Next Article in Journal
Investigation of Crystallization and Salt Formation of Poorly Water-Soluble Telmisartan for Enhanced Solubility
Next Article in Special Issue
Brain Delivery of Thyrotropin-Releasing Hormone via a Novel Prodrug Approach
Previous Article in Journal
Three-Step Thermal Drawing for Rapid Prototyping of Highly Customizable Microneedles for Vascular Tissue Insertion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuronanomedicine: An Up-to-Date Overview

by
Daniel Mihai Teleanu
1,
Cristina Chircov
2,3,
Alexandru Mihai Grumezescu
3,4,* and
Raluca Ioana Teleanu
5
1
Emergency University Hospital, “Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania
2
Faculty of Engineering in Foreign Languages, University Politehnica of Bucharest, 060042 Bucharest, Romania
3
Department of Science and Engineering of Oxide Materials and Nanomaterials, Faculty of Applied Chemistry and Materials Science, Politehnica University of Bucharest, 060042 Bucharest, Romania
4
ICUB—Research Institute of University of Bucharest, University of Bucharest, 36-46 M. Kogalniceanu Blvd., 050107 Bucharest, Romania
5
“Dr. Victor Gomoiu” Clinical Children’s Hospital, “Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Pharmaceutics 2019, 11(3), 101; https://doi.org/10.3390/pharmaceutics11030101
Submission received: 18 January 2019 / Revised: 21 February 2019 / Accepted: 22 February 2019 / Published: 26 February 2019
(This article belongs to the Special Issue Drug Delivery to the Brain)

Abstract

:
The field of neuronanomedicine has recently emerged as the bridge between neurological sciences and nanotechnology. The possibilities of this novel perspective are promising for the diagnosis and treatment strategies of severe central nervous system disorders. Therefore, the development of nano-vehicles capable of permeating the blood–brain barrier (BBB) and reaching the brain parenchyma may lead to breakthrough therapies that could improve life expectancy and quality of the patients diagnosed with brain disorders. The aim of this review is to summarize the recently developed organic, inorganic, and biological nanocarriers that could be used for the delivery of imaging and therapeutic agents to the brain, as well as the latest studies on the use of nanomaterials in brain cancer, neurodegenerative diseases, and stroke. Additionally, the main challenges and limitations associated with the use of these nanocarriers are briefly presented.

1. Introduction

Neuroscience is a multidisciplinary field that studies the macro- and microscale neuroanatomy, the functional organization of specific brain areas, and the electrophysiology of neurons and synapses. Moreover, neuroscience represents the means to understand the underlying mechanisms involving the structure and function of individual channels and receptor proteins, the development and repair signaling, and the assembly of proteins into molecular machines that regulate neuronal functions [1]. A complete knowledge of brain function is fundamental for developing novel and efficient strategies that will allow for the long-term and minimally invasive diagnosis and treatment of neurological diseases [1,2].
Neurological disorders comprise a wide variety of sporadic and hereditary [3] pathological conditions, including brain cancer, neurodegenerative diseases, multiple sclerosis, and stroke, which can manifest mild to severe symptoms [4,5]. Due to the increase in elderly populations, the prevalence of these diseases is becoming a great concern [5]. Neurological pathologies are characterized by processes involving protein aggregation which subsequently lead to neurodegeneration or dysregulation of immune mechanisms, or by the progressive loss of neuronal structure and function, associated with abnormalities in brain development and function and neuronal death [6,7]. However, diagnosis, management, and monitoring strategies for neurological disorders are currently unsuccessful mostly due to the complexity of the nervous system [8,9]. Additionally, their diagnosis and treatment involve high precision, dedication, and experience [9].
Furthermore, the presence of the blood–brain barrier (BBB) and the blood–cerebrospinal fluid barrier (BCSFB) represents the main cause for limitations in the management of neurological diseases [7]. The BBB comprises the neurovascular unit which includes specialized endothelial cells, pericytes, astrocytes, neurons, and the extracellular matrix [10]. The BBB represents the dynamic interface between the brain and the circulating blood, acting as a gateway to protect the brain from toxins and cells and to maintain its proper microenvironment [11,12] through the tight junctions, an intricate system of proteins between the endothelial cells [10]. By contrast, the BCSFB consists in epithelial cells only, which are responsible for the physical and chemical properties. Similarly, the tight junctions between the epithelial cells prevent the paracellular diffusion of molecules into the cerebrospinal fluid. If the molecules penetrate the barrier, they may enter into the interstitial fluid of the brain [13]. Since the delivery of contrast and therapeutics is restricted by the two barriers, the need to design novel approaches that can effectively target and reach the central nervous system is fundamental for the diagnosis and treatment of brain disorders [7].
Advances in nanotechnology have allowed for a better understanding of the pathological conditions of the nervous system and the development of formulations that could enhance the therapy of neurological diseases [14]. Comprising knowledge from multiple disciplines, including chemistry, physics, engineering, and biology [15,16], nanotechnology is defined as the field which aims to control matter at atomic and molecular levels [17,18,19]. The nanotechnology processes, also termed as nanofacture or ultraprecision engineering [20], have allowed for the development of non-invasive approaches for the delivery of therapeutic and imaging agents across the brain barriers [21,22]. Therefore, the combination of nanotechnology, specifically nanomedicine, and neuroscience has led to the birth of a novel field, neuronanomedicine, through which nanomaterials, nanoformulations, and nanofacturing processes are effectively employed in neurology for understanding physiological and pathological mechanisms and for diagnosing and treating the disorders of the central nervous system.
The scientific interest in designing nanotechnology-based approaches, including nanoparticles, liposomes, dendrimers, micelles, carbon nanotubes, quantum dots, viral vectors, and extracellular vesicles, which have the potential to deliver the appropriate amount of the therapeutic and imaging agent to the brain, has grown rapidly [21,23]. Hence, the subject of this paper is to review the nanocarriers presently used to diagnose and treat the most prevalent neurological disorders (Figure 1).

2. Nanocarriers for Brain Targeting

Delivering drugs to the brain represents a challenge because conventional neuropharmaceuticals do not possess the appropriate physicochemical characteristics regarding molecular size, lipid solubility, and surface charge [24]. Hence, owing to the capacity to modulate their interactions with endothelial cells in the brain through surface functionalization, various nanocarriers have been employed [25]. Therefore, the encapsulation of therapeutic and imaging agents into specific nanocarriers might overcome the challenges associated with the conventional delivery methods across the BBB [26]. Additionally, after intravenous administration, nanocarriers are capable of crossing the tissues in the organism and reach the central nervous system [27]. A summary of the advantages and disadvantages associated with each nanocarrier type, as well as the surface functionalization strategies is presented in Table 1. Furthermore, the specific pathways to cross the BBB for each nanocarrier is presented in Figure 2.

2.1. Organic Nanocarriers

2.1.1. Polymeric Nanoparticles

Polymeric nanoparticles used as nanocarriers involve matrix architectures, most commonly in the form of nanocapsules and nanospheres [28,29]. The most widely used polymers for manufacturing these nanocarriers are biocompatible and biodegradable and of synthetic origin, such as polylactic acid, polyglycolic acid, polylactide-co-polyglycolic acid, poly(ε-caprolactone), and polymethyl methacrylate, and of natural origin, such as chitosan, alginate, gelatin, and albumin [28,30]. The pharmacokinetics of the encapsulated agents is mainly influenced by the structure of the polymer and the entrapping method [31].
The mechanisms for brain uptake and drug release of polymeric nanoparticles have been intensively studied in order to design nanocarriers that can efficiently deliver therapeutics to the central nervous system through systemic and local administration. Hence, the main strategies involve endocytosis or transcytosis through the endothelial cells, accumulation in the brain capillaries resulting in the transfer to the brain parenchyma owing to the high concentration gradient, membrane fluidization through lipid solubilization due to the surfactant effect, tight junctions opening [32], and restricted efflux phenomenon by coating polymers with polysorbates [31]. Furthermore, to improve the transcytosis across the BBB, the surface of the polymeric nanoparticles can be functionalized by the conjugation of targeting peptides or cell-penetrating ligands [32].

2.1.2. Solid-Lipid Nanoparticles

Solid-lipid nanoparticles are the new generation of colloidal nanocarriers consisting of surfactant-stabilized triglycerides, monoglycerides, hard fats, complex glyceride mixtures, or waxes, that are solid at both room and body temperatures [28,33,34]. Their structure usually involves a hydrophobic solid matrix core in which phospholipids are embedded through the hydrophobic tail regions. Therefore, the entrapment efficiency for hydrophobic drugs in the core is higher than conventional nanocarriers [33]. Commonly used solid lipids for the formulation of these nanocarriers are stearic acid, cetyl alcohol, cholesterol butyrate, carnauba wax, beeswax, and emulsifying wax [34,35].
With a size similar to other nanocarriers, between 50 and 1000 nm [35], solid-lipid nanoparticles combine the advantages of liposomes and polymeric nanoparticles, while outcoming the associated individual disadvantages [33]. Hence, solid-lipid nanoparticles are characterized by biocompatibility, high physical stability, bioavailability, drug protection, strict control of release, ease of preparation, good tolerance, and biodegradability without generating toxic by-products [36,37].
As they have the capacity to target the central nervous system and naturally cross the BBB due to their highly lipophilic nature, solid-lipid nanoparticles have been extensively used as nano-vehicles for the delivery of chemotherapeutic drugs into the central nervous system [37,38]. The main mechanisms involved in the brain uptake of solid-lipid nanoparticles are the paracellular pathway through the opening of the tight junctions in the brain microvasculature, passive diffusion, active transport, and endocytosis [39]. Moreover, since apolipoprotein E receptors are predominantly expressed in the brain, the functionalization of solid-lipid nanoparticles with this protein has become an important strategy in enhancing the brain targeted drug delivery [38,40,41].

2.1.3. Liposomes

Liposomes are artificial and spherical vesicles, consisting of single or multiple amphiphilic lipid bilayers which surround an aqueous solution core [42,43,44,45]. As they can entrap both hydrophilic and hydrophobic compounds in the aqueous core and in the phospholipid bilayers, respectively [42,43], they have been extensively used as drug delivery systems to improve the safety and efficiency of therapeutics targeting. Hence, liposomes have been formulated as nanocarriers to efficiently deliver therapeutic molecules, including drugs, vaccines, enzymes, proteins, and nucleic acids [42], and imaging agents for diagnostics [46].
Furthermore, liposomes have demonstrated their potential in neurological applications as they can cross the BBB through passive or active targeting and deliver the appropriate quantity of therapeutic and diagnosis agents to the brain [47]. The main paths for liposomes to reach the brain parenchyma include the adsorption-mediated transcytosis, the receptor-mediated endocytosis, and the disruption of the BBB through external forces [48].

2.1.4. Dendrimers

Dendrimers are a class of nanoscaled artificial, highly branched, globular macromolecules [49,50]. Their tree-like topological structure includes an initiator core, branched repeat units from the core, and functional terminal groups on the external layer of the repeat units [49]. The most common molecules for dendrimer formulations are polyamidoamine, polypropylenimine, and polyaryl ether [50]. As they are able to encapsulate both hydrophilic and hydrophobic molecules, these unrivalled polymer-based nanostructures [51] have been extensively used as nanocarriers to transport various therapeutic and imaging agents [50,52].
Dendrimers possess the capacity to overcome the BBB, and therefore, they have been widely applied in the therapy of central nervous system disorders [53]. Furthermore, they have the ability to cross various cell membranes or biological barriers through the endocytosis-mediated cellular internalization. Specifically, the cellular uptake is mediated by the reversible modulation of the tight junction proteins, such as occludin and actin. Moreover, specific ligands can be conjugated to the surface of the dendrimers for an enhanced brain targeting and facilitated transport across the BBB [54].

2.1.5. Micelles

Micelles are amphiphilic nanocarriers with a particle size within the range of 5–50 nm that spontaneously form under certain conditions of concentration and temperature of the aqueous solution [55]. The mechanisms for generating micelles mainly involve the self-assembly of amphiphilic molecules. Their architecture is characterized by a core formed by the hydrophobic/non-polar regions of the molecules, known as the tail, and the outside surface comprising the hydrophilic/polar regions of the molecules, known as the head. Therefore, micelles have gained great scientific interest due to their ability to deliver poorly water-soluble and lipophilic compounds and their potential to improve drug bioavailability by providing chemical and physical stability and a sustained and controlled release [35].
Micelles penetrate the BBB mainly through the mechanisms of endocytosis and/or transcytosis. Furthermore, the penetration capacity can be enhanced by conjugating specific ligands and antibodies or by applying external thermal or mechanical forces to disrupt the BBB [56].

2.2. Inorganic Nanocarriers

2.2.1. Inorganic Nanoparticles

The significant amount of work in the area of inorganic nanoparticles synthesis and surface modification has contributed immensely to their applicability in the medical field [57]. Moreover, their unique intrinsic optical, electrical, and magnetic properties have paved the way for novel biomedical applications, such as targeted drug delivery, cancer therapy, bioimaging, and biosensing [57,58].
Inorganic nanoparticles, specifically metal, semiconductor, and metal oxide nanoparticles, have attracted great scientific interest owing to the possibility of tuning their size, shape, composition, structure, and porosity and to decorate their surface to facilitate the conjugation of ligands and polymers, thus enhancing their biological performances [58,59]. As they lack the property of biodegradability, silver, iron oxide, and titanium oxide have been mostly applied for tissue bioimaging in disease diagnosis. However, several inorganic nanoparticles, such as gold and silica nanoparticles, have been used as nanocarriers across the BBB [60]. Moreover, as superparamagnetic iron oxide nanoparticles (SPIONs) are relatively large in size with a mean particle diameter higher than 50 nm, and exhibit unfavorable pharmacokinetic behavior that leads to liver and spleen accumulation due to the opsonization and scavenging by the mononuclear phagocyte system, ultra-small SPIONs (USPIONs) have been developed [61] for drug delivery applications.
Furthermore, to increase brain uptake of these nanoparticles, the application of external stimuli, including near-infrared radiation and magnetic field, could enhance the on-demand drug release across the BBB and improve tissue imaging. Moreover, inorganic nanoparticles are characterized by a prolonged enhanced permeability and retention effect which makes them a great candidate for brain cancer therapy [59].

2.2.2. Carbon Nanotubes

Carbon nanotubes are the most commonly used among the class of carbon-based nanomaterials, comprising graphite sheets rolled into tubes with diameters within the nanoscale. Depending on their architecture, carbon nanotubes can be single-walled or multi-walled, with open ends or closed with fullerene caps [62]. Carbon nanotubes have gained great scientific attention in various fields owing to their unique structure, exceptional electrical, mechanical, optical, and thermal properties, and high surface area [63,64]. Their main nanomedical applications involve drug, hormone, and enzyme delivery, gene therapy, tissue engineering, and biosensing [64,65]. Owing to the possibility of functionalization using specific chemical compounds to modify their physical and biological properties, carbon nanotubes have been applied as nanocarrier systems [66]. As they cannot cross the BBB through passive diffusion, the conjugation of compounds that could facilitate the active transport to the brain is essential for the emerging applications in neuronanomedicine.

2.2.3. Quantum Dots

Quantum dots are zero-dimensional nanomaterials which have attracted considerable scientific interest owing to their exceptional optical and electrical properties [67]. Their application in the fields of medicine and biology has emerged as nanoscaled systems for drug delivery, targeted cancer therapy, bioimaging, and transplanted cell labeling and tracking [68]. Similar to carbon nanotubes, quantum dots require subsequent surface functionalizations through which brain targeting and BBB crossing could be possible. Thus, the mechanisms for reaching the brain parenchyma mostly involve the carrier-mediated transport.

2.3. Biological Vectors

2.3.1. Viral Vectors

The use of viral vectors is based on the ability of viruses to enter and insert genetic material into the host’s cells. The application of viral vectors in the central nervous system is mainly represented by gene therapy, which usually involves the delivery of a normal copy of a defective gene and the reduction of the deleterious functions [69]. The most intensively studied and commonly applied in clinical trials for gene therapy and cancer oncolytic therapy are retrovirus vectors, lentivirus vectors, adenovirus vectors, herpes simplex virus type 1, and adeno-associated virus vectors, which possess transgene capacity and expression properties [70].
The main strategies for the transportation of viral vectors across the BBB are the receptor-mediated pathway across the endothelial cells by transcytosis and the transient disruption of the BBB, which allows for the paracellular transport into the brain parenchyma. One method of disruption involves the intravenously administration of a highly concentrated mannitol solution which will result in the osmotic shrinkage of the cells [71].
Studies reported the use of herpes simplex viral vectors to combat stroke by repairing or replacing genes that lead to neuronal damage. Similar results were obtained by using the adenovirus-mediated vectors. Although there are promising solutions and results, gene therapy is still in its infancy due to ethical issues and high risks of therapy failure [72].

2.3.2. Extracellular Vesicles

Extracellular vesicles represent a heterogenous class of cell-derived membrane structures, originating from the endosomal system, termed as exosomes, or shedding from the plasma membrane, termed as microvesicles. Extracellular vesicles can be ubiquitously found in biological fluids and have a role in various physiological and pathological processes [73]. Recently, they have attracted great attention as an additional mechanism for intercellular communication throughout the body, allowing for protein, lipid, and genetic material exchange [73,74].
In the central nervous system, extracellular vesicles are involved in the maintenance of normal neuronal functions and the development of neurodegenerative disorders. The pathways of BBB crossing by extracellular vesicles are mainly through adsorptive-mediated transcytosis or receptor-mediated transcytosis. However, the underlying and precise mechanisms of crossing in physiological and pathological conditions are not completely understood [74].
The applications of extracellular vesicles as nanocarriers for brain disorders include the use of autologous exosomes containing glyceraldehyde-3-phosphate dehydrogenase that can deliver small-interfering RNA to neurons, microglia, and oligodendrocytes. Moreover, they were also used for the delivery of the APP cleaving enzyme for the downregulation of the BACE1 protein. The intranasal administration of curcumin-containing exosomes for the inhibition of brain inflammation and autoimmune responses has been reported. Exosomes can also be used to deliver Stat3 inhibitor JSI-124 to inhibit tumor growth in a glioblastoma model [75].
Considering the abovementioned characteristics, solid-lipid nanoparticles might be preferred for the treatment applications using hydrophobic drugs as they are highly lipophilic, biodegradable, there are no neurotoxic effects reported in the literature, and they can cross the BBB through the paracellular pathway. Nevertheless, if the nature of the therapeutic agent is hydrophilic, polymeric nanoparticles can be applied, but their reported neurotoxicity must be considered. However, for the bioimaging applications, inorganic nanoparticles are the preferred nanocarrier type, as carbon nanotubes and quantum dots exhibit more serious neurotoxic effects.

3. Nanomedicine in Central Nervous System Disorders

The central nervous system comprises hundreds of various highly organized subtypes of neurons and glia, thus being the most complex and specialized body system. Consequently, diseases associated with the central nervous system are equally complex, causing various diagnostically definitive disruptions in behavior [76]. As nanomaterials are considerably advantageous in regard to their effective targeting, non-invasiveness, stability, biodegradability, and possibility to control the encapsulation and release of the drugs, they have gained a great interest in the area of neuromedicine [48]. Therefore, significant advances have been made in the development of nanotherapeutics capable of crossing the BBB for the diagnosis and/or treatment of the central nervous system disorders (Figure 3) [77], which will be thoroughly described (Table 2 and Table 3).

3.1. Brain Cancer

The transport of anti-cancer drugs through the BBB for the treatment of brain tumors remains one of the major challenges in brain cancer therapy. Thus, the development of nanotechnology-based strategies for an efficient brain uptake and controlled release of the active compounds is essential. The characteristics of the nanocarriers mostly depend on the cancer type, tumor characteristics, stage, and location [78].
Novel approaches for the diagnosis of brain cancer involve the use of nanoparticles, liposomes, micelles, and quantum dots for various neuroimaging techniques. Therefore, silica shells double-coated with semiconducting polymer layers were synthesized for fluorescence and photoacoustic brightness imaging [79]. Moreover, phosphonate polyethylene glycol and cyclo RGD functionalized iron-oxide nanoparticles [80] or bovine serum albumin and tumor-specific folic acid [81] functionalized iron-oxide nanoparticles were applied for magnetic resonance imaging. Another strategy for the diagnosis of glioblastoma involves coating gold nanoparticles with the CBP4 peptide for an enhanced binding to the CD133 biomarker [82]. Furthermore, liposomes incorporating heptamethine cyanine dye IR780 [46] and iron-oxide nanoparticles and a near-infrared fluorescence dye [83] were studied for near-infrared fluorescence imaging and magnetic resonance imaging. Magnetic resonance imaging using gadolinium-incorporated micelles was used for the quantitative hemorrhage-risk evaluation due to the correlation between the extravasation of micelles and the hemorrhagic edema site [84]. Another study focused on the use of polyethylene glycol-coated quantum dots as neuroimaging systems at the tumor site through IVIS imaging system. However, to acquire the images, it was necessary to remove the skulls of the mice [85].
Studies regarding the treatment of brain cancer focused on the delivery of various anti-cancer drugs using nanocarriers. Specifically, poly(lactide-co-glycolic) acid nanoparticles containing doxorubicin, cisplatin, and boldine resulted in an efficient internalization into the glioma cells, inducing cytotoxic effects [86] and an effective target-specific delivery [87]. Other polymeric nanocarriers for the treatment of brain cancer include polyethylene glycol and poly(ω-pentadecalactone-co-p-dioxanone) [88] or polyethylene glycol and poly(lactic-co-glycolic) acid [89] block copolymer nanoparticles which led to an improved drug release efficiency and a decrease in tumor size. Moreover, the administration of amphiphilic polymer-lipid nanoparticles containing docetaxel led to the in vivo accumulation at the tumor site, with an enhanced tumor growth inhibition and increased median survival compared to the equivalent clinical dose of docetaxel solution [90]. Liposomal formulations have also been utilized for the delivery of various anti-cancer drugs, namely methotrexate [91], doxorubicin, erlotinib [92], 5-fluorouracil [93], and paclitaxel [94]. The passage through the BBB has been enhanced by coating the liposomes with different molecules. The results showed an extended blood-circulation time by coating with polyethylene glycol [91], an enhanced translocation across the BBB by attaching transferrin for receptor targeting [92,93], and a higher accumulation of the nanocarriers at the tumor site by conjugating the glucose-vitamin C complex [94]. As previously mentioned, the conjugation of dendrimers with molecules such as polyethylene glycol and glioma homing peptides [95], or sialic acid, glucosamine, and concanavalin A could significantly increase tumor penetration and consequently the amount of drug at the tumor site and reduce the efflux of the nanocarriers [96]. Other chemically functionalized nanocarrier systems for targeting brain tumors are micellar formulations for the delivery of curcumin [97] and multi-walled carbon nanotubes, which showed an increased tumor uptake for the targeted systems [98]. Additionally, the in vitro cytotoxic effects of USPIONs on glioblastoma multiforme was evaluated using rat CNS-1 cell cultures. Results showed that the USPIONs entered the cells through clatherin-coated pits which were further internalized in vacuoles and the effects of USPIONs on cell viability and mitopotential were dose- and time-dependent [99].

3.2. Neurodegenerative Diseases

Neurodegenerative diseases are increasingly prevalent age-dependent disorders which represent a major threat to human health. The most common neurodegenerative diseases are Alzheimer’s disease, Parkinson’s disease, Huntington’s disease, and multiple sclerosis, each characterized by their own pathophysiology, from memory and cognitive impairments to motor dysfunctions, affecting the ability to move, speak, or breathe. As effective treatment strategies are urgently needed, extensive studies regarding the potential of nanotechnology have been performed [100].
Characterized by conformational changes of native proteins which lead to the aggregation and formation of insoluble amyloid fibrils, neurodegenerative diseases therapy mostly relies on the development of adequate platforms that could detect the amyloid formations [101]. Neuroimaging applications of nanotechnology for diagnosing these diseases mainly focus on the use of inorganic nanomaterials as imaging nanocarriers, including iron-oxide nanoparticles [102], gadolinium-based nanoparticles [103], and plasmonic nanoparticles [101]. Furthermore, the administration of carbon nanotubes conjugated with the Pittsburgh Compound B could lead to a more effective early diagnosis of Alzheimer’s disease and therapy monitoring [104].
The treatment of neurodegenerative diseases through nanotechnology approaches focuses on both organic and inorganic nanocarriers. Polyethylene glycol and/or poly(lactic-co-glycolic) acid biodegradable polymeric nanoparticles functionalized with specific antibodies [105,106] or oligopeptide drugs [107] have been applied for the elimination of amyloid fibrils in Alzheimer’s disease. Furthermore, in vitro studies showed that the use of polymeric nanocarriers for the delivery of curcumin resulted in an enhanced drug delivery with reduced oxidative stress, inflammation, and plaque load [108]. In vivo studies regarding the administration of chitosan nanoparticles for the delivery of saxagliptin demonstrated the capacity to prevent premature release and enhanced site targeting compared to the equivalent dose of free saxagliptin in solution [109]. The treatment of Parkinson’s disease through the administration of polymeric nanocarriers has also been studied. Thus, administering chitosan nanoparticles through the intranasal route for the delivery of selegiline [110] and pramipexole [111] increases the brain targeting efficiency and the amount of drug reaching the brain by decreasing the pre-systemic metabolism. Moreover, the in vitro and in vivo studies on the delivery of apolipoprotein E and α-mangostin by using phosphatidic acid-conjugated liposomes [112] and transferrin-modified liposomes [113], respectively, confirmed the potential for an enhanced BBB permeation and efficient drug delivery. Polyamidoamine dendrimers containing carbamazepine [114] and micelles containing curcumin [115] have been reported for the treatment of Alzheimer’s disease. Inorganic nanomaterials, such as gold nanoparticles and carbon nanotubes have been studied for reducing the β-amyloid induced Alzheimer’s disease. Therefore, the functionalization of gold nanoparticles with β-amyloid specific peptides led to an enhanced BBB permeation for the in vitro models [116] and adsorption of berberine onto the surface of the multi-walled carbon nanotubes increased the amount of the drug in the brain [117]. Additionally, gold nanoparticles in the form of l-DOPA functionalized multi-branched nanoflower-like gold nanoparticles have shown potential in the treatment of Parkinson’s disease [118]. Furthermore, improvement of motor dysfunctions and decreased apoptosis could be achieved through the administration of cerium oxide nanoparticles, which have the capacity to protect neurons against reactive oxygen species-induced damage [119].

3.3. Stroke

As it can cause disability or even death, stroke represents a major concerning medical emergency. There are two types of stroke characterized by different mechanisms for triggering. Thus, a cerebral blood vessel blockage is the main cause for the ischemic stroke, while the hemorrhagic stroke is triggered by the rupture of the cerebral blood vessel [120]. Although the prevalence of the ischemic stroke is higher, causing inflammation, damages to the neurovascular unit and even neurological death, the available treatments are limited [120,121]. Current strategies for both emergency treatment and recovery focus on the development of inorganic and organic nanoparticles, such as metal and metal oxide nanoparticles, polymeric nanoparticles, and liposomes [122,123].
One pilot study performed on a gel brain phantom, New Zealand rabbits, and a middle-aged human male reported the potential of administering SPIONs for the rapid diagnosis of the emergent stroke through microwave imaging. Injection of the nanoparticles resulted in the possibility to approximate an area of reduced attenuation difference associated with ischemic hypo-perfusion of the left carotid circulation [124].
Studies for stroke therapy reported the use of poly(lactic-co-glycolic) acid nanoparticles functionalized with chlorotoxin as a targeting ligand for the co-delivery of Lexiscan and Nogo-66, for the simultaneous improvement of the BBB permeability and effective targeting of the stroke site. This system has proved its potential for stroke therapy as results showed an increased stroke survival [125]. Furthermore, polyethylene glycol conjugated polyamidoamine dendrimers have been applied as nanoplatforms for the delivery of drugs that could eliminate blood clots from the vessel [126]. The delivery of siRNA and endothelial progenitor cells represent a promising strategy for ischemic stroke therapy. Therefore, SPIONs have been used as nano-vehicles for gene therapy and for cell tracking, simultaneously. In addition, to further increase the migration and the survival rate of the cells, hif-prolyl hydroxylase 2 silencing might be used [127].

3.4. Clinical Applications

Due to the unmet medical need in the treatment of brain diseases, AstraZeneca has been focusing on key aspects of neurodegenerative diseases, analgesia, and psychiatry. Recent works have been studying the MEDI1814, a monoclonal antibody as a potential disease-modifying treatment for Alzheimer’s disease. This strategy is based on the ability of MEDI1814 to selectively target β-amyloid 42 [128], which is highly associated with Alzheimer’s disease [129]. As early trials have proved that MEDI1814 can reduce the levels of the β-amyloid 42, AstraZeneca and Lily are co-developing it as part of the BACE alliance.
Scientists at BiOasis have developed the xB3 patented platform, formerly known as Transcend-peptide, for applications in neuromedicine. This platform involves the use of a human transport protein found circulating at low levels in the blood, which has shown high efficiency in delivering molecules across the BBB through receptor-mediated transcytosis. Preclinical studies proved the capacity of xB3 to transport molecules such as monoclonal antibodies, enzymes, small-interfering RNA, and other types of gene therapies into the brain, thus having a great potential to treat brain cancers and metabolic and neurodegenerative diseases.
Moreover, the Cerense® technology (Pharmidex Pharmaceutical Service, London, UK), formerly LipoBridge® by Genzyme, and G-Technology® (to-BBB, Leiden, The Netherlands), using BBB targeting delivery systems are in the clinical development pipeline. On one hand, the Cerense® technology utilizes short-chain oligoglycerophospholipids that can transiently open the BBB tight junctions and facilitate the drug transport. On the other hand, the G-Technology comprises liposomes coated with polyethylene glycol and covered with glutathione to facilitate drug transport across the BBB through receptor-mediated transcytosis [130].

4. Challenges and Limitations

The continuous emergence of nanotechnology in the biomedical field has raised some concerns regarding the potential health risks as opposed to the associated benefits. In certain conditions, the numerous advantageous physicochemical properties of nanomaterials, such as reduced size, reactive surface, high surface to volume ratio, or tunable shape might cause serious toxic effects [131].
On one hand, their unique tunable characteristics might provoke unpredictable biological responses when introduced into the body [131]. As it leads to highly reactive and colloidal instability, the high surface to volume ratio is a cause for nanomaterial aggregation [132]. Consequently, as they form clusters no longer in the nanoscale, the cellular uptake is decreased, and apoptosis might be induced. Furthermore, research studies suggest the capacity of nanomaterials to translocate from the administration site to secondary vital organs, including the brain, liver, heart, lungs, and kidneys [133]. As nanomaterials could exert serious toxic effects at the accumulation site, thorough in vivo experiments that investigate organ toxicity and carcinogenicity are vital [134].
Moreover, the administration of nanomaterials for neurological purposes might lead to immediate and direct neurotoxic effects. Therefore, oxidative stress, induced cell apoptosis and necrosis, and immune responses and inflammation as main neurotoxic effects, result in the activation of specific signaling pathways that will further affect the function of the BBB. Additionally, the neurotoxic effect could directly alter the structure and activity of the neurons, or, due to the activation of glial cells and interactions between glial and neuronal cells, it might result in a cascade of effects. Neurotoxic effects manifest immediately or after certain periods of time, leading to reversible or permanent consequences that can affect parts of the nervous system or the whole system [135]. Table 4 summarizes the main neurotoxic effects of the previously described nanocarriers.
Another challenge for the administration of nanomaterials in biological systems is the formation of the biocorona on the surface. The biocorona might lead to alterations of the physicochemical properties, functionality, and biodistribution, and induce highly toxic effects [136].
On the other hand, the application of nanotechnology in the field of biomedicine is limited by the lack of standardized model systems, experimental assays, and in vivo monitoring systems to accurately determine the toxic effects of nanomaterials. Current BBB models involve the use of primary co-cultures of mouse brain endothelial cells and astrocytes, primary mono-, co-, and triple-cultures of rat endothelial cells/astrocytes/pericytes, bovine co-cultures of endothelial cells and astrocytes, porcine monocultures of endothelial cells, and human cultures using either the cCMEC/D3 endothelial cell line or stem cells. As the BBB is a highly dynamic barrier and its properties change in various physiological and pathological conditions, these models should be further refined to allow for the translation of results to the in vivo settings [137]. Additionally, the mechanisms underlying the impact of nanomaterials on biological systems are incompletely understood and further research work is fundamental in order to limit the risks associated with the neuronanomedicinal strategies [131]. One possible solution to accelerate the transition from the in vitro and animal model studies is the implementation of recently developed techniques of lab-on-a-chip and organ cultures. This strategy could allow for more rapid and accurate results regarding the efficiency of the nanomedical approach and the safety of applying it to the human body.

5. Conclusions and Perspectives

Neuronanomedicine has merged the fields of neuroscience and nanotechnology, through which nanomaterials, nanoformulations, and nanofacturing processes are effectively applied in neurology to further understand the physiological and pathological mechanisms. By this means, novel strategies for diagnosing and treating the disorders of the central nervous system have emerged. Specifically, nanotechnology mostly focuses on the development of organic and inorganic nanocarriers, but also on the use of biological entities, such as viral vectors or extracellular vesicles, that could efficiently deliver imaging and therapeutic agents across the BBB, into the brain parenchyma. Although studies are showing promising results, there are several limitations regarding the immediate and long-term interactions of these nanocarriers with the biological tissues. Additional coatings using extracellular matrix-derived polymers or anti-microbial materials and antibiotic treatments might offer possible solutions to reduce neurotoxic effects and to avoid bacteria adherence on the surface of the nanocarrier.
Besides the development of standardized experimental assays, future perspectives might also focus on the development of nanotechnology approaches for neuronal cell regeneration and reconstruction. Furthermore, there are promising possibilities in the field neurosurgery that could benefit from the advantages of nanotechnologies. One example is represented by the implementation of nanorobotics in neurosurgery that involves several manipulation technologies, such as the assembly of nanosized objects and biological cell and molecules manipulation. Moreover, this could also represent a solution for the early diagnosis and therapy monitoring of central nervous system disorders.

Author Contributions

D.M.T., C.C., A.M.G., and R.I.T. designed and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kennedy, M.B. Biochemistry and neuroscience: The twain need to meet. Curr. Opin. Neurobiol. 2017, 43, 79–86. [Google Scholar] [CrossRef] [PubMed]
  2. Rivnay, J.; Wang, H.; Fenno, L.; Deisseroth, K.; Malliaras, G.G. Next-generation probes, particles, and proteins for neural interfacing. Sci. Adv. 2017, 3, e1601649. [Google Scholar] [CrossRef] [PubMed]
  3. Mader, B.J.; Boulis, N.M. Chapter 9—Viral Vectors and Other Modulatory Biologics. In Innovative Neuromodulation; Arle, J., Shils, J., Eds.; Academic Press: San Diego, CA, USA, 2017; pp. 171–205. [Google Scholar]
  4. Crean, P.M.; Tirupathi, S. 24-Essentials of Neurology and Neuromuscular Disorders. In A Practice of Anesthesia for Infants and Children, 6th ed.; Coté, C.J., Lerman, J., Anderson, B.J., Eds.; Elsevier: Philadelphia, PA, USA, 2019; pp. 561.e4–580.e4. [Google Scholar]
  5. Barnabas, W. Drug targeting strategies into the brain for treating neurological diseases. J. Neurosci. Methods 2018, 311, 133–146. [Google Scholar] [CrossRef] [PubMed]
  6. Di Luca, M.; Nutt, D.; Oertel, W.; Boyer, P.; Jaarsma, J.; Destrebecq, F.; Esposito, G.; Quoidbach, V. Towards earlier diagnosis and treatment of disorders of the brain. Bull. World Health Organ. 2018, 96, 298. [Google Scholar] [CrossRef] [PubMed]
  7. Komane, P.P.; Choonara, Y.E.; Toit, L.C.d.; Kumar, P.; Kondiah, P.P.D.; Modi, G.; Pillay, V. Diagnosis and Treatment of Neurological and Ischemic Disorders Employing Carbon Nanotube Technology. J. Nanomater. 2016, 2016, 34. [Google Scholar] [CrossRef]
  8. Dong, X. Current Strategies for Brain Drug Delivery. Theranostics 2018, 8, 1481–1493. [Google Scholar] [CrossRef] [PubMed]
  9. Siuly, S.; Zhang, Y. Medical Big Data: Neurological Diseases Diagnosis Through Medical Data Analysis. Data Sci. Eng. 2016, 1, 54–64. [Google Scholar] [CrossRef]
  10. Gupta, S.; Dhanda, S.; Sandhir, R. Anatomy and physiology of blood-brain barrier. In Brain Targeted Drug Delivery System; Gao, H., Gao, X., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 7–31. [Google Scholar]
  11. Alexander, J.J. Blood-brain barrier (BBB) and the complement landscape. Mol. Immunol. 2018, 102, 26–31. [Google Scholar] [CrossRef] [PubMed]
  12. Tumani, H.; Huss, A.; Bachhuber, F. Chapter 2—The cerebrospinal fluid and barriers—Anatomic and physiologic considerations. In Handbook of Clinical Neurology; Deisenhammer, F., Teunissen, C.E., Tumani, H., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 21–32. [Google Scholar]
  13. Herrera, E.M.; Phillips-Farfán, B.V.; Ospina, G.G. Endothelial Cell Plasticity in the Normal and Injured Central Nervous System; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  14. Edwards, C.A.; Kouzani, A.; Lee, K.H.; Ross, E.K. Neurostimulation Devices for the Treatment of Neurologic Disorders. Mayo Clin. Proc. 2017, 92, 1427–1444. [Google Scholar] [CrossRef] [PubMed]
  15. Kaur, M.; Singh, G.; Khanna, K.; Kaur, N. Nanotechnology: A Review, Second National Conference on Advances in Manufacturing Systems; S B S State Technical Campus: Ferozepur, India, 2015. [Google Scholar]
  16. Husain, Q. Carbon nanotubes mediated immobilized glucose oxidase biosensors as an effective and sensitive analytical tool. Biointerface Res. Appl. Chem. 2018, 8, 3060–3074. [Google Scholar]
  17. Kargozar, S.; Mozafari, M. Nanotechnology and Nanomedicine: Start small, think big. Mater. Today Proc. 2018, 5, 15492–15500. [Google Scholar] [CrossRef]
  18. Afchangi, L.; Fazli, M. Investigation of the nanotubes-H2O complex by molecular mechanics and semi empirical methods. Biointerface Res. Appl. Chem. 2018, 8, 3700–3704. [Google Scholar]
  19. Naskar, A.; Bera, S.; Bhattacharya, R.; Roy, S.S.; Jana, S. Solution based PEG and PVP capped maghemite–reduced graphene oxide nanocomposites: Cell viability study. Biointerface Res. Appl. Chem. 2018, 8, 3751–3757. [Google Scholar]
  20. Ramsden, J.J. Chapter 1—What is Nanotechnology? In Applied Nanotechnology, 3rd ed.; Ramsden, J.J., Ed.; William Andrew Publishing: Amsterdam, The Netherlands, 2018; pp. 3–13. [Google Scholar]
  21. Çetin, M.; Aytekin, E.; Yavuz, B.; Bozdağ-Pehlivan, S. Chapter 7—Nanoscience in Targeted Brain Drug Delivery. In Nanotechnology Methods for Neurological Diseases and Brain Tumors; Gürsoy-Özdemir, Y., Bozdağ-Pehlivan, S., Sekerdag, E., Eds.; Academic Press: Cambridge, MA, USA, 2017; pp. 117–147. [Google Scholar]
  22. Rasoolzadeh, R.; Mehrnejad, F.; Taghdir, M.; Yaghmaei, P. Theoretical investigation of interaction energies between carbon and BN nanotubes with human hepcidin peptides: Insights into the semi empirical and Monte Carlo methods. Biointerface Res. Appl. Chem. 2018, 8, 3594–3601. [Google Scholar]
  23. Feldman, D. Polymer nanocomposites for tissue engineering, antimicrobials and drug delivery. Biointerface Res. Appl. Chem. 2018, 8, 3153–3160. [Google Scholar]
  24. Rahman, H.; Krishnamoorthy, B.; Tamilselvan, N.; Siram, K.; Karthik, S.; Hariprasad, R. Chapter 7—Nanomaterials in drug delivery: Existing scenario and potential scope. In Nanobiomaterials in Drug Delivery; Grumezescu, A.M., Ed.; William Andrew Publishing: Amsterdam, The Netherlands, 2016; pp. 197–228. [Google Scholar]
  25. Comoglu, T.; Arisoy, S.; Akkus, Z.B. Nanocarriers for Effective Brain Drug Delivery. Curr. Top. Med. Chem. 2017, 17, 1490–1506. [Google Scholar] [CrossRef] [PubMed]
  26. Niu, X.; Chen, J.; Gao, J. Nanocarriers as a powerful vehicle to overcome blood-brain barrier in treating neurodegenerative diseases: Focus on recent advances. Asian J. Pharm. Sci. 2018. [Google Scholar] [CrossRef]
  27. Claudio, P.; Reatul, K.; Brigitte, E.; Geraldine, P. Chapter 10—Drug-delivery nanocarriers to cross the blood–brain barrier. In Nanobiomaterials in Drug Delivery; Grumezescu, A.M., Ed.; William Andrew Publishing: Amsterdam, The Netherlands, 2016; pp. 333–370. [Google Scholar]
  28. Pund, S.; Joshi, A. Chapter 23—Nanoarchitectures for Neglected Tropical Protozoal Diseases: Challenges and State of the Art. In Nano- and Microscale Drug Delivery Systems; Grumezescu, A.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 439–480. [Google Scholar]
  29. Jeevanandam, J.; Aing, Y.S.; Chan, Y.S.; Pan, S.; Danquah, M.K. Chapter 3—Nanoformulation and Application of Phytochemicals as Antimicrobial Agents. In Antimicrobial Nanoarchitectonics; Grumezescu, A.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 61–82. [Google Scholar]
  30. Sengel-Turk, C.T.; Gumustas, M.; Uslu, B.; Ozkan, S.A. Chapter 10—Nanosized Drug Carriers for Oral Delivery of Anticancer Compounds and the Importance of the Chromatographic Techniques. In Nano- and Microscale Drug Delivery Systems; Grumezescu, A.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 165–195. [Google Scholar]
  31. Singh, N.; Joshi, A.; Toor, A.P.; Verma, G. Chapter 27—Drug delivery: Advancements and challenges. In Nanostructures for Drug Delivery; Andronescu, E., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 865–886. [Google Scholar]
  32. El-Say, K.M.; El-Sawy, H.S. Polymeric nanoparticles: Promising platform for drug delivery. Int. J. Pharm. 2017, 528, 675–691. [Google Scholar] [CrossRef] [PubMed]
  33. Kammari, R.; Das, N.G.; Das, S.K. Chapter 6—Nanoparticulate Systems for Therapeutic and Diagnostic Applications. In Emerging Nanotechnologies for Diagnostics, Drug Delivery and Medical Devices; Mitra, A.K., Cholkar, K., Mandal, A., Eds.; Elsevier: Boston, MA, USA, 2017; pp. 105–144. [Google Scholar]
  34. Chifiriuc, M.C.; Kamerzan, C.; Lazar, V. Chapter 12—Essential Oils and Nanoparticles: New Strategy to Prevent Microbial Biofilms. In Nanostructures for Antimicrobial Therapy; Ficai, A., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 279–291. [Google Scholar]
  35. Joseph, M.; Trinh, H.M.; Mitra, A.K. Chapter 7—Peptide and Protein-Based Therapeutic Agents. In Emerging Nanotechnologies for Diagnostics, Drug Delivery and Medical Devices; Mitra, A.K., Cholkar, K., Mandal, A., Eds.; Elsevier: Boston, MA, USA, 2017; pp. 145–167. [Google Scholar]
  36. Mihai, M.M.; Holban, A.M.; Călugăreanu, A.; Orzan, O.A. Chapter 11—Recent advances in diagnosis and therapy of skin cancers through nanotechnological approaches. In Nanostructures for Cancer Therapy; Ficai, A., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 285–306. [Google Scholar]
  37. Mandal, A.; Bisht, R.; Pal, D.; Mitra, A.K. Chapter 4—Diagnosis and Drug Delivery to the Brain: Novel Strategies. In Emerging Nanotechnologies for Diagnostics, Drug Delivery and Medical Devices; Mitra, A.K., Cholkar, K., Mandal, A., Eds.; Elsevier: Boston, MA, USA, 2017; pp. 59–83. [Google Scholar]
  38. Neves, A.R.; Queiroz, J.F.; Weksler, B.; Romero, I.A.; Couraud, P.O.; Reis, S. Solid lipid nanoparticles as a vehicle for brain-targeted drug delivery: Two new strategies of functionalization with apolipoprotein E. Nanotechnology 2015, 26, 495103. [Google Scholar] [CrossRef] [PubMed]
  39. Gastaldi, L.; Battaglia, L.; Peira, E.; Chirio, D.; Muntoni, E.; Solazzi, I.; Gallarate, M.; Dosio, F. Solid lipid nanoparticles as vehicles of drugs to the brain: Current state of the art. Eur. J. Pharm. Biopharm. 2014, 87, 433–444. [Google Scholar] [CrossRef] [PubMed]
  40. Neves, A.R.; Queiroz, J.F.; Lima, S.A.C.; Reis, S. Apo E-Functionalization of Solid Lipid Nanoparticles Enhances Brain Drug Delivery: Uptake Mechanism and Transport Pathways. Bioconjugate Chem. 2017, 28, 995–1004. [Google Scholar] [CrossRef] [PubMed]
  41. Magro, R.D.; Ornaghi, F.; Cambianica, I.; Beretta, S.; Re, F.; Musicanti, C.; Rigolio, R.; Donzelli, E.; Canta, A.; Ballarini, E.; et al. ApoE-modified solid lipid nanoparticles: A feasible strategy to cross the blood-brain barrier. J. Control. Release Off. J. Control. Release Soc. 2017, 249, 103–110. [Google Scholar] [CrossRef] [PubMed]
  42. Abbina, S.; Parambath, A. 14—PEGylation and its alternatives: A summary. In Engineering of Biomaterials for Drug Delivery Systems; Parambath, A., Ed.; Woodhead Publishing: Sawston, UK, 2018; pp. 363–376. [Google Scholar]
  43. Kulkarni, V.S.; Shaw, C. Chapter 4—Formulating Creams, Gels, Lotions, and Suspensions. In Essential Chemistry for Formulators of Semisolid and Liquid Dosages; Kulkarni, V.S., Shaw, C., Eds.; Academic Press: Boston, MA, USA, 2016; pp. 29–41. [Google Scholar]
  44. Karandikar, S.; Mirani, A.; Waybhase, V.; Patravale, V.B.; Patankar, S. Chapter 10—Nanovaccines for oral delivery-formulation strategies and challenges. In Nanostructures for Oral Medicine; Andronescu, E., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 263–293. [Google Scholar]
  45. Rai, M.; Ingle, A.P.; Bansod, S.; Kon, K. Chapter 9—Tackling the Problem of Tuberculosis by Nanotechnology: Disease Diagnosis and Drug Delivery. In Nanotechnology in Diagnosis, Treatment and Prophylaxis of Infectious Diseases; Rai, M., Kon, K., Eds.; Academic Press: Boston, MA, USA, 2015; pp. 133–149. [Google Scholar]
  46. Li, S.; Johnson, J.; Peck, A.; Xie, Q. Near infrared fluorescent imaging of brain tumor with IR780 dye incorporated phospholipid nanoparticles. J. Transl. Med. 2017, 15, 18. [Google Scholar] [CrossRef] [PubMed]
  47. Montesinos, R.N. Liposomal Drug Delivery to the Central Nervous System. In Liposomes; Catala, A., Ed.; IntechOpen: London, UK, 2017. [Google Scholar]
  48. Zhou, Y.; Peng, Z.; Seven, E.S.; Leblanc, R. Crossing the blood-brain barrier with nanoparticles. J. Control. Release 2018, 270, 290–303. [Google Scholar] [CrossRef] [PubMed]
  49. Priya, L.B.; Baskaran, R.; Padma, V.V. Chapter 21—Phytonanoconjugates in oral medicine. In Nanostructures for Oral Medicine; Andronescu, E., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 639–668. [Google Scholar]
  50. Verma, G.; Rajagopalan, M.D.; Valluru, R.; Sridhar, K.A. Chapter 7—Nanoparticles: A Novel Approach to Target Tumors. In Nano- and Microscale Drug Delivery Systems; Grumezescu, A.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 113–129. [Google Scholar]
  51. Gumustas, M.; Sengel-Turk, C.T.; Gumustas, A.; Ozkan, S.A.; Uslu, B. Chapter 5—Effect of Polymer-Based Nanoparticles on the Assay of Antimicrobial Drug Delivery Systems. In Multifunctional Systems for Combined Delivery, Biosensing and Diagnostics; Grumezescu, A.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 67–108. [Google Scholar]
  52. Acharya, G.; Mitra, A.K.; Cholkar, K. Chapter 10—Nanosystems for Diagnostic Imaging, Biodetectors, and Biosensors. In Emerging Nanotechnologies for Diagnostics, Drug Delivery and Medical Devices; Mitra, A.K., Cholkar, K., Mandal, A., Eds.; Elsevier: Boston, MA, USA, 2017; pp. 217–248. [Google Scholar]
  53. Srinageshwar, B.; Peruzzaro, S.; Andrews, M.; Johnson, K.; Hietpas, A.; Clark, B.; McGuire, C.; Petersen, E.; Kippe, J.; Stewart, A.; et al. PAMAM Dendrimers Cross the Blood-Brain Barrier When Administered through the Carotid Artery in C57BL/6J Mice. Int. J. Mol. Sci. 2017, 18, 628. [Google Scholar] [CrossRef] [PubMed]
  54. Xu, L.; Zhang, H.; Wu, Y. Dendrimer advances for the central nervous system delivery of therapeutics. ACS Chem. Neurosci. 2014, 5, 2–13. [Google Scholar] [CrossRef] [PubMed]
  55. Rana, S.; Bhattacharjee, J.; Barick, K.C.; Verma, G.; Hassan, P.A.; Yakhmi, J.V. Chapter 7—Interfacial engineering of nanoparticles for cancer therapeutics. In Nanostructures for Cancer Therapy; Ficai, A., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 177–209. [Google Scholar]
  56. Wang, X.-J.; Gao, Y.-P.; Lu, N.-N.; Li, W.-S.; Xu, J.-F.; Ying, X.-Y.; Wu, G.; Liao, M.-H.; Tan, C.; Shao, L.-X.; et al. Endogenous Polysialic Acid Based Micelles for Calmodulin Antagonist Delivery against Vascular Dementia. ACS Appl. Mater. Interfaces 2016, 8, 35045–35058. [Google Scholar] [CrossRef] [PubMed]
  57. Atabaev, T.S. Chapter 8—Multimodal inorganic nanoparticles for biomedical applications. In Nanobiomaterials in Medical Imaging; Grumezescu, A.M., Ed.; William Andrew Publishing: Amsterdam, The Netherlands, 2016; pp. 253–278. [Google Scholar]
  58. Jiao, M.; Zhang, P.; Meng, J.; Li, Y.; Liu, C.; Luo, X.; Gao, M. Recent advancements in biocompatible inorganic nanoparticles towards biomedical applications. Biomater. Sci. 2018, 6, 726–745. [Google Scholar] [CrossRef] [PubMed]
  59. Tsou, Y.H.; Zhang, X.Q.; Zhu, H.; Syed, S.; Xu, X. Drug Delivery to the Brain across the Blood-Brain Barrier Using Nanomaterials. Small 2017, 13, 1701921. [Google Scholar] [CrossRef] [PubMed]
  60. Teleanu, D.M.; Chircov, C.; Grumezescu, A.M.; Volceanov, A.; Teleanu, R.I. Blood-Brain Delivery Methods Using Nanotechnology. Pharmaceutics 2018, 10, 269. [Google Scholar] [CrossRef] [PubMed]
  61. Zarschler, K.; Rocks, L.; Licciardello, N.; Boselli, L.; Polo, E.; Garcia, K.P.; de Cola, L.; Stephan, H.; Dawson, K.A. Ultrasmall inorganic nanoparticles: State-of-the-art and perspectives for biomedical applications. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 1663–1701. [Google Scholar] [CrossRef] [PubMed]
  62. Subramani, K.; Mehta, M. Chapter 19—Nanodiagnostics in microbiology and dentistry. In Emerging Nanotechnologies in Dentistry, 2nd ed.; Subramani, K., Ahmed, W., Eds.; William Andrew Publishing: Amsterdam, The Netherlands, 2018; pp. 391–419. [Google Scholar]
  63. Xue, Y. Chapter 11—Carbon Nanotubes for Biomedical Applications. In Industrial Applications of Carbon Nanotubes; Peng, H., Li, Q., Chen, T., Eds.; Elsevier: Boston, MA, USA, 2017; pp. 323–346. [Google Scholar]
  64. Ajitha, A.; Akhina, H.; Aswathi, M.; LovelyMathew, P.; Sabu, T. Carbon Nanotubes: An Ideal Candidate for Biomedical Applications. JSM Nanotechnol. Nanomed. 2018, 6, 1065. [Google Scholar]
  65. Alshehri, R.; Ilyas, A.M.; Hasan, A.; Arnaout, A.; Ahmed, F.; Memic, A. Carbon Nanotubes in Biomedical Applications: Factors, Mechanisms, and Remedies of Toxicity. J. Med. Chem. 2016, 59, 8149–8167. [Google Scholar] [CrossRef] [PubMed]
  66. Ahmed, W.; Elhissi, A.; Dhanak, V.; Subramani, K. Chapter 18—Carbon nanotubes: Applications in cancer therapy and drug delivery research. In Emerging Nanotechnologies in Dentistry, 2nd ed.; Subramani, K., Ahmed, W., Eds.; William Andrew Publishing: Amsterdam, The Netherlands, 2018; pp. 371–389. [Google Scholar]
  67. Granada-Ramírez, D.A.; Arias-Cerón, J.S.; Rodriguez-Fragoso, P.; Vázquez-Hernández, F.; Luna-Arias, J.P.; Herrera-Perez, J.L.; Mendoza-Álvarez, J.G. 16—Quantum dots for biomedical applications. In Nanobiomaterials; Narayan, R., Ed.; Woodhead Publishing: Sawston, UK, 2018; pp. 411–436. [Google Scholar]
  68. Aswathi, M.; Ajitha, A.; Akhina, H.; Lovely, M.; Thomas, S. Quantum Dots: A Promising Tool for Biomedical application. JSM Nanotechnol. Nanomed. 2018, 6, 1066. [Google Scholar]
  69. Hunter, J.E.; Ramos, L.; Wolfe, J.H. Viral Vectors in the CNS. In Reference Module in Neuroscience and Biobehavioral Psychology; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar]
  70. Choudhury, S.R.; Hudry, E.; Maguire, C.A.; Sena-Esteves, M.; Breakefield, X.O.; Grandi, P. Viral vectors for therapy of neurologic diseases. Neuropharmacology 2017, 120, 63–80. [Google Scholar] [CrossRef] [PubMed]
  71. Fu, H.; McCarty, D.M. Crossing the blood-brain-barrier with viral vectors. Curr. Opin. Virol. 2016, 21, 87–92. [Google Scholar] [CrossRef] [PubMed]
  72. Pulicherla, K.K.; Verma, M.K. Targeting therapeutics across the blood brain barrier (BBB), prerequisite towards thrombolytic therapy for cerebrovascular disorders-an overview and advancements. AAPS PharmSciTech 2015, 16, 223–233. [Google Scholar] [CrossRef] [PubMed]
  73. Van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef] [PubMed]
  74. Matsumoto, J.; Stewart, T.; Banks, W.A.; Zhang, J. The Transport Mechanism of Extracellular Vesicles at the Blood-Brain Barrier. Curr. Pharm. Des. 2017, 23, 6206–6214. [Google Scholar] [CrossRef] [PubMed]
  75. András, I.E.; Toborek, M. Extracellular Vesicles of the Blood-Brain Barrier. Tissue Barriers 2015. [Google Scholar] [CrossRef] [PubMed]
  76. Dougherty, J.D.; Yang, C.; Lake, A.M. Systems biology in the central nervous system: A brief perspective on essential recent advancements. Curr. Opin. Syst. Biol. 2017, 3, 67–76. [Google Scholar] [CrossRef] [PubMed]
  77. Soni, S.; Ruhela, R.K.; Medhi, B. Nanomedicine in Central Nervous System (CNS) Disorders: A Present and Future Prospective. Adv. Pharm. Bull. 2016, 6, 319–335. [Google Scholar] [CrossRef] [PubMed]
  78. Saeedi, M.; Eslamifar, M.; Khezri, K.; Dizaj, S.M. Applications of nanotechnology in drug delivery to the central nervous system. Biomed. Pharm. 2019, 111, 666–675. [Google Scholar] [CrossRef] [PubMed]
  79. Zhen, X.; Feng, X.; Xie, C.; Zheng, Y.; Pu, K. Surface engineering of semiconducting polymer nanoparticles for amplified photoacoustic imaging. Biomaterials 2017, 127, 97–106. [Google Scholar] [CrossRef] [PubMed]
  80. Richard, S.; Boucher, M.; Lalatonne, Y.; Mériaux, S.; Motte, L. Iron oxide nanoparticle surface decorated with cRGD peptides for magnetic resonance imaging of brain tumors. Biochim. Et Biophys. Acta Gen. Subj. 2017, 1861, 1515–1520. [Google Scholar] [CrossRef] [PubMed]
  81. Wang, X.; Tu, M.; Tian, B.; Yi, Y.; Wei, Z.; Wei, F. Synthesis of tumor-targeted folate conjugated fluorescent magnetic albumin nanoparticles for enhanced intracellular dual-modal imaging into human brain tumor cells. Anal. Biochem. 2016, 512, 8–17. [Google Scholar] [CrossRef] [PubMed]
  82. Cho, J.-H.; Kim, A.R.; Kim, S.-H.; Lee, S.-J.; Chung, H.; Yoon, M.-Y. Development of a novel imaging agent using peptide-coated gold nanoparticles toward brain glioma stem cell marker CD133. Acta Biomater. 2017, 47, 182–192. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, L.; Habib, A.A.; Zhao, D. Phosphatidylserine-targeted liposome for enhanced glioma-selective imaging. Oncotarget 2016, 7, 38693–38706. [Google Scholar] [CrossRef] [PubMed]
  84. Shiraishi, K.; Wang, Z.; Kokuryo, D.; Aoki, I.; Yokoyama, M. A polymeric micelle magnetic resonance imaging (MRI) contrast agent reveals blood–brain barrier (BBB) permeability for macromolecules in cerebral ischemia-reperfusion injury. J. Control. Release 2017, 253, 165–171. [Google Scholar] [CrossRef] [PubMed]
  85. Huang, N.; Cheng, S.; Zhang, X.; Tian, Q.; Pi, J.; Tang, J.; Huang, Q.; Wang, F.; Chen, J.; Xie, Z.; et al. Efficacy of NGR peptide-modified PEGylated quantum dots for crossing the blood–brain barrier and targeted fluorescence imaging of glioma and tumor vasculature. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 83–93. [Google Scholar] [CrossRef] [PubMed]
  86. Malinovskaya, Y.; Melnikov, P.; Baklaushev, V.; Gabashvili, A.; Osipova, N.; Mantrov, S.; Ermolenko, Y.; Maksimenko, O.; Gorshkova, M.; Balabanyan, V.; et al. Delivery of doxorubicin-loaded PLGA nanoparticles into U87 human glioblastoma cells. Int. J. Pharm. 2017, 524, 77–90. [Google Scholar] [CrossRef] [PubMed]
  87. Mondal, J.; Patra, M.; Panigrahi, A.K.; Khuda-Bukhsh, A.R. Boldine-loaded PLGA nanoparticles have improved efficiency of drug carriage and protective potential against Cisplatin-induced toxicity. J. Ayurveda Integr. Med. 2018. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, E.M.; Quijano, A.R.; Seo, Y.-E.; Jackson, C.; Josowitz, A.D.; Noorbakhsh, S.; Merlettini, A.; Sundaram, R.K.; Focarete, M.L.; Jiang, Z.; et al. Biodegradable PEG-poly(ω-pentadecalactone-co-p-dioxanone) nanoparticles for enhanced and sustained drug delivery to treat brain tumors. Biomaterials 2018, 178, 193–203. [Google Scholar] [CrossRef] [PubMed]
  89. Hua, H.; Zhang, X.; Mu, H.; Meng, Q.; Jiang, Y.; Wang, Y.; Lu, X.; Wang, A.; Liu, S.; Zhang, Y.; et al. RVG29-modified docetaxel-loaded nanoparticles for brain-targeted glioma therapy. Int. J. Pharm. 2018, 543, 179–189. [Google Scholar] [CrossRef] [PubMed]
  90. He, C.; Cai, P.; Li, J.; Zhang, T.; Lin, L.; Abbasi, A.Z.; Henderson, J.T.; Rauth, A.M.; Wu, X.Y. Blood-brain barrier-penetrating amphiphilic polymer nanoparticles deliver docetaxel for the treatment of brain metastases of triple negative breast cancer. J. Control. Release 2017, 246, 98–109. [Google Scholar] [CrossRef] [PubMed]
  91. Hu, Y.; Rip, J.; Gaillard, P.J.; de Lange, E.C.M.; Hammarlund-Udenaes, M. The Impact of Liposomal Formulations on the Release and Brain Delivery of Methotrexate: An In Vivo Microdialysis Study. J. Pharm. Sci. 2017, 106, 2606–2613. [Google Scholar] [CrossRef] [PubMed]
  92. Lakkadwala, S.; Singh, J. Co-delivery of doxorubicin and erlotinib through liposomal nanoparticles for glioblastoma tumor regression using an in vitro brain tumor model. Colloids Surf. B Biointerfaces 2019, 173, 27–35. [Google Scholar] [CrossRef] [PubMed]
  93. Lakkadwala, S.; Singh, J. Dual Functionalized 5-Fluorouracil Liposomes as Highly Efficient Nanomedicine for Glioblastoma Treatment as Assessed in an In Vitro Brain Tumor Model. J. Pharm. Sci. 2018, 107, 2902–2913. [Google Scholar] [CrossRef] [PubMed]
  94. Peng, Y.; Zhao, Y.; Chen, Y.; Yang, Z.; Zhang, L.; Xiao, W.; Yang, J.; Guo, L.; Wu, Y. Dual-targeting for brain-specific liposomes drug delivery system: Synthesis and preliminary evaluation. Bioorg. Med. Chem. 2018, 26, 4677–4686. [Google Scholar] [CrossRef] [PubMed]
  95. Jiang, Y.; Lv, L.; Shi, H.; Hua, Y.; Lv, W.; Wang, X.; Xin, H.; Xu, Q. PEGylated Polyamidoamine dendrimer conjugated with tumor homing peptide as a potential targeted delivery system for glioma. Colloids Surf. B Biointerfaces 2016, 147, 242–249. [Google Scholar] [CrossRef] [PubMed]
  96. Patel, H.K.; Gajbhiye, V.; Kesharwani, P.; Jain, N.K. Ligand anchored poly(propyleneimine) dendrimers for brain targeting: Comparative in vitro and in vivo assessment. J. Colloid Interface Sci. 2016, 482, 142–150. [Google Scholar] [CrossRef] [PubMed]
  97. Tian, C.; Asghar, S.; Xu, Y.; Chen, Z.; Zhang, J.; Ping, Q.; Xiao, Y. Tween 80-modified hyaluronic acid-ss-curcumin micelles for targeting glioma: Synthesis, characterization and their in vitro evaluation. Int. J. Biol. Macromol. 2018, 120, 2579–2588. [Google Scholar] [CrossRef] [PubMed]
  98. Kafa, H.; Wang, J.T.-W.; Rubio, N.; Klippstein, R.; Costa, P.M.; Hassan, H.A.F.M.; Sosabowski, J.K.; Bansal, S.S.; Preston, J.E.; Abbott, N.J.; et al. Translocation of LRP1 targeted carbon nanotubes of different diameters across the blood–brain barrier in vitro and in vivo. J. Control. Release 2016, 225, 217–229. [Google Scholar] [CrossRef] [PubMed]
  99. Pang, R.T.-W.; Fatseas, G.; Koina, M.E.; Adamson, S.; Eamegdool, S.S.; Weible, M.W.; Pham, N.; Pham, B.T.T.; Hawkett, B.S.; Chan-Ling, T. Mechanism of Ultrasmall Superparamagnetic Iron Oxide Nanoparticles-Induced Glioblastoma Multiforme Cytotoxicity: Effects on Mitochondrial Function; Neuro Oncology and Rehabilitation, Brain Disord and Therapy: Brisbane, Australia, 2016; p. 56. [Google Scholar]
  100. Gitler, A.D.; Dhillon, P.; Shorter, J. Neurodegenerative disease: Models, mechanisms, and a new hope. Dis. Models Mech. 2017, 10, 499–502. [Google Scholar] [CrossRef] [PubMed]
  101. Kumar, J.; Eraña, H.; López-Martínez, E.; Claes, N.; Martín, V.F.; Solís, D.M.; Bals, S.; Cortajarena, A.L.; Castilla, J.; Liz-Marzán, L.M. Detection of amyloid fibrils in Parkinson’s disease using plasmonic chirality. Proc. Natl. Acad. Sci. USA 2018, 115, 3225–3230. [Google Scholar] [CrossRef] [PubMed]
  102. Pansieri, J.; Gerstenmayer, M.; Lux, F.; Mériaux, S.; Tillement, O.; Forge, V.; Larrat, B.; Marquette, C. Magnetic Nanoparticles Applications for Amyloidosis Study and Detection: A Review. Nanomaterials 2018, 8, 740. [Google Scholar] [CrossRef] [PubMed]
  103. Plissonneau, M.; Pansieri, J.; Heinrich-Balard, L.; Morfin, J.-F.; Stransky-Heilkron, N.; Rivory, P.; Mowat, P.; Dumoulin, M.; Cohen, R.; Allémann, É.; et al. Gd-nanoparticles functionalization with specific peptides for ß-amyloid plaques targeting. J. Nanobiotechnol. 2016, 14, 60. [Google Scholar] [CrossRef] [PubMed]
  104. Costa, P.M.; Wang, J.T.-W.; Morfin, J.-F.; Khanum, T.; To, W.; Sosabowski, J.; Tóth, E.; Al-Jamal, K.T. Functionalised Carbon Nanotubes Enhance Brain Delivery of Amyloid-Targeting Pittsburgh Compound B (PiB)-Derived Ligands. Nanotheranostics 2018, 2, 168–183. [Google Scholar] [CrossRef] [PubMed]
  105. Carradori, D.; Balducci, C.; Re, F.; Brambilla, D.; le Droumaguet, B.; Flores, O.; Gaudin, A.; Mura, S.; Forloni, G.; Ordoñez-Gutierrez, L.; et al. Antibody-functionalized polymer nanoparticle leading to memory recovery in Alzheimer’s disease-like transgenic mouse model. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 609–618. [Google Scholar] [CrossRef] [PubMed]
  106. Loureiro, J.A.; Gomes, B.; Fricker, G.; Coelho, M.A.N.; Rocha, S.; Pereira, M.C. Cellular uptake of PLGA nanoparticles targeted with anti-amyloid and anti-transferrin receptor antibodies for Alzheimer’s disease treatment. Colloids Surf. B Biointerfaces 2016, 145, 8–13. [Google Scholar] [CrossRef] [PubMed]
  107. Zheng, X.; Zhang, C.; Guo, Q.; Wan, X.; Shao, X.; Liu, Q.; Zhang, Q. Dual-functional nanoparticles for precise drug delivery to Alzheimer’s disease lesions: Targeting mechanisms, pharmacodynamics and safety. Int. J. Pharm. 2017, 525, 237–248. [Google Scholar] [CrossRef] [PubMed]
  108. Barbara, R.; Belletti, D.; Pederzoli, F.; Masoni, M.; Keller, J.; Ballestrazzi, A.; Vandelli, M.A.; Tosi, G.; Grabrucker, A.M. Novel Curcumin loaded nanoparticles engineered for Blood-Brain Barrier crossing and able to disrupt Abeta aggregates. Int. J. Pharm. 2017, 526, 413–424. [Google Scholar] [CrossRef] [PubMed]
  109. Fernandes, J.; Ghate, M.V.; Mallik, S.B.; Lewis, S.A. Amino acid conjugated chitosan nanoparticles for the brain targeting of a model dipeptidyl peptidase-4 inhibitor. Int. J. Pharm. 2018, 547, 563–571. [Google Scholar] [CrossRef] [PubMed]
  110. Sridhar, V.; Gaud, R.; Bajaj, A.; Wairkar, S. Pharmacokinetics and pharmacodynamics of intranasally administered selegiline nanoparticles with improved brain delivery in Parkinson’s disease. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 2609–2618. [Google Scholar] [CrossRef] [PubMed]
  111. Raj, R.; Wairkar, S.; Sridhar, V.; Gaud, R. Pramipexole dihydrochloride loaded chitosan nanoparticles for nose to brain delivery: Development, characterization and in vivo anti-Parkinson activity. Int. J. Biol. Macromol. 2018, 109, 27–35. [Google Scholar] [CrossRef] [PubMed]
  112. Kuo, Y.-C.; Chen, I.Y.; Rajesh, R. Use of functionalized liposomes loaded with antioxidants to permeate the blood–brain barrier and inhibit β-amyloid-induced neurodegeneration in the brain. J. Taiwan Inst. Chem. Eng. 2018, 87, 1–14. [Google Scholar] [CrossRef]
  113. Chen, Z.-L.; Huang, M.; Wang, X.-R.; Fu, J.; Han, M.; Shen, Y.-Q.; Xia, Z.; Gao, J.-Q. Transferrin-modified liposome promotes α-mangostin to penetrate the blood–brain barrier. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 421–430. [Google Scholar] [CrossRef] [PubMed]
  114. Igartúa, D.E.; Martinez, C.S.; Temprana, C.F.; Alonso, S.d.V.; Prieto, M.J. PAMAM dendrimers as a carbamazepine delivery system for neurodegenerative diseases: A biophysical and nanotoxicological characterization. Int. J. Pharm. 2018, 544, 191–202. [Google Scholar] [CrossRef] [PubMed]
  115. Desai, P.P.; Patravale, V.B. Curcumin Cocrystal Micelles—Multifunctional Nanocomposites for Management of Neurodegenerative Ailments. J. Pharm. Sci. 2018, 107, 1143–1156. [Google Scholar] [CrossRef] [PubMed]
  116. Ruff, J.; Hüwel, S.; Kogan, M.J.; Simon, U.; Galla, H.-J. The effects of gold nanoparticles functionalized with ß-amyloid specific peptides on an in vitro model of blood–brain barrier. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1645–1652. [Google Scholar] [CrossRef] [PubMed]
  117. Lohan, S.; Raza, K.; Mehta, S.K.; Bhatti, G.K.; Saini, S.; Singh, B. Anti-Alzheimer’s potential of berberine using surface decorated multi-walled carbon nanotubes: A preclinical evidence. Int. J. Pharm. 2017, 530, 263–278. [Google Scholar] [CrossRef] [PubMed]
  118. Gonzalez-Carter, D.A.; Ong, Z.Y.; McGilvery, C.M.; Dunlop, I.E.; Dexter, D.T.; Porter, A.E. L-DOPA functionalized, multi-branched gold nanoparticles as brain-targeted nano-vehicles. Nanomed. Nanotechnol. Biol. Med. 2019, 15, 1–11. [Google Scholar] [CrossRef] [PubMed]
  119. Hegazy, M.A.; Maklad, H.M.; Samy, D.M.; Abdelmonsif, D.A.; el Sabaa, B.M.; Elnozahy, F.Y. Cerium oxide nanoparticles could ameliorate behavioral and neurochemical impairments in 6-hydroxydopamine induced Parkinson’s disease in rats. Neurochem. Int. 2017, 108, 361–371. [Google Scholar] [CrossRef] [PubMed]
  120. Zhou, Z.; Lu, J.; Liu, W.-W.; Manaenko, A.; Hou, X.; Mei, Q.; Huang, J.-L.; Tang, J.; Zhang, J.H.; Yao, H.; et al. Advances in stroke pharmacology. Pharm. Ther. 2018, 191, 23–42. [Google Scholar] [CrossRef] [PubMed]
  121. Barthels, D.; Das, H. Current advances in ischemic stroke research and therapies. Biochim. Et Biophys. Acta Mol. Basis Dis. 2018. [Google Scholar] [CrossRef]
  122. Sarmah, D.; Saraf, J.; Kaur, H.; Pravalika, K.; Tekade, R.K.; Borah, A.; Kalia, K.; Dave, K.R.; Bhattacharya, P. Stroke management: An emerging role of nanotechnology. Micromachines 2017, 8, 262. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, L.; Gao, X. The application of nanoparticles for neuroprotection in acute ischemic stroke. Ther. Deliv. 2017, 8, 915–928. [Google Scholar] [CrossRef] [PubMed]
  124. Hudson, J.S.; Chung, T.K.; Prout, B.S.; Nagahama, Y.; Raghavan, M.L.; Hasan, D.M. Iron nanoparticle contrast enhanced microwave imaging for emergent stroke: A pilot study. J. Clin. Neurosci. Off. J. Neurosurg. Soc. Australas. 2019, 59, 284–290. [Google Scholar] [CrossRef] [PubMed]
  125. Han, L.; Cai, Q.; Tian, D.; Kong, D.K.; Gou, X.; Chen, Z.; Strittmatter, S.M.; Wang, Z.; Sheth, K.N.; Zhou, J. Targeted drug delivery to ischemic stroke via chlorotoxin-anchored, lexiscan-loaded nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 1833–1842. [Google Scholar] [CrossRef] [PubMed]
  126. Santos, S.D.; Xavier, M.; Leite, D.M.; Moreira, D.A.; Custódio, B.; Torrado, M.; Castro, R.; Leiro, V.; Rodrigues, J.; Tomás, H.; et al. PAMAM dendrimers: Blood-brain barrier transport and neuronal uptake after focal brain ischemia. J. Control. Release 2018, 291, 65–79. [Google Scholar] [CrossRef] [PubMed]
  127. Wang, C.; Lin, G.; Luan, Y.; Ding, J.; Li, P.-C.; Zhao, Z.; Qian, C.; Liu, G.; Ju, S.; Teng, G.-J. HIF-prolyl hydroxylase 2 silencing using siRNA delivered by MRI-visible nanoparticles improves therapy efficacy of transplanted EPCs for ischemic stroke. Biomaterials 2018, 197, 229–243. [Google Scholar] [CrossRef] [PubMed]
  128. Billinton, A.; Newton, P.; Lloyd, C.; Groves, M.; Welsh, F.; Bogstedt, A.; Eketjall, S.; McFarlane, M.; Perkinton, M.; Narwal, R.; et al. Preclinical Discovery and Development of MEDI1814, A Monoclonal Antibody Selectively Targeting Beta-Amyloid 42 (Aβ42). Alzheimer’s Dement. 2017, 13, P266. [Google Scholar] [CrossRef]
  129. Gao, Q.; Fang, Y.; Zhang, S.; Wong, H.S.H.; Chan, Y.E.; Wong, S.S.M.; Yung, K.K.L.; Lai, K.W.C. Dynamic effect of beta-amyloid 42 on cell mechanics. J. Biomech. 2019. [Google Scholar] [CrossRef] [PubMed]
  130. Catalan-Figueroa, J.; Palma-Florez, S.; Alvarez, G.; Fritz, H.F.; Jara, M.O.; Morales, J.O. Nanomedicine and nanotoxicology: The pros and cons for neurodegeneration and brain cancer. Nanomedicine 2016, 11, 171–187. [Google Scholar] [CrossRef] [PubMed]
  131. Patel, S.; Nanda, R.; Sahoo, S. Nanotechnology in Healthcare: Applications and Challenges. Med. Chem. 2015, 5, 528–533. [Google Scholar] [CrossRef]
  132. Zhang, W. Nanoparticle aggregation: Principles and modeling. Adv. Exp. Med. Biol. 2014, 811, 19–43. [Google Scholar] [PubMed]
  133. Kermanizadeh, A.; Balharry, D.; Wallin, H.; Loft, S.; Moller, P. Nanomaterial translocation--the biokinetics, tissue accumulation, toxicity and fate of materials in secondary organs—A review. Crit. Rev. Toxicol. 2015, 45, 837–872. [Google Scholar] [CrossRef] [PubMed]
  134. Wu, T.; Tang, M. Review of the effects of manufactured nanoparticles on mammalian target organs. J. Appl. Toxicol. 2018, 38, 25–40. [Google Scholar] [CrossRef] [PubMed]
  135. Teleanu, D.M.; Chircov, C.; Grumezescu, A.M.; Teleanu, R.I. Neurotoxicity of Nanomaterials: An Up-to-Date Overview. Nanomaterials 2019, 9, 96. [Google Scholar] [CrossRef] [PubMed]
  136. Shannahan, J. The biocorona: A challenge for the biomedical application of nanoparticles. Nanotechnol. Rev. 2017, 6, 345–353. [Google Scholar] [CrossRef] [PubMed]
  137. Helms, H.C.; Abbott, N.J.; Burek, M.; Cecchelli, R.; Couraud, P.O.; Deli, M.A.; Forster, C.; Galla, H.J.; Romero, I.A.; Shusta, E.V.; et al. In vitro models of the blood-brain barrier: An overview of commonly used brain endothelial cell culture models and guidelines for their use. J. Cereb. Blood Flow Metab. Off. J. Int. Soc. Cereb. Blood Flow Metab. 2016, 36, 862–890. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A summary of the various types of nanocarriers for brain targeting.
Figure 1. A summary of the various types of nanocarriers for brain targeting.
Pharmaceutics 11 00101 g001
Figure 2. The main pathways of crossing the BBB for each type of the described nanocarriers.
Figure 2. The main pathways of crossing the BBB for each type of the described nanocarriers.
Pharmaceutics 11 00101 g002
Figure 3. The applications of nanomaterials as nanocarriers in brain cancer, neurodegenerative diseases, and stroke therapy.
Figure 3. The applications of nanomaterials as nanocarriers in brain cancer, neurodegenerative diseases, and stroke therapy.
Pharmaceutics 11 00101 g003
Table 1. The main advantages, disadvantages, and surface functionalization strategies for the organic and inorganic nanocarriers.
Table 1. The main advantages, disadvantages, and surface functionalization strategies for the organic and inorganic nanocarriers.
Nanocarrier TypeAdvantagesDisadvantagesSurface Functionalization Strategies
Polymeric nanoparticlesbiocompatibility, biodegradability, drug protection, ease of preparation, good tolerance
controlled pharmacokinetics
tunable physicochemical properties
neurotoxicitypolysorbate 80
RVG29 peptide
anti-Aβ1-42 antibody
monoclonal antibody (OX26)
anti-Aβ (DE2B4)
g7 ligand
TGN peptides
QSH peptides
l-valine
chlorotoxin
Solid-lipid nanoparticlesbiocompatibility, high physical stability, bioavailability, drug protection, strict control of release, ease of preparation, good tolerance, and biodegradability without generating toxic by-products
no neurotoxic effects reported
hydrophobic drug entrapment efficiency
lipophilicity
possibility of passively cross the BBB
reduced hydrophilic drug entrapment efficiency
sterilization difficulties
apolipoprotein E
Liposomespossibility of entrapping both hydrophilic and hydrophobic compounds
improved drug protection and targeting efficiency
lipophilicity
possibility of passively cross the BBB
neurotoxicity
physicochemical instability
tendency of fusion
rapid clearance
sterilization difficulties
phosphatidylserine-targeting antibody
polyethylene glycol transferrin
PFVYLI peptide
penetratin peptide
glucose-vitamin C complex
phosphatidic acid
apolipoprotein E
Dendrimerspossibility of entrapping both hydrophilic and hydrophobic compounds
biodegradability
stimuli-responsiveness
enhanced targeting efficiency
neurotoxicity
synthesis variability
rapid clearance
organ accumulation
polyethylene glycol
glioma homing peptides
sialic acid
glucosamine
concanavalin A
Micellesno neurotoxic effects reported
improved drug bioavailability
physicochemical stability
sustained and controlled release
use only for lipophilic drugs
low drug loading capacity
Tween 80
Inorganic nanoparticlesunique optical, electrical, and magnetic properties
tunable size, shape, composition, structure, and porosity
prolonged enhanced permeability and retention effect
enhanced on-demand drug release by applying external stimuli (near-infrared radiation and magnetic field)
neurotoxicity
high tendency of aggregation
non-degradableorgan accumulation
need further functionalization for BBB crossing
cyclo RGD peptides
phosphonate polyethylene glycol
bovine serum albumin folic acid
CBP4 peptide
KLVFF and LPFFD peptides
CLPFFD peptides
l-DOPA
hif-prolyl hydroxylase 2 silencing
Carbon nanotubesunique structure, exceptional electrical, mechanical, optical, and thermal properties, and high surface areaneurotoxicity
need further functionalization for BBB crossing
Pittsburgh Compound B
polysorbate and phospholipid coating
Quantum dotsexceptional optical and electrical propertiesneurotoxicity
need further functionalization for BBB crossing
polyethylene glycol
asparagine-glycine-arginine peptides
Table 2. A summary of the nanotechnology-based neuroimaging approaches for the diagnosis of brain cancer, neurodegenerative diseases, and stroke.
Table 2. A summary of the nanotechnology-based neuroimaging approaches for the diagnosis of brain cancer, neurodegenerative diseases, and stroke.
Central Nervous System DisorderNanocarrier TypeFunctionalizationImaging AgentNeuroimaging TechniqueStudy ModelReference
Brain cancersilica shells double coated with semiconducting polymer layerscyclo RGD peptides-fluorescence and photoacoustic brightness imagingin vitro—4T1 human breast cancer epithelial cells
in vivo—tumor-bearing female mice
[79]
iron oxide nanoparticlesphosphonate polyethylene glycol and cyclo RGD peptides-magnetic resonance imagingin vitro—U87-MG cells
in vivo—tumor-bearing nude mice
[80]
bovine serum albumin and tumor-specific folic acidfluorescein isothiocyanatemagnetic resonance imagingin vitro—U251 cells[81]
gold nanoparticlesCBP4 peptidefluorescein isothiocyanateconfocal microscopyin vitro—U373 human glioma cells[82]
liposomes-heptamethine cyanine dye IR780near-infrared fluorescence imagingin vitro—U87MG human glioma cells and T98G human glioblastoma cells
in vivo—glioblastoma mouse models
[46]
phosphatidylserine-targeting antibodyiron oxide nanoparticles and a near-infrared fluorescence dyenear-infrared fluorescence imaging and magnetic resonance imagingin vitro—U87MG human glioma cells
in vivo—tumor-bearing nude mice
[83]
micelles-gadoliniummagnetic resonance imagingin vivo—Wistar male rats[84]
quantum dotspolyethylene glycol and asparagine–glycine–arginine peptides-IVIS imagingin vitro—primary rat BCECs, astrocytes and C6 glioma cells
in vivo—Sprague–Dawley male rats
[85]
Neurodegenerative diseasesgadolinium-based nanoparticlesKLVFF and LPFFD peptides-fluorescence microscopyin vivo—APPswe/PS1A246E/TTR mouse model[103]
carbon nanotubesPittsburgh Compound Bgadolinium complexessingle photon emission computed tomography/computed tomography and γ-scintigraphyin vivo—female C57BL/6 mice[104]
StrokeIron-oxide nanoparticles--microwave imagingin vitro—gel brain phantom
in vivo—New Zealand rabbits and a middle-aged human male
[124]
Table 3. A summary of the nanotechnology-based treatment strategies for brain cancer, neurodegenerative diseases, and stroke.
Table 3. A summary of the nanotechnology-based treatment strategies for brain cancer, neurodegenerative diseases, and stroke.
Central Nervous System DisorderNanocarrier TypeFunctionalizationActive CompoundStudy ModelReference
Brain cancerpoly(lactide-co-glycolic) nanoparticlespoloxamer 188doxorubicinin vitro—U-87 MG, ATCC cell line[86]
-cisplatin and boldinein vivo – tumor-bearing swiss albino mice[87]
polyethylene glycol and poly(ω-pentadecalactone-co-p-dioxanone) nanoparticles-VE822in vitro—RG2 cells
in vivo —Tumor-bearing male Fischer 344 rats
[88]
polyethylene glycol and poly(lactic-co-glycolic) acid nanoparticlesRVG29 peptidedocetaxelin vitro—C6 cells
in vivo—tumor-bearing adult Sprague–Dawley male rats
[89]
amphiphilic polymer-lipid nanoparticlespolysorbate 80docetaxelin vitro—MDA-MB-231 cells
in vivo—tumor-bearing severe combined immune deficiency mice
[90]
liposomespolyethylene glycolmethotrexatein vivo – male Sprague–Dawley rats[91]
transferrin and PFVYLI peptidedoxorubicin and erlotinibin vitro—U87 tumor cells, brain endothelial cells, and glial cells[92]
transferrin and penetratin peptide5-fluorouracilin vitro—U87 tumor cells and brain endothelial cells[93]
glucose-vitamin C complexpaclitaxelin vitro—C6 cells
in vivo—C6 glioma-bearing Kunming mice
[94]
dendrimerspolyethylene glycol and glioma homing peptides-in vitro—U87MG cells
in vivo—U87MG tumor-bearing BALB/c nude mice
[95]
sialic acid, glucosamine, and concanavalin Apaclitaxelin vitro—U373MG human astrocytoma cell line
in vivo—Sprague–Dawley rats
[96]
micellesTween 80curcuminin vitro—G422 cells[97]
multi-walled carbon nanotubesAngiopep-2-in vitro—primary porcine brain endothelial cells and primary rat astrocytes
in vivo—GL261 glioma-bearing female C57/Bl6 mice
[98]
USPIONS--in vitro—rat CNS-1 cells[99]
Neurodegenerative diseasespolyethylene glycol nanoparticlesanti-Aβ1-42 antibody-in vivo—NIHS adult male mice[105]
poly(lactic-co-glycolic) acid nanoparticlesmonoclonal antibody (OX26) and anti-Aβ (DE2B4)-in vitro—porcine brain capillary endothelial cells[106]
poly(lactic-co-glycolic) acid nanoparticlesg7 ligandcurcuminin vitro—primary hippocampal cultures from rat brains[108]
polyethylene glycol-polylactic acid nanoparticlesTGN peptides and QSH peptidescoumarin-6 and H102in vitro—brain endothelial cells
in vivo—5XFAD transgenic mice
[107]
chitosan nanoparticlesL-valinesaxagliptinin vivo—female Wistar rats[109]
-selegilineex vivo—male Sprague–Dawley rats[110]
-pramipexole dihydrochlorideex vivo—goat nasal mucosa
in vivo—male Sprague–Dawley rats
[111]
liposomesphosphatidic acid and apolipoprotein Equercetin and rosmarinic acidin vitro—brain microvascular endothelial cells and Aβ1-42-insulted SK-N-MC cells[112]
transferrinα-mangostinin vitro—brain endothelial cells
in vivo—Sprague–Dawley rats
[113]
polyamidoamine dendrimers-carbamazepineex vivo—human red blood cells
in vitro—N2a cell linein vivo—zebrafish
[114]
micelles-curcuminin vitro—U87MG cell line
in vivo—female Sprague–Dawley rats
[115]
gold nanoparticles CLPFFD peptides, neutral methoxy terminated polyethylene glycol ligands, and negatively-charged monosulfonated triphenylphosphine ligands-in vitro—porcine brain capillary endothelial[116]
L-DOPA-in vitro—human brain endothelial cell line hCMEC/D3, brain microvascular endothelial cells, and mouse microglia N9 cell line[118]
multi-walled carbon nanotubespolysorbate and phospholipid coatingberberinein vitro—human red blood cells and SH-SY5Y cells
in vivo—male Wistar rats
[117]
cerium oxide nanoparticles--in vivo—adult male Wistar rats[119]
Strokepoly(lactic-co-glycolic) acid nanoparticleschlorotoxinLexiscan and Nogo-66in vivo—male C57BL/6 mice[125]
polyamidoamine dendrimerspolyethylene glycol-in vitro—rat primary astrocytes and mouse brain endothelial cells
in vivo—male C57BL/6 mice
[126]
iron oxide nanoparticleshif-prolyl hydroxylase 2 silencingsiRNAin vivo—female BALB/c nude mice[127]
Table 4. A summary of the main neurotoxic effects of the organic and inorganic nanocarriers for BBB crossing [135].
Table 4. A summary of the main neurotoxic effects of the organic and inorganic nanocarriers for BBB crossing [135].
Nanocarrier TypeNeurotoxic Effect
Polymeric nanoparticlesneuronal apoptosis; neuroinflammation; increased oxidative stress
Liposomesnecrosis; neuroinflammation; hemorrhage; macrophage infiltration
Dendrimerscell proliferation and migration inhibition; abnormal mitochondrial activity; apoptosis; affected neuronal differentiation; increased oxidative stress; DNA damage; decreased locomotor function
Gold nanoparticlesincreased oxidative stress; cognition defects; astrogliosis
Silver nanoparticlesincreased oxidative stress; apoptosis; necrosis; neuroinflammation
Iron oxide nanoparticlessynaptic transmission and nerve conduction alterations; neuroinflammation; apoptosis; macrophage infiltration
Titanium oxide nanoparticlesincreased oxidative stress; neuroinflammation; apoptosis; synaptic transmission alterations and plasticity; genotoxicity
Silica nanoparticlescognitive dysfunctions and impairment; neurodegeneration; synaptic transmission alterations
Carbon nanotubesneuroinflammation; cell proliferation inhibition; apoptosis; increased oxidative stress; mitochondrial membrane potential reduction; lipid peroxidization; astrocyte function reduction; neurobehavioral toxicity
Quantum dotsincreased oxidative stress; cell function damage; neurobehavioral toxicity; cognitive impairment

Share and Cite

MDPI and ACS Style

Teleanu, D.M.; Chircov, C.; Grumezescu, A.M.; Teleanu, R.I. Neuronanomedicine: An Up-to-Date Overview. Pharmaceutics 2019, 11, 101. https://doi.org/10.3390/pharmaceutics11030101

AMA Style

Teleanu DM, Chircov C, Grumezescu AM, Teleanu RI. Neuronanomedicine: An Up-to-Date Overview. Pharmaceutics. 2019; 11(3):101. https://doi.org/10.3390/pharmaceutics11030101

Chicago/Turabian Style

Teleanu, Daniel Mihai, Cristina Chircov, Alexandru Mihai Grumezescu, and Raluca Ioana Teleanu. 2019. "Neuronanomedicine: An Up-to-Date Overview" Pharmaceutics 11, no. 3: 101. https://doi.org/10.3390/pharmaceutics11030101

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop