Evidence-Based Imaging - part 4 pps

60 148 0
Evidence-Based Imaging - part 4 pps

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

iron exposed to surrounding water molecules in the form of deoxyhemo- globin creates signal loss, making it easy to identify on susceptibility- weighted and T2-weighted (T2W) sequences (21,22). Thus the earliest detection of hemorrhage depends on the conversion of oxyhemoglobin to deoxyhemoglobin, which was believed to occur after the first 12 to 24 hours (20,23). However, this early assumption has been questioned with reports of intraparenchymal hemorrhage detected by MRI within 6 hours, and as early as 23 minutes from symptom onset (24–26). One of the studies prospectively demonstrated that MRI detected all nine patients with CT-confirmed intracerebral hemorrhage (ICH), suggesting the potential of MRI for the hyperacute evaluation of stroke (limited evidence) (24–26). More recently, a blinded study comparing MRI (diffusion-, T2-, and T2*- weighted images) to CT for the evaluation of ICH within 6 hours of onset demonstrated that ICH was diagnosed with 100% sensitivity and 100% accuracy by expert readers using MRI; CT-detected ICH was used as the gold standard (strong evidence) (9). Data regarding the detection of acute subarachnoid and intraventricular hemorrhage using MRI is limited. While it is possible that the conversion of blood to deoxyhemoglobin occurs much earlier than expected in hypoxic tissue, this transition may not occur until much later in the oxygen-rich environment of the CSF (20,27). Thus the susceptibility-weighted sequence may not be sensitive enough to detect subarachnoid blood in the hyper- acute stage. This problem is further compounded by severe susceptibility artifacts at the skull base, limiting detection in this area. The use of the fluid-attenuated inversion recovery (FLAIR) sequence has been advocated to overcome this problem. Increased protein content in bloody CSF appears hyperintense on FLAIR and can be readily detected. Three case-control series using FLAIR in patients with CT-documented subarachnoid or intra- ventricular hemorrhage demonstrated a sensitivity of 92% to 100% and specificity of 100% compared to CT and was superior to CT during the sub- acute to chronic stages (limited evidence) (28–30). Hyperintense signal in the CSF on FLAIR can be seen in areas associated with prominent CSF pul- sation artifacts (i.e., third and fourth ventricles and basal cisterns) and in other conditions that elevate protein in the CSF such as meningitis or after gadolinium administration (level III) (31–33); however, these conditions are not usually confused with clinical presentations suggestive of subarach- noid hemorrhage. At later time points in hematoma evolution (subacute to chronic phase) when the clot demonstrates nonspecific isodense to hypodense appearance on CT, MRI has been shown to have a higher sensitivity and specificity than CT (limited evidence) (28,34,35). The heightened sensitivity of MRI susceptibility-weighted sequences to microbleeds that are not otherwise detected on CT makes interpretation of hyperacute scans difficult, espe- cially when faced with decisions regarding thrombolysis (Fig. 9.1). Patient outcome regarding the use of thrombolytic treatment in this subgroup of patients with microbleeds is not known; however, in one series of 41 patients who had MRI prior to intraarterial tPA, one of five patients with microbleeds on MRI developed major symptomatic hemorrhage compared to three of 36 without (36), raising the possibility that the presence of microbleeds may predict the subsequent development of symptomatic hemorrhage following tPA treatment. As this finding was not statistically significant, a larger study is required for confirmation. 164 K.D. Vo et al. II. What Are the Imaging Modalities of Choice for the Identification of Brain Ischemia and the Exclusion of Stroke Mimics? Summary of Evidence: Based on moderate evidence (level II), MRI (diffusion-weighted imaging) is superior to CT for positive identification of ischemic stroke within the first 24 hours of symptom onset, allowing exclusion of stroke mimics. However, some argue that despite its superi- ority, positive identification merely confirms a clinical diagnosis and does not necessarily influence acute clinical decision making or outcome. Supporting Evidence A. Computed Tomography Computed tomography images are frequently normal during the acute phase of ischemia and therefore the diagnosis of ischemic stroke is con- Chapter 9 Neuroimaging in Acute Ischemic Stroke 165 Figure 9.1. Microhemorrhages. Top row: Two sequential magnetic resonance (MR) images of T2* sequence show innumerable small low signal lesions scattered throughout both cerebral hemispheres compatible with microhemorrhages. Bottom row: Noncontrast axial computed tomography (CT) at the same anatomic levels does not show the microhemorrhages. tingent upon the exclusion of stroke mimics, which include postictal state, systemic infection, brain tumor, toxic-metabolic conditions, positional vertigo, cardiac disease, syncope, trauma, subdural hematoma, herpes encephalitis, dementia, demyelinating disease, cervical spine fracture, con- version disorder, hypertensive encephalopathy, myasthenia gravis, and Parkinson disease (37). Based purely on history and physical examination alone without confirmation by CT, stroke mimics can account for 13% to 19% of cases initially diagnosed with stroke (37,38). Sensitivity of diagno- sis improves when noncontrast CT is used but still 5% of cases are misdi- agnosed as stroke, with ultimate diagnoses including paresthesias or numbness of unknown cause, seizure, complicated migraine, peripheral neuropathy, cranial neuropathy, psychogenic paralysis, and others (39). An alternative approach to excluding stroke mimics, which may account for the presenting neurologic deficit, is to directly visualize ischemic changes in the hyperacute scan. Increased scrutiny of hyperacute CT scans, especially following the early thrombolytic trials, suggests that some patients with large areas of ischemia may demonstrate subtle early signs of infarction, even if imaged within 3 hours after symptom onset. These early CT signs include parenchymal hypodensity, loss of the insular ribbon (40), obscuration of the lentiform nucleus (41), loss of gray–white matter differentiation, blurring of the margins of the basal ganglia, subtle efface- ment of the cortical sulci, and local mass effect (Fig. 9.2). It was previously believed that these signs of infarction were not present on CT until 24 hours after stroke onset; however, early changes were found in 31% of CTs per- formed within 3 hours of ischemic stroke (moderate evidence) (42). In addi- 166 K.D. Vo et al. Figure 9.2. Early CT signs of infarction. A: Noncontrast axial CT performed at 2 hours after stroke onset shows a large low-attenuated area involving the entire right middle cerebral artery distribution (bounded by arrows) with associated effacement of the sulci and sylvian fissure. There is obscuration the right lentiform nucleus (*) and loss of the insular ribbon (arrowhead). B: Follow-up noncontrast axial image 4 days later confirms the infarction in the same vascular distribution. There is hemorrhagic conversion (*) in the basal ganglia with mass effect and subfalcine herniation. tion, early CT signs were found in 81% of patients with CTs performed within 5 hours of middle cerebral artery (MCA) stroke onset (demonstrated angiographically) (moderate evidence) (43). Early CT signs, however, can be very subtle and difficult to detect even among very experienced readers (moderate evidence) (44–46). Moreover, the presence of these early ischemic changes in only 31% of hyperacute strokes precludes its reliabil- ity as a positive sign of ischemia. Early CT signs of infarction, especially involving more than 33% of the MCA distribution, have been reported to be associated with severe stroke, increased risk of hemorrhagic transformation (46–49), and poor outcome (50). Because of these associations, several trials involving thrombolytic therapy including the European Cooperative Acute Stroke Study (ECASS) excluded patients with early CT signs in an attempt to avoid treatment of patients at risk for hemorrhagic transformation (8,46,51,52). Although ECASS failed to demonstrate efficacy of intravenous tPA administered within 6 hours of stroke onset, a marginal treatment benefit was observed in the target population (post-hoc analysis), excluding patients with early CT signs that were inappropriately enrolled in the trial (46). The National Institute of Neurological Disorders and Stroke (NINDS) t-PA stroke trial (7), which did demonstrate efficacy, did not exclude patients with early CT signs, and retrospective analysis of the data showed that early CT signs were associated with stroke severity but not with increased risk of adverse outcome after t-PA treatment (42). Thus, based on current data, early CT signs should not be used to exclude patients who are otherwise eligible for thrombolytic treatment within 3 hours of stroke onset (strong and moder- ate evidence) (7,42). B. Magnetic Resonance Imaging Unlike CT and conventional MR, new functional MR techniques such as diffusion-weighted imaging (DWI) allow detection of the earliest phy- siologic changes of cerebral ischemia. Diffusion-weighted imaging, a sequence sensitive to the random brownian motion of water, is capable of demonstrating changes within minutes of ischemia in rodent stroke models (53–55). Moreover, the sequence is sensitive, detecting lesions as small as 4mm in diameter (56). Although the in vivo mechanism of signal alteration observed in DWI after acute ischemia is unclear, it is believed that ischemia-induced energy depletion increases the influx of water from the extracellular to the intracellular space, thereby restricting water motion, resulting in a bright signal on DW images (57,58). Diffusion-weighted imaging has become widely employed for clinical applications due to improvements in gradient capability, and it is now possible to acquire DW images free from artifacts with an echo planar approach. Because DW images are affected by T1 and T2 contrast, stroke lesions becomes pro- gressively brighter due to concurrent increases in brain water content, leading to the added contribution of hyperintense T2W signal known as “T2 shine-through.” To differentiate between true restricted diffusion and T2 shine-through, bright lesions on DWI should always be confirmed with apparent diffusion coefficient (ADC) maps, which exclusively measure dif- fusion. For stroke lesions in adults, although there is wide individual vari- ability, ADC signal remains decreased for 4 days, pseudo-normalizes at 5 Chapter 9 Neuroimaging in Acute Ischemic Stroke 167 to 10 days, and increases thereafter (56). This temporal evolution of DWI signal allows one to determine the age of a stroke. The high sensitivity and specificity of DWI for the detection of ischemia make it an ideal sequence for positive identification of hyperacute stroke, thereby excluding stroke mimics. Two studies evaluating DWI for the detection of ischemia within 6 hours of stroke onset reported an 88% to 100% sensitivity and 95% to 100% specificity with a positive predictive value (PPV) of 98.5% and negative predictive value (NPV) of 69.5%, using final clinical diagnosis as the gold standard (moderate and limited evi- dence) (59,60). In another study, 50 patients were randomized to DWI or CT within 6 hours of stroke onset, and subsequently received the other imaging modality with a mean delay of 30 minutes. Sensitivity and speci- ficity of infarct detection among blinded expert readers was significantly better when based on DWI (91% and 95%, respectively) compared to CT (61% and 65%) (moderate evidence) (61). The presence of restricted diffu- sion is highly correlated with ischemia, but its absence does not rule out ischemia: false negatives have been reported in transient ischemic attacks and small subcortical infarctions (moderate evidence) (60,62–64). False- positive DWI signals have been reported in brain abscesses (65), herpes encephalitis (66,67), Creutzfeldt-Jacob disease (68), highly cellular tumors such as lymphoma or meningioma (69), epidermoid cysts (70), seizures (71), and hypoglycemia (72) (limited evidence). However, the clinical history and the appearance of these lesions on conventional MR should allow for exclusion of these stroke mimics. Within the first 8 hours of onset, the stroke lesion should be seen only on DWI, and its presence on conventional MR sequences suggests an older stroke or a nonstroke lesion. The DWI images, therefore, should not be interpreted alone but in con- junction with conventional MR sequences and within the proper clinical context. Acute DWI lesion volume has been correlated with long-term clinical outcome, using various assessment scales including the National In- stitutes of Health Stroke Scale (NIHSS), the Canadian Neurologic Scale, the Barthel Index, and the Rankin Scale (moderate evidence) (73–77). This correlation was stronger for strokes involving the cortex and weaker for subcortical strokes (73,74), which is likely explained by a discordance between infarct size and severity of neurologic deficit for small subcorti- cal strokes. In addition to DWI, MR perfusion-weighted imaging (PWI) approaches have been employed to depict brain regions of hypoperfusion. They involve the repeated and rapid acquisition of images prior to and follow- ing the injection of contrast agent using a two-dimensional (2D) gradient echo or spin echo EPI sequence (78,79). Signal changes induced by the first passage of contrast in the brain can be used to obtain estimates of a variety of hemodynamic parameters, including cerebral blood flow (CBF), cerebral blood volume (CBV), and mean transit time (MTT, the mean time for the bolus of contrast agent to pass through each pixel) (79–81). These parame- ters are often reported as relative values since accurate measurement of the input function cannot be determined. However, absolute quantification of CBF has also been reported (82). Thus, hypoperfused brain tissue result- ing from ischemia demonstrates signal changes in perfusion-weighted images, and may provide information regarding regional hemodynamic status during acute ischemia (insufficient evidence). 168 K.D. Vo et al. III. What Imaging Modality Should Be Used for the Determination of Tissue Viability—the Ischemic Penumbra? Summary of Evidence: Determination of tissue viability using functional imaging has tremendous potential to individualize therapy and extend the therapeutic time window for some. Several imaging modalities, including MRI, CT, PET, and SPECT, have been examined in this role. Operational hurdles may limit the use of some of these modalities in the acute setting of stroke (e.g., PET and SPECT), while others such as MRI show promise (limited evidence). Rigorous testing in large randomized controlled trials that can clearly demonstrate that reestablishment of perfusion to regions “at risk” prevents progression to infarction is needed prior to their use in routine clinical decision making. Supporting Evidence A. Magnetic Resonance Imaging The combination of DWI and PWI techniques holds promise in identify- ing brain tissue at risk for infarction. It has been postulated that brain tissue dies over a period of minutes to hours following arterial occlusion. Initially, a core of tissue dies within minutes, but there is surrounding brain tissue that is dysfunctional but viable, comprising the ischemic penumbra. If blood flow is not restored in a timely manner, the brain tissue at risk dies, completing the infarct (83). The temporal profile of signal changes seen on DWI and PWI follows a pattern that is strikingly similar to the theoretical construct of the penumbra described above. On MR images obtained within hours of stroke onset, the DWI lesion is often smaller than the area of perfusion defect (on PWI), and smaller than the final infarct (defined by T2W images obtained weeks later). If the arterial occlusion persists, the DWI lesion grows until it eventually matches the initial perfusion defect, which is often similar in size and location to the final infarct (chronic T2W lesion) (Fig. 9.3) (limited evidence) (84,85). The area of normal DWI signal but abnormal PWI signal is known as the diffusion-perfusion mismatch and has been postulated to represent the ischemic penumbra. Diffusion- perfusion mismatch has been reported to be present in 49% of stroke patients during the hyperacute period (0 to 6 hours) (limited evidence) (86). Growth of the DWI lesion over time has been documented in a random- ized trial testing the efficacy of the neuroprotective agent citicoline. Mean lesion volume in the placebo group increased by 180% from the initial DWI scan (obtained within 24 hours of stroke onset) to the final T2W scan obtained 12 weeks later. Interestingly, lesion volume grew by only 34% in the citicoline-treated group, suggesting a treatment effect (moderate evi- dence) (87). However, efficacy of the agent was not definitively demon- strated using clinical outcome measures (88). The ultimate test of the hypothesis that mismatch represents “penumbra,” will come from studies that correlate initial mismatch with salvaged tissue after effective treat- ment. One small prospective series of 10 patients demonstrated that patients with successful recanalization after intraarterial thrombolysis showed larger areas of mismatch that were salvaged compared to patients that were not successfully recanalized (limited evidence) (89). Chapter 9 Neuroimaging in Acute Ischemic Stroke 169 The promise of diffusion-perfusion mismatch is that it will provide an image of ischemic brain tissue that is salvageable, and thereby individual- ize therapeutic time windows for acute treatments. The growth of the lesion to the final infarct volume may not occur until hours or even days later in some individuals (limited evidence) (84,85), suggesting that tissue may be salvaged beyond the 3-hour window in some. One of the assump- tions underlying the hypothesis that diffusion-perfusion mismatch repre- sents salvageable tissue is that the acute DWI lesion represents irreversibly injured tissue. However, it has been known for some time that DWI lesions are reversible after transient ischemia in animal stroke models (90,91), and reversible lesions in humans have been reported following a transient ischemic attack (TIA) (92) or after reperfusion (93). These data suggest that at least some brain tissue within the DWI lesion may represent reversibly injured tissue. 170 K.D. Vo et al. Figure 9.3. Evolution of the right middle cerebral distribution infarction on mag- netic resonance imaging (MRI). A,B: MRI at 3 hours after stroke onset shows an area of restricted diffusion on diffusion-weighted imaging (DWI) (A) with a larger area of perfusion defect on perfusion-weighted imaging (PWI) (B). The area of normal DWI but abnormal PWI represents an area of diffusion-perfusion mismatch. C,D: Follow-up MRI at 3 days postictus shows interval enlargement of the DWI lesion (C) to the same size as the initial perfusion deficit (B). There is now a matched dif- fusion-perfusion (C,D). Additional new experimental MR techniques such as proton MR spec- troscopy (MRS) and T2 Blood Oxygen Level Dependent (BOLD) and 2D multiecho gradient echo/spin echo have also been explored for the iden- tification of salvageable tissue (94,95). Magnetic resonance spectroscopy is an MR technique that measures the metabolic and biochemical changes within the brain tissues. The two metabolites that are commonly measured following ischemia are lactate and N-acetylaspartate (NAA). Lactate signal is not detected in normal brain but is elevated within minutes of ischemia in animal models, remaining elevated for days to weeks (96). The lactate signal can normalize with immediate reperfusion (97). N-acetylaspartate, found exclusively in neurons, decreases more gradually over a period of hours after stroke onset in animal stroke models (98). It has been suggested that an elevation in lactate with a normal or mild reduction in NAA during the acute period of ischemia may represent the ischemic penumbra (94), though this has not been examined in a large population of stroke patients. The cerebral metabolic rate of oxygen consumption (CMRO 2 ) has been measured in acute stroke patients using MRI, and a threshold value has been proposed to define irreversibly injured brain tissue (level III) (82). Though preliminary, these results appear to be in agreement with data obtained using PET (see below) (99,100). Measurement of CMRO 2 has theoretical advantages over other measures (e.g., CBF, CBV), as the threshold value for irreversible injury is not likely to be time-dependent (101). Clearly research into the identification of viable ischemic brain tissue is at a preliminary stage. However, such techniques may be important for future acute stroke management. These new imaging approaches will require extensive validation and assessment in well-designed clinical trials. B. Computed Tomography In addition to anatomic information, CT is capable of providing some physiologic information, accomplished with either intravenous injection of nonionic contrast or inhalation of xenon gas. Like PWI, perfusion parameters can be obtained by tracking a bolus of contrast or inhaled xenon gas in blood vessels and brain parenchyma with sequential CT imaging. Using spiral CT technology, the study can be completed in 6 minutes. Stable xenon (Xe) has been employed as a means to obtain quantitative estimates of CBF in vivo. Xenon, an inert gas with an atomic number similar to iodine, can attenuate x-rays like contrast material. However, unlike CT contrast, the gas is freely diffusable and can cross the blood–brain barrier. Sequential imaging permits the tracking of progres- sive accumulation and washout of the gas in brain tissue, reflected by changes in Hounsfield units over time, and quantitative CBF and CBV maps can be calculated (102). The quantitative CBF value from xenon- enhanced CT has been shown to be highly accurate compared with radioactive microsphere and iodoantipyrine techniques under different physiologic conditions and wide range of CBF rates in baboons (correla- tion coefficient r = 0.67 to 0.92, p < .01 and <.001) (103,104). The major advantage of the xenon CT is that it allows absolute quantification of the CBF, which may help to define a threshold value from reversible to Chapter 9 Neuroimaging in Acute Ischemic Stroke 171 irreversible cerebral injury. Low CBF (<15mL/100g/min) correlated with early CT signs of infarction, proximal M1 occlusion, severe edema, and life- threatening herniation. Very low CBF values (<7mL/100g/min) predicted irreversibly injured tissue (105,106). In addition, xenon CT has been shown to be effective in obtaining cerebral vascular reserve (CVR) in patients with occlusive disease (107). Poor CVR has been shown to be a risk factor for stroke in patients with high-grade carotid stenosis or occlusion (108). However, to ensure a sufficient signal-to-noise ratio for Xe-CT perfusion, a high concentration of Xe is needed, which itself may cause respiratory depression, cerebral vasodilation, and thus confound the measurements of CBF (109). In addition to inhalation xenon gas, bolus nonionic contrast can also be used to generate a CT perfusion map. Rapid repeated serial images of the brain are acquired during the first-pass passage of intravenous contrast to generate relative CBF, CBV, and MTT. The CT perfusion maps obtained within 6 hours of stroke onset in patients with MCA occlusion had signif- icantly higher sensitivity for the detection of stroke lesion volume com- pared to noncontrast CT, and the perfusion volume correlated with clinical outcome (limited evidence) (105,110). Cerebral blood flow maps generated by CT perfusion in 70 acute stroke patients predicted the extent of cerebral infarction with a sensitivity of 93% and a specificity of 98% (limited evidence) (111). A major limitation to this technique is that only relative CBF map can be obtained, thus precluding exact determination of the tran- sition from ischemia to infarction. C. Positron Emission Tomography (PET) Positron emission tomography imaging has provided fundamental infor- mation on the pathophysiology of human cerebral ischemia. Quantitative measurements of cerebral perfusion and metabolic parameters can be obtained, namely CBF, CBV, MTT, oxygen extraction fraction (OEF), and CMRO 2 , using multiple tracers and serial arterial blood samplings. Based on the values of these hemodynamic parameters, four distinct successive pathophysiologic phases of ischemic stroke have been identified: autoreg- ulation, oligemia, ischemia, and irreversible injury (112). Oligemia (low CBF, elevated OEF with normal CMRO 2 ) and ischemia (low CBF, elevated OEF but decreased CMRO 2 ) are sometimes termed misery perfusion, and have been postulated to represent the ischemic penumbra (113). During misery perfusion, a decline in CMRO 2 heralds the beginning of a transi- tion from reversible to irreversible injury. Irreversible injury is reflected in tissue with CMRO 2 below 1.4mL/100g/min (99,100). In three serial obser- vational studies of acute ischemic stroke, elevation of OEF in the setting of low CBF has been suggested to be the marker of tissue viability in ischemic tissue (level II) (114–116). The CBF in ischemic tissue with elevated OEF is between 7 and 17mL/100g/min. Elevated OEF has been observed to persist up to 48 hours after stroke onset (115). Progression to irreversible injury is reflected in decreased OEF (114,115). Furthermore, in a prospec- tive blinded longitudinal cohort study of 81 patients with carotid occlu- sion, elevated OEF was found to be an independent predictor for subsequent stroke and potentially defining a subgroup of patients who may benefit from revascularization (moderate evidence) (117). However, 172 K.D. Vo et al. confirmation of tissue viability in the region of elevated OEF is best accom- plished by large randomized controlled trials that can clearly demonstrate that reestablishment of perfusion to this region prevents progression to infarction. Such studies have not been done and are difficult to implement since PET is limited to major medical centers and requires considerable expertise and time. Moreover, the requirement for intraarterial line place- ment precludes its use for evaluating thrombolytic candidates. Despite these hurdles one study assessed PET after thrombolysis in 12 ischemic stroke patients within 3 hours of symptoms onset (118). Due to the above- mentioned hurdles, only relative CBF was obtained prior to and follow- ing intravenous thrombolysis (118). In all patients, early reperfusion of severely ischemic tissue (<12mL/110g/min in gray matter) predicted better clinical outcome and limited infarction. D. Single Photon Emission Computed Tomography (SPECT) The most commonly used radiopharmaceutical agent for SPECT perfu- sion study is technetium-99m pertechnetate hexamethyl-propylene amine oxime (99m Tc-HMPAO). This lipophilic substance readily crosses the blood–brain barrier and interacts with intracellular glutathione, which pre- vents it from diffusing back. However, due to technical problems includ- ing incomplete first-pass extraction from blood, incomplete binding to glutathione leading to back diffusion, and metabolism within the brain, absolute quantification of the CBF cannot be determined. However, SPECT technology is much more accessible than PET and is more readily avail- able. In a multicenter prospective trial with 99mTc-bicisate (99mTc-ECD, an agent with better brain-to-background contrast) of 128 patients with ischemic stroke and 42 controls, SPECT had a sensitivity of 86% and specificity of 98% for localization of stroke compared with final clinical, diagnostic, and laboratory studies (119). The sensitivity decreased to 58% for lacunar stroke (119). Perfusion studies with HMPAO-SPECT in early ischemic stroke demonstrated that patients with severe hypoperfu- sion on admission had poor outcome at 1 month (120). Furthermore, reperfusion of ischemic tissue with 65% to 85% reduction of regional CBF (rCBF) compared to the contralateral hemisphere decreased the final infarct volume but had no affect on regions with reduction greater than 85% (121). IV. What Is the Role of Noninvasive Intracranial Vascular Imaging? Summary of Evidence: With the development of different delivery approaches for thrombolysis in acute ischemic stroke, there is a new demand for noninvasive vascular imaging modalities. While some data are available comparing magnetic resonance angiography (MRA) and com- puted tomography angiography (CTA) to digital substraction angiography (DSA) (moderate and limited evidence), strong evidence in support of the use of such approaches for available therapies is lacking. Prospective studies examining clinical outcome after the use of screening vascular imaging approaches to triage therapy are needed. Chapter 9 Neuroimaging in Acute Ischemic Stroke 173 [...]... 2002;50(1):28– 34; discussion 34 35 Minematsu K, et al Stroke 1992;23(9):13 04 1310; discussion 1310–1311 Hasegawa Y, et al Neurology 19 94; 44( 8): 148 4– 149 0 Kidwell CS, et al Ann Neurol 2000 ;47 (4) :46 2 46 9 Fiehler J, et al Stroke 2002;33(1):79–86 Barker PB, et al Radiology 19 94; 192(3):723–732 Grohn OH, Kauppinen RA NMR Biomed 2001; 14( 7–8) :43 2 44 0 Decanniere C, et al Magn Reson Med 1995; 34( 3): 343 –352 Bizzi... 1995;26(12):2238–2 241 40 Truwit CL, et al Radiology 1990;176(3):801–806 41 Tomura N, et al Radiology 1988;168(2) :46 3 46 7 42 Patel SC, et al JAMA 2001;286(22):2830–2838 43 von Kummer R, et al AJNR 19 94; 15(1):9–15; discussion 16–18 44 Schriger DL, et al JAMA 1998;279(16):1293–1297 45 Grotta JC, et al Stroke 1999;30(8):1528–1533 46 Hacke W, et al JAMA 1995;2 74( 13):1017–1025 47 Toni D, et al Neurology 1996 ;46 (2): 341 – 345 ... MP, et al Radiology 1996;199(2) :40 3 40 8 Kidwell CS, et al Stroke 1999;30(6):11 74 1180 Ay H, et al Neurology 1999;52(9):17 84 1792 Ebisu T, et al Magn Reson Imaging 1996; 14( 9):1113–1116 Ohta K, et al J Neurol 1999; 246 (8):736–738 Sener RN Comput Med Imaging Graph 2001;25(5):391–397 Bahn MM, et al Arch Neurol 1997; 54( 11): 141 1– 141 5 Gauvain KM, et al AJR 2001;177(2) :44 9 45 4 Chen S, et al AJNR 2001;22(6):1089–1096... J 1993 ;44 :189–193 Prager JM, Mikulis DJ Med Clin North Am 1991;75:525– 544 Edelman RR, Warach S N Engl J Med 1993;328:708–715 Lledo A, Calandre L, Martinez-Menendez B, et al Headache 19 94; 34: 172–1 74 Scott Morey S Am Fam Physician 2000;62:1699–1701 Joseph R, Cook GE, Steiner TJ, et al Practitioner 1985;229 :47 7 48 1 Weingarten S, Kleinman M, Elperin L, et al Arch Intern Med 1992; 152: 245 7– 246 2 24 Quality... Firlik AD, et al J Neurosurg 1998;89(2): 243 – 249 Pindzola RR, et al Stroke 2001;32(8):1811–1817 Webster MW, et al J Vasc Surg 1995;21(2):338– 344 ; discussion 344 – 345 Chapter 9 Neuroimaging in Acute Ischemic Stroke 109 110 111 112 113 1 14 115 116 117 118 119 120 121 122 123 1 24 125 126 127 128 129 130 131 132 133 1 34 135 136 137 138 139 Plougmann J, et al J Neurosurg 19 94; 81(6):822–828 Lev MH, et al Stroke... 1996 ;46 (2): 341 – 345 48 Larrue V, et al Stroke 1997;28(5):957–960 49 Larrue V, et al Stroke 2001;32(2) :43 8 44 1 50 von Kummer R, et al Radiology 1997;205(2):327–333 51 Hacke W, et al Lancet 1998;352(9136):1 245 –1251 52 Clark WM, et al JAMA 1999;282(21):2019–2026 177 178 K.D Vo et al 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96... 19 94; 81(6):822–828 Lev MH, et al Stroke 2001;32(9):2021–2028 Mayer TE, et al AJNR 2000;21(8): 144 1– 144 9 Baron JC, et al J Cereb Blood Flow Metab 1989;9(6):723– 742 Baron JC, et al Stroke 1981;12 (4) :45 4 45 9 Wise RJ, et al Brain 1983;106(pt 1):197–222 Heiss WD, et al J Cereb Blood Flow Metab 1992;12(2):193–203 Furlan M, et al Ann Neurol 1996 ;40 (2):216–226 Grubb RL Jr, et al JAMA 1998;280(12):1055–1060 Heiss WD, et al J... al JAMA 1999;282(21):2003–2011 9 Fiebach JB, et al Stroke 20 04; 35(2):502–506 10 New PF, Aronow S Radiology 1976;121(3 pt 1):635– 640 11 Smith WP Jr, Batnitzky S, Rengachary SS AJR 1981;136(3): 543 – 546 12 Culebras A, et al Stroke 1997;28(7): 148 0– 149 7 13 Beauchamp NJ Jr, et al Radiology 1999;212(2):307–3 24 14 Paxton R, Ambrose J Br J Radiol 19 74; 47(561):530–565 15 Jacobs L, Kinkel WR, Heffner RR Jr Neurology... 2001;58(6):993–998 Hasegawa Y, et al Stroke 1996;27(9):1 648 –1655; discussion 1655–1656 Lovblad KO, et al Ann Neurol 1997 ;42 (2):1 64 170 Schwamm LH, et al Stroke 1998;29(11):2268–2276 Barber PA, et al Neurology 1998;51(2) :41 8 42 6 Tong DC, et al Neurology 1998;50 (4) :8 64 870 van Everdingen KJ, et al Stroke 1998;29(9):1783–1790 Villringer A, et al Magn Reson Med 1988;6(2):1 64 1 74 Rosen BR, et al Magn Reson Med 1991;19(2):285–292... et al Magn Reson Med 1990; 14( 2): 249 –265 Ostergaard L, et al Magn Reson Med 1996;36(5):726–736 Lin W, et al J Magn Reson Imaging 2001; 14( 6):659–667 Astrup J, Siesjo BK, Symon L Stroke 1981;12(6):723–725 Baird AE, et al Ann Neurol 1997 ;41 (5):581–589 Beaulieu C, et al Ann Neurol 1999 ;46 (4) :568–578 Perkins CJ, et al Stroke 2001;32(12):27 74 2781 Warach S, et al Ann Neurol 2000 ;48 (5):713–722 Clark WM, et . et al. Neurology 19 94; 44( 8): 148 4– 149 0. 92. Kidwell CS, et al. Ann Neurol 2000 ;47 (4) :46 2 46 9. 93. Fiehler J, et al. Stroke 2002;33(1):79–86. 94. Barker PB, et al. Radiology 19 94; 192(3):723–732. 95 Neurol 1999; 246 (8):736–738. 67. Sener RN. Comput Med Imaging Graph 2001;25(5):391–397. 68. Bahn MM, et al. Arch Neurol 1997; 54( 11): 141 1– 141 5. 69. Gauvain KM, et al. AJR 2001;177(2) :44 9 45 4. 70. Chen. 1995;2 74( 13):1017–1025. 47 . Toni D, et al. Neurology 1996 ;46 (2): 341 – 345 . 48 . Larrue V, et al. Stroke 1997;28(5):957–960. 49 . Larrue V, et al. Stroke 2001;32(2) :43 8 44 1. 50. von Kummer R, et al. Radiology 1997;205(2):327–333. 51.

Ngày đăng: 11/08/2014, 01:22

Tài liệu cùng người dùng

  • Đang cập nhật ...

Tài liệu liên quan