Skip to main content

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

  • 2455 Accesses

Abstract

Endovascular thrombectomy for stroke has become an integral part of the management of acute ischemic stroke. This chapter discusses mechanical thrombectomy and pharmacological thrombolysis. Patient selection, technical issues, and post-procedure care are covered. The appendix covers imaging for stroke.

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

Access this chapter

Institutional subscriptions

References

  1. Powers WJ, Derdeyn CP, Biller J, et al. American Heart Association/American Stroke Association Focused Update of the 2013 Guidelines for the Early Management of Patients with Acute Ischemic Stroke Regarding Endovascular Treatment: A Guideline for Healthcare Professionals From the American Heart Association/American Stroke Association. Stroke. 2015;46:3020–35.

    CAS  PubMed  Google Scholar 

  2. Berkhemer OA, Fransen PS, Beumer D, et al. A randomized trial of intraarterial treatment for acute ischemic stroke. N Engl J Med. 2015;372:11–20.

    PubMed  Google Scholar 

  3. van den Berg LA, Dijkgraaf MG, Berkhemer OA, et al. Two-year outcome after endovascular treatment for acute ischemic stroke. N Engl J Med. 2017;376:1341–9.

    PubMed  Google Scholar 

  4. Al-Ali F, Berkhemer OA, Yousman WP, et al. The capillary index score as a marker of viable cerebral tissue: proof of concept-the capillary index score in the MR CLEAN (Multicenter Randomized Clinical Trial of Endovascular Treatment for Acute Ischemic Stroke in the Netherlands) Trial. Stroke. 2016;47:2286–91.

    PubMed  Google Scholar 

  5. Grotta JC, Hacke W. Stroke neurologist’s perspective on the new endovascular trials. Stroke. 2015;46:1447–52.

    PubMed  Google Scholar 

  6. Goyal M, Demchuk AM, Menon BK, et al. Randomized assessment of rapid endovascular treatment of ischemic stroke. N Engl J Med. 2015;372:1019–30.

    CAS  PubMed  Google Scholar 

  7. Jovin TG, Chamorro A, Cobo E, et al. Thrombectomy within 8 hours after symptom onset in ischemic stroke. N Engl J Med. 2015;372:2296–306.

    CAS  PubMed  Google Scholar 

  8. Saver JL, Goyal M, Bonafe A, et al. Stent-retriever thrombectomy after intravenous t-PA vs. t-PA alone in stroke. N Engl J Med. 2015;372:2285–95.

    CAS  PubMed  Google Scholar 

  9. Campbell BC, Mitchell PJ, Kleinig TJ, et al. Endovascular therapy for ischemic stroke with perfusion-imaging selection. N Engl J Med. 2015;372:1009–18.

    CAS  PubMed  Google Scholar 

  10. Goyal M, Menon BK, van Zwam WH, et al. Endovascular thrombectomy after large-vessel ischaemic stroke: a meta-analysis of individual patient data from five randomised trials. Lancet. 2016;387:1723–31.

    PubMed  Google Scholar 

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

    Google Scholar 

  12. 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.

    CAS  PubMed  Google Scholar 

  13. 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.

    CAS  PubMed  Google Scholar 

  14. 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.

    PubMed  Google Scholar 

  15. Mazighi M, Meseguer E, Labreuche J, et al. Dramatic recovery in acute ischemic stroke is associated with arterial recanalization grade and speed. Stroke. 2012;43:2998–3002.

    PubMed  Google Scholar 

  16. 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.

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  18. Adams HP, Jr., 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;34:1056-1083.

    Google Scholar 

  19. Manual on Contrast Media. 5.0 ed. Reston, VA: American College of Radiology; 2004.

    Google Scholar 

  20. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Lima HN, Cabral NL, Goncalves AR, Hauser A, Pecoits-Filho R. Association between albuminuria, glomerular filtration rate and mortality or recurrence in stroke patients. Nephron Clin Pract. 2011;117:c246–52.

    CAS  PubMed  Google Scholar 

  22. 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.

    Google Scholar 

  23. 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.

    CAS  PubMed  Google Scholar 

  24. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  26. Hoffmeister HM, Jur M, Ruf-Lehmann M, Helber U, Heller W, Seipel L. Endothelial tissue-type plasminogen activator release in coronary heart disease: Transient reduction in endothelial fibrinolytic reserve in patients with unstable angina pectoris or acute myocardial infarction. J Am Coll Cardiol. 1998;31:547–51.

    CAS  PubMed  Google Scholar 

  27. Haley EC Jr, 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  29. 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.

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  31. 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.

    PubMed  Google Scholar 

  32. 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.

    PubMed  Google Scholar 

  33. 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.

    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.

    CAS  PubMed  Google Scholar 

  35. 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. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  38. Vergouwen MD, Algra A, Pfefferkorn T, et al. Time is brain(stem) in basilar artery occlusion. Stroke. 2012;43:3003–6.

    PubMed  Google Scholar 

  39. Higashida RT, Furlan AJ. Trial design and reporting standards for intra-arterial cerebral thrombolysis for acute ischemic stroke. Stroke. 2003;34:109e–37.

    Google Scholar 

  40. Zaidat OO, Yoo AJ, Khatri P, et al. Recommendations on angiographic revascularization grading standards for acute ischemic stroke: a consensus statement. Stroke. 2013;44:2650–63.

    PubMed  PubMed Central  Google Scholar 

  41. Yoo AJ, Simonsen CZ, Prabhakaran S, et al. Refining angiographic biomarkers of revascularization: improving outcome prediction after intra-arterial therapy. Stroke. 2013;44:2509–12.

    PubMed  PubMed Central  Google Scholar 

  42. 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.

    CAS  PubMed  Google Scholar 

  43. 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.

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  45. 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.

    PubMed  Google Scholar 

  46. 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.

    CAS  PubMed  Google Scholar 

  47. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  49. 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.

    PubMed  Google Scholar 

  50. 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.

    CAS  PubMed  Google Scholar 

  51. 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.

    PubMed  Google Scholar 

  52. 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.

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  57. 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.

    CAS  PubMed  Google Scholar 

  58. 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.

    CAS  PubMed  Google Scholar 

  59. van den Berg LA, Koelman DL, Berkhemer OA, et al. Type of anesthesia and differences in clinical outcome after intra-arterial treatment for ischemic stroke. Stroke. 2015;46:1257–62.

    PubMed  Google Scholar 

  60. Abou-Chebl A, Yeatts SD, Yan B, et al. Impact of general anesthesia on safety and outcomes in the endovascular arm of interventional management of stroke (IMS) III trial. Stroke. 2015;46(8):2142.

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Jagani M, Brinjikji W, Rabinstein AA, Pasternak JJ, Kallmes DF. Hemodynamics during anesthesia for intra-arterial therapy of acute ischemic stroke. J Neurointerv Surg. 2016;8:883–8.

    PubMed  Google Scholar 

  62. 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.

    PubMed  Google Scholar 

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

    Google Scholar 

  64. 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.

    PubMed  Google Scholar 

  65. 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.

    PubMed  Google Scholar 

  66. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Nogueira RG, Lutsep HL, Gupta R, et al. Trevo versus Merci retrievers for thrombectomy revascularisation of large vessel occlusions in acute ischaemic stroke (TREVO 2): a randomised trial. Lancet. 2012;380:1231–40.

    PubMed  PubMed Central  Google Scholar 

  68. Saver JL, Jahan R, Levy EI, et al. Solitaire flow restoration device versus the Merci Retriever in patients with acute ischaemic stroke (SWIFT): a randomised, parallel-group, non-inferiority trial. Lancet. 2012;380:1241–9.

    PubMed  Google Scholar 

  69. 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. https://doi.org/10.1227/NEU.0b013e3182315f22.

    Article  PubMed  Google Scholar 

  70. Takahira K, Kataoka T, Ogino T, Endo H, Nakamura H. Efficacy of a coaxial system with a compliant balloon catheter for navigation of the Penumbra reperfusion catheter in tortuous arteries: technique and case experience. J Neurosurg. 2017;126:1334–8.

    PubMed  Google Scholar 

  71. Turk AS, Frei D, Fiorella D, et al. ADAPT FAST study: a direct aspiration first pass technique for acute stroke thrombectomy. J Neurointerv Surg. 2014;6(4):260.

    PubMed  Google Scholar 

  72. Blanc R, Redjem H, Ciccio G, et al. Predictors of the aspiration component success of a direct aspiration first pass technique (ADAPT) for the endovascular treatment of stroke reperfusion strategy in anterior circulation acute stroke. Stroke. 2017;48:1588–93.

    PubMed  Google Scholar 

  73. Jankowitz B, Grandhi R, Horev A, et al. Primary manual aspiration thrombectomy (MAT) for acute ischemic stroke: safety, feasibility and outcomes in 112 consecutive patients. J Neurointerv Surg. 2015;7:27–31.

    PubMed  Google Scholar 

  74. Lee HC, Kang DH, Hwang YH, Kim YS, Kim YW. Forced arterial suction thrombectomy using distal access catheter in acute ischemic stroke. Neurointervention. 2017;12:45–9.

    PubMed  PubMed Central  Google Scholar 

  75. Stampfl S, Kabbasch C, Muller M, et al. Initial experience with a new distal intermediate and aspiration catheter in the treatment of acute ischemic stroke: clinical safety and efficacy. J Neurointerv Surg. 2016;8:714–8.

    PubMed  Google Scholar 

  76. Neurovascular thrombus retrieval catheters and guide catheters used during neurological interventional procedures: differences in FDA review and intended use-Letter to Health Care Providers. U.S. Food and Drug Administration; 2017. 2017, at https://www.fda.gov/MedicalDevices/Safety/LetterstoHealthCareProviders/ucm543890.htm.

  77. Dippel DW, Majoie CB, Roos YB, et al. Influence of device choice on the effect of intra-arterial treatment for acute ischemic stroke in MR CLEAN (Multicenter Randomized Clinical Trial of Endovascular Treatment for Acute Ischemic Stroke in the Netherlands). Stroke. 2016;47:2574–81.

    PubMed  Google Scholar 

  78. 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. Intervent Neuroradiol. 2003;9:391–3.

    CAS  Google Scholar 

  79. 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.

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  81. Haussen DC, Rebello LC, Nogueira RG. Optimizating clot retrieval in acute stroke: the push and fluff technique for closed-cell stentrievers. Stroke. 2015;46:2838–42.

    CAS  PubMed  Google Scholar 

  82. Raoult H, Redjem H, Bourcier R, et al. Mechanical thrombectomy with the ERIC retrieval device: initial experience. J Neurointerv Surg. 2016;

    Google Scholar 

  83. Schwaiger BJ, Kober F, Gersing AS, et al. The pREset stent retriever for endovascular treatment of stroke caused by MCA occlusion: safety and clinical outcome. Clin Neuroradiol. 2016;26:47–55.

    CAS  PubMed  Google Scholar 

  84. 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.

    PubMed  Google Scholar 

  85. Behme D, Kowoll A, Mpotsaris A, et al. Multicenter clinical experience in over 125 patients with the Penumbra separator 3D for mechanical thrombectomy in acute ischemic stroke. J Neurointerv Surg. 2016;8:8–12.

    PubMed  Google Scholar 

  86. Rohde S, Haehnel S, Herweh C, et al. Mechanical thrombectomy in acute embolic stroke: preliminary results with the revive device. Stroke. 2011;42:2954–6.

    PubMed  Google Scholar 

  87. Kara B, Selcuk HH, Yildiz O, Cetinkaya D. Revascularization of acute basilar artery occlusion using the Tigertriever adjustable clot retriever. Clin Neuroradiol. 2017;27(2):241–3. https://doi.org/10.1007/s00062-016-0532-1.

    Article  PubMed  Google Scholar 

  88. Kallenberg K, Solymosi L, Taschner CA, et al. Endovascular stroke therapy with the Aperio thrombectomy device. J Neurointerv Surg. 2016;8:834–9.

    PubMed  Google Scholar 

  89. Fargen KM, Mocco J, Gobin YP. The Lazarus Funnel: a blinded prospective randomized in vitro trial of a novel CE-marked thrombectomy assist device. J Neurointerv Surg. 2016;8:66–8.

    PubMed  Google Scholar 

  90. Baek JH, Kim BM, Kim DJ, Heo JH, Nam HS, Yoo J. Stenting as a rescue treatment after failure of mechanical thrombectomy for anterior circulation large artery occlusion. Stroke. 2016;47:2360–3.

    PubMed  Google Scholar 

  91. 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.

    PubMed  PubMed Central  Google Scholar 

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

    Google Scholar 

  93. 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.

    Google Scholar 

  94. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  95. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  Google Scholar 

  97. 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.

    CAS  PubMed  Google Scholar 

  98. 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.

    PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  101. Kerber CW, Wanke I, Bernard J Jr, 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  102. 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  PubMed Central  Google Scholar 

  103. 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  PubMed Central  Google Scholar 

  104. 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.

    CAS  PubMed  Google Scholar 

  105. Cross DT 3rd, 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  PubMed Central  Google Scholar 

  106. 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.

    CAS  PubMed  Google Scholar 

  107. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  108. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  110. 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.

    CAS  PubMed  Google Scholar 

  111. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  113. 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.

    CAS  PubMed  Google Scholar 

  114. The IMSSI. Combined intravenous and intra-arterial recanalization for acute ischemic stroke: The Interventional Management of Stroke Study. Stroke. 2004;35:904–11.

    Google Scholar 

  115. 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(9):802.

    CAS  PubMed  Google Scholar 

  116. Rubiera MMDP, Ribo MMDP, Pagola JMDP, 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.

    CAS  PubMed  Google Scholar 

  117. 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.

    CAS  PubMed  Google Scholar 

  118. 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.

    PubMed  Google Scholar 

  119. Friedman SG, Pellerito JS, Scher L, Faust G, Burke B, Safa T. Ultrasound-guided thrombin injection is the treatment of choice for femoral pseudoaneurysms. Arch Surg. 2002;137(4):462.

    CAS  PubMed  Google Scholar 

  120. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Howard G, Goff DC. Population shifts and the future of stroke: forecasts of the future burden of stroke. Ann N Y Acad Sci. 2012;1268:14–20.

    PubMed  PubMed Central  Google Scholar 

  122. Chandra RV, Leslie-Mazwi TM, Oh DC, et al. Elderly patients are at higher risk for poor outcomes after intra-arterial therapy. Stroke. 2012;43:2356–61.

    PubMed  Google Scholar 

  123. Di Carlo A, Lamassa M, Pracucci G, et al. Stroke in the very old : clinical presentation and determinants of 3-month functional outcome: A European perspective. European BIOMED Study of Stroke Care Group. Stroke. 1999;30:2313–9.

    PubMed  Google Scholar 

  124. Castonguay AC, Zaidat OO, Novakovic R, et al. Influence of age on clinical and revascularization outcomes in the North American Solitaire Stent-Retriever Acute Stroke Registry. Stroke. 2014;45:3631–6.

    PubMed  Google Scholar 

  125. Willey JZ, Ortega-Gutierrez S, Petersen N, et al. Impact of acute ischemic stroke treatment in patients >80 years of age: the specialized program of translational research in acute stroke (SPOTRIAS) consortium experience. Stroke. 2012;43:2369–75.

    PubMed  PubMed Central  Google Scholar 

  126. Satti S, Chen J, Sivapatham T, Jayaraman M, Orbach D. Mechanical thrombectomy for pediatric acute ischemic stroke: review of the literature. J Neurointerv Surg. 2016;

    Google Scholar 

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

    PubMed  Google Scholar 

  128. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  131. 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.

    CAS  PubMed  Google Scholar 

  132. 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.

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  134. 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.

    CAS  PubMed  Google Scholar 

  135. 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

    Google Scholar 

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

    PubMed  Google Scholar 

  137. Rangaraju S, Jovin TG, Frankel M, et al. Neurologic examination at 24 to 48 hours predicts functional outcomes in basilar artery occlusion stroke. Stroke. 2016;47:2534–40.

    PubMed  PubMed Central  Google Scholar 

  138. Heck DV, Brown MD. Carotid stenting and intracranial thrombectomy for treatment of acute stroke due to tandem occlusions with aggressive antiplatelet therapy may be associated with a high incidence of intracranial hemorrhage. J Neurointerv Surg. 2015;7:170–5.

    PubMed  Google Scholar 

  139. 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.

    CAS  PubMed  Google Scholar 

  140. 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.

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  142. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  146. 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  PubMed Central  Google Scholar 

  147. 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 

  148. Leavitt JA, Larson TA, Hodge DO, Gullerud RE. The incidence of central retinal artery occlusion in Olmsted County, Minnesota. Am J Ophthalmol. 2011;

    Google Scholar 

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

    CAS  PubMed  Google Scholar 

  150. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  152. 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.

    CAS  PubMed  Google Scholar 

  153. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  154. 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  PubMed Central  Google Scholar 

  155. 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.

    PubMed  Google Scholar 

  156. 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.

    PubMed  Google Scholar 

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

    Google Scholar 

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

    CAS  PubMed  Google Scholar 

  159. 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

    CAS  PubMed  Google Scholar 

  160. 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;117:1367–75. e1

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  162. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  163. 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;

    Google Scholar 

  164. 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.

    CAS  PubMed  Google Scholar 

  165. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  167. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  168. 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.

    CAS  PubMed  Google Scholar 

  169. 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.

    PubMed  PubMed Central  Google Scholar 

  170. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  171. 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 

  172. 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.

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  174. 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;32:438-441.

    Google Scholar 

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

    CAS  PubMed  Google Scholar 

  176. 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

    PubMed  Google Scholar 

  177. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  179. 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.

    Google Scholar 

  180. 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.

    CAS  PubMed  Google Scholar 

  181. 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.

    CAS  PubMed  Google Scholar 

  182. 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.

    CAS  PubMed  Google Scholar 

  183. 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.

    PubMed  Google Scholar 

  184. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  Google Scholar 

  186. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  188. 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.

    CAS  PubMed  Google Scholar 

  189. 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.

    CAS  PubMed  Google Scholar 

  190. 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.

    CAS  PubMed  Google Scholar 

  191. 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.

    PubMed  Google Scholar 

  192. 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.

    PubMed  Google Scholar 

  193. Menon BK, d'Esterre CD, Qazi EM, et al. Multiphase CT angiography: a new tool for the imaging triage of patients with acute ischemic stroke. Radiology. 2015;275:510–20.

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  195. 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.

    CAS  PubMed  Google Scholar 

  196. 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.

    CAS  PubMed  Google Scholar 

  197. 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.

    CAS  PubMed  Google Scholar 

  198. 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.

    CAS  PubMed  Google Scholar 

  199. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  201. 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.

    CAS  PubMed  Google Scholar 

  202. 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.

    CAS  PubMed  Google Scholar 

  203. 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.

    CAS  PubMed  Google Scholar 

  204. 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.

    PubMed  Google Scholar 

  205. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  206. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  208. 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 

  209. 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.

    CAS  PubMed  Google Scholar 

  210. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  211. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  213. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  214. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  215. 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.

    CAS  PubMed  Google Scholar 

  216. 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  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  218. 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.

    Google Scholar 

  219. 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  PubMed Central  Google Scholar 

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

    Google Scholar 

  221. 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.

    PubMed  PubMed Central  Google Scholar 

  222. 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.

    CAS  PubMed  Google Scholar 

  223. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  Google Scholar 

  226. 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.

    PubMed  Google Scholar 

  227. 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.

    PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  229. 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.

    PubMed  Google Scholar 

  230. 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.

    CAS  PubMed  Google Scholar 

  231. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  232. 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.

    PubMed  Google Scholar 

  233. 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.

    PubMed  Google Scholar 

  234. 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.

    CAS  PubMed  Google Scholar 

  235. Campbell BC, Christensen S, Levi CR, et al. Cerebral blood flow is the optimal CT perfusion parameter for assessing infarct core. Stroke. 2011;

    Google Scholar 

  236. 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.

    CAS  PubMed  Google Scholar 

  237. Straka M, Albers GW, Bammer R. Real-time diffusion-perfusion mismatch analysis in acute stroke. J Magn Reson Imaging. 2010;32:1024–37.

    PubMed  PubMed Central  Google Scholar 

  238. Mokin M, Levy EI, Saver JL, et al. Predictive value of RAPID assessed perfusion thresholds on final infarct volume in SWIFT PRIME (Solitaire With the Intention for Thrombectomy as Primary Endovascular Treatment). Stroke. 2017;48:932–8.

    PubMed  Google Scholar 

  239. 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(6):544.

    PubMed  Google Scholar 

  240. Kohrmann M, Struffert T, Frenzel T, Schwab S, Doerfler A. The hyperintense acute reperfusion marker on fluid-attenuated inversion recovery magnetic resonance imaging is caused by gadolinium in the cerebrospinal fluid. Stroke. 2012;43:259–61.

    PubMed  Google Scholar 

  241. 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.

    CAS  PubMed  Google Scholar 

  242. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  244. 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.

    PubMed  Google Scholar 

  245. Diffusion Imaging: From Basic Physics to Practical Imaging. RSNA, 1999. (Accessed February 17,2007, 2007, at http://ej.rsna.org/ej3/0095-98.fin/index.htm.)

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

    CAS  PubMed  Google Scholar 

  247. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  250. 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.

    PubMed  Google Scholar 

  251. 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.

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  254. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  256. 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.

    CAS  PubMed  PubMed Central  Google Scholar 

  257. 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.

    CAS  PubMed  Google Scholar 

  258. 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.

    PubMed  Google Scholar 

  259. 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.

    CAS  PubMed  Google Scholar 

  260. 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.

    CAS  PubMed  Google Scholar 

  261. 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.

    CAS  PubMed  Google Scholar 

  262. 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.

    PubMed  Google Scholar 

  263. 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 Neurointervent Surg. 2011;

    Google Scholar 

  264. 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.

    CAS  PubMed  Google Scholar 

  265. 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.

    PubMed  Google Scholar 

  266. 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.

    CAS  PubMed  Google Scholar 

  267. 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(7):1581.

    CAS  PubMed  Google Scholar 

  268. 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  PubMed Central  Google Scholar 

  269. 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.

    CAS  PubMed  Google Scholar 

  270. 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.

    PubMed  Google Scholar 

  271. 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.

    CAS  PubMed  Google Scholar 

  272. 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.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  274. 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.

    CAS  PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  276. 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. 1070353

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  279. Stenver DI. Investigation of the safety of MRI contrast medium Omniscan.: Danish Medicines Agency; 2006.

    Google Scholar 

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

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  284. 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.

    PubMed  Google Scholar 

  285. 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.

    CAS  PubMed  Google Scholar 

  286. Brott T, Adams HP, Jr., Olinger CP, et al. Measurements of acute cerebral infarction: a clinical examination scale. Stroke 1989;20:864-870.

    Google Scholar 

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

    CAS  PubMed  Google Scholar 

  288. Adams HP Jr, 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.

    CAS  PubMed  Google Scholar 

  289. 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.

    CAS  PubMed  Google Scholar 

  290. 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.

    CAS  PubMed  Google Scholar 

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

    Google Scholar 

  292. 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.

    PubMed  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 8.1: Primer on Imaging in Stroke Joel K. Curé

Imaging goals in ischemic stroke:

  1. 1.

    Confirming a diagnosis of ischemic stroke and exclusion of nonvascular (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

CT is the most practical initial brain imaging test for stroke. 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 [169].

Stroke location and distribution at imaging reflect mechanism [170]. 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 Non-contrast 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 [171]. 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% [169]. CT is insensitive for small acute infarcts, especially in the posterior fossa [172], and is less sensitive than diffusion-weighted MRI in the acute setting [173]. 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 [171], portends an increased risk of hemorrhagic transformation in patients treated with rTPA [174], and is associated with an increased risk of fatal brain edema [175]. 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 [176]. 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 (Fig. 8.3). 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 7 predicted patients who were unlikely to achieve independent functional outcome [177]. Sensitivity of CT for acute intracranial hemorrhage approaches 100% [178].

Fig. 8.3
figure 3

ASPECTS (Alberta Stroke Program Early Computed Tomography Score). This CT shows a patient with a left MCA-territory ischemic stroke with an ASPECTS score of 6. The right hemisphere is unaffected. A normal ASPECTS score (i.e., no evidence of ischemic stroke) is 10. Subtract one point for each territory showing evidence of ischemia in two standard CT cuts (level of the basal ganglia, left, and above the ganglia, right). In this case, the four areas of ischemic change in the left hemisphere are the internal capsule, lentiform nucleus, internal capsule, and M2. Abbreviation: Region of brain affected showing ischemic change: C Caudate, I Insular ribbon, IC Internal capsule, L Lentiform nucleus, M1 Anterior MCA cortex, M2 MCA, M3 Posterior MCA cortex, M4 Superior territory: Anterior MCA cortex, M5 Superior territory: Lateral MCA cortex, M6 Superior territory: Posterior MCA cortex

1.2 Early CT Findings 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 [179]. 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 [180].

    2. (b)

      Obscuration of the lentiform nucleus reflects cytotoxic edema within the involved basal ganglia, again decreasing gray–white matter contrast [181].

  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 (Fig. 8.4).

  4. 4.

    Sylvian dot sign.” Distal MCA (M2 or M3 branches) occlusion indicated by hyperattenuation in the Sylvian fissure [182]. Sensitivity 38%, specificity 100%, positive predictive value 100%, negative predictive value 68% [183].

Fig. 8.4
figure 4

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

The combined presence of the insular ribbon sign (Fig. 8.5), hemispheric sulcal effacement, and decreased attenuation of the lentiform nucleus is predictive of ICA occlusion [184].

Fig. 8.5
figure 5

Insular ribbon sign. Insular ribbon sign. Loss of gray–white differentiation in the insula (arrowheads) can be an early sign of MCA territory ischemia

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 nonresponders to whom endovascular therapies might be offered [185].

1.3 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 [186]. Compared to catheter angiography, CTA has sensitivity and specificity for the detection of large-vessel occlusion of 98.4% and 98.1%, respectively [186]. CTA may be prone to false-positive results; in two series of CTA in acute stroke, a minority of patients were found to have lesions on CTA that could not be found with catheter angiography [187, 188]. CTA can be particularly useful in assessment of vertebrobasilar occlusion [128], 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 [189]. CTA combined with CT perfusion shows good agreement with MRI in the assessment of infarct size, cortical involvement, and intracranial cerebral artery occlusion [190]. Finally, some authors have found application of ASPECTS [177] 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 [191, 192]. Multiphase CTA is another option; [193] this technique was used to screen patients for enrollment in the ESCAPE trial [6].

1.4 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 [194], 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.

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 [195]. 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 [196,197,198]. Metabolic failure and cell death occur at CBF ≤10 mL per 100 g/min [11]. 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 [199, 200]. CBV can be decreased or increased during cerebral ischemia, depending on the efficacy of cerebral autoregulation and patency of collateral arterial pathways [200,201,202,203].

1.5 CT Perfusion Technique

1.5.1 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/100mL/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.

1.6 Concepts

There are two commonly applied methods of CT perfusion (Table 8.4). These methods, known as the first-pass bolus tracking techniques, are based on the indicator dilution principle and provide information about CBF, CBV, 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 8.4 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:

$$ {C}_t(t)=\mathrm{CBF}\cdot \left[{C}_a(t)\otimes R(t)\right] $$

where Ct(t) is the tissue time-attenuation curve; Ca(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 Ct(t) and Ca(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:

$$ \mathrm{CBF}=\mathrm{CBV}/\mathrm{MTT} $$

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 artery [204] 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 [205].

In the maximum slope method, the maximum slope of the time-attenuation curve is used to calculate CBF (Fig. 8.6) [206,207,208]. 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 [206, 209, 210]. 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 [206, 211].

Fig. 8.6
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 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

1.6.1 Validation

Quantitative CBF measurement by CT perfusion has been validated by comparison to other techniques for measuring CBF such as microspheres; [212], xenon CT [213, 214], and PET [215, 216]. CT perfusion imaging using the deconvolution technique has been shown to demonstrate little variability within individuals [215]. The use of CT perfusion in the identification of cerebral ischemia has been validated in experimental models [217,218,219]. 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.

1.6.2 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 overestimation of CBF in regions that include large vessels, such as around the Sylvian fissure [220]. 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 [205, 216]. 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 [221].

1.7 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 [222, 223]. Assessment of cerebral perfusion based on absolute values for CBF and CBV should be made with caution [224, 225].

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 [226]. Using the maximum slope method, a reduction of CBV of 60%, compared to nonischemic regions of the brain, best identified cerebral ischemia [223].

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 [227]. 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 [223]. 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 [228].

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 [229]. 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 [229]. Moreover, CBF and CBV data are overestimated when the ROI includes major vessels, such as MCA branches [216]. 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.

1.8 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% [230] and 29% [231] 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 [229, 232].

CT perfusion can potentially identify salvageable tissue at risk of infarction [229, 233]. Using the deconvolution method, a mismatch between regional MTT, CBF, and CBV maps can indicate the presence of ischemic but potentially salvageable brain (penumbra) [217, 222, 228, 234]. 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) [226]. A more recent study found that CBF <31% of the mean contralateral hemispheric CBF best predicted infarct core [235].

“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. 8.7) [236].

Fig. 8.7
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 ischemic core (dark threshold area—red on the color image—black arrow) is defined as absolute CBV <2 mL per 100 g, and the penumbra (light threshold area—green on the color image—white arrows) is defined as the region of brain with MTT values 1.45 times the MTT values in the corresponding area of the opposite hemisphere [226]. The follow-up CT shows an infarction that corresponds to the ischemic core region

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 [223]. The lowest relative CBF and CBV values among brain regions not developing infarctions were 29% and 40% of normal values, respectively.

The Rapid software [237] is an automated technique to generate ischemic threshold maps. Based on results from the SWIFT PRIME trial, the following regional thresholds provide the most accurate prediction of infarct volume (values are expressed as the fraction of the values of the opposite, unaffected hemisphere [i.e., >70% reduction in regional CBF = rCBF < 0.3]): [238]

  1. 1.

    rCBF: 0.30–0.34

  2. 2.

    rCBV: 0.32–0.34

1.9 Validation of CT Perfusion in Ischemic Stroke

The deconvolution method has been validated in the diagnosis of acute ischemic stroke by comparison to CT imaging [227] and to MR T2-weighted imaging [227], diffusion imaging [229, 236], and perfusion imaging [227]. 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) [236]. 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) [229]. 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 [239].

The maximum slope method has been validated in acute stroke by comparison to CT, MRI, and SPECT [228]. 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) [208]. 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 [228]. 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 [223].

1.10 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 are available (Table 8.5). 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. The hyperintense acute reperfusion marker (HARM) on FLAIR is a sign of early blood–brain barrier disruption and is caused by leakage of gadolinium into the CSF [240].

Table 8.5 MRI signal characteristics of cerebral infarction

1.11 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. 8.8) [241, 242]. 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 [172]. Sensitivities of 88–100% for acute stroke detection with DWI have been reported, with specificity from 86 to 100% [243]. Analyzed pooled data from several studies yielded a PPV of 100% and a NPV of 90.6% [244]

Fig. 8.8
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); a lacunar stroke (D) and posterior circulation territory embolic stroke (E)

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 [245]. Two otherwise identical image sets are obtained, one with a low (but nonzero) b value and one with a b value = 1000 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 [246]. 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. Later, in the subacute phase, 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 [247]. 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 [248].

Diffusion-weighted imaging can be useful in the workup of patients with TIA [249]. “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 [250]. 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.” [251] Patients with clinically transient neurological events in whom asymptomatic diffusion abnormalities are discovered have a high risk of early completed stroke [252].

Diffusion-weighted imaging and ADC maps are dynamic. Areas of ischemic injury may enlarge by 43% in the first 52 h after onset [253], although in most patients lesion size appears to reach a maximum by 24 h [254]. 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 [255]. Tissues with 75–90% of ADC values in normal brain are likely to proceed to infarction, whereas tissues with ADC values >90% of normal are more likely to recover [256]. Nevertheless, DWI hyperintensity is a necessary stage on the path to infarction [249], and the volume of DWI abnormalities does correlate with clinical severity [253, 257].

1.12 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 selection [258], and controversy about the optimal perfusion parameters for the identification of affected tissue [259]. The value of MRI perfusion imaging lies in the perfusion–diffusion mismatch hypothesis. This holds that abnormal regions on perfusion images which 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 [260], is strongly associated with proximal MCA occlusion [260], and resolves on reperfusion [261, 262]. 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, especially those with large infarctions [263], who underwent attempted intra-arterial thrombolysis.

The penumbra on perfusion imaging has been defined as regions where DWI is normal and TTP >4 s [264], although, for practical purposes, any region that is abnormal on perfusion imaging but normal on DWI may represent salvageable tissue [265]. Among MRI perfusion parameters, CBF, MTT, and TTP appear to best identify all affected tissue (and thereby distinguish penumbra when compared to DWI) [266, 267], and CBV seems to best predict the final infarct volume [268]. 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 [268].

Together, perfusion imaging and DWI can identify tissue that is at the risk of infarction but amenable to salvage with revascularization [269]. 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 [270]. Perfusion–diffusion imaging has been used in clinical trials to select patients for thrombolysis. Intravenous desmoplasia was given only to patients with a DWI-PWI mismatch ≥20%, and the drug was found to be potentially effective in improving clinical outcomes [271, 272].

1.13 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.

      Advantage: No contrast agent is used.

      1. (i)

        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).

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 [273]. 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 [274].

Nephrogenic systemic fibrosis (aka nephrogenic fibrosing dermopathy) is strongly associated with gadodiamide (OmniscanTM; GE Healthcare, Princeton, NJ) [275, 276]. Although most patients have a history of exposure to gadodiamide, other gadolinium-based agents have been implicated [277]. It appears to occur only in patients with renal insufficiency, generally in those requiring dialysis [277], and is dose dependent [276]. 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 [278]. 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 [279].

Management consists of correction of renal function (usually dialysis), which may result in a cessation or reversal of symptoms [280].

1.14 Identification of Hemorrhage on MRI

MRI is as sensitive for acute hemorrhage as CT [169]. 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 T1-weighted 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) depends 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 be 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 [281].

Acute hemorrhage characteristics on MRI are summarized in Table 8.6. Susceptibility-weighted MRI can help identify acute cerebral hemorrhage, microbleeds, and intravascular clot [282]. 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 [249]. 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 [283]. 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 [284].

Table 8.6 MRI signal characteristics of cerebral hemorrhage

Appendix 8.2: NIH Stroke Scale

The National Institutes of Health Stroke Scale (NIHSS) is widely used and it provides important prognostic information [285,286,287,288].

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 8.7 and 8.8). A score of 16 predicts a high probability of death or severe disability whereas a score of ≥6 predicts a good recovery [288]. Some 60–70% of acute ischemic stroke patients with a baseline NIHSS score <10 will have a favorable outcome after 1 year, compared to only 4–16% of patients with a score >20 [289].

Table 8.7 NIH stroke scale
Table 8.8 NIH stroke scale score severity

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Harrigan, M.R., Deveikis, J.P. (2018). Treatment of Acute Ischemic Stroke. In: Handbook of Cerebrovascular Disease and Neurointerventional Technique. Contemporary Medical Imaging. Humana, Cham. https://doi.org/10.1007/978-3-319-66779-9_8

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-66779-9_8

  • Published:

  • Publisher Name: Humana, Cham

  • Print ISBN: 978-3-319-66777-5

  • Online ISBN: 978-3-319-66779-9

  • eBook Packages: MedicineMedicine (R0)

Publish with us

Policies and ethics