandable stents in the management of intracranial arterial diseases: a 5-year single-center experience. AJNR Am J Neuroradiol. 2005;26:2342–8.

    PubMed  Google Scholar 

  123. Kiyosue H, Okahara M, Yamashita M, Nagatomi H, Nakamura N, Mori H. Endovascular stenting for restenosis of the intracranial vertebrobasilar artery after balloon angioplasty: two case reports and review of the literature. Cardiovasc Intervent Radiol. 2004;27:538–43.

    PubMed  Google Scholar 

  124. Leavitt JA, Larson TA, Hodge DO, Gullerud RE. The incidence of central retinal artery occlusion in Olmsted County, Minnesota. Am J Ophthalmol. 2011;152(5):820–3.

    Article  Google Scholar 

  125. Hayreh SS, Zimmerman MB, Kimura A, Sanon A. Central retinal artery occlusion. Retinal survival time. Exp Eye Res. 2004;78:723–36.

    Article  PubMed  CAS  Google Scholar 

  126. Beatty S, Au Eong KG. Local intra-arterial fibrinolysis for acute occlusion of the central retinal artery: a meta-analysis of the published data. Br J Ophthalmol. 2000;84:914–6.

    Article  PubMed  CAS  Google Scholar 

  127. Weber J, Remonda L, Mattle HP, et al. Selective intra-arterial fibrinolysis of acute central retinal artery occlusion. Stroke. 1998;29:2076–9.

    Article  PubMed  CAS  Google Scholar 

  128. Butz B, Strotzer M, Manke C, Roider J, Link J, Lenhart M. Selective intraarterial fibrinolysis of acute central retinal artery occlusion. Acta Radiol. 2003;44:680–4.

    PubMed  CAS  Google Scholar 

  129. Arnold M, Koerner U, Remonda L, et al. Comparison of intra-arterial thrombolysis with conventional treatment in patients with acute central retinal artery occlusion. J Neurol Neurosurg Psychiatry. 2005;76:196–9.

    Article  PubMed  CAS  Google Scholar 

  130. Schmidt DP, Schulte-Monting J, Schumacher M. Prognosis of central retinal artery occlusion: local intraarterial fibrinolysis versus conservative treatment. AJNR Am J Neuroradiol. 2002;23:1301–7.

    PubMed  Google Scholar 

  131. Aldrich EM, Lee AW, Chen CS, et al. Local intraarterial fibrinolysis administered in aliquots for the treatment of central retinal artery occlusion: the Johns Hopkins Hospital experience. Stroke. 2008;39:1746–50.

    Article  PubMed  Google Scholar 

  132. Mueller AJ, Neubauer AS, Schaller U, Kampik A. Evaluation of minimally invasive therapies and rationale for a prospective randomized trial to evaluate selective intra-arterial lysis for clinically complete central retinal artery occlusion. Arch Ophthalmol. 2003;121:1377–81.

    Article  PubMed  Google Scholar 

  133. Fraser SG, Adams W. Interventions for acute non-arteritic central retinal artery occlusion. Cochrane Database Syst Rev. 2009:CD001989.

    Google Scholar 

  134. Schmidt D, Schumacher M, Wakhloo AK. Microcatheter urokinase infusion in central retinal artery occlusion. Am J Ophthalmol. 1992;113:429–34.

    PubMed  CAS  Google Scholar 

  135. Atebara NH, Brown GC, Cater J. Efficacy of anterior chamber paracentesis and Carbogen in treating acute nonarteritic central retinal artery occlusion. Ophthalmology. 1995;102:2029–34; discussion 34–5.

    PubMed  CAS  Google Scholar 

  136. Schumacher M, Schmidt D, Jurklies B, et al. Central retinal artery occlusion: local intra-arterial fibrinolysis versus conservative treatment, a multicenter randomized trial. Ophthalmology. 2010;117:1367.e1–75.e1.

    Article  Google Scholar 

  137. Ros MA, Magargal LE, Uram M. Branch retinal-artery obstruction: a review of 201 eyes. Ann Ophthalmol. 1989;21:103–7.

    PubMed  CAS  Google Scholar 

  138. Paques M, Vallee JN, Herbreteau D, et al. Superselective ophthalmic artery fibrinolytic therapy for the treatment of central retinal vein occlusion. Br J Ophthalmol. 2000;84:1387–91.

    Article  PubMed  CAS  Google Scholar 

  139. Kilani R, Marshall L, Koch S, Fernandez M, Postel E. DWI findings of optic nerve ischemia in the setting of central retinal artery occlusion. J Neuroimaging. 2011 Jun 23. doi: 10.1111/j.1552-6569.2011.00601.x. [Epub ahead of print]

    Google Scholar 

  140. Feltgen N, Neubauer A, Jurklies B, et al. Multicenter study of the European Assessment Group for Lysis in the Eye (EAGLE) for the treatment of central retinal artery occlusion: design issues and implications. EAGLE Study report no. 1: EAGLE Study report no. 1. Graefes Arch Clin Exp Ophthalmol. 2006;244:950–6.

    Article  PubMed  CAS  Google Scholar 

  141. Richard G, Lerche RC, Knospe V, Zeumer H. Treatment of retinal arterial occlusion with local fibrinolysis using recombinant tissue plasminogen activator. Ophthalmology. 1999;106:768–73.

    Article  PubMed  CAS  Google Scholar 

  142. Hayreh SS. Acute retinal arterial occlusive disorders. Prog Retin Eye Res. 2011;30:359–94.

    Article  PubMed  Google Scholar 

  143. Beatty S, Au Eong KG. Acute occlusion of the retinal arteries: current concepts and recent advances in diagnosis and management. J Accid Emerg Med. 2000;17:324–9.

    Article  PubMed  CAS  Google Scholar 

  144. Chen CS, Lee AW, Campbell B, et al. Efficacy of intravenous tissue-type plasminogen activator in central retinal artery occlusion: report from a randomized, controlled trial. Stroke. 2011;42:2229–34.

    Article  PubMed  CAS  Google Scholar 

  145. Adams Jr HP, del Zoppo G, Alberts MJ, et al. Guidelines for the early management of adults with ischemic stroke: a guideline from the American Heart Association/American Stroke Association Stroke Council, Clinical Cardiology Council, Cardiovascular Radiology and Intervention Council, and the Atherosclerotic Peripheral Vascular Disease and Quality of Care Outcomes in Research Interdisciplinary Working Groups: the American Academy of Neurology affirms the value of this guideline as an educational tool for neurologists. Stroke. 2007;38:1655–711.

    Article  PubMed  Google Scholar 

  146. Chalela JA, Kidwell CS, Nentwich LM, et al. Magnetic resonance imaging and computed tomography in emergency assessment of patients with suspected acute stroke: a prospective comparison. Lancet. 2007;369:293–8.

    Article  PubMed  Google Scholar 

  147. Wessels T, Wessels C, Ellsiepen A, et al. Contribution of diffusion-weighted imaging in determination of stroke etiology. AJNR Am J Neuroradiol. 2006;27:35–9.

    PubMed  CAS  Google Scholar 

  148. von Kummer R, Bourquain H, Bastianello S, et al. Early prediction of irreversible brain damage after ischemic stroke at CT. Radiology. 2001;219:95–100.

    Google Scholar 

  149. Mullins ME, Schaefer PW, Sorensen AG, et al. CT and conventional and diffusion-weighted MR imaging in acute stroke: study in 691 patients at presentation to the emergency department. Radiology. 2002;224:353–60.

    Article  PubMed  Google Scholar 

  150. Lansberg MG, Albers GW, Beaulieu C, Marks MP. Comparison of diffusion-weighted MRI and CT in acute stroke. Neurology. 2000;54:1557–61.

    Article  PubMed  CAS  Google Scholar 

  151. Larrue V, von Kummer RR, Muller A, Bluhmki E. Risk factors for severe hemorrhagic transformation in ischemic stroke patients treated with recombinant tissue plasminogen activator: a secondary analysis of the European-Australasian Acute Stroke Study (ECASS II). Stroke. 2001;32:438–41.

    Article  PubMed  CAS  Google Scholar 

  152. Kasner SE, Demchuk AM, Berrouschot J, et al. Predictors of fatal brain edema in massive hemispheric ischemic stroke. Stroke. 2001;32:2117–23.

    Article  PubMed  CAS  Google Scholar 

  153. Menon BK, Puetz V, Kochar P, Demchuk AM. ASPECTS and other neuroimaging scores in the triage and prediction of outcome in acute stroke patients. Neuroimaging Clin N Am. 2011;21:407–23, xii.

    Article  PubMed  Google Scholar 

  154. Barber PA, Demchuk AM, Zhang J, Buchan AM. Validity and reliability of a quantitative computed tomography score in predicting outcome of hyperacute stroke before thrombolytic therapy. ASPECTS Study Group. Alberta Stroke Programme Early CT Score. Lancet. 2000;355:1670–4.

    Article  PubMed  CAS  Google Scholar 

  155. Castillo PR, Miller DA, Meschia JF. Choice of neuroimaging in perioperative acute stroke management. Neurol Clin. 2006;24:807–20.

    Article  PubMed  Google Scholar 

  156. von Kummer R, Nolte PN, Schnittger H, Thron A, Ringelstein EB. Detectability of cerebral hemisphere ischaemic infarcts by CT within 6 h of stroke. Neuroradiology. 1996;38:31–3.

    Article  Google Scholar 

  157. Truwit CL, Barkovich AJ, Gean-Marton A, Hibri N, Norman D. Loss of the insular ribbon: another early CT sign of acute middle cerebral artery infarction. Radiology. 1990;176:801–6.

    PubMed  CAS  Google Scholar 

  158. Tomura N, Uemura K, Inugami A, Fujita H, Higano S, Shishido F. Early CT finding in cerebral infarction: obscuration of the lentiform nucleus. Radiology. 1988;168:463–7.

    PubMed  CAS  Google Scholar 

  159. Gacs G, Fox AJ, Barnett HJ, Vinuela F. CT visualization of intracranial arterial thromboembolism. Stroke. 1983;14: 756–62.

    Article  PubMed  CAS  Google Scholar 

  160. Tomsick TA, Brott TG, Olinger CP, et al. Hyperdense middle cerebral artery: incidence and quantitative significance. Neuroradiology. 1989;31:312–5.

    Article  PubMed  CAS  Google Scholar 

  161. Barber PA, Demchuk AM, Hill MD, et al. The probability of middle cerebral artery MRA flow signal abnormality with quantified CT ischaemic change: targets for future therapeutic studies. J Neurol Neurosurg Psychiatry. 2004;75:1426–30.

    Article  PubMed  CAS  Google Scholar 

  162. Barber PA, Demchuk AM, Hudon ME, Pexman JH, Hill MD, Buchan AM. Hyperdense sylvian fissure MCA “dot” sign: A CT marker of acute ischemia. Stroke. 2001;32:84–8.

    Article  PubMed  CAS  Google Scholar 

  163. Leary MC, Kidwell CS, Villablanca JP, et al. Validation of computed tomographic middle cerebral artery “dot” sign: an angiographic correlation study. Stroke. 2003;34:2636–40.

    Article  PubMed  Google Scholar 

  164. Koga M, Saku Y, Toyoda K, Takaba H, Ibayashi S, Iida M. Reappraisal of early CT signs to predict the arterial occlusion site in acute embolic stroke. J Neurol Neurosurg Psychiatry. 2003;74:649–53.

    Article  PubMed  CAS  Google Scholar 

  165. Lyden PD. Advanced brain imaging studies should not be performed in patients with suspected stroke presenting within 4.5 hours of symptom onset. Stroke. 2011;42:2668–9.

    Article  PubMed  Google Scholar 

  166. Selim MH, Molina CA. Conundra of the penumbra and acute stroke imaging. Stroke. 2011;42:2670–1.

    Article  PubMed  Google Scholar 

  167. Axel L. Cerebral blood flow determination by rapid sequence computed tomography. Radiology. 1980;137:679–86.

    PubMed  CAS  Google Scholar 

  168. Mies G, Ishimaru S, Xie Y, Seo K, Hossmann KA. Ischemic thresholds of cerebral protein synthesis and energy state following middle cerebral artery occlusion in rat. J Cereb Blood Flow Metab. 1991;11:753–61.

    Article  PubMed  CAS  Google Scholar 

  169. Astrup J, Symon L, Branston NM, Lassen NA. Cortical evoked potential and extracellular K+ and H+ at critical levels of brain ischemia. Stroke. 1977;8:51–7.

    Article  PubMed  CAS  Google Scholar 

  170. Morawetz RB, Crowell RH, DeGirolami U, Marcoux FW, Jones TH, Halsey JH. Regional cerebral blood flow thresholds during cerebral ischemia. Fed Proc. 1979;38:2493–4.

    PubMed  CAS  Google Scholar 

  171. Morawetz RB, DeGirolami U, Ojemann RG, Marcoux FW, Crowell RM. Cerebral blood flow determined by hydrogen clearance during middle cerebral artery occlusion in unanesthetized monkeys. Stroke. 1978;9:143–9.

    Article  PubMed  CAS  Google Scholar 

  172. Sakai F, Nakazawa K, Tazaki Y, et al. Regional cerebral blood volume and hematocrit measured in normal human volunteers by single-photon emission computed tomography. J Cereb Blood Flow Metab. 1985;5:207–13.

    Article  PubMed  CAS  Google Scholar 

  173. Muizelaar JP, Fatouros PP, Schroder ML. A new method for quantitative regional cerebral blood volume measurements using computed tomography. Stroke. 1997;28:1998–2005.

    Article  PubMed  CAS  Google Scholar 

  174. Nabavi DG, Cenic A, Dool J, et al. Quantitative assessment of cerebral hemodynamics using CT: stability, accuracy, and precision studies in dogs. J Comput Assist Tomogr. 1999;23:506–15.

    Article  PubMed  CAS  Google Scholar 

  175. Hatazawa J, Shimosegawa E, Toyoshima H, et al. Cerebral blood volume in acute brain infarction: a combined study with dynamic susceptibility contrast MRI and 99mTc-HMPAO-SPECT. Stroke. 1999;30:800–6.

    Article  PubMed  CAS  Google Scholar 

  176. Todd NV, Picozzi P, Crockard HA. Quantitative measurement of cerebral blood flow and cerebral blood volume after cerebral ischaemia. J Cereb Blood Flow Metab. 1986;6:338–41.

    Article  PubMed  CAS  Google Scholar 

  177. Latchaw RE, Yonas H, Hunter GJ, et al. Guidelines and recommendations for perfusion imaging in cerebral ischemia: a scientific statement for healthcare professionals by the writing group on perfusion imaging, from the Council on Cardiovascular Radiology of the American Heart Association. Stroke. 2003;34:1084–104.

    Article  PubMed  Google Scholar 

  178. Konstas AA, Goldmakher GV, Lee TY, Lev MH. Theoretic basis and technical implementations of CT perfusion in acute ischemic stroke, part 1: theoretic basis. AJNR Am J Neuroradiol. 2009;30:662–8.

    Article  PubMed  CAS  Google Scholar 

  179. Klotz E, Konig M. Perfusion measurements of the brain: using dynamic CT for the quantitative assessment of cerebral ischemia in acute stroke. Eur J Radiol. 1999;30:170–84.

    Article  PubMed  CAS  Google Scholar 

  180. Miles K. Measurement of tissue perfusion by dynamic computed tomography. Br J Radiol. 1991;64:409–12.

    Article  PubMed  CAS  Google Scholar 

  181. Koenig M, Klotz E, Heuser L. Perfusion CT in acute stroke: characterization of cerebral ischemia using parameter images of cerebral blood flow and their therapeutic relevance. Electromedica. 1998;66:61–6.

    Google Scholar 

  182. Steiger HJ, Aaslid R, Stooss R. Dynamic computed tomographic imaging of regional cerebral blood flow and blood volume. A clinical pilot study. Stroke. 1993;24:591–7.

    Article  PubMed  CAS  Google Scholar 

  183. Hunter GJ, Hamberg LM, Ponzo JA, et al. Assessment of cerebral perfusion and arterial anatomy in hyperacute stroke with three-dimensional functional CT: early clinical results. AJNR Am J Neuroradiol. 1998;19:29–37.

    PubMed  CAS  Google Scholar 

  184. Wintermark M, Maeder P, Thiran JP, Schnyder P, Meuli R. Quantitative assessment of regional cerebral blood flows by perfusion CT studies at low injection rates: a critical review of the underlying theoretical models. Eur Radiol. 2001;11:1220–30.

    Article  PubMed  CAS  Google Scholar 

  185. Gobbel G, Cann C, Fike J. Measurement of regional cerebral blood flow using ultrafast computed tomography. Theoretical aspects. Stroke. 1991;22:768–71.

    Article  PubMed  CAS  Google Scholar 

  186. Gobbel GT, Cann CE, Fike JR. Comparison of xenon-enhanced CT with ultrafast CT for measurement of regional cerebral blood flow. AJNR Am J Neuroradiol. 1993;14:543–50.

    PubMed  CAS  Google Scholar 

  187. Wintermark M, Thiran JP, Maeder P, Schnyder P, Meuli R. Simultaneous measurement of regional cerebral blood flow by perfusion CT and stable xenon CT: a validation study. AJNR Am J Neuroradiol. 2001;22:905–14.

    PubMed  CAS  Google Scholar 

  188. Gillard JH, Minhas PS, Hayball MP, et al. Assessment of quantitative computed tomographic cerebral perfusion imaging with H2(15)O positron emission tomography. Neurol Res. 2000;22:457–64.

    PubMed  CAS  Google Scholar 

  189. Kudo K, Terae S, Katoh C, et al. Quantitative cerebral blood flow measurement with dynamic perfusion CT using the vascular-pixel elimination method: comparison with H2(15)O positron emission tomography. AJNR Am J Neuroradiol. 2003;24:419–26.

    PubMed  Google Scholar 

  190. Nabavi DG, Cenic A, Craen RA, et al. CT assessment of cerebral perfusion: experimental validation and initial clinical experience. Radiology. 1999;213:141–9.

    PubMed  CAS  Google Scholar 

  191. Nabavi DG, Cenic A, Henderson S, Gelb AW, Lee TY. Perfusion mapping using computed tomography allows accurate prediction of cerebral infarction in experimental brain ischemia. Stroke. 2001;32:175–83.

    Article  PubMed  CAS  Google Scholar 

  192. Hamberg LM, Hunter GJ, Maynard KI, et al. Functional CT perfusion imaging in predicting the extent of cerebral infarction from a 3-hour middle cerebral arterial occlusion in a primate stroke model. AJNR Am J Neuroradiol. 2002;23:1013–21.

    PubMed  Google Scholar 

  193. Roberts H. Neuroimaging techniques in cerebrovascular disease: computed tomography angiography/computed tomography perfusion. Semin Cerebrovasc Dis Stroke. 2001;1:303–16.

    Article  Google Scholar 

  194. Kamalian S, Maas MB, Goldmacher GV, et al. CT cerebral blood flow maps optimally correlate with admission diffusion-weighted imaging in acute stroke but thresholds vary by postprocessing platform. Stroke. 2011;42:1923–8.

    Article  PubMed  Google Scholar 

  195. Rother J, Jonetz-Mentzel L, Fiala A, et al. Hemodynamic assessment of acute stroke using dynamic single-slice computed tomographic perfusion imaging. Arch Neurol. 2000;57:1161–6.

    Article  PubMed  CAS  Google Scholar 

  196. Koenig M, Kraus M, Theek C, Klotz E, Gehlen W, Heuser L. Quantitative assessment of the ischemic brain by means of perfusion-related parameters derived from perfusion CT. Stroke. 2001;32:431–7.

    Article  PubMed  CAS  Google Scholar 

  197. Sorensen AG. What is the meaning of quantitative CBF? AJNR Am J Neuroradiol. 2001;22:235–6.

    PubMed  CAS  Google Scholar 

  198. Tomandl BF, Klotz E, Handschu R, et al. Comprehensive imaging of ischemic stroke with multisection CT. Radiographics. 2003;23:565–92.

    Article  PubMed  Google Scholar 

  199. Wintermark M, Flanders AE, Velthuis B, et al. Perfusion-CT assessment of infarct core and penumbra: receiver operating characteristic curve analysis in 130 patients suspected of acute hemispheric stroke. Stroke. 2006;37:979–85.

    Article  PubMed  Google Scholar 

  200. Eastwood JD, Lev MH, Azhari T, et al. CT perfusion scanning with deconvolution analysis: pilot study in patients with acute middle cerebral artery stroke. Radiology. 2002;222:227–36.

    Article  PubMed  Google Scholar 

  201. Mayer TE, Hamann GF, Baranczyk J, et al. Dynamic CT perfusion imaging of acute stroke. AJNR Am J Neuroradiol. 2000;21:1441–9.

    PubMed  CAS  Google Scholar 

  202. Wintermark M, Reichhart M, Thiran JP, et al. Prognostic accuracy of cerebral blood flow measurement by perfusion computed tomography, at the time of emergency room admission, in acute stroke patients. Ann Neurol. 2002;51:417–32.

    Article  PubMed  Google Scholar 

  203. Chamorro A, Sacco RL, Mohr JP, et al. Clinical-computed tomographic correlations of lacunar infarction in the Stroke Data Bank. Stroke. 1991;22:175–81.

    Article  PubMed  CAS  Google Scholar 

  204. Derex L, Tomsick TA, Brott TG, et al. Outcome of stroke patients without angiographically revealed arterial occlusion within four hours of symptom onset. AJNR Am J Neuroradiol. 2001;22:685–90.

    PubMed  CAS  Google Scholar 

  205. Ezzeddine MA, Lev MH, McDonald CT, et al. CT angiography with whole brain perfused blood volume imaging: added clinical value in the assessment of acute stroke. Stroke. 2002;33:959–66.

    Article  PubMed  Google Scholar 

  206. Koroshetz WJ, Lev MH. Contrast computed tomography scan in acute stroke: “You can’t always get what you want but…you get what you need”. Ann Neurol. 2002;51:415–6.

    Article  PubMed  Google Scholar 

  207. Cenic A, Nabavi DG, Craen RA, Gelb AW, Lee TY. Dynamic CT measurement of cerebral blood flow: a validation study. AJNR Am J Neuroradiol. 1999;20:63–73.

    PubMed  CAS  Google Scholar 

  208. Campbell BC, Christensen S, Levi CR, et al. Cerebral blood flow is the optimal CT perfusion parameter for assessing infarct core. Stroke. 2011;42(12):3435–40.

    Article  PubMed  Google Scholar 

  209. Wintermark M, Reichhart M, Cuisenaire O, et al. Comparison of admission perfusion computed tomography and qualitative diffusion- and perfusion-weighted magnetic resonance imaging in acute stroke patients. Stroke. 2002;33:2025–31.

    Article  PubMed  CAS  Google Scholar 

  210. Zhao L, Barlinn K, Bag AK, et al. Computed tomography perfusion prognostic maps do not predict reversible and irreversible neurological dysfunction following reperfusion therapies. Int J Stroke. 2011;6:544–6.

    Article  PubMed  Google Scholar 

  211. Lev MH, Farkas J, Rodriguez VR, et al. CT angiography in the rapid triage of patients with hyperacute stroke to intraarterial thrombolysis: accuracy in the detection of large vessel thrombus. J Comput Assist Tomogr. 2001;25:520–8.

    Article  PubMed  CAS  Google Scholar 

  212. Verro P, Tanenbaum LN, Borden NM, Sen S, Eshkar N. CT angiography in acute ischemic stroke: preliminary results. Stroke. 2002;33:276–8.

    Article  PubMed  CAS  Google Scholar 

  213. Wildermuth S, Knauth M, Brandt T, Winter R, Sartor K, Hacke W. Role of CT angiography in patient selection for thrombolytic therapy in acute hemispheric stroke. Stroke. 1998;29:935–8.

    Article  PubMed  CAS  Google Scholar 

  214. Graf J, Skutta B, Kuhn FP, Ferbert A. Computed tomographic angiography findings in 103 patients following vascular events in the posterior circulation: potential and clinical relevance. J Neurol. 2000;247:760–6.

    Article  PubMed  CAS  Google Scholar 

  215. Wintermark M, Meuli R, Browaeys P, et al. Comparison of CT perfusion and angiography and MRI in selecting stroke patients for acute treatment. Neurology. 2007;68:694–7.

    Article  PubMed  CAS  Google Scholar 

  216. Coutts SB, Lev MH, Eliasziw M, et al. ASPECTS on CTA source images versus unenhanced CT: added value in predicting final infarct extent and clinical outcome. Stroke. 2004;35:2472–6.

    Article  PubMed  Google Scholar 

  217. Camargo EC, Furie KL, Singhal AB, et al. Acute brain infarct: detection and delineation with CT angiographic source images versus nonenhanced CT scans. Radiology. 2007;244:541–8.

    Article  PubMed  Google Scholar 

  218. Li F, Silva MD, Sotak CH, Fisher M. Temporal evolution of ischemic injury evaluated with diffusion-, perfusion-, and T2-weighted MRI. Neurology. 2000;54:689–96.

    Article  PubMed  CAS  Google Scholar 

  219. Moseley ME, Cohen Y, Mintorovitch J, et al. Early detection of regional cerebral ischemia in cats: comparison of diffusion- and T2-weighted MRI and spectroscopy. Magn Reson Med. 1990;14:330–46.

    Article  PubMed  CAS  Google Scholar 

  220. Kunst MM, Schaefer PW. Ischemic stroke. Radiol Clin North Am. 2011;49:1–26.

    Article  PubMed  Google Scholar 

  221. Davis DP, Robertson T, Imbesi SG. Diffusion-weighted magnetic resonance imaging versus computed tomography in the diagnosis of acute ischemic stroke. J Emerg Med. 2006;31:269–77.

    Article  PubMed  Google Scholar 

  222. Radiological Society of North America. Diffusion imaging: from basic physics to practical imaging. RSNA, 1999. 2007. http://ej.rsna.org/ej3/0095-98.fin/index.htm. Accessed 17 Feb 2007

  223. Schaefer PW, Grant PE, Gonzalez RG. Diffusion-weighted MR imaging of the brain. Radiology. 2000;217:331–45.

    PubMed  CAS  Google Scholar 

  224. Lansberg MG, Thijs VN, O’Brien MW, et al. Evolution of apparent diffusion coefficient, diffusion-weighted, and T2-weighted signal intensity of acute stroke. AJNR Am J Neuroradiol. 2001;22:637–44.

    PubMed  CAS  Google Scholar 

  225. Lovblad KO, Bassetti C, Schneider J, et al. Diffusion-weighted MR in cerebral venous thrombosis. Cerebrovasc Dis. 2001;11:169–76.

    Article  PubMed  CAS  Google Scholar 

  226. Sitburana O, Koroshetz WJ. Magnetic resonance imaging: implication in acute ischemic stroke management. Curr Atheroscler Rep. 2005;7:305–12.

    Article  PubMed  Google Scholar 

  227. Gass A, Ay H, Szabo K, Koroshetz WJ. Diffusion-weighted MRI for the “small stuff”: the details of acute cerebral ischaemia. Lancet Neurol. 2004;3:39–45.

    Article  PubMed  Google Scholar 

  228. Easton JD, Saver JL, Albers GW, et al. Definition and evaluation of transient ischemic attack: a scientific statement for healthcare professionals from the American Heart Association/American Stroke Association Stroke Council; Council on Cardiovascular Surgery and Anesthesia; Council on Cardiovascular Radiology and Intervention; Council on Cardiovascular Nursing; and the Interdisciplinary Council on Peripheral Vascular Disease. The American Academy of Neurology affirms the value of this statement as an educational tool for neurologists. Stroke. 2009;40:2276–93.

    Article  PubMed  Google Scholar 

  229. Ay H, Koroshetz WJ, Benner T, et al. Transient ischemic attack with infarction: a unique syndrome? Ann Neurol. 2005;57:679–86.

    Article  PubMed  Google Scholar 

  230. Baird AE, Warach S. Magnetic resonance imaging of acute stroke. J Cereb Blood Flow Metab. 1998;18:583–609.

    Article  PubMed  CAS  Google Scholar 

  231. Schwamm LH, Koroshetz WJ, Sorensen AG, et al. Time course of lesion development in patients with acute stroke: serial diffusion- and hemodynamic-weighted magnetic resonance imaging. Stroke. 1998;29:2268–76.

    Article  PubMed  CAS  Google Scholar 

  232. Fiehler J, Knudsen K, Kucinski T, et al. Predictors of apparent diffusion coefficient normalization in stroke patients. Stroke. 2004;35:514–9.

    Article  PubMed  Google Scholar 

  233. Desmond PM, Lovell AC, Rawlinson AA, et al. The value of apparent diffusion coefficient maps in early cerebral ischemia. AJNR Am J Neuroradiol. 2001;22:1260–7.

    PubMed  CAS  Google Scholar 

  234. Kidwell CS, Saver JL, Mattiello J, et al. Thrombolytic reversal of acute human cerebral ischemic injury shown by diffusion/perfusion magnetic resonance imaging. Ann Neurol. 2000;47:462–9.

    Article  PubMed  CAS  Google Scholar 

  235. Thijs VN, Somford DM, Bammer R, Robberecht W, Moseley ME, Albers GW. Influence of arterial input function on hypoperfusion volumes measured with perfusion-weighted imaging. Stroke. 2004;35:94–8.

    Article  PubMed  Google Scholar 

  236. Rivers CS, Wardlaw JM, Armitage PA, et al. Do acute diffusion- and perfusion-weighted MRI lesions identify final infarct volume in ischemic stroke? Stroke. 2006;37:98–104.

    Article  PubMed  CAS  Google Scholar 

  237. Barber PA, Davis SM, Darby DG, et al. Absent middle cerebral artery flow predicts the presence and evolution of the ischemic penumbra. Neurology. 1999;52:1125–32.

    Article  PubMed  CAS  Google Scholar 

  238. Staroselskaya IA, Chaves C, Silver B, et al. Relationship between magnetic resonance arterial patency and perfusion-diffusion mismatch in acute ischemic stroke and its potential clinical use. Arch Neurol. 2001;58:1069–74.

    Article  PubMed  CAS  Google Scholar 

  239. Seitz RJ, Meisel S, Moll M, Wittsack HJ, Junghans U, Siebler M. Partial rescue of the perfusion deficit area by thrombolysis. J Magn Reson Imaging. 2005;22:199–205.

    Article  PubMed  Google Scholar 

  240. Sandhu GS, Parikh PT, Hsu DP, Blackham KA, Tarr RW, Sunshine JL. Outcomes of intra-arterial thrombolytic treatment in acute ischemic stroke patients with a matched defect on diffusion and perfusion MR images. J Neurointerv Surg. 2012;4(2):105–9.

    Article  PubMed  Google Scholar 

  241. Sobesky J, Zaro Weber O, Lehnhardt FG, et al. Which time-to-peak threshold best identifies penumbral flow? A comparison of perfusion-weighted magnetic resonance imaging and positron emission tomography in acute ischemic stroke. Stroke. 2004;35:2843–7.

    Article  PubMed  CAS  Google Scholar 

  242. Kidwell CS, Alger JR, Saver JL. Beyond mismatch: evolving paradigms in imaging the ischemic penumbra with multimodal magnetic resonance imaging. Stroke. 2003;34:2729–35.

    Article  PubMed  Google Scholar 

  243. Sorensen AG, Copen WA, Ostergaard L, et al. Hyperacute stroke: simultaneous measurement of relative cerebral blood volume, relative cerebral blood flow, and mean tissue transit time. Radiology. 1999;210:519–27.

    PubMed  CAS  Google Scholar 

  244. Parsons MW, Yang Q, Barber PA, et al. Perfusion magnetic resonance imaging maps in hyperacute stroke: relative cerebral blood flow most accurately identifies tissue destined to infarct. Stroke. 2001;32:1581–7.

    Article  PubMed  CAS  Google Scholar 

  245. Schaefer PW, Hunter GJ, He J, et al. Predicting cerebral ischemic infarct volume with diffusion and perfusion MR imaging. AJNR Am J Neuroradiol. 2002;23:1785–94.

    PubMed  Google Scholar 

  246. Neumann-Haefelin T, Wittsack HJ, Wenserski F, et al. Diffusion- and perfusion-weighted MRI. The DWI/PWI mismatch region in acute stroke. Stroke. 1999;30:1591–7.

    Article  PubMed  CAS  Google Scholar 

  247. Chalela JA, Kang DW, Luby M, et al. Early magnetic resonance imaging findings in patients receiving tissue plasminogen activator predict outcome: insights into the pathophysiology of acute stroke in the thrombolysis era. Ann Neurol. 2004;55:105–12.

    Article  PubMed  Google Scholar 

  248. Prince MR, Arnoldus C, Frisoli JK. Nephrotoxicity of high-dose gadolinium compared with iodinated contrast. J Magn Reson Imaging. 1996;6:162–6.

    Article  PubMed  CAS  Google Scholar 

  249. Murphy KP, Szopinski KT, Cohan RH, Mermillod B, Ellis JH. Occurrence of adverse reactions to gadolinium-based contrast material and management of patients at increased risk: a survey of the American Society of Neuroradiology Fellowship Directors. Acad Radiol. 1999;6:656–64.

    Article  PubMed  CAS  Google Scholar 

  250. Thomsen HS. Nephrogenic systemic fibrosis: a serious late adverse reaction to gadodiamide. Eur Radiol. 2006;16:2619–21.

    Article  PubMed  Google Scholar 

  251. Collidge TA, Thomson PC, Mark PB, et al. Gadolinium-enhanced MR Imaging and nephrogenic systemic fibrosis: Retrospective Study of a Renal Replacement Therapy Cohort. Radiology. 2007;245(1):168–75.

    Article  PubMed  Google Scholar 

  252. Kuo PH, Kanal E, Abu-Alfa AK, Cowper SE. Gadolinium-based MR contrast agents and nephrogenic systemic fibrosis. Radiology. 2007;242:647–9.

    Article  PubMed  Google Scholar 

  253. Cowper SE, Boyer PJ. Nephrogenic systemic fibrosis: an update. Curr Rheumatol Rep. 2006;8:151–7.

    Article  PubMed  Google Scholar 

  254. Stenver DI. Investigation of the safety of MRI contrast medium Omniscan. Danish Medicines Agency. 2006. http://www.dkma.dk/1024/visUKLSArtikel.asp?artikelID=8931. Published 29 May 2006. Accessed 7 Dec 2006.

  255. Cowper SE. Nephrogenic systemic fibrosis: the nosological and conceptual evolution of nephrogenic fibrosing dermopathy. Am J Kidney Dis. 2005;46:763–5.

    Article  PubMed  Google Scholar 

  256. Barkovich AJ, Atlas SW. Magnetic resonance imaging of intracranial hemorrhage. Radiol Clin North Am. 1988;26:801–20.

    PubMed  CAS  Google Scholar 

  257. Hermier M, Nighoghossian N. Contribution of susceptibility-weighted imaging to acute stroke assessment. Stroke. 2004;35:1989–94.

    Article  PubMed  Google Scholar 

  258. Kidwell CS, Saver JL, Villablanca JP, et al. Magnetic resonance imaging detection of microbleeds before thrombolysis: an emerging application. Stroke. 2002;33:95–8.

    Article  PubMed  Google Scholar 

  259. Fiehler J, Albers GW, Boulanger JM, et al. Bleeding risk analysis in stroke imaging before thromboLysis (BRASIL): pooled analysis of T2*-weighted magnetic resonance imaging data from 570 patients. Stroke. 2007;38:2738–44.

    Article  PubMed  Google Scholar 

  260. Muir KW, Weir CJ, Murray GD, Povey C, Lees KR. Comparison of neurological scales and scoring systems for acute stroke prognosis. Stroke. 1996;27:1817–20.

    Article  PubMed  CAS  Google Scholar 

  261. Brott T, Adams Jr HP, Olinger CP, et al. Measurements of acute cerebral infarction: a clinical examination scale. Stroke. 1989;20:864–70.

    Article  PubMed  CAS  Google Scholar 

  262. Goldstein LB, Bertels C, Davis JN. Interrater reliability of the NIH stroke scale. Arch Neurol. 1989;46:660–2.

    Article  PubMed  CAS  Google Scholar 

  263. Adams Jr HP, Davis PH, Leira EC, et al. Baseline NIH stroke scale score strongly predicts outcome after stroke: a report of the Trial of Org 10172 in Acute Stroke Treatment (TOAST). Neurology. 1999;53:126–31.

    Article  PubMed  CAS  Google Scholar 

  264. Kwiatkowski TG, Libman RB, Frankel M, et al. Effects of tissue plasminogen activator for acute ischemic stroke at one year. N Engl J Med. 1999;340:1781–7.

    Article  PubMed  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 1: Primer on Imaging in Stroke by Joel K. Curé, M.D.

Imaging goals in ischemic stroke:

  1. 1.

    Confirming a diagnosis of ischemic stroke and exclusion of non-vascular (e.g. tumor) causes of the clinical ictus

  2. 2.

    Exclusion of hemorrhage and estimation of the risk of hemorrhagic transformation

  3. 3.

    Selection of patients for reperfusion therapy by distinguishing ischemic but viable (i.e., the penumbra) from infarcted tissue and excluding those for whom the therapeutic risk exceeds the anticipated benefit

  4. 4.

    Identifying large-vessel occlusion that may complicate or represent a target for therapy

The 2007 Guidelines for the Management of Patients with Acute Ischemic Stroke noted: “Both CT and MRI are options for imaging the brain, but for most cases and at most institutions, CT remains the most practical initial brain imaging test”.145 This may change with increasing availability of MRI. MRI is more sensitive than CT for detection of acute infarction, is able to detect both acute and chronic hemorrhage as effectively as CT, and demonstrates higher interobserver and intraobserver reliability than CT for ischemic stroke diagnosis, even in readers with little experience.146

Stroke location and distribution at imaging reflects mechanism.147 Most strokes are thromboembolic and their imaging appearance reflects the territory of the occluded artery less any portion of that territory that is adequately perfused by collateral blood supply. Solitary or multiple unilateral cortical or cortical/subcortical infarcts may be secondary to either cardiogenic emboli or large arterial occlusion. Cardiogenic emboli typically account for bilateral acute infarcts in both the anterior and posterior circulations, especially in the absence of definable intracranial arterial occlusions on CTA, MRA, or transcranial Doppler. However, multiple synchronous infarcts may be encountered in patients with occlusive vasculopathies (e.g. CNS vasculitis) or coagulopathies. Lacunar infarcts, due to small arterial occlusion, are typically small (<1.5 cm) and imaging abnormalities correspond to the territory of the occluded perforating artery. These infarcts most commonly occur in the basal ganglia, thalamus, brainstem, or deep cerebellar white matter. Arterial border zone infarcts occur in regions of brain that lie between major arterial territories. These include deep cerebral hemispheric regions such as the centrum semiovale, corona radiata, and cortical zones between the ACA and MCA and the MCA and PCA territories. Arterial border zone infarcts may occur bilaterally in cases of global cerebral hypoperfusion or unilaterally in patients with severe ICA stenosis or MCA stenosis plus A1 segment hypoplasia.

1.1 Noncontrast CT Diagnosis of Acute Infarction

In CT scan interpretation the terms, “hypoattenuation” and “hyperattenuation” are preferred to “hypodense” and “hyperdense”. Attenuation indicates the degree of x-ray absorption that occurs within tissue. In patients with stroke, hypoattenuating tissue tends to be edematous, and hyperattenuating tissue tends to be hemorrhagic.

Brain edema associated with hemispheric stroke may be detectable within 1–2 h of stroke onset. CT identifies ischemic lesions with a sensitivity of 65% and a specificity of 90% within 6 h of stroke onset.148 However, the sensitivity of CT for acute ischemic stroke within the first 3 h of symptom onset has been reported to be as low as 7%.146 CT is insensitive for small acute infarcts, especially in the posterior fossa149 and is less sensitive than diffusion weighted MRI in the acute setting.150 The peak period for identifying brain ischemia on CT is 3–10 days after the ictus, well beyond the thrombolytic time window. The value of early CT in acute stroke is therefore not chiefly diagnostic, but prognostic. A large hypoattenuating area detected within 6 h of stroke onset is an indication of irreversible tissue injury,148 portends an increased risk of hemorrhagic transformation in patients treated with rTPA,151 and is associated with an increased risk of fatal brain edema.152 ECASS-1 and subsequent analyses of its CT and patient data led to the “one third rule”. Patients with CT-identified early ischemic changes (EIC) involving less than one third of the MCA territory had an improved functional outcome after IV thrombolysis compared to patients with EIC in more than one third of the MCA territory or who had no EIC on CT.153 However, the unreliability of volume estimation with the one third rule and lack of demonstrable evidence of an effect on treatment modification led to development of the ASPECTS scoring system. This scoring system assigns one point each to ten regions within the MCA territory. A point is deducted for each of the ten regions demonstrating EIC. In patients undergoing intravenous thrombolysis at less than 3 h from symptom onset, a baseline ASPECTS score less than or equal to seven predicted patients who were unlikely to achieve independent functional outcome.154 Sensitivity of CT for acute intracranial hemorrhage approaches 100%.155

1.1.1 Early CT Finding That Suggest Infarction

  1. 1.

    Loss of gray–white differentiation may be detected within 6 h of onset of stroke symptoms in 82% of patients with MCA territory ischemia.156 Cytotoxic edema reduces the attenuation of gray matter into the range of white matter, thereby decreasing gray-white matter contrast.

    1. (a)

      Insular ribbon sign”: Loss of gray-white matter differentiation in the insular cortex can be an early sign of MCA ischemia.157

    2. (b)

      Obscuration of the lentiform nucleus reflects cytotoxic edema within the involved basal ganglia, again decreasing gray–white matter contrast.158

  2. 2.

    Cortical sulcal effacement due to swelling of edematous gyri.

  3. 3.

    MCA sign.” Hyperattenuation of the M1 segment (or other intracranial arteries, e.g. the posterior cerebral artery) due to thromboembolism.159161(Fig. 9.4)

    Fig. 9.4
    figure 4

    MCA sign. Hyperattenuation of the M1 segment due to thromboembolism (arrow).

  4. 4.

    Sylvian dot sign. Distal MCA (M2 or M3 branches) occlusion indicated by hyperattenuation in the Sylvian fissure.162 Sensitivity 38%, specificity 100%, positive predictive value 100%, negative predictive value 68%.163

The combined presence of the insular ribbon sign (Fig. 9.5), hemispheric sulcal effacement, and decreased attenuation of the lentiform nucleus is predictive of ICA occlusion.164

Fig. 9.5
figure 5

Insular ribbon sign. Loss of gray-white differentiation in the insula (arrow heads) can be an early sign of MCA territory ischaemia.

Since “time is brain”, it is held that eligible patients presenting with acute stroke within the 4.5 h time window should receive intravenous tPA prior to performance of additional advanced imaging (including vascular or “penumbral” imaging with CT or MR perfusion techniques).165 The identification of ischemic penumbra may be useful in three scenarios:

  1. 1)

    Offering treatment to patients who do not qualify for treatment under current guidelines (e.g. beyond the “time window”).

  2. 2)

    Identifying patients for whom treatment within the current time window is likely to be futile.

  3. 3)

    Identifying IV-tPA non-responders to whom endovascular therapies might be offered.166

1.2 CT Perfusion

CT perfusion provides quantitative data about CBF, and is becoming widely available on multidetector CT scanners. CT perfusion involves repeated (“cine”) helical CT imaging of the brain during the transit of an injected bolus of iodinated contrast through the intracranial vasculature. Measurements of the change in tissue attenuation during passage of the contrast bolus are used to generate quantitative information about CBF as well as CBV and time-to-peak (TTP) or mean transit time (MTT). Acquisition and processing of the data are accomplished seconds to minutes. The concept of CT perfusion was introduced more than 20 years ago,167 but had to await the development of high-speed helical CT scanners, fast computers, and software capable of rapid data analysis to make the technique clinically useful.

1.3 Normal Values of CBF and CBV

Cerebral blood flow is normally maintained within a narrow range by autoregulation. Normal CBF is approximately 80 mL per 100 g/min in human gray matter and approximately 20 mL per 100 g/min in white matter. Global CBF, as well as average CBF in the cortical mantle, which is roughly a 50:50 mix of gray matter and white matter, is approximately 50 mL per 100 g/min. Protein synthesis in neurons ceases when CBF falls below 35 mL per 100 g/min.168 At a CBF ≤20 mL per 100 g/min, however, electrical failure occurs and synaptic transmission between neurons is disturbed, leading to loss of function of still-viable neurons.169171 Metabolic failure and cell death occur at CBF ≤12 mL per 100 g/min.170 CBV is defined as the amount of blood in a given quantity of brain tissue. Normal CBV is approximately 4–5 mL per 100 g.172,173 CBV can be decreased or increased during cerebral ischemia, depending on the efficacy of cerebral autoregulation and patency of collateral arterial pathways.173176

1.3.1 CT Perfusion Technique

1.3.1.1 Parameters

CT perfusion produces the following data:

  1. 1.

    Cerebral blood flow (CBF), measured in mL per 100 g of brain tissue per minute (mL/100 g/min) or as mL per 100 mL of brain tissue per minute (mL/100 mL/min).

  2. 2.

    Cerebral blood volume (CBV), measured in mL/100 g or mL/100 mL.

  3. 3.

    Time to peak (TTP) is defined as the time delay (in seconds) between the first arrival of contrast within major arteries included in the section imaged and the peak attenuation of the brain tissue.

  4. 4.

    Mean transit time (MTT) indicates the time (in seconds) required for contrast material to pass from the arterial side to the venous side of the intracranial circulation. Blood and intravascular contrast material pass through vascular pathways of varying length and complexity in the brain’s vascular network. The average of all of these possible transit times is MTT.

    1. (a)

      TTP and MTT are parameters unique to CBF techniques (e.g., CT perfusion and MRI perfusion) that utilize an intravascular indicator and track the passage of the indicator through the brain over the course of time to determine CBF.

1.3.1.2 Concepts

There are two commonly applied methods of CT perfusion (Table 9.6). These methods, known as the first pass bolus tracking techniques, are based on the indicator dilution principle and provide information about CBF, CBV, and MTT, and TTP. A known amount of a nondiffusible tracer (e.g., iodinated contrast material) is injected into an antecubital vein, and its concentration is repeatedly measured during its first pass through an intracranial vessel. Contrast transit through the intracranial vasculature produces a transient change in brain tissue attenuation. This change is linearly proportional to the serum concentration of the contrast agent. With helical CT scanning, these changes can be graphed as a time-attenuation curve for every voxel in a CT-imaging slice.

Table 9.6 CT perfusion methods

Two different mathematical approaches are commonly used to calculate CT perfusion data from the time-attenuation curve: deconvolution and maximum slope. With deconvolution methodology, the attenuation values of an artery in the field of view (the arterial input function), such as the anterior cerebral artery, are integrated with time-attenuation information of the brain tissue on a voxel-by-voxel basis in a mathematical operation called deconvolution. In mathematical terms: Ct(t) = CBF · [Ca (t) ⊗ R (t)] where C t(t) is the tissue time-attenuation curve; C a(t) is the arterial time-attenuation curve; R(t) is the impulse residue function, and  ⊗  is the convolution operator. The impulse residue function is an idealized tissue time-attenuation curve that would result if the entire bolus (the impulse) of contrast material was administered instantaneously into the artery supplying a given area of the brain. The plateau of impulse residue function reflects the length of time during which the contrast material (the residue) is passing through the capillary network. Both C t(t) and C a(t) can be measured, and the deconvolution process uses the information to calculate CBF and CBV. MTT is then derived by using the central volume principle, which relates CBF, CBV, and MTT in the following relationship: CBF = CBV/MTT

The accuracy of this method depends upon an intact blood–brain barrier, as leakage of the contrast material out of the intravascular space can lead to artifactually high perfusion parameters. Accuracy can also be influenced by the choice of the reference artery177 and recirculation of contrast material. The venous output function serves as a reference against which the CTP parametric values are normalized and scaled. Since CBV values are affected by the choice of the venous output function, the chosen ROI for the venous output function should include the voxel demonstrating the maximum area under the time/attenuation curve and the least amount of partial volume averaging.178

In the maximum slope method, the maximum slope of the time-attenuation curve is used to calculate CBF (Fig. 9.6).179181 Values for CBV are calculated from the maximum-enhancement ratio, which is the maximum enhancement of the time-attenuation curve in a given voxel compared to that of the superior sagittal sinus.179,182,183 Software using this method reports TTP rather than MTT. The accuracy of this method depends on a rapid bolus injection of contrast material, because a delay in the appearance of contrast material in the intracranial vasculature will lead to a decrease in the maximum slope of the time-attenuation curve, and CBF will be underestimated.179,184

Fig. 9.6
figure 6

Maximum slope method. In CT perfusion with the maximum slope method (Siemens), regions of completed infarction show up as a “black hole” – dense, black areas on CBF, CBV, and TTP images (black arrow). Adjacent areas of abnormality on that are not black (white arrows) indicate regions of salvageable tissue. The CT done 2 weeks later shows the black hole region as a completed infarction.

1.3.2 Validation

Quantitative CBF measurement by CT perfusion has been validated by comparison to other techniques for measuring CBF such as microspheres;,185 xenon CT,186,187 and PET.188,189 CT perfusion imaging using the deconvolution technique has been shown to demonstrate little variability within individuals.188 The use of CT perfusion in the identification of cerebral ischemia has been validated in experimental ­models.190192 Further validation of CT perfusion in human subjects by comparison with other brain imaging techniques in the setting of acute stroke, has been extensive and is discussed below.

1.3.3 Limitations

CT perfusion imaging has several practical limitations. Brain regions close to the skull are difficult to image because of bone artifact. A peripheral IV is required for intravenous administration of the contrast material, which can be a nuisance for some intensive care unit patients. The study requires iodinated contrast, which can be problematic in patients with renal insufficiency or contrast allergy.

An important limitation concerns the use of an intravascular indicator in first-pass CT perfusion methods. As opposed to older techniques like xenon-CT and PET, in which diffusible tracers are used and only capillary perfusion is measured, all intracranial vessels are included in CT perfusion. This difference leads to an over-estimation of CBF in regions that include large vessels, such as around the Sylvian fissure.193 Moreover, this aspect of CT perfusion makes it difficult to compare CT perfusion results to CBF values obtained by the use of other methods. This situation may be ameliorated by vessel removal using threshold-based segmentation algorithms.178,189 Finally, variability in quantification between different CT perfusion post-processing software packages limits the ability to generalize parametric thresholds (e.g. CBF threshold representing the infarct core) between platforms.194

1.3.4 Interpretation of CT Perfusion Data

Validity has been demonstrated for the commonly used mathematical techniques for CT perfusion by comparison with other CBF measurement techniques. However, each method has inherent limitations and sources of systematic error; hence, the description of CT perfusion as being “semiquantitative” by some authors.195,196 Assessment of cerebral perfusion based on absolute values for CBF and CBV should be made with caution.197,198

In CT perfusion using the deconvolution method, some have found that MTT values >145% the contralateral hemisphere correlate best with tissue at risk for infarction in cases of persistent arterial occlusion, compared to DWI/FLAIR MRI.199 Using the maximum slope method, a reduction of CBV of 60%, compared to non-ischemic regions of the brain, best identified cerebral ischemia.196

Mean transit time is prolonged in regions of cerebral ischemia. In a series of patients with acute ischemic MCA stroke, Eastwood and colleagues found average MTT to be 7.6 s in the affected MCA territories and 3.6 s in the unaffected MCA territories.200 Areas of reduced perfusion were defined as MTT >6 s because that value represented at least three standard deviations greater than the average MTT values in unaffected MCA territories.

TTP is typically <8 s in normal brain tissue with unimpaired antegrade flow. In ischemic regions, TTP is prolonged, reflecting delayed tracer arrival through alternative pathways such as leptomeningeal vessels. TTP maps are useful for accurate identification of areas of impaired perfusion.196 A regional TTP >8 s raises the suspicion of cerebral ischemia. However, TTP maps can provide false-positive findings when TTP is prolonged due to carotid stenosis or occlusion and regional CBF is compensated for by collateral vessels.201

Both MTT and TTP maps can be used to identify cerebral ischemia. MTT maps offer advantages over CBF and CBV maps. MTT appears to be affected by ischemia at an earlier stage than CBF or CBV, although it is less specific.202 Color-coded TTP and MTT maps appear to demonstrate regions of cerebral ischemia more readily than CBF and CBV maps. TTP and MTT are usually homogenous in normal areas of brain tissue, permitting easy identification of abnormal hemodynamics.202 Moreover, CBF and CBV data are over-estimated when the ROI includes major vessels, such as MCA branches.189 In comparison, TTP and MTT do not seem to be influenced by the presence of large vessels within ROIs. The absence of regions of extended TTP or MTT is usually a reliable indication that ischemia is not present.

1.3.4.1 CT Perfusion in Ischemic Stroke

CT perfusion can be done at the same time as the initial screening CT scan in patients with acute ischemic stroke and can distinguish viable tissue from regions of completed infarction.

  1. 1.

    CT perfusion can be used to exclude poor candidates for thrombolysis, such as patients with lacunar strokes and patients with no arterial occlusions, which account for up to 25%203 and 29%204 of patients with acute stroke, respectively.

  2. 2.

    CT perfusion imaging can provide prognostic information because patients with profound, widespread ischemia can be expected to have poorer outcomes than those with borderline ischemia.202,205

CT perfusion can potentially identify salvageable tissue at risk of infarction.202,206 Using the deconvolution method, a mismatch between regional MTT, CBF, and CBV maps can indicate the presence of ischemic but potentially salvageable brain (penumbra).190,195,201,207 Studies attempting to define the parameter that best identifies the infarct core have yielded different results. In a series of patients with acute stroke studied by Wintermark and colleagues, CBV <2.0 mL per 100 g best identified the irreversibly injured infarct core. Regions demonstrating MTT >145% compared to mirror-image voxels in the contralateral hemisphere optimally conformed to the ischemic region (infarct core  +  penumbra).199 A more recent study by Campbell, et al found that CBF <31% of the mean contralateral hemispheric CBF best predicted infarct core.208

“Prognostic maps” co-demonstrating the ischemic zone (e.g. MTT  >  145% contralateral mirror image voxel values  =  core  +  penumbra) in green and the infarct core (CBV <2.0 mL per 100 g) in red can be generated to provide an at-a-glance image of these parameters (Fig. 9.6).209

Using the maximum slope method, the relative values of CBF and CBV can be used to distinguish infarcted from ischemic tissue. In a series of patients undergoing CT perfusion studies within 6 h of stroke onset, the thresholds for best discrimination between infarcted and non-infarcted tissue were 48% of normal values for CBF, and 60% of normal values for CBV.196 The lowest relative CBF and CBV values among brain regions not developing infarctions were 29% and 40% of normal values, respectively.

1.3.4.2 Validation of CT Perfusion in Acute Ischemic Stroke

The deconvolution method has been validated in the diagnosis of acute ischemic stroke by comparison to CT imaging200 and to MR T2-weighted imaging,200 diffusion imaging,202,209 and perfusion imaging (Fig. 9.7).200 In a series of patients with acute ischemic stroke, undergoing both CT perfusion and MRI diffusion studies on admission, infarct size on CBF maps correlated highly with the size of the abnormality on the diffusion-weighted imaging (DWI) map (r  =  0.968).209 Similarly, infarct size assessed by CT perfusion studies done on admission in patients with ischemic stroke, correlated highly with infarct size measured by follow-up MRI-DWI maps obtained an average of 3 days after admission (r  =  0.958).202 However, a recent study of treated patients who had complete early reperfusion found that CTP prognostic maps were not predictive for irreversibly or reversibly lost neurologic function.210

Fig. 9.7
figure 7

Deconvolution method. In deconvolution CT perfusion (General Electric, Philips), the “black hole” technique is not a reliable way to identify regions of completed infarction. Threshold maps, however, can provide the same information. Here, the ischaemic core (dark threshold area – red on the colour image) is defined as absolute CBV <2 mL per 100 g, and the penumbra (light threshold area – green on the colour image) is defined as the region of brain with MTT values 1.45 times the MTT values in the corresponding area of the opposite hemisphere.199 The follow-up CT shows an infarction that corresponds to the ischaemic core region.

The maximum slope method has been validated in acute stroke by comparison to CT, MRI, and SPECT.201 In a series of patients with acute stroke, who underwent both CT perfusion and SPECT studies on admission, the areas of ischemia indicated by CT perfusion CBF maps correlated well with those indicated by SPECT imaging (r  =  0.81).181 In a series in which ischemic areas on admission of CT perfusion images, were compared to the follow-up CT or MR images showing final infarctions, infarction was found to develop in all patients with >70% CBF reduction and in 50% of patients with 40–70% CBF reduction.201 Based on a threshold of CBF <60% (compared with CBF in normal vascular territories), CBF maps predicted the extent of infarction with high sensitivity (93%) and specificity (98%). Similarly, TTP >3 s predicted infarction with a sensitivity of 91% and a specificity of 93%. Notably, in the same study, a negative predictive value for TTP >3 s of 99% was found, indicating that the absence of extended TTP is usually accurate in excluding the presence of ischemia. In a series of CT perfusion studies done in patients with acute stroke <6 h after onset, and compared to the follow-up CT or MRI, threshold values of 48% and 60% of normal, for CBF and CBV, respectively, were found to discriminate best between the areas of infarction and the areas of non-infarction.196

1.4 CT Angiography

CT angiography (CTA) is useful in identifying large vessel occlusion, and can complement CT perfusion. The time required for acquisition, processing, and analysis of CTA studies of patients with acute ischemic stroke, averages 15 min.211 Compared to catheter angiography, CTA has sensitivity and specificity for the detection of large vessel occlusion of 98.4 and 98.1%, respectively.211 CTA may be prone to false-positive results; in two series of CTA in acute stroke, a minority of patients was found to have lesions on CTA that could not be found with catheter angiography.212,213 CTA can be particularly useful in assessment of vertebrobasilar occlusion,106 as CT perfusion imaging of the posterior circulation territory is limited because of bone artifact. However, basilar artery lesions can be better assessed with CTA than vertebral artery lesions.214 CTA combined with CT perfusion shows good agreement with MRI in the assessment of infarct size, cortical involvement, and intracranial cerebral artery occlusion.215 Finally, some authors have found application of ASPECTS.154 scoring to CTA source images a robust method (and superior to ASPECT analysis of routine non-contrast brain CT) for early detection of irreversible ischemia and prediction of final infarct volume.216,217

1.5 MRI

Magnetic resonance imaging is based on the interaction between a powerful, uniform magnetic field, radiofrequency (RF) energy, and body tissues. Protons absorb energy from pulsed RF waves (excitation) and are thereby deflected from their alignment with the main magnetic field. As the nuclei return to rest, energy is released and signals are induced in a receiver and converted into diagnostic images. During the process of energy release, spatially encoded voxel-specific relaxation constants can be obtained and, in conjunction with Fourier transform reconstruction, used to construct images that demonstrate specific tissues. A wide array of MRI imaging sequences is available (Table 9.7). Most MRI images accentuate T1 or T2 relaxation; T1 is longitudinal, or spin–lattice relaxation time and T2 is transverse, or spin-spin relaxation time. In T1-weighted images, fat has increased signal (short T1 relaxation) and water appears dark (long T1 relaxation). In T2-weighted images, water has increased signal relative to brain (long T2 relaxation). Brain tissue water content is typically increased in regions of edema, ischemia, and hemorrhage, thus changing the appearance of the tissue on MRI. T2-weighted images usually show only tissue changes caused by severe and prolonged ischemia – apparent only after some 6–24 h following stroke onset – and are therefore not optimal for evaluating acute ischemia.

Table 9.7 MRI signal characteristics of cerebral infarction

1.5.1 Diffusion-Weighted Imaging

Diffusion-weighted imaging (DWI) measures the Brownian motion of water protons in tissue. Normal random motion of water protons leads to a loss of signal on diffusion weighted images. Ischemic failure of the ATP-dependent sodium-potassium cellular membrane pumps leads to water migration from the extracellular space to the intracellular space. Random water proton motion in the constricted extracellular space is reduced. Severely ischemic brain tissue appears bright on DWI due to signal preservation in these areas of decreased Brownian motion. These changes occur within minutes after ischemic stroke (Fig. 9.8).218,219 Areas of ischemia appear bright on DWI in the acute phase and become unapparent or dark after about 2 weeks. DWI images are superior to CT and conventional MRI in the detection of acute ischemia.149 Sensitivities of 88–100% for acute stroke detection with DWI have been reported, with specificity from 86–100%.220 Analyzed pooled data from several studies yielded a PPV of 100% and a NPV of 90.6%.221

Fig. 9.8
figure 8

Patterns of acute ischemic stroke on MRI. Diffusion-weighted images of an embolic stroke (a); an arterial border zone (aka watershed territory or “rosary” pattern) stroke (b); a large artery (MCA) stroke (c); and a lacunar stroke (d).

Diffusion-weighted images are influenced by other parameters including spin density and T1 and T2 relaxation effects. Calculation of the apparent diffusion coefficient (ADC) eliminates these influences and provides “pure” diffusion information.222 Two otherwise identical image sets are obtained, one with a low (but non-zero) b value and one with a b value  =  1,000 s/mm2. The natural logarithm of signal intensity vs. b value for these two values is plotted and the slope of this line is used to determine ADC for each voxel in the image.223 The resulting “map” demonstrates the calculated ADC for each pixel in the image, with signal intensity proportional to the magnitude of the ADC. Areas of acute infarction (restrained diffusion) appear bright on diffusion weighted images and have low ADC values (appear dark) on ADC maps. Subacutely, signal on diffusion weighted images within infarcted areas may appear bright due to “T2 shine through”, but correlation with the ADC maps will demonstrate that this is a T2 effect, and does not reflect true diffusion restraint. After about 2 weeks, diffusion becomes facilitated. Signal in the infarcted area decreases on DWI and increases on ADC maps as a result. Decreased ADC values indicate with good sensitivity (88%) and specificity (90%) that, an infarct is less than 10 days old.224 Venous infarctions, in contrast, cause an increase in ADC values in the acute phase because of vasogenic edema, although in later stages, the ADC map appearance becomes complex because of the coexistence of cytotoxic and vasogenic edema and the presence of hemorrhage.225

Diffusion-weighted imaging can be useful in the workup of patients with TIA.226 “Dots of hyperintensity” on DWI, indicating microinfarctions that are too small to cause permanent neurological symptoms, are found in some 40–50% of patients with traditionally defined TIA.227 This has led to a recommendation for a change of the definition of TIA from a clinical to a tissue-based definition, specifically: “A transient episode of neurological dysfunction caused by focal brain, spinal cord, or retinal ischemia, without acute infarction”.228 Patients with clinically transient neurological events in whom asymptomatic diffusion abnormalities are discovered have a high risk of early completed stroke.229

Diffusion-weighted imaging and ADC maps are dynamic. Areas of ischemic injury may enlarge by 43% in the first 52 h after onset,230 although in most patients, lesion size appears to reach a maximum by 24 h.231 Conversely, all areas demonstrating DWI hyperintensity and ADC map hypointensity may not necessarily be infarcted, as bright regions on DWI can be reversed by reperfusion. In a series of patients with acute stroke, 19.7% demonstrated “normalization” of ADC abnormalities after reperfusion.232 Tissue with ADC values 75–90% of ADC values in normal brain are likely to proceed to infarction, whereas tissue with ADC values >90% of normal are more likely to recover.233 Nevertheless, DWI hyperintensity is a necessary stage on the path to infarction,226 and the volume of DWI abnormalities do correlate with clinical severity.230,234

1.5.2 Perfusion Imaging

MRI perfusion imaging employs a first pass tracking technique and ­deconvolution method for calculating the brain perfusion parameters. MRI perfusion uses the same deconvolution technique as CT perfusion, which is described in detail above. In MRI perfusion, a bolus of gadolinium is injected rapidly into a peripheral vein, and tissue- and arterial-input curves are used to generate CBF, CBV, TTP and MTT images. The information is not quantitative because, the MR signal change after IV administration of gadolinium is not proportionally related to the plasma concentration of gadolinium. MRI perfusion is subjected to many of the limitations of CT perfusion, such as the dependence of lesion volume on arterial input function selection235, and controversy about the optimal perfusion parameters for the identification of affected tissue.236 The value of MRI perfusion imaging lies in the perfusion–diffusion mismatch hypothesis. This holds that abnormal regions on perfusion images that appear normal on diffusion weighted imaging, are equal to the penumbra, and represent potentially salvageable tissue. A perfusion–diffusion mismatch pattern is present in some 70% of patients with anterior circulation stroke scanned within 6 h of onset,237 is strongly associated with proximal MCA occlusion,237 and resolves on reperfusion.238,239 A recent study reported low favorable clinical responses and high mortality rates in a small group of patients (N  =  8) with matched perfusion and diffusion abnormalities who underwent attempted intraarterial thrombolysis, especially those with large infarcts.240

The penumbra on perfusion imaging has been defined as regions where DWI is normal and TTP >4 s,241 although, for practical purposes, any region that is abnormal on perfusion imaging but normal on DWI may represent salvageable tissue.242 Among MRI perfusion parameters, CBF, MTT and TTP appear to best identify all affected tissue (and thereby distinguish penumbra when compared to DWI),243,244 and CBV seems to best predict the final infarct volume.245 Compared to final infarct volumes imaged on MRI, the sensitivities of CBF, CBV, and MTT for detection of perfusion abnormalities were 84%, 74%, and 84%, respectively, and the specificities were 96%, 100%, and 96%, respectively.245

Together, perfusion imaging and DWI can identify tissue that is at the risk of infarction but amenable to salvage with revascularization.246 In a series of patients receiving IV thrombolytics for acute ischemic stroke and imaged both before and after 2 h of treatment, 78% of patients had complete resolution of perfusion lesions and 41% had resolution of DWI lesions.247 Perfusion–diffusion imaging has been used in clinical trials to select patients for thrombolysis. Intravenous desmoplase was given only to patients with a DWI-PWI mismatch ≥20%, and the drug was found to be potentially effective in improving clinical outcomes.19,20

1.6 MR Angiography

MRA techniques fall into three categories:

  1. 1)

    Time of flight. Very common MRA technique.

    1. a)

      Depends on a strong signal from blood flowing into a plane where stationary tissue signal has been saturated.

      1. i)

        Advantage: No contrast agent is used.

      2. ii)

        Disadvantages: Spin dephasing in areas of turbulent flow or magnetic susceptibility (near paramagnetic blood products, ferromagnetic objects, and air/bone interfaces) causes signal loss that may lead to overestimation of stenosis.

  2. 2)

    Phase contrast. Not often used.

    1. a)

      Image contrast from the differences in phase accumulated by moving spins in a magnetic field gradient. Stationary spins accumulate no net phase.

      1. i)

        Advantages: No contrast agent is used. Less likely to confuse fresh clot for flowing blood as it is strictly flow-dependent.

      2. ii)

        Disadvantages: Acquisition times are relatively long.

  3. 3)

    Contrast-enhanced MRA. Common MRA technique.

    1. a)

      Based on a combination of rapid 3D imaging and the T1-shortening effect of IV gadolinium.

      1. i)

        Advantages: High signal-to-noise ratios, robustness irrespective of blood flow patterns or velocities, and fast image acquisition, allowing for the evaluation of larger anatomic segments (from the aortic arch to the circle of Willis).

      2. ii)

        Disadvantage: Requires IV gadolinium, which carries a small risk of complications, particularly in patients with renal insufficiency (see below).

1.6.1 Gadolinium and Nephrogenic Systemic Fibrosis

Gadolinium is a chemical element with an atomic number of 64. It has seven unpaired electrons in its outer shell which hasten T1 relaxation and increase signal in the area of interest. Gadolinium alone is toxic, but not when combined with a chelating agent. Several FDA-approved gadolinium preparations are available. A study of high-dose gadolinium administration in a population with a high prevalence of baseline renal insufficiency, showed no renal failure associated with its administration.248 The rate of anaphylactic reactions is also very low; in a survey of >700,000 patients receiving gadolinium, the rate of serious allergic reactions was <0.01% and most reactions were limited to mild nausea or urticaria.249

Nephrogenic systemic fibrosis (aka nephrogenic fibrosing dermopathy) is strongly associated with gadodiamide (Omniscan™; GE Healthcare, Princeton, NJ).250,251 Although most patients have a history of exposure to gadoliamide, other gadolinium-based agents have been implicated.252 It appears to occur only in patients with renal insufficiency, generally in those requiring dialysis,252 and is dose-dependent.251 The condition consists of thickening and hardening of the skin of the extremities, due to increased skin deposition of collagen. The condition may develop rapidly and result in wheelchair-dependence within weeks. There may also be involvement of other tissue such as the lungs, skeletal muscle, heart, diaphragm, and esophagus.253 The mechanism is not understood. An estimate of the incidence of this syndrome comes from an internet-based medical advisory originating in Denmark, which reported that, of about 400 patients with severely impaired renal function, 5% were subsequently diagnosed with nephrogenic systemic fibrosis.254

Management consists of correction of renal function (usually dialysis), which may result in a cessation or reversal of symptoms.255

1.6.2 Identification of Hemorrhage on MRI

MRI is as sensitive for acute hemorrhage as CT.146 The appearance of intracranial hemorrhage changes with time as the hemoglobin moiety changes from non-paramagnetic oxyhemoglobin through the paramagnetic forms (deoxyhemoglobin, methemoglobin, and hemosiderin). Subacute blood appearing hyperintense on T1weighted images is in the methemoglobin form. The characteristic T1 shortening here is due to a phenomenon known as “dipole-dipole relaxation enhancement (PEDDRE).” T2 shortening (and the associated signal loss) depend on the presence of an intact cell membrane sequestering paramagnetic hemoglobin moieties from the extracellular space and thereby establishing a local magnetic gradient. Red blood cells usually undergo lysis in the subacute phase (i.e. methemoglobin) of parenchymal hemorrhage evolution. Early in the pre-lysis phase, blood appears “bright” on T1 (PEDDRE) and “dark” on T2 weighted images (paramagnetic effect). After RBC lysis, methemoglobin-dominant hematomas still appear bright on T1 (again, PEDDRE), but become “bright” on T2 weighted images due to disruption of the paramagnetic effect by RBC lysis. Deoxyhemoglobin (acute) and hemosiderin (chronic) share similar appearances on T1 (isointense to gray matter) and T2 (hypointense to gray matter) MRI. However, acute hemorrhage is typically associated with vasogenic edema, while chronic hemorrhage is not (the latter may associated with cavitation, gliosis, and focal atrophy). The recurrence of T2 shortening in chronic (hemosiderin) hematomas long after RBC lysis is due to the ingestion of hemosiderin by macrophages.256

Acute hemorrhage characteristics on MRI are summarized in Table 9.8. Susceptibility weighted MRI can help identify acute cerebral hemorrhage, “microbleeds,” and intravascular clot.257 Asymptomatic microbleeds are caused by hypertension and amyloid angiopathy, and are found in up to 6% of elderly patients and 26% of patients with prior ischemic stroke.226 The finding of microbleeds in patients with acute ischemic stroke, may predict an increased risk of hemorrhage transformation after thrombolysis. In a study of patients undergoing IA thrombolysis for acute ischemic stroke, microbleeds were found in 12% of patients prior to treatment.258 Symptomatic hemorrhages occurred in 20% of patients with an evidence of prior microbleeds, compared to 11% of patients without prior microbleeds. The Bleeding Risk Analysis in Stroke Imaging Before Thrombolysis (BRASIL) study found that the risk of intracranial hemorrhage attributable to microbleeds was small and unlikely to exceed the benefits of thrombolytic therapy. This study could not draw conclusions about the risk of hemorrhage in patients with multiple microbleeds, however.259

Table 9.8 MRI signal characteristics of cerebral haemorrhage

Appendix 2: NIH Stroke Scale

The National Institutes of Health Stroke Scale (NIHSS) is widely used and it provides important prognostic information.260263

A detailed description of the NIHSS can be downloaded at www.ninds.nih.gov/disorders/stroke/strokescales.htm

Higher scores indicate greater stroke severity (Tables 9.9 and 9.10). A score of ≥16 predicts a high probability of death or severe disability whereas a score of ≥6 predicts a good recovery.263 Some 60–70% of acute ischaemic stroke patients with a baseline NIHSS score <10 will have a favourable outcome after 1 year, compared to only 4–16% of patients with a score >20.264

Table 9.9 NIH stroke scale
Table 9.10 NIH stroke scale score severity

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Science+Business Media New York

About this chapter

Cite this chapter

Harrigan, M.R., Deveikis, J.P. (2013). Treatment of Acute Ischaemic Stroke. In: Handbook of Cerebrovascular Disease and Neurointerventional Technique. Contemporary Medical Imaging, vol 1. Humana Press, Totowa, NJ. https://doi.org/10.1007/978-1-61779-946-4_9

Download citation

  • DOI: https://doi.org/10.1007/978-1-61779-946-4_9

  • Published:

  • Publisher Name: Humana Press, Totowa, NJ

  • Print ISBN: 978-1-61779-945-7

  • Online ISBN: 978-1-61779-946-4

  • eBook Packages: MedicineMedicine (R0)

Publish with us

Policies and ethics