You are on page 1of 8

Journal of the Neurological Sciences 419 (2020) 117217

Contents lists available at ScienceDirect

Journal of the Neurological Sciences


journal homepage: www.elsevier.com/locate/jns

Review Article

Strategies to prevent hemorrhagic transformation after reperfusion


therapies for acute ischemic stroke: A literature review
Yutaka Otsu a, 1, Masaki Namekawa a, 1, Masafumi Toriyabe a, b, 1, Itaru Ninomiya a,
Masahiro Hatakeyama a, Masahiro Uemura a, Osamu Onodera a, Takayoshi Shimohata c,
Masato Kanazawa a, *
a
Department of Neurology, Brain Research Institute, Niigata University, Niigata, Japan
b
Department of Medical Technology, Graduate School of Health Sciences, Niigata University, Niigata, Japan
c
Department of Neurology, Gifu University Graduate School of Medicine, Gifu, Japan

A R T I C L E I N F O A B S T R A C T

Keywords: Background: Reperfusion therapies by tissue plasminogen activator (tPA) and mechanical thrombectomy (MT)
tPA have ushered in a new era in the treatment of acute ischemic stroke (AIS). However, reperfusion therapy-related
Thrombectomy HT remains an enigma.
Reperfusion
Aim: To provide a comprehensive review focused on emerging concepts of stroke and therapeutic strategies,
Therapeutic time window
including the use of protective agents to prevent HT after reperfusion therapies for AIS.
Hemorrhagic transformation
Methods: A literature review was performed using PubMed and the ClinicalTrials.gov database.
Results: Risk of HT increases with delayed initiation of tPA treatment, higher baseline glucose level, age, stroke
severity, episode of transient ischemic attack within 7 days of stroke onset, and hypertension. At a molecular
level, HT that develops after thrombolysis is thought to be caused by reactive oxygen species, inflammation,
remodeling factor-mediated effects, and tPA toxicity. Modulation of these pathophysiological mechanisms could
be a therapeutic strategy to prevent HT after tPA treatment. Clinical mechanisms underlying HT after MT are
thought to involve smoking, a low Alberta Stroke Program Early CT Score, use of general anesthesia, unfavorable
collaterals, and thromboembolic migration. However, the molecular mechanisms are yet to be fully investigated.
Clinical trials with MT and protective agents have also been planned and good outcomes are expected.
Conclusion: To fully utilize the easily accessible drug—tPA—and the high recanalization rate of MT, it is
important to reduce bleeding complications after recanalization. A future study direction could be to investigate
the recovery of neurological function by combining reperfusion therapies with cell therapies and/or use of
pleiotropic protective agents.

1. Introduction with tPA administration is a class I recommendation included in the


American Heart Association/American Stroke Association guidelines
Acute ischemic stroke (AIS) is a leading cause of death and disability (causative occlusion of the internal carotid artery or middle cerebral
worldwide. The 25th anniversary of the first tPA clinical trial and the 5th artery segment 1 [M1]; to be administered within 6 h of symptom onset)
anniversary of the first MT trial ushered in a new era in the treatment of [3]. The frequency of tPA treatment increased from 9.9% in 2006 to
AIS. Currently, intravenous (IV) administration of tissue plasminogen 21.8% in 2018 according to the Austrian Stroke Unit registry [9].
activator (tPA) is the only thrombolytic therapy approved to treat AIS However, in the United States, between 3.4% and 5.2% of all patients
within 4.5 h of stroke onset [1–3]. In addition, mechanical thrombec­ with AIS are eligible for tPA treatment because of the huge geographical
tomy (MT) or endovascular therapy has shown great benefit to AIS pa­ area and differences in healthcare systems [10]. MT was performed in
tients when performed within 6 or 8 h of stroke onset in patients with only 3.0% of patients with AIS in the Catalonian registry (REVASCAT
large vessel occlusion [4–8]. MT with a stent retriever in combination trial) [11]. In other words, patients eligible for reperfusion treatment

* Corresponding author at: Department of Neurology, Brain Research Institute, Niigata University, 1-757 Asahimachi-dori, Chuoku, Niigata 951-8585, Japan.
E-mail address: masa2@bri.niigata-u.ac.jp (M. Kanazawa).
1
These authors contributed equally to the work.

https://doi.org/10.1016/j.jns.2020.117217
Received 1 September 2020; Received in revised form 9 October 2020; Accepted 29 October 2020
Available online 4 November 2020
0022-510X/© 2020 Elsevier B.V. All rights reserved.
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

still account for only 5%–10% of all patients with AIS. In the pooled outcome or ongoing studies (Fig. 1).
analysis, tPA (alteplase) increased hemorrhagic transformation (HT),
especially when administered from 3 h to 4.5 h after onset of symptoms 3. Results
[12,13]. Even within 4.5 h of stroke onset, early treatment is very
important for patients with severe stroke because of the increasing risk 3.1. HT pathophysiology after thrombolysis
of symptomatic HT or intracranial hemorrhage [13]. Therefore, strate­
gies to prevent HT that develop after reperfusion therapies might in­ To prevent the onset of thrombolysis-associated HT, it is necessary to
crease the number of patients who are eligible for reperfusion. understand its pathophysiology. The risk of symptomatic HT also in­
Reperfusion within the therapeutic time window can reduce brain creases with several factors (Table 1). However, the risk of HT
damage by salvaging a reversibly damaged penumbra of tissue. Ische­ completely offset by a long-term survival benefit if treatment is
mia–reperfusion–induced injury has been demonstrated in a variety of administered within 4.5 h [20–22]. These risk factors may have caused
organ systems, caused by many potential mechanisms that can disrupt BBB disruption and HT.
the neurovascular unit and increase the blood-brain barrier (BBB) The molecular mechanisms underlying HT are thought to be caused
permeability, thus leading to HT. Very recently, among patients with AIS by the following: (1) cerebral ischemia/reperfusion injury leading to the
who underwent diffusion-weighted imaging (DWI) with fluid- activation of several reactive oxygen species (ROS), (2) inflammation,
attenuated inversion recovery (FLAIR) and perfusion mismatch se­ (3) remodeling factor-mediated effects, and (4) direct toxicity of tPA
quences between 4.5 and 9 h of stroke onset, tPA treatment improved (Fig. 2) [23–25]. HT-induced mediators change chronologically and
functional outcomes [14,15]. However, the incidence of symptomatic dynamically, whereas in rat models, (1) ROS production was higher in
HT has increased [14,15]. Another meta-analysis suggested that eligible the reperfusion group than in the middle cerebral artery occlusion
patients presenting with unknown symptom onset time and DWI-FLAIR (MCAO) models [26]. In permanent MCAO models, ROS gradually
mismatch or patients with symptom onset outside the conventional time increased over the course of up to 3 h. By contrast, reperfusion after 1 h
window of 4.5 h with evidence of viable tissue on penumbral imaging of MCAO resulted in a significant and sustained increase in ROS for 3 h.
could benefit from prompt intravenous (IV) tPA administration; how­ ROS, presumably, play a role in the early stages of reperfusion injury. (2)
ever, the HT increased [16]. Moreover, although the rate of symptom­ Inflammation is also an important factor mediating reperfusion injury.
atic HT did not differ significantly between the MT and control groups in tPA-induced reperfusion resulted in increased numbers of neutrophils,
the DAWN and DEFUSE 3 trials, the symptomatic HT ratio in the MT microglia and macrophages [27–30], and proinflammatory M1 macro­
group was double than that in the control group (asymptomatic HT was phages after 3 h [31]. Further, inflammation mediates BBB disruption
unknown) [17,18]. From the findings of the Safe Implementation of and induces HT. (3) Angiogenesis after ischemia could be a pathological
Treatments in Stroke-International Stroke Thrombolysis Registry (SITS- process contributing to neuropil injury after ischemia, or might be a
ISTR) (N = 50,726), early neurological deterioration (at 24 h from mechanism to limit the ischemic injury, or both, or perhaps a mecha­
baseline scale) was most associated with HT and death (odds ratio, 3.72 nism to recover tissue function [32,33]. During angiogenesis, remodel­
and 7.32, respectively) [19]. Therefore, HT attenuation after tPA ing factors, such as vascular endothelial growth factor (VEGF),
treatment and MT is an important and safe therapeutic strategy to treat angiopoietin-1, and matrix metalloproteinase (MMP)-9, play impor­
AIS. It might enable extension of the therapeutic time window and the tant roles in the chronic phase of ischemia. However, alternation in the
probability of achieving more excellent outcomes. levels of these factors in the acute phase induces HT [34–36]. In hy­
This review focuses on emerging concepts of stroke and the thera­ perglycemic mice, tPA treatment increased peroxisome proliferator-
peutic strategies to improve the outcomes of patients with AIS, including activated receptor-γ upregulation, activation of inflammasomes, and
administration of protective agents to prevent HT after reperfusion VEGF and MMP-9 signaling [37]. HT after thrombolytic therapies might
therapies for AIS. be coupled with VEGF and MMP-9 signaling [31,34,37,38]. Remodeling
factors mediate BBB disruption and delay HT (> 18–24 h). Inhibition of
2. Methods the VEGF signaling pathway and the administration of angiopoietin-1
attenuates HT after tPA treatment of ischemic stroke [34,35].
A literature review was performed using PubMed and the National Although this treatment can enable vascular protection, it cannot reduce
Institutes of Health clinical trial database (ClinicalTrials.gov). We the cerebral infarct volume [34,35]. Pleiotropic target, which includes
searched articles published between April 1995 and August 2020 using vascular protection, neuroprotection, and anti-inflammation, is an ideal
the following search terms: “stroke,” “cerebral ischemia,” “thrombol­ therapeutic strategy for ischemic stroke. It has been reported that the
ysis,” “tissue plasminogen activator, tPA,” “mechanical thrombectomy, growth factor, progranulin, could protect against acute focal cerebral
MT,” and “hemorrhagic transformation, HT,” and “pleiotropic mecha­ ischemia by neuroprotection, suppression of neuroinflammation, and
nism.” Based on the data of the clinical studies, we selected positive attenuation of BBB disruption via VEGF inhibition [27]. Intravenously
administered recombinant progranulin significantly reduced the vol­
umes of cerebral infarction and edema, suppressed HT, and improved
motor outcomes in thromboembolic rats with delayed administration of

Table 1
Risk factors of hemorrhagic transformation after tPA treatment.
Risk factors

Delayed initiation of tPA (alteplase) treatment


Higher baseline glucose level
Diabetes mellitus
Hypertension
Older age
Greater stroke severity
Large infarct volume
Excessive alcohol consumption
Episode of TIA within 7 days

Fig. 1. Flowchart for study inclusions. tPA, tissue plasminogen activator; TIA, transient ischemic attack.

2
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

Fig. 3. Drug candidates to prevent hemorrhagic transformation after tPA


treatment and their underlying therapeutic mechanisms. MMP, matrix metal­
loproteinase; tPA, tissue plasminogen activator.

against HT [44]. Another clinical trial reported that the administration


Fig. 2. Mechanisms underlying intracerebral hemorrhagic transformation after
of edaravone before tPA did not affect the rate of early recanalization,
tPA treatment and the therapeutic targets (revised figure from Kanazawa et al.
symptomatic HT, or favorable outcomes after tPA treatment [45].
[25]). BBB, blood-brain barrier; ROS, reactive oxygen species; tPA, tissue
plasminogen activator.
Because edaravone is widely used in Japan, it was difficult to create a
control group (tPA without edaravone treatment). Although edaravone
may be a good candidate for combined therapy with tPA and MT, a
tPA treatment. It has also been reported that progesterone, a sex steroid
worldwide clinical trial is needed to assess the efficacy of edaravone and
hormone, attenuates HT after tPA treatment in an ischemic stroke model
tPA with and without MT.
of rats [31,38]. Intraperitoneal administration of progesterone reduces
Fingolimod is a sphingosine analog that binds to the sphingosine-1-
volumes of cerebral infarction and edema, suppresses HT via inhibition
phosphate receptors. It inhibits the release of lymphocytes from lymph
of VEGF, MMP-9 and MMP-3 activity, and inflammation. Progesterone
nodes and limits their recirculation [46]. Inflammation caused by brain
may be also effective for suppression of HT. Although a phase IV clinical
ischemia contributes to HT and reperfusion injury and worsens the
trial is currently ongoing in hemorrhagic stroke with progesterone
clinical outcomes of stroke associated with tPA treatment. In animal
(ClinicalTrials.gov; NCT04143880), clinical trials for ischemic stroke
models of thromboembolic MCAO, fingolimod and tPA attenuated the
are not planned. Pleiotropic growth factors could inhibit HT and effi­
neurological deficit and reduced infarct volume and HT [47]. In addi­
ciently promote functional recovery. (4) Recently, two clinical trials
tion, although a small number of patients received tPA and fingolimod
investigating the direct effects of MT without tPA demonstrated a lower
combined therapy 4.5 to 6 h after ischemic onset, improved functional
incidence of HT compared to MT with tPA [39,40]. Thus, tPA (alteplase)
outcomes were observed [48]. Fingolimod is approved for use and is
could directly accelerate HT. Modulating these variable pathophysio­
being used as a disease-modifying drug for multiple sclerosis; therefore,
logical mechanisms is a direct and promising therapeutic strategy to
drug repurposing is easy. However, the neutrophil-to-lymphocyte ratio
prevent HT. Several mechanisms underlie an ischemic stroke. Therefore,
is associated with a greater risk of HT in tPA-treated patients with AIS
therapeutic agents with pleiotropic protective mechanisms, which target
[29,30]. Combinating fingolimod with alteplase bridging with MT in
the underlying molecular pathways, are ideal to prevent HT.
acute ischemic stroke (FAMTAIS) trial was withdrawn on March 2020
(ClinicalTrials.gov; NCT02956200). Therefore, based on these results,
3.2. Agents with potential benefits in clinical studies of patients treated the effects of combined treatment with tPA and fingolimod remain
with tPA unknown.
Minocycline, apart from its antibacterial properties, has neuro­
Studies have investigated the concomitant use of various agents, protective effects in various neurological disease processes [49]. Mino­
which reduce hemorrhagic events and exert neuroprotective effects, cycline decreases plasma MMP-9 levels and reduce infarction and HT
alongside tPA. Because monotherapies might be insufficient to attenuate [36]. Two pilot clinical trials have shown that minocycline, in combi­
HT after tPA treatment and improve the therapeutic outcome of AIS nation with tPA, is potentially effective to treat AIS [50,51]. In 2018,
patients, the efficacy of several—but not single—potential target drugs two systematic reviews showed that use of minocycline with tPA had
have been evaluated (Fig. 3). The drugs that have shown potential favorable functional outcomes [52,53]. However, phase III large clinical
benefits in clinical trials are described below. trials are needed to confirm the potential treatment effect of minocycline
Edaravone, a free radical scavenger, exerts a neuroprotective effect in AIS therapy.
in ischemic brain models by inhibiting vascular endothelial cell injury Natalizumab is a monoclonal antibody targeting α4 integrin within
and ameliorating neuronal damage [41]. Combined treatment with tPA the adhesion molecule very late antigen-4 (VLA-4), leading to reduced
and edaravone might suppress MMP-9 expression in and around the transmigration of leucocytes, excluding neutrophils, across the vascular
cerebral microvessels and inhibit BBB damage in a rat focal cerebral endothelium [54]. Natalizumab is approved for the treatment of mul­
ischemic model. Moreover, edaravone might protect cerebral micro­ tiple sclerosis and Crohn’s disease and is highly effective in reducing
vascular integrity because it safeguards the basement membrane from inflammatory lesions [55–57]. Safety and efficacy of intravenous nata­
excess free radicals and MMP-9, leading to a subsequent decrease in HT lizumab in AIS (ACTION and ACTION II) trials were performed
and functional improvement. Edaravone was approved for use in pa­ (ClinicalTrials.gov; NCT01955707 [58] and NCT02730455 [59],
tients with AIS in Japan [42] in 2001 and the compound is widely used respectively). The administration of natalizumab was combined with
in China [43], although combined use of tPA with a free-radical­ tPA and MT. However, natalizumab did not reduce infarct volume
–trapping agent—NXY-059—did not show any evidence of efficacy growth from baseline to day 5 compared with that of the placebo.

3
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

Although no difference was noted between the natalizumab and placebo use with tPA to treat AIS. Phase III clinical trials are inevitably needed to
groups in the National Institutes of Health Stroke Scale (NIHSS) and confirm the possibilities of preventing HT and to identify the functional
incidence of HT, more patients in the natalizumab group than in the efficacy of the candidate drugs.
placebo group had modified Rankin scores of 0 or 1 at day 30 (p = 0.024)
and day 90 (p = 0.18) in ACTION trial [58]. However, the effect of two
doses of natalizumab on functional outcomes was not observed in the 3.3. HT in MT
phase IIb ACTION II trial [59]. Therefore, larger trials are currently not
planned. It is important to review the significance of reperfusion injuries
Otaplimastat is a small molecule with a quinazoline-2,4-dione scaf­ presenting after MT. Five trials (MR CLEAN [4], REVASCAT [5],
fold that improves neurological outcomes through anti-excitotoxic ef­ ESCAPE [6], SWIFT PRIME [7], and EXTEND-IA [8]) showed superior
fects via N-methyl-(D)-aspartate (NMDA) receptor-mediated efficacy and safety of MT compared to medical (tPA) treatment in pa­
excitotoxicity [60]. In a phase IIa trial examining the safety and efficacy tients with AIS [77]. Another recent study reported a high reperfusion
of combined otaplimastat and tPA treatment in AIS patients, otaplima­ ratio (80%) after MT [40]. HT after MT occurred at a rate of 11.6% in
stat was found to be safe for use and tended to show better functional another multicenter study [78]. However, there was no increase in
outcomes than the placebo group [61]. Currently, a phase IIb trial is mortality or symptomatic HT between the IV tPA and MT with tPA
ongoing to obtain more evidence of the clinical efficacy of otaplimastat. groups [79]. Although HT is common because of high reperfusion after
However, the functional efficacy of otaplimastat needs to be investi­ MT, symptomatic HT after MT is uncommon.
gated in larger trials. Several risk factors are thought to cause HT after MT (Table 2).
IV glyburide (glibenclamide) is a specific blocker of the adenosine Significant factors that cause HT after MT include current smoking, a
triphosphate (ATP) sensitive potassium (K+) channel (KATP channel) low Alberta Stroke Program Early CT Score (ASPECTS), use of general
and sulfonylurea receptor 1 (SUR1)-transient receptor potential mela­ anesthesia, unfavorable collaterals on angiography, and thromboem­
statin 4 (TRPM4) channel. Results of preclinical studies suggest that bolic migration [78,80]. The TICI-ASPECTS-glucose (TAG) score, which
blockade of the SUR1-TRPM4 channel in neurons, astrocytes, and is the combined score of thrombolysis in cerebral ischemia (TICI), AS­
endothelium substantially reduces cerebral edema in rodent models of PECTS, and glucose level, was associated with symptomatic HT after MT
stroke [62,63]. In addition, it was shown that, in activated brain endo­ [81]. Other factors causing HT include vitamin K antagonists (VKAs) and
thelium, tPA induced phasic secretion of MMP-9 and causes activation of high UA levels [82]. An advanced age and a high NIHSS score are
SUR1-TRPM4 channels [64]. Therefore, glyburide acts as a partial in­ relatively significant causative factors for HT after MT [78]. However,
hibitor of MMP-9, which is expected to be beneficial in the treatment of prior use of direct oral anticoagulation (DOAC) [83] and the presence of
stroke, especially in patients treated with tPA. A phase II trial of a severe leukoaraiosis [84,85] were not significantly associated with
combination of tPA and glyburide was performed. Midline shift, an symptomatic HT after MT.
established marker of brain swelling, was significantly reduced in gly­ Presence of unfavorable collaterals on angiography is one of the
buride group vs. controls at 72–96 h (P = 0.0006) [65]. Thus, the plasma main reasons for HT after MT. Collateral flow by angiogenesis and
levels of MMP-9 in glyburide group was lower than those in controls at arteriogenesis after ischemic stroke is stimulated by various chemokines
24–72 h (P = 0.006) [66]. However, primary outcome, modified Rankin and growth factors and supports ischemic border and functional re­
Scale 0–4, and the frequency of HT was no different between the groups covery after ischemic stroke [32,86]. However, the presence of and
(ClinicalTrials.gov; NCT01794182) [65,66]. Although the results of mechanisms underlying collateral development prior to ischemic stroke
primary efficacy outcome and suppression of HT were disappointing, are still unknown. Angiographic early venous filling is also a predictive
glyburide will serve to guide a phase III trial. factor of HT after MT [87,88]. Severe HT is also more likely to occur
Statins, 3-hydroxy-3-methylglutaryl-coenzyme A (HMG-CoA) when recanalization has failed [89]. This indicates that acute reperfu­
reductase inhibitors, have multifaceted effects, such as anti- sion stress may demand collateral flow and is more likely to induce HT
inflammatory and antioxidant effects [67]. Statins reduce the risk of under poor collateral flow. A recent study that used direct arterial
intracranial hemorrhage [68]. In addition, there is increasing evidence damage model with rough suture insertion showed that VEGF and MMP-
that early use of simvastatin and atorvastatin in animal models reduces
HT through anti-inflammatory effects, such as suppression of MMP Table 2
expression and protective effects on the BBB [69,70]. Low dose of statins Risk factors of hemorrhagic transformation after mechanical thrombectomy.
(for example, simvastatin; 20 mg, once daily) may be useful in reducing Risk factors References
HT [71]. In contrast, the STARS trial showed no HT suppression or Significant TAG score Montalvo et al. [81]
therapeutic effect of combined simvastatin (40 mg, once daily) and tPA Angiographic early venous Ohta et al. [87], Cartmell et al.
therapy [72]. Therefore, the effect of thrombolytic therapy on HT re­ filling [88]
Smoking Ohta et al. [87], Cartmell et al.
mains clinically unknown. Furthermore, the effects of statins with tPA in
[88]
the acute phase varied depending on each condition. ASPECTS Ohta et al. [87], Cartmell et al.
Uric acid (UA) is an endogenous antioxidant derived from purine [88]
metabolism. Administration of UA in the MCAO mouse model reduced General anesthesia Boisseau et al. [78]
infarction volume and ROS production [73]. UA also prevented the Angiographic unfavorable Boisseau et al. [78] Bang et al.,
collaterals [80]
production of superoxide in the ischemic arterial wall, suggesting that it
Thromboembolic migration Boisseau et al. [78]
might protect the cerebral vasculature after AIS [74]. Chamorro et al. VKAs Meinel et al. [83]
first described that higher endogenous levels of UA in patients with AIS Higher uric acid level Yuan et al. [82]
at clinical onset were associated with better stroke outcomes [75]. A Relatively Failed recanalization Parrilla et al. [89]
significant High age Boisseau et al. [78]
phase IIb/III trial (URICO-ICTUS) that recruited patients with AIS on tPA
Higher NIHSS Boisseau et al. [78]
treatment with UA or placebo did not demonstrate a significant differ­ Not significant DOAC Meinel et al. [83]
ence in clinical outcomes between the two patient cohorts. However, Severe leukoaraiosis Atchaneeyasakul et al. [84],
secondary analysis showed that UA administration decreased the inci­ Boulouis et al. [85]
dence of early neurological worsening [76]. Thus, a larger study is TAG, the combination score of thrombolysis in cerebral ischemia (TICI), AS­
planned to analyze the effects of combined therapy with UA and MT PECTS, and glucose level; ASPECTS, Alberta Stroke Program Early CT Score;
[74]. NIHSS, National Institutes of Health Stroke Scale; VKAs, vitamin K antagonists;
In summary, several candidates have been clinically evaluated for DOAC, direct oral anticoagulants.

4
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

9 upregulation may cause MT-induced BBB disruption in a rat stroke study might reveal the therapeutic potential of SMTP.
model [90]. The association of collateral angiogenesis by VEGF and Tenecteplase is a genetically engineered variant of tPA with a longer
MMP-9 may be the cause of HT after MT. However, further studies half-life and more fibrin specificity than tPA. These properties confer
mimicking human MT models are essential to investigate the precise tenecteplase with more complete clot lysis abilities and fewer bleeding
mechanisms underlying HT and unfavorable collaterals. complications. Among patients with AIS who received tenecteplase or
MT alone (without tPA) may counter one of the mechanisms of alteplase treatment within 4.5 h after the onset of the ischemic stroke
reperfusion injury. Results of the SKIP study, a randomized study of MT (EXTEND-IA TNK trial, a phase III clinical trial), the tenecteplase group
with and without IV tPA in patients with AIS, demonstrated that the had greater reperfusion rates and better clinical improvement at 24 h
favorable outcome rates at 90 days were not significantly different be­ than the tPA group [97]. The other phase III clinical trial, the Norwegian
tween the two groups (p = 0.18 for non-inferiority). On the other hand, Tenecteplase Stroke Trial (NOR-TEST), showed a similar efficacy and
HT was significantly lower in the MT group without IV tPA (direct MT) safety of tenecteplase, compared with alteplase, in larger groups of pa­
than in the MT group with IV tPA (33.7% and 50.5%, respectively, p = tients [98]. However, intravenous tenecteplase administration prior to
0.02) [39]. In addition, another study demonstrated that the functional MT in patients with large vessel occlusion ischemic stroke (The
outcome at 90 days in the direct MT group within 206 min of stroke EXTEND-IA TNK Part 2 Randomized Clinical Trial) showed that a high
onset was non-inferior to that in the MT group with IV tPA within 215 dose of tenecteplase did not significantly improve cerebral reperfusion
min of stroke onset [40]. This study demonstrated that asymptomatic prior to MT as compared to a low dose of tenecteplase [99]. Neverthe­
HT in the direct MT (33.3%) and MT with IV tPA (36.2%) groups, and less, a low dose of tenecteplase was safe, and high reperfusion was noted
symptomatic HT in the direct MT (4.3%) and MT with IV tPA (6.1%) before MT in patients with large vessel occlusion ischemic stroke. Ten­
groups, were not statistically different. Although the ratio of asymp­ ecteplase, which has fewer bleeding complications than other agents, is
tomatic HT in both groups was higher than that reported in other studies clinically replacing alteplase, and a next-generation thrombolytic agent
in the HERMES reanalysis (4%) [91], tPA had no effect on HT within the is also being investigated.
therapeutic time window. These results show that MT with and without
tPA within 4.5 h after symptom onset did not increase the risk of HT. A 4.3. Combination treatments from experimental studies to clinical
time-dependent mechanism might be very important for HT after MT. applications

4. Discussion To conduct successful clinical trials of protective agents, animal


models and experimental methods that mimic real patient populations
4.1. Next-generation therapeutic strategies are essential. Although preclinical studies of protective agents in animal
models have shown many positive outcomes, only a few successful drugs
Patients with AIS, which developed 4.5–9 h from stroke onset, or can be assessed clinically [100,101]. One of the reasons for this is the
those with wake-up stroke with salvageable brain tissue identified by dissociation between the animal model and the actual patient popula­
DWI-FLAIR mismatch or perfusion image who were treated with alte­ tion. Existing models of animal experiments mostly involved young
plase achieved better functional outcomes than those who received males with no complications [100]. In many cases, the endpoint of
placebo [14,15]. However, the rate of symptomatic HT was high with treatment response was the infarction volume, which is dissociated from
alteplase treatment, although this increase did not negate the overall net clinical trials that often evaluate treatment efficacy using the modified
benefit of thrombolysis. The guidelines for MT were revised to extend Rankin score at 90 days [101]. Models and methods of animal experi­
the therapeutic time window (within 16 or 24 h after the onset of ments should be considered to find effective therapeutic agents in
symptoms) if patients met the DAWN or DEFUSE 3 eligibility criteria clinical practice. Long-term functional assessment is needed in both
[17,18]. The next-generation therapeutic strategies are as follows: (1) male and female animal models with older age and complications. For
development of new thrombolytic agents, and (2) combination treat­ clinical application of combination treatments, evaluation by suitable
ments with protective and thrombolytic agents or with thrombolytic models should be essential in pre-clinical studies.
agents and MT. Theoretically, the increased risk of HT after MT does not differ
significantly from that after tPA treatment. However, some recent
4.2. New thrombolytic agents studies have shown that the risk of HT in MT without tPA may be lower
than that with tPA treatment alone [39,40]. Thus, the neuroprotectant,
Some new thrombolytic agents that have superior effects than alte­ nerinetide, which interferes with post-synaptic density protein 95, did
plase have been described recently. not show efficiency in combination with IV tPA and MT in the primary
Desmoteplase (Desmodus rotundus salivary plasminogen activator) endpoint [102]. However, the MT without IV tPA group showed sig­
was thought to be an alternative to alteplase. It has demonstrated nificant functional improvement following the administration of ner­
minimal neurotoxicity, fibrin specificity, and a long half-life. However, a inetide (59.3%) compared to placebo (49.8%). The absence of benefit of
randomized, placebo-controlled, phase III clinical trial (the prematurely nerinetide in the alteplase group was likely because of enzymatic
terminated DIAS-4 trial), together with a post hoc pooled analysis of the cleavage of nerinetide by plasmin, leading to subtherapeutic concen­
concomitant DIAS-3, DIAS-4, and DIAS-J (Japan) trials that enrolled trations of nerinetide. From these results, the effects of MT with candi­
patients with AIS presenting within 3–9 h after onset, did not show date protectants without tPA should be further investigated.
beneficial outcomes when administered to patients with ischemic stroke
[92]. 5. Conclusion and future research direction
Stachybotrys microspora triphenyl phenol-7 (SMTP-7) may be a
candidate thrombolytic agent. The small SMTP-7 molecule promotes Clinical trials for MT showed that 71 to 80% of patients presented
thrombolysis and suppresses inflammation after reperfusion by inhib­ recanalization [40,91]. However, only half or fewer patients presented
iting the effects of proinflammatory cytokines [93,94]. Although SMTP- favorable neurological outcomes. These results suggest that MT provides
7 decreased infarction volume, HT, mortality, and neurological deficits insufficient therapeutic efficacy. Patients eligible for tPA should receive
in rodents and primate focal cerebral ischemia models [95], another IV tPA even if MT is being considered. To take advantage of the easily
report showed that a small number of embolic strokes in monkeys ten­ accessible drug tPA and the high recanalization rate of MT, it is
ded to cause HT-associated premature death [96]. However, a phase II important to reduce bleeding complications after recanalization. In
clinical trial is currently under investigation (JapicCTI-183,842; Japan addition, cell therapies are gaining attention in the treatment of stroke
Pharmaceutical Information Center Clinical Trials Information). This because they seem to have a protective function [103]. The

5
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

endovascular suture model in rodents with occlusion of the MCA may population-based stroke registry, Stroke. 46 (2015) 3437–3442, https://doi.org/
10.1161/STROKEAHA.115.011050.
closely mimic the situation of MT [90,100]. Cell therapies involving
[12] K.R. Lees, E. Bluhmki, R. von Kummer, T.G. Brott, D. Toni, J.C. Grotta, et al.,
protective phenotypes in a pre-clinical suture model facilitate functional Time to treatment with intravenous alteplase and outcome in stroke: an updated
recovery after post-reperfusion treatment [104,105]. It is unknown pooled analysis of ECASS, ATLANTIS, NINDS, and EPITHET trials, Lancet 375
whether administrations of both bone marrow-derived progenitor cells (2010) 1695–1703, https://doi.org/10.1016/S0140-6736(10)60491-6.
[13] W.N. Whiteley, J. Emberson, K.R. Lees, L. Blackwell, G. Albers, E. Bluhmki, et al.,
(MASTERS) [106] and stem cells [107] in phase II clinical trials show Risk of intracerebral haemorrhage with alteplase after acute ischaemic stroke: a
functional improvement and suppression of HT. However, a half of the secondary analysis of an individual patient data meta-analysis, Lancet Neurol. 15
patients underwent IV tPA and MT in these trials. The safety of the (2016) 925–933, https://doi.org/10.1016/S1474-4422(16)30076-X.
[14] G. Thomalla, C.Z. Simonsen, F. Boutitie, G. Andersen, Y. Berthezene, B. Cheng, et
combination of reperfusion therapies and cell therapies has previously al., MRI-guided thrombolysis for stroke with unknown time of onset, N. Engl. J.
been shown. Following the results, clinical trials are ongoing to evaluate Med. 379 (2018) 611–622, https://doi.org/10.1056/nejmoa1804355.
the safety and efficacy of administration of bone marrow-derived pro­ [15] B.C. Campbell, H. Ma, P.A. Ringleb, M.W. Parsons, L. Churilov, M. Bendszus, et
al., Extending thrombolysis to 4⋅5-9 h and wake-up stroke using perfusion
genitor cells and umbilical cord mesenchymal stem cells after IV tPA or imaging: a systematic review and meta-analysis of individual patient data, Lancet.
MT (ClinicalTrials.gov; NCT02961504 (MASTERS-2) and 394 (2019) 139–147, https://doi.org/10.1016/s0140-6736(19)31053-0.
NCT04434768, respectively). Future research should facilitate the re­ [16] G. Tsivgoulis, A.H. Katsanos, K. Malhotra, A. Sarraj, A.D. Barreto, M. Köhrmann,
et al., Thrombolysis for acute ischemic stroke in the unwitnessed or extended
covery of neurological function by combining reperfusion therapies with therapeutic time window, Neurology. 94 (2020) e1241–e1248, https://doi.org/
cell therapies and/or investigate pleiotropic protective agents for 10.1212/WNL.0000000000008904.
treatment of HT. [17] R.G. Nogueira, A.P. Jadhav, D.C. Haussen, A. Bonafe, R.F. Budzik, P. Bhuva,
DAWN Trial Investigators, et al., Thrombectomy 6 to 24 hours after stroke with a
mismatch between deficit and infarct, New Engl. J. Med. 378 (2018) 11–21,
https://doi.org/10.1056/NEJMoa1706442.
Declaration of competing interest
[18] G.W. Albers, M.P. Marks, S. Kemp, S. Christensen, J.P. Tsai, S. Ortega-Gutierrez,
et al., Thrombectomy for stroke at 6 to 16 hours with selection by perfusion
TS is an academic adviser for ShimoJani LLC, a biotechnology imaging, N. Engl. J. Med. 378 (2018) 708–718, https://doi.org/10.1056/
NEJMoa1713973.
company.
[19] W.M. Yu, A.H. Abdul-Rahim, A.C. Cameron, J. Kõrv, P. Sevcik, D. Toni, K.R. Lees,
S.I.T.S. Scientific Committee, The incidence and associated factors of early
Acknowledgments neurological deterioration after thrombolysis: results from SITS registry, Stroke.
51 (2020) 2705–2714, https://doi.org/10.1161/STROKEAHA.119.028287.
[20] S. Yaghi, J.Z. Willey, B. Cucchiara, J.N. Goldstein, N.R. Gonzales, P. Khatri, et al.,
This work was supported by a Grant-in-Aid for Scientific Research Treatment and outcome of hemorrhagic transformation after intravenous
(Research Project Number: 18K07493), a grant from SENSHIN Medical alteplase in acute ischemic stroke: a scientific statement for healthcare
professionals from the American Heart Association/American Stroke Association,
Research Foundation, Takeda Science Foundation and Astellas Foun­
Stroke. 48 (2017) e343–e361, https://doi.org/10.1161/
dation for Research on Metabolic Disorders, and Medical Research STR.0000000000000152.
Encouragement Prize of the Japan Medical Association and Niigata [21] F. Caparros, G. Kuchcinski, A. Drelon, B. Casolla, S. Moulin, N. Dequatre-
Ponchelle, et al., Use of MRI to predict symptomatic haemorrhagic transformation
Medical Association (Dr. Kanazawa).
after thrombolysis for cerebral ischaemia, J. Neurol. Neurosurg. Psychiatry 91
(2020) 402–410, https://doi.org/10.1136/jnnp-2019-321904.
References [22] G. Tsivgoulis, A.H. Katsanos, D. Mavridis, V. Lambadiari, C. Roffe, M.J. Macleod,
et al., Association of baseline hyperglycemia with outcomes of patients with and
without diabetes with acute ischemic stroke treated with intravenous
[1] The National Institute of Neurological Disorders and Stroke rt-PA Stroke Study
thrombolysis: a propensity score-matched analysis from the SITS-ISTR registry,
Group, Tissue plasminogen activator for acute ischemic stroke, N. Engl. J. Med.
Diabetes. 68 (2019) 1861–1869, https://doi.org/10.2337/db19-0440.
333 (1995) 1581–1587, https://doi.org/10.1056/NEJM199512143332401.
[23] T. Shimohata, M. Kanazawa, K. Kawamura, T. Takahashi, M. Nishizawa,
[2] G.J. del Zoppo, J.L. Saver, E.C. Jauch, H.P. Adams Jr., Expansion of the time
Therapeutic strategies to attenuate hemorrhagic transformation after tissue
window for treatment of acute ischemic stroke with intravenous tissue
plasminogen activator treatment for acute ischemic stroke, Neurol. Clin.
plasminogen activator: a science advisory from the American Heart Association/
Neurosci. 1 (2013) 201–208, https://doi.org/10.1111/ncn3.63.
American Stroke Association, Stroke. 40 (2009) 2945–2948, https://doi.org/
[24] G.C. Jickling, D. Liu, B. Stamova, B.P. Ander, X. Zhan, A. Lu, et al., Hemorrhagic
10.1161/STROKEAHA.109.192535.
transformation after ischemic stroke in animals and humans, J. Cereb. Blood Flow
[3] W.J. Powers, A.A. Rabinstein, T. Ackerson, O.M. Adeoye, N.C. Bambakidis,
Metab. 34 (2014) 185–199, https://doi.org/10.1038/jcbfm.2013.203.
K. Becker, et al., 2018 guidelines for the early Management of Patients with Acute
[25] M. Kanazawa, T. Takahashi, M. Nishizawa, T. Shimohata, Therapeutic strategies
Ischemic Stroke : a guideline for healthcare professionals from the American
to attenuate Hemorrhagic transformation after tissue plasminogen activator
Heart Association/American Stroke Association, Stroke. 49 (2018) e46–e110,
treatment for acute ischemic stroke, J. Atheroscler. Thromb. 24 (2017) 240–253,
https://doi.org/10.1161/STR.0000000000000158.
https://doi.org/10.5551/jat.RV16006.
[4] O.A. Berkhemer, P.S. Fransen, D. Beumer, L.A. van den Berg, H.F. Lingsma, A.
[26] O. Peters, T. Back, U. Lindauer, C. Busch, D. Megow, J. Dreier, et al., Increased
J. Yoo, et al., A randomized trial of intraarterial treatment for acute ischemic
formation of reactive oxygen species after permanent and reversible middle
stroke, N. Engl. J. Med. 372 (2015) 11–20, https://doi.org/10.1056/
cerebral artery occlusion in the rat, J. Cereb. Blood Flow Metab. 18 (1998)
NEJMoa1411587.
196–205, https://doi.org/10.1097/00004647-199802000-00011.
[5] T.G. Jovin, A. Chamorro, E. Cobo, M.A. de Miquel, C.A. Molina, A. Rovira, et al.,
[27] M. Kanazawa, K. Kawamura, T. Takahashi, M. Miura, Y. Tanaka, M. Koyama, et
Thrombectomy within 8 hours after symptom onset in ischemic stroke, N. Engl. J.
al., Multiple therapeutic effects of progranulin on experimental acute ischaemic
Med. 372 (2015) 2296–2306, https://doi.org/10.1056/NEJMoa1503780.
stroke, Brain. 138 (2015) 1932–1948, https://doi.org/10.1093/brain/awv079.
[6] M. Goyal, A.M. Demchuk, B.K. Menon, M. Eesa, J.L. Rempel, J. Thornton, et al.,
[28] M. Kanazawa, I. Ninomiya, M. Hatakeyama, T. Takahashi, T. Shimohata,
Randomized assessment of rapid endovascular treatment of ischemic stroke,
Microglia and monocytes/macrophages polarization reveal novel therapeutic
N. Engl. J. Med. 372 (2015) 1019–1030, https://doi.org/10.1056/
mechanism against stroke, Int. J. Mol. Sci. 18 (2017) 2135, https://doi.org/
NEJMoa1414905.
10.3390/ijms18102135.
[7] J.L. Saver, M. Goyal, A. Bonafe, H.C. Diener, E.I. Levy, V.M. Pereira, et al., Stent-
[29] J. Shi, H. Peng, S. You, Y. Liu, J. Xu, Y. Xu, et al., Increase in neutrophils after
retriever thrombectomy after intravenous t-PA vs. t-PA alone in stroke, N. Engl. J.
recombinant tissue plasminogen activator thrombolysis predicts poor functional
Med. 372 (2015) 2285–2295, https://doi.org/10.1056/NEJMoa1415061.
outcome of ischaemic stroke: a longitudinal study, Eur. J. Neurol. 25 (2018)
[8] B.C. Campbell, P.J. Mitchell, T.J. Kleinig, H.M. Dewey, L. Churilov, N. Yassi, et
687–e45, https://doi.org/10.1111/ene.13575.
al., Endovascular therapy for ischemic stroke with perfusion-imaging selection,
[30] L. Wang, Q. Song, C. Wang, S. Wu, L. Deng, Y. Li, et al., Neutrophil to lymphocyte
N. Engl. J. Med. 372 (2015) 1009–1018, https://doi.org/10.1056/
ratio predicts poor outcomes after acute ischemic stroke: a cohort study and
NEJMoa1414792.
systematic review, J. Neurol. Sci. 406 (2019) 116445, https://doi.org/10.1016/j.
[9] M. Marko, A. Posekany, S. Szabo, S. Scharer, S. Kiechl, M. Knoflach, et al., Trends
jns.2019.116445.
of r-tPA (recombinant tissue-type plasminogen activator) treatment and
[31] S. Won, J.K. Lee, D.G. Stein, Recombinant tissue plasminogen activator promotes,
treatment-influencing factors in acute ischemic stroke, Stroke. 51 (2020)
and progesterone attenuates, microglia/macrophage M1 polarization and
1240–1247, https://doi.org/10.1161/STROKEAHA.119.027921.
recruitment of microglia after MCAO stroke in rats, Brain Behav. Immun. 49
[10] D. Mozaffarian, E.J. Benjamin, A.S. Go, D.K. Arnett, M.J. Blaha, M. Cushman, et
(2015) 267–279, https://doi.org/10.1016/j.bbi.2015.06.007.
al., Executive summary: heart disease and stroke statistics-2016 update: a report
[32] M. Kanazawa, T. Takahashi, M. Ishikawa, O. Onodera, T. Shimohata, G.J. del
from the American Heart Association, Circulation. 133 (2016) 447–454, https://
Zoppo, Angiogenesis in the ischemic core: a potential treatment target? J. Cereb.
doi.org/10.1161/CIR.0000000000000366.
Blood Flow Metab. 39 (2019) 753–769, https://doi.org/10.1177/
[11] X. Urra, S. Abilleira, L. Dorado, M. Ribó, P. Cardona, M. Millán, et al., Mechanical
0271678X19834158.
thrombectomy in and outside the REVASCAT trial insights from a concurrent

6
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

[33] M. Hatakeyama, I. Ninomiya, M. Kanazawa, Angiogenesis and neuronal [55] S. Ghosh, E. Goldin, F.H. Gordon, H.A. Malchow, J. Rask-Madsen, P. Rutgeerts,
remodeling after ischemic stroke, Neural Regen. Res. 15 (2020) 16–19, https:// P. Vyhnálek, Z. Zádorová, T. Palmer, S. Donoghue, Natalizumab for active
doi.org/10.4103/1673-5374.264442. Crohn’s disease, N. Engl. J. Med. 348 (2003) 24–32, https://doi.org/10.1056/
[34] M. Kanazawa, H. Igarashi, K. Kawamura, T. Takahashi, A. Kakita, H. Takahashi, NEJMoa020732.
T. Nakada, M. Nishizawa, T. Shimohata, Inhibition of VEGF signaling pathway [56] D.H. Miller, O.A. Khan, W.A. Sheremata, L.D. Blumhardt, G.P.A. Rice, M.
attenuates hemorrhage after tPA treatment, J. Cereb. Blood Flow Metab. 31 A. Libonati, A.J. Willmer-Hulme, C.M. Dalton, K.A. Miszkiel, P.W. O’Connor,
(2011) 1461–1474, https://doi.org/10.1038/jcbfm.2011.9. A controlled trial of natalizumab for relapsing multiple sclerosis, N. Engl. J. Med.
[35] K. Kawamura, T. Takahashi, M. Kanazawa, H. Igarashi, T. Nakada, M. Nishizawa, 348 (2003) 15–23, https://doi.org/10.1056/NEJMoa020696.
et al., Effects of angiopoietin-1 on hemorrhagic transformation and cerebral [57] L. Kappos, D. Bates, G. Edan, M. Eraksoy, A. Garcia-Merino, N. Grigoriadis, et al.,
edema after tissue plasminogen activator treatment for ischemic stroke in rats, Natalizumab treatment for multiple sclerosis: updated recommendations for
PLoS One 9 (2014), e98639, https://doi.org/10.1371/journal.pone.0098639. patient selection and monitoring, Lancet Neurol. 10 (2011) 745–758, https://doi.
[36] Y. Murata, A. Rosell, R.H. Scannevin, K.J. Rhodes, X. Wang, E.H. Lo, Extension of org/10.1016/S1474-4422(11)70149-1.
the thrombolytic time window with minocycline in experimental stroke, Stroke. [58] J. Elkins, R. Veltkamp, J. Montaner, S.C. Johnston, A.B. Singhal, K. Becker, et al.,
39 (2008) 3372–3377, https://doi.org/10.1161/STROKEAHA.108.514026. Safety and efficacy of natalizumab in patients with acute ischaemic stroke
[37] S. Ismael, S. Nasoohi, A. Yoo, H.A. Ahmed, T. Ishrat, Tissue plasminogen activator (ACTION): a randomised, placebo-controlled, double-blind phase 2 trial, Lancet
promotes TXNIP-NLRP3 inflammasome activation after hyperglycemic stroke in Neurol. 16 (2017) 217–226, https://doi.org/10.1016/S1474-4422(16)30357-X.
mice, Mol. Neurobiol. 57 (2020) 2495–2508, https://doi.org/10.1007/s12035- [59] M. Elkind, R. Veltkamp, J. Montaner, S.C. Johnston, A.B. Singhal, K. Becker, et
020-01893-7. al., Natalizumab in acute ischemic stroke (ACTION II): a randomized, placebo-
[38] S. Won, J.H. Lee, B. Wali, D.G. Stein, I. Sayeed, Progesterone attenuates controlled trial, Neurology. 95 (2020) e1091–e1104, https://doi.org/10.1212/
hemorrhagic transformation after delayed tPA treatment in an experimental WNL.0000000000010038.
model of stroke in rats: involvement of the VEGF-MMP pathway, J. Cereb. Blood [60] S.J. Noh, J.M. Lee, K.S. Lee, H.S. Hong, C.K. Lee, I.H. Cho, et al., SP-8203 shows
Flow Metab. 34 (2014) 72–80, https://doi.org/10.1038/jcbfm.2013.163. neuroprotective effects and improves cognitive impairment in ischemic brain
[39] K. Suzuki, K. Kimura, M. Takeuchi, M. Morimoto, R. Kanazawa, Y. Kamiya, , et al. injury through NMDA receptor, Pharmacol. Biochem. Behav. 100 (2011) 73–80,
SKIP study investigators, The randomized study of endovascular therapy with https://doi.org/10.1016/j.pbb.2011.07.018.
versus without intravenous tissue plasminogen activator in acute stroke with ICA [61] J.S. Kim, K.B. Lee, J.H. Park, S.M. Sung, K. Oh, E.G. Kim, et al., Safety and efficacy
and M1 occlusion (SKIP study), in: International Stroke Conference 2020, Los of otaplimastat in patients with acute ischemic stroke requiring tPA (SAFE-TPA):
Angeles, CA, USA, 2020. https://www.professional.heart.org/professional/Scie a multicenter, randomized, double-blind, placebo-controlled phase 2 study, Ann.
nceNews/UCM_505646_SKIP-Study-Clinical-Trial-Details.jsp. Neurol. 87 (2020) 233–245, https://doi.org/10.1002/ana.25644.
[40] P. Yang, Y. Zhang, L. Zhang, Y. Zhang, K.M. Treurniet, W. Chen, et al., [62] J.M. Simard, M. Chen, K.V. Tarasov, S. Bhatta, S. Ivanova, L. Melnitchenko, et al.,
Endovascular thrombectomy with or without intravenous alteplase in acute Newly expressed SUR1-regulated NC(ca-ATP) channel mediates cerebral edema
stroke, N. Engl. J. Med. 382 (2020) 1981–1993, https://doi.org/10.1056/ after ischemic stroke, Nat. Med. 12 (2006) 433–440, https://doi.org/10.1038/
NEJMoa2001123. nm1390.
[41] T. Yamashita, T. Kamiya, K. Deguchi, T. Inaba, H. Zhang, J. Shang, et al., [63] J.M. Simard, N. Tsymbalyuk, O. Tsymbalyuk, S. Ivanova, V. Yurovsky,
Dissociation and protection of the neurovascular unit after thrombolysis and V. Gerzanich, Glibenclamide is superior to decompressive craniectomy in a rat
reperfusion in ischemic rat brain, J. Cereb. Blood Flow Metab. 29 (2009) model of malignant stroke, Stroke. 41 (2010) 531–537, https://doi.org/10.1161/
715–725, https://doi.org/10.1038/jcbfm.2008.164. STROKEAHA.109.572644.
[42] T. Yamaguchi, H. Awano, H. Matsuda, N. Tanahashi, Edaravone with and without [64] V. Gerzanich, M.S. Kwon, S.K. Woo, A. Ivanov, J.M. Simard, SUR1-TRPM4
.6 mg/kg alteplase within 4.5 hours after ischemic stroke: A prospective cohort channel activation and phasic secretion of MMP-9 induced by tPA in brain
study (PROTECT4.5), J. Stroke Cerebrovasc. Dis. 26 (2017) 756–765, https://doi. endothelial cells, PLoS One 13 (2018), e0195526, https://doi.org/10.1371/
org/10.1016/j.jstrokecerebrovasdis.2016.10.011. journal.pone.0195526.
[43] S. Feng, Q. Yang, M. Liu, W. Li, W. Yuan, S. Zhang, et al., Edaravone for acute [65] K.N. Sheth, J.J. Elm, B.J. Molyneaux, H. Hinson, L.A. Beslow, G.K. Sze, et al.,
ischaemic stroke, Cochrane Database Syst. Rev. 12 (2011), CD007230, https:// Safety and efficacy of intravenous glyburide on brain swelling after large
doi.org/10.1002/14651858.CD007230.pub2. hemispheric infarction (GAMES-RP): a randomised, double-blind, placebo-
[44] A. Shuaib, K.R. Lees, P. Lyden, J. Grotta, A. Davalos, S.M. Davis, H.C. Diener, controlled phase 2 trial, Lancet Neurol. 15 (2016) 1160–1169, https://doi.org/
T. Ashwood, W.W. Wasiewski, U. Emeribe, SAINT. Trial Investigators, NXY-059 10.1016/S1474-4422(16)30196-X.
for the treatment of acute ischemic stroke, New Engl. J. Med. 357 (2007) [66] W.T. Kimberly, M.B. Bevers, R. von Kummer, A.M. Demchuk, J.M. Romero, J.
562–571, https://doi.org/10.1056/NEJMoa070240. J. Elm, et al., Effect of IV glyburide on adjudicated edema endpoints in the
[45] J. Aoki, K. Kimura, N. Morita, M. Harada, N. Metoki, Y. Tateishi, et al., YAMATO GAMES-RP Trial, Neurology 91 (2018), https://doi.org/10.1212/
study (tissue-type plasminogen activator and edaravone combination therapy), WNL.0000000000006618 e1–e7.
Stroke. 48 (2017) 712–719, https://doi.org/10.1161/STROKEAHA.116.015042. [67] L. García-Bonilla, M. Campos, D. Giralt, D. Salat, P. Chacón, Hernández-
[46] L. Kappos, J. Antel, G. Comi, X. Montalban, P. O’Connor, C.H. Polman, et al., Oral Guillamon, et al., Evidence for the efficacy of statins in animal stroke models: a
fingolimod (FTY720) for relapsing multiple sclerosis, N. Engl. J. Med. 355 (2006) meta-analysis, J. Neurochem. 122 (2012) 233–243, https://doi.org/10.1111/
1124–1140, https://doi.org/10.1056/NEJMoa052643. j.1471-4159.2012.07773.x.
[47] F. Campos, T. Qin, J. Castillo, J.H. Seo, K. Arai, E.H. Lo, et al., Fingolimod reduces [68] W. Saliba, H.S. Rennert, O. Barnett-Griness, N. Gronich, J. Molad, G. Rennert, et
hemorrhagic transformation associated with delayed tissue plasminogen activator al., Association of statin use with spontaneous intracerebral hemorrhage: a cohort
treatment in a mouse thromboembolic model, Stroke. 44 (2013) 505–511, study, Neurology. 91 (2018) e400–e409, https://doi.org/10.1212/
https://doi.org/10.1161/STROKEAHA.112.679043. WNL.0000000000005907.
[48] D.C. Tian, K. Shi, Z. Zhu, J. Yao, X. Yang, L. Su, et al., Fingolimod enhances the [69] B. Reuter, C. Rodemer, S. Grudzenski, S. Meairs, P. Bugert, M.G. Hennerici, et al.,
efficacy of delayed alteplase administration in acute ischemic stroke by Effect of simvastatin on MMPs and TIMPs in human brain endothelial cells and
promoting anterograde reperfusion and retrograde collateral flow, Ann. Neurol. experimental stroke, Transl. Stroke Res. 6 (2015) 156–159, https://doi.org/
84 (2018) 717–728, https://doi.org/10.1002/ana.25352. 10.1007/s12975-014-0381-7.
[49] V. Brundula, N.B. Rewcastle, L.M. Metz, C.C. Bernard, V.W. Yong, Targeting [70] X. Fang, D. Tao, J. Shen, Y. Wang, X. Dong, X. Ji, Neuroprotective effects and
leukocyte MMPs and transmigration: minocycline as a potential therapy for dynamic expressions of MMP9 and TIMP1 associated with atorvastatin
multiple sclerosis, Brain. 125 (2002) 1297–1308, https://doi.org/10.1093/brain/ pretreatment in ischemia - reperfusion rats, Neurosci. Lett. 603 (2015) 60–65,
awf133. https://doi.org/10.1016/j.neulet.2015.07.013.
[50] Y. Lampl, M. Boaz, R. Gilad, M. Lorberboym, R. Dabby, Rapoport, et al., [71] Y. Wu, D. Lu, A. Xu, The effect of HMG-CoA reductase inhibitors on thrombolysis-
Minocycline treatment in acute stroke: an open-label, evaluator-blinded study, induced haemorrhagic transformation, J. Clin. Neurosci. 69 (2019) 1–6, https://
Neurology. 69 (2007) 1404–1410, https://doi.org/10.1212/01. doi.org/10.1016/j.jocn.2019.08.074.
wnl.0000277487.04281.db. [72] J. Montaner, A. Bustamante, S. García-Matas, M. Martínez-Zabaleta, C. Jiménez,
[51] M.V.P. Srivastava, A. Bhasin, R. Bhatia, A. Garg, S. Gaikwad, K. Prasad, et al., J. de la Torre, et al., Combination of thrombolysis and statins in acute stroke is
Efficacy of minocycline in acute ischemic stroke: a single-blinded, placebo- safe, Stroke. 47 (2016) 2870–2873, https://doi.org/10.1161/
controlled trial, Neurol. India 60 (2012) 23–28, https://doi.org/10.4103/0028- strokeaha.116.014600.
3886.93584. [73] F. Haberman, S.C. Tang, T.V. Arumugam, D.H. Hyun, Q.S. Yu, R.G. Cutler, et al.,
[52] K. Malhotra, J.J. Chang, A. Khunger, D. Blacker, J.A. Switzer, N. Goyal, et al., Soluble neuroprotective antioxidant uric acid analogs ameliorate ischemic brain
Minocycline for acute stroke treatment: a systematic review and meta-analysis of injury in mice, NeuroMolecular Med. 9 (2007) 315–323, https://doi.org/
randomized clinical trials, J. Neurol. 265 (2018) 1871–1879, https://doi.org/ 10.1007/s12017-007-8010-1.
10.1007/s00415-018-8935-3. [74] Á. Chamorro, Neuroprotectants in the era of reperfusion therapy, J. Stroke. 20
[53] Z. Sheng, Y. Liu, H. Li, W. Zheng, B. Xia, X. Zhang, et al., Efficacy of minocycline (2018) 197–207, https://doi.org/10.5853/jos.2017.02901.
in acute ischemic stroke: a systematic review and meta-analysis of rodent and [75] Á. Chamorro, V. Obach, A. Cervera, M. Revilla, R. Deulofeu, J.H. Aponte,
clinical studies, Front. Neurol. 9 (2018) 1103, https://doi.org/10.3389/ Prognostic significance of uric acid serum concentration in patients with acute
fneur.2018.01103. ischemic stroke, Stroke. 33 (2002) 1048–1052, https://doi.org/10.1161/
[54] T.A. Yednock, C. Cannon, L.C. Fritz, F. Sanchez-Madrid, L. Steinman, N. Karin, hs0402.105927.
Prevention of experimental autoimmune encephalomyelitis by antibodies against [76] S. Amaro, C. Laredo, A. Renú, L. Llull, S. Rudilosso, V. Obach, et al., Uric acid
alpha 4 beta 1 integrin, Nature. 356 (1992) 63–66, https://doi.org/10.1038/ therapy prevents early ischemic stroke progression: a tertiary analysis of the
356063a0. URICO-ICTUS trial (efficacy study of combined treatment with uric acid and r-tPA

7
Y. Otsu et al. Journal of the Neurological Sciences 419 (2020) 117217

in acute ischemic stroke), Stroke. 47 (2016) 2874–2876, https://doi.org/ [92] R. von Kummer, E. Mori, T. Truelsen, J.S. Jensen, B.A. Grønning, J.B. Fiebach, et
10.1161/STROKEAHA.116.014672. al., Desmoteplase 3 to 9 hours after major artery occlusion stroke: the DIAS-4 trial
[77] J.L. Saver, M. Goyal, A. van der Lugt, B.K. Menon, C.B. Majoie, D.W. Dippel, et al., (efficacy and safety study of desmoteplase to treat acute ischemic stroke), Stroke.
Time to treatment with endovascular thrombectomy and outcomes from ischemic 47 (2016) 2880–2887, https://doi.org/10.1161/STROKEAHA.116.013715.
stroke: a meta-analysis, JAMA. 316 (2016) 1279–1288, https://doi.org/10.1001/ [93] T. Miyazaki, Y. Kimura, H. Ohata, T. Hashimoto, K. Shibata, K. Hasumi, et al.,
jama.2016.13647. Distinct effects of tissue-type plasminogen activator and SMTP-7 on
[78] W. Boisseau, R. Fahed, B. Lapergue, J.P. Desilles, K. Zuber, N. Khoury, et al., cerebrovascular inflammation following thrombolytic reperfusion, Stroke 42
Predictors of parenchymal hematoma after mechanical thrombectomy: a (2011) 1097–1104, https://doi.org/10.1161/STROKEAHA.110.598359.
multicenter study, Stroke. 50 (2019) 2364–2370, https://doi.org/10.1161/ [94] H. Sawada, N. Nishimura, E. Suzuki, J. Zhuang, K. Hasegawa, H. Takamatsu, et
STROKEAHA.118.024512. al., SMTP-7, a novel small-molecule thrombolytic for ischemic stroke: a study in
[79] C.K. Bush, D. Kurimella, L.J. Cross, K.R. Conner, S. Martin-Schild, J. He, et al., rodents and primates, J. Cereb. Blood Flow Metab. 34 (2014) 235–241, https://
Endovascular treatment with stent-retriever devices for acute ischemic stroke: a doi.org/10.1038/jcbfm.2013.191.
meta-analysis of randomized controlled trials, PLoS One 11 (2016), e0147287, [95] A. Ito, K. Niizuma, H. Shimizu, M. Fujimura, K. Hasumi, T. Tominaga, SMTP-7, a
https://doi.org/10.1371/journal.pone.0147287. new thrombolytic agent, decreases hemorrhagic transformation after transient
[80] O.Y. Bang, J.L. Saver, S.J. Kim, G.M. Kim, C.S. Chung, B. Ovbiagele, et al., middle cerebral artery occlusion under warfarin anticoagulation in mice, Brain
Collateral flow averts hemorrhagic transformation after endovascular therapy for Res. 1578 (2014) 38–48, https://doi.org/10.1016/j.brainres.2014.07.004.
acute ischemic stroke, Stroke. 42 (2011) 2235–2239, https://doi.org/10.1161/ [96] E. Suzuki, N. Nishimura, T. Yoshikawa, Y. Kunikiyo, K. Hasegawa, K. Hasumi,
STROKEAHA.110.604603. Efficacy of SMTP-7, a small-molecule anti-inflammatory thrombolytic, in embolic
[81] M. Montalvo, E. Mistry, A.D. Chang, A. Yakhkind, K. Dakay, I. Azher, et al., stroke in monkeys, Pharmacol. Res. Perspect. 6 (2018), e00448, https://doi.org/
Predicting symptomatic intracranial haemorrhage after mechanical 10.1002/prp2.448.
thrombectomy: the TAG score, J. Neurol. Neurosurg. Psychiatry 90 (2019) [97] B.C. Campbell, P.J. Mitchell, L. Churilov, N. Yassi, T.J. Kleinig, R.J. Dowling, et
1370–1374, https://doi.org/10.1136/jnnp-2019-321184. al., Tenecteplase versus alteplase before thrombectomy for ischemic stroke, New
[82] K. Yuan, X. Zhang, J. Chen, S. Li, D. Yang, Y. Xie, et al., Uric acid level and risk of Engl. J. Med. 378 (2018) 1573–1582, https://doi.org/10.1056/NEJMoa1716405.
symptomatic intracranial haemorrhage in ischaemic stroke treated with [98] N. Logallo, V. Novotny, J. Assmus, C.E. Kvistad, L. Alteheld, O.M. Rønning, et al.,
endovascular treatment, Eur. J. Neurol. 27 (2020) 1048–1055, https://doi.org/ Tenecteplase versus alteplase for management of acute ischaemic stroke (NOR-
10.1111/ene.14202. TEST): a phase 3, randomised, open-label, blinded endpoint trial, Lancet Neurol.
[83] T.R. Meinel, J.U. Kniepert, D.J. Seiffge, J. Gralla, S. Jung, E. Auer, et al., 16 (2017) 781–788, https://doi.org/10.1016/S1474-4422(17)30253-3.
Endovascular stroke treatment and risk of intracranial hemorrhage in [99] B.C. Campbell, P.J. Mitchell, L. Churilov, N. Yassi, T.J. Kleinig, R.J. Dowling, et
anticoagulated patients, Stroke. 51 (2020) 892–898, https://doi.org/10.1161/ al., Effect of intravenous tenecteplase dose on cerebral reperfusion before
STROKEAHA.119.026606. thrombectomy in patients with large vessel occlusion ischemic stroke: the
[84] K. Atchaneeyasakul, T. Leslie-Mazwi, K. Donahue, A.K. Giese, N.S. Rost, White EXTEND-IA TNK part 2 randomized clinical trial, JAMA 323 (2020) 1257–1265.
matter hyperintensity volume and outcome of mechanical thrombectomy with Advance online publication, https://doi.org/10.1001/jama.2020.1511.
stentriever in acute ischemic stroke, Stroke. 48 (2017) 2892–2894, https://doi. [100] C.J. Sommer, Ischemic stroke: experimental models and reality, Acta
org/10.1161/STROKEAHA.117.018653. Neuropathol. 133 (2017) 245–261, https://doi.org/10.1007/s00401-017-1667-0.
[85] G. Boulouis, N. Bricout, W. Benhassen, M. Ferrigno, G. Turc, M. Bretzner, et al., [101] A. Schmidt-Pogoda, N. Bonberg, M.H.M. Koecke, et al., Why Most acute stroke
White matter hyperintensity burden in patients with ischemic stroke treated with studies are positive in animals but not in patients: a systematic comparison of
thrombectomy, Neurology. 93 (2019) e1498–e1506, https://doi.org/10.1212/ preclinical, early phase, and phase 3 clinical trials of neuroprotective agents, Ann.
WNL.0000000000008317. Neurol. 87 (2020) 40–51, https://doi.org/10.1002/ana.25643.
[86] L. Wei, J.P. Erinjeri, C.M. Rovainen, T.A. Woolsey, Collateral growth and [102] M.D. Hill, M. Goyal, B.K. Menon, R.G. Nogueira, R.A. McTaggart, A.M. Demchuk,
angiogenesis around cortical stroke, Stroke. 32 (2001) 2179–2184, https://doi. et al., Efficacy and safety of nerinetide for the treatment of acute ischaemic stroke
org/10.1161/hs0901.094282. (ESCAPE-NA1): a multicentre, double-blind, randomised controlled trial, Lancet.
[87] H. Ohta, S. Nakano, K. Yokogami, T. Iseda, T. Yoneyama, S. Wakisaka, 395 (2020) 878–887, https://doi.org/10.1016/S0140-6736(20)30258-0.
Appearance of early venous filling during intra-arterial reperfusion therapy for [103] M. Hatakeyama, I. Ninomiya, Y. Otsu, K. Omae, Y. Kimura, O. Onodera, et al., Cell
acute middle cerebral artery occlusion: a predictive sign for hemorrhagic therapies under clinical trials and polarized cell therapies in pre-clinical studies to
complications, Stroke. 35 (2004) 893–898, https://doi.org/10.1161/01. treat ischemic stroke and neurological diseases: a literature review, Int. J. Mol.
STR.0000119751.92640.7F. Sci. 21 (2019) 6194, https://doi.org/10.3390/ijms21176194.
[88] S. Cartmell, R.L. Ball, R. Kaimal, N.A. Telischak, M.P. Marks, H.M. Do, et al., Early [104] M. Kanazawa, M. Miura, M. Toriyabe, M. Koyama, M. Hatakeyama, M. Ishikawa,
cerebral vein after endovascular ischemic stroke treatment predicts symptomatic et al., Microglia preconditioned by oxygen-glucose deprivation promote
reperfusion hemorrhage, Stroke. 49 (2018) 1741–1746, https://doi.org/10.1161/ functional recovery in ischemic rats, Sci. Rep. 7 (2017) 42582, https://doi.org/
STROKEAHA.118.021402. 10.1038/srep42582.
[89] G. Parrilla, B. García-Villalba, M. Espinosa de Rueda, J. Zamarro, E. Carrión, [105] M. Hatakeyama, M. Kanazawa, I. Ninomiya, K. Omae, Y. Kimura, T. Takahashi, et
F. Hernández-Fernández, et al., Hemorrhage/contrast staining areas after al., A novel therapeutic approach using peripheral blood mononuclear cells
mechanical intra-arterial thrombectomy in acute ischemic stroke: imaging preconditioned by oxygen-glucose deprivation, Sci. Rep. 9 (2019) 16819, https://
findings and clinical significance, Am. J. Neuroradiol. 33 (2012) 1791–1796, doi.org/10.1038/s41598-019-53418-5.
https://doi.org/10.3174/ajnr.A3044. [106] D.C. Hess, L.R. Wechsler, W.M. Clark, S.I. Savitz, G.A. Ford, D. Chiu, et al., Safety
[90] R. Sasaki, T. Yamashita, K. Tadokoro, N. Matsumoto, E. Nomura, Y. Omote, et al., and efficacy of multipotent adult progenitor cells in acute ischaemic stroke
Direct arterial damage and neurovascular unit disruption by mechanical (MASTERS): a randomised, double-blind, placebo-controlled, phase 2 trial, Lancet
thrombectomy in a rat stroke model, J. Neurosci. Res. (2020), https://doi.org/ Neurol. 16 (2017) 360–368, https://doi.org/10.1016/S1474-4422(17)30046-7.
10.1002/jnr.24671. Advance online publication. [107] S.I. Savitz, D. Yavagal, G. Rappard, W. Likosky, N. Rutledge, C. Graffagnino, et al.,
[91] M. Goyal, B.K. Menon, W.H. van Zwam, D.W. Dippel, P.J. Mitchell, A. A phase 2 randomized, sham-controlled Trial of internal carotid artery infusion of
M. Demchuk, et al., Endovascular thrombectomy after large-vessel ischaemic autologous bone marrow-derived ALD-401 cells in patients with recent stable
stroke: a meta-analysis of individual patient data from five randomised trials, ischemic stroke (RECOVER-stroke), Circulation. 139 (2019) 192–205, https://
Lancet 387 (2016) 1723–1731, https://doi.org/10.1016/S0140-6736(16)00163- doi.org/10.1161/CIRCULATIONAHA.117.030659.
X.

You might also like