Open access peer-reviewed chapter

Drug Repurposing in Neurological Diseases: Opportunities and Challenges

Written By

Xiao-Yuan Mao

Submitted: 14 January 2020 Reviewed: 01 June 2020 Published: 26 June 2020

DOI: 10.5772/intechopen.93093

From the Edited Volume

Drug Repurposing - Hypothesis, Molecular Aspects and Therapeutic Applications

Edited by Farid A. Badria

Chapter metrics overview

1,013 Chapter Downloads

View Full Metrics

Abstract

Drug repurposing or repositioning refers to “studying of clinically approved drugs in one disease to see if they have therapeutic value and do not trigger side effects in other diseases.” Nowadays, it is a vital drug discovery approach to explore new therapeutic benefits of existing drugs or drug candidates in various human diseases including neurological disorders. This approach overcomes the shortage faced during traditional drug development in grounds of financial support and timeline. It is especially hopeful in some refractory diseases including neurological diseases. The feature that structure complexity of the nervous system and influence of blood–brain barrier permeability often becomes more difficult to develop new drugs in neuropathological conditions than diseases in other organs; therefore, drug repurposing is particularly of utmost importance. In this chapter, we discuss the role of drug repurposing in neurological diseases and make a summarization of repurposing candidates currently in clinical trials for neurological diseases and potential mechanisms as well as preliminary results. Subsequently we also outline drug repurposing approaches and limitations and challenges in the future investigations.

Keywords

  • drug repurposing
  • brain injury
  • neurological diseases
  • therapeutics

1. Introduction

Neurological disorders are devastating diseases which usually occur in the brain, spinal cord, cranial nerves, peripheral nerves, and so on. It has reported that there are more than 600 kinds of neuropathological conditions including epilepsy, brain tumor, Parkinson’s disease, Alzheimer’s disease, and stroke. Nowadays, it is estimated that more than 1 billion people suffer from neurological disorders, seriously affecting people’s life quality [1]. These kinds of diseases are especially prevalent in developing countries at any stage of age [2, 3]. There are several factors contributing to etiology of neurological disorders such as aggravating tendency of aging population, irregular diet, and insufficient exercise [4].

Drug therapy is an important way for curing neurological diseases in the clinic. Nevertheless, serious neurological disorders such as Alzheimer’s disease (AD) and Parkinson’s disease (PD) are usually incurable in late stages of diseases with current therapeutic intervention [5, 6]. In the meantime, drug treatment often becomes less effective and causes serious side effects due to individual differences. Taking epilepsy as example, nearly 30% of epileptic patients are unable to obtain seizure control following treatment with marketed drugs [7, 8]. In addition, they have no significant effect on the improvement of cognitive dysfunction in patients with severe epilepsy [9]. Thus, it is essential for investigation of more effective and/or less toxic CNS targeted drugs.

Drug repurposing, also known as drug reprofiling or drug repositioning, includes the development of new uses and dosage forms for existing drugs or drug candidates. It is regarded as an economic and practical strategy [10]. Drug repurposing avoids the defects of new drug development. Compared to the drug repurposing, development of new drugs consumes much more time and huge investments. It is roughly reported that the cost from basic research for a new drug to clinical trials is 2.6 billion US dollars [11] and it often takes an average of 13–15 years [12]. Although more and more drug candidates are developed, many cases have failed in recent years [13]. Most of new drugs are withdrawn from the market due to unsatisfactory efficacy or intolerable side effects [14, 15]. Therefore, reusing existing drugs, namely, drug repurposing, has attracted great attention, as this approach has the capacity of saving cost and expediting drug development process.

The purpose of this chapter is to discuss the role of drug repurposing in human diseases especially neurological diseases and summarize repurposing candidates currently in clinical trials for neurological diseases and potential mechanisms as well as preliminary results. Subsequently we also list drug repurposing approaches and limitations and challenges in the future investigations.

Advertisement

2. Repurposed drugs in neurological diseases

Prior to development of repurposed drugs for neurological diseases therapeutics, it is emphasized how the drug reposition process is carried out. Generally, there are three stages in drug repurposing. First, diverse approaches including serendipitous clinical observation, cellular drug activity assays, in silico drug screens, and data mining of clinical drug interaction are employed to obtain drug candidates [16]. The detailed illustrations in grounds of methodologies are summarized as mentioned above [17]. Second, preclinical investigations including in vivo rodent models and in vitro cell lines for these drugs are conducted in neurological diseases [18]. Finally, large-scale and multicenter clinical trials are implemented for evaluating efficacy and safety of repurposed drugs [19]. Up to date, there are plenty of drugs which are repurposed in neurological diseases through the above approaches. Then, in the following section, we also cite several repurposed drugs to elaborate how they function in neurological diseases. Table 1 summarizes various repurposed drugs in the treatment of neurological disorders.

Name of drug Original indication Novel indication Target Summarization of evidence
Verapamil Hypertension
Angina pectoris
Arrhythmia
Intractable epilepsy
Subarachnoid hemorrhage
Stroke
Resistant depression
P-glycoprotein
  1. Improving life quality in drug-resistant epileptic patients

  2. Preventing behavior phenotype in a mouse model of focal ischemia

  3. Showing no adverse effect in patients with stroke

Bumetanide Liver disease
Heart failure
Stubborn edema
Acute and chronic renal failure
Epilepsy
Autism
NKCC1 protein
  1. Improving anticonvulsant effect of phenobarbital in hypoxic rats

  2. Decreasing neuronal discharge in vitro and in vivo

Minocycline Antibacterial Epilepsy
Spinal cord injury
Brain inflammation
Neurodegenerative diseases
Activated microglia
IL-6, TNF-α
TrkB/BDNF
PPAR-γ/NF-κB
LKB1/AMPK
  1. Reducing seizure duration in rats

  2. Inhibiting inflammatory cytokines and cell death in kainic acid-induced epilepsy models

Fenfluramine Simple obesity
Diabetes
Hypertension
Epilepsy
Parkinson’s disease
5-HT receptors
  1. Alleviating epilepsy in patients with Dravet syndrome

  2. Anticonvulsant effects on photosensitive or induced convulsions

Propranolol Hypertension
Supraventricular tachycardia
Prolonged Q-T interval
Thyrotoxicosis
Migraine
Traumatic brain injury
Parkinson’s disease
IL-6
β-adrenergic
  1. Alleviating headache in patients with angina pectoris

  2. Reducing mortality within 24 h of admission in patients with TBI

  3. Preventing neuronal necrosis in a pig model of TBI

Sunitinib Gastrointestinal stromal tumor
Non-small-cell lung cancer
Renal cell carcinoma
Glioma
Pheochromocytoma
Alzheimer’s disease’
Acetylcholinesterase
CGNs, SH-SY5Y
  1. Penetrating the blood–brain barrier in clinical studies

  2. Alleviating glioma progression and glioma-induced neurodegeneration in vivo

  3. Preventing neuronal death induced by neurotoxins in vivo

Angiotensin receptor blockers Essential hypertension
Renal disease
Diabetes
Alzheimer’s disease
Episodic migraine
AT1 receptor
Angiotensin II
  1. Reducing Aβ accumulation and aggregation in vivo

  2. Alleviating AD in epidemiological studies and RCTs

Amantadine Antiviral Parkinson’s disease
Chronic traumatic brain injury
N-methyl-D-aspartate (NMDA)
Anticholinergic
  1. Improving motor symptoms in a female PD patient

  2. Activating the dopamine system in several preclinical data demonstrate

Table 1.

List of repurposed drugs in neurological disease.

2.1 Verapamil

Verapamil, a classical calcium channel blocker, is mainly used in the treatment of hypertension, angina pectoris, arrhythmia, and other diseases, especially for paroxysmal supraventricular tachycardia [20]. It has been found that administration of verapamil greatly improves seizure control in drug-resistant epileptic patients via inhibiting P-glycoprotein (Pgp). Pgp is responsible for the transport of antiepileptic drug (AED) into the blood vessels through the blood–brain barrier (BBB). And there is evidence supporting that overexpression of Pgp in the brain represents a major mechanism underlying drug resistance in epileptic patients [21]. Verapamil is found to suppress Pgp expression and subsequently facilitates the entry of this drug into epileptogenic zones. As a marketed drug, verapamil treatment in patients with intractable epilepsy can doubtfully alleviate brain injury caused by repetitive seizures [22]. Actually, in clinical trials, verapamil has previously shown to exhibit great efficacy in intractable depression or mania via inhibiting the function of Pgp [23, 24]. Moreover, it is documented that verapamil has been approved to treat cerebral vasospasm secondary to subarachnoid hemorrhage due to its vasodilatory effects [25]. Intra-arterial (IA) treatment with verapamil, which was physiologically feasible, safe, and neuroprotective as a therapeutic adjunct in stroke, significantly reduces infarct volume and improved functional outcome [26], although there are still some mysteries about the mechanism.

2.2 Bumetanide

As a potent diuretic agent, bumetanide, which is mainly employed to cure liver disease, heart failure, and various kinds of stubborn edema in clinic [27], is a specific inhibitor of Na+-K+-2Cl cotransporter isoform 1 (NKCC1) [28]. Mechanically, NKCC1 significantly modulates the content of intracellular Cl. Upregulation of NKCC1 leads to elevation of intracellular concentration of Cl, which is associated with pathogenesis of neurological diseases. It has been unequivocally proven that many of the available drugs have anti-seizure potential via activating GABAA-mediated hyperpolarization due to accumulation of neuronal Cl [29]. Indeed, current investigations have confirmed that bumetanide exerts antiepileptic effect via switching the GABA-mediated inhibitory postsynaptic potential in neurons from depolarization to hyperpolarization, resulting in decreased neuronal discharge [30, 31]. In addition, previous work reinforces that bumetanide can enhance the anticonvulsant effect of phenobarbital in hypoxic rats [32]. It suggests that the combination of phenobarbital and bumetanide may provide a promising therapeutic strategy for ceasing seizures in neonatal epilepsy and may increase the neuroprotective effect of hypothermia on asphyxiated newborns [33]. Persuasively, a current clinically pilot study further demonstrated that bumetanide, as a specific NKCC1 antagonist, considerably reduced seizure frequency in adult patients with temporal lobe epilepsy [34]. Additionally, as a consequence of a randomized controlled trial, bumetanide may also be effective for treatment of autism [35]. It should be considered that there are two obstacles for bumetanide treatment in neurological disorders [31, 36]. It has been shown that the highly potent diuretic effect of bumetanide can lead to hypokalemic alkalosis and the poor penetration into brain exists. This indicates that reuse of bumetanide in neurological diseases brings about opportunities and challenges in the future.

2.3 Minocycline

Minocycline is the second generation of semisynthetic broad-spectrum antibacterial tetracycline analogues. It has immunomodulatory, anti-inflammatory, and anti-apoptosis effects. Minocycline has neuroprotective effects in rodent models of ischemia, spinal cord injury, and infection [37]. It can efficiently penetrate the BBB and has a good effect on activated microglia, which indicates a possible role in the treatment of epilepsy. Minocycline may have synergistic effects with other compounds in manipulating epilepsy. Minocycline has been found to remarkably obviate epileptic conditions and reduce seizure-induced brain impairment at early stage [38]. In addition, minocycline also inhibits pro-inflammatory cytokines through caspase-dependent and caspase-independent pathways, thus inhibiting cell death in kainic acid-induced status epilepticus [39]. An obvious improvement of seizure phenotype is also observed in a rat model of amygdala kindling [40]. Additionally, increasing studies have reported the neuroprotective effects of minocycline in neurologic diseases, such as ischemic stroke, multiple sclerosis (MS), and traumatic brain injury (TBI) [41, 42, 43]. In in vivo animal model, minocycline promotes M2 microglia polarization via activation of tyrosine kinase receptor B (TrkB)/brain-derived neurotrophic factors (BDNF) pathway and facilitates neurogenesis after intracerebral hemorrhage (ICH) [44]. In the process of acute cerebral infarct, minocycline also effectively inhibits oxidative stress via elevating the activity of superoxide dismutase (SOD) and activating the liver kinase B1 (LKB1)/adenosine 5′-monophosphate (AMP)-activated protein kinase (AMPK) signaling pathway [45]. However, repurposing of minocycline in treating neurological diseases requires to be re-evaluated as there is a clinical study showing serious neurodegeneration TBI [46].

2.4 Fenfluramine

Fenfluramine, which has been successfully applied in obesity, diabetes, and hypertension [47], is a potent 5-hydroxytryptamine (5-HT) releaser activating multiple 5-HT receptor subtypes. Of note, elevation of extracellular 5-HT levels inhibits focal and generalized seizures, while depletion of 5-HT lowers the threshold of epileptic seizures [48]. Therefore, 5-HT agonist fenfluramine is assessed for treatment of epilepsy. In a small-scale retrospective study, it has reported that adjuvant treatment with fenfluramine has evidently obtained seizure control in patients with Dravet syndrome. As the side effects is not serious, it does not lead to the termination of treatment [49]. This drug may have anticonvulsant effects on other severe epilepsy syndromes, especially those characterized by photosensitive or induced convulsions [50, 51]. Encouragingly, a recent investigation has unveiled that fenfluramine significantly reduces convulsive seizure frequency compared with placebo and exhibits good tolerance [52]. It indicates that fenfluramine could be functioned as a potent novel therapeutic regime for patients with Dravet syndrome. It is noteworthy that fenfluramine also alleviates L-DOPA-induced dyskinesia via stimulation of 5-HT1A receptor in PD [53].

2.5 Propranolol

Propranolol as a β-adrenoceptor antagonist (b-blocker) has been commonly used in hypertension, supraventricular tachycardia, prolonged Q-T interval, and thyrotoxicosis in clinic [54]. Since 1996, in patients who were being treated for angina pectoris, Rabkin et al. has disclosed the therapeutic effect of propranolol on migraine headache [55]. Meanwhile, further clinical studies have noted that administration of propranolol within 24 h of admission after TBI triggers lower mortality [56]. The evidence also arises from a recent study that propranolol blocks the upregulation of IL-6 and prevents neuronal cell necrosis in CA1 and CA3 hippocampus in a pig model of TBI [57]. Given that propranolol has neuroprotective potential in neuropathological conditions, it is likely to serve as a neuroprotective drug in epilepsy. Additionally, both clinical and experimental studies have demonstrated the potential of propranolol to resist dyskinesia in PD, as modulation of β-adrenergic receptors (βAR), which is abundantly, expressed in striatum, is involved L-DOPA-induced dyskinesia (LID) [58, 59].

2.6 Sunitinib

Sunitinib, which is an oral, small molecule receptor tyrosine kinase inhibitor approved by the US Food and Drug Administration, has been currently implemented in the treatment of various cancers such as gastrointestinal stromal tumor (GIST), non-small-cell lung cancer, and renal cell carcinoma [60]. Clinical evidence has revealed that oral administration of sunitinib penetrates the BBB and subsequently facilitates the entry into central nervous system [61]. Furthermore, on the basis of its potent antiangiogenic and antitumoral characteristics, it has discovered that sunitinib can alleviate glioma-induced neurodegeneration and glioma progression in vivo models [60]. Meanwhile, sunitinib has been found to exert therapeutic effects on learning and memory deficits in a mouse model of AD through inhibition of acetylcholinesterase (AChE) [62]. Additionally, sunitinib has also demonstrated to prevent neuronal death induced by neurotoxins via inhibiting NO overproduction in cerebellar granule neurons (CGNs) and SH-SY5Y cells following exposure with low potassium or 1-methyl-4-phenylpyridinium ion (MPP+)-induced neuronal apoptosis [63]. It indicates that sunitinib may improve brain dysfunction via inhibition of oxidative stress.

2.7 Angiotensin receptor blockers

In in vitro studies, angiotensin receptor blockers (ARBs) are generally known to treat essential hypertension by influencing the level of angiotensin II (Ang II) via two distinct pathways, namely, through interrupting the AT1 receptor and augmentation of Ang II processing which plays a critical role in cognition regulation [64]. For example, valsartan, which has previously been found to penetrate BBB and elicit antihypertensive responses in the brain, has been demonstrated to reduce Aβ accumulation and aggregation in vivo and in vitro [65]. Actually, similar situation exists in losartan and telmisartan, which are also classical ARBs [66, 67]. Overall, it indicates ARBs are potential candidates for treating AD. Significantly, several clinically epidemiological studies and RCTs certify the efficacy of ARBs in AD. A large-scale retrospective cohort study has revealed an obvious reduction of dementia in patients treated with ARBs compared with other cardiovascular agents [68]. Likewise, the further UK-based study also reports a similar trend, with a 50% reduction in AD after ARBs treatment [69]. In brief, ARBs, the conventional cardiovascular medicine, have been confirmed to exert a vital effect in AD, and it is further deserved to identify the most suitable dosage in clinic.

2.8 Amantadine

Amantadine is a classic antiviral compound which has been found to moderately ameliorate impaired motor behavior in Parkinson’s disease [70]. Intriguingly, in 1969, it was coincident that Schwab et al. found an improvement of motor symptoms in a female PD patient, who took 200 mg amantadine daily for antiviral prophylaxis [71]. Subsequently, three potential mechanisms have been proposed to explain the efficacy of amantadine in PD. Several preclinical data demonstrate an activation of the dopamine system’s both presynaptic and postsynaptic actions [72], and amantadine also inhibits the N-methyl-D-aspartate (NMDA) subtype of glutamate receptors [72, 73]. The mild anticholinergic effect is also involved [74]. Surprisingly, PD is well known to be frequently associated with depression, and antagonism of NMDA receptors is also a promising target for new antidepressants, although there is no definite evidence to certify its efficacy in depressive disorder.

Advertisement

3. Approaches to drug repurposing

There are three important stages in the field of drug repurposing: generation of candidate compounds, preclinical investigation, and clinical trial. Determination of appropriate drugs for potential indications is crucial for production of candidate compounds. At present, two approaches are widely used for drug repurposing including experimental screening approaches and molecular docking by computer. In the following items, we make a detailed description of these two methods in drug repurposing process.

3.1 Experimental approaches

Experimental screening approaches are usually regarded as the first stage in the process of drug discovery and drug repurposing. Proteomic techniques such as affinity chromatography and mass spectrometry have been widely employed to identify drug candidates [75]. Nowadays, drug target analysis and drug repurposing are inseparable. Drug repurposing is distinct from drug discovery in terms of alteration of drug target. Cellular thermo stability assay technique can predict the affinity of drug ligands by mapping the contact patterns of intracellular targets [76]. The molecular on and off targets have been disclosed for many clinically approved drugs via this method. Especially in the field of kinases, new targets of well-known drugs are obtained through affinity matrices [77, 78]. For example, imatinib, a tyrosine kinase inhibitor, has been successfully reused in the treatment of gastrointestinal stromal tumors [79].

In addition, chemical compounds with disease-related effects can be defined in the model through phenotype screening [80]. Phenotype screening has always been more successful than target screening in the facet of drug development [81, 82]. In the case of drug repurposing, if the compounds selected through phenotypical assays are approved clinical drugs or ongoing clinical trials, they are probable to reuse. Several drugs approved for tobacco dependence have been evaluated, and it has been found that topiramate changes nicotine- or ethanol-induced behavior in zebrafish models [83]. However, there are some challenges that the efficacy of drug candidates in in vitro experiments require to be validated in human diseases [84].

3.2 Computational approaches

Molecular docking by a computer is also an important method for evaluating drug target binding kinetics and drug residence times of existing drugs or drug candidates [85]. Large amounts of computational drug repositioning methods choose transcriptomic data to identify potential new indications for drugs. Furthermore, these methods have applied techniques such as comparison of gene expression profiles between a disease model and drug-treated condition [86], network integration [87], prediction of drug-protein interactions [88], and utilization of genotype–phenotype associations. Recently, a proteotranscriptomic-based computational drug repositioning method named Drug Repositioning Perturbation Score/Class (DRPS/C) for Alzheimer’s disease occurs on the basis of inverse associations between disease-induced or drug-induced gene and protein perturbation patterns [89]. Briefly, these approaches can be applicable to discovery of drug targets or biomarkers.

It should be considered that for many neurological disorders, drugs require good penetration into BBB. Then, the therapeutic approaches of targeting brain have been classified as invasive and noninvasive categories [90, 91]. The invasive approaches contain the temporary increase of BBB permeability, and noninvasive approaches involve modification of drug molecule via physiological, chemical, or colloidal carrier system approach. Meanwhile, these methods are also related to computational approaches. Influx clearance into the brain (K in), which is the unidirectional influx constant from the blood to brain, can be used to calculate the transport of drugs in the brain. Similar computational approaches conclude the permeability surface area (PS), brain/plasma ratio (K p), brain uptake index (BUI), and apparent permeability (P app) [92, 93, 94, 95]. Consequently, drug repurposing in neurological diseases covers various manners to participate in integrating the role of transporters and pathophysiological complexity of BBB to establish a suitable model for high-throughput screening.

Advertisement

4. Concluding remarks and perspectives

Drug repurposing is a vital strategy for developing new therapeutic values of existing drugs or drug candidates due to its ability to save time and reduce cost [96]. This type of innovative concept will undoubtedly expedite the drug development process. Meanwhile, some limitations need to be considered during drug repurposing process in neurological diseases. Owing to complex molecular and cellular signaling mechanisms in neuropathological states, drug repurposing may be difficult. Additionally, drugs not only respond to a single target but also affect multiple targets [97], causing a variety of adverse reactions. A comprehensive assessment of the advantages and disadvantages of these side effects can help us understand drug repositioning from a more all-round perspective [98, 99].

In order to overcome limitations faced during drug repurposing, we make proposals in the following descriptions. Firstly, it is foremost to establish a comprehensive data analysis platform to maximize data sharing. Information science services and artificial intelligence can help unlock and reanalyze the large amount of data accumulated by approved drugs or drug candidates to clinical trials. These data may be stored in a diversified way. Storage locations, formats, and types may vary, including different storage locations, formats, and types. The data obtained from clinical trials and biological databases are too large and complex that the traditional data processing methods cannot deal with it, which leads to the bottleneck in the research process [99]. Big data can significantly improve our understanding of the disease and make more accurate disease-related strategies. However, there is a big gap between generating biomedical data and data analysis [99, 100]. To ensure the efficiency of research, it takes time, energy, and expertise to find technical solutions to integrate them. Secondly, it is encouraged to provide more financial support for clinical trials of drug repurposing, including technical support. The preclinical research of drug repurposing requires financial support to obtain the data in clinical trials. In this case, drugs that can be developed to treat rare diseases are more likely to apply in clinical neurological diseases therapeutics [101]. Finally, in order to facilitate drug repurposing process, we advocate it is indispensable to solve patent restrictions and take reasonable supervision. All applications of drug repurposing should be accompanied by a risk management plan. Drug’s safety can be supported by clinical trial data or post marketing data.

In conclusion, drug repurposing is a novel approach for expediting drug development process in neurological diseases. Repurposed drugs may provide an efficient avenue for improving a plethora of pathological conditions including neurological disorders. In the future, it is essential to exploit molecular mechanisms during drug repurposing processes due to the possibility that targets of repurposed drugs in neurological diseases are distinct from original targets in treating other diseases, in order to make these drugs more effective and safe.

Advertisement

Acknowledgments

The authors apologize to all the investigators whose work cannot be cited in this paper due to space constraint. This work was partly supported by the National Natural Science Foundation of China (No. 81974502 and 81671293).

Advertisement

Conflict of interest

There is no potential conflict of interest.

Advertisement

Abbreviations

CNScentral nervous system
ADAlzheimer’s disease
PDParkinson’s disease
AEDantiepileptic drug
BBBblood–brain barrier
PgpP-glycoprotein
NKCC1Na+-K+-2Cl-cotransporter isoform 1
GABAAgamma-aminobutyric acid
MSmultiple sclerosis
TBItraumatic brain injury
TrkBtyrosine kinase receptor B
BDNFbrain-derived neurotrophic factors
ICHintracerebral hemorrhage
SODsuperoxide dismutase
LKB1liver kinase B1
AMPKadenosine 5′-monophosphate (AMP)-activated protein kinase
5-HT5-hydroxytryptamine
LIDL-DOPA-induced dyskinesia
βARβ-adrenergic receptors
AChEacetylcholinesterase
CGNscerebellar granule neurons
ARBsangiotensin receptor blockers
Ang IIangiotensin II
NMDAN-methyl-D-aspartate

References

  1. 1. Cottler LB, Zunt J, Weiss B, Kamal AK, Vaddiparti K. Building global capacity for brain and nervous system disorders research. Nature. 2015;527:S207-S213. DOI: 10.1038/nature16037
  2. 2. Chin JH, Vora N. The global burden of neurologic diseases. Neurology. 2014;83:349-351. DOI: 10.1212/wnl.0000000000000610
  3. 3. Whiteford HA et al. Global burden of disease attributable to mental and substance use disorders: Findings from the global burden of disease study 2010. Lancet. 2013;382:1575-1586. DOI: 10.1016/s0140-6736(13)61611-6
  4. 4. Nugent RA, Yach D, Feigl AB. Non-communicable diseases and the Paris declaration. Lancet. 2009;374:784-785. DOI: 10.1016/s0140-6736(09)61589-0
  5. 5. Lane CA, Hardy J, Schott JM. Alzheimer’s disease. European Journal of Neurology. 2018;25:59-70. DOI: 10.1111/ene.13439
  6. 6. Radder DLM et al. Physical therapy and occupational therapy in Parkinson’s disease. The International Journal of Neuroscience. 2017;127:930-943. DOI: 10.1080/00207454.2016.1275617
  7. 7. Shorvon SD. The epidemiology and treatment of chronic and refractory epilepsy. Epilepsia. 1996;37(Suppl 2):S1-S3. DOI: 10.1111/j.1528-1157.1996.tb06027.x
  8. 8. Löscher W, Brandt C. Prevention or modification of epileptogenesis after brain insults: Experimental approaches and translational research. Pharmacological Reviews. 2010;62:668-700. DOI: 10.1124/pr.110.003046
  9. 9. Holmes GL, Noebels JL. The epilepsy spectrum: Targeting future research challenges. Cold Spring Harbor Perspectives in Medicine. 2016;6:1-12. DOI: 10.1101/cshperspect.a028043
  10. 10. Hemphill CS, Sampat BN. Evergreening, patent challenges, and effective market life in pharmaceuticals. Journal of Health Economics. 2012;31:327-339. DOI: 10.1016/j.jhealeco.2012.01.004
  11. 11. DiMasi JA, Grabowski HG, Hansen RW. Innovation in the pharmaceutical industry: New estimates of R&D costs. Journal of Health Economics. 2016;47:20-33. DOI: 10.1016/j.jhealeco.2016.01.012
  12. 12. Strittmatter SM. Overcoming drug development bottlenecks with repurposing: Old drugs learn new tricks. Nature Medicine. 2014;20:590-591. DOI: 10.1038/nm.3595
  13. 13. Scannell JW, Blanckley A, Boldon H, Warrington B. Diagnosing the decline in pharmaceutical R&D efficiency. Nature Reviews. Drug Discovery. 2012;11:191-200. DOI: 10.1038/nrd3681
  14. 14. DiMasi JA, Feldman L, Seckler A, Wilson A. Trends in risks associated with new drug development: Success rates for investigational drugs. Clinical Pharmacology and Therapeutics. 2010;87:272-277. DOI: 10.1038/clpt.2009.295
  15. 15. Arrowsmith J, Miller P. Trial watch: Phase II and phase III attrition rates 2011-2012. Nature Reviews. Drug Discovery;12:569, 2013. DOI: 10.1038/nrd4090
  16. 16. O’Connor KA, Roth BL. Finding new tricks for old drugs: An efficient route for public-sector drug discovery. Nature Reviews. Drug Discovery. 2005;4:1005-1014. DOI: 10.1038/nrd1900
  17. 17. Kumar R et al. Exploring the new horizons of drug repurposing: A vital tool for turning hard work into smart work. European Journal of Medicinal Chemistry. 2019;182:111602. DOI: 10.1016/j.ejmech.2019.111602
  18. 18. Turanli B et al. Drug repositioning for effective prostate cancer treatment. Frontiers in Physiology. 2018;9:500. DOI: 10.3389/fphys.2018.00500
  19. 19. Lago SG, Bahn S. Clinical trials and therapeutic rationale for drug repurposing in schizophrenia. ACS Chemical Neuroscience. 2019;10:58-78. DOI: 10.1021/acschemneuro.8b00205
  20. 20. Delaney B, Loy J, Kelly AM. The relative efficacy of adenosine versus verapamil for the treatment of stable paroxysmal supraventricular tachycardia in adults: A meta-analysis. European Journal of Emergency Medicine: Official Journal of the European Society for Emergency Medicine. 2011;18:148-152. DOI: 10.1097/MEJ.0b013e3283400ba2
  21. 21. Robey RW, Lazarowski A, Bates SE. P-glycoprotein—A clinical target in drug-refractory epilepsy? Molecular Pharmacology. 2008;73:1343-1346. DOI: 10.1124/mol.108.046680
  22. 22. Summers MA, Moore JL, McAuley JW. Use of verapamil as a potential P-glycoprotein inhibitor in a patient with refractory epilepsy. The Annals of Pharmacotherapy. 2004;38:1631-1634. DOI: 10.1345/aph.1E068
  23. 23. de Klerk OL et al. Locally increased P-glycoprotein function in major depression: A PET study with [11C]verapamil as a probe for P-glycoprotein function in the blood-brain barrier. The International Journal of Neuropsychopharmacology. 2009;12:895-904. DOI: 10.1017/s1461145709009894
  24. 24. Barton BM, Gitlin MJ. Verapamil in treatment-resistant mania: An open trial. Journal of Clinical Psychopharmacology. 1987;7:101-103
  25. 25. Keuskamp J, Murali R, Chao KH. High-dose intraarterial verapamil in the treatment of cerebral vasospasm after aneurysmal subarachnoid hemorrhage. Journal of Neurosurgery. 2008;108:458-463. DOI: 10.3171/jns/2008/108/3/0458
  26. 26. Fraser JF et al. Intra-arterial verapamil post-thrombectomy is feasible, safe, and neuroprotective in stroke. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 2017;37:3531-3543. DOI: 10.1177/0271678x17705259
  27. 27. Xiong L et al. Evaluation of severe myalgia induced by continuous-infusion bumetanide in patients with acute heart failure. Pharmacotherapy. 2019;39:854-860. DOI: 10.1002/phar.2297
  28. 28. Kaila K, Price TJ, Payne JA, Puskarjov M, Voipio J. Cation-chloride cotransporters in neuronal development, plasticity and disease. Nature Reviews. Neuroscience. 2014;15:637-654. DOI: 10.1038/nrn3819
  29. 29. Hochman DW. The extracellular space and epileptic activity in the adult brain: Explaining the antiepileptic effects of furosemide and bumetanide. Epilepsia. 2012;53(Suppl 1):18-25. DOI: 10.1111/j.1528-1167.2012.03471.x
  30. 30. Rheims S et al. Excitatory GABA in rodent developing neocortex in vitro. Journal of Neurophysiology. 2008;100:609-619. DOI: 10.1152/jn.90402.2008
  31. 31. Löscher W, Puskarjov M, Kaila K. Cation-chloride cotransporters NKCC1 and KCC2 as potential targets for novel antiepileptic and antiepileptogenic treatments. Neuropharmacology. 2013;69:62-74. DOI: 10.1016/j.neuropharm.2012.05.045
  32. 32. Dzhala VI, Brumback AC, Staley KJ. Bumetanide enhances phenobarbital efficacy in a neonatal seizure model. Annals of Neurology. 2008;63:222-235. DOI: 10.1002/ana.21229
  33. 33. Liu Y, Shangguan Y, Barks JD, Silverstein FS. Bumetanide augments the neuroprotective efficacy of phenobarbital plus hypothermia in a neonatal hypoxia-ischemia model. Pediatric Research. 2012;71:559-565. DOI: 10.1038/pr.2012.7
  34. 34. Eftekhari S et al. Bumetanide reduces seizure frequency in patients with temporal lobe epilepsy. Epilepsia. 2013;54:e9-e12. DOI: 10.1111/j.1528-1167.2012.03654.x
  35. 35. Lemonnier E et al. A randomised controlled trial of bumetanide in the treatment of autism in children. Translational Psychiatry. 2012;2:e202. DOI: 10.1038/tp.2012.124
  36. 36. Töllner K et al. A novel prodrug-based strategy to increase effects of bumetanide in epilepsy. Annals of Neurology. 2014;75:550-562. DOI: 10.1002/ana.24124
  37. 37. Rosenblat JD, McIntyre RS. Efficacy and tolerability of minocycline for depression: A systematic review and meta-analysis of clinical trials. Journal of Affective Disorders. 2018;227:219-225. DOI: 10.1016/j.jad.2017.10.042
  38. 38. Abraham J, Fox PD, Condello C, Bartolini A, Koh S. Minocycline attenuates microglia activation and blocks the long-term epileptogenic effects of early-life seizures. Neurobiology of Disease. 2012;46:425-430. DOI: 10.1016/j.nbd.2012.02.006
  39. 39. Heo K et al. Minocycline inhibits caspase-dependent and -independent cell death pathways and is neuroprotective against hippocampal damage after treatment with kainic acid in mice. Neuroscience Letters. 2006;398:195-200. DOI: 10.1016/j.neulet.2006.01.027
  40. 40. Beheshti Nasr SM, Moghimi A, Mohammad-Zadeh M, Shamsizadeh A, Noorbakhsh SM. The effect of minocycline on seizures induced by amygdala kindling in rats. Seizure. 2013;22:670-674. DOI: 10.1016/j.seizure.2013.05.005
  41. 41. Kumar A et al. NOX2 drives M1-like microglial/macrophage activation and neurodegeneration following experimental traumatic brain injury. Brain, Behavior, and Immunity. 2016;58:291-309. DOI: 10.1016/j.bbi.2016.07.158
  42. 42. Perego C et al. Macrophages are essential for maintaining a M2 protective response early after ischemic brain injury. Neurobiology of Disease. 2016;96:284-293. DOI: 10.1016/j.nbd.2016.09.017
  43. 43. Wan S et al. Microglia activation and polarization after intracerebral hemorrhage in mice: The role of protease-activated receptor-1. Translational Stroke Research. 2016;7:478-487. DOI: 10.1007/s12975-016-0472-8
  44. 44. Miao H, Li R, Han C, Lu X, Zhang H. Minocycline promotes posthemorrhagic neurogenesis via M2 microglia polarization via upregulation of the TrkB/BDNF pathway in rats. Journal of Neurophysiology. 2018;120:1307-1317. DOI: 10.1152/jn.00234.2018
  45. 45. Cai Z, Wang C, Chen Y, He W. An antioxidant role by minocycline via enhancing the activation of LKB1/AMPK signaling in the process of cerebral ischemia injury. Current Molecular Medicine. 2018;18:142-151. DOI: 10.2174/1566524018666180907161504
  46. 46. Scott G et al. Minocycline reduces chronic microglial activation after brain trauma but increases neurodegeneration. Brain: A Journal of Neurology. 2018;141:459-471. DOI: 10.1093/brain/awx339
  47. 47. Aman MG, Kern RA. Review of fenfluramine in the treatment of the developmental disabilities. Journal of the American Academy of Child and Adolescent Psychiatry. 1989;28:549-565. DOI: 10.1097/00004583-198907000-00014
  48. 48. Bagdy G, Kecskemeti V, Riba P, Jakus R. Serotonin and epilepsy. Journal of Neurochemistry. 2007;100:857-873. DOI: 10.1111/j.1471-4159.2006.04277.x
  49. 49. Ceulemans B et al. Successful use of fenfluramine as an add-on treatment for Dravet syndrome. Epilepsia. 2012;53:1131-1139. DOI: 10.1111/j.1528-1167.2012.03495.x
  50. 50. Boel M, Casaer P. Add-on therapy of fenfluramine in intractable self-induced epilepsy. Neuropediatrics. 1996;27:171-173. DOI: 10.1055/s-2007-973781
  51. 51. Pierce JG, Mithal DS. Fenfluramine: New treatment for seizures in Dravet syndrome. Pediatric Neurology Briefs. 2020;34:8. DOI: 10.15844/pedneurbriefs-34-8
  52. 52. Lagae L et al. Fenfluramine hydrochloride for the treatment of seizures in Dravet syndrome: A randomised, double-blind, placebo-controlled trial. Lancet. 2020;394:2243-2254. DOI: 10.1016/s0140-6736(19)32500-0
  53. 53. Bishop C et al. MDMA and fenfluramine reduce L-DOPA-induced dyskinesia via indirect 5-HT1A receptor stimulation. The European Journal of Neuroscience. 2006;23:2669-2676. DOI: 10.1111/j.1460-9568.2006.04790.x
  54. 54. Bidabadi E, Mashouf M. A randomized trial of propranolol versus sodium valproate for the prophylaxis of migraine in pediatric patients. Paediatric Drugs. 2010;12:269-275. DOI: 10.2165/11316270-000000000-00000
  55. 55. Rabkin R, Stables DP, Levin NW, Suzman MM. The prophylactic value of propranolol in angina pectoris. The American Journal of Cardiology. 1966;18:370-383. DOI: 10.1016/0002-9149(66)90056-7
  56. 56. Ko A et al. Early propranolol after traumatic brain injury is associated with lower mortality. The Journal of Trauma and Acute Care Surgery. 2016;80:637-642. DOI: 10.1097/ta.0000000000000959
  57. 57. Armstead WM, Vavilala MS. Propranolol protects cerebral autoregulation and reduces hippocampal neuronal cell death through inhibition of interleukin-6 upregulation after traumatic brain injury in pigs. British Journal of Anaesthesia. 2019;123:610-617. DOI: 10.1016/j.bja.2019.07.017
  58. 58. Barnum CJ et al. Effects of noradrenergic denervation on L-DOPA-induced dyskinesia and its treatment by α- and β-adrenergic receptor antagonists in hemiparkinsonian rats. Pharmacology, Biochemistry, and Behavior. 2012;100:607-615. DOI: 10.1016/j.pbb.2011.09.009
  59. 59. Waeber C, Rigo M, Chinaglia G, Probst A, Palacios JM. Beta-adrenergic receptor subtypes in the basal ganglia of patients with Huntington’s chorea and Parkinson’s disease. Synapse. 1991;8:270-280. DOI: 10.1002/syn.890080405
  60. 60. Hatipoglu G et al. Sunitinib impedes brain tumor progression and reduces tumor-induced neurodegeneration in the microenvironment. Cancer Science. 2015;106:160-170. DOI: 10.1111/cas.12580
  61. 61. Addeo R, Caraglia M. The oral tyrosine kinase inhibitors lapatinib and sunitinib: New opportunities for the treatment of brain metastases from breast cancer? Expert Review of Anticancer Therapy. 2011;11:139-142. DOI: 10.1586/era.10.190
  62. 62. Huang L et al. Sunitinib, a clinically used anticancer drug. Is a potent AChE inhibitor and attenuates cognitive impairments in mice. ACS Chemical Neuroscience. 2016;7:1047-1056. DOI: 10.1021/acschemneuro.5b00329
  63. 63. Cui W et al. Sunitinib produces neuroprotective effect via inhibiting nitric oxide overproduction. CNS Neuroscience & Therapeutics. 2014;20:244-252. DOI: 10.1111/cns.12203
  64. 64. Wright JW, Harding JW. Brain renin-angiotensin—a new look at an old system. Progress in Neurobiology. 2011;95:49-67. DOI: 10.1016/j.pneurobio.2011.07.001
  65. 65. Wang J et al. Valsartan lowers brain beta-amyloid protein levels and improves spatial learning in a mouse model of Alzheimer disease. The Journal of Clinical Investigation. 2007;117:3393-3402. DOI: 10.1172/jci31547
  66. 66. Mogi M et al. Telmisartan prevented cognitive decline partly due to PPAR-gamma activation. Biochemical and Biophysical Research Communications. 2008;375:446-449. DOI: 10.1016/j.bbrc.2008.08.032
  67. 67. Danielyan L et al. Protective effects of intranasal losartan in the APP/PS1 transgenic mouse model of Alzheimer disease. Rejuvenation Research. 2010;13:195-201. DOI: 10.1089/rej.2009.0944
  68. 68. Li NC et al. Use of angiotensin receptor blockers and risk of dementia in a predominantly male population: Prospective cohort analysis. British Medical Journal (Clinical Research Edition). 2010;340:b5465. DOI: 10.1136/bmj.b5465
  69. 69. Davies NM, Kehoe PG, Ben-Shlomo Y, Martin RM. Associations of anti-hypertensive treatments with Alzheimer’s disease, vascular dementia, and other dementias. Journal of Alzheimer’s Disease. 2011;26:699-708. DOI: 10.3233/jad-2011-110347
  70. 70. Müller T, Kuhn W, Möhr JD. Evaluating ADS5102 (amantadine) for the treatment of Parkinson’s disease patients with dyskinesia. Expert Opinion on Pharmacotherapy. 2019;20:1181-1187. DOI: 10.1080/14656566.2019.1612365
  71. 71. Schwab RS, England AC Jr, Poskanzer DC, Young RR. Amantadine in the treatment of Parkinson’s disease. JAMA. 1969;208:1168-1170
  72. 72. Bailey EV, Stone TW. The mechanism of action of amantadine in parkinsonism: A review. Archives Internationales de Pharmacodynamie et de Thérapie. 1975;216:246-262
  73. 73. Chase TN, Bibbiani F, Oh JD. Striatal glutamatergic mechanisms and extrapyramidal movement disorders. Neurotoxicity Research. 2003;5:139-146. DOI: 10.1007/bf03033378
  74. 74. Nastuk WL, Su P, Doubilet P. Anticholinergic and membrane activities of amantadine in neuromuscular transmission. Nature. 1976;264:76-79. DOI: 10.1038/264076a0
  75. 75. Brehmer D et al. Cellular targets of gefitinib. Cancer Research. 2005;65:379-382
  76. 76. Martinez Molina D et al. Monitoring drug target engagement in cells and tissues using the cellular thermal shift assay. Science (New York, N.Y.). 2013;341:84-87. DOI: 10.1126/science.1233606
  77. 77. Klaeger S et al. Chemical proteomics reveals ferrochelatase as a common off-target of kinase inhibitors. ACS Chemical Biology. 2016;11:1245-1254. DOI: 10.1021/acschembio.5b01063
  78. 78. Troutman S et al. Crizotinib inhibits NF2-associated schwannoma through inhibition of focal adhesion kinase 1. Oncotarget. 2016;7:54515-54525. DOI: 10.18632/oncotarget.10248
  79. 79. Blanke CD et al. Long-term results from a randomized phase II trial of standard- versus higher-dose imatinib mesylate for patients with unresectable or metastatic gastrointestinal stromal tumors expressing KIT. Journal of Clinical Oncology : Official Journal of the American Society of Clinical Oncology. 2008;26:620-625. DOI: 10.1200/jco.2007.13.4403
  80. 80. Moffat JG, Vincent F, Lee JA, Eder J, Prunotto M. Opportunities and challenges in phenotypic drug discovery: An industry perspective. Nature Reviews. Drug Discovery. 2017;16:531-543. DOI: 10.1038/nrd.2017.111
  81. 81. Swinney DC, Anthony J. How were new medicines discovered? Nature Reviews. Drug Discovery. 2011;10:507-519. DOI: 10.1038/nrd3480
  82. 82. Eder J, Sedrani R, Wiesmann C. The discovery of first-in-class drugs: Origins and evolution. Nature Reviews. Drug Discovery. 2014;13:577-587. DOI: 10.1038/nrd4336
  83. 83. Cousin MA et al. Larval zebrafish model for FDA-approved drug repositioning for tobacco dependence treatment. PLoS One. 2014;9:e90467. DOI: 10.1371/journal.pone.0090467
  84. 84. Horvath P et al. Screening out irrelevant cell-based models of disease. Nature Reviews. Drug Discovery. 2016;15:751-769. DOI: 10.1038/nrd.2016.175
  85. 85. De Benedetti PG, Fanelli F. Computational modeling approaches to quantitative structure-binding kinetics relationships in drug discovery. Drug Discovery Today. 2018;23:1396-1406. DOI: 10.1016/j.drudis.2018.03.010
  86. 86. Chen B et al. Reversal of cancer gene expression correlates with drug efficacy and reveals therapeutic targets. Nature Communications. 2017;8:16022. DOI: 10.1038/ncomms16022
  87. 87. Luo Y et al. A network integration approach for drug-target interaction prediction and computational drug repositioning from heterogeneous information. Nature Communications. 2017;8:573. DOI: 10.1038/s41467-017-00680-8
  88. 88. Yang L, Agarwal P. Systematic drug repositioning based on clinical side-effects. PLoS One. 2011;6:e28025. DOI: 10.1371/journal.pone.0028025
  89. 89. Lee SY et al. A proteotranscriptomic-based computational drug-repositioning method for Alzheimer’s disease. Frontiers in Pharmacology. 2019;10:1653. DOI: 10.3389/fphar.2019.01653
  90. 90. Gabathuler R. Approaches to transport therapeutic drugs across the blood-brain barrier to treat brain diseases. Neurobiology of Disease. 2010;37:48-57. DOI: 10.1016/j.nbd.2009.07.028
  91. 91. Alam MI et al. Strategy for effective brain drug delivery. European Journal of Pharmaceutical Sciences: Official Journal of the European Federation for Pharmaceutical Sciences. 2010;40:385-403. DOI: 10.1016/j.ejps.2010.05.003
  92. 92. van Rooy I et al. In vivo methods to study uptake of nanoparticles into the brain. Pharmaceutical Research. 2011;28:456-471. DOI: 10.1007/s11095-010-0291-7
  93. 93. Reichel A. Addressing central nervous system (CNS) penetration in drug discovery: Basics and implications of the evolving new concept. Chemistry & Biodiversity. 2009;6:2030-2049. DOI: 10.1002/cbdv.200900103
  94. 94. Bonate PL. Animal models for studying transport across the blood-brain barrier. Journal of Neuroscience Methods. 1995;56:1-15. DOI: 10.1016/0165-0270(94)00081-q
  95. 95. Artursson P. Epithelial transport of drugs in cell culture. I: A model for studying the passive diffusion of drugs over intestinal absorptive (Caco-2) cells. Journal of Pharmaceutical Sciences. 1990;79:476-482. DOI: 10.1002/jps.2600790604
  96. 96. Raju TN et al. Lancet. 2000;355:1022. DOI: 10.1016/s0140-6736(05)74775-9
  97. 97. Vogt I, Mestres J. Drug-target networks. Molecular Informatics. 2010;29:10-14. DOI: 10.1002/minf.200900069
  98. 98. Reddy AS, Zhang S. Polypharmacology: Drug discovery for the future. Expert Review of Clinical Pharmacology. 2013;6:41-47. DOI: 10.1586/ecp.12.74
  99. 99. Yildirim MA, Goh KI, Cusick ME, Barabási AL, Vidal M. Drug-target network. Nature Biotechnology. 2007;25:1119-1126. DOI: 10.1038/nbt1338
  100. 100. Chen Y, Elenee Argentinis JD, Weber G. IBM Watson: How cognitive computing can be applied to big data challenges in life sciences research. Clinical Therapeutics. 2016;38:688-701. DOI: 10.1016/j.clinthera.2015.12.001
  101. 101. Eisenstein M. Big data: The power of petabytes. Nature. 2015;527:S2-S4. DOI: 10.1038/527S2a

Written By

Xiao-Yuan Mao

Submitted: 14 January 2020 Reviewed: 01 June 2020 Published: 26 June 2020