Next Article in Journal
Antioxidant and Anti-Inflammatory Activities of Six Flavonoids from Smilax glabra Roxb
Previous Article in Journal
Virtual Screening and In Vitro Evaluation of PD-L1 Dimer Stabilizers for Uncoupling PD-1/PD-L1 Interaction from Natural Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nano Carrier Drug Delivery Systems for the Treatment of Neuropsychiatric Disorders: Advantages and Limitations

1
Department Basic and Applied Neurobiology, V.P. Serbsky Federal Medical Research Centre of Psychiatry and Narcology, 119034 Moscow, Russia
2
Healthcare Department, Mental-Health Clinic No. 1 Named after N.A. Alexeev of Moscow, 117152 Moscow, Russia
3
Department of Biology, Lomonosov Moscow State University, 119992 Moscow, Russia
4
D. Mendeleev University of Chemical Technology of Russia, 125047 Moscow, Russia
5
Department of Medical Nanobiotechnology, Pirogov Russian National Research Medical University, 117997 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(22), 5294; https://doi.org/10.3390/molecules25225294
Submission received: 20 October 2020 / Revised: 9 November 2020 / Accepted: 11 November 2020 / Published: 13 November 2020
(This article belongs to the Section Nanochemistry)

Abstract

:
Neuropsychiatric diseases are one of the main causes of disability, affecting millions of people. Various drugs are used for its treatment, although no effective therapy has been found yet. The blood brain barrier (BBB) significantly complicates drugs delivery to the target cells in the brain tissues. One of the problem-solving methods is the usage of nanocontainer systems. In this review we summarized the data about nanoparticles drug delivery systems and their application for the treatment of neuropsychiatric disorders. Firstly, we described and characterized types of nanocarriers: inorganic nanoparticles, polymeric and lipid nanocarriers, their advantages and disadvantages. We discussed ways to interact with nerve tissue and methods of BBB penetration. We provided a summary of nanotechnology-based pharmacotherapy of schizophrenia, bipolar disorder, depression, anxiety disorder and Alzheimer’s disease, where development of nanocontainer drugs derives the most active. We described various experimental drugs for the treatment of Alzheimer’s disease that include vector nanocontainers targeted on β-amyloid or tau-protein. Integrally, nanoparticles can substantially improve the drug delivery as its implication can increase BBB permeability, the pharmacodynamics and bioavailability of applied drugs. Thus, nanotechnology is anticipated to overcome the limitations of existing pharmacotherapy of psychiatric disorders and to effectively combine various treatment modalities in that direction.

1. Introduction

According to statistics, one in four people in the world afflicts with various mental or neurological illnesses at some point of their lives. Around 450 million people currently suffer from such conditions, placing mental disorders among the leading causes of ill and disability worldwide [1].
There is a great clinical need for development a new-generation of psychoactive drugs, because the global burden of mental illness continues to increase. Despite the wide range of existing drugs with psychotropic and neurotropic activity, there has been no significant success in treating mental illness. This may be because it takes time to achieve an effective concentration of drugs, with which side effects are associated. This, in turn, forces clinicians to change their treatment methods frequently.
Preparations in their usual form (pills or injections) have many obstacles in their way to achieve the desired effect. These include poor bioavailability, low absorption (food effect), first-pass metabolism, drug toxicity and dose-dependent side effects.
The concept of nanotherapy includes systems with sizes of 10–100 nm and unique physiological and chemical properties. Nanopreparations are distinguished by their plasticity and modified surface properties, flexibility and controlled release of the active drug substance. Nanotherapeutics provide targeted delivery of central nervous system (CNS) active drugs to the brain [2], which is essential for drugs aimed at treating mental illness. An example of such preparations is biodegradable, biocompatible polymers, which are non-immunogenic and non-toxic to cells. Their increased solubility and targeted delivery result in an increased concentration of drugs in cells and tissues, thereby increasing the bioavailability of drug molecules. Therefore, nanoparticles (NPs) application demonstrates a variety of advantages in comparison with other methods of drug administration as intranasal, transdermal, oral and intravenous administration (Table 1).
Currently, various types of nanotherapy for mental disorders, including polymeric nanoparticles, solid lipid nanoparticles (SLN), nanostructured lipid carriers (NLC), nanoemulsions, nanogels, carbon nanotubes and liquid crystalline nanoparticles with neuroprotective properties [3] are being investigated.
In our review we will present current data on promising areas of nanotherapy in the field of mental and neurodegenerative illnesses, such as schizophrenia, bipolar disorder, depression, anxiety and Alzheimer’s disease.

2. Drug Transport through the Blood Brain Barrier

Despite significant progress in understanding the molecular and cellular mechanisms of brain disorders and the development of strategies for their treatment, efficient drug delivery to the central nervous system is an important problem today [4]. One of the most significant barriers in the central nervous system that interferes with drug delivery compounds is the blood brain barrier (BBB). Its functions include not only transporting nutrients and oxygen to the brain from the bloodstream, but also protection of the CNS from neurotoxic substances [5]. The BBB consists of several types of cells such as: endothelial cells, astrocytes, perivascular macrophages and pericytes [6]. The impermeability of the BBB for many molecules is due to the tight contacts between endothelial cells, which in turn form the proteins occludins, claudins and adhesion molecules [7]. Together with tight junctions of endothelial cells, a specialized extracellular matrix and a basement membrane from collagen IV type, laminin, fibronectin, tenascin and proteoglycans provide controlled BBB permeability [8]. The most studied pathways of substances through the BBB are the so-called passive or transcellular diffusion, facilitated diffusion and active transport. Transcellular transport of substances occurs through active and passive diffusion. In the case of active transport in the transfer of molecules, there are three types of transport systems in the BBB: receptor-mediated transport, which carries out the transfer of macromolecules; carrier proteins that transport sugars, amino acids, organic anions and cations, neurotransmitters and metabolites; active transport by proteins of the adenosine triphosphate binding cassette family [9]. Proteins of the adenosine triphosphate binding cassette family, such as P-glycoprotein, make the BBB impermeable to drugs [10].
One of the key problems today is the development of drugs for the treatment of CNS diseases, which will pass through the BBB. For effective therapeutic response the necessary concentration of the drug should accumulate in the right body region and remain maintained enough time to achieve that response. In other organs and tissues, the drug concentration has to stay minimal for side effects reduction. In that case, the presence of BBB significantly complicates the delivery of drugs into brain tissue.
To solve that problem, many methods have been developed to date, such as: violation of the integrity of tight contacts using hypertonic solutions (arabinose and mannitol) [11]; vasoactive drugs (leukotriene C4, IL-2, TNF-α, IFN-γ and bradykinin) [12] and directed ultrasound [13].
Other methods of overcoming the BBB are direct injection of drugs into the brain tissue. In these techniques, it is possible to distinguish intrathecal or intraventricular administration of drugs for direct delivery into the cerebrospinal fluid, intranasal administration and administration of biodegradable substances [14].
High concentrations of drugs in the brain can be achieved by intrathecal administration. Compared to systemic administration, this method allows the use of low doses of drugs and does not require changes in the integrity of BBB. However, this is an invasive strategy and can lead to increased intracranial pressure and neurotoxicity [15].
Intranasal delivery is another method of direct transport of a therapeutic agent for treating various diseases and CNS disorders. This method is based on the connection between the olfactory nasal lining mucosa and the circulation of cerebrospinal fluid next to the olfactory bulbs [16]. Intranasal delivery of pharmaceuticals is a non-invasive, safe and promising method of treatment.
One of the promising methods to overcome BBB is the loading of drugs into nanocontainers that are characterized by relatively small size and ability to cross the BBB. Using this method of drug delivery significantly increases substances bioavailability and reduces the side effects.
The mechanisms of NPs transportation through BBB can be classified into tree main types [17]:
  • Transient opening of BBB is induced by a stimulus derived from bioactive components on the surface of nanoparticles, or by a stimulus derived from the “nanoeffects” or “nanotoxicity” of nanoparticles. This opening promotes the diffusion of drug conjugates into brain tissue;
  • The adsorption of nanocarrier conjugates on the surface of capillary endothelium cells facilitates the release of the drug from carriers on the cell surface. This process increases the concentration gradient of the drug and promotes the diffusion of substances into the brain;
  • Transcytosis, endocytosis and exocytosis of nanocarriers by capillary endothelial cells of the brain provides direct penetration of the substances into brain tissue.
Figure 1 describes the main pathways of nanoparticles entering the brain: mechanisms for BBB penetration and transport across the cell membrane.
Basically, there are three main mechanisms of NPs cellular uptake [18]. The first one is pinocytosis, which is divided into macro- and micropinocytosis, depending on the nanocarrier size and the type of targeted cell. Another method of nanotherapeutics uptake is clathrin-mediated endocytosis or receptor-mediated endocytosis [18]. This process is based on the formation of clathrin-coated endocytic vesicles. The third way to load NPs into cells is caveolae-dependent endocytosis with calveolar vesicles generation. This type of endocytosis is commonly found in endothelial and muscle cells [18].
In the brain different types of neural cells, including neurons, astrocytes, oligodendrocytes and microglia can represent a target for drug delivery. Targeting ability of nanotherapeutic agents depends on NPs functionality and sometimes requires the presence of specific molecules. Neuron-specific delivery is rather difficult because neurons have a non-phagocytic nature in comparison to glial cells. In some studies, neuron-specific targeting was achieved by the properties of connecting ability of loaded substances. To enhance the neuronal targeting, active strategies such as utilizing ligands with dopamine [19], and rabies virus glycoprotein peptide (RVG and RDP peptide) [20] have being used. Receptor-mediated endocytosis strategies are commonly used to induce neuron uptake of NPs. For, example, nanocarriers coated with apolipoprotein E can be uptaken by neurons through the agency of the lipoprotein-binding receptors on neurons surface [21].
Astrocytes are also very promising targets for nanomedicine. Commonly NPs can be transported into astrocytes via phagocytosis. Their targeting can affect the neurotransmitters release (e.g., glutamate, γ-aminobutyric acid, glycine and histamine), which is crucial for the maintenance of neuronal excitability [22].
Oligodendrocytes are specialized neural cells, where their role is to myelinate the axons in CNS. Properties of their surface can simplify drug delivery via nanopreparations. In particular, NG-2 chondroitin sulphate proteoglycans are expressed on oligodendrocyte precursor cells, which represent useful targets for specific NPs like poly(lactic-co-glycolic acid (PLGA). Such nanocarriers loaded with leukemia inhibitory factor can increase myelin repair [23].
Targeting of microglia or macrophages is also a promising direction for nanomedicine. It is known, that clathrin-mediated endocytosis serves as the major pathway for microglia to uptake NPs. NPs with a larger surface area demonstrate higher uptake efficiency, because clathrin-mediated endocytosis is carried out by the binding of clathrin receptor with the surface of nanocarriers. This property was manifested by the experiments demonstrated that gold NPs with “urchin-mimicking” geometry were absorbed more effectively by microglial cells compared to spherical and rod-like gold NPs [24]. In addition to clathrin-mediated endocytosis, mannose receptors and macrophage scavenger receptors have also been implicated in playing a role in nanocarriers uptake in non-activated microglia [25,26]. There are two main ways to affect microglia or macrophages with nanotherapy. The first one allows controlling microglia/macrophages population by inhibition of infiltrating monocytes/macrophages. For these reason cytotoxic drugs, such as clodronate, glucocorticoids, doxorubicin and methotrexate are usually applied. For example, it was demonstrated that locally delivered methylprednisolone encapsulated into PLGA nanoparticles dramatically reduced the number of macrophages/reactive microglia [27]. The second way to modulate glial cells functions is the use of control release of cytokines via NPs.
Therefore, various research groups are developing nanocontainer systems for drug delivery. There are several types of using nanoparticles with different advantages and disadvantages. In the following we will consider the types of nanoparticles and their way of penetration through BBB, which is necessary for the treatment of neuropsychiatric disorders.

3. Categories of Nanocarriers

Currently the spectrum of nanoparticles used for targeted drug delivery is quite wide. Such nanocontainers can be divided into three main groups: polymeric nanoparticles (NPs), lipid-based NPs and inorganic NPs (Table 2).

3.1. Polymeric Nanoparticles

This class includes polymer-based nanoparticles and micelles and dendrimers (Table 2).
Nanoparticles based on polylactide acid (PLA) and poly (lactic-co-glycolic acid) (PLGA) developed over several decades have proved to be a promising alternative system for delivering drugs to the brain, primarily due to their high biocompatibility and low toxicity. They have uncomplicated synthesis methods (emulsification, evaporation, solvent replacement method (deposition), solvent diffusion method, etc.) and are metabolized into lactic and glycolic acids using a simple mechanism. So, they are biodegradable, with optimal sizes ranging from 100 to 200 nm. PLGA/PLA nanoparticles are U.S. Food and Drug Administration (FDA) certified and can be used in implants, biodegradable scaffolds and human medicines.
The therapeutic molecules encapsulated in polymer nanoparticles can form the following thermodynamically stable structures, depending on the loading method. Either they are nanospheres where the drug is loaded across the entire polymer matrix or only on its surface, or they are nanocapsules where the active substance is surrounded by a polymer shell.
The surface of such particles is easily modifiable, both with ligands for target delivery and molecules that promote adsorption on target cells. Thus, using the example of the anticonvulsant Clonazepam contained in the PLGA matrix, it was possible to achieve effective delivery of the drug through BBB by conjugating the poloxamer-188 to the surface of the NP [35]. Conclusion in PLGA nanocarriers curcumin solved the problem of its insolubility in water, and conjugation with Tet-1 made it possible to increase nanoparticles capture in the treatment of Alzheimer’s disease (AD) [36].
Polymer nanoparticles penetrate the BBB through endocytosis. They can also avoid the phagocytosis of the reticular endothelial system, thereby increasing the concentration of drugs in the brain [37]. In vitro studies have shown that the use of polymer nanoparticles increases the delivery of drugs to the brain. For example, improved delivery of curcumin (in the treatment of AD) is more effective in reducing oxidative stress, inflammation and plaque load.

3.1.1. Dendrimers

Dendrimers are monodisperse symmetrical macromolecules consisting of a series of branched blocks around the inner nucleus. The spherical structure of dendrimers consists of the core, layers of branched, repeating blocks coming out of the core and functional end groups on the outer layer [17].
In the treatment of brain diseases, polyamidoamine dendrimers are most commonly used among dendrimers. For example, there is a study on the use of encapsulated carbamazepine (an antiepileptic drug) for the treatment of AD [38].

3.1.2. Micelles

Micelles consist of amphiphilic block copolymers. They aggregate in aqueous solutions to form stable spheroidal nanostructures that have a hydrophobic core and hydrophilic surface.
The micelles are promising agents for delivering drugs to the brain. This is made possible by the dissolution of poorly soluble substances in the hydrophobic nucleus area and the ability to conjugate with certain target ligands [17].

3.2. Lipid-Based NPs

This type of nanoparticles includes liposomes, solid lipid nanoparticles, nanoemulsions and exosomes (Table 2).

3.2.1. Liposomes

Liposomes are the family of nanostructures, the main characteristic of which is a spherical shape and lipid bilayer. Optimal sizes range from 100 to 200 nm. The first FDA approved nanodrug in 1995 was Doxil represented a preparation based on liposomes [39]. Over the next 2 decades all types of liposome forms, methods of construction and modifications have been developed [40]. For example, liposomal nanoparticles produced by microfluidics were constructed. These nanocarriers allow increasing the percentage of drug loading and obtaining a balanced size of nanocontainer [41]. Generally, liposomes are characterized by the high biocompatibility, biodegradability, improved bioavailability and stability of therapeutic agents [42]. Moreover, they are able to modify the surfaces for active targeting [43]. Nanoparticles structure allows the incorporation of auxiliary lipids such as dioleylphosphatidylethanolamine (DOPE) or cholesterol. DOPE is able to conjugate with other lipids when exposed to low pH, which promotes increased efficiency of loading into cytosol and favor endosomes construction [44]. Cholesterol provides structural stability. The linkers are sensitive to different biological stimuli and contribute to the release of the drug under certain conditions. For example, modifications of gold nanoshells liposomes mediating by chitosan can provide liver protection [45]. Opsonization and capture by macrophages of the organs of the mononuclear phagocytic system obstructs the delivery of drugs via unmodified liposomes. Modification with polyethylene glycol (PEG) increases nanoparticles circulation time and reduces cell capture [17].
The using of liposomal constructions for BBB crossing has been investigated for decades [43]. Liposomes are modified for targeting transcytosis through CNS endothelium via the liposomal surface functionalization with biologically active ligands, such as peptides, antibodies or small molecules. G-Technology® (FDA approved technology) is another method for transformation of the liposomal surface. It represents a strategy for using glutathione-PEGylated liposomes, which are able to bind with glutathione transporter [46].
Thus, liposomes are promising candidates for BBB penetration and targeting delivery of therapeutic agents.

3.2.2. Solid Lipid Nanoparticles

Unlike liposomes, this type of nanoparticles does not possess a lipid layer, but is characterized by the presence of a solid lipid core matrix that can encapsulate lipophilic molecules. Solid lipid nanoparticles (SLNs) are nanoscale suspensions of biologically compatible lipids such as triglycerides, fatty acids or waxes stabilized by bioactive substances possessing a hydrophile-lipophile balance [17].
Such nanoparticles are also subjected to surface modification, which significantly increases the efficiency of drug transporting through BBB [47]. As containers for drug delivery, they demonstrate a high level of biocompatibility, biodegradability and the ability to transform the surface with target bioactive molecules. However, their hydrophobicity promotes easier clearance by the reticuloendothelial system (RES), which alleviates nanoparticles effectiveness.

3.2.3. Nanoemulsion

Nanoemulsions are heterogeneous dispersions of the composition “oil in water” (O/W) or “water in oil” (W/O), stabilized by surfactants. The diameter of the internal phase in such nanosystems is reduced to a nanometric scale. Due to their properties, nanoemulsions demonstrate the ability to carrier both hydrophobic (O/W) or hydrophilic (W/O) drugs. The surface of these nanoparticles can also be modified with bioactive ligands. Delivery of drugs via nanoemulsions is carried out through receptor-mediated endocytosis of cells.
Compared to other nanocarriers, the preponderance of nanoemulsions for BBB penetration is the ability to use safe oils. In addition, the application of this type of nanoparticles possesses a number of useful biological properties due to the presence of essential omega-3 and omega-6 fatty acids in the oils [17].

3.3. Inorganic NPs

Inorganic nanoparticles represent solid nanosize objects consisting of inorganic materials such as silicon dioxide, carbon nanotubes, metal or metal oxide. This type of nanoparticles is widely used for visualization research, including neuroimaging.
To facilitate penetration through BBB, they are covered with several polymers such as polysaccharides, polyacrylamide, poly(vinyl alcohol), poly(N-vinyl-2-pyrrolidone), PEG and PEG-containing copolymers. The most commonly used covering polymer is PEG. These surface modifications enhance nanoparticles stability, improve water solubility and allow modifying of the particle surface with vector molecules [48]. In addition, such PEGylation of nanocontainers averts the capture of particles by the RES and prevents non-specific interactions with opsonin proteins [49].
Besides PEG covering inorganic nanoparticles are also modified with lactoferrin (Lf). It is interesting that this conjunction demonstrated better efficiency during BBB penetration than PEG because of Lf-receptor-mediated transcytosis of cerebral endothelial cells [50].
In addition to chemical modification, the physical properties of inorganic nanoparticles can also be changed for an increase of penetration through BBB. For example, Fe3O4 nanoparticles demonstrate the ability to be magnetized in the presence of a magnetic field. It has been shown that superparamagnetic iron oxide nanoparticles (SPIONs) contained magnetite (Fe3O4) and maghemite (c-Fe2O3) can cross human brain microvascular endothelial cells via an external magnet pressure. Therefore, this mechanism demonstrates the ability of BBB penetration and targeted drug delivery to the brain to be carried out by the magnetic force-mediated dragging of SPIONs through the BBB. Such property of iron oxide nanoparticles as magnetism can be applied for diagnostics and simultaneous therapy of neurodegenerative diseases [51]. Along with iron nanocarriers gold nanoparticles can also penetrate through BBB and can be administered for therapy of neuropsychiatric diseases [52].
Nevertheless, despite the advantages of these types of nanoparticles, they exhibit potential toxicity due to their nondegradability. For this reason, their application requires further research.

4. Drug Delivery for Depression

Depression is a widespread disorder, affecting almost 350 million people worldwide [53]. Depressive symptoms influence quality of life of patients and aggravate the course of comorbid diseases. Depression pathogenesis affects various neurobiological processes including HPA axis activity, monoamine systems functioning, neurogenesis, neuroinflammation, etc. [54]. Traditional medications for the treatment of depression are antidepressants with different mechanisms of action affecting monoamine neurotransmission in the brain.
Antidepressants are divided into several classes according to its mechanisms [55].
  • Tricyclic antidepressants (TCAs) block the reuptake of monoamine neurotransmitters (primarily serotonin and norepinephrine) presynaptically.
  • Monoamine oxidase inhibitors (MAOIs) amplify concentration of monoamines in the synaptic cleft by the blockage of monoamine oxidase enzymes.
  • Selective serotonin reuptake inhibitors (SSRIs) selectively block the reuptake of serotonin. This group of drugs demonstrates decreased side-effects in comparison to other classes of antidepressants.
  • Serotonin and norepinephrine reuptake inhibitors (SNRIs) produce dual inhibition of serotonin and norepinephrine reuptake.
  • Norepinephrine-dopamine reuptake inhibitors block the action of norepinephrine and dopamine transporters thereby inducing reuptake suppression.
  • Norepinephrine reuptake inhibitors block norepinephrine and epinephrine reuptake via norepinephrine transporter inhibition.
  • Serotonin antagonists and reuptake inhibitors (SARIs) act by antagonizing serotonin receptors such as 5-HT2A and inhibiting the reuptake of serotonin, norepinephrine and/or dopamine. Most of them also act as α1-adrenergic receptor antagonists. The majority of the currently marketed SARIs belong to the phenylpiperazine class of substances.
  • NMDA receptor antagonists deactivate NMDA glutamate receptors. This class of antidepressants characterizes as rapid-action drugs and includes such substances as ketamine and esketamine.
In spite of the variety of therapy methods, depression treatment is still a serious concern among psychiatrists. A major study by the National Institute of Mental Health found that less than 50% of people are completely recovered from symptoms of depression even after taking two different antidepressants. Moreover, many of those who respond to treatment soon become depressed again, even though they continue to take their medication [55].
Effectiveness of antidepressant therapy depends on the concentration of the drug in the brain. Conventional oral and parenteral therapies are limited because it requires the drug to cross the BBB.
Antidepressants in dosage forms that manageably release the drug directly into the brain reduce their systemic effects, side effects, drug interactions and ultimately bypass pharmacoresistant problems.
At present, quite a few research groups are busy studying the pharmacokinetics, pharmacodynamics and efficacy of antidepressant nanodrugs. Their research are at various stages: from nanoproducts design to preclinical and clinical stages of testing. So far, none of the antidepressants with targeted delivery to the brain have passed the FDA.
Literature analysis suggests that the intranasal administration of nanotherapy (where the drug is encapsulated in nanoparticles) is the most advantageous way to deliver a whole set of antidepressants of different classes to the brain [56,57]. Moreover, this is consistent with the assumption that a faster start of action accompanied by prolonged circulation of active molecules increases efficiency. Polymer nanoparticles are the most frequently investigated among the various nanosystems (Table 3).

4.1. TCAs

Doxepin is one of the tricyclic antidepressants used for depression treatment. During the last few years doxepin was also investigated as a nanoform. In particular, doxepin in the form of thermoreversible biogel or chitosan and glycerophosphate or polyethylenenglycol demonstrated lower local toxicity and significant effectiveness in the forced swim test [113].

4.2. MAOIs

Antidepressants from MAOIs class have been also investigated for targeting drug delivery. For example, a selective inhibitor of MAO type B—Selegiline was loaded into nanoparticles for depression treatment. It was demonstrated that thiolated chitosan nanoparticles developed by Singh et al. (2016) enhanced the nasal delivery of selegiline and produced antidepressant-like effect in rats [69].

4.3. SSRIs

SSRIs have also being studied as potential nanopreparations. Paroxetine as nanoemulsion form [114] and fluoxetine as ion sensitive in situ nasal gel [115], showed significant responses in locomotor activity level and immobility in injected rats compared to control groups.
Citalopram is also one of the SSRIs. It belongs to the few drugs that can be used safely in childhood psychiatric disorders treatment. Along with other SSRIs Citalopram was loaded in interpenetrating polyelectrolytes nanocomplexes (composed of chitosan: pectin in a 3:1 ratio) constructed by Kamel and colleagues [116]. During experiment researchers demonstrated extended drug release pattern, higher serotonin brain level up to 24 h, more rapid onset of action and more extended behavioral effects in rats with depressive-like behavior treated with these complexes compared to pure citalopram.
Sertraline is another SSRI drug characterized by poor water-solubility and low oral bioavailability. Sertraline-loaded solid lipid nanoparticles improved the drug releasing. In particular, the maximum plasma concentration after nanoparticles injection was 10-fold higher compared to pure drug. Moreover, Sertraline-loaded SLN were found to be stable at 4 °C for 6 months of the study period [117].

4.4. SNRIs

One of the promising antidepressants for targeting delivery is venflaksin hydrochloride (SNRI). There are several ways to deliver it into the brain, for example by Lutrol F127 [118], nanoparticle-based NLC gel [56], alginate–chitosan nanoparticles (AGNPs) [119]. It was shown that nanoparticles loaded with venlafaxine hydrochloride produced a more pronounced effect after intranasal administration in vivo in comparison to oral administration of equivalent dose [118,119]. Desvenlafaxine is another preparation that belongs to that class of drugs. It represents an active metabolite of venlafaxine also investigated in PLGA-chitosan nanoparticles [120].
Duloxetine (SNRI) is another antidepressant that demonstrated better response in NLC [47]. The scintigraphy images are consistent with the biodistribution and pharmacokinetic studies, having revealed a high uptake of duloxetine into the rat brain [47].

4.5. SARIs

Serotonin antagonists and reuptake inhibitors are drugs characterized by antidepressant, anxiolytics and hypnotics effects. This class of substances includes trazodone and nefazodone. These drugs were also loaded into nanoform. In particular, Casolaro et al. (2015) constructed hydrogels bearing l-phenylalanine (Phe-Nip3) or l-valine (Ava2) residues, with citalopram and trazodone. Hydrogel form showed improved release of these drugs compared to pure substances [121].

4.6. Other Drugs

Agomelatine is atypical antidepressant that both stimulates melatonin receptors (MT1 and MT2) and blocks 5-HT2C-receptors. This drug is characterized by low side-effects because it does not counteract with other serotonin receptors, does not affect reuptake of monoamines and does not interact with adrenergic, cholinergic, dopamine, histamine and GABA receptors. Nevertheless, agomelatine has a low biological half-life coupled with extensive hepatic first-pass metabolism [59]. In that context the loading of preparation into nanoform could improve its therapeutic effect. Thus, it was demonstrated that intranasal administration of Agomelatine-loaded poly-lactic-co-glycolic acid nanoparticles shown significant antidepressant-like effect in rats [58]. Another type of nanoparticles was developed by Shinde et al. (2020) and represented polymeric nanoparticles with Agomelatine for transdermal use. These nanocarriers demonstrated good permeability across rat skin [59].
Another type of irregular antidepressant is flavonoid Baicalein. Chen and colleagues constructed solid lipid NPs loaded with Baicalein [60]. These NPs were modified with N-Acetyl Pro-Gly-Pro (PGP) peptide, which is characterized by high specific binding affinity to neutrophils through the CXCR2 receptor. When associated with neutrophils, these particles can penetrate through the BBB and reach targets in the brain.
Along with research of classical antidepressants in nanoforms, it is quite interesting to study various peptides and neurotrophins as antidepressant treatment. In this regard, challenging data concern the proven neuroprotective and antidepressant properties of the brain-derived neurotrophic factor (BDNF). In vitro experiments showed that BDNF delivery using the adenoviral vector produce neuroprotective effects. AAV-Syn-BDNF-EGFP virus vector maintains brain cell viability and preserves the neural network functional structure in the late posthypoxic period [122].
It was found that intranasal administration of BDNF-HA2TAT/AAV to normal mice produce an antidepressant effect in the forced swimming test, but only with chronic treatment. No statistical difference was found in the forced swim test after acute intranasal injection of the AAV to the mice. After the BDNF-HA2TAT/AAV administration to the animals with the chronic mild stress model of depression the successful therapeutic effect was demonstrated in the tail suspension test and forced swim test, but not in open field test [123]. Moreover, the female mice were more sensitive to the drug administration. One of the possible explanations is that the prevalence of depression in females is much higher than that in males [124].
Therefore, different nanocontainer forms of drugs have been developed for the treatment of depression. Such systems contain not only the most popular antidepressants, such as SNRI and SSRI, but also various original forms including vector systems for target delivery in the brain. Application of nanocarriers could help to reduce side effects and to improve drugs bioavailability.

5. Drug Delivery for Anxiety Disorders

Anxiety disorders are the largest group of mental disorders in most Western societies. Essential features of anxiety disorders are excessive and persistent fear, avoidance of perceived threats and panic attacks in certain instances. Although the pathophysiology of anxiety disorders is not fully understood, targeting systems for their treatment are of interest. Drugs for treating anxiety disorders include medications that affect serotonergic, adrenergic, glutamatergic, endocannabinoid systems and various neuropeptides [125]. First of all, tranquilizers with anxiolytic, myorelaxing, sedative and anticonvulsant activity are used for the treatment of anxiety disorders. Many anxiolytic drugs are derived from benzodiazepine and affect benzodiazepine receptors (second generation anxiolytics). However, there is also a spectrum of tranquilizers aimed at other therapeutic targets. Examples include such drugs as buspirone, meprobamate, benactin, etc. Significant progress has been made in the development of formulations and psychosocial treatment strategies for anxiety disorders. However, symptoms persist in many patients despite improvements in therapy. Randomized controlled trials of anxiety disorders pharmacotherapy report a response rate of 40–70% and a remission rate of 20–47%. Anxiety disorders are considered as resistant to treatment when there are residual symptoms or when symptoms do not improve at all after some form of therapeutic intervention. The optimal use of available medication and the development of new pharmacological strategies and agents promise to improve this situation [126].
Polymer nanocarriers are used to develop delivery forms for therapeutic drugs against anxiety disorders. Nanoparticles created from biodegradable polymers such as PLGA are widely studied as drug delivery systems because of their two important properties, such as biocompatibility and controlled drug release. For example, replicable spherical diazepam PLGA nanoparticles have been developed and characterized by Bohrey and colleagues [81]. Chitosan is another natural carbohydrate polymer with ideal for nanoparticles polymer carrier properties such as biocompatibility, biodegradability, non-toxicity and low cost. It exhibits bioadhesive properties and the ability to significantly increase permeability for hydrophilic compounds. Chitosan with thiol groups provides much higher adhesion properties.
Buspirone is a drug belonging to the group of tranquilizers using for anxiety disorders treatment. It is a complete agonist of 5-HT1A presynaptic serotonin receptors, which leads to a decrease of serotonin release and the resulting reduction of anxiety symptoms. Buspirone also acts as a partial agonist of the post-synaptic serotonin receptors, which contribute to the release of serotonin in synapses, resulting in increased serotonin concentration and reduced depression. Additionally, buspirone inhibits the dopamine receptors D2 [127]. Bari et al. (2015) loaded buspirone hydrochloride into polymeric thiolated chitosan nanoparticles for intranasal administration. Ex vivo study carried out via the porcine mucin binding efficiency test demonstrated excellent bioadhesion. Moreover, the drug concentration in the brain after the intranasal injection was significantly higher than after pure Buspirone hydrochloride administration [79]. Along with chitosan nanoparticles construction two types of nanovesicular in situ gels (based on carbopol 974P and poloxamer 407) were developed for delivery of buspirone hydrochloride. The carbopol composition showed higher permeability compared to the poloxamer. It was demonstrated that buspirone hydrochloride has low bioavailability of about 4% after oral administration due to first pass metabolism. Nevertheless, the bioavailability of buspirone hydrochloride increased by 3.26 times when used in the composition of the nanovesicular gel of carbopol compared to the nasal solution of a pure drug [80].
Gallic acid has significant anxiolytic activity, which may be due to its antioxidant properties and reduced nitrite levels in plasma. Chitosan nanoparticles coated with Tween 80 ligand were used to deliver gallic acid to the brain. Behavioral studies have shown that GA nanoparticles significantly improve anxiolytic activity in mice. Nitrite levels decrease in plasma was most pronounced in the group that received the nanoparticle with loaded with drugs. This may have been due to the reduction in plasma nitrite levels by gallic acid. This effect was improved by the increase of GA brain absorption through the nanoparticles application [82].
Inorganic nanoparticles are also used to create delivery systems for anxiety disorders. Many peptides and their analogues cannot effectively cross the BBB. For that reason, Vinzant and colleagues conjugated peptide antisauvagine-30 (ASV-30) with iron oxide nanoparticles [78]. It was shown that ASV-30 peptide affects corticotrophin releasing factor type 2 receptors when directly injected into the brain and thus reduces anxious behavior in animals. Imaging studies demonstrated that iron oxide + ASV-30 particles were present in the brain and were associated with neurons, including those that express CRF2 receptors after intraperitoneal injection. Brain distribution studies were affirmed nanoparticles anxiolytic activity in rats [78].
Currently, the range of application of anxiolytic drugs and their assortment are quite wide. These drugs include preparations of benzodiazepine-line and non-benzodiazepine anxiolytics. Nevertheless, not many nanocontainer forms have been developed for the standard medicines using for anxiety disorders treatment. The loading of other types of drugs in nanocarriers besides the usual forms of anxiolytics could help to improve the treatment of anxiety disorders and to prevent aggravation into depression or severe generalized form.

6. Drug Delivery for Schizophrenia and Bipolar Disorders

Schizophrenia is a complex endogenous mental disorder characterized by cognitive and emotional symptoms that significantly reduce the quality of life of patients [128]. The main drugs used to treat the pathology are antipsychotic drugs that reduce the increased dopaminergic transmission in the brain of patients. Atypical antipsychotics of a new generation, characterized by a narrower range of side effects compared to the first generation, are also used to treat bipolar disorder, along with lithium and antiepileptic drugs [129].
The first drug for the treatment of schizophrenia, chlorpromazine, was first used in 1952, but already in the 60s researchers tried to develop long-acting substances consisting of an antipsychotic agent and fat-soluble solution [130].
Despite the fact that new antipsychotic drugs have appeared on the market, providing greater efficiency with significantly reduced side effects, delivery systems of antipsychotic drugs are still in the stage of conventional drug forms, such as tablets, capsules and solutions, and need to be dosed 2–4 times a day. There is no doubt that new drug delivery systems, such as stable and controlled release systems, will be useful for antipsychotics. They should reduce the frequency of dosing, increase bioavailability of the drug and improve adherence of the therapy course [131].
The drug delivery system produces less pronounced side effects than the original drugs, because they provide lower inpatient concentrations of therapeutic drugs and more stable release characteristics than oral drugs. Thus, such combined drugs increase exposure to oral medications, further reducing morbidity and mortality [132]. As nanocarriers bypass the gastrointestinal tract, such a type of therapy reduces the amount of medication needed and can minimize some peripheral side effects, including hepatotoxicity and hyperprolactinemia [130].
Nanocontainer systems based on PLA were developed in 1989 for chlorpromazine—the first antipsychotic agent [131].
Since then atypical neuroleptics have appeared and become widespread in medical practice. A lot of publications about nanocontainers involve drugs with olanzapine—a second-generation or atypical antipsychotic, which selectively binds to central dopamine D2 and serotonin (5-HT(2c)) receptors. It has poor bioavailability due to hepatic first-pass metabolism and low permeability into the brain due to efflux by P-glycoproteins [133]. Due to its low oral bioavailability, researchers have created various container preparations [134]. Joseph et al. 2017 conducted research on glyceryl monostearate nanoparticles with olanzapine [133]. A longer release of the drug within 48 hours was shown, compared with 8 hours of the antipsychotic action of pure olanzapine solution. Pervaiz and colleagues described the release of olanzapine loaded in PNA (poly(N-isopropylacrylamide-co-acrylic acid)) microgels for 72 h [135]. In vitro bioaccessibility research of Jawahar and colleagues indicated 5½-fold increase in oral bioavailability for nano structured lipid carriers with olanzapine [134]. Cyclodextrins containers were also used for this neuroleptic [136]. An in vitro study on kinetics of clozapine solid lipid nanoparticles showed an increase in lymphatic bioavailability of the drug and its steady release up to 48 h [137]. The study done by Shafaat and coworkers demonstrated that clozapine nanoemulsions resulted in 4.4-fold increase in bioavailability and better drug loading capacity of hydrophobic drug molecules compared with marketed drug formulation [138]. There are also studies on the intranasal introduction of nanoparticles with olanzapine [94]. Nanoparticles containing olanzapine (PLGA) have been shown to be more effectively delivered to the brain by intranasal administration of poly(lactic-co-glycolic acid) nanoparticles [97]. Olanzapine-SLNs showed a 23-fold increase in the relative bioavailability of olanzapine in the brain after nasal administration [92]. The innovative therapeutic approach represents a new form of pharmaceuticals—transdermal drug delivery systems—“patches” containing lipid nanocontainers with this neuroleptic [139]. In addition to improving bioavailability, a study [140] with biodegradable polymeric lipid-core nanocapsules and a decrease in extrapyramidal side effects were shown for polymeric nanoparticles polycaprolactone [141]. Animal model studies have demonstrated improved therapeutic efficacy compared to pure olanzapine [141].
Lurasidone (another atypical antipsychotic) was loaded into SLNs. Its improved therapeutic effect in MK-801 induced schizophrenia model in rats after oral administration was shown [89].
Aripiprazole, a second-generation (or atypical antipsychotic), is poorly soluble and undergoes extensive hepatic metabolism and P-glycoprotein efflux, which lead to reduced in vivo efficacy and increased dose-related side effects. Solid lipid nanoparticles with aripiprazole enhance bioaccessibility of this neuroleptic [83]. In another study, intranasal administration of poly-caprolactone nanoparticles showed a 2-fold increase in aripiprazole distribution in rat brain when compared to aripiprazole nanoparticles administered intravenously [84]. An interesting study was devoted to buccoadhesive chitosan films with aripiprazole. Ex vivo nanocrystal loaded buccal films FAPZ 14 have the potential to provide a faster availability of aripiprazole. The aim of the present study was to design a mucoadhesive dosage form for buccal delivery of aripiprazole, which could provide a rapid drug delivery to the systemic circulation [142].
The same mucoadhesive drug delivery systems were created for riperidone, another atypical neuroleptic used to treat schizophrenia and bipolar disorder. A mucoadhesive buccal tablet formulation of risperidone allowed 90% release of the drug over 8 h, thereby providing direct systemic delivery of risperidone [143]. Risperidone-loaded solid lipid nanoparticles drug release and transport studies illustrated the advantages of oral delivery over poorly water-soluble drugs such as risperidone [144,145]. Pharmacokinetic studies of NLCs of iloperidone 2-fold increase in the % drug targeting efficiency relative to iloperidone pure drug suspension [146]. The pharmacokinetic study revealed almost 8.30-fold increase in oral bioavailability of iloperidone NLCs as compared to the plain drug solution. The nanoemulsions of riperidone also showed improved bioavailability and brain uptake [147]. Intranasal administration of risperidone-loaded mucoadhesive nanoemulsion demonstrated higher efficacy, enhanced brain targeting and higher drug transport to the brain in Swiss albino rats, when compared with intranasal risperidone solution and risperidone nanoemulsion intravenously [105]. In another study, riperidone was enclosed in biodegradable proteinoids. These systems were synthesized by thermal step-growth polymerization from the amino acids l-glutamic acid, l-phenylalanine and l-histidine and poly (l-lactic acid). Behavioral studies on mice found enhanced antipsychotic activity compared to free risperidone.
Quetiapine fumarate, a second generation atypical antipsychotic drug has a plasma half-life of 6h, hence requires frequent administration to maintain effective therapeutic concentration. The increase in bioavailability of mucoadhesive microemulsion Quetiapine during intranasal injection is shown [111].
Azenapine appeared on market in 2009. It was a new atypical antipsychotic for the treatment of schizophrenia and bipolar disorder. However, it had the same disadvantage—low bioavailability when used orally (less than 2%). Most of this drug is metabolized in the liver. The antipsychotic activity of azenapine is thought to be due to antagonistic activity on D2 dopamine and 5-HT2A serotonin receptors. To improve its bioavailability some of the researchers developed sublingual film, intranasal and injectable formulations of azenapine [86]. Shreya et al. (2016) used transfersomes, deformable elastic liposomes loaded with azenapine and transdermal plasters, as a drug formulation. Transfersomes as a drug carrier showed increasing the transdermal permeation and bioavailability of azenapine during in vitro and in vivo experiments.
Lithium medications—psychotropic drugs from the group of normotimists—are historically the first drugs of this group, discovered in 1949. However, they remain essential in the treatment of affective disorders, primarily manic and hypomaniacal phases of bipolar disorder [148]. They are also used to prevent its exacerbation and to treat severe and resistant depression. Their important property is to get rid of suicidal thoughts, so they can prevent suicide.
Lithium ions have a variety of effects on the nervous system, particularly as an antagonist of sodium ions in nerve and muscle cells. In this way, they weaken the nerve impulse (this also explains one of the frequent side effects of lithium preparations: muscle weakness). Lithium also affects the metabolism and transport of monoamines (noradrenaline and serotonin) and increases the sensitivity of some areas of the brain to dopamine. According to some data, the main role in the mechanism of action of lithium is its ability to block the activity of enzymes involved in the synthesis of inositol, which plays a role in regulating the sensitivity of neurons.
The treatment with lithium bipolar disorders is based on its property to selectively inhibit kinase 3 glycogen synthase activity [149]. In addition, it was found that in mania, there is an irregular increase in the activity of proteinkinase C, and a recent study has shown that lithium inhibits its activity. Thus, lithium, like other PkC inhibitors, exhibits antimaniac properties.
Formulations with lithium are the standard drugs for the treatment of bipolar disorders [148]. In one work, the researchers concluded lithium carbonate in chitosan nanoparticles to improve bioavailability. It is postulated that the encapsulation of lithium carbonate within chitosan nanoparticles enables the drug to overcome the intracellular degradative blocks such as endosomes and lysosomes. The ion chelation route was used for encapsulating lithium carbonate [88].
Neuroleptics therapy frequently used for schizophrenia and bipolar disorder treatment requires constant use of drugs and thereby induces various side effects. The loading of drugs into nanocontainers for the controlled release is modern and relevant approach. In addition to targeted delivery and increased bioavailability, it reduces the doses and side effects of antipsychotics, often accompanied its application.

7. Drug Delivery for Alzheimer’s Disease

Currently, the most intensive development of drugs based on nanocontainer systems is being carried out to treat AD. The progress of modern medicine has significantly increased average life expectancy. At the same time, the risks of neurodegenerative diseases rise with age. For example, the percent of people suffering from the most frequent neurodegenerative pathology, AD, in the ages from 65 to 74 is 3% and in the ages from 75 to 84 is already 17%. AD is a progressive disorder characterized by memory, cognition and behavioral impairment (dementia), which eventually leads to mood fluctuation and fatal delirium [150].
The action of the existing drug groups is directed on the pathogenesis blocks presumed as the hypotheses of AD pathogenesis.
Hypotheses of AD development can be divided into several main causes [32,151].
  • Accumulation of β-amyloid (Aβ(1 → 40) and Aβ(1 → 42)), which is resulted from pathologic proteolysis of amyloid precursor protein (APP). The aggregates of this protein are toxic and induce neurodegeneration, cytotoxicity and inevitably leading to dementia manifestations.
  • Tau accumulation. It is known that a normal mature neuron produces three microtubule-associated protein (MAP) taus; MAP1A, MAP1B and MAP2. These proteins are responsible for the promotion of the assembly and stability of microtubules. The biological activity of tau is regulated by its phosphorylation level. For example, in the brain of AD patients, tau hyperphosphorylates abnormally, which impairs its binding to microtubules; leading to the accumulation of neurofibrillary tangles and dementia development. Several anti-tau therapies including lithium, valproate and nicotinamide have been already investigated for the prevention of Tau protein hyperphosphorylation.
  • Progressive decline of acetylcholine concentration. This is the earliest hypothesis of AD progression. It was demonstrated that AD is accompanied by substantial presynaptic cholinergic deficit, which expresses as a reduced choline uptake and acetylcholine release in the brain. That leads to memory and cognition impairment. Drugs that eliminate these symptoms include acetylcholinesterase inhibitors (AChEs) such as tacrine, donepezil, rivastigmine and galantamine.
  • NMDA excitotoxicity. Some experimental and neuropathological evidence suggests that glutamatergic system is also involved in the neurodegenerative progression. NMDA receptor is essential for the control of synaptic plasticity and memory function. It is activated by glutamate and glycine, allowing Ca2+ and Na+ influx. It was demonstrated that the hyperexcitability of NMDA receptors induces Ca2+ overload, eventually leading to apoptosis and producing cognitive impairments. The drug that affects this pathogenesis block is memantine (NMDA inhibitor).
  • Abnormality of mitochondria energy metabolism. It is suggested that genetic mutations alter the regulation of the electron transport chain complex enzymes, which are capable of generating ROS. That leads to cell apoptosis and neurodegeneration.
Figure 2 describes the main pathological pathway of Alzheimer’s disease and therapy methods, targeted to these mechanisms.
Currently, nanoparticles with various therapeutic mechanism associated with AD pathogenesis have been developed (Table 4).

7.1. Nanoparticles with High Affinity to Aβ1–42 Peptide

This type of nanocarriers represents a completely new approach to treating AD. Such a method includes application of multifunctional nanoparticles modified by antitransferrin antibodies in combination with MAbs-IgG domains and transferrin ligands for delivery of the iAβ5 peptide, inhibiting the aggregates associated with AD [228]. Another strategy involves monoclonal antibodies (MAbs) targeting Aβ fibril formation [164]. For example, Kuo with coauthors investigated various modifications of gold nanoparticle systems loaded with the neuron growth factor in combination with transferrin, lactoferrin, p-aminophenyl-alpha-d-mannopyranoside and apolipoprotein E and added peptide sequence [218,219,220]. This sequence interacts with the transferrin receptor present in the microvascular endothelial cells of the blood–brain barrier, thus causing an increase in the permeability of the conjugate in brain, as shown by experiments in vitro and in vivo. It is highly relevant for the therapeutic applications of gold nanoparticles for molecular surgery in the treatment of neurodegenerative diseases such as AD. By irradiation with weak microwaves, Ab aggregates bound to the conjugate peptide sequence with a gold nanoparticle were destroyed by local dissipation of the absorbed energy in vitro [165]. Along with gold nanoparticles PLGA nanoparticles with surface functionalized by antitransferrin receptor monoclonal antibody (OX26) and anti-Aβ (DE2B4) have also been developed for the delivery of encapsulated iAβ5 into the brain. This system efficacy and toxicity were evaluated with porcine brain capillary endothelial cells (PBCECs) used as a BBB model [158].
Another form of nanocarriers [162] presents the construction of Aβ-targeted stealth liposomal nanoparticles using the Aβ-targeted lipid conjugate DSPE-PEG with Methoxy-XO4, which interconnect with amyloid. It was demonstrated that these nanoparticles selectively bind to amyloid plaques in the brain tissue of APP/PSEN1 transgenic mice. Ex vivo analyses of brain samples showed that injected nanoparticles efficiently bind both parenchymal plaques and cerebral amyloid angiopathy associated amyloid throughout the brain [162]. Along with other nanocarriers PEG coated and anti-Ab antibody-conjugated antioxidant (cerium oxide) nanoparticles were developed, and their effects on neuronal survival and BDNF signaling pathway were examined. For this nanoparticles protective effect against oxidative stress and Ab-mediated neurodegeneration were reported. In particular, these nanoforms specifically targets the Ab aggregates, produce an antioxidant effect and thereby perform neuroprotection [168].
β-sheet breaker peptides compose a class of compounds that are highly potent in Aβ1−42- or α-synuclein-inflicted cell toxicity neutralization. β-sheet breaker peptides are highly effective for Aβ amyloidogenesis inhibition as they are homologous to regions of the β-sheet hydrophobic carboxyl segments [229]. These peptides represent oligopeptides structurally analogous to specific epitopes of Aβ. Such substances have been also loaded in nanocarriers for Aβ-targeted delivery. It was demonstrated that the use of gold nanoparticles loaded with LPFFD peptide induces Aβ-aggregate inhibition, Aβ fibrils dissociation and Aβ-mediated peroxidase reduction [166]. TGN TGNYKALHPHNG and QSH D-enantiomeric peptide QSHYRHISPAQV were also conjugated to the surface of the nanoparticles for blood–brain barrier transport and Aβ42 targeting. Nanoparticles implication enhanced bioavailability of drug in the brain and produced neuroprotective effects in the mouse model of AD in vivo [159]. In another work [230] researchers developed a dual-functional nanoparticle drug delivery system loaded with β-sheet breaker peptide H102 (HKQLPFFEED). This peptide was encapsulated into liposomes to reduce the degradation level and was administrated nasally to increase brain penetration. It was demonstrated that the nanosystem application recovered spatial memory impairment in AD model rats both in a low dose and in a high dose [230].
Liposomes are frequently used for drug delivery during targeted AD therapy. For example, liposome nanoparticles are loaded with Aβ1–42 ligands with high affinity, phosphatidic acid and cardiolipin phospholipids [160]. Another research demonstrates the induction of amyloid degradation via administration of apolipoprotein E3-reconstituted high-density lipoprotein nanoparticles bond with Aβ oligomers [163]. Therapeutic effect was also reported for bifunctional liposomes, loaded with phosphatidic acid and peptide derived from apolipoprotein-E receptor-binding domain. In particular, peptide implication allowed penetrating through BBB and phosphatidic acid use caused Aβ aggregates degradation in APP/PS1 transgenic mice [161]. Thus, liposome administration reduced amyloid concentration and improved cognitive function in transgenic mice [161].
Nanoparticles using Congo Red/Rutin/MNP are another method of AD treatment. It was shown that the intranasal administration of Congo Red/Rutina-MNP contributed to the dissipation of amyloid plaques in MRI, increased drug delivery and reduced the production of active oxygen forms and Aβ-induced cytotoxicity in vitro. The intravenous administration of such nanosystems also increased the detection of amyloid aggregates, prevented memory deficit and neurological disorders in transgenic APPswe/PS1dE9 mice in vivo [231].
Multifunctional nanoparticles are being also used for diagnostic and therapy of other pathologies associated with Aβ accumulation. One of them is cerebral amyloid angiopathy described as a condition resulted from Aβ protein lodgment within the walls of cerebral vascular cells. The peptide accumulation leads to vascular inflammation and periodic hemorrhagic strokes. Therapeutic nano-transportable vehicles containing a polymeric core of chitosan-conjugated magnetivist (magnetic resonance imaging (MRI) contrast agent) have been developed for the treatment and early diagnosis cerebral amyloid angiopathy [157]. Cyclophosphamine, an agent that suppresses inflammation of brain vessels, was also loaded into the THB nanonucleus, and a Putreskin-modified F(ab’)2 fragment of an antiamyloid antibody, IgG4.1 (pF(ab’)24.1), was bound to the surface of the nanonucleus to target the cerebrovascular amyloid. Similar systems have already been tested in vitro with human endothelial cells (hCMEC/D3) and in vivo in laboratory mice [157]. It has been shown that other anti-inflammatory and antiamyloidogenic agents capable of reducing the inflammation of cerebrovascular vessels can also be loaded into such nanotransport systems [212].

7.2. Drugs Directed on the Inhibition of Tau Aggregation

Tau represents an intracellular protein, playing a crucial role in the microtubules stabilization. It was demonstrated that tau protein could respond to various factors via aggregates formation. The accumulation of these aggregates induces neurofibrillary tangles conglomeration in neurons and neurodegeneration. As the peptide contains two hexapeptide motifs in the repeat region, after pathological impact tau can randomly coil into the β-sheet structure and perform aggregation.
By now serious research has been done in the field of protein aggregation prevention in neurodegenerative diseases. In spite of investigation of various drugs from different classes like peptide inhibitors, natural substances, etc., little success was achieved. Some research has been performed in the area of nanoparticles, which offers a challenge in AD treatment improvement.
For example, a tau aggregation suppressor phenothiazine was loaded in polymer-based (PLGA-PEG) hydrophobic–glutathione coated nanoparticles [170]. Moreover, for some types of nanocarriers aggregation inhibitory activity were demonstrated of its own accord. The cadmium sulfide and iron oxide (Fe3O4) nanoparticles inhibit tau fibrillization in vitro. Despite its antiaggregation activity and excellent delivery properties, the in vivo experiments are hampered by interfering molecular species [171].
Red blood cell membrane-coated PLGA particles with T807 can also be used as a resource for tau aggregation inhibition. As T807 demonstrates high permeability for neurons and pronounced p-tau-specific tracing activity, it can be considered as a therapeutic agent of targeted delivery for AD treatment. One more virtue of that type of nanocarriers is red blood cell membranes, characterized by low immunogenicity and long half-life in the circulation, which allow performing effective drug delivery. Synergetic action of T807 with triphenylphosphine (TPP) produces an inhibitory effect on key components of AD pathogenesis associated with tau-phosphorylation and prevents neurodegeneration. Thus, the systemic administration red blood cell PLGA nanoparticles loaded with curcumin recover cognitive impairments in the mouse model of AD [232].

7.3. Acetylcholinesterase Inhibitors (AChE)

This class of drugs is characterized by many side effects due to their action on the peripheral tissues, high doses of administration and hepatic metabolism. These side effects are aggravated by the low pharmacokinetic (half-life) profiles of the drugs [150]. The loading of such preparation into nanocontainers would help to eliminate these side effects and to improve their bioavailability to the brain via an intranasal form.
One of the drugs used to treat AD is rivastigmine, which inhibits both the AChE and butyrylcholinesterase enzymes in brain and boosts the cholinergic function. It improves cognition and memory by increasing the ACh concentration in the brain. Nevertheless, the half-life period of rivastigmine in blood circulation is only 1.5 h, which complicates its therapeutic action. Loading drugs in nanocontainers increases its half-life and decreases the necessary dosage, thereby reducing side effects. For example, it has been already demonstrated for AChE inhibitors represented as naturally derived polyphenols, such as epigallocatechin-3-gallate and catechin placed into the sodium-taurocholate liposomal carrier [192], cholesterol/soy-lecithinated liposomal NP formulation [190], PLGA, PEG and gold NPs [233]. Yang and colleagues have also performed analysis of rivastigmine-liposome and CPP27 modified rivastigmine-loaded liposome for intranasal administration and showed the improvement of drug activity [191]. In particular, it was demonstrated that CPP modification increases the drug diffusion across the BBB by enhanced transcytosis and cell membrane permeability. It was acknowledged by the pharmacokinetic study indicated a significant increase of drug concentration in the various regions of the brain like cortex, hippocampus, and olfactory regional. Additionally, that modification has not exerted any toxic effects in the nasal cavity and in the brain. Improved drug properties were also demonstrated for rivastigmine via various nanoparticles. For example, rivastigmine-loaded mucoadhesive NP developed by Fazil et al., PLGA, PBCA NP by Joshi et al. and chitosan (polysorbate 80-coated) NP by Khemariya et al. showed an increased bioavailability [193,194,195]. In was also shown that gelling cationic nanostructured lipid carriers enhance brain distribution and pharmacodynamics [188]. Moreover, liposomes with RHN increased half-life, decreased the toxicity and improved memory and neurological deficits in Male albino rat with the model of AD [189].
Another AChE inhibitor using for AD treatment is galantamine. Some studies reported that galantamine could also reduce Aβ aggregation and neurotoxicity. Nevertheless, the application of galantamine is limited by its side effects such as nausea, vomiting, dizziness, drowsiness, loss of appetite, etc. That makes it inconvenient for the long-term oral treatment. In addition, galantamine is not able to cross the BBB efficiently [33]. For this reason, NP can improve response and bioavailability of galantamine. For example, Hanafy and colleagues constructed chitosan nanoparticle and cationic nanoparticles for intranasal drug administration [175]. Li et al. developed liposomal system, as a drug carrier for efficient delivery of galantamine through intranasal administration [173]. Other research groups created PLGA [174] and solid lipid NP [172] with galantamine, which have not only increased its bioavailability but also improved cognitive function and memory in Wistar rats.
Donepezil is another frequently used drug for AD treatment, represented as a centrally acting reversible AChE inhibitor. It is characterized by high pharmacokinetic properties with 100% bioavailable and a biological half-life of 70 h. However, the long-term administration leads to severe side effects, especially when taken in a high dose (>10 mg/day) [234]. The development of chitosan nanosuspension for intranasal administration [178] and the chitosan intranasal form of donepezil [235] promoted the usage of lower drug dose. Along with these nanoparticles LGA-b-PEG by Baysal et al. and liposomes by Al Asmari et al. were constructed for this drug [177,180].
Tacrine is AChE inhibitor acting as a histamine N-methyltransferase inhibitor. This drug produces various side effects including not only nausea, vomiting, dizziness, drowsiness, hallucination, confusion, insomnia and anxiety, but also hepatotoxicity, convulsion, urinary tract infection, bradycardia, depression, hypotension, etc. For the problem solution Qian and colleagues have developed in situ gelling systems for intranasal administration of tacrine [185]. That drug form enhanced the substance retention time in the nasal cavity and improved the nasomucosal absorbance. Wilson et al. have constructed a tacrine loaded chitosan nanoparticle coated with polysorbate 80 and demonstrated the improvement of its biodistribution [197]. Luppi et al. has also developed albumin nanoparticles carrying cyclodextrins for nasal delivery and drug properties advance [186].
Huperzine is another preparation used for AD treatment. For improvement of it bioavailability some research groups also constructed NP. For example, Yang et al. created nanostructured lipid carriers [181]. The treatment with huperzine lipid NP performed by Patel et al. demonstrated behavioral and memory improvement in the mouse model of AD [182].

7.4. Memantine, the NMDA Receptors Antagonist

Memantine hydrochloride (MEM) is the only anti-AD drug used in the US and Europe for the moderate-to-severe form of AD. The labeled formulation for memantine hydrochloride is Namenda, which is prescribed in tablet or capsule form. MEM prevents the aggregation of Aβ1–40 and enhances memory in rodents. Despite the best results in improving cognitive impairments in patients, the clinical applicability of MEM is limited. MEM should be administered daily, which leads to poor compliance with the drug regimen [205]. The development of nanocarriers with memantine can help to solve these problems. Toxicologically safe lipid nanoparticles with lipoyl-memantine have been elaborated and can be used to deliver memantine to the brain [206]. Gothwal et al. (2019) developed memantine-loaded polyamidoamine dendrimers. He used lactoferrin as a targeting ligand. The profile of memantine release from nanoparticles was slow and persisted for up to 48 h. The brain absorption of NPs was significantly higher than that of pure memantine. The in vivo research in an AlCl3-induced AD mice model demonstrated a considerable improving in behavioral tests [205].

7.5. Other Drugs for AD Treatment

Many literature data suggest that various growth factors may have therapeutic potential in the AD treatment. In particular, growth factors such as basic fibroblast growth factor, neurotrophins (nerve growth factors; glial-derived neurotrophic factor and BDNF and insulin-like growth factor (IGF-1/IGF-2)) and bone morphogenetic proteins are promising molecules [236]. In such a way PEG-PLGA nanoparticles loaded with basic fibroblast growth factor have been shown to improve cognitive function in the Morris test in rats exposed to intraventricular administration of β-amyloid 25–35 and ionic acid [221].
Metal chelators are another potential drugs to treat AD because they block the effects of metal ions stimulating the formation of Aβ aggregates and active oxygen forms. However, their use is limited due to non-specific interactions with metal ions and the inability to pass through BBB. Such problems have been solved by using controlled release systems of gold nanoparticles with mesoporous silica based on H2O2. Loading metal chelator CQ in these nanocontainers allowed targeted CQ release in the brain areas where the level of H2O2 is increased, for example, in Aβ-plaques [201].
Several recent studies have demonstrated that some antidiabetic drugs also have a therapeutic potential in AD patients. For example, Hsieh and colleagues has shown the inhibitory effect of gold nanoparticles on the fibrillogenesis of insulin fibrils [167]. Gold nanoparticles, coincubated with insulin, retarded the structural transformation of insulin fibrils into amyloid-like fibrils for about a week [167]. The neuroprotective effect was also demonstrated for nanoparticles with peptide NAPVSIPQ as a model drug and lactoferrin as a targeting ligand, which improved cognitive function in mice. As lactoferrin is a natural iron binding protein, whose receptor is highly expressed in both respiratory epithelial cells and neurons, its administration facilitates the nose-to-brain drug delivery of neuroprotection peptides like NAPVSIPQ [200].

7.6. Others Methods of AD Treatment

Currently, siRNA gene therapy is a promising method of treating of many brain disorders. Potentially it can also be used for the treatment of AD. An example of such a therapy method is the inhibition of the Beta-Secretase 1 (BACE1), which leads to suppression of disease development. Nanoparticles of the solid lipid with and without chitosan were used for directed delivery of BACE1 siRNA to the brain. These nanoparticles contained a RVG-9R peptide derived from the rabies virus glycoprotein, which enhanced the transcellular delivery of nanosystems to the brain during their intranasal injection. It was demonstrated that the investigated siRNA permeates the monolayer of the cell culture Caco-2 to a greater extent when released from nanocarriers, primarily from chitosan-coated solid lipid nanoparticles, than from bare siRNA [153].
The same research strategies were developed by other researchers. Wang and colleagues designed siRNA nanocarriers based on PEGylated poly(2-(N,N-dimethylamino) ethyl methacrylate) (PEG-PDMAEMA) modified with both the CGN peptide for BBB penetration and the Tet1 peptide for neuron-specific binding [154]. These nanocomplexes have been shown to successfully enter the cytoplasm of neurons, effectively suppressing the expression of BACE1 mRNA (by about 50%) and preventing synaptic damage caused by Aβ25-35. These results were also confirmed by studies of transgenic APP/PS1 mice, where the administration of nanoparticles stimulated neurogenesis in the hippocampus, inhibited the formation of amyloid plaques and tau-protein by repression of BACE1 mRNA. The administration of the nanocomplexes restored the cognitive function of the AD transgenic mice to the level of wild-type control without significant adverse effects on myelination [154].
Downregulation of the key enzyme in amyloid-β formation by delivering non-coding RNA plasmid is another potential way of AD treatment. It was shown that simultaneous delivery of the therapeutic peptide into the brain leads to the suppression of neurofibrillary tangles formation. In particular, a multifunctional nanocarrier contained the peptide to achieve the therapeutic gene, based on the PEGylated dendrigraphte of polylysines, has been developed for the treatment of AD [155]. Directional nanoparticle delivery has been investigated in transgenic AD mice [155].
Treatment of AD is extremely difficult because no precise mechanisms have been discovered for its development and pathogenesis. Despite that fact, a large number of different drug forms and therapy methods are being investigated. Among others, many nanocontainer systems have been constructed for different groups of drugs, inducing targeted delivery to pathological brain centers. Various clinical data provide an opportunity to hope that these strategies will bring their results.

8. Conclusions

The development of nanocontainer systems for the treatment of various diseases is a modern, promising and actively evolving field of science. Among other things, nanoparticles also contain agents for the treatment of neuropsychiatric diseases. All mental illnesses require long-term therapy, and patients need constant medication for months or even years. For this reason, it is important that the drug is delivered using nanocontainer systems that facilitate a longer release of the medicine. Various nanocontainers are opening up prospects for new forms of medication, such as subcutaneous transdermal drug delivery systems—“patches”. Non-compliance with therapy in patients is a very common and serious problem, which can be solved by prescribing injections of long-acting drugs or implants. Many neuroleptics have low bioavailability when used orally due to liver metabolism and low brain permeability, and drug delivery systems can improve these indicators. Prolonged therapy inevitably leads to side effects. Studies of some combined medications for the treatment of neuropsychiatric diseases have shown a reduction in side effects compared to pure drugs. Thus, the use of nanocontainer systems for the delivery of antidepressants, antipsychotics and drugs for the treatment of neurodegenerative diseases can contribute to more effective therapy and improve patients’ quality of life.

Author Contributions

Original draft preparation, Y.Z., O.A., V.U., A.M. (Anna Morozova); review and editing, E.Z.; review and editing of the part about types of nanocarriers P.M.; review and editing of the part about BBB penetration M.V.; supervision, V.C.; project administration, A.M. (Alexander Majouga); funding acquisition, A.M. (Alexander Majouga). All authors have read and agreed to the published version of the manuscript.

Funding

Work was supported by grant of Ministry of Science and Higher Education of Russian Federation No 075-15-2020-792 (Unique identifier RF----190220X0031).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AChEacetylcholinesterase inhibitors
ADAlzheimer’s disease
BACE1beta-Secretase 1
BBBblood–brain barrier
BDNFbrain-derived neurotrophic factor
DOPEdioleylphosphatidylethanolamine
FDAU.S. Food and Drug Administration
Lflactoferrin
MAOIsmonoamine oxidase inhibitors
MEMmemantine hydrochloride
MRImagnetic resonance imaging
NMDAN-methyl-d-aspartate
NPsnanoparticles
PEGpolyethylene glycol
PLApolylactide acid
PLGApoly lactic-co-glycolic acid
RESreticuloendothelial system
SARIsserotonin antagonists and reuptake inhibitors
SLNssolid lipid nanoparticles
SNRIsserotonin and norepinephrine reuptake inhibitors
SPIONssuperparamagnetic iron oxide nanoparticles
SSRIsselective serotonin reuptake inhibitors
TCAstricyclic antidepressants

References

  1. James, S.L.; Abate, D.; Abate, K.H.; Abay, S.M.; Abbafati, C.; Abbasi, N.; Abbastabar, H.; Abd-Allah, F.; Abdela, J.; Abdelalim, A.; et al. Global, regional, and national incidence, prevalence, and years lived with disability for 354 diseases and injuries for 195 countries and territories, 1990–2017: A systematic analysis for the Global Burden of Disease Study 2017. Lancet 2018, 392, 1789–1858. [Google Scholar] [CrossRef] [Green Version]
  2. Agrahari, V. The exciting potential of nanotherapy in brain-tumor targeted drug delivery approaches. Neural Regen. Res. 2017, 12, 197–200. [Google Scholar] [CrossRef] [PubMed]
  3. Rakotoarisoa, M.; Angelov, B.; Garamus, V.M.; Angelova, A. Curcumin- and fish oil-loaded spongosome and cubosome nanoparticles with neuroprotective potential against H2O2-induced oxidative stress in differentiated human SH-SY5Y cells. ACS Omega 2019, 4, 3061–3073. [Google Scholar] [CrossRef] [Green Version]
  4. Deeken, J.F.; Löscher, W. The blood-brain barrier and cancer: Transporters, treatment, and Trojan horses. Clin. Cancer Res. 2007, 13, 1663–1674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Van Tellingen, O.; Yetkin-Arik, B.; De Gooijer, M.; Wesseling, P.; Wurdinger, T.; De Vries, H.E. Overcoming the blood–brain tumor barrier for effective glioblastoma treatment. Drug Resist. Updates 2015, 19, 1–12. [Google Scholar] [CrossRef] [PubMed]
  6. Villabona-Rueda, A.; Erice, C.; Pardo, C.A.; Stins, M.F. The evolving concept of the blood brain barrier (BBB): From a single static barrier to a heterogeneous and dynamic relay center. Front. Cell. Neurosci. 2019, 13. [Google Scholar] [CrossRef]
  7. Abbott, N.J.; Patabendige, A.A.K.; Dolman, D.E.M.; Yusof, S.R.; Begley, D.J. Structure and function of the blood–brain barrier. Neurobiol. Dis. 2010, 37, 13–25. [Google Scholar] [CrossRef] [PubMed]
  8. Weidle, U.H.; Niewöhner, J.; Tiefenthaler, G. The blood-brain barrier challenge for the treatment of brain cancer, secondary brain metastases, and neurological diseases. Cancer Genom. Proteom. 2015, 12, 167–177. [Google Scholar]
  9. Pardridge, W.M. Drug and gene targeting to the brain with molecular Trojan horses. Nat. Rev. Drug Discov. 2002, 1, 131–139. [Google Scholar] [CrossRef]
  10. Agarwal, S.; Sane, R.; Oberoi, R.; Ohlfest, J.R.; Elmquist, W.F. Delivery of molecularly targeted therapy to malignant glioma, a disease of the whole brain. Expert Rev. Mol. Med. 2011, 13, e17. [Google Scholar] [CrossRef] [Green Version]
  11. Lesniak, W.G.; Chu, C.; Jablonska, A.; Du, Y.; Pomper, M.G.; Walczak, P.; Janowski, M. A Distinct advantage to intraarterial delivery of 89Zr-bevacizumab in PET imaging of mice with and without osmotic opening of the blood–brain barrier. J. Nucl. Med. 2018, 60, 617–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Côté, J.; Bovenzi, V.; Savard, M.; Dubuc, C.; Fortier, A.; Neugebauer, W.; Tremblay, L.; Müller-Esterl, W.; Tsanaclis, A.-M.; Lepage, M.; et al. Induction of selective blood-tumor barrier permeability and macromolecular transport by a biostable kinin B1 receptor agonist in a glioma rat model. PLoS ONE 2012, 7, e37485. [Google Scholar] [CrossRef] [PubMed]
  13. Hendricks, B.K.; Cohen-Gadol, A.A.; Miller, J.C. Novel delivery methods bypassing the blood-brain and blood-tumor barriers. Neurosurg. Focus 2015, 38, E10. [Google Scholar] [CrossRef] [PubMed]
  14. Hersh, D.S.; Wadajkar, A.S.; Roberts, N.B.; Perez, J.G.; Connolly, N.P.; Frenkel, V.; Winkles, J.A.; Woodworth, G.F.; Kim, A.J. Evolving drug delivery strategies to overcome the blood brain barrier. Curr. Pharm. Des. 2016, 22, 1177–1193. [Google Scholar] [CrossRef] [Green Version]
  15. Alkhani, A.M.; Al-Shaar, H.A. Intrathecal baclofen therapy for spasticity: A compliance-based study to indicate effectiveness. Surg. Neurol. Int. 2016, 7, S539–S541. [Google Scholar] [CrossRef] [Green Version]
  16. Lochhead, J.J.; Thorne, R.G. Intranasal delivery of biologics to the central nervous system. Adv. Drug Deliv. Rev. 2012, 64, 614–628. [Google Scholar] [CrossRef]
  17. Li, X.; Tsibouklis, J.; Weng, T.; Zhang, B.; Yin, G.; Feng, G.; Cui, Y.; Savina, I.N.; Mikhalovska, L.; Sandeman, S.R.; et al. Nano carriers for drug transport across the blood–brain barrier. J. Drug Target. 2017, 25, 17–28. [Google Scholar] [CrossRef] [Green Version]
  18. Adjei, I.M.; Sharma, B.; Labhasetwar, V. Nanoparticles: Cellular uptake and cytotoxicity. Atherosclerosis 2014, 811, 73–91. [Google Scholar] [CrossRef]
  19. Zhang, H.; Jiang, Y.; Zhao, S.-G.; Jiang, L.-Q.; Meng, Y.; Liu, P.; Kim, M.O.; Li, S. Selective neuronal targeting, protection and signaling network analysis via dopamine-mediated mesoporous silica nanoparticles. MedChemComm 2015, 6, 1117–1129. [Google Scholar] [CrossRef]
  20. Alvarez-Erviti, L.; Seow, Y.; Yin, H.; Betts, C.; Lakhal, S.; Wood, M.J. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat. Biotechnol. 2011, 29, 341–345. [Google Scholar] [CrossRef]
  21. Hasadsri, L.; Kreuter, J.; Hattori, H.; Iwasaki, T.; George, J.M. Functional protein delivery into neurons using polymeric nanoparticles. J. Biol. Chem. 2009, 284, 6972–6981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Scuderi, C.; Stecca, C.; Iacomino, A.; Steardo, L. Role of astrocytes in major neurological disorders: The evidence and implications. IUBMB Life 2013, 65, 957–961. [Google Scholar] [CrossRef] [PubMed]
  23. Rittchen, S.; Boyd, A.; Burns, A.; Park, J.; Fahmy, T.M.; Metcalfe, S.; Williams, A. Myelin repair in vivo is increased by targeting oligodendrocyte precursor cells with nanoparticles encapsulating leukaemia inhibitory factor (LIF). Biomaterials 2015, 56, 78–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hutter, E.; Boridy, S.; Labrecque, S.; Lalancette-Hébert, M.; Kriz, J.; Winnik, F.M.; Maysinger, D. Microglial response to gold nanoparticles. ACS Nano 2010, 4, 2595–2606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Patel, S.K.; Janjic, J.M. Macrophage targeted theranostics as personalized nanomedicine strategies for inflammatory diseases. Theranostics 2015, 5, 150–172. [Google Scholar] [CrossRef] [Green Version]
  26. Papa, S.; Rossi, F.; Ferrari, R.; Mariani, A.; De Paola, M.; Caron, I.; Fiordaliso, F.; Bisighini, C.; Sammali, E.; Colombo, C.; et al. Selective nanovector mediated treatment of activated proinflammatory microglia/macrophages in spinal cord injury. ACS Nano 2013, 7, 9881–9895. [Google Scholar] [CrossRef]
  27. Kim, Y.-T.; Caldwell, J.-M.; Bellamkonda, R.V. Nanoparticle-mediated local delivery of methylprednisolone after spinal cord injury. Biomaterials 2009, 30, 2582–2590. [Google Scholar] [CrossRef] [Green Version]
  28. Fan, Y.; Chen, M.; Zhang, J.; Maincent, P.; Xia, X.; Wu, W. Updated progress of nanocarrier-based intranasal drug delivery systems for treatment of brain diseases. Crit. Rev. Ther. Drug Carr. Syst. 2018, 35, 433–467. [Google Scholar] [CrossRef]
  29. Fonseca-Santos, B.; Chorilli, M.; Gremião, M.P.D. Nanotechnology-based drug delivery systems for the treatment of Alzheimer’s disease. Int. J. Nanomed. 2015, 10, 4981–5003. [Google Scholar] [CrossRef] [Green Version]
  30. Wilczewska, A.Z.; Niemirowicz, K.; Markiewicz, K.H.; Car, H. Nanoparticles as drug delivery systems. Pharmacol. Rep. 2012, 64, 1020–1037. [Google Scholar] [CrossRef]
  31. Zhou, Y.; Peng, Z.; Seven, E.S.; Leblanc, R.M. Crossing the blood-brain barrier with nanoparticles. J. Control. Release 2018, 270, 290–303. [Google Scholar] [CrossRef]
  32. Wen, M.M.; El-Salamouni, N.S.; El-Refaie, W.M.; Hazzah, H.A.; Ali, M.M.; Tosi, G.; Farid, R.M.; Blanco-Prieto, M.J.; Billa, N.; Hanafy, A.S. Nanotechnology-based drug delivery systems for Alzheimer’s disease management: Technical, industrial, and clinical challenges. J. Control. Release 2017, 245, 95–107. [Google Scholar] [CrossRef] [PubMed]
  33. Agrawal, M.; Saraf, S.; Saraf, S.K.; Antimisiaris, S.G.; Hamano, N.; Li, S.-D.; Chougule, M.; Shoyele, S.A.; Gupta, U.; Ajazuddin; et al. Recent advancements in the field of nanotechnology for the delivery of anti-Alzheimer drug in the brain region. Expert Opin. Drug Deliv. 2018, 15, 589–617. [Google Scholar] [CrossRef] [PubMed]
  34. Majumder, J.; Taratula, O.; Minko, T. Nanocarrier-based systems for targeted and site specific therapeutic delivery. Adv. Drug Deliv. Rev. 2019, 144, 57–77. [Google Scholar] [CrossRef] [PubMed]
  35. Shah, P.; Chavda, K.; Vyas, B.; Patel, S. Formulation development of linagliptin solid lipid nanoparticles for oral bioavailability enhancement: Role of P-gp inhibition. Drug Deliv. Transl. Res. 2020, 1–20. [Google Scholar] [CrossRef] [PubMed]
  36. Mathew, A.; Fukuda, T.; Nagaoka, Y.; Hasumura, T.; Morimoto, H.; Yoshida, Y.; Maekawa, T.; Venugopal, K.; Kumar, D.S. Curcumin loaded-PLGA nanoparticles conjugated with Tet-1 peptide for potential use in Alzheimer’s disease. PLoS ONE 2012, 7, e32616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. 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] [Green Version]
  38. 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] [Green Version]
  39. Barenholz, Y. Doxil®—The first FDA-approved nano-drug: Lessons learned. J. Control. Release 2012, 160, 117–134. [Google Scholar] [CrossRef]
  40. Yu, B.; Lee, R.J.; Lee, L.J. Microfluidic methods for production of liposomes. Methods Enzymol. 2009, 465, 129–141. [Google Scholar] [CrossRef] [Green Version]
  41. Carugo, D.; Bottaro, E.; Owen, J.; Stride, E.; Nastruzzi, C. Liposome production by microfluidics: Potential and limiting factors. Sci. Rep. 2016, 6, 25876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Deshpande, P.P.; Biswas, S.; Torchilin, V. Current trends in the use of liposomes for tumor targeting. Nanomedicine 2013, 8, 1509–1528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Stenehjem, D.D.; Hartz, A.M.; Bauer, B.; Anderson, G.W. Novel and emerging strategies in drug delivery for overcoming the blood–brain barrier. Futur. Med. Chem. 2009, 1, 1623–1641. [Google Scholar] [CrossRef] [PubMed]
  44. Hattori, Y.; Suzuki, S.; Kawakami, S.; Yamashita, F.; Hashida, M. The role of dioleoylphosphatidylethanolamine (DOPE) in targeted gene delivery with mannosylated cationic liposomes via intravenous route. J. Control. Release 2005, 108, 484–495. [Google Scholar] [CrossRef]
  45. Luo, L.; Bian, Y.; Liu, Y.; Zhang, X.; Wang, M.; Xing, S.; Li, L.; Gao, D. Gold nanoshells: Combined near infrared photothermal therapy and chemotherapy using gold nanoshells coated liposomes to enhance antitumor effect (Small 30/2016). Small 2016, 12, 4102. [Google Scholar] [CrossRef]
  46. Lindqvist, A.; Rip, J.; Van Kregten, J.; Gaillard, P.J.; Hammarlund-Udenaes, M. In vivo functional evaluation of increased brain delivery of the opioid peptide DAMGO by glutathione-PEGylated liposomes. Pharm. Res. 2016, 33, 177–185. [Google Scholar] [CrossRef]
  47. Tapeinos, C.; Battaglini, M.; Ciofani, G. Advances in the design of solid lipid nanoparticles and nanostructured lipid carriers for targeting brain diseases. J. Control. Release 2017, 264, 306–332. [Google Scholar] [CrossRef]
  48. Karakoti, A.S.; Das, S.; Thevuthasan, S.; Seal, S. PEGylated inorganic nanoparticles. Angew. Chem. Int. Ed. 2011, 50, 1980–1994. [Google Scholar] [CrossRef]
  49. Joralemon, M.J.; McRae, S.; Emrick, T. PEGylated polymers for medicine: From conjugation to self-assembled systems. Chem. Commun. 2010, 46, 1377–1393. [Google Scholar] [CrossRef] [Green Version]
  50. Qiao, R.; Jia, Q.; Hüwel, S.; Xia, R.; Liu, T.; Gao, F.; Galla, H.-J.; Gao, M. Receptor-mediated delivery of magnetic nanoparticles across the blood–brain barrier. ACS Nano 2012, 6, 3304–3310. [Google Scholar] [CrossRef]
  51. Luo, S.; Ma, C.; Zhu, M.-Q.; Ju, W.-N.; Yang, Y.; Wang, X. Application of iron oxide nanoparticles in the diagnosis and treatment of neurodegenerative diseases with emphasis on Alzheimer’s disease. Front. Cell. Neurosci. 2020, 14, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Frigell, J.; García, I.; Gómez-Vallejo, V.; Llop, J.; Penadés, S. 68Ga-Labeled gold glyconanoparticles for exploring blood–brain barrier permeability: Preparation, biodistribution studies, and improved brain uptake via neuropeptide conjugation. J. Am. Chem. Soc. 2013, 136, 449–457. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, Q.; Timberlake, M.A., II; Prall, K.; Dwivedi, Y. The recent progress in animal models of depression. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2017, 77, 99–109. [Google Scholar] [CrossRef]
  54. McQuaid, R.J.; McInnis, O.A.; Abizaid, A.; Anisman, H. Making room for oxytocin in understanding depression. Neurosci. Biobehav. Rev. 2014, 45, 305–322. [Google Scholar] [CrossRef]
  55. O’Leary, O.F.; Dinan, T.G.; Cryan, J.F. Faster, better, stronger: Towards new antidepressant therapeutic strategies. Eur. J. Pharmacol. 2015, 753, 32–50. [Google Scholar] [CrossRef] [PubMed]
  56. Cirri, M.; Maestrini, L.; Maestrelli, F.; Mennini, N.; Mura, P.; Ghelardini, C.; Mannelli, L.D.C. Design, characterization and in vivo evaluation of nanostructured lipid carriers (NLC) as a new drug delivery system for hydrochlorothiazide oral administration in pediatric therapy. Drug Deliv. 2018, 25, 1910–1921. [Google Scholar] [CrossRef]
  57. Vitorino, C.; Silva, S.; Bicker, J.; Falcão, A.; Fortuna, A. Antidepressants and nose-to-brain delivery: Drivers, restraints, opportunities and challenges. Drug Discov. Today 2019, 24, 1911–1923. [Google Scholar] [CrossRef]
  58. Jani, P.; Vanza, J.; Pandya, N.; Tandel, H. Formulation of polymeric nanoparticles of antidepressant drug for intranasal delivery. Ther. Deliv. 2019, 10, 683–696. [Google Scholar] [CrossRef]
  59. Shinde, M.; Bali, N.R.; Rathod, S.; Karemore, M.N.; Salve, P. Effect of binary combinations of solvent systems on permeability profiling of pure agomelatine across rat skin: A comparative study with statistically optimized polymeric nanoparticles. Drug Dev. Ind. Pharm. 2020, 46, 826–845. [Google Scholar] [CrossRef]
  60. Chen, B.; Luo, M.; Liang, J.; Zhang, C.; Gao, C.; Wang, J.; Wang, J.; Li, Y.; Xu, D.; Liu, L.; et al. Surface modification of PGP for a neutrophil–nanoparticle co-vehicle to enhance the anti-depressant effect of baicalein. Acta Pharm. Sin. B 2018, 8, 64–73. [Google Scholar] [CrossRef]
  61. Harris, N.M.; Ritzel, R.; Mancini, N.S.; Jiang, Y.; Yi, X.; Manickam, D.S.; Banks, W.A.; Kabanov, A.V.; McCullough, L.D.; Verma, R. Nano-particle delivery of brain derived neurotrophic factor after focal cerebral ischemia reduces tissue injury and enhances behavioral recovery. Pharmacol. Biochem. Behav. 2016, 48–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. He, X.; Zhu, Y.; Wang, M.; Jing, G.; Zhu, R.; Wang, S. Antidepressant effects of curcumin and HU-211 coencapsulated solid lipid nanoparticles against corticosterone-induced cellular and animal models of major depression. Int. J. Nanomed. 2016, 11, 4975–4990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. He, X.; Yang, L.; Wang, M.; Zhuang, X.; Huang, R.; Zhu, R.; Wang, S. Targeting the endocannabinoid/CB1 receptor system for treating major depression through antidepressant activities of curcumin and dexanabinol-loaded solid lipid nanoparticles. Cell Physiol. Biochem. 2017, 42, 2281–2294. [Google Scholar] [CrossRef] [PubMed]
  64. Fidelis, E.M.; Savall, A.S.P.; Abreu, E.D.L.; Carvalho, F.; Teixeira, F.E.G.; Haas, S.E.; Sampaio, T.B.; Pinton, S. Curcumin-loaded nanocapsules reverses the depressant-like behavior and oxidative stress induced by β-amyloid in mice. Neuroscience 2019, 423, 122–130. [Google Scholar] [CrossRef]
  65. Alam, M.I.; Baboota, S.; Ahuja, A.; Ali, M.; Ali, J.; Sahni, J.K.; Bhatnagar, A. Pharmacoscintigraphic evaluation of potential of lipid nanocarriers for nose-to-brain delivery of antidepressant drug. Int. J. Pharm. 2014, 470, 99–106. [Google Scholar] [CrossRef]
  66. Ravouru, N.; Kondreddy, P.; Korakanchi, D. Formulation and evaluation of niosomal nasal drug delivery system of folic acid for brain targeting. Curr. Drug Discov. Technol. 2013, 10, 270–282. [Google Scholar] [CrossRef]
  67. Margret, A.A.; Dhayabaran, V.V.; Kumar, A.G. Nanoparticulated polymeric composites enfolding lithium carbonate as brain drug in persuading depression: An in vivo study. Prog. Biomater. 2017, 6, 165–173. [Google Scholar] [CrossRef] [Green Version]
  68. Nagpal, K.; Singh, S.K.; Mishra, D. Evaluation of safety and efficacy of brain targeted chitosan nanoparticles of minocycline. Int. J. Biol. Macromol. 2013, 59, 20–28. [Google Scholar] [CrossRef]
  69. Singh, D.P.; Rashid, M.; Hallan, S.S.; Mehra, N.K.; Prakash, A.; Mishra, N. Pharmacological evaluation of nasal delivery of selegiline hydrochloride-loaded thiolated chitosan nanoparticles for the treatment of depression. Artif. Cells Nanomed. Biotechnol. 2015, 44, 1–13. [Google Scholar] [CrossRef]
  70. Ashraf, A.; Mahmoud, P.A.; Reda, H.; Mansour, S.; Helal, M.H.; Michel, H.E.; Nasr, M. Silymarin and silymarin nanoparticles guard against chronic unpredictable mild stress induced depressive-like behavior in mice: Involvement of neurogenesis and NLRP3 inflammasome. J. Psychopharmacol. 2019, 33, 615–631. [Google Scholar] [CrossRef]
  71. Alam, M.; Najmi, A.K.; Ahmad, I.; Ahmad, F.J.; Akhtar, J.; Imam, S.S.; Akhtar, M. Formulation and evaluation of nano lipid formulation containing CNS acting drug: Molecular docking, in-vitro assessment and bioactivity detail in rats. Artif. Cells Nanomed. Biotechnol. 2018, 46, 46–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Kaur, P.; Garg, T.; Vaidya, B.; Prakash, A.; Rath, G.; Goyal, A.K. Brain delivery of intranasalin situgel of nanoparticulated polymeric carriers containing antidepressant drug: Behavioral and biochemical assessment. J. Drug Target. 2014, 23, 275–286. [Google Scholar] [CrossRef] [PubMed]
  73. Qin, J.; Yang, X.; Zhang, R.-X.; Luo, Y.-X.; Li, J.-L.; Hou, J.; Zhang, C.; Li, Y.-J.; Shi, J.; Lu, L.; et al. Monocyte mediated brain targeting delivery of macromolecular drug for the therapy of depression. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 391–400. [Google Scholar] [CrossRef] [PubMed]
  74. Haque, S.; Shadab; Fazil, M.; Kumar, M.; Sahni, J.K.; Ali, J.; Baboota, S. Venlafaxine loaded chitosan NPs for brain targeting: Pharmacokinetic and pharmacodynamic evaluation. Carbohydr. Polym. 2012, 89, 72–79. [Google Scholar] [CrossRef] [PubMed]
  75. Cayero-Otero, M.; Gomes, M.J.; Martins, C.; Álvarez-Fuentes, J.; Fernández-Arévalo, M.; Sarmento, B.; Martín-Banderas, L. In vivo biodistribution of venlafaxine-PLGA nanoparticles for brain delivery: Plain vs. functionalized nanoparticles. Expert Opin. Drug Deliv. 2019, 16, 1413–1427. [Google Scholar] [CrossRef]
  76. Haque, S.; Shadab; Sahni, J.K.; Ali, J.; Baboota, S. Development and evaluation of brain targeted intranasal alginate nanoparticles for treatment of depression. J. Psychiatr. Res. 2014, 48, 1–12. [Google Scholar] [CrossRef]
  77. Grabrucker, A.M.; Garner, C.C.; Boeckers, T.M.; Bondioli, L.; Ruozi, B.; Forni, F.; Vandelli, M.A.; Tosi, G. Development of novel Zn2+ loaded nanoparticles designed for cell-type targeted drug release in CNS neurons: In vitro evidences. PLoS ONE 2011, 6, e17851. [Google Scholar] [CrossRef] [Green Version]
  78. Vinzant, N.; Scholl, J.L.; Wu, C.-M.; Kindle, T.; Koodali, R.; Forster, G.L. Iron oxide nanoparticle delivery of peptides to the brain: Reversal of anxiety during drug withdrawal. Front. Neurosci. 2017, 11, 608. [Google Scholar] [CrossRef] [Green Version]
  79. Bari, N.K.; Fazil, M.; Hassan, Q.; Haider, R.; Gaba, B.; Narang, J.K.; Baboota, S.; Ali, J. Brain delivery of buspirone hydrochloride chitosan nanoparticles for the treatment of general anxiety disorder. Int. J. Biol. Macromol. 2015, 81, 49–59. [Google Scholar] [CrossRef]
  80. Abdelnabi, D.M.; Abdallah, M.H.; Elghamry, H.A. Buspirone hydrochloride loaded in situ nanovesicular gel as an anxiolytic nasal drug delivery system: In vitro and animal studies. AAPS PharmSciTech 2019, 20, 134. [Google Scholar] [CrossRef]
  81. Bohrey, S.; Chourasiya, V.; Pandey, A. Polymeric nanoparticles containing diazepam: Preparation, optimization, characterization, in-vitro drug release and release kinetic study. Nano Converg. 2016, 3, 1–7. [Google Scholar] [CrossRef] [Green Version]
  82. Nagpal, K.; Singh, S.K.; Mishra, D. Optimization of brain targeted gallic acid nanoparticles for improved antianxiety-like activity. Int. J. Biol. Macromol. 2013, 57, 83–91. [Google Scholar] [CrossRef]
  83. Sinha, S.V. Enhancement of in vivo efficacy and oral bioavailability of aripiprazole with solid lipid nanoparticles. AAPS PharmSciTech 2018, 19, 1264–1273. [Google Scholar] [CrossRef]
  84. Sawant, K.; Pandey, A.; Patel, S. Aripiprazole loaded poly(caprolactone) nanoparticles: Optimization and in vivo pharmacokinetics. Mater. Sci. Eng. C 2016, 66, 230–243. [Google Scholar] [CrossRef]
  85. Singh, S.K.; Hidau, M.K.; Gautam, S.; Gupta, K.; Singh, K.P.; Singh, S.K. Glycol chitosan functionalized asenapine nanostructured lipid carriers for targeted brain delivery: Pharmacokinetic and teratogenic assessment. Int. J. Biol. Macromol. 2018, 108, 1092–1100. [Google Scholar] [CrossRef]
  86. Shreya, A.B.; Managuli, R.S.; Menon, J.; Kondapalli, L.; Hegde, A.R.; Avadhani, K.; Shetty, P.K.; Amirthalingam, M.; Kalthur, G.; Mutalik, S. Nano-transfersomal formulations for transdermal delivery of asenapine maleate: In vitro and in vivo performance evaluations. J. Liposome Res. 2015, 26, 221–232. [Google Scholar] [CrossRef]
  87. Piazza, J.; Hoare, T.; Molinaro, L.; Terpstra, K.; Bhandari, J.; Selvaganapathy, P.R.; Gupta, B.; Mishra, R.K. Haloperidol-loaded intranasally administered lectin functionalized poly(ethylene glycol)–block-poly(d,l)-lactic-co-glycolic acid (PEG–PLGA) nanoparticles for the treatment of schizophrenia. Eur. J. Pharm. Biopharm. 2014, 87, 30–39. [Google Scholar] [CrossRef]
  88. Narayan, S. Lithium entrapped chitosan nanoparticles to reduce toxicity and increase cellular uptake of lithium. Environ. Toxicol. Pharmacol. 2018, 61, 79–86. [Google Scholar] [CrossRef]
  89. Patel, M.H.; Mundada, V.P.; Sawant, K.K. Fabrication of solid lipid nanoparticles of lurasidone HCl for oral delivery: Optimization, in vitro characterization, cell line studies and in vivo efficacy in schizophrenia. Drug Dev. Ind. Pharm. 2019, 45, 1242–1257. [Google Scholar] [CrossRef]
  90. Jazuli, I.; Nabi, B.; Moolakkadath, T.; Alam, T.; Baboota, S.; Ali, J.; Annu, I.J. Optimization of nanostructured lipid carriers of lurasidone hydrochloride using box-behnken design for brain targeting: In vitro and in vivo studies. J. Pharm. Sci. 2019, 108, 3082–3090. [Google Scholar] [CrossRef]
  91. Pokharkar, V.; Suryawanshi, S.; Dhapte-Pawar, V. Exploring micellar-based polymeric systems for effective nose-to-brain drug delivery as potential neurotherapeutics. Drug Deliv. Transl. Res. 2019, 10, 1019–1031. [Google Scholar] [CrossRef] [PubMed]
  92. Natarajan, J.; Baskaran, M.; Humtsoe, L.C.; Vadivelan, R.; Justin, A. Enhanced brain targeting efficacy of Olanzapine through solid lipid nanoparticles. Artif. Cells Nanomed. Biotechnol. 2016, 45, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Gadhave, D.G.; Tagalpallewar, A.A.; Kokare, C.R. Agranulocytosis-protective olanzapine-loaded nanostructured lipid carriers engineered for CNS delivery: Optimization and hematological toxicity studies. AAPS PharmSciTech 2019, 20, 22. [Google Scholar] [CrossRef] [PubMed]
  94. Abdelbary, G.A.; Tadros, M.I. Brain targeting of olanzapine via intranasal delivery of core–shell difunctional block copolymer mixed nanomicellar carriers: In vitro characterization, ex vivo estimation of nasal toxicity and in vivo biodistribution studies. Int. J. Pharm. 2013, 452, 300–310. [Google Scholar] [CrossRef] [PubMed]
  95. Fonseca, F.N.; Betti, A.H.; Carvalho, F.C.U.; Gremiao, M.P.D.U.; Dimer, F.A.; Guterres, S.S.; Tebaldi, M.L.; Rates, S.M.K.; Pohlmann, A.R. Mucoadhesive amphiphilic methacrylic copolymer-functionalized poly(ε-caprolactone) nanocapsules for nose-to-brain delivery of olanzapine. J. Biomed. Nanotechnol. 2015, 11, 1472–1481. [Google Scholar] [CrossRef] [PubMed]
  96. Baltzley, S.; Mohammad, A.; Malkawi, A.H.; Al-Ghananeem, A.M. Intranasal drug delivery of olanzapine-loaded chitosan nanoparticles. AAPS PharmSciTech 2014, 15, 1598–1602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Seju, U.; Kumar, A.; Sawant, K. Development and evaluation of olanzapine-loaded PLGA nanoparticles for nose-to-brain delivery: In vitro and in vivo studies. Acta Biomater. 2011, 7, 4169–4176. [Google Scholar] [CrossRef]
  98. Helal, H.M.; Mortada, S.M.; Sallam, M.A. Paliperidone-loaded nanolipomer system for sustained delivery and enhanced intestinal permeation: Superiority to polymeric and solid lipid nanoparticles. AAPS PharmSciTech 2016, 18, 1946–1959. [Google Scholar] [CrossRef]
  99. Sherje, A.P.; Londhe, V. Development and evaluation of pH-responsive cyclodextrin-based in situ gel of paliperidone for intranasal delivery. AAPS PharmSciTech 2017, 19, 384–394. [Google Scholar] [CrossRef]
  100. Muthu, M.S.; Sahu, A.K.; Sonali; Abdulla, A.; Kaklotar, D.; Rajesh, C.V.; Singh, S.; Pandey, B.L. Solubilized delivery of paliperidone palmitate by d-alpha-tocopheryl polyethylene glycol 1000 succinate micelles for improved short-term psychotic management. Drug Deliv. 2014, 23, 230–237. [Google Scholar] [CrossRef]
  101. Patel, M.R.; Patel, R.B.; Bhatt, K.K.; Patel, B.G.; Gaikwad, R.V. Paliperidone microemulsion for nose-to-brain targeted drug delivery system: Pharmacodynamic and pharmacokinetic evaluation. Drug Deliv. 2014, 23, 346–354. [Google Scholar] [CrossRef] [PubMed]
  102. Lugasi, L.; Grinberg, I.; Sabag, R.; Madar, R.; Einat, H.; Margel, S. Proteinoid nanocapsules as drug delivery system for improving antipsychotic activity of risperidone. Molecules 2020, 25, 4013. [Google Scholar] [CrossRef] [PubMed]
  103. Rukmangathen, R.; Yallamalli, I.M.; Yalavarthi, P.R. Formulation and biopharmaceutical evaluation of risperidone-loaded chitosan nanoparticles for intranasal delivery. Drug Dev. Ind. Pharm. 2019, 45, 1342–1350. [Google Scholar] [CrossRef] [PubMed]
  104. Patel, S.; Chavhan, S.; Soni, H.; Babbar, A.; Mathur, R.; Mishra, A.; Sawant, K. Brain targeting of risperidone-loaded solid lipid nanoparticles by intranasal route. J. Drug Target. 2010, 19, 468–474. [Google Scholar] [CrossRef] [PubMed]
  105. Kumar, M.; Misra, A.; Mishra, A.K.; Mishra, P.; Pathak, K. Mucoadhesive nanoemulsion-based intranasal drug delivery system of olanzapine for brain targeting. J. Drug Target. 2008, 16, 806–814. [Google Scholar] [CrossRef] [PubMed]
  106. Narayan, R.; Singh, M.; Ranjan, O.; Nayak, Y.; Garg, S.; Shavi, G.V.; Nayak, U.Y. Development of risperidone liposomes for brain targeting through intranasal route. Life Sci. 2016, 163, 38–45. [Google Scholar] [CrossRef]
  107. Qureshi, M.; Aqil, M.; Imam, S.S.; Ahad, A.; Sultana, Y. Formulation and evaluation of neuroactive drug loaded chitosan nanoparticle for nose to brain delivery: In-vitro characterization and in-vivo behavior study. Curr. Drug Deliv. 2018, 16, 123–135. [Google Scholar] [CrossRef]
  108. Li, J.; Zhang, W.-J.; Zhu, J.-X.; Zhu, N.; Zhang, H.-M.; Wang, X.; Zhang, J.; Wang, Q.-Q. Preparation and brain delivery of nasal solid lipid nanoparticles of quetiapine fumarate in situ gel in rat model of schizophrenia. Int. J. Clin. Exp. Med. 2015, 8, 17590–17600. [Google Scholar]
  109. Boche, M.; Pokharkar, V.B. Quetiapine nanoemulsion for intranasal drug delivery: Evaluation of brain-targeting efficiency. AAPS PharmSciTech 2016, 18, 686–696. [Google Scholar] [CrossRef]
  110. Upadhyay, P.; Trivedi, J.; Pundarikakshudu, K.; Sheth, N. Direct and enhanced delivery of nanoliposomes of anti schizophrenic agent to the brain through nasal route. Saudi Pharm. J. 2017, 25, 346–358. [Google Scholar] [CrossRef] [Green Version]
  111. Shah, B.; Khunt, D.; Misra, M.; Padh, H. Non-invasive intranasal delivery of quetiapine fumarate loaded microemulsion for brain targeting: Formulation, physicochemical and pharmacokinetic consideration. Eur. J. Pharm. Sci. 2016, 91, 196–207. [Google Scholar] [CrossRef] [PubMed]
  112. Shah, B.; Khunt, D.; Misra, M.; Padh, H. Application of box-behnken design for optimization and development of quetiapine fumarate loaded chitosan nanoparticles for brain delivery via intranasal route. Int. J. Biol. Macromol. 2016, 89, 206–218. [Google Scholar] [CrossRef] [PubMed]
  113. Naik, A.; Nair, H. Formulation and evaluation of thermosensitive biogels for nose to brain delivery of doxepin. BioMed Res. Int. 2014, 2014, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Pandey, Y.R.; Kumar, S.; Gupta, B.K.; Ali, J.; Baboota, S. Intranasal delivery of paroxetine nanoemulsion via the olfactory region for the management of depression: Formulation, behavioural and biochemical estimation. Nanotechnology 2016, 27, 25102. [Google Scholar] [CrossRef] [PubMed]
  115. Pathan, I.B.; Mene, H.; Bairagi, S. Quality by design (QbD) approach to formulate in situ gelling system for nose to brain delivery of Fluoxetine hydrochloride: Ex-vivo and In-vivo study. ARS Pharm. 2017, 58, 107–114. [Google Scholar]
  116. Kamel, R.; Abbas, H.; El-Naa, M. Composite carbohydrate interpenetrating polyelectrolyte nano-complexes (IPNC) as a controlled oral delivery system of citalopram HCl for pediatric use: In-vitro/in-vivo evaluation and histopathological examination. Drug Deliv. Transl. Res. 2018, 8, 657–669. [Google Scholar] [CrossRef]
  117. Rahman, M.A.; Harwansh, R.K.; Iqbal, Z. Systematic development of sertraline loaded solid lipid nanoparticle (SLN) by emulsification-ultrasonication method and pharmacokinetic study in sprague-dawley rats. Pharm. Nanotechnol. 2019, 7, 162–176. [Google Scholar] [CrossRef]
  118. Bhandwalkar, M.J.; Avachat, A.M. Thermoreversible nasal in situ gel of venlafaxine hydrochloride: Formulation, characterization, and pharmacodynamic evaluation. AAPS PharmSciTech 2012, 14, 101–110. [Google Scholar] [CrossRef] [Green Version]
  119. Aranaz, I.; Paños, I.; Peniche, C.; Heras, A.; Acosta, N. Chitosan spray-dried microparticles for controlled delivery of venlafaxine hydrochloride. Molecules 2017, 22, 1980. [Google Scholar] [CrossRef] [Green Version]
  120. Tong, G.-F.; Qin, N.; Sun, L.-W. Development and evaluation of Desvenlafaxine loaded PLGA-chitosan nanoparticles for brain delivery. Saudi Pharm. J. 2016, 25, 844–851. [Google Scholar] [CrossRef]
  121. Casolaro, M.; Casolaro, I. Controlled release of antidepressant drugs by multiple stimuli-sensitive hydrogels based on α-aminoacid residues. J. Drug Deliv. Sci. Technol. 2015, 30, 82–89. [Google Scholar] [CrossRef]
  122. Mitroshina, E.V.; Mishchenko, T.A.; Usenko, A.V.; Epifanova, E.A.; Yarkov, R.S.; Gavrish, M.S.; Babaev, A.A.; Vedunova, M.V. AAV-Syn-BDNF-EGFP virus construct exerts neuroprotective action on the hippocampal neural network during hypoxia in vitro. Int. J. Mol. Sci. 2018, 19, 2295. [Google Scholar] [CrossRef] [Green Version]
  123. Ma, X.-C.; Liu, P.; Zhang, X.-L.; Jiang, W.-H.; Jia, M.; Wang, C.-X.; Dong, Y.-Y.; Dang, Y.-H.; Gao, C.-G. intranasal delivery of recombinant AAV containing BDNF fused with HA2TAT: A potential promising therapy strategy for major depressive disorder. Sci. Rep. 2016, 6, 22404. [Google Scholar] [CrossRef] [Green Version]
  124. Mill, J.; Petronis, A. Molecular studies of major depressive disorder: The epigenetic perspective. Mol. Psychiatry 2007, 12, 799–814. [Google Scholar] [CrossRef] [Green Version]
  125. Bandelow, B.; Michaelis, S.; Wedekind, D. Treatment of anxiety disorders. Dialogues Clin. Neurosci. 2017, 19, 93–107. [Google Scholar]
  126. Pollack, M.; Otto, M.W.; Roy-Byrne, P.P.; Coplan, J.D.; Rothbaum, B.O.; Simon, N.M.; Gorman, J.M. Novel treatment approaches for refractory anxiety disorders. Depression Anxiety 2008, 25, 467–476. [Google Scholar] [CrossRef]
  127. Loane, C.; Politis, M. Buspirone: What is it all about? Brain Res. 2012, 1461, 111–118. [Google Scholar] [CrossRef] [PubMed]
  128. Mendrek, A.; Mancini-Marïe, A. Sex/gender differences in the brain and cognition in schizophrenia. Neurosci. Biobehav. Rev. 2016, 67, 57–78. [Google Scholar] [CrossRef] [PubMed]
  129. Clark, C.T.; Wisner, K.L. Treatment of peripartum bipolar disorder. Obstet. Gynecol. Clin. N. Am. 2018, 45, 403–417. [Google Scholar] [CrossRef]
  130. Rabin, C.R.; Siegel, S.J. Antipsychotic dosing and drug delivery. In Behavioral Neurobiology of Schizophrenia and Its Treatment; Swerdlow, N., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 4, pp. 141–177. [Google Scholar]
  131. Cheng, Y.-H.; Illum, L.; Davis, S.S. Schizophrenia and drug delivery systems. J. Drug Target. 2000, 8, 107–117. [Google Scholar] [CrossRef]
  132. Kane, J.M.; García-Ribera, C. Clinical guideline recommendations for antipsychotic long-acting injections. Br. J. Psychiatry 2009, 195, s63–s67. [Google Scholar] [CrossRef] [PubMed]
  133. Joseph, E.; Reddi, S.; Rinwa, V.; Balwani, G.; Saha, R. Design and in vivo evaluation of solid lipid nanoparticulate systems of Olanzapine for acute phase schizophrenia treatment: Investigations on antipsychotic potential and adverse effects. Eur. J. Pharm. Sci. 2017, 104, 315–325. [Google Scholar] [CrossRef] [PubMed]
  134. Jawahar, N.; Hingarh, P.K.; Arun, R.; Selvaraj, J.; Anbarasan, A.; Sathianarayanan, S.; Nagaraju, G. Enhanced oral bioavailability of an antipsychotic drug through nanostructured lipid carriers. Int. J. Biol. Macromol. 2018, 110, 269–275. [Google Scholar] [CrossRef] [PubMed]
  135. Pervaiz, F.; Ahmad, M.; Hussain, T.; Idrees, A.; Yaqoob, A.; Abbas, K. Development of olanzapine loaded PNA microgels for depot drug delivery in treatment of schizophrenia: In vitro an in vivo release profile. Acta Pol. Pharm. Drug Res. 2016, 73, 175–181. [Google Scholar]
  136. De Freitas, M.R.; Rolim, L.A.; Soares, M.F.D.L.R.; Rolim-Neto, P.J.; De Albuquerque, M.M.; Soares-Sobrinho, J.L. Inclusion complex of methyl-β-cyclodextrin and olanzapine as potential drug delivery system for schizophrenia. Carbohydr. Polym. 2012, 89, 1095–1100. [Google Scholar] [CrossRef]
  137. Venkateswarlu, V.; Manjunath, K. Preparation, characterization and in vitro release kinetics of clozapine solid lipid nanoparticles. J. Control. Release 2004, 95, 627–638. [Google Scholar] [CrossRef]
  138. Shafaat, K.; Kumar, B.; Das, S.K.; Hasan, R.U.; Prajapati, S.K. Novel nanoemulsion as vehicles for transdermal delivery of clozapine, in vitro in vivo studies. Int. J. Pharm. Pharm. Sci. 2013, 5, 126–134. [Google Scholar]
  139. Iqbal, N.; Vitorino, C.; Taylor, K.M.G. How can lipid nanocarriers improve transdermal delivery of olanzapine? Pharm. Dev. Technol. 2017, 22, 587–596. [Google Scholar] [CrossRef]
  140. Dimer, F.A.; Ortiz, M.; Pase, C.S.; Roversi, K.; Friedrich, R.B.; Pohlmann, A.R.; Burger, M.E.; Guterres, S.S. Nanoencapsulation of olanzapine increases its efficacy in antipsychotic treatment and reduces adverse effects. J. Biomed. Nanotechnol. 2014, 10, 1137–1145. [Google Scholar] [CrossRef]
  141. Joseph, E.; Reddi, S.; Rinwa, V.; Balwani, G.; Saha, R. DoE based Olanzapine loaded poly-caprolactone nanoparticles decreases extrapyramidal effects in rodent model. Int. J. Pharm. 2018, 541, 198–205. [Google Scholar] [CrossRef]
  142. Al-Dhubiab, B.E. Aripiprazole nanocrystal impregnated buccoadhesive films for schizophrenia. J. Nanosci. Nanotechnol. 2017, 17, 2345–2352. [Google Scholar] [CrossRef] [PubMed]
  143. Çelik, B. Risperidone mucoadhesive buccal tablets: Formulation design, optimization and evaluation. Drug Des. Dev. Ther. 2017, 11, 3355–3365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Silva, A.C.; Kumar, A.; Wild, W.; Ferreira, D.; Santos, D.; Forbes, B. Long-term stability, biocompatibility and oral delivery potential of risperidone-loaded solid lipid nanoparticles. Int. J. Pharm. 2012, 436, 798–805. [Google Scholar] [CrossRef] [PubMed]
  145. Alzubaidi, A.F.A.; El-Helw, A.-R.M.; Ahmed, T.A.; Ahmed, O.A.A. The use of experimental design in the optimization of risperidone biodegradable nanoparticles: In vitro and in vivo study. Artif. Cells Nanomed. Biotechnol. 2016, 45, 1–8. [Google Scholar] [CrossRef] [PubMed]
  146. Mandpe, L.; Pokharkar, V. Targeted brain delivery of iloperidone nanostructured lipid carriers following intranasal administration: In vivo pharmacokinetics and brain distribution studies. J. Nanopharm. Drug Deliv. 2013, 1, 212–225. [Google Scholar] [CrossRef]
  147. Đorđević, S.M.; Santrač, A.; Cekić, N.D.; Markovic, B.D.; Divović, B.; Ilić, T.M.; Savić, M.M.; Savic, S. Parenteral nanoemulsions of risperidone for enhanced brain delivery in acute psychosis: Physicochemical and in vivo performances. Int. J. Pharm. 2017, 533, 421–430. [Google Scholar] [CrossRef]
  148. Shan, G.W.; Makmor-Bakry, M.; Omar, M.S. Long term use of lithium and factors associated with treatment response among patients with bipolar disorder. Psychiatr. Danub. 2016, 28, 146–153. [Google Scholar]
  149. Won, E.; Kim, Y.-K. An oldie but goodie: Lithium in the treatment of bipolar disorder through neuroprotective and neurotrophic mechanisms. Int. J. Mol. Sci. 2017, 18, 2679. [Google Scholar] [CrossRef] [Green Version]
  150. Karthivashan, G.; Ganesan, P.; Park, S.-Y.; Kim, J.-S.; Choi, D.-K. Therapeutic strategies and nano-drug delivery applications in management of ageing Alzheimer’s disease. Drug Deliv. 2018, 25, 307–320. [Google Scholar] [CrossRef] [Green Version]
  151. Swerdlow, R.H.; Burns, J.M.; Khan, S.M. The Alzheimer’s disease mitochondrial cascade hypothesis: Progress and perspectives. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2014, 1842, 1219–1231. [Google Scholar] [CrossRef] [Green Version]
  152. Jojo, G.M.; Kuppusamy, G.; De, A.; Karri, V.V.S.N.R. Formulation and optimization of intranasal nanolipid carriers of pioglitazone for the repurposing in Alzheimer’s disease using Box-Behnken design. Drug Dev. Ind. Pharm. 2019, 45, 1061–1072. [Google Scholar] [CrossRef] [PubMed]
  153. Rassu, G.; Soddu, E.; Posadino, A.M.; Pintus, G.; Sarmento, B.; Giunchedi, P.; Gavini, E. Nose-to-brain delivery of BACE1 siRNA loaded in solid lipid nanoparticles for Alzheimer’s therapy. Colloids Surf. B Biointerfaces 2017, 152, 296–301. [Google Scholar] [CrossRef] [PubMed]
  154. Wang, P.; Zheng, X.; Guo, Q.; Yang, P.; Pang, X.; Qian, K.; Lu, W.; Zhang, Q.; Jiang, X. Systemic delivery of BACE1 siRNA through neuron-targeted nanocomplexes for treatment of Alzheimer’s disease. J. Control. Release 2018, 279, 220–233. [Google Scholar] [CrossRef] [PubMed]
  155. Liu, Y.; An, S.; Li, J.; Kuang, Y.; He, X.; Guo, Y.; Ma, H.; Zhang, Y.; Ji, B.; Jiang, C. Brain-targeted co-delivery of therapeutic gene and peptide by multifunctional nanoparticles in Alzheimer’s disease mice. Biomaterials 2016, 80, 33–45. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, C.; Gu, Z.; Shen, L.; Liu, X.; Lin, H. A dual targeting drug delivery system for penetrating blood-brain barrier and selectively delivering siRNA to neurons for Alzheimer’s disease treatment. Curr. Pharm. Biotechnol. 2018, 18, 1124–1131. [Google Scholar] [CrossRef] [PubMed]
  157. Agyare, E.; Jaruszewski, K.M.; Curran, G.L.; Rosenberg, J.T.; Grant, S.C.; Lowe, V.J.; Ramakrishnan, S.; Paravastu, A.K.; Poduslo, J.F.; Kandimalla, K.K. Engineering theranostic nanovehicles capable of targeting cerebrovascular amyloid deposits. J. Control. Release 2014, 185, 121–129. [Google Scholar] [CrossRef] [PubMed]
  158. Loureiro, J.A.; Gomes, B.; Fricker, G.; Coelho, M.A.; 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]
  159. Zhang, C.; Zheng, X.; Wan, X.; Shao, X.; Liu, Q.; Zhang, Q.; Zhang, Q. The potential use of H102 peptide-loaded dual-functional nanoparticles in the treatment of Alzheimer’s disease. J. Control. Release 2014, 192, 317–324. [Google Scholar] [CrossRef]
  160. Gobbi, M.; Re, F.; Canovi, M.; Beeg, M.; Gregori, M.F.; Sesana, S.; Sonnino, S.; Brogioli, D.; Musicanti, C.; Gasco, P.; et al. Lipid-based nanoparticles with high binding affinity for amyloid-β1–42 peptide. Biomaterials 2010, 31, 6519–6529. [Google Scholar] [CrossRef]
  161. Balducci, C.; Mancini, S.; Minniti, S.; La Vitola, P.; Zotti, M.; Sancini, G.; Mauri, M.; Cagnotto, A.; Colombo, L.; Fiordaliso, F.; et al. Multifunctional liposomes reduce brain—amyloid burden and ameliorate memory impairment in Alzheimer’s disease mouse models. J. Neurosci. 2014, 34, 14022–14031. [Google Scholar] [CrossRef]
  162. Tanifum, E.A.; Dasgupta, I.; Srivastava, M.; Bhavane, R.C.; Sun, L.; Berridge, J.; Pourgarzham, H.; Kamath, R.; Espinosa, G.; Cook, S.C.; et al. Intravenous delivery of targeted liposomes to amyloid-β pathology in APP/PSEN1 transgenic mice. PLoS ONE 2012, 7, e48515. [Google Scholar] [CrossRef] [PubMed]
  163. Song, Q.; Huang, M.; Yao, L.; Wang, X.; Gu, X.; Chen, J.; Huang, J.; Hu, Q.; Kang, T.; Rong, Z.; et al. Lipoprotein-based nanoparticles rescue the memory loss of mice with Alzheimer’s disease by accelerating the clearance of amyloid-beta. ACS Nano 2014, 8, 2345–2359. [Google Scholar] [CrossRef]
  164. Poduslo, J.F.; Hultman, K.L.; Curran, G.L.; Preboske, G.M.; Chamberlain, R.; Marjańska, M.; Garwood, M.; Jack, C.R., Jr.; Wengenack, T.M. Targeting vascular amyloid in arterioles of Alzheimer disease transgenic mice with amyloid β protein antibody-coated nanoparticles. J. Neuropathol. Exp. Neurol. 2011, 70, 653–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Prades, R.; Guerrero, S.; Araya, E.; Molina, C.; Salas, E.; Zurita, E.; Selva, J.; Egea, G.; López-Iglesias, C.; Teixidó, M.; et al. Delivery of gold nanoparticles to the brain by conjugation with a peptide that recognizes the transferrin receptor. Biomaterials 2012, 33, 7194–7205. [Google Scholar] [CrossRef] [PubMed]
  166. Gao, N.; Sun, H.; Dong, K.; Ren, J.; Qu, X. Gold-nanoparticle-based multifunctional amyloid-β inhibitor against Alzheimer’s disease. Chem. A Eur. J. 2015, 21, 829–835. [Google Scholar] [CrossRef] [PubMed]
  167. Hsieh, S.; Chang, C.-W.; Chou, H.-H. Gold nanoparticles as amyloid-like fibrillogenesis inhibitors. Colloids Surf. B Biointerfaces 2013, 112, 525–529. [Google Scholar] [CrossRef]
  168. Cimini, A.; D’Angelo, B.; Das, S.; Gentile, R.; Benedetti, E.; Singh, V.; Monaco, A.; Santucci, S.; Seal, S. Antibody-conjugated PEGylated cerium oxide nanoparticles for specific targeting of Aβ aggregates modulate neuronal survival pathways. Acta Biomater. 2012, 8, 2056–2067. [Google Scholar] [CrossRef]
  169. Vakilinezhad, M.A.; Amini, A.; Javar, H.A.; Zarandi, B.F.B.B.; Montaseri, H.; Dinarvand, R. Nicotinamide loaded functionalized solid lipid nanoparticles improves cognition in Alzheimer’s disease animal model by reducing Tau hyperphosphorylation. DARU J. Pharm. Sci. 2018, 26, 165–177. [Google Scholar] [CrossRef]
  170. Jinwal, U.K.; Groshev, A.; Zhang, J.; Grover, A.; Sutariya, V.B. Preparation and characterization of methylene blue nanoparticles for Alzheimer’s disease and other tauopathies. Curr. Drug Deliv. 2014, 11, 541–550. [Google Scholar] [CrossRef]
  171. Sonawane, S.K.; Ahmad, A.; Chinnathambi, S. Protein-capped metal nanoparticles inhibit Tau aggregation in Alzheimer’s disease. ACS Omega 2019, 4, 12833–12840. [Google Scholar] [CrossRef] [Green Version]
  172. Misra, S.; Chopra, K.; Sinha, V.R.; Medhi, B. Galantamine-loaded solid–lipid nanoparticles for enhanced brain delivery: Preparation, characterization, in vitro and in vivo evaluations. Drug Deliv. 2015, 23, 1434–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Li, W.; Zhou, Y.-Q.; Zhao, N.; Hao, B.; Wang, X.; Kong, P. Pharmacokinetic behavior and efficiency of acetylcholinesterase inhibition in rat brain after intranasal administration of galanthamine hydrobromide loaded flexible liposomes. Environ. Toxicol. Pharmacol. 2012, 34, 272–279. [Google Scholar] [CrossRef] [PubMed]
  174. Fornaguera, C.; Feiner-Gracia, N.; Calderó, G.; García-Celma, M.J.; Solans, C. Galantamine-loaded PLGA nanoparticles, from nano-emulsion templating, as novel advanced drug delivery systems to treat neurodegenerative diseases. Nanoscale 2015, 7, 12076–12084. [Google Scholar] [CrossRef] [PubMed]
  175. Hanafy, A.S.; Farid, R.M.; Helmy, M.W.; Elgamal, S.S. Pharmacological, toxicological and neuronal localization assessment of galantamine/chitosan complex nanoparticles in rats: Future potential contribution in Alzheimer’s disease management. Drug Deliv. 2016, 23, 3111–3122. [Google Scholar] [CrossRef] [PubMed]
  176. Hanafy, A.S.; Farid, R.M.; Elgamal, S.S. Complexation as an approach to entrap cationic drugs into cationic nanoparticles administered intranasally for Alzheimer’s disease management: Preparation and detection in rat brain. Drug Dev. Ind. Pharm. 2015, 41, 2055–2068. [Google Scholar] [CrossRef]
  177. Ullah, Z.; Al Asmari, A.K.; Tariq, M.; Fatani, A. Preparation, characterization, and in vivo evaluation of intranasally administered liposomal formulation of donepezil. Drug Des. Dev. Ther. 2016, 205–215. [Google Scholar] [CrossRef] [Green Version]
  178. Bhavna, M.D.S.; Ali, M.; Ali, R.; Bhatnagar, A.; Baboota, S.; Ali, J. Donepezil nanosuspension intended for nose to brain targeting: In vitro and in vivo safety evaluation. Int. J. Biol. Macromol. 2014, 67, 418–425. [Google Scholar] [CrossRef]
  179. Jakki, S.L.; Ramesh, Y.V.; Gowthamarajan, K.; Senthil, V.; Jain, K.; Sood, S.; Pathak, D. Novel anionic polymer as a carrier for CNS delivery of anti-Alzheimer drug. Drug Deliv. 2016, 23, 3471–3479. [Google Scholar] [CrossRef] [Green Version]
  180. Baysal, I.; Ucar, G.; Gultekinoglu, M.; Ulubayram, K.; Yabanoglu-Ciftci, S. Donepezil loaded PLGA-b-PEG nanoparticles: Their ability to induce destabilization of amyloid fibrils and to cross blood brain barrier in vitro. J. Neural Transm. 2017, 124, 33–45. [Google Scholar] [CrossRef]
  181. Yang, C.-R.; Zhao, X.-L.; Hu, H.-Y.; Li, K.-X.; Sun, X.; Li, L.; Chen, D.-W. Preparation, optimization and characteristic of huperzine a loaded nanostructured lipid carriers. Chem. Pharm. Bull. 2010, 58, 656–661. [Google Scholar] [CrossRef] [Green Version]
  182. Patel, P.A.; Patil, S.C.; Kalaria, D.R.; Kalia, Y.N.; Patravale, V. Comparative in vitro and in vivo evaluation of lipid based nanocarriers of Huperzine A. Int. J. Pharm. 2013, 446, 16–23. [Google Scholar] [CrossRef] [PubMed]
  183. Jiang, Y.; Liu, C.; Zhai, W.; Zhuang, N.; Han, T.; Ding, Z. The optimization design of lactoferrin loaded hupa nanoemulsion for targeted drug transport via intranasal route. Int. J. Nanomed. 2019, 14, 9217–9234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Meng, Q.; Wang, A.; Hua, H.; Jiang, Y.; Wang, Y.; Mu, H.; Wu, Z.; Sun, K. Intranasal delivery of Huperzine A to the brain using lactoferrin-conjugated N-trimethylated chitosan surface-modified PLGA nanoparticles for treatment of Alzheimer’s disease. Int. J. Nanomed. 2018, 13, 705–718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Qian, S.; Wong, Y.C.; Zuo, Z. Development, characterization and application of in situ gel systems for intranasal delivery of tacrine. Int. J. Pharm. 2014, 468, 272–282. [Google Scholar] [CrossRef]
  186. Luppi, B.; Bigucci, F.; Corace, G.; Delucca, A.; Cerchiara, T.; Sorrenti, M.; Catenacci, L.; Di Pietra, A.M.; Zecchi, V. Albumin nanoparticles carrying cyclodextrins for nasal delivery of the anti-Alzheimer drug tacrine. Eur. J. Pharm. Sci. 2011, 44, 559–565. [Google Scholar] [CrossRef]
  187. Wilson, B.; Samanta, M.K.; Santhi, K.; Kumar, K.S.; Ramasamy, M.; Suresh, B. Chitosan nanoparticles as a new delivery system for the anti-Alzheimer drug tacrine. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 144–152. [Google Scholar] [CrossRef]
  188. Wavikar, P.; Pai, R.; Vavia, P. Nose to brain delivery of rivastigmine by in situ gelling cationic nanostructured lipid carriers: Enhanced brain distribution and pharmacodynamics. J. Pharm. Sci. 2017, 106, 3613–3622. [Google Scholar] [CrossRef]
  189. Ismail, M.F.; ElMeshad, A.N.; Salem, N.A.-H. Potential therapeutic effect of nanobased formulation of rivastigmine on rat model of Alzheimer’s disease. Int. J. Nanomed. 2013, 8, 393–406. [Google Scholar] [CrossRef]
  190. Arumugam, K.; Subramanian, G.S.; Mallayasamy, S.R.; Averineni, R.K.; Reddy, M.S.; Udupa, N. A study of rivastigmine liposomes for delivery into the brain through intranasal route. Acta Pharm. 2008, 58, 287–297. [Google Scholar] [CrossRef]
  191. Yang, Z.-Z.; Zhang, Y.-Q.; Wang, Z.-Z.; Wu, K.; Lou, J.-N.; Qi, X.-R. Enhanced brain distribution and pharmacodynamics of rivastigmine by liposomes following intranasal administration. Int. J. Pharm. 2013, 452, 344–354. [Google Scholar] [CrossRef]
  192. Mutlu, N.B.; Değim, Z.; Yilmaz, Ş.; Eşsiz, D.; Nacar, A. New perspective for the treatment of Alzheimer diseases: Liposomal rivastigmine formulations. Drug Dev. Ind. Pharm. 2011, 37, 775–789. [Google Scholar] [CrossRef] [PubMed]
  193. Joshi, S.; Chavhan, S.S.; Sawant, K.K. Rivastigmine-loaded PLGA and PBCA nanoparticles: Preparation, optimization, characterization, in vitro and pharmacodynamic studies. Eur. J. Pharm. Biopharm. 2010, 76, 189–199. [Google Scholar] [CrossRef] [PubMed]
  194. Khemariya, R.P.; Khemariya, P.S. New-fangled approach in the management of Alzheimer by formulation of polysorbate 80 coated chitosan nanoparticles of rivastigmine for brain delivery and their in vivo evaluation. Int. J. Cur. Res. Med. Sci. 2016, 2, 18–29. [Google Scholar]
  195. Fazil, M.; Md, S.; Haque, S.; Kumar, M.; Baboota, S.; Sahni, J.K.; Ali, J. Development and evaluation of rivastigmine loaded chitosan nanoparticles for brain targeting. Eur. J. Pharm. Sci. 2012, 47, 6–15. [Google Scholar] [CrossRef]
  196. Casadomé-Perales, Á.; De Matteis, L.; Alleva, M.; Infantes-Rodríguez, C.; Palomares-Pérez, I.; Saito, T.; Saido, T.C.; Esteban, J.A.; Nebreda, A.R.; De La Fuente, J.M.; et al. Inhibition of p38 MAPK in the brain through nasal administration of p38 inhibitor loaded in chitosan nanocapsules. Nanomedicine 2019, 14, 2409–2422. [Google Scholar] [CrossRef]
  197. Dara, T.; Vatanara, A.; Sharifzadeh, M.; Khani, S.; Vakilinezhad, M.A.; Vakhshiteh, F.; Meybodi, M.N.; Malvajerd, S.S.; Hassani, S.; Mosaddegh, M.H. Improvement of memory deficits in the rat model of Alzheimer’s disease by erythropoietin-loaded solid lipid nanoparticles. Neurobiol. Learn. Mem. 2019, 166, 107082. [Google Scholar] [CrossRef]
  198. Elnaggar, Y.S.R.; Etman, S.M.; Abdelmonsif, D.A.; Abdallah, O.Y. Novel piperine-loaded Tween-integrated monoolein cubosomes as brain-targeted oral nanomedicine in Alzheimer’s disease: Pharmaceutical, biological, and toxicological studies. Int. J. Nanomed. 2015, 10, 5459–5473. [Google Scholar] [CrossRef] [Green Version]
  199. Huang, M.; Hu, M.; Song, Q.; Song, H.; Huang, J.; Gu, X.; Wang, X.; Chen, J.; Kang, T.; Feng, X.; et al. GM1-modified lipoprotein-like nanoparticle: Multifunctional nanoplatform for the combination therapy of Alzheimer’s disease. ACS Nano 2015, 9, 10801–10816. [Google Scholar] [CrossRef]
  200. Liu, Z.; Jiang, M.; Kang, T.; Miao, D.; Gu, G.; Song, Q.; Yao, L.; Hu, Q.; Tu, Y.; Pang, Z.; et al. Lactoferrin-modified PEG-co-PCL nanoparticles for enhanced brain delivery of NAP peptide following intranasal administration. Biomaterials 2013, 34, 3870–3881. [Google Scholar] [CrossRef]
  201. Yang, L.; Yin, T.; Liu, Y.; Sun, J.; Zhou, Y.; Liu, J. Gold nanoparticle-capped mesoporous silica-based H2O2-responsive controlled release system for Alzheimer’s disease treatment. Acta Biomater. 2016, 46, 177–190. [Google Scholar] [CrossRef]
  202. Yin, T.; Yang, L.; Liu, Y.; Zhou, X.; Sun, J.; Liu, J. Sialic acid (SA)-modified selenium nanoparticles coated with a high blood–brain barrier permeability peptide-B6 peptide for potential use in Alzheimer’s disease. Acta Biomater. 2015, 25, 172–183. [Google Scholar] [CrossRef] [PubMed]
  203. Loureiro, J.A.; Andrade, S.; Duarte, A.; Neves, A.; Queiroz, J.F.; Nunes, C.; Sevin, E.; Fenart, L.; Gosselet, F.; Coelho, M.A.; et al. Resveratrol and grape extract-loaded solid lipid nanoparticles for the treatment of Alzheimer’s disease. Molecules 2017, 22, 277. [Google Scholar] [CrossRef] [PubMed]
  204. Sun, J.; Wei, C.; Liu, Y.; Xie, W.; Xu, M.; Zhou, H.; Liu, J. Progressive release of mesoporous nano-selenium delivery system for the multi-channel synergistic treatment of Alzheimer’s disease. Biomaterials 2019, 197, 417–431. [Google Scholar] [CrossRef]
  205. Gothwal, A.; Kumar, H.; Nakhate, K.T.; Ajazuddin; Dutta, A.; Borah, A.; Gupta, U. Lactoferrin coupled lower generation PAMAM dendrimers for brain targeted delivery of memantine in aluminum-chloride-induced Alzheimer’s disease in mice. Bioconjugate Chem. 2019, 30, 2573–2583. [Google Scholar] [CrossRef] [PubMed]
  206. Laserra, S.; Basit, A.; Sozio, P.; Marinelli, L.; Fornasari, E.; Cacciatore, I.; Ciulla, M.; Turkez, H.; Geyikoğglu, F.; Di Stefano, A. Solid lipid nanoparticles loaded with lipoyl–memantine codrug: Preparation and characterization. Int. J. Pharm. 2015, 485, 183–191. [Google Scholar] [CrossRef]
  207. Alawdi, S.H.; Eidi, H.; Safar, M.M.; Abdel-Wahhab, M.A. Loading amlodipine on diamond nanoparticles: A novel drug delivery system. Nanotechnol. Sci. Appl. 2019, 12, 47–53. [Google Scholar] [CrossRef] [Green Version]
  208. Meng, F.; Asghar, S.; Gao, S.; Su, Z.; Song, J.; Huo, M.; Meng, W.; Ping, Q.; Xiao, Y. A novel LDL-mimic nanocarrier for the targeted delivery of curcumin into the brain to treat Alzheimer’s disease. Colloids Surf. B Biointerfaces 2015, 134, 88–97. [Google Scholar] [CrossRef]
  209. Gao, C.; Wang, Y.; Sun, J.; Han, Y.; Gong, W.; Li, Y.; Feng, Y.; Wang, H.; Yang, M.; Li, Z.; et al. Neuronal mitochondria-targeted delivery of curcumin by biomimetic engineered nanosystems in Alzheimer’s disease mice. Acta Biomater. 2020, 108, 285–299. [Google Scholar] [CrossRef]
  210. Kuo, Y.-C.; Lin, C.-Y.; Li, J.-S.; Lou, Y.-I. Wheat germ agglutinin-conjugated liposomes incorporated with cardiolipin to improve neuronal survival in Alzheimer’s disease treatment. Int. J. Nanomed. 2017, 12, 1757–1774. [Google Scholar] [CrossRef] [Green Version]
  211. Huo, X.; Zhang, Y.; Jin, X.; Li, Y.; Zhang, L. A novel synthesis of selenium nanoparticles encapsulated PLGA nanospheres with curcumin molecules for the inhibition of amyloid β aggregation in Alzheimer’s disease. J. Photochem. Photobiol. B Biol. 2019, 190, 98–102. [Google Scholar] [CrossRef]
  212. Jaruszewski, K.M.; Curran, G.L.; Swaminathan, S.K.; Rosenberg, J.T.; Grant, S.C.; Ramakrishnan, S.; Lowe, V.J.; Poduslo, J.F.; Kandimalla, K.K. Multimodal nanoprobes to target cerebrovascular amyloid in Alzheimer’s disease brain. Biomaterials 2013, 35, 1967–1976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Bilia, A.R.; Nardiello, P.; Piazzini, V.; Leri, M.; Bergonzi, M.C.; Bucciantini, M.; Casamenti, F. Successful brain delivery of andrographolide loaded in human albumin nanoparticles to TgCRND8 mice, an Alzheimer’s disease mouse model. Front. Pharmacol. 2019, 10, 910. [Google Scholar] [CrossRef] [PubMed]
  214. Al-Azzawi, S.; Masheta, D.; Guildford, A.; Phillips, G.; Santin, M. Dendrimeric poly(epsilon-lysine) delivery systems for the enhanced permeability of flurbiprofen across the blood-brain barrier in Alzheimer’s disease. Int. J. Mol. Sci. 2018, 19, 3224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Muntimadugu, E.; Dhommati, R.; Jain, A.; Challa, V.G.S.; Shaheen, M.; Khan, W. Intranasal delivery of nanoparticle encapsulated tarenflurbil: A potential brain targeting strategy for Alzheimer’s disease. Eur. J. Pharm. Sci. 2016, 92, 224–234. [Google Scholar] [CrossRef]
  216. Saffari, P.M.; Alijanpour, S.; Takzaree, N.; Sahebgharani, M.; Etemad-Moghadam, S.; Noorbakhsh, F.; Partoazar, A. Metformin loaded phosphatidylserine nanoliposomes improve memory deficit and reduce neuroinflammation in streptozotocin-induced Alzheimer’s disease model. Life Sci. 2020, 255, 117861. [Google Scholar] [CrossRef]
  217. Fernandes, J.; Ghate, V.M.; Mallik, S.B.; Lewis, S. 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]
  218. Kuo, Y.-C.; Wang, C.-T. Protection of SK-N-MC cells against β-amyloid peptide-induced degeneration using neuron growth factor-loaded liposomes with surface lactoferrin. Biomaterials 2014, 35, 5954–5964. [Google Scholar] [CrossRef]
  219. Kuo, Y.-C.; Lin, C.-Y. Targeting delivery of liposomes with conjugated p-aminophenyl-α-d-manno-pyranoside and apolipoprotein E for inhibiting neuronal degeneration insulted with β-amyloid peptide. J. Drug Target. 2014, 23, 147–158. [Google Scholar] [CrossRef]
  220. Kuo, Y.-C.; Chou, P.-R. Neuroprotection against degeneration of SK-N-MC cells using neuron growth factor-encapsulated liposomes with surface cereport and transferrin. J. Pharm. Sci. 2014, 103, 2484–2497. [Google Scholar] [CrossRef]
  221. Zhang, C.; Chen, J.; Feng, C.; Shao, X.; Liu, Q.; Zhang, Q.; Pang, Z.; Jiang, X. Intranasal nanoparticles of basic fibroblast growth factor for brain delivery to treat Alzheimer’s disease. Int. J. Pharm. 2014, 461, 192–202. [Google Scholar] [CrossRef]
  222. Angelova, A.; Drechsler, M.; Garamus, V.M.; Angelov, B. Pep-lipid cubosomes and vesicles compartmentalized by micelles from self-assembly of multiple neuroprotective building blocks including a large peptide hormone PACAP-DHA. ChemNanoMat 2019, 5, 1381–1389. [Google Scholar] [CrossRef]
  223. Leblanc, R.M.; Liyanage, P.Y.; Devadoss, D.; Guevara, L.R.R.; Cheng, L.; Graham, R.M.; Chand, H.S.; Al-Youbi, A.O.; Bashammakh, A.S.; El-Shahawi, M.S.; et al. Nontoxic amphiphilic carbon dots as promising drug nanocarriers across the blood–brain barrier and inhibitors of β-amyloid. Nanoscale 2019, 11, 22387–22397. [Google Scholar] [CrossRef]
  224. Chen, Y.; Fan, H.; Xu, C.; Hu, W.; Yu, B. Efficient cholera toxin B subunit-based nanoparticles with MRI capability for drug delivery to the brain following intranasal administration. Macromol. Biosci. 2018, 19, e1800340. [Google Scholar] [CrossRef]
  225. Cui, N.; Lu, H.; Li, M. Magnetic nanoparticles associated PEG/PLGA block copolymer targeted with anti-transferrin receptor antibodies for Alzheimer’s disease. J. Biomed. Nanotechnol. 2018, 14, 1017–1024. [Google Scholar] [CrossRef] [PubMed]
  226. Do, T.D.; Amin, F.U.; Noh, Y.; Kim, M.O.; Yoon, J.; Duc, D.T.; Ul, A.F.; YeongIl, N.; Ok, K.M.; Jungwon, Y. Guidance of magnetic nanocontainers for treating Alzheimer’s disease using an electromagnetic, targeted drug-delivery actuator. J. Biomed. Nanotechnol. 2016, 12, 569–574. [Google Scholar] [CrossRef] [PubMed]
  227. Sanati, M.; Khodagholi, F.; Aminyavari, S.; Ghasemi, F.; Gholami, M.; Kebriaeezadeh, A.; Sabzevari, O.; Hajipour, M.J.; Imani, M.; Mahmoudi, M.; et al. Impact of gold nanoparticles on amyloid β-induced Alzheimer’s disease in a rat animal model: Involvement of STIM proteins. ACS Chem. Neurosci. 2019, 10, 2299–2309. [Google Scholar] [CrossRef]
  228. Pardridge, W.M. Blood–brain barrier drug delivery of IgG fusion proteins with a transferrin receptor monoclonal antibody. Expert Opin. Drug Deliv. 2014, 12, 207–222. [Google Scholar] [CrossRef]
  229. Liu, W.; Sun, F.; Wan, M.; Jiang, F.; Bo, X.; Lin, L.; Tang, H.; Xu, S. β-Sheet breaker peptide-HPYD for the treatment of Alzheimer’s disease: Primary studies on behavioral test and transcriptional profiling. Front. Pharmacol. 2018, 8, 969. [Google Scholar] [CrossRef] [Green Version]
  230. Zheng, X.; Shao, X.; Zhang, C.; Tan, Y.; Liu, Q.; Wan, X.; Zhang, Q.; Xu, S.; Jiang, X. Intranasal H102 peptide-loaded liposomes for brain delivery to treat Alzheimer’s disease. Pharm. Res. 2015, 32, 3837–3849. [Google Scholar] [CrossRef]
  231. Hu, B.; Dai, F.; Fan, Z.; Ma, G.; Tang, Q.; Zhang, X. Nanotheranostics: Congo Red/Rutin-MNPs with enhanced magnetic resonance imaging and H2O2-responsive therapy of alzheimer’s disease in APPswe/PS1dE9 transgenic mice. Adv. Mater. 2015, 27, 5499–5505. [Google Scholar] [CrossRef]
  232. Gao, C.; Chu, X.; Gong, W.; Zheng, J.; Xie, X.; Wang, Y.; Yang, M.; Li, Z.; Gao, C.; Yang, Y. Neuron tau-targeting biomimetic nanoparticles for curcumin delivery to delay progression of Alzheimer’s disease. J. Nanobiotechnol. 2020, 18, 1–23. [Google Scholar] [CrossRef] [PubMed]
  233. Singh, N.A.; Mandal, A.K.A.; Khan, Z.A. Potential neuroprotective properties of epigallocatechin-3-gallate (EGCG). Nutr. J. 2015, 15, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Agrawal, M.; Saraf, S.; Saraf, S.; Antimisiaris, S.G.; Chougule, M.B.; Shoyele, S.A.; Alexander, A. Nose-to-brain drug delivery: An update on clinical challenges and progress towards approval of anti-Alzheimer drugs. J. Control. Release 2018, 281, 139–177. [Google Scholar] [CrossRef] [PubMed]
  235. Md, S.; Haque, S.; Fazil, M.; Kumar, M.; Baboota, S.; Narang, J.K.; Ali, J. Optimised nanoformulation of bromocriptine for direct nose-to-brain delivery: Biodistribution, pharmacokinetic and dopamine estimation by ultra-HPLC/mass spectrometry method. Expert Opin. Drug Deliv. 2014, 11, 827–842. [Google Scholar] [CrossRef] [Green Version]
  236. Lauzon, M.-A.; Daviau, A.; Marcos, B.; Faucheux, N. Nanoparticle-mediated growth factor delivery systems: A new way to treat Alzheimer’s disease. J. Control. Release 2015, 206, 187–205. [Google Scholar] [CrossRef]
Figure 1. The main pathways of nanoparticles entering the brain. (A) The main mechanisms for blood brain barrier (BBB) penetration. (B) The mechanisms of nanoparticle transport across the cell membrane. BBB—blood-brain barrier; NPs—nanoparticles.
Figure 1. The main pathways of nanoparticles entering the brain. (A) The main mechanisms for blood brain barrier (BBB) penetration. (B) The mechanisms of nanoparticle transport across the cell membrane. BBB—blood-brain barrier; NPs—nanoparticles.
Molecules 25 05294 g001
Figure 2. The main pathological pathway of Alzheimer’s disease and therapy methods, targeted to these mechanisms. Ach—acetylcholine, AChE—acetylcholine esterase, ROS—reactive oxygen species, NMDA—N-methyl-D-aspartate.
Figure 2. The main pathological pathway of Alzheimer’s disease and therapy methods, targeted to these mechanisms. Ach—acetylcholine, AChE—acetylcholine esterase, ROS—reactive oxygen species, NMDA—N-methyl-D-aspartate.
Molecules 25 05294 g002
Table 1. Advantages of nanoparticles (NPs) administration (adm.) compared to other methods of drug delivery for CNS treatment.
Table 1. Advantages of nanoparticles (NPs) administration (adm.) compared to other methods of drug delivery for CNS treatment.
PropertyNPsIntranasal Adm.Transdermal Adm.Oral
Adm.
Intravenous
Adm.
Controlled releaseAvailableNot availableAvailableNot availableNot available
BBB penetration
  • Does not depend on the properties of drug
  • Target ligands
  • Transcytosis, endocytosis and exocytosis and others.
  • Does not depend on the properties of drug
  • Direct transport
Depends on the properties of drugDepends on the properties of drugDepends on the properties of drug
Ability to targeted deliveryHighNot availableNot availableNot availableNot available
BioavailabilityHighHighPoorPoorPoor
Hepatic first-pass metabolismPreventPreventPreventNot preventNot prevent
ToxicityLess pronounced side effects because of lower concentrations of therapeutic drugs and more stable release characteristics.Less pronounced side effects because of lower concentrations of therapeutic drugs and direct transport through BBB.Less pronounced side effects because of lower concentrations of therapeutic drugs and more stable release characteristicHigh chance and dose-depended side effectsHigh chance and dose-depended side effects
Table 2. Various categories of nanocarriers.
Table 2. Various categories of nanocarriers.
NanocarrierDescriptionMaterialsBBB PenetrationAdvantagesDisadvantagesReferences
Polymer-Based Nanocarrier Systems
Polymeric NPsThe colloidal carriers are obtained from biodegradable and biocompatible natural or synthetic polymers into which drugs are loaded in either solid-state or solution.Alginate, chitosan, gelatin, cellulose, polyacrylate, polyanhydride, PLGA, PACA, PCL, PEI, PLA, PEG.
  • The NPs are permeated the BBB through the tight junctions. The NPs are retained in the brain blood capillaries and are adsorbed of the NPs to the capillary walls.
  • Endocytosis by the endothelial cells.
  • Transcytosis.
  • The inhibition of the efflux system by using polysorbate 80 as the coating agent.
  • The surface functionalization with targeting ligands.
  • The significant potential for brain drug delivery across the BBB.
  • Biocompatibility and biodegradability
  • Good stability.
  • Low cost.
  • Less toxicity.
  • Ease in production.
  • Low immunogenic response.
  • The mucoadhesive polymers (alginate, chitosan, cellulose, gelatin) can be used intranasally to bypass the BBB. Can be used as gelling/viscosity building agents to conquer nasal mucociliary clearance.
  • Controlled drug release.
  • The surface groups can be conjugated with targeting ligands.
  • Uncertain potential toxicity.
  • Slow degradability.
[28,29,30,31,32,33]
Polymeric micellesThe core–shell structure from amphiphilic blocks copolymers that aggregate in aqueous solutions to form spheroidal NPs with a hydrophobic core and hydrophilic surface.PEG, PLGA, cholesterol conjugated polyoxyethylene sorbitol oleate.The surface functionalization with targeting ligands.
  • Biocompatibility.
  • Possibility of solubilizing lipophilic drugs.
  • The surface groups can be conjugated with targeting ligands.
The efficiency delivering across the intact BBB still needs further investigation.[17]
DendrimersThe monodisperse symmetric molecules that comprise a series of branching units around an inner core, which have a spheroidal shape and radially crowded layers.Poly(amido amide) (PAMAM)The internalization by brain capillary endothelial cells through a clathrin- and caveolane-mediated energy-depending endocytosis, also partly through macropinocytosis.
  • The core is loosely packed in comparison to the periphery and is suitable for the entrapment of drugs.
  • The presence of numerous surface groups allows for high drug payload and multifunctionality.
  • The surface groups can be conjugated with targeting ligands.
Potential toxicity.[17,30,31,32]
Lipid-Based Nanocarrier System
Less or no toxicity, biodegradability, and ability to successfully deliver biomolecules, DNA, RNA, genes, antibodies, etc.
LiposomesThe spherical vesicles with different sizes (20 nm to 100 μm), with an aqueous inner core enclosed by unilamellar or multilamellar phospholipid bilayers. Liposomes may have different surface charges and uni-, bi- or multi-lamellar structures.Sphingomyelin, phosphatidyl choline, glycerophospholipids, cholesterol, phosphatidylcholine, cardiolipin.
  • Receptor-mediated and adsorption-mediated endocytosis (for small liposomes with a diameter not larger than 100 nm)
  • The surface functionalization with targeting ligands
  • Cationization of the conjugated ligands is another method to improve BBB transport rate of liposomes.
  • Interactions of liposomes with cells can be realized by: adsorption, fusion, endocytosis, and lipid transfer.
  • Good biocompatibility.
  • Non-toxic.
  • Widely investigated.
  • Encapsulate both hydrophilic and lipophilic drugs.
  • Increase the solubility of drugs.
  • Improve pharmacokinetic properties and protect the drugs from enzymatic degradation.
  • Reduction of harmful side effects of drugs.
  • Improving therapeutic effectiveness of drugs.
  • Opsonization and rapid clearance by macrophages of the mononuclear phagocytic system organs. This can be fixed by using surface coatings, such as PEG.
  • Low drug transport rate.
  • Poor stability.
  • Difficulty of binding ligands to the surface as a result of the small number of available surface groups and steric hindrance.
[28,29,30,32,33]
Solid lipid NPsNano-sized dispersions of biocompatible lipidsThe lipid core consists of triglycerides, diglycerides, monoglycerides, fatty acids, steroids, stearic acid or waxes stabilized by various surfactants
  • The surface functionalization with targeting ligands
  • Can be used intranasally to bypass the BBB. The use of cationic lipids can improve mucoadhesion in the nasal cavity by promoting electrostatic interactions with mucus.
  • Endocytosis by the endothelial cells.
  • Adsorptive-mediated transcytosis of cationic NPs.
  • Good physical stability
  • The capability to deliver lipophilic drugs dispersed in the hydrophobic core into cells.
  • Improves bioavailability of drugs.
  • Increases drug loading ability.
  • Controlled drug release.
  • Reduced cytotoxicity of drugs.
  • Reduces the RES uptake, increases the retention and circulation time and protect drugs from degradation.
  • The surface groups can be conjugated with targeting ligands.
  • Easy clearance by the reticuloendothelial system
  • Low loading capacity
  • Drug expulsion after crystallization Relatively high water content of the dispersions.
[29,30,33]
NanoemulsionsColloidal droplet system of oil-in-water (O/W) or water-in-oil (W/O) formulations stabilized with surface-active agents
Oil droplet size of nanoemulsions ranges from 10 to 100 nm, making them interesting systems to improve drug delivery
Edible oils, such as flaxseed oil, pine-nut oil, hemp oil, fish oil as well as safflower oil and wheat-germ oil, biocompatible surfactants such as egg phosphatidylcholine which is one of the components of cell membrane lipids, deoxycholic acid, stearylamine, dioleoyltrimethyl ammonium propane.
  • The surface functionalization with targeting ligands
  • Receptor-mediated endocytosis of cells
  • Biocompatibility.
  • The ability to solubilize hydrophobic or hydrophilic drugs.
  • Improves the bioavailability of the drug.
  • The surface groups can be conjugated with targeting ligands.
  • Thermodynamic instability.
  • Instability upon storage.
  • Immediate drug release.
[17,32,33]
Inorganic Nanoparticles
Mesoporous Silica NPsSilica xerogels and mesoporous silica nanoparticlesN-Cetyltrimethylammonium bromide, TetraethoxysilaneTranscytosis of vascular endothelial cells (PEGylated NPs)
  • Good biocompatibility.
  • Stability.
  • Highly porous framework.
  • High drug loading efficiency.
  • Controlled drug release.
  • Potential toxicity and adverse effect showed by recent studies
  • Contribute to the level increase of reactive oxygen species
  • Can induce the elevated production of malondialdehyde and decreased glutathione level
[30,34]
Gold NPs
  • Adsorption-mediated endocytosis
  • Diffusion
  • Can easily penetrate cells because of their small size, high efficiency in cellular uptake.
  • Easy conjugation to biomolecules.
  • Can be easy functionalized by surface modification (small size, large surface area and pore volume, high reactivity).
Drug delivery efficiency needs further investigation.[34]
Carbon
nanotubes
Graphitic sheets rolled into single-walled or multiple walled tubes with an enormous surface areaGraphitic sheets
  • The surface functionalization with targeting ligands.
  • PEG-modified carbon nanotubes also exhibited the capability of crossing the BBB.
  • Biocompatibility.
  • Efficient loading of multiple molecules.
  • Potential toxicity.
  • Drug delivery efficiency needs further investigation.
  • Contribute to the level increase of reactive oxygen species.
  • Similarity in carcinogenic potential between carbon nanotubes and asbestos.
  • Can cause necrosis or apoptosis of macrophage cell lines and changes in cell morphology.
  • Can induce proinflammatory response.
  • Cannot be functionally integrated into the biological systems unless surface functionalization.
[30]
Iron oxide NPsIron oxide nanoparticles, which are also known as super paramagnetic nanoparticlesIron oxideAbility to cross the blood-brain barrier.
Mechanisms of internalization and overall biodistribution are closely associated with their surface chemistry and hydrodynamic sizes.
  • Easy biodegradable (degraded iron can be absorbed by hemoglobin).
  • Non-toxicity.
  • Chemical stability in physiological conditions.
  • Can be visualized by magnetic resonance imaging and it can be used for MR-imaging.
  • Generate heat after exposure to an alternating magnetic field and can be used for magnetic hyperthermia and temperature-triggered drug release
  • The possibility of using passive and active drug delivery strategies.
  • Possibility of chemical modification by coating the iron oxide cores with various shells.
Tend to aggregate into larger clusters.[17,30,31,34]
PLGA—polylactic acid-co-glycolic acid, PACA—poly (alkyl cyanoacrylate), PCL—polycaprolactone, PEI, polyethyleneimine, PLA—poly(lactic acid), PEG—poly(ethylene glycol), RES—reticular endothelial system.
Table 3. Summary of nanotechnology-based systems applied in the treatment of depression, anxiety and schizophrenia.
Table 3. Summary of nanotechnology-based systems applied in the treatment of depression, anxiety and schizophrenia.
DrugNanocarrierComponentsPS, nmZP, mVEE, %LC, %PDIIn Vivo Route of AdministrationOutcomesReferences
Depression
AgomelatinePLGA NPsPLGA, PM407116,06 ± 3523−22.796-<0.3i.n.Prominent antidepressant activity[58]
AgomelatinePolymeric NPsPropylene glycol, Plu F68, Ethanol, Transcutol HP.107.64−1581.91-0.209t.d.Penetration enhancement of drug[59]
BaicaleinSolid lipid NPs.N-Acetyl Pro-Gly-Pro (to binding to neutrophil CXCR2 receptor), GMS, PM188, 1,2-dipalmitoyl-sn-glycero-3-phospho choline.<100−13.5, −12.698.7, 99.1--i.v.The enhanced concentration of drug in the basolateral amygdala. Antidepressant effect in vitro and in vivo.[60]
BDNFPolyion complex formulation of BDNF (nano-BDNF)Human recombinant BDNF, PEG-poly(l-glutamate) diblock copolymer95---0.165i.v.Significant reduction in cerebral tissue loss. Improved memory/cognition, reduced post-stroke depressive phenotypes, and maintained myelin basic protein and brain BDNF levels.[61]
HU-211 and curcuminSolid lipid dual-drug NPsPolyoxyethylene (40) stearate, stearic acid, and lecithin.58.77 ± 1.7−21.7 ± 0.4-0.74 ± 0.02,
18.34 ± 1.06
-i.p.Protection of PC12 cells from corticosterone-induced apoptosis. Antidepressant-like effect and enhanced fall latency in rotarod test, improved level of dopamine in the mice blood. NPs can deliver more curcumin to the brain and thus produce a significant increase in neurotransmitters level in brain tissue, especially in the hippocampus and striatum.[62]
Curcumin and dexanabinolSolid lipid NPsStearic acid, lecithin, polyoxyethylene stearate.-−22.6 ± 0.919.12 ± 1.43,
0.81 ± 0.04
---NPs exerted antidepressant activity by targeting the endocannabinoid/CB1 receptor system.[63]
CurcuminNanocapsulesOrganic phase PCL, capric caprylic triglycerides, 106 sorbitan monoestearate, acetone291–312From −22 to −36Close to 100%--via gavageAntidepressant-like and antioxidant effects in a mouse model of Alzheimer’s disease. NPs were more effective than the free curcumin.[64]
DuloxetineNanostructured lipid carriersGMS, capryol 136 PGMC, Plu F-68, sodium taurocholate80.17–127.73----i.n., i.v.Better brain targeting efficiency than DLX solution when administered intranasally. Decreased side effects.[65]
Folic acidNiosomesDifferent nonionic surfactants (Span 20, Span 60, Span 80, Tween 20, Tween 80, CL)3050–5625-69.42--i.n.Niosomes prepared with span 60 and CL in the ratio of 1:1 (50 mg: 50 mg) showed higher EE and better in vitro drug release of 64.2% at the end of 12 hrs and therefore considered as optimized formulation. About 48.15% of the drug was found to be absorbed through the nasal cavity at the end of 6 hrs.[66]
Lithium carbonateCS nanocompositesCS, sodium tripolyphosphate anions, Tween 80.90.68–220.81+37.987 ± 1.2128.87-p.o.Reversed degenerative changes and gliosis in depression-induced animal models. Fortified targeted drug delivery and restrained adverse effects.[67]
Minocycline hydrochloride (MH)CS NPsTween 80, CS.
Two types of NPs: MH loaded NPs, Tween 80 coated MH encapsulated NPs.
-----i.p.NPs improved the therapeutic efficacy as well as safety of MH.[68]
Selegiline hydrochlorideThiolated CS NPsCS, thioglycolic acid.215 ± 34.71+17.0670 ± 2.71-0.214 ± 0.042i.n.Attenuation of the oxidative stress and restoring of the activity of the mitochondrial complex. Antidepressant-like effect in vivo[69]
SilymarinNanostructured lipid carriersPrecirol solid lipid, Labrafac Lipophile oil.519.00 ± 28.67−12.95 ± 1.5890.00 ± 3.20-0.66 ± 0.05p.o.Antidepressant-like effect, comparable with fluoxetine in mice. Significantly higher brain concentration by 12.46 fold superior to silymarin.[70]
Thymoquinone (TQ)Solid lipid NPsTween 80, GMS, PM188.188.66 ± 8.94−12.32 ± 1.0468.60 ± 4.82-0.319 ± 0.04p.o.Higher amount of TQ reached the target region after administration. Higher levels of monoamines 5 hydroxytryptamine, dopamine and norepinephrine as compared to TQ suspension were demonstrated.[71]
Tramadol HClCS NPs with mucoadhesive thermo-reversible gel.CS, Thermo reversible mucoadhesive Plu-HPMC152.0 ± 9.56+31 ± 2.2185 ± 3.23-0.143 ± 0.003i.n.Antidepressant-like effect in rat model of depression.[72]
Trefoil factor 3 (TFF3)cRGD-modified liposomesCyclic RGD (cRGD) peptide with high affinity for integrin receptors of leukocytes, soybean phosphatidylcholine, DSPE-PEG (PEG 2000), DSPG, CL133−21.827.6--i.p.Brain targeted delivery in a murine model. Antidepressant-like effect of direct intra-basolateral amygdala administration of TFF3 solution in rats subjected to chronic mild stress.[73]
VenlafaxineCS NPsCS glutamate.167 ± 6.5+23.83 ± 1.7679.3 ± 2.632.25 ± 1.630.367 ± 0.045i.n.Enhanced uptake of venlafaxine to the brain.[74]
VenlafaxinePLGA NPsPLGA, two ligands (Tf and TfRp) against transferrin receptor (TfR) to enhance access to brain across BBB.206.3 ± 3.7−2548–5010–120.041 ± 0.017i.n.Plain NPs demonstrated the highest ability to reach the brain vs. functionalized NPs.[75]
VenlafaxineAlginate NPs------i.n., i.v., p.o.Improved antidepressant-like activity in comparison with the venlafaxine (i.n. and oral form). The greater brain/blood ratios for VNPs (i.n.)[76]
Zn(2+)PLGA NPs conjugated with glycopeptidesAntibodies against NCAM1 and CD44, PLGA conjugated with tetramethylrhodamine and glycopeptides.190–210From −0.5 to −10----NPs were able to cross the BBB and to deliver Zn(2+) ions at non-toxic concentrations. Easily modified for preferential targeting of specific cell populations.[77]
Anxiety Disorders
Peptide antisauvagine-30 (ASV-30)Iron oxide NPsFe2O3, 3-aminopropyltriethoxysilane5 ± 1----i.p.Systemically administered NPs were observed in the brain. Association with neurons and reduced amphetamine withdrawal-induced anxiety in rats without affecting locomotion were demonstrated.[78]
BUHThiolated CS NPsCS, thiolated CS.226.7 ± 2.52-81.13 ± 2.849.67 ± 5.5-i.n.The brain concentration achieved after intranasal administration was significantly higher than after administration of BUH (i.v. and i.n.). Excellent bioadhesion.[79]
BUHNanovesiclesTwo types of nanovesicular in situ gels (P407-based and carbopol 974P-based).
Span 20, Span 40, Span 60, CL, stearylamine, dicetylphosphate.
--56.67–70.57--i.n.Higher level of permeability for carbopol formulation in comparison to PM formulations. A 3.26 times increase of BUH bioavailability when loaded into the carbopol nanovesicular in situ gel in comparison to control (i.n.).[80]
DiazepamPLGA NPsPLGA, PVA.250−23.366---In vitro drug release analysis showed sustained release of drug from nanoparticles. Correspondence to Korsmeyer–Peppas model.[81]
Gallic acid (GA)CS NPsCS, Tween 80 (for coating (cGANP batch)).
Two types of NPs: Ligand coated NPs (cGANP) and uncoated NPs (GANP).
103.33 ± 2.28-93.62--i.p.Improved anxiolytic activity in mice. The plasma nitrite level decreased in GA, GANP and cGANP (most pronounced for cGANP) treated group as compared to saline treated control group while no change in plasma corticosterone levels was observed after any treatment.[82]
Schizophrenia and Bipolar Disorders
Aripiprazole (ARP)Solid lipid NPsTristearin, Tween 80, sodium taurocholate.85.05–2042From -4.25 to −17.557.65–81.709.28–31.540.241–0.682p.o.Improved bioavailability of aripiprazole (1.6-fold compared to plain drug suspension). Enhancement of absorption and minimizing of the first-pass metabolism.[83]
AripiprazolePCL NPsPCL199.2 ± 5.65−21.4 ± 4.669.2 ± 2.34--i.n., i.v.The enhanced brain targeting efficiency[84]
Asenapine maleateCS coated nanostructured lipid carrierGMS, oleic acid, Tween 80, glycol CS.184.2 ± 5.59+18.83 ± 1.1983.52 ± 2.59--i.n.2.3 and 4 fold higher systemic and brain bioavailability compared to pure asenapine solution (i.n.).[85]
Asenapine maleateTransfersomesSPC and sodium deoxycholate (SDC).126.0−43.754.96-0.232t.d.Significant increase of bioavailability after transdermal application compared with oral route.[86]
HaloperidolLectin-functionalized, PEG–PLGA NPsMixture of Maleimide-PEG–PLGA and Me-PEG–PLGA, Solanum tuberosum lectin (STL).132 ± 2014.4 ± 0.173.2 ± 0.80.85 ± 0.010.07–0.22i.n.Increase of the brain tissue haloperidol concentrations by 1.5-3-fold compared to non-STL-NPs and other routes of administration.[87]
Lithium carbonateCS NPsCS162–419+ 30----CS increased cellular uptake of lithium.[88]
Lurasidone hydrochloride (LH)Solid lipid NPsGMS, PM188, sodium deoxycholate.139.8 ± 5.5−30.8 ± 3.579.10 ± 2.50--p.o.Improved bioavailability of LH via lymphatic uptake along with improved therapeutic effect in MK-801 induced schizophrenia model in rats.[89]
Lurasidone hydrochlorideNanolipid carrierCapryol 90, Tween 80, and Transcutol P.207.4 ± 1.5-92.12 ± 1.0-0.392 ± 0.15i.n.2-fold increase of the drug concentration in the brain when compared with pure drug solution (i.n.)[90]
LurasidoneMixed polymeric micellesPlu F127, Gelucire 44/14.175-97.8--i.n., i.v.Improved brain distribution as well as kinetics of lurasidone via the intranasal route. The sustained release of lurasidone hydrochloride from micelles with better permeability and bioavailability.[91]
OlanzapineSolid lipid NPsTwo types of NPs with GMS and tripalmitin. Stearylamine, Tween 80.165,
110
+66.50,
+35.3
96.3,
67.2
-0.34,
0.726
i.v.23-fold increased bioavailability of olanzapine in brain and decreased clearance.[92]
OlanzapineNanostructured lipid carriersMucoadhesive NPs was prepared by using carbopol 974P (MNLC (C)) and the combination of PM 407 and of HPMC K4M (MNLC (P + H)).88.95 ± 1.7−22.62 ± 1.988.94 ± 3.9-0,31 ± 0,01i.n.Nose-to-brain delivery of olanzapine MNLC (P + H) was considered as an effective and safe for CNS disorders.[93]
OlanzapineMicellar nanocarriersPlu® mixture of L121 and P123 in different ratios.58.55 ± 2.47-75.03 ± 2.351.84 ± 0.060.27 ± 0.03i.n., i.v.The micelles (i.n.) demonstrated a conciliation between kinetic and thermodynamic stability, controlled drug-release characteristics and evoked minor histopathological changes in sheep nasal mucosa.[94]
OlanzapineNanocapsulesCopolymer-functionalized PCL254.9 ± 12.1+22.2 ± 1.299.00 ± 0.05-0.03 ± 0.01i.n.Controlled release, improved retention of the drug on the nasal mucosa under continuous wash was demonstrated. Increased brain uptake and improved prepulse inhibition deficit induced by apomorphine in rats.[95]
OlanzapineCS NPsCS. NPs were prepared with 20 or 60% loading.208 ± 29
322 ± 18
-86.7 ± 7.1,
87.6 ± 5.2
17.2 ± 1.4,
52.3 ± 3.1
-i.n., i.v.Olanzapine administered via intranasal CS NPs demonstrated the potential to improve the efficacy of systemic absorption thereby offering an efficient method of administration in noncompliant patients.[96]
OlanzapinePLGA NPsPLGA, PM407, acetone, acetonitrile, tetrahydrofuran.91.2 ± 5.2−23.7 ± 2.168.91 ± 2.318.613 ± 0.2880.120 ± 0.018i.n., i.v.6.35 and 10.86 times higher uptake than pure olanzapine solution (i.v. and i.n.) in vivo[97]
PaliperidoneNanolipomersPCL as a polymeric core, Lipoid S75, and Gelucire® 50/13 as a lipid shell and PVA as a stabilizing agent.168.2 ± 0.7−23.187.27 ± 0.098-0.22p.o.Sustained release up to 24 h and better ex vivo intestinal permeation for paliperidone compared to the corresponding polymeric, solid lipid NPs and drug suspension.[98]
Paliperidonein situ GelsCarbopol 934, HPMC K4M, HP-β-CD in the form of inclusion complex of PLPD as nasal absorption enhancer.-----i.n.In vitro and ex vivo drug permeation, exhibited mucoadhesion and sustained drug release. The formulation containing HP-β-CD complex of paliperidone demonstrated higher level of drug permeation through sheep nasal mucosa without injury of nasal mucosa architecture.[99]
Paliperidone palmitateMicellesd-alpha-tocopheryl polyethylene glycol 1000 succinate26.5 ± 4.892.61 ± 2.5---i.m.Improved antipsychotic effect and decreased adverse effects after micellar formulation application.[100]
PaliperidoneMicroemulsionMuco-adhesive polymer, oleic acid27.31 ± 1.86−38.65 ± 2.39--0.241 ± 0.05i.n., i.v.Higher brain paliperidone concentrations compared to pure drug (i.v.)[101]
RisperidoneProteinoid NPsAmino acids l-glutamic acid, l-phenylalanine, l-histidine, poly (l-lactic acid), PEG86 ± 3−16 ± 1-20 ± 0.1-i.v.Enhanced antipsychotic activity compared to pure risperidone in mice[102]
RisperidoneCS NPsCS, tween 80, PM 188.86+36.677.96 ± 1.5013–370.287i.n.Reduced stereotypical behavior score in experimental animals and reversed amphetamine effect.[103]
RisperidoneSolid lipid NPsCompritol 888 ATO, Plu F-127.0.148 ± 0.028−25.35 ± 0.4559.65 ± 1.1859.65 ± 1.180.148 ± 0.03i.n., i.v.Effective brain targeting in mice in vivo[104]
RisperidoneNanoemulsionCS, capmul MCM, transcutol, propylene glycol16.7 ± 1.21−9.15 ± 2.1498.86 ± 1.21-0.191 ± 0.04i.n., i.v.Effective delivery of significant amount of risperidone to the brain after intranasal administration.[105]
RisperidoneFunctionalized liposomesConventional liposomes consisting of SPC/CL, cationic liposomes containing SPC/CL/stearylamine, PEGylated liposomes consists of SPC/CL/distearylphosphatidylethanolamine-mPEG-200098.51 ± 6.82−28.6 ± 3.6258.86 ± 1.38-0.103–0.324i.n., i.v.Liposomal formulations provided enhanced bioavailability, less clearance rate, higher mean residential time and better response in vivo compared to conventional formulations.[106]
RisperidoneCS lipid NPsCS132.7--7.6-i.n., i.v.Increased nose to brain drug delivery compared to pure drug suspension of equivalent dose.[107]
Quetiapine fumarateSolid lipid NPs in situ gelHeat-melting GMS, Span 80, butanol, PM407, PM188.307.1 ± 17.7+ 57.2 ± 0.2497.6 ± 0.58--i.n., i.v., p.o.Stable and effective brain delivery, amelioration of the damages induced by MK-801 in rat model of schizophrenia.[108]
Quetiapine fumarateNanoemulsion systemHLBs of Emalex LWIS 10, PEG 400, Transcutol P, Capmul MCM, Tween 80.144 ± 0.5−8.131 ± 1.8--0.193 ± 0.04i.n.2-fold increase of the drug release compared to pure drug. Better direct nose-to-brain drug transport.[109]
Quetiapine FumarateLiposomeEgg phosphatidylcholine, CL139.6−32.175.63 ± 3.77--i.n.Higher level of diffusion. Better potential to deliver drugs to the brain than by the pure solution.[110]
Quetiapine fumarateMicroemulsion with and without CSCS, methyl-β-cyclodextrin, Capmul MCM EP, labrasol, Tween 80, Transcutol-P.35.31 ± 1.7120.29 ± 1.23--0.249 ± 0.03i.n., i.v.Enhanced brain uptake of quetiapine and improved bioavailability.[111]
Quetiapine fumarateCS NPsCS, sodium tripolyphosphate131.08 ± 7.45+ 34.4 ± 1.8789.93 ± 3.85-0.252 ± 0.064i.n., i.v.Significantly higher brain/blood ratio and 2 folds higher nasal bioavailability in the brain in comparison to pure drug solution (i.n.).[112]
BBB—Blood–brain barrier; NPs—nanoparticles; PS—particle size; ZP—zeta potential; EE—drug entrapment efficiency; LC—loading capacity; PDI—polydispersity index; i.m. —Intramuscular route of administration; i.n.—intranasal route of administration; i.p.—intraperitoneal route of administration; i.v.—intravenous route of administration; p.o. —per oral route of administration; t.d.—transdermal route of administration; BDNF—brain-derived neurotrophic factor; BUH—Buspirone hydrochloride; CL—cholesterol; CS—chitosan; GMS—glyceryl monostearate; HPMC—hydroxypropyl methyl cellulose K4M; PCL—poly(ε-caprolactone); PEG—(poly(ethylene glycol); PLGA—poly(lactic-co-glycolic acid); Plu—pluronic; PM—poloxamer; PVA—polyvinyl alcohol; SPC—soya-phosphatidylcholine.
Table 4. Summary of nanotechnology-based systems applied in the treatment of Alzheimer’s disease.
Table 4. Summary of nanotechnology-based systems applied in the treatment of Alzheimer’s disease.
DrugNanocarrierTargeting LigandComponentsPS, nmZP, mVEE, %LC, %PDIIn Vivo Route of AdministrationOutcomesReferences
Amyloid—Related Nanoparticles
PioglitazoneNano lipid carriers-Tripalmitin, MCM, stearyl amine211.4 ± 3.54+14.9 ± 1.0970.18 ± 4.5-0.257 ± 0.108i.n.The NC significantly improved the nasal permeability of pioglitazone ex-vivo.[152]
BACE1 siRNASolid lipid NPsPeptide derived from rabies virus glycoprotein (RVG-9R)CS-coated and uncoated NPs419.47 ± 24.36, 469.71 ± 49.07−12.52 ± 0.99, +14.47 ± 0.19--0.26 ± 0.09, 0.30 ± 0.04i.n.The siRNA permeated the monolayer (Caco-2) to a greater extent when released from any of the studied formulations than from bare siRNA, and primarily from CS-coated NPs.[153]
siRNA against BACE1Nano complexes CT/siRNA, composed of CGN-PEG-PDMAEMA and Tet1-PEG-PDMAEMA (1:1)CGN peptide, Tet1 peptidePEGylated poly(2-(N,N-dimethylamino) ethyl methacrylate) (PEG-PDMAEMA)70–80+10---i.v.The NC entered the cytoplasm of the neuron cells, inducing effective gene silence (about 50% decrease in BACE1 mRNA levels) and reversing synaptic injury. The NC significantly decreased BACE1 mRNA and the amyloid plaques, suppressed phosphorylated tau protein levels, and promoted hippocampal neurogenesis. NC restored the cognitive performance of the AD transgenic mice to the level of wild-type control.[154]
Plasmid DNA encoding BACE1-AS siRNA, d-peptidePEGylated DGL NPsPeptide from rabies virus glycoprotein (RVG29)Dendrigraft poly-l-lysines 10 (DGL), α-Malemidyl-ω-N-hydroxysuccinimidylpolyethyleneglycol.110+ 7.72 ± 2.80--0.3i.v.Successful codelivery of drugs crossed the BBB. Simultaneous delivery of the therapeutic peptide into brain led to the reduction of neurofibrillary tangles. The memory loss rescue in AD mice was also observed.[155]
BACE1 siRNADendrimer NPsApolipoprotein A-I (ApoA-I), NL4 peptide.Dendrigraft poly-l-lysines, α-Maleimide-ω-N-hydroxysuccinimidylPEG79.26+3.5397.05-0.216-The NC effectively targeted both BBB and PC12 cells and down-regulated BACE1 gene expression in PC12 cells.[156]
CyclophosphamideTheranostic nanovehiclesPutrescine modified F(ab′)2 fragment of antiamyloid antibody, IgG4CS, MRI contrast agent, pentasodiumtripolyphosphate,239 ± 4.111.9 ± 0.5-21,7 ± 1,31-i.v.NPs successfully targeted cerebrovascular amyloid.[157]
Peptide iA5PLGA NPsAnti-transferrin receptor monoclonal antibody (OX26), anti-A (DE2B4)PLGA166 ± 2−13 ± 163 ± 9-0.10 ± 0.04-The uptake of NPs with a controlled delivery of the peptide iA5 was substantially increased compared to the NPs without monoclonal antibody functionalization.[158]
Peptide H102PEG-PLA NPsTGN and QSH peptides.PEG-PLA125.5.10 ± 2.26−29.33 ± 0.1558.49 ± 0.860.540.127. ± 0.010i.v.This dual-functional drug delivery system effectively increased the H102 accumulation at brain Aβ42 concentrated in the hippocampal region and provided better neuroprotective effects in the AD model mice.[159]
-Liposomes and solid lipid NPsPhosphatidicacid, cardiolipinMonosialogangliosides, disialogangliosides, trisialoganglioside, sphingomyelin, CL, diphosphatidylglycerol, 1-palmitoyl-oleoyl-PC, phosphatidylethanolamine, Sephadex G75.145, 76−37.89, −43.30----NPs with surfacephosphatidic acid and cardiolipin demonstrated pronounced affinity to Aβ1e42 fibrils[160]
-LiposomesPeptide from the apolipoprotein-E receptor- binding domain, phosphatedic acid.Dimyristoyl-PA, sphingomyelin, CL.121 ± 7−18.7 ± 4--0.15i.p.Decrease of the total brain-insoluble amyloid peptide was demonstrated. Amelioration of mice impaired memory was observed.[161]
-LiposomesMethoxyXO4DSPE-PEG, CL, DPPC.150----i.v.The NPs appeared to cross the BBB and to bind with Aβ plaque deposits, marketing the parenchymal amyloid deposits and vascular amyloid.[162]
-Apolipoprotein E3 NPsHigh density lipoproteinLipid free ApoE3, DMPC.27.9 ± 8.9−4.07 ± 0.83--0.30i.v.Four-week daily treatment with NPs decreased Aβ deposition, attenuated microgliosis, ameliorated neurologic changes, and rescued memory deficits in an AD animal model.[163]
-Monodisperse iron Oxide NPsMonoclonal antibody against fibrillar human amyloid-β-8----i.v.The targeting ability of NPs to cerebrovascular amyloid was demonstrated.[164]
-Gold NPspeptide CLPFFD, peptide sequence THRPPMWSPVWPGold (III) chloride hydrate.13 ± 1.7−41 ± 2---i.p.Increase of the permeability of the conjugate into the brain was shown.[165]
-Gold NPs-Gold NPs, polyoxometalate with Wells–Dawson structure, Aβ1-40 peptide21.7−36.8---i.v.NPs inhibited Aβ aggregation, dissociated Aβ fibrils and decreased Aβ -mediated peroxidase activity and Aβ -induced cytotoxicity.[166]
-Gold NPs-Gold colloid solution10 ± 2-----NPs disrupted insulin amyloid fibrillation resulting in construction of fibrils that are shorter and more compact.[167]
-PEG-coated cerium oxide NPsAβ antibodyPEG-−37----The rescue of neuronal survival was demonstrated.[168]
Tau–Related Nanoparticles
Nicotin-amideSolid lipid NPs-Stearic acid, phospholipon® 90G, sodium taurocholate.
Three types of particles functionalized by phosphatidic acid, polysorbate 80, phosphatidylserine.
124 ± 0.8−46.1 ± 0.6541.3 ± 0.41--i.p., i.v.The phosphatidylserine-functionalized NPs improved the cognition, preserving the neuronal cells and reducing tau hyperphosphorylation in a rat model of AD.[169]
Methylene bluePLGA NPs-PLGA-b-PEG136.5 ± 4.4-----The reduction of both endogenous and over expressed tau protein levels in human neuroblastoma SHSY-5Y and HeLa cells was shown.[170]
-Protein-capped metal NPs-Two types of NPs: iron oxide (Fe3O4) NPsare capped with hydrolytic proteins from fungi, cadmium sulfide (CdS) NPs are capped with the mixture of four different proteins.10–20-----CdS NPs demonstrated dual properties of inhibition and disaggregation of Tau.[171]
Acetylcholinesterase Inhibitors
GalSolid lipid NPs-Plu F127, Tween 80, glyceryl behenate.92.0 ± 3.51−17.22 ± 1.183.42 ± 0.63-0.380 ± 0.16p.o.NPs restored memory in cognitive deficit rats. Increased bioavailability compared to the plain drug was demonstrated.[172]
GalFlexible liposomes-Propylene glycol, soya PC, CL112 ± 8−49.2 ± 0.783.6 ± 1.8--i.n., p.o.The efficiency of acetylcholinesterase inhibition was greatly enhanced by i.n. administration compared with p.o. administration, especially after loading of the drug in flexible liposomes.[173]
GalNano-emulsions-PLGA21.5 ± 0.25−11.18 ± 0.8998.47 ± 0.4356.87 ± 3.48--Non-cytotoxic drug-loaded NPs have been obtained with high encapsulation efficiencies and sustained drug release, maintaining drug pharmacological activity.[174]
GalCS complex NPs-CS, Tween 80, sodium tripolyphosphate-----i.n.NPs exhibited a significant decrease of AChE protein level and activity in rat brains.[175]
GalCS NPs-CS, sodium tripolyphosphate.190 ± 1.159+ 31.6 ± 9.7523.349.860.276 ± 0.006i.n.NPs were detected in the olfactory bulb, hippocampus, orbitofrontal and parietal cortices 1 h after i.n. administration.[176]
DonepezilLiposomes-CL, PEG, DSPC.102 ± 3.3−28.31 ± 0.8584.91 ± 3.31-0.28 ± 0.03i.n.The bioavailability of drug in the plasma and in the brain increased significantly. The formulation was safe and free from toxicity[177]
DonepezilNano suspension-CS, tripolyphosphate.150–200-92–9640–480.341i.n.No mortality, hematological changes, body weight variations and toxicity in animals were observed.[178]
DonepezilMango gum polymeric NPs-CS, mango gum, Span 80. NPs were prepared by two methods.135. 55, 95.1 2+ 8.2 5, + 25.9 272 ± 3.62, 85 ± 2.1451 ± 5.42, 76 ± 2.260.53, 0.26i.v.The brain targeting was achieved.[179]
DonepezilPLGA-b-PEG nanoparticles-PLGA, PEG, Plu F68.174-240From −11.32 to −20.4952–61-0.20–0.36-NPs crossed the BBB and showed a controlled release profile in this system. NPs administration caused a significant dose-dependent decrease in both gene and protein expression levels of IL-1b, IL-6, GM-CSF and TNF-a.[180]
HupNano structured lipid carriers-Cetyl Palmitate, Miglyol®812, soybean PC, Solutol HS15®120−22.93 ± 0.9189.180.281.460.05--A burst release at the initial stage and followed by a prolonged release of drug from NPs was up to 96 h.[181]
HupMicro emulsion, solid lipid NPs, nano-structured lipid carriers.-Glyceryl monostearate, glyceryl mono-dicaprylate, TranscutolP, Tween 80, triethanolamine.119–148From −4 to −2060.05 ± 5.84-0.3–0.422t.d.In vitro permeation profiles in rat skin exhibited zero-order kinetics. The significant improvement in cognitive function in mice was observed.[182]
HupNanoemulsionLfIsopropyl myristate, Capryol 90, Cremophor EL, Labrasol.15.24 ± 0.67−4.48 ± 0.97--0.128 ± 0.025i.n.Intranasal Lf-nanoemulsion significantly enhanced drug delivery to the brain compared to pure nanoemulsion.[183]
HupMucoadhesive and targeted PLGA NPsLfN-trimethylated CS, PLGA 5050 2A.153.2 ± 13.7+35.6 ± 5.273.8 ± 5.7-0.229 ± 0.078i.n.The NPs facilitated the distribution of drug in the brain.[184]
TacrineIn situ gels-Plu F127, Plu F68, CS, PEG 8000-----i.n.The enhanced nasal residence time, improved bioavailability, increased brain uptake of the drug and decreased exposure of metabolites were shown.[185]
TacrineAlbumin NPs carrying beta cyclodextrin and two beta cyclodextrin derivatives-Beta cyclodextrinderivatives, hydroxypropyl beta cyclodextrin, sulphobutylether beta cyclodextrin.177–266From −10 to −10.985–9112.5–22.00.228–0.327i.n.The presence of the different beta cyclodextrins affected drug loading and differently modulated NPs mucoadhesiveness and drug permeation.[186]
TacrineCS NPs-CS, Polysorbate 8041 ± 7+34.7 ± 1.577–8310,86 ± 0,30-i.v.The NPs coated with 1% Polysorbate 80 altered the biodistribution pattern of NPs.[187]
RivNano structured lipid carriers-Glyceryl monosterate, Capmul MCM C8, lecithin, stearylamine, Tween 80123.2 ± 2.332 ± 1.268.3 ± 3.4--i.n., i.v.The faster regain of memory loss in amnesic mice with 5-fold decrease in escape latency with NC compared to plain rivastigmine solution was shown.[188]
RivLiposomes-1,2-Diacyl-sn-glycero-3-phosphocholine, dihexadecyl phosphate, CL67,51–528.7From −6.6 to −25.175–97-0.612–0.755SubcutaneouslyNPs resulted in faster memory regain and amelioration of metabolic disturbances in AD rats.[189]
RivLiposomes-Soya lecithin, CL10.0 ± 2.8 µm-80.0 ± 5.0--i.n., p.o.Intranasal liposomes demonstrated a longer half-life in the brain than intranasally or orally administered pure drug.[190]
RivLiposomesCell penetrating peptideDSPE-PEG, CL, egg PC178.9 ± 11.7−8.6 ± 2.430.5 ± 8.0-0.333 ± 0.032i.n., i.v.i.n. administration demonstrated the capacity to improve rivastigmine distribution and adequate retention in CNS regions compared to i.v. administration.[191]
RivLiposomes-CL, DPPC, methyl cellulose, dimethyl-β-CD, sodium taurocholate3.37 ± 0.00 µm−4.30 ± 0.6625,2--p.o., i.p.The highest acetylcholinesterase inhibition was observed for rivastigmine-sodium-taurocholate liposomes.[192]
RivPLGA NPs, PBCA NPs-PLGA, PBCA135.6 ± 4.2, 146.8 ± 2.7−23.7 ± 1.18, −3.9 ± 0.58 m74.46 ± 0.76, 57.32 ± 0.91--i.v.The faster regain of memory loss in amnesic mice with both PLGA and PBCA NPs compared to rivastigmine solution was demonstrated.[193]
RivCS NPs-CS, Polysorbate 8045,16 ± 1,5635.08 ± 1. 583.26 ± 3.143.48 ± 1.3 0-i.v.The Polysorbate 80 Coated CS NPs of rivastigmine was formulated and evaluated for brain delivery.[194]
RivCS NPs-CS185.4 ± 8.438.4 ± 2.8585.3 ± 3.543.37 ± 3.90.391 ± 0.065i.n., i.v.NPs demonstrated better brain targeting efficiency and represented a promising approach for i.n. delivery of rivastigmine for the treatment and prevention of AD.[195]
Others
p38 MAPK inhibitor PH797804Nanoemulsion-based CS nanocapsules-Span 85, oleic acid, Tween 20, CS.406.1−18.5 ± 2.921.5 ± 2.73.5 ± 0.50.287i.n.The p38 MAPK inhibitor encapsulated in CS nanocapsules reduces p38 MAPK activity in brain cells in the cerebral cortex and hippocampus to a lower extent.[196]
Erythropoie-tinSolid lipid NPs-Glycerin monostearate, Span 80, Span 60, dichloromethane, Tween 80219.9 ± 15.6−22.4 ± 0.8--(0.187 ± 0.03i.p.NPs reduced the oxidative stress, and beta-amyloid plaque deposition in the hippocampus more effectively than the pure drug. The memory was significantly restored in cognitive deficit rats treated with NPs.[197]
PiperineMonoolein cubosomes-Peceol®, Tween 80, poloxamer, Cremophor.167.00 ± 10.49−34.60 ± 0.4786.67 ± 0.62-0.18 ± 0.01p.o.The NPs significantly enhanced drug cognitive effect and even restored cognitive function to the normal level.[198]
NAPHigh density lipoprotein nanostructureMono sialotetrahexosyl ganglioside (GM1)apoE3-reconstituted high density lipoprotein (rHDL), 1,2-dimyristoyl-sn-glycero3-phosphocholine (DMPC) liposome25.42 ± 1.18−15.70 ± 0.9364.39 ± 12.84-0.218i.n.Protection of neurons from cell toxicity, a reduction of Aβ deposition and a rescue of memory in AD model mice were demonstrated.[199]
NAPPEG-PCL NPsLfPEG-PCL, coumarin-6.88.4 ± 7.823.56 ± 0.9647.61 ± 2.360.62 ± 0.0130.22 ± 0.033i.n.The neuroprotective and memory improvement effect of Lf-NP was observed. These results were also confirmed by the evaluation of acetylcholinesterase, choline acetyltransferase activity and neuronal degeneration in the mice hippocampus.[200]
Metal chelatorclio-quinolGold NP-capped mesoporous silica-Mesoporous silica NPs (N-Cetyl trimethylammonium bromide, tetraethoxysilane)52.13+26.7---i.v.The NPs reduced the cell membrane disruption, microtubular defects and apoptosis.[201]
Selenium, sialic acidSialic acid modified selenium NPsPeptide-B6Sodiumselenite,95−14.4----The high permeability across the BBB was shown. The effectively inhibition Aβ aggregation was demonstrated. Therefore, NC did not disaggregate preformed Aβ fibrils into non-toxic amorphous oligomers.[202]
Resveratrol, grape extractSolid lipid NPsAntitransferrin receptor mono clonal antibody (OX26 mAb)Cetylpalmitate, polysorbate 80.254 ± 17−4.0 ± 0.192 ± 7, 95 ± 2-0.23 ± 0.05-The cellular uptake of the OX26 NPs was substantially more efficient than that of normal NPs and NPs functionalized with an unspecific antibody.[203]
Resveratrol (Res)Delivery system MSe-Res/Fc-β-CD/Bor-Mesoporous nano-selenium (MSe), β-CD, borneolBor)160->70--i.v.NC inhibited aggregation of Aβ, mitigated oxidative stress, suppressed tau hyperphosphorylation and improving memory impairment in AD mice.[204]
MemantinePolyamido amine (PAMAM) dendrimersLfPAMAM131.72 ± 4.73+ 20.13 ± 0.9471.1 ± 4.84-0.16 ± 0.025i.v.A significant improvement in behavioral responses was demonstrated.[205]
Lipoyl–memantineSolid lipid NPs-Stearic acid170−33.888-0.072Through the gastrointestinal tractThe NPs demonstrated low toxicity.[206]
AmlodipineDiamond NPs-Nanodiamond powder31.1 ± 8.2-41---The highest percentage of loaded amlodipine onto nanodiamond particles was achieved in alkaline medium using 2 mMNaOH at a corresponding pH of 8.5.[207]
CurcuminLow density lipoprotein mimic nanostructured lipid carrierLfCarboxylated PEG (100) monostearate, CL, glycerol trioleate103.8 ± 0.6−5.80 ± 0.7396.51 ± 1.872.60 ± 0.170.15 ± 0.022i.v.The Lf NPs could effectively permeate BBB and preferentially accumulate in the brain. Superior efficacy of Lf NPs in controlling the damage associated with AD.[208]
CurcuminRBC membrane camouflaged HSANPsT807, tri phenyl phosphine (TPP)Amino-T807, carboxy-TPP, PEG, red blood cell (RBC) membrane, human serum albumin (HAS).<120-88.43 ± 1.254.87 ± 1.04-i.v.The NPs relieved AD symptoms by mitigating mitochondrial oxidative stress and suppressing neuronal death.[209]
Curcumin, NGFLiposomesWheat germ agglutinin (WGA), cardiolipin (CL)CL, soybean PC, 1,2-Dipalmitoylsn-glycero-3-phosphocholine.122–142−5.2 to −18.320.5–56.7--i.v.The NPs inhibited the expression of phosphorylated p38, prevented neurodegeneration of cells, enhanced the quantities of p-neurotrophic tyrosine kinase receptor type 1 and p-extracellular signal-regulated kinase 5.[210]
Curcumin, seleniumPLGA nanospheres-PLGA, selenium NPs160 ± 5--11.5Lowi.v.The NPs provided the enhanced therapeutic efficacy in AD lesions.[211]
Curcumin or dexamethasonePolymeric nanocoreAntiamyloid antibody IgG4.1CS, gadolinium-diethylene triamine pentaacetic acid, hydroxypropyl-beta-cyclodextran, Magnevist®, 125I145 ± 5,4, 157,6 ± 3,47,7 ± 0,4, 4,5 ± 0,5---i.v.The NPs targeted cerebrovascular amyloid deposits selectively.[212]
AndrographolideHuman albumin NPs-Human albumin210.4 ± 3.2−20.3 ± 1.599.1 ± 0.2-0.10 ± 0.01i.v., i.p.In the step-down inhibitory avoidance test, NPs improved mice performance. The presence of NPs both in the pE3-Aβ plaque surroundings and inside the pE3-Aβ plaque was observed. The anti-inflammatory activity was shown.[213]
FlurbiprofenDendrimer NPs-Phenylalanine------Efficiency of NC influence on the γ-secretase enzyme in target cells was shown. Eventual drug release by hydrolysis of the carrier was demonstrated.[214]
Tarenflurbil (TFB)Solid lipid NPs, PLGA NPs-Monostearate, stearic acid, soya lecithin, Tween 20, PLGA, DMAB, PF-68, PVA169.87 ± 10.98, 133.13 ± 7.82−23.13 ± 2.32, −30.25 ± 2.1157.81 ± 5.32, 64.11 ± 2.21−100.24 ± 0.04, 0.21 ± 0.02i.n.The absolute bioavailability of the NPs was higher than TFB solution suspension.[215]
MetforminNano liposomes-Phosphatidylserine.148.3 ± 5.6−34.8 ± 2.837 ± 2.3-<0.3i.p.The learning and memory parameters significantly improved in AD-rats treated with NPs. The decreased cytokine levels of IL1-β, TNF-α and TGF-β in hippocampal tissues were shown. A reduction in inflammatory and necrotic neural cells, and an increase neurogenesis was demonstrated.[216]
SaxagliptinCS-l-valine based NPsl-valinCS, BocL-valine385 ± 21+ 0.554 ± 0.11056.23 ± 13.44-0.574 ± 0.125i.p.The NPs were highly stable in the plasma releasing only a minute after administration of the drug. Pronounced accumulation of drug from the NPs was demonstrated. The pure substance showed no detectable amount of the drug after 24 h.[217]
NGFLiposomesLfCL, DSPE-PEG, DPPC, PEPEGAbout 110From −3 to −11----Lf/NGF-liposomes comprising CL and DPPC were physically stable with high biocompatibility to HBMECs and HAs cells.[218]
NGFLiposomesP-aminophenyl-a D manno-pyrano-side, apolipo protein ESoybean PC, phosphatidic acid, DPPC, cardiolipin, DSPE-PEG, CL<170From −4 to −1035.1 ± 3.2---NPs were capable to enhance the NGF delivery across the BBB. NPs recognized a low-density lipoprotein receptor expressed by SK-N-MC cells and yielded neuroprotective effect on the neuronal degeneration induced by fibrillar Aβ1–42.[219]
NGFLiposomesCereport trans-ferrinl-α-PC, DPPC, CL, DSPE-PEG-carboxy.152-189from −4.5 to −8.531.2 ± 2.8---Covering of NPs with cereport and transferrin was effective in carrying NGF across the BBB and rescued the apoptosis of SK-N-MC cells after neurotoxic treatment with Aβ1–42.[220]
Fibroblast growth factorPEG-PLGA NPsSolanum tubero-sum lectinPEG-PLGA118.7−31,1869.210.0462-i.n.Neuroprotective effect in AD rats was demonstrated.[221]
Pituitary adenylate cyclaseactivating
Polypeptide (PACAP38) coupled to a docosahexaenoic acid (DHA: an ω-3 polyunsaturated fatty acid)
Pep-lipid nanostructures and liquid crystalline nanocarriers-PACAP-DHA/MO (nonlamellar lipid monoolein)/DHA/vitamin E/VPGS-PEG1000 and MO/DHA/vitamin E/VPGS-PEG1000~100-----Multicompartment nanocarriers with PACAP and DHA are promising therapy strategies for neurodegenerative disorders.
PACAP is neuropeptide ligand of the PAC1 membrane receptor (a class B GPCR). PACAP demonstrate antiapoptotic effects in neuronal and non-neuronal cells, neuroprotive and neurotransmitter properties, can act as neurotrophic factor, attenuates Aβ amyloid toxicity.
DHA plays a crucial role in the control of the phospholipid membrane organization, act as lipid trophic factor, demonstrated antiapoptotic, neuroprotective effect, take part in inflammatory cell signaling.
Vitamin E reduces oxidative stress, demonstrates anti-inflammatory, cardio and neuroprotective effects, inhibit lipid peroxidation.
[222]
-Amphiphilic yellow-emissive carbon dots (Y-CDs)-Citric acid, o-phenylenediamine.3.4 ± 1.0−15.3---Injected into the heart.Y-CDs entered cells to inhibit the overexpression of human amyloid precursor protein and β-amyloid.[223]
-Nanosystem CB-Gd-Cy5.5Holera toxin B subunit (CB)Chelated gadolinium (Gd), Cyanine5.5 (Cy5.5).110−11.0---i.n.The nanosystem was accumulated in the hippocampus and demonstrated good magnetic resonance imaging capability satisfying the monitoring of AD at the different stages.[224]
-Fe3O4 NPs loaded with PEG-PLGA nano compositeAnti-transferrin mono clonal antibody (OX26) receptorFe3O4NPs, PEG, PLGA.95-76.218.1--The significant in vitro drug release and cell viability were shown.[225]
-Carboxyl magnetic nano containers-Fluorescent Carboxyl Magnetic Particles, Yellow, 1% w/v700–900----i.v.The magnetic NPs crossed the normal BBB in mice after subjection to external electromagnetic fields of 28 mT (0.43 T/m) and 79.8 mT (1.39 T/m).[226]
-Gold NPs-Gold NPs suspension5−47.7 ± 10.9---i.p.The NPs improved the acquisition and retention of spatial learning and memory in Aβ treated rats. Expression of BDNF, cAMP, CREB and stromal interaction molecules, e.g., STIM1 and STIM2 was increased.[227]
AD—Alzheimer’s disease; Aβ—amyloid-β peptide; BBB—Blood–brain barrier; NC—nanocarrier; NPs—nanoparticles; PS—particle size; ZP—zeta potential; EE—drug entrapment efficiency; LC—loading capacity; PDI—polydispersity index; i.m.—Intramuscular route of administration; i.n.—intranasal route of administration; i.p.—intraperitoneal route of administration; i.v.—intravenous route of administration; p.o.—per oral route of administration; t.d.—transdermal route of administration; BDNF—brain-derived neurotrophic factor; CL—cholesterol; CS—chitosan; DPPC—1,2-dipalmitoyl-sn-glycero-3- phosphocholine; DMPC—1,2-dimyristoyl-sn-glycero-3-phosphocholine; DSPC—1,2-distearyl-sn-glycero-3-phosphocholine; DSPE-PEG—1,2-distearoyl-sn-glycero-3- phosphoethanolamine-N-[carboxy(polyethylene glycol); Gal—galantamine hydrobromide; Hup—huperzine A; Lf—lactoferrin; NAP—neuroprotective peptide NAPVSIPQ; NGF—neuron growth factor; PC—phosphatidylcholine; PEG—(poly(ethylene glycol); PEG-PCL—poly(ethylene glycol)-co-poly(ε-caprolactone) copolymer; PEPEG—1,2-dipalmitoyl-sn-glycero-3-phosphoethanol- amine-N-[methoxy(polyethylene glycol)-2000]; PLA—poly(lactic acid); PLGA—poly(lactic-co-glycolic acid); PVA—polyvinyl alcohol; siRNA—small interfering RNAs; Plu—pluronic; Riv—rivastigmine;.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zorkina, Y.; Abramova, O.; Ushakova, V.; Morozova, A.; Zubkov, E.; Valikhov, M.; Melnikov, P.; Majouga, A.; Chekhonin, V. Nano Carrier Drug Delivery Systems for the Treatment of Neuropsychiatric Disorders: Advantages and Limitations. Molecules 2020, 25, 5294. https://doi.org/10.3390/molecules25225294

AMA Style

Zorkina Y, Abramova O, Ushakova V, Morozova A, Zubkov E, Valikhov M, Melnikov P, Majouga A, Chekhonin V. Nano Carrier Drug Delivery Systems for the Treatment of Neuropsychiatric Disorders: Advantages and Limitations. Molecules. 2020; 25(22):5294. https://doi.org/10.3390/molecules25225294

Chicago/Turabian Style

Zorkina, Yana, Olga Abramova, Valeriya Ushakova, Anna Morozova, Eugene Zubkov, Marat Valikhov, Pavel Melnikov, Alexander Majouga, and Vladimir Chekhonin. 2020. "Nano Carrier Drug Delivery Systems for the Treatment of Neuropsychiatric Disorders: Advantages and Limitations" Molecules 25, no. 22: 5294. https://doi.org/10.3390/molecules25225294

Article Metrics

Back to TopTop