Next Article in Journal
Microneedle-Assisted Transfersomes as a Transdermal Delivery System for Aspirin
Next Article in Special Issue
Maltodextrin-Nanoparticles as a Delivery System for Nasal Vaccines: A Review Article
Previous Article in Journal
Maximizing Anticancer Response with MPS1 and CENPE Inhibition Alongside Apoptosis Induction
Previous Article in Special Issue
The Efficacy and Safety of Intranasal Formulations of Ketamine and Esketamine for the Treatment of Major Depressive Disorder: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Intranasal Drug Delivery by Nanotechnology: Advances in and Challenges for Alzheimer’s Disease Management

1
Department of Pharmaceutics, SVKM’s Dr. Bhanuben Nanavati College of Pharmacy, University of Mumbai, Mumbai 400056, India
2
Indoco Remedies Private Limited, Mumbai 400098, India
3
The Novel Drug & Vaccine Delivery Systems Facility, Department of Chemistry and Biochemistry, Laurentian University, Sudbury, ON P3E 2C6, Canada
*
Authors to whom correspondence should be addressed.
Pharmaceutics 2024, 16(1), 58; https://doi.org/10.3390/pharmaceutics16010058
Submission received: 18 September 2023 / Revised: 11 October 2023 / Accepted: 12 December 2023 / Published: 29 December 2023

Abstract

:
Alzheimer’s disease, a progressive neurodegenerative condition, is characterized by a gradual decline in cognitive functions. Current treatment approaches primarily involve the administration of medications through oral, parenteral, and transdermal routes, aiming to improve cognitive function and alleviate symptoms. However, these treatments face limitations, such as low bioavailability and inadequate permeation. Alternative invasive methods, while explored, often entail discomfort and require specialized assistance. Therefore, the development of a non-invasive and efficient delivery system is crucial. Intranasal delivery has emerged as a potential solution, although it is constrained by the unique conditions of the nasal cavity. An innovative approach involves the use of nano-carriers based on nanotechnology for intranasal delivery. This strategy has the potential to overcome current limitations by providing enhanced bioavailability, improved permeation, effective traversal of the blood–brain barrier, extended retention within the body, and precise targeting of the brain. The comprehensive review focuses on the advancements in designing various types of nano-carriers, including polymeric nanoparticles, metal nanoparticles, lipid nanoparticles, liposomes, nanoemulsions, Quantum dots, and dendrimers. These nano-carriers are specifically tailored for the intranasal delivery of therapeutic agents aimed at combatting Alzheimer’s disease. In summary, the development and utilization of intranasal delivery systems based on nanotechnology show significant potential in surmounting the constraints of current Alzheimer’s disease treatment strategies. Nevertheless, it is essential to acknowledge regulatory as well as toxicity concerns associated with this route; meticulous consideration is required when engineering a carrier. This comprehensive review underscores the potential to revolutionize Alzheimer’s disease management and highlights the importance of addressing regulatory considerations for safe and effective implementations. Embracing this strategy could lead to substantial advancements in the field of Alzheimer’s disease treatment.

Graphical Abstract

1. Introduction

Alzheimer’s disease represents an advanced neurodegenerative condition characterized by compromised cognition, challenges in daily tasks, and difficulties related to learning, speech, and language [1,2]. Projections indicate that by 2050, dementia will impact over 100 million individuals worldwide, with associated costs estimated to escalate to USD 1 trillion in the coming years. Dementia, a prominent manifestation of Alzheimer’s disease, displays age-related progression, doubling approximately every five years past the age of 65 and increasing by about 50% beyond the age of 85. The distinctive molecular features of Alzheimer’s disease encompass the accumulation of Aβ, leading to the formation of senile plaques, excessive tau phosphorylation resulting in neurofibrillary tangles (NFTs), compromised glial function, neuronal inflammation, and irregularities in vascular activity [3,4].
It is also recognized as a protein-conformational disorder (PCD), as the misfolding of neuronal proteins leads to altered conformations that transform soluble forms into insoluble aggregates [5]. AD is acknowledged as a multifactorial ailment, yet current knowledge of the disease highlights NFTs and Aβ plaques as primary contributors to its onset and progression [6]. For decades, research efforts have been directed toward unravelling the biology and mechanisms of the Aβ peptide in AD’s pathogenesis [7]. To simplify the intricate pathology, various hypotheses such as the amyloid cascade hypothesis, tauopathies, and the cholinergic hypothesis [8,9,10,11,12] have been proposed by investigating the disease at both the cellular and molecular levels [13]. Additionally, a mounting body of evidence supports the substantial role of oxidative stress [14,15], neuroinflammation [16], neuron-associated astrocytes, and metal ions such as aluminium in the initiation and advancement of AD [17,18,19,20,21,22]. Brain imaging studies utilizing PET scans in AD patients have revealed heightened levels of activated microglia [23,24,25], along with inflammatory cytokines. Moreover, research has demonstrated that Aβ activates the innate immune response [26,27]. Similarly, dysregulated glutamatergic signaling and the hyperactivation of NMDA receptors result in calcium dysregulation, which is one of the underlying mechanisms that causes AD to progress. Among all of these discoveries, the cholinergic hypothesis and the role of NMDA receptors marked a significant breakthrough in Alzheimer’s disease research, as they form the basis of current conventional pharmacological treatments for AD.
The current available treatments for AD can be categorized into pharmacological interventions targeting altered disease-related neurotransmitters (e.g., acetylcholinesterase inhibitors (AChEIs)such as galantamine and N-Methyl-D-aspartate receptor (NMDA) antagonists such as memantine) and non-pharmacological strategies primarily focusing on behavioural aspects [28]. The elevated level of AChEs in the brains of people with AD prompted researchers to identify AChEIs to substantiate their cholinergic activity, yet research has underscored the significance of both AChE and BuChE in the progression of AD [29]. As a result, there are two categories of cholinesterase inhibitors: non-specific inhibitors that act on both AChE and BuChE; and specific inhibitors that target acetylcholinesterases exclusively or are classified based on the degree and type of inhibition, such as reversible (donepezil, galantamine), irreversible, and pseudo-irreversible inhibitors (rivastigmine).Given the multifaceted nature of the disease, tackling its progression or achieving a cure with a single therapeutic agent is challenging. Consequently, numerous investigations have explored combinations of AChEIs with other agents, such as choline precursors, NMDA antagonists [30], and antioxidants [29]. In this context, several preclinical studies have demonstrated synergistic activity when combining donepezil (AChEI) actions by inhibiting AChEs and memantine (an NMDA antagonist), which execute anti-Alzheimer’s disease action by regulating the Ca2+ influx, glutamanergic signalling, etc., leading to overall improved cognition [31,32]. Based on substantial evidence, a fixed-dose combination of donepezil and memantine, known as Namzaric TM, received approval from the FDA in 2014 [33,34]. Despite promising results in providing symptomatic care, these medications have shown inconsistent effects as disease-modifying therapies. Moreover, they can induce serious side effects, such as nausea, diarrhoea, dizziness, and appetite loss [35].
Therefore, the pursuit of novel treatments that alter the course of the disease is currently a top global research priority. The undeniable role of Aβ plaques and tau proteins in the pathology of AD has led research efforts to predominantly focus on these as unique targets for disease-modifying therapies [36,37]. A significant breakthrough in AD research occurred with the recent FDA approval (2021) of the first disease-modifying monoclonal antibody, aducanumab (Aduhelm®) [38], which targets Aβ plaques, including both insoluble fibrils and oligomers [39]. In a double-blinded clinical trial, a 1-year infusion of aducanumab demonstrated a controlled reduction in Aβ plaques based on dosage and time [40].These findings were supported by two Phase 3 randomized trials, ENGAGE and EMERGE [41]. However, the accelerated approval of aducanumab was controversial due to safety concerns, and serious side effects such as the development of ARIA, brain oedema, microhaemorrhages, and vertigo, etc., led to its initial disapproval [42,43,44,45]. Despite these concerns, the drug was eventually re-approved as no fatalities were reported. Further, the US FDA mandates post-approval clinical trials to validate the anticipated benefits of aducanumab [46]. Likewise, two humanized monoclonal antibodies, lecanemab (Leqembi®) [47] and gantenerumab, obtained FDA approval in 2023 [48].Both of these antibodies demonstrated a high binding affinity to Aβ protofibrils, a potential reduction in Aβ burden, and the deceleration of disease progression in early-stage patients [49]. Further biweekly infusions of lecanemab in Phase 2 trials showcased a time-dependent attenuation of ARIA, with more pronounced occurrences in the ApoE4-positive homozygous population [50,51].In the latter case of gantenerumab, two separate Phase 3 trials (SCarlet RoAD and Marguerite RoAD) were conducted to assess the safety profile and therapeutic efficacy of low-dose subcutaneous gantenerumab [52], and an open-label extension (OLE) study was performed at an escalated dose(up to 1200 mg), which revealed a significant Aβ reduction [53]. Currently, a randomized, double-blind Phase 3 trial, GRADUATE I and II, is underway to evaluate the safety and efficacy of subcutaneous gantenerumab compared to a placebo in early AD populations [54].
Despite promising preclinical results, many Aβ-directed therapies have failed to show efficacy in clinical trials [55]. Consequently, research has shifted towards exploring other potential targets, such as tau proteins and neuroinflammation. This shift has led to the investigation of a wide array of immunotherapies targeting Aβ fibrils and tau proteins for AD treatment, some of which are enumerated in Table 1. However, the efficacy of anti-tau therapy is influenced by factors like the mode of action, existing tau forms, and the epitope and form of tau that spreads to other cells [56]. A growing body of studies indicate discrepancies between pathogeneses, disease severity, and diagnoses, which impacts the success of treatments. Furthermore, the chosen approach for delivering therapeutics to the brain is a pivotal determinant in the success of immunotherapy [57]. Consequently, achieving an efficient and safe delivery of both conventional approved therapeutics and immunotherapies remains a formidable challenge in AD treatment.
Recent advancements in nanotechnology have positioned it as a promising domain for brain targeting, and numerous studies have demonstrated its potential in managing Alzheimer’s disease (AD) [66]. Furthermore, a variety of factors including physiological barriers, brain anatomy, and physicochemical properties significantly impact the therapeutic effectiveness of conventional anti-AD drugs [67]. Thus, adopting a nanocarrier-based delivery approach holds promise for enhancing the efficacy of existing treatments [68]. Loaded with drugs, these nanocarriers elevate the drug concentration in the brain, thus reducing the required dosage and associated side effects [69]. Additionally, nanomedicines contribute to improved stability, biocompatibility, biodegradability, reduced toxicity, an extended half-life, controlled release, and the enhanced solubility of poorly soluble drugs [70].Nanocarriers follow various transport mechanisms to traverse the blood–brain barrier (BBB), including simple diffusion, transcytosis, receptor-mediated endocytosis, and exocytosis [71]. NP diffusion is facilitated through two mechanisms: the first involves stimuli (generated by the “nano-effect” or bioactive substances adsorbed on NP surfaces) mediating the transient opening of tight junctions, followed by diffusion. The second mechanism entails NP adsorption on endothelial cell surfaces, leading to drug release, the creation of a concentration gradient, and the subsequent promotion of diffusion [72]. Moreover, lipid nanoparticles with small molecular weights (<400 Da) and sizes (<100 nm) can effortlessly diffuse through the BBB due to their inherent lipidic nature [73]. Furthermore, active targeting through receptors can be achieved by modifying the surface of nanocarriers with various ligands, such as peptides, polysaccharides, antibodies, and more [74]. This enables tailored nanocarrier systems to achieve specific tissue accumulation in the brain through passive or active targeting mechanisms [75].In spite of such phenomenal characteristics when administered via the conventional route, only 5% of the dose reaches the brain while the remaining 95% accumulates in non-targeted/peripheral tissues, causing potential toxicity to the reticuloendothelial system, etc. Hence, research pipelines have tended toward exploring novel strategies to improve the delivery of nanocarriers to intricate organs, including the brain [76,77,78]. In recent years, intranasal drug delivery has surfaced as a non-invasive, safer, and efficacious alternative to traditional routes of brain targeting [79]. Figure 1 presents various intranasal treatment approaches for Alzheimer’s disease management based on nanocarriers. The potential of the intranasal route of brain targeting is exceptional and can be attributed to unique olfactory and trigeminal pathways that provide direct access to the brain. However, there still exist some anatomical and structural challenges associated with the IN route, e.g., limited volume, mucociliary clearance, etc., which affect the targeting potential [80]. One of the ground-breaking strategies to overcome the aforementioned challenges is integrating nanoscale carriers with the intranasal route of brain targeting. Several studies have demonstrated that nanocarriers administered via the IN route accumulate in a higher concentration at the olfactory bulb and pons, suggesting nanocarriers can readily transverse across the BBB via the intranasal route [80]. Due to the significance of intranasal nanocarriers in brain targeting, diverse polymer-, lipid-, and metal-based carriers have been explored for managing AD [81]. Figure 2 illustrates the trajectory of the delivery system after transport through distinct intranasal pathways.

2. Exploring Nanocarriers for Alzheimer’s Disease Therapy

2.1. Polymeric Nanoparticles

A burgeoning and innovative approach to delivering therapeutics to the brain in the context of Alzheimer’s disease (AD) involves the utilization of polymeric nanoparticles. These nanoparticles can be synthesized from monomers or polymers using various polymerization methods [82]. The physicochemical properties of polymeric nanoparticles can be customized according to their intended application. For AD, a diverse range of synthetic polymers (such as PACA and PLGA), natural polymers (including chitosan and alginate), and hybrid polymers have been employed [83]. Diverging from vesicular carriers like liposomes and micelles, polymeric nanoparticles offer distinct advantages such as enhanced stability, reduced drug exposure, and tuneable properties achievable through composition and structural modifications [84]. The traversing of nanoparticles across the blood–brain barrier (BBB) can be facilitated by functionalizing them with ligands, which can occur through various mechanisms: 1. absorbing macromolecules from the bloodstream, enabling interaction with specific receptors (e.g., tween 80) [85]; 2. direct binding to receptors (e.g., lactoferrin) [86]; 3. increasing hydrophobicity and charge (e.g., amphiphilic peptides) [87]; and 4. prolonging circulation time (e.g., PEG) [88]. Additionally, absorptive-mediated transcytosis can be promoted by attaching cationic peptides to the surface of nanoparticles or using cationic polymers (e.g., chitosan) [89]. These cationic nanoparticles engage in electrostatic interactions with negatively charged capillary endothelial cells, facilitating adsorptive-mediated transcytosis transport [90]. However, the exact transport mechanism of nanoparticles remains incompletely understood, and the influence of physicochemical properties on transport remains to be fully elucidated [91].
Upon successful transport, it is subsequently crucial to consider the mechanism of drug release from the carrier. The predominant mechanisms through which polymeric systems achieve controlled release encompass drug diffusion through aqueous pores, matrix diffusion, osmotic-driven release, and erosion mechanisms. Several factors, including the molecular weight, mechanical strength, solubility, nature of the polymer, and glass transition temperature (Tg), affect the drug release profile from polymeric nanoparticles [92]. Various polymers are being investigated for effectively targeting different anti-AD agents, as summarized in Table 2. Despite promising outcomes, clinical applications of polymeric nanoparticles face challenges posed by oxidative stress, cytotoxicity, and genotoxicity, often linked to the quantum dimensions of the nanoparticles [93].

2.2. Lipid-Based Nanocarriers

Lipid-based nanocarriers present an innovative avenue for brain targeting, attributed to their lipophilic nature, biocompatibility, biodegradability, and their ability to bypass P-glycoprotein (P-gp) efflux [98]. A significant advantage of lipid nanocarriers lies in their ability to tailor structural properties based on the physicochemical attributes of small drug moieties and excipients. Moreover, the incorporation of lipids as fundamental constituents contributes to achieving distinct controlled release and non-toxic degradation products, in contrast to polymeric nanoparticles, which often exhibit an initial burst release, instability, and toxicity of degradation products [99]. The ease of preparation, avoidance of first-pass metabolism, reduced use of organic solvents, and potential for scale-up further elevate the appeal of lipid nanoparticles over polymeric alternatives [100]. Prominent among lipid nanocarriers for brain targeting are solid lipid nanoparticles (SLNs), nanostructured carriers, and liposomes, largely due to their capacity to circumvent the BBB [101]. Table 3 summarizes various lipid based nanocarriers that have been investigated for targeting drug for effective therapy of Alzheimer’s disease.
Liposomes, a lipid-based vesicular nanocarrier, have versatile applications, including gene delivery, therapeutic administration, and nucleic acid delivery to the brain [109]. Liposomes serve as ideal carriers for gene delivery, benefiting from the incorporation of ionizable or fusogenic lipids, which enhance endosomal escape, target specificity, diminish immunogenicity, and extend circulation time [110]. However, the presence of lipids in liposomes yields a dual-edge characteristic, conferring biocompatibility while also increasing susceptibility to peroxidation and leakage, leading to compromised stability and shelf life. Moreover, challenges such as limited drug loading and entrapment efficiency hinder their clinical application [111]. Therefore, research efforts are directed towards enhancing the stability of existing liposomes and devising novel carriers with an improved stability.
Solid lipid nanoparticles, as the first generation of lipid nanocarriers, were designed to surmount the limitations of liposomes by utilizing lipids to replace the aqueous core, thereby preventing active drug interactions [112]. SLNs also possess the ability to evade the brain’s reticuloendothelial system [113]. The choice of surfactants significantly impacts SLNs’ formation, influencing their particle size, distribution, and targeting efficiency [114]. Some studies have demonstrated increased brain uptake with surfactant-coated SLNs, notably Polysorbate-80 coating, possibly due to the stimulation of endocytosis by transporters such as apolipoprotein E present at the BBB [115]. Additionally, coating SLNs with cationic polymers like chitosan has been shown to enhance drug loading, overcome initial burst release, and improve stability [116]. SLNs have been extensively investigated to enhance bioavailability, BBB transport, and brain targeting for AD management [117].
While SLNs have demonstrated broad applications in brain targeting, nanostructured lipid carriers (NLCs) are preferred from a formulation perspective, offering a high payload due to their imperfect structure, enhanced stability, and reduced risk of drug expulsion [118]. In addition to improved BBB transport, NLCs exhibit a high affinity for Aβ plaques, followed by degradation [119]. NLC surfaces can be tailored through surfactants and ligands like lactoferrin for active targeting [120]. Unlike solid lipid carriers, nanostructured carriers exhibit a dual release mechanism, involving an initial rapid release followed by a sustained release. This characteristic is advantageous for brain targeting [76]. The literature indicates that NLCs can enhance the pharmacokinetic properties and therapeutic efficacy of various anti-AD agents, including donepezil, rivastigmine, antioxidants (e.g., ubiquinone), and ECGCs [121,122,123].
Although lipid-based nanocarriers, particularly LNPs, hold substantial promise for brain targeting, challenges remain in terms of scale-up due to issues such as instability, polymorphism, aggregation, safety concerns, and sterilization-related problems [124].
Nanoemulsions, a biphasic emulsion system, exhibit broad applicability in enhancing bioavailability and targeting across various administration routes [125]. Nanoemulsions offer advantages over microemulsions, such as maintaining globule size regardless of dilution or temperature changes and achieving spherical and smaller globule sizes (<200 nm) [126]. The narrow particle size distribution and inherent lipid nature of nanoemulsions contribute to improved brain uptake across the BBB. Furthermore, they enhance drug stability against degradation, ultimately reducing the required dose and associated side effects [127]. Notably, conventional anti-Alzheimer’s disease drugs like memantine have been delivered to the brain using nanoemulsion formulations, demonstrating enhanced brain uptake with a sustained release of up to 80% [128]. The functionalization of nanoemulsions with ligands, such as shuttle peptides, can further augment uptake and contribute to active targeting [129].

2.3. Metal Nanoparticles

Metal nanoparticles have garnered substantial interest due to their distinctive physicochemical attributes and their potential for theragnostic applications in Alzheimer’s disease (AD) management [130]. Various metal nanoparticles, such as gold nanoparticles, silver nanoparticles, iron nanoparticles, and more, have been explored for their anti-Alzheimer’s effects. An intriguing aspect of metal nanoparticles is their inherent ability to permeate the blood–brain barrier (BBB) without requiring additional functionalization, primarily achieved through endocytosis involving both pinocytosis and phagocytosis mechanisms [131,132].
Among these, gold nanoparticles (AuNPs) have captured significant attention owing to their exceptional optical properties, electrical conductance, enhanced stability, and low toxicity. They have demonstrated the potential to counteract memory impairment, as well as inhibit and disaggregate Aβ aggregates [133]. The anti-Aβ properties of AuNPs are influenced by their physicochemical characteristics, such as their size, shape, and charge [134]. Some studies have revealed that rod-shaped, cationic gold nanoparticles exhibit a superior binding affinity to Aβ plaques compared to cube-shaped, anionic gold nanoparticles [135]. Additionally, selenium nanoparticles have demonstrated neuroprotective effects attributed to their reduced toxicity and antioxidant properties [136].
However, while metal nanoparticles exhibit promises as theragnostic tools for AD, studies have also highlighted significant toxicity associated with certain metal nanoparticles, like mercury, aluminium, and copper, and their potential correlation with AD pathogeneses [137]. The primary mechanism underlying this toxicity involves the generation of oxidative stress which damages macromolecules and cells [138]. Consequently, efforts are being directed toward mitigating metal toxicity through various approaches, such as the biogenic synthesis method [139].
Recently, a fusion of metals and organic ligands has led to the formation of “metal-organic frameworks” (MOFs), which offer biocompatibility, stability, improved delivery efficiency, and diagnostic applications [140]. Numerous studies have investigated the role of metal nanoparticles in enhancing brain targeting for AD management, with some of these studies summarized in Table 4.
Advancements in nanocarrier-based delivery systems have ushered in a significant breakthrough in enhancing the clinical effectiveness of treating complex disorders like Alzheimer’s disease. Leveraging their distinctive physicochemical properties and structural attributes, nanocarriers have demonstrated the potential to elevate therapeutic efficacy and enhance the brain uptake of conventional anti-Alzheimer’s drugs. While the solubility and bioavailability benefits offered by nanocarriers are unquestionable, the extent of improvement critically hinges on the chosen administration route [146].
Oral administration is less conducive for brain targeting due to inherent limitations, such as unpredictable or reduced bioavailability, increased dosage requirements and frequency, enzymatic degradation leading to an insufficient drug concentration reaching the brain, and more [147]. Overcoming the BBB and achieving targeted drug delivery have prompted the exploration of various invasive and non-invasive routes [148]. Invasive methods to breach the BBB encompass osmotic, chemical, ultrasound-mediated disruption, intra-cerebro-ventricular, and intrathecal infusions [149]. While effective in conditions like glioblastoma, these approaches entail significant drawbacks, including pathological changes in the brain, perturbed glucose uptake and homeostasis, toxicity to cerebral tissues, and disrupted brain function. Additionally, several of these techniques require high drug doses, potentially leading to toxicity [150,151].
Hence, non-invasive alternatives are under investigation. These include enhancing intracellular transport using transport carriers [152] and inhibiting efflux transporters, although initial inhibitors demonstrated notable toxicity risks [153]. Another strategy involves modifying drug structures to enhance lipid solubility (prodrugs) by limiting polar groups or attaching hydrophilic moieties to lipophilic side chains [154]. While this approach can enhance uptake to some degree, it often necessitates intricate compound engineering.
Further non-invasive methods encompass the Trojan horse approach, chimeric peptides, monoclonal antibody (MAB) fusion proteins, nanoparticle-based delivery, and intranasal delivery [155,156,157]. Each approach presents its own merits and limitations, but combining two or more approaches could potentially yield superior outcomes through dual targeting [158].
Intranasal drug delivery stands out as a well-recognized and established non-invasive strategy for treating various brain disorders [159]. The nasal cavity provides a direct route to the brain through olfactory and trigeminal pathways, while the highly vascularized nasal mucosa enables rapid drug absorption [160]. Enhanced brain targeting via intranasal delivery can reduce necessary dosage levels and minimize exposure to peripheral organs, thus mitigating toxicity [161]. Furthermore, compared to the oral route, intranasal administration offers a rapid onset of action, bypasses first-pass metabolism, and attenuates dose-related side effects [162]. Nonetheless, it is imperative to comprehensively grasp the physiological intricacies of intranasal targeting before formulating a dosage form.

3. Transport Mechanisms of Intranasal Route

At present, the treatment of Alzheimer’s disease primarily relies on systemic drug administration, usually in the form of oral or intravenous dosage forms. However, these conventional delivery methods come with several limitations, such as poor bioavailability, extensive first-pass metabolism, a slow onset of action, limited permeability, and restricted access to the brain due to the presence of the blood–brain barrier. In response, the intranasal route of administration has emerged as a promising avenue for addressing various brain-related disorders. The nasal cavity offers a direct pathway for nose-to-brain drug delivery via the olfactory and trigeminal pathways. The highly vascularized nasal mucosa facilitates rapid drug absorption and opens the door for a potential dose reduction through improved brain targeting. While intranasal delivery shows potential as a route for various therapeutic agents, including those for Alzheimer’s disease treatment, a thorough understanding of the physiological aspects of nasal drug delivery is crucial before developing a dosage form.
Brain targeting through the intranasal route predominantly occurs through three pathways: the respiratory pathway (an indirect route), the olfactory pathway, and the trigeminal pathway (a direct route) [163]. Intranasally administered drugs can travel through different pathways, including absorption by the nasal mucosa into the systemic circulation, axonal transport to the olfactory bulb, or direct entry through the trigeminal nerve [164]. Both the olfactory and trigeminal pathways are considered effective and safe routes for delivering active substances to the brain [165]. Gaining a comprehensive understanding of the mechanisms underlying these pathways is essential for devising effective therapeutic strategies for Alzheimer’s disease.
The olfactory neuronal pathway encompasses intra- and extra-neuronal mechanisms [166], spanning the olfactory epithelium, olfactory bulb, and lamina propria. Administered drugs reach the olfactory bulb from the olfactory region through a transcellular mechanism [167]. Additionally, various mechanisms such as paracellular transport, transcytosis, and diffusion, as well as the involvement of efflux transporters [168], can come into play based on the physicochemical properties of the drug. The olfactory bulb serves as a direct conduit for distributing the drug to different brain regions, including the piriform cortex, hypothalamus, and amygdala [150].
Another significant route for delivering active agents to the brain is the trigeminal pathway [169]. These nerves are present in the nasal epithelium of the respiratory region and extend to the brain via the pons, connecting with the olfactory bulb [170]. Within intranasal delivery, the ophthalmic and maxillary divisions of the trigeminal nerves play a pivotal role, as neurons in these areas directly traverse the nasal mucosa [150]. The segment of the trigeminal nerve that passes through the cribriform plates may contribute to drug delivery to the forebrain [171]. While this pathway is as equally significant as the olfactory pathway for delivering drugs to the anterior and other important brain regions, distinguishing the exact contribution of each pathway can be challenging [150].
Mucus within the nasal cavity plays a vital role in drug delivery and absorption. Mucin, a protein present in mucus, has the potential to bind with solutes, thereby influencing the diffusion process. Multiple mechanisms, including paracellular and transcellular routes [172], are involved in nasal delivery and absorption through the mucosa.
Intranasal drug delivery for neurological diseases has garnered significant attention. However, achieving targeted drug delivery to specific areas of interest remains a challenge due to a multitude of factors, encompassing the drug’s physicochemical properties, experimental conditions, and anatomical and structural characteristics [173]. Thorough investigations into and control of the therapeutic’s physicochemical attributes, including its nature, molecular weight, lipophilicity, shape, and size, are essential for successful formulation development via the intranasal route [174]. For instance, Huang et al. discovered that the ester form of L-tyrosine exhibited greater nasal absorption than that of L-tyrosine [175]. It has also been observed that nasal absorption is enhanced with lower-molecular-weight, cyclic molecule shapes [175]. Nevertheless, when the molecular weight of the active component surpasses 300 Da, permeability challenges may arise [176]. In the context of brain targeting, effective drug deposition within the olfactory epithelium hinges on dosing conditions, including head positioning, the administration technique, and the volume delivered [177]. Alongside dosing considerations, physiological factors such as the blood flow, enzyme activity, and mucociliary clearance of the nasal cavity can impact the absorption, therapeutic stability, and residence time.
Kushwaha et al. (2011) established a direct relationship between absorption and residence time, inversely linked to mucociliary clearance [175]. To surmount challenges associated with physiological and physicochemical factors, diverse strategies have been explored. These include the utilization of varied formulations (like dendrimers and vesicular systems) and permeation enhancers that modify the nasal cavity’s epithelial barrier. The nasal delivery of peptides, such as insulin, was limited due to degradation and a short half-life. To address these concerns, researchers delved into the prodrug approach [178], which not only provides protection but also enhances lipophilicity [179]. The incorporation of absorption enhancers has also proven effective in augmenting nasal delivery and targeting [180]. For instance, Chavanpatil et al. [181] examined the use of hydroxypropyl β-cyclodextrin, sodium deoxycholate, sodium caprate, sodium tauroglycocholate, and EDTA as penetration enhancers for the intranasal delivery of acyclovir. However, these approaches are not without drawbacks, including potential nasal toxicity, nasal mucosa damage [182,183], and limited success in breaching the BBB and precisely localizing therapeutics in the brain.
Therefore, a pressing need exists for a delivery system that can effectively traverse central nervous system barriers and guide the active ingredient to its intended target site without disrupting the physiology and structure of the nasal epithelium or the blood–brain barrier. Nanocarrier-based drug delivery systems present a promising alternative to traditional intranasal delivery methods [166,184,185]. Polymers, metal- and lipid-based particulate systems, vesicular carriers, and miscellaneous carriers such as nanoemulsions, nanosponges, dendrimers, and quantum dots are extensively explored nanocarrier-based platforms for intranasal drug delivery in the context of Alzheimer’s disease. The following section delves into various nanocarrier approaches reported for brain targeting via the intranasal route, aiming to effectively manage Alzheimer’s disease.

4. Intra-Nasal Nanoparticulate System for Alzheimer’s Disease Treatment

4.1. Nanoparticle-Based System

The utilization of nanoparticle-driven drug delivery has demonstrated its effectiveness in enhancing the absorption of nasal therapeutics. By encapsulating the drug within nanoparticles and safeguarding it from enzymatic degradation, therapeutic concentrations are elevated at the target site [186,187]. While the blood–brain barrier typically restrains particles exceeding 200 nm in size, nanoparticles with dimensions smaller than 200 nm can traverse the olfactory pathway [188]. Polymeric and metal nanoparticles have garnered attention for their potential in managing Alzheimer’s disease, offering several advantages such as a heightened loading capacity, degradation protection, enhanced stability, precise targeting, a reduction in dosage, and the potential for affinity enhancement for Aβ proteins, a hallmark of AD [189]. Furthermore, the surface modification of these nanoparticles can enhance their interaction with Aβ proteins.
Biodegradable and biocompatible polymers like chitosan, poly D, L-lactic-co-glycolic acid (PLGA), and polyvinyl alcohol (PVA) have been harnessed for intranasal drug delivery due to their controlled-release properties [190]. Chitosan, with its bio-adhesive nature, low toxicity, resistance to mucociliary clearance, and ability to prolong nasal residence time in the olfactory region, stands out as a preferred choice for nanoparticle formulations. This prolonged residence is attributed to interactions between the chitosan polymer’s polysaccharide moiety and the corresponding saccharide groups of the nasal mucosa. Chitosan also has the capacity to perturb intercellular tight junctions, thereby enhancing drug permeability.
Studies have delved into chitosan-based nanoparticles for intranasal drug delivery targeting Alzheimer’s disease. For instance, Wilson et al. developed chitosan nanoparticle-encapsulated sitagliptin and found a five-fold increase in the sitagliptin concentration compared to that of free sitagliptin. Furthermore, sitagliptin-loaded nanoparticles exhibited enhanced brain accumulation, potentially due to chitosan’s ability to modulate tight junctions [191]. In another study, Kandil et al. administered intranasal galantamine–chitosan complex nanoparticles to Wistar rats. This intervention led to reduced levels of MDA (malondialdehyde) and tumour necrosis factor-α in the brain extracts of nanoparticle-treated subjects in comparison to those of the control group. Conversely, higher levels of superoxide dismutase and glutathione were observed in the group treated with the galantamine–chitosan nanoparticles [192].
Zhang Li et al. conducted an in vitro/in vivo correlation (IVIVC) comparison between intranasally administered curcumin-loaded chitosan-coated PLGA nanoparticles and curcumin–hydroxypropyl-beta-cyclodextrin (HP-β-CD) inclusion complexes. The curcumin–HP-β-CD complex exhibited an improved cellular uptake and reduced cytotoxicity and demonstrated an antioxidant effect at a 20µM concentration in BV-2 cells, as compared to that of the curcumin–chitosan-PLGA nanoparticles [193].
Pawar et al. observed an enhanced uptake and reduced nasal clearance in glycol- and chitosan-coated PLGA nanoparticles. The glycol-coated nanoparticles displayed a superior uptake and nasal retention time compared to those of the chitosan-coated PLGA nanoparticles, potentially attributed to their surface charge density and polymer molecular weight [194].
Lastly, Sunena et al. evaluated the in vivo pharmacodynamics of intranasally administered galantamine-loaded thiolated chitosan nanoparticles. Their results underscored the significant delivery advantage of intranasal galantamine–chitosan nanoparticles for oral and nasal routes, highlighting the therapeutic superiority of intranasal administration [195].
The therapeutic potential of piperine (PIP), an alkaloid with cognitive improvement properties, is hindered by its poor aqueous solubility and low bioavailability, necessitating a high-dose regimen. In response, Elnaggar et al. devised a solution by developing intranasal chitosan nanoparticles (CS-NPs) encapsulating PIP, utilizing the ionic gelation technique. These CS-NPs exhibited a spherical morphology with optimal attributes, including a particle size of 248.50 nm, PDI of 0.24, zeta potential of +56.30 mV, and encapsulation efficiency (EE) of 81.70%. Their controlled-release behaviour was evident, with a 92% release achieved by the 24h mark. Comprehensive evaluations confirmed the safety of CS-NPs regarding nasal irritation and brain toxicity. Notably, PIP-NPs matched the effectiveness of standard donepezil injections in enhancing cognitive function, while displaying a remarkable 20-fold reduction in effective dosage compared to the conventional oral dosage. These nanoparticles also exhibited a dual mechanism involving anti-apoptosis and anti-inflammatory effects [196].
Similarly, Fazil et al. employed a similar approach to prepare nasal chitosan nanoparticles loaded with rivastigmine (CS-RHT NPs). Their characterization encompassed parameters like the zeta potential (ZP), particle size, PDI, and %EE. The brain-targeting capability of placebo NPs was assessed using rhodamine-123-based laser scanning microscopy. Pharmacokinetic and distribution investigations revealed a higher brain concentration of rivastigmine with CS-NPs (i.n.) (966 ± 20.66 ng mL−1; tmax of 60 min) compared to that of an intranasal drug solution (508.66 ± 22.50 ng mL−1; tmax of 60 min) or the intravenous administration of CS-NPs (387 ± 29.51 ng mL−1; tmax of 30 min). The drug transport efficiency of the CS-RHT NPs via nasal administration reached 355 ± 13.52%, with the direct transport percentage being approximately 71.80 ± 6.71%. An examination of the brain/blood ratio indicated the highest ratio for the CS-RHT NPs via the intra-nasal administration. Additionally, the study demonstrated the higher permeability of CS-RHT NPs compared to that of the pure drug solution. Overall, these findings underscored the brain-targeting potential of chitosan nanoparticles administered via the intranasal route [197].
A wide range of synthetic polymers, including poly(L-lactide-co-glycolic) acid, poly (lactic acid), and poly (glycolic acid), have been extensively explored for delivering drugs to the brain through the nasal route. The modification of these polymeric nanoparticles using compounds such as PEG (polyethylene glycol) or poloxamers can enhance drug loading, stability, and penetration across the nasal mucosa [195]. Musumeci T. et al. advanced this concept by developing PLGA nanoparticles and NLC-based nanosystems for adsorbing a neutralizing monoclonal antibody targeting TNF-related apoptosis-inducing ligand (TRAIL). Pharmacokinetics and dynamics studies in an AD mouse model demonstrated a high entrapment efficiency (99%) for both formulations, as confirmed by an ELISA. Notably, the intranasal administration of the antibody–nanocarrier complex led to significantly higher brain levels compared to those of the free anti-TRAIL antibody [198].
In a separate investigation, Yu Su et al. devised PEG-PLA nanoparticles loaded with miR132, a crucial molecule for sustaining neuronal survival in the brain. However, due to miRNA’s net anionic charge and low solubility in aqueous media, bare miRNA molecules are prone to rapid degradation or mucosal elimination following nasal administration. Thus, the quest for a carrier that ensures safety, an enhanced stability, and the target specificity level remains. The amalgamation of polylactic acid (PLA) and polyethylene glycol (PEG) generates a core-shell structure in aqueous environments, bolstering nasal permeability while diminishing mucociliary clearance. Animal studies have yielded augmented expressions of SYN and PSD-95, along with the inhibition of neuronal cell apoptosis in peripheral nerve cells and the cerebral cortex, signifying the neuroprotective effect of PLGA nanoparticles [199].
A comparison between intranasal curcumin- and bismethoxycurcumin-loaded PLGA nanoparticles showcased curcumin’s superior anti-inflammatory potential, interacting with molecular targets like amyloid peptide plaques and the cyclooxygenase2 enzyme, responsible for inflammatory reactions within the disease. Nanaki et al. constructed hybrid nanoparticles for nose-to-brain galantamine delivery, for which PLGA nanoparticles exhibited a greater uptake through olfactory unsheathing cells than that of PLA nanoparticles. Successful brain targeting was indicated by strong fluorescence in the hippocampus post intranasal administration, with an observed acceptable level of safety and no toxicity [200].
Protamine-coated PLGA nanoparticles within a Carbopol gel were formulated by Shamarekh et al. for Tacrine brain targeting via intranasal administration. This nanocomposite gel displayed higher Cmax and AUC values after 0–12h in the brain compared to those of i.v. and i.n. drug solutions. A histopathological analysis indicated no damage, suggesting their potential for neurodegenerative disease treatment [201]. Meng et al. developed lactoferrin-functionalized intranasal PLGA nanoparticles modified with N-trimethylated chitosan for effective Huperzine A brain targeting. In vivo imaging showcased prolonged brain fluorescence, with successful targeting evident in the olfactory bulb, cerebellum, cerebrum, and hippocampus following the nasal nanoparticles’ administration [202]. To enhance targetability and minimize mucociliary clearance, researchers have explored nanoparticle surface modifications with specific ligands, which demonstrate superior targeting compared to that of unmodified nanoparticles.
The field of nanomedicine in Alzheimer’s disease (AD) therapy has been a burgeoning area of exploration, particularly in the realm of metallic nanoparticles (NPs) for BBB-targeted delivery. However, the use of metallic NPs is hampered by chemical synthesis methods. Nonetheless, cerium, gold, selenium, and iron metallic NPs have demonstrated potent anti-AD capabilities, finding applications in theranostics, gene delivery, and stimulus-responsive therapies like photothermal treatments for diverse diseases, including cancer. Gold nanoparticles, particularly relevant for crossing the BBB, are being investigated for theranostic AD management. Bastus et al. [203] engineered gold nanoparticles targeting and solubilizing amyloid fibrillar aggregates, indicating their potential for dissolution via microwave-generated thermal energy. Controlled binding with the target through the energy input was established. While promising, exclusive AuNP targeting is imperative to mitigate cytotoxicity associated with amyloid beta oligomer species. Kogan et al.’s non-invasive investigation and amyloid beta aggregate manipulation technique seem advantageous for AD therapy [204]. Moreover, metallic nanoparticles have been explored for diagnostic purposes in detecting β-amyloid plaques in animal models.
Resveratrol, a promising neuroprotective stilbenoid, has the potential to enhance cognitive function in Alzheimer’s disease. However, its clinical efficacy is hindered by its extensive metabolism and poor bioavailability. To overcome these limitations, Salem et al. designed resveratrol-loaded transferosomes and nanoemulsions, incorporating gold nanoparticles (GNPs) for an improved delivery. Various physicochemical properties were assessed, along with dynamic studies such as water maze tests, to analyse spatial memory recovery. The results revealed memory improvements in all treated groups, with the transferosome–GNP gel group matching the normal group. Notably, the transferosome–GNPs exhibited enhanced permeation (81.29±2.64%) and symptom alleviation, with increased gold nanoparticle accumulation [205].
Iron oxide nanoparticles, another category of metal nanoparticles, are widely employed in AD therapeutic management. Zhang et al. devised super paramagnetic iron oxide NPs (SPIONs) modified with1,1-dicyano-2-[6-(dimethylamino)-naphthalene-2-yl] propene carboxyl. This disease model displayed a reduced signal strength in the hippocampal region [206]. Mahmoudi et al. explored the influence of SPIONs’ surface charge and coating thickness on beta amyloid fibrillary dynamics, revealing a direct correlation between the SPION concentration and fibrillation rate. Positively charged SPIONs induced fibrillation at lower concentrations compared to neutrally/negatively charged ones. Leveraging the magnetic properties of amyloid beta fibrils, FDA-approved AD drugs can be coupled with SPIONs or similar metal nanoparticles for targeted intranasal delivery [207].
Addressing the reactive oxygen species (ROS) concentration in the brain represents another crucial AD treatment avenue. Selenium (II), sodium selenite (IV and VI), are potent ROS inhibitors, pivotal in curbing oxidative stress and cellular cytotoxicity. Selenium- and selenite-containing nanoparticles have biomedical applications [208]. Yin et al. synthesized sialic acid (SA)-functionalized selenium (Se) nanoparticles, further linked with substitute peptide-B6 peptide (B6-SA-SeNPs). These nanoparticles showcased enhanced BBB transport, promising a nanomedicine-based strategy for AD modification. Uptake studies and transport capability assessments using cerebral endothelial cells (bEnd.3) and inductively coupled plasma atomic emission (ICP-AES) highlighted B6-SA-SeNPs’three-fold higher uptake compared to that of SA-SeNPs. The transwell method and PC12 co-culture models demonstrated the B6-SA-SeNPs’ superior transport ability. These findings indicate B6 peptide’s potential in enhancing brain delivery, suggesting B6-SA-SeNPs as a favourable platform, particularly for intranasal AD treatment.
Metal nanoparticles present a versatile platform for the intranasal targeting of various therapeutic agents in AD management. Ongoing research endeavours focus on harnessing green-chemistry-based synthesis methods to optimize these nanoparticles for future AD treatments.

4.2. Lipid Nanocarriers

Lipid nanocarriers, consisting of solid lipid matrices (SLNs) or combinations of solid lipid and oil matrices (NLCs), have garnered significant attention as versatile delivery systems. These nanocarriers offer benefits such as prolonged retention, reduced clearance, enhanced solubilization and permeation, improved stability, and compatibility within the nasomucosal region. Researchers have extensively explored SLNs and NLCs for intranasal delivery, showcasing improved brain-targeting efficacy. Solid lipid nanoparticles (SLNs) and nanostructured lipid carriers (NLCs) represent lipid-based nanocarriers that excel in delivering both hydrophobic and hydrophilic drugs [209].
Addressing the limitations of risperidone, an anti-psychotic drug commonly used to treat Alzheimer’s-related agitation, Patel et al. engineered solid lipid nanoparticles (RSLNs) using Compritol 888 ATO and Pluronic F-127. These RSLNs exhibited a high entrapment efficiency (59.65% ± 1.18%) and a narrow PdI of 0.148 ± 0.028, indicating formulation stability. Pharmacodynamic assessments using hindlimb retraction time (HRT) in a mouse model demonstrated the superior antipsychotic potential and brain targeting of RSLNs compared to risperidone solution (RS) and a control. The intranasal administration of RSLNs yielded a brain/blood ratio 10-fold higher than that of their intravenous administration, highlighting improved brain concentration [210]. Deepshi et al. utilized a solvent evaporation diffusion method to design rivastigmine tartrate-loaded SLNs, achieving optimized particle size, entrapment efficiency, and drug content. These rivastigmine-loaded SLNs showcased sustained release and improved ex-vivo nasal mucosa flow and diffusion coefficients compared to rivastigmine solution [211]. Similarly, Yasir et al. created donepezil-entrapped solid lipid nanocarriers using glyceryl behenate, exhibiting enhanced targeting potential and improved brain bioavailability [212].
The surface modification of SLNs, akin to polymeric nanoparticles, enhances their target specificity [78]. Yusuf et al. explored surface-modified SLNs for the enhanced bioavailability and brain targeting of piperine. Surface-coated SLNs demonstrated reduced superoxide dismutase values and cholinergic degradation, with a sustained brain concentration and improved bioavailability compared to those of free drug [213]. Saini et al. incorporated ferulic acid into SLNs, enhancing their permeability across lipophilic barriers, and further surface-modified the SLNs with chitosan. The chitosan-coated SLNs showcased a superior drug concentration in the brain, improved cognition, and improved biochemical factor levels in the cortex and hippocampus [214].
While solid lipid nanoparticles (SLNs) have shown potential, their limitations have led to the emergence of nanostructured lipid nanocarriers (NLCs). Anand et al. developed NLCs loaded with rivastigmine hydrogen tartrate for dementia treatment. The NLCs displayed controlled release, enhanced penetration, and decreased acetylcholinesterase expressions, suggesting their potential for Alzheimer’s management [215].
In the realm of Alzheimer’s therapy, lipid nanocarriers hold great promise, offering a transformative approach to drug delivery and targeting within the brain.
The pioneering work of Musumeci et al. [198] aimed to surmount challenges in Alzheimer’s disease (AD) treatment, including high dosage regimens and low transport efficiency. To achieve this, they devised nanostructured lipid carriers (NLCs) through a phase inversion technique without organic solvents (the PIT method). The NLCs were then coated with TRAIL and subjected to freeze-drying using glucose as a cryoprotectant. Immunofluorescence studies on 3xTg-AD and wild-type mice demonstrated that NANO-A and NANO-B complexes, upon being injected intranasally, effectively traversed the BBB of 3xTg-AD mice. This showed successful TRAIL targeting, known to be abundant in hippocampal inflammatory sites. Blocking TRAIL yielded cognitive enhancements and the halting of disease progression and brain degeneration.
In a quest to enhance brain targeting and nasal retention, Vavia et al. [216] delved into an in-situ gel loaded with rivastigmine nanostructured lipid carriers (NLCs). The incorporation of stearylamine (SA) into the NLCs facilitated nasal retention by overcoming mucociliary drainage. Pharmacokinetic and distribution studies revealed NLCs’ sustained release, improved brain penetration, and BBB penetration. This led to cognitive recovery in amnesic mice through intravenous and intranasal administration [216]. Similarly, Jojo et al. devised intranasal pioglitazone NLCs using the micro-emulsion method. The optimized NLCs exhibited increased permeability, flux, and permeability coefficients compared to those of a drug solution. In vivo studies showcased elevated brain/blood ratios, demonstrating the potential of NLCs in clinical AD management via intranasal administration [217].
Moreover, lipid nanoparticles, including solid lipid nanoparticles (SLNs) and NLCs, have demonstrated efficacy as effective carriers for brain-targeted drug delivery. The ingenious utilization of lipid nanocarriers holds immense promise in revolutionizing Alzheimer’s treatment. Through ingenious engineering and innovative delivery strategies, these nanocarriers pave the way for targeted and enhanced drug delivery to the brain, offering renewed hope in the battle against this debilitating disease.
Recent studies have demonstrated the potential of liposomal formulations in revolutionizing Alzheimer’s disease (AD) treatment. In the study by Li et al. (2022) [163], encapsulating Hydroxy-α-sanshool (HAS) within liposomes led to a superior targeting efficacy compared to that of free HAS. Liposomes, owing to their versatility, are capable of encapsulating hydrophilic, hydrophobic, and amphipathic therapeutics. However, overcoming challenges posed by limited blood–brain barrier (BBB) penetration and oral bioavailability is essential for effective AD drug delivery. To tackle this, Rompicherla et al. [218] compared intranasal rivastigmine-loaded liposomes to PLGA nanoparticles. Their results highlighted that liposomal formulations exhibited rapid action and higher concentrations, achieving notable acetylcholinesterase inhibition in plasma and brain homogenate samples. Sokolik VV et al. conducted a comparative analysis between solubilized and liposomal curcumin formulations in an AD model [219].
Curcumin, renowned for its anti-inflammatory properties and potential in reducing Alzheimer’s symptoms, has faced limitations due to its stability and low bioavailability. Overcoming these challenges, intranasal liposomal curcumin displayed enhanced cognitive responses and a greater reduction in cytokine biomarkers, offering a promising avenue for AD treatment [79]. Galantamine hydrobromide, an AD-approved drug, has shown adverse effects when administered through oral and parenteral routes. Seeking an alternative, Li et al. explored an intranasal galantamine hydrobromide (GH)-loaded flexible liposomal formulation. Characterized by highly elastic fluid membranes, flexible liposomes are optimal for efficiently delivering hydrophilic compounds across cell membranes. GH-loaded flexible liposomes demonstrated favourable characteristics, including size and zeta potential. Pharmacokinetic studies indicated superior brain concentrations for formulations administered nasally, with flexible liposomes showing the highest concentration [220].
Furthermore, liposomes have shown great potential as carriers for neurotrophic factors, attributed to their cellular uptake enhancement, lipophilicity, and degradation protection. Cationic liposomes, particularly, have displayed improved protein passage across the nasal epithelium. Migliore et al. developed cationic liposomes loaded with ovalbumin (OVAL), which exhibited persistent brain delivery, highlighting their viability for protein transport [221].
The therapeutic potential of liposomes extends to targeting H102 peptide, which cleaves β-sheets. Zheng et al. developed H102-peptide-based liposomes that demonstrated enhanced brain penetration and reduced degradation, significantly improving spatial memory and enzyme activities in AD-induced rat models [222]. Similarly, Yang et al. explored rivastigmine-loaded liposomes modified with PEGylated poly-arginine CPP derivatives to enhance stability and brain targeting through improved transcytosis [223]. Another strategy by El-Helaly et al. involved introducing a positive charge using dodecyl dimethyl ammonium bromide to maintain stability. Further coupling with PEGylated lipids yielded stable electrostatic stealth long-circulating liposomes, with an increased drug concentration observed in both plasma and the brain [224]. Collectively, these recent studies underscore the potential of liposomes in enhancing Alzheimer’s treatment. Their versatility, stability improvement, and targeted delivery capabilities make them a promising tool in the fight against this debilitating disease.
Arumugam and colleagues ventured into the realm of Alzheimer’s disease (AD) treatment by developing liposomes incorporating rivastigmine. They embarked on a comparative study to discern rivastigmine concentrations in plasma after administering free drugs via oral and nasal routes, orally administered liposomes, and liposomes delivered intranasally. Intriguingly, intranasal liposome administration displayed a remarkable five-fold increase in the area under the curve (AUC) compared to that of orally administered free drugs, and a three-fold rise compared to that of intranasal free drug administration. Furthermore, rivastigmine-loaded liposomes (IN) exhibited a notable 5.6-fold surge in brain concentration and a prolonged half-life (T1/2) compared to those of free drug solutions via the nasal and oral routes. This enhancement in absorption can be attributed to effective brain targeting facilitated through the nasal olfactory pathway, with the physicochemical attributes of the drug also playing a pivotal role in breaching the BBB [225].
In addition to the targeting strategies discussed earlier, liposomal carriers can be harnessed with Aβ targeting ligands or brain-penetrating peptides for heightened brain-specific delivery. A new avenue lies in multifunctional liposomes, catering to both therapeutic and diagnostic roles. Mourtas et al. delved into this frontier, crafting DPS–curcumin surface immobilized nanoliposomes for AD treatment. These nanoliposomes exhibited a dual functionality: labelling Aβ deposition with a high efficiency and instigating the inhibition of amyloid beta-42 aggregates. Intriguingly, these multifunctional nanoliposomes could switch between activated and inactivated states, granting them a theranostic capability [226].
Indeed, multifunctional nanoliposomes are gaining attention from various researchers for their potential in brain targeting and the management of Alzheimer’s disease. Table 5 and Table 6 provide an overview of research endeavours concerning polymeric nanoparticles, lipid nanoparticles, and liposomes in the context of Alzheimer’s disease treatment.

4.3. Nanoemulsions and Microemulsions

Nanoemulsions are a specialized drug delivery system composed of two non-miscible phases held together by surfactants, resulting in a stable and uniform solution. These formulations typically range in size from 20 to 200 nanometres [123]. Intranasal nanoemulsions have shown promising results in experimental studies, allowing for the direct delivery of small molecules to the brain. This approach addresses challenges related to poor aqueous solubility, limited bioavailability, degradation, and a slow onset of action. The addition of mucoadhesive polymers can prevent the rapid nasal clearance of nanoemulsions [241]. However, these systems are kinetically stable and require a high amount of energy for manufacturing. In contrast, microemulsions (MEs) are pseudo-ternary formulations comprising oil, aqueous media, surfactants, and co-surfactants, forming spontaneously and remaining thermodynamically stable. AnME system usually has a size range from 10 to 100 nm, enabling passive targeting. Both nanoemulsions and microemulsions are biodegradable, biocompatible, and display nanometric sizes. Nevertheless, these formulations can experience sedimentation, creaming, and Ostwald ripening. Proper formulation design can lead to the creation of stable nanoemulsions and microemulsions for extended periods [242]. This section discusses several experimental studies involving the intranasal delivery of Alzheimer’s therapeutics using nanoemulsions and microemulsions.
Atinderpal et al. developed a nasal nanoemulsion containing memantine through a combination of pressure homogenization and ultrasonication. The resulting nanoemulsion’s average size, zeta potential (ZP), polydispersity index (PdI), and entrapment efficiency (% EE) were characterized. In vitro diffusion studies conducted in simulated nasal fluid (SNF) at a pH of 5, phosphate buffer saline (PBS) at a pH of 7.4, and artificial cerebrospinal fluid (ACSF) at a pH of 7.3 demonstrated 80%, 60%, and 40% drug release after 6 h, respectively. The prepared nanoemulsion exhibited first-order release kinetics in SNF and adhered to the Peppas kinetic model in PBS and ACSF. The nanoemulsion exhibited a strong antioxidant potential in FRAP and DPPH assays and displayed a higher reducing potential, which is beneficial for Alzheimer’s treatment. In vivo studies using radiolabelled memantine revealed the highest radioactivity percentage in the brain after intranasal administration. Biodistribution studies and gamma images indicated direct nose-to-brain targeting across the blood–brain barrier (BBB) [128].
The study by Kaur et al. demonstrated the brain-targeting potential and antioxidant activity of intranasal nanoemulsions. Specifically, a technetium pertechnetate (99mTc) labelled donepezil nanoemulsion exhibited successful intranasal brain delivery, as confirmed by scintigraphy imaging. This nanoemulsion showed no adverse effects on cell morphology but displayed dose-dependent cytotoxicity and radical scavenging activity percentage (%RSA) [243].
Furthermore, nanoemulsion systems have exhibited versatility in enhancing brain-targeting efficacy for a wide range of molecules, including poorly soluble drugs such as osthole and resveratrol. For instance, Song et al. formulated a nasal nanoemulsion of osthole, a natural coumarin with potential therapeutic properties. The resulting OST-NE formulation demonstrated significant improvements in spatial memory, decreased cholinesterase activity, increased anticholine content, and neuroprotective effects in mouse models, rendering it a promising option for Alzheimer’s therapy [244].
Similarly, Kota et al. developed a coconut oil-based resveratrol nanoemulsion, and Vasdev et al. formulated a low-energy nanoemulsion using rosemary oil and donepezil for Alzheimer’s treatment. The safety of these nanoemulsions was confirmed through ex-vivo mucosal ciliotoxicity and permeation studies. The low energy requirement of these formulations suggests their potential scalability [245].
Comparative pharmacokinetic studies between a nanoemulsion and suspension, as conducted by Kotta et al., revealed that the nanoemulsion exhibited a higher maximum concentration (Cmax), a shorter time to reach maximum concentration (Tmax), and a larger area under the curve (AUC) compared to those of the suspension, in terms of both plasma and brain distribution. These findings indicate the potential of the developed nanoemulsion as a suitable candidate for targeted drug delivery to the brain [246].
In addition to passive brain targeting, ligand-modified nanoemulsions have been explored for active targeting to the brain. For instance, Jiang et al. optimized a lactoferrin-loaded HupA intranasal nanoemulsion, demonstrating enhanced brain uptake through specific carriers and transcytosis. An in vivo analysis confirmed its successful delivery to the central nervous system, signifying its potential for Alzheimer’s treatment [247]. Recent research efforts have shifted towards exploring the brain-targeting ability of thermodynamically stable dispersion systems like microemulsions. Wen et al. developed an ibuprofen-based microemulsion for managing Alzheimer’s, resulting in a significantly increased brain uptake compared to that of intravenous and oral administrations of ibuprofen. Additionally, Zussy et al. demonstrated that an intranasal microemulsion of nanovectorized docosahexaenoic acid (DHA) improved cognitive ability and reduced tau phosphorylation in AD mouse models [248,249]. Various targeting approaches using microemulsions have been investigated to enhance brain uptake and therapeutic efficacy. For instance, Chen et al. formulated a dual-responsive intranasal microemulgel for the delivery of Huperzine A, exhibiting significantly improved drug exposure in the brain. Another study by Khunt et al. utilized omega-3 fatty acids and butter oil for the targeted delivery of donepezil hydrochloride in a microemulsion formulation via the intranasal route, achieving a superior bioavailability compared to that of the solution [250,251].
Furthermore, Shah et al. conducted a comparative study between a plain microemulsion (ME) and a chitosan-based bioadhesive microemulsion (MME) for the intranasal delivery of rivastigmine. Their results showed that the MME exhibited higher diffusion through the nasal mucosa and an increased concentration of the therapeutic agent in the brain, surpassing the performance of the ME and the solution [252].
In a separate study, Pathak et al. developed a mucoadhesive microemulsion of nimodipine using Carbopol 934. This formulation demonstrated a rapid burst release followed by a sustained release, leading to an increased concentration of the therapeutic agent in the brain [253]. Collectively, these studies underscore the potential of intranasal nanoemulsions and microemulsions as feasible, cost-effective, and scalable approaches for delivering both synthetic and natural treatments for Alzheimer’s disease. The summarized research findings are presented in Table 7.

4.4. Miscellaneous Nanocarriers

Nano suspensions represent a promising approach for intranasal drug delivery, especially for poorly soluble agents, by utilizing surfactant-stabilized small-scale dispersions. These systems retain their crystalline structure, have an enhanced drug loading capacity, and can be engineered to avoid phagocytosis for targeted delivery [259].
The challenges of curcumin, a potent neuroprotective compound against Aβ plaques, which include its rapid metabolism and poor absorption, have been addressed by Dibaei et al. They developed a surface-engineered nano-suspension of curcumin using stabilizers such as D-α-tocopheryl polyethylene glycol 1000 succinate and Tween 80. Through high-pressure homogenization and probe sonication, nanocrystals were formed. Their study reported enhanced brain concentrations with the Tween 80-coated curcumin, indicating better ApoE absorption than that of the TPGS-coated NS. However, the TPGS-NS exhibited a higher brain distribution than that of the plain curcumin solution [260]. Bhavna et al. employed an ionic cross-linking technique to create a chitosan-based intranasal nanosuspension of donepezil. The nanosuspension exhibited a size range of 150–200 nm with a PDI of 0.341. Safety evaluations showed no toxicity and no mortality in vivo. Additionally, a higher fraction of donepezil was detected in the brain (147.54 ± 25.08 ng/mL) at a 0.5 mg/mL dose. These findings underscore nanosuspensions’ potential as Alzheimer’s disease treatment carriers, suggesting that surface modifications or coatings could enhance their targeting efficiency [261].
Nanocrystals, pure drug crystals with no carriers, offer several benefits such as a higher surface-to-volume ratio, an enhanced dissolution rate, and versatile administration routes, ultimately leading to improved bioavailability and therapeutic effectiveness [262,263]. Paeoniflorin, a neuroprotective agent with poor oral bioavailability, was transformed into nanocrystals by Wu et al. using the anti-solvent precipitation method. The nanocrystals exhibited an average size of 139.6 ± 1.3 nm with a zeta potential of −23.2 ± 0.529 mV. In vitro studies demonstrated enhanced release and brain uptake, along with neuroprotective effects on damaged SHSY5Y cells mediated by MPP+ [264]. Stahr et al. highlighted the importance of nanocrystal size in targeting efficiency. Hesperidin nanocrystals of varying sizes were developed, indicating that a smaller size (<200 nm) improved their dissolution rate and solubility, while surface modifications with ligands further enhanced their targeted delivery [265].
Zhu et al. (2021) introduced a nanocrystal-based hydrogel to increase the solubility and permeation of armodafinil, known for cognitive enhancement. Utilizing PVP-K90 and lecithin, they incorporated armodafinil nanocrystals into the hydrogel. This formulation exhibited a high stability due to intermolecular hydrogen bonding. A pharmacokinetic analysis indicated significantly higher brain concentrations with the intranasal nanocrystal-based hydrogel (Cmax = 9533.0 ± 2327.9 ng/mL, Tmax = 0.21 ± 0.08 h) compared to those of its oral administration (Cmax = 4170.0 ± 388.3 ng/mL, Tmax = 0.25 ± 0.00 h). The relative brain-targeting index was 1.99, reflecting the hydrogel’s enhanced brain-targeting ability [266].
Quantum dots (QDs), semiconductor nanocrystals with unique optical and electronic properties, offer a superior stability and multi-functionality for diagnostics. Gao et al. developed CdSe/ZnS QDs coated with poly (ethylene glycol)-poly(lactic acid) nanoparticles (QDs-NPs) to enhance nasal QD delivery. The details are mentioned in Table 8. By further modifying QDs-NPs with wheat germ agglutinin (WGA), they created a WGA-QDs-NP system. Biodistribution studies showed fluorescence signals in the brain region, indicating effective nasal delivery. Fomicheva et al. demonstrated the potential of QDs in detecting Alzheimer’s disease biomarkers, while Thakur et al. explored the capacity of QDs to influence fibrillation [234,267,268]. Diverse nanocarriers, from nano suspensions to nanocrystals and quantum dots, hold promise for enhancing Alzheimer’s disease treatment, offering solutions to challenges such as poor solubility and targeting efficacy.
Dendrimers represent advanced nano-scale systems featuring three-dimensional polymeric cores, which can be tailored for a range of applications. Their capability to traverse cellular membranes, including the blood–brain barrier, has rendered them increasingly valuable in the realm of drug delivery. Notable types of dendrimers encompass poly(amidoamine), PEGylation, pH-sensitive, and peptide dendrimers [276]. Specifically, poly(amidoamine) (PAMAM), polypropylene polybenzylisocyanate (PPI), polylysine (PLL), and carbosilane dendrimers are frequently harnessed for brain targeting.
Dendrimers like Poly(amidoamine) (PAMAM), Polypropylene polybenzylisocyanate (PPI), Polylysine (PLL), and carbosilane are commonly utilized for brain targeting owing to their branched structure, which facilitates the functionalization of active agents and ligand-like peptides derived from ApoE. This enhancement in recognition by LDL receptors present on the endothelial cells of the central nervous system leads to improved uptake and targeting [277]. To overcome the limited half-life and suboptimal permeability of flurbiprofen across the blood–brain barrier, Al-azzawi et al. devised flurbiprofen dendrimers using the solid-phase peptide method for Alzheimer’s disease treatment. The synthesized dendrons loaded with FP were successfully characterized via mass spectrometry and FTIR. Their biocompatibility was evaluated through cytotoxicity assays, and a notable level of permeability (14.79 ± 2.06) was observed for an in vitro model of the BBB as analysed via HPLC [278].

4.5. InSitu Gelling System

In-situ gels can be classified into various types based on the external stimuli employed, including temperature-sensitive, ion-sensitive, and pH-sensitive gelling systems. These formulations undergo a transformation from a sol to a gel state in the presence of specific external triggers. In the context of intranasal drug delivery, in-situ nasal gels offer several advantages, such as an extended residence time, enhanced drug penetration, increased payload capacity, improved elasticity, and robust stability due to the gels’ crosslinking property. Moreover, they enable sustained drug release, as elucidated by Hamano et al. in 2018 [123]. Agrawal et al. (2020) elucidated the existence of diverse types of in-situ gelling systems, encompassing temperature-sensitive and ionic cross-linking systems. Depending on the specific type of in-situ gel, various polymers such as xyloglucan, EHEC, pluronic, poloxamer, carbopol, gellan gum, and chitosan are employed [279]. In efforts to mitigate the side effects associated with oral administration.
Patil et al. devised a mucoadhesive in-situ gel incorporating cubosomes containing donepezil. The cubosomes were formulated using glycerol mono-oleate and poloxamer 407, optimized through a central composite design. The optimum cubosome formulation comprised 2% glyceryl mono-oleate and 1.5% poloxamer 407. Additionally, gellan gum (0.3%) and konjac gum (0.03%) were utilized as the gel-forming and mucoadhesive components, respectively. The optimized in-situ gel underwent characterization for various parameters, encompassing the zeta potential, size, PDI, and % EE. The prepared cubosome-based in-situ gel exhibited a drug content of 90.16 ± 1.02%, along with a pH of 6.4 ± 1.29. Notably, the viscosity of the cubosome-based in-situ gel was measured at 180 ± 9.5 cps, accompanied by a gel strength of 34 ± 2.11 s. The in vitro drug release showcased an initial burst release of 24.52% at 2 h, followed by 53.73% at the end of 6 h. Biodistribution studies conducted in vivo exhibited the highest CMAX values for the brain with the cubosome-based in-situ gel (24.01 ± 7.32 µg/mL), followed by the cubosome dispersion (14.34 ± 6.31 µg/mL) and the plain drug solution (3.96 ± 2.38 µg/mL). The Tmax (minutes) value of the dispersion and in-situ gel was 60 ± 0.0. The AUC (0–240 min) for the in-situ gel was 2460.19 ± 4.42 (µg·min·mL−1), whereas for the plain dispersion, it was 2002.55 ± 5.56 (µg·min·mL−1). Thus, the in-situ gel demonstrated enhanced brain targeting via the nasal route [280].
Cunha and colleagues developed an in-situ gel loaded with rivastigmine (RVG) employing nanoemulsion and nanostructured lipid carriers (NLCs) to prolong the residence time within the nasal cavity. Through meticulous optimization involving different percentages of the thermosensitive polymer, the final batch containing 17% Kolliphor® P407 and 0.3% MethocelTM K4M was identified as the most effective. The RVG-loaded nanoemulsion exhibited a particle size of 141.70 ± 0.40 nm with a PDI 0.45 ± 0.00, while the RVG-loaded NLCs possessed a particle size of 146.10 ± 1.73 nm with a PDI 0.43 ± 0.02. A texture analysis revealed that the NLC-loaded gel demonstrated superior firmness and bioadhesion compared to the nanoemulsion-based gel. Both formulations exhibited enhanced firmness and adhesiveness compared to the plain gel, indicating the potential of nanosystem-based in-situ gels to enhance the retention time within the nasal cavity [281].
Chen et al. pioneered the development of a dual-responsive (pH and temperature) in-situ gel of huperzine A using chitosan as a pH-responsive mucoadhesive polymer and pluronic F127 as a temperature-sensitive agent. This gel was designed to address challenges associated with low bioavailability and efficacy. The hup A microemulsion (ME) and hup AME temperature- and pH-sensitive in-situ gel (TPISG) demonstrated an average size of 21.26 nm and 20.53 nm, respectively. The optimized hup AME TPISG formulation exhibited a clear, free-flowing liquid state at room temperature (560 ±10 mPa s), transitioning to a highly viscous gel (5200 ± 100 mPa s) under nasal conditions. The gelation time was 89 s, with a gelation temperature of 29–34°C. Invitro studies comparing different formulations with initial burst releases (hup AME: 10.7%, hup A solution: 8.8%, hup AMTISG: 9.0%, and optimized hup AMETPISG: 10.6%) over a 0.5h period revealed sustained release in the case of the optimized hup AMETPISG, with 90.52% release at 24 h, attributable to the presence of pluronic F127. An in-vivo evaluation using the microdialysis method indicated improved brain targeting and patient compliance (Table 9).

5. Toxicity and Safety Aspects of Nanoparticulate Delivery

Nanoparticles possess the potential to introduce toxicity at various levels, ranging from organs, tissues, and cells to even subcellular components, owing to their distinct physicochemical attributes [288,289]. Table 10 provides detailed preview of detailed toxicity studies conducted on plethora of nanomaterials. Certain metal particles have demonstrated heightened toxicity as their size diminishes, despite their inert nature. Nanoparticles engage with enzymes and proteins within cells, disrupting antioxidant defence mechanisms, leading to the generation of reactive oxygen species and eliciting inflammatory responses, ultimately resulting in necrosis [290]. The toxicity associated with nanoparticles is contingent upon a spectrum of physicochemical factors, including dose, size, surface area, concentration, crystalline structure, aspect ratio, surface coating, and functionalization [291], as well as chemical stability [292,293]. Given that industrial nanoparticles predominantly comprise heavy metals, the compatibility and toxicity factors warrant careful consideration. Although the bioavailability of heavy metal NPs may be restricted, their inherent toxicity remains substantial [292]. For instance, superparamagnetic iron oxide NPs have been documented to modulate gene expression, cellular response, homeostasis, and even cell cycle dynamics [294]. While an abundance of research is dedicated to unravelling the impact of nanoparticle toxicity on human health, some studies have delved into the ecological ramifications to foster the sustainable utilization of this innovative material [290].

6. Regulatory Aspects/Challenges of Intranasal Nanocarrier Drug Delivery

Despite the emergence of numerous approved nanomedicines in recent decades, many countries still lack well-defined regulations governing the marketing and utilization of nanocarrier-based formulations. This regulatory gap has constrained the full clinical potential of nanomedicines, underscoring the urgency of collaborative initiatives among global regulatory bodies to establish a robust framework for nanocarrier development. In the absence of explicit guidelines, certain assessments related to the safety, toxicity, and compatibility of nanoformulations are conducted following strategies akin to those employed for conventional therapies [298,299].
The regulation of biologics-based nanomedicines falls within the purview of the framework devised by the European Medicines Agency (EMA). For formulations encompassing proteins and antibodies, the manufacturer is mandated to adhere to the regulations governing new chemical entities (NCEs) and biological medicinal products [300,301]. Conversely, the EMA employs case-by-case analyses for non-biological complex drugs (NBCDs). In specific scenarios, regulatory guidelines for NBCDs can align with the biological framework [302].
The development of nanomedicines presents a substantial challenge owing to the necessity for an extensive characterization of their attributes, which can be easily influenced by even minor modifications. Researchers have been actively pursuing targeted drug delivery through ligand attachment, receptor engagement, or conjugation with diagnostic imaging agents. In such instances, novel approaches are required to assess their physicochemical properties and performance, encompassing considerations such as biocompatibility, protein interactions, and drug metabolism, among others [303].
In India, the requisites for quality, safety, and efficacy data differ based on the approval status of drugs and nanocarriers. All nanopharmaceutical formulations are treated as Investigational New Drugs (INDs), yet their scrutiny may vary depending on diverse categories. If both the drug and nanocarrier are novel and lack prior approval, they are treated as an IND, necessitating adherence to Schedule Y of the Drug and Cosmetics Rule, 1945. Similar guidelines apply to approved nanocarriers paired with new drugs, albeit independent studies specific to nanocarriers may not be mandatory. For fresh nanocarriers paired with traditional or conventional drugs, complete adherence to the Schedule Y IND guidelines might not be requisite, but documented evidence of safety and efficacy remains imperative. In cases in which both the drug and nanocarrier have been previously sanctioned, abbreviated studies are undertaken. The data requisites for nanopharmaceuticals are approached on a case-by-case basis, factoring parameters such as their biological and physicochemical attributes, alongside other considerations, such as the available data for the drug or nanocarrier, encompassing nonclinical proof of concept to clinical challenges.
The FDA and EMA have formulated evaluation guidelines for intranasal formulations, encompassing diverse factors like physical characterization, plume geometry, resting time effects, agitation requirements, particle size distribution, photo-stability, and microbial challenges. The specific tests mandated might vary based on factors such as the formulation type (e.g., suspension, drops, or powder), the device employed (e.g., metered dose container), and the intended application (e.g., single or multiple sprays). While the FDA and EMA have provided suitable methodologies for conducting these evaluations, there exists a scarcity of comprehensive procedural guidelines for their execution. Despite significant strides in nanomedicine-based drug delivery systems, their scaling up and advancement have encountered obstacles due to the absence of universally accepted and harmonized regulatory directives for their assessment and process control. The current intricate and sophisticated nanostructured formulations present additional regulatory complexities [299]. To surmount these challenges, regulatory agencies must collaborate to establish a universally recognized protocol for intranasal nanocarrier systems. This protocol should encompass comprehensive evaluation guidelines and also encompass considerations for process-related variables that might impact the final performance of the product.

7. Conclusions

Intranasal drug delivery utilizing nanotechnology has emerged as a promising approach for addressing the challenges posed by Alzheimer’s disease management. The convergence of nanotechnology and pharmaceutical sciences has yielded innovative strategies to enhance the efficacy, bioavailability, and targeted delivery of therapeutic agents across the blood–brain barrier. This review has explored the substantial progress made in this field, shedding light on the advancements, challenges, and future prospects for intranasal drug delivery in Alzheimer’s disease management. The utilization of nanocarriers, such as liposomes, solid lipid nanoparticles, nanoemulsions, and dendrimers, has enabled the precise encapsulation and controlled release of Alzheimer’s disease therapeutics. These nanocarriers offer improved drug stability, a prolonged residence time, and the potential for targeted brain delivery. Furthermore, their versatile nature allows for surface modification, ligand conjugation, and multi-functionalization, thereby enhancing their ability to cross biological barriers and interact with specific cellular receptors. While the potential of intranasal nanocarrier drug delivery for Alzheimer’s disease treatment is significant, several challenges warrant attention. Toxicity and safety concerns associated with nanomaterials necessitate comprehensive evaluations and standardized regulatory guidelines to ensure patient safety. Additionally, the scalability of nanotechnology-based formulations remains a pivotal concern, as their transition from laboratory research to large-scale production requires rigorous optimization and cost-effectiveness’s, the multi-facet pathology of AD poses a significant challenge in the clinical translation of ongoing research. There still exists a gap in our understanding of the etiology and identification of potential targets. Based on this understanding, small-molecule and associated formulations can be developed. Thus, fostering inter-disciplinary collaborations among researchers, clinicians, and regulatory bodies could provide valuable insights for tackling these problems.

8. Future Prospects

The future trajectory of intranasal drug delivery for Alzheimer’s disease management via nanotechnology is imbued with profound promise. To harness this potential to its fullest extent, it is essential to foster collaborations among researchers, clinicians, and regulatory agencies. The following avenues hold considerable promise for advancement:
Precision Targeting: Delving into advanced targeting strategies involving ligands, peptides, or biomolecules could yield precise and potent brain delivery. Tailoring nanocarriers to selectively engage relevant receptors promises to amplify their therapeutic efficacy.
Personalized Therapies: By harnessing patient-specific nanomedicines, the treatment landscape could undergo a paradigm shift. Accounting for individual genetic and physiological nuances through tailored therapies could optimize outcomes and curtail adverse effects.
Theranostic Platforms: Orchestrating diagnostic and therapeutic functionalities within a singular nanocarrier configuration can furnish real-time insights into drug delivery and treatment response. This integrated theranostic approach could furnish personalized treatment paradigms.
Regulatory Framework Enhancement: Crafting comprehensive, well-defined regulatory guidelines tailored to the realm of nanomedicine-based intranasal drug delivery is of paramount importance. Such harmonized regulations on a global scale are instrumental in the seamless translation of research findings into clinical practice.
Biomarker Integration: Enmeshing biomarker insights into the early diagnosis and monitoring of Alzheimer’s disease could be transformative. The nanotechnology-enabled intranasal delivery of diagnostic agents might herald more accurate disease assessments.
Combination Therapies: Leveraging the versatility of nanocarriers to facilitate the co-delivery of multiple therapeutic agents presents an avenue to harness synergistic effects and tackle the multifaceted nature of Alzheimer’s disease.
Long-Term Safety Endeavours: Rigorous, long-term investigations are indispensable to ensuring the safety and compatibility of nanocarriers, including their potential cumulative effects over extended periods.
In summary, the integration of nanotechnology into intranasal drug delivery opens up new possibilities for effective Alzheimer’s disease management. While challenges exist, collaborative efforts among researchers, clinicians, and regulatory authorities are pivotal in realizing the full potential of this ground-breaking approach. By skillfully addressing these challenges and moving towards future prospects, intranasal nanotechnology-based therapies may usher in a transformative era characterized by personalized, effective, and safer interventions for Alzheimer’s disease.

Author Contributions

Conceptualization: S.S. and S.D.; writing—original draft preparation: S.D. and S.J.; writing—review and editing: S.D., S.J., A.O., M.M. and S.S.; supervision: A.O., M.M. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

Sunil Jog is a registered PhD student (Reg No:371/Date: 26 Nov 2021) at SVKM’s Dr Bhanuben Nanavati College of Pharmacy. The manuscript being compiled is part of his PhD coursework. The manuscript under consideration is a review article. It is not related in any way to the current activities being conducted at Indoco Remedies 1.td, and hence, there are no conflicts of interest. The remaining authors declare no conflict of interest.

Abbreviation

ADAlzheimer’s disease
CNSCentral nervous system
BBBBlood–brain barrier
APOEApolipoprotein E
PSENPresenilin
APPAmyloid precursor protein
NMDAN-methyl-D-aspartate receptor
NFTNeurofibrillary tangles
CS-NPChitosan nanoparticles
ChATCholine acetyltransferase
PIPPiperine
PAMAMPoly (amidoamine)
PVAPolyvinyl alcohol
PLGApoly lactic co glycolic acid
HPMCHydroxypropyl methylcellulose
i.nIntranasal
i.vIntravenous
NPNanoparticles
SLNSolid lipid nanoparticles
NLCNanostructured lipid carrier
Lf-TMC-NPLactoferrin conjugated N-methylated chitosan nanoparticles
RSLNRisperidone solid lipid nanoparticles
PDIPolydispersibility Index
EEEntrapment efficiency
BuChEButyrylcholinesterases
AChEIAcetylcholinesterase inhibitor
ROSReactive oxygen species
ARIAAmyloid-related imaging abnormality
FDCFixed dose combination
OLEOpen label extension
PETPhoton emission topography
PEGPolytheylene glycol
AMTAdsorption mediated transcytosis
TgGlass transition temperature
PLGAPolylactide-co-glycolide
LNPLipid nanoparticles
MOFMetal organic framework
MDAMalonyldialdehyde
IVIVCIn vitro in vivo correlation
RHTRivastigmine
TNFTumour necrosis factor
ELISAEnzyme-linked immunosorbent assay
TRAILTNF-related apoptosis-inducing ligand
mi-RNAMicro ribonucleic acid
PLAPolylactic acid
AuNPGold nanoparticles
SPIONSuper paramagnetic iron oxide nanoparticles
HASHydroxy-α-sanshool
GHGalantamine hydroxide
OVALOvalalbumin
AUCArea under the curve
SNFSimulated nasal fluid
ACSFArtificial cerebrospinal fluid
PBSPhosphate buffer saline
DPPH2,2-diphenyl-1-picrylhydrazyl
FRAPFerric reducing ability of plasma
NENanoemulsion
GQRG alpha subunits
DTEDrug transport efficiency
DTPDrug targeting potential
MPP1-Methyl-4-phenylpyridinium
QDQuantum dots
WGAWheat germ agglutinin
ApoE4Apolipoprotein E4
GQDGraphene quantum dots
FTIRFourier transform infrared spectroscopy
HPLCHigh-performance liquid chromatography
CmaxMaximum concentration
EMAEuropean Medicines Agency
NBCDNon-biological complex drugs
INDInvestigational new drug
Aβ4242-amino acidβ amyloid

References

  1. Guo, T.; Zhang, D.; Zeng, Y.; Huang, T.Y.; Xu, H.; Zhao, Y. Molecular and cellular mechanisms underlying the pathogenesis of Alzheimer’s disease. Mol. Neurodegener. 2020, 15, 40. [Google Scholar] [CrossRef] [PubMed]
  2. Vinicius, M.; De Mello, C.; Vieira, L.; de Souza, L.C.; Gomes, K.; Carvalho, M. Alzheimer’s disease: Risk factors and potentially protective measures. J. Biomed. Sci. 2019, 26, 33. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6507104/ (accessed on 16 April 2021).
  3. Wang, L.; Yin, Y.-L.; Liu, X.-Z.; Shen, P.; Zheng, Y.-G.; Lan, X.-R.; Lu, C.-B.; Wang, J.-Z. Current understanding of metal ions in the pathogenesis of Alzheimer’s disease. Transl. Neurodegener. 2020, 9, 10. [Google Scholar] [CrossRef] [PubMed]
  4. Haque, R.U.; Levey, A.I. Alzheimer’sdisease: A clinical perspective and future nonhuman primate research opportunities. Proc. Natl. Acad. Sci. USA 2019, 116, 26224–26229. [Google Scholar] [CrossRef] [PubMed]
  5. Tiwari, S.; Atluri, V.; Kaushik, A.; Yndart, A.; Nair, M. Alzheimer’s disease: Pathogenesis, diagnostics, and therapeutics. Int. J. Nanomed. 2019, 14, 5541–5554. [Google Scholar] [CrossRef] [PubMed]
  6. Scheltens, P.; De Strooper, B.; Kivipelto, M.; Holstege, H.; Chételat, G.; Teunissen, C.E.; Cummings, J.; van der Flier, W.M. Alzheimer’s disease. Lancet 2021, 397, 1577–1590. [Google Scholar] [CrossRef] [PubMed]
  7. Long, J.M.; Holtzman, D.M. Alzheimer Disease: An Update on Pathobiology and Treatment Strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef] [PubMed]
  8. Kumar, A.; Singh, A. A review on Alzheimer’s disease pathophysiology and its management: An update. Pharmacol. Rep. 2015, 67, 195–203. [Google Scholar] [CrossRef]
  9. Avila, J.; Lucas, J.J.; Pérez, M.; Hernández, F. Role of Tau Protein in Both Physiological and Pathological Conditions. Physiol. Rev. 2004, 84, 361–384. [Google Scholar] [CrossRef]
  10. Pérez, M.J.; Jara, C.; Quintanilla, R.A. Contribution of Tau pathology to mitochondrial impairment in neurodegeneration. Front. Neurosci. 2018, 12, 441. [Google Scholar] [CrossRef]
  11. Fan, L.; Mao, C.; Hu, X.; Zhang, S.; Yang, Z.; Hu, Z.; Sun, H.; Fan, Y.; Dong, Y.; Yang, J.; et al. New Insights into the Pathogenesis of Alzheimer’s Disease. Front. Neurol. 2020, 10, 1312. [Google Scholar] [CrossRef] [PubMed]
  12. Sanabria-Castro, A.; Alvarado-Echeverría, I.; Monge-Bonilla, C. Molecular Pathogenesis of Alzheimer’s Disease: An Update. Ann. Neurosci. 2017, 24, 46–54. [Google Scholar] [CrossRef] [PubMed]
  13. Trejo-Lopez, J.A.; Yachnis, A.T.; Prokop, S. Neuropathology of Alzheimer’s Disease. Neurotherapeutics 2022, 19, 173–185. [Google Scholar] [CrossRef] [PubMed]
  14. Tchekalarova, J.; Tzoneva, R. Oxidative Stress and Aging as Risk Factors for Alzheimer’s Disease and Parkinson’s Disease: The Role of the Antioxidant Melatonin. Int. J. Mol. Sci. 2023, 24, 3022. [Google Scholar] [CrossRef]
  15. Cassidy, L.; Fernandez, F.; Johnson, J.B.; Naiker, M.; Owoola, A.G.; Broszczak, D.A. Oxidative stress in alzheimer’s disease: A review on emergent natural polyphenolic therapeutics. Complement. Ther. Med. 2019, 49, 102294. [Google Scholar] [CrossRef]
  16. Wang, W.Y.; Tan, M.S.; Yu, J.T.; Tan, L. Role of pro-inflammatory cytokines released from microglia in Alzheimer’s disease. Ann. Transl. Med. 2015, 3, 136. [Google Scholar] [CrossRef]
  17. Kawahara, M.; Kato-Negishi, M. Link between aluminum and the pathogenesis of Alzheimer’s disease: The integrationof the aluminum and amyloid cascadehypotheses. Int. J. Alzheimers Dis. 2011, 2011, 276393. [Google Scholar] [CrossRef]
  18. Chukwu, L.C.; Ekenjoku, J.A.; Ohadoma, S.C.; Olisa, C.L.; Okam, P.C.; Okany, C.C.; Ramalam, M.A.; Innocent, O.C. Advances in the pathogenesis of Alzheimer’s disease: A reevaluation of the Amyloid cascade hypothesis. World J. Adv. Res. Rev. 2023, 17, 882–904. [Google Scholar] [CrossRef]
  19. Sutinen, E.M.; Pirttilä, T.; Anderson, G.; Salminen, A.; Ojala, J.O. Pro-inflammatory interleukin-18 increases Alzheimer’s disease-associated amyloid-β production in human neuron-like cells. J. Neuroinflamm. 2012, 9, 199. [Google Scholar] [CrossRef]
  20. Heppner, F.L.; Ransohoff, R.M.; Becher, B. Immune attack: The role of inflammation in Alzheimer disease. Nat. Rev. Neurosci. 2015, 16, 358–372. [Google Scholar] [CrossRef]
  21. Briyal, S.; Ranjan, A.K.; Gulati, A. Oxidative stress: A target to treat Alzheimer’s disease and stroke. Neurochem. Int. 2023, 165, 105509. [Google Scholar] [CrossRef] [PubMed]
  22. Olufunmilayo, E.O.; Gerke-Duncan, M.B.; Holsinger, R.M.D. Oxidative Stress and Antioxidants in Neurodegenerative Disorders. Antioxidants 2023, 12, 517. [Google Scholar] [CrossRef] [PubMed]
  23. Shen, Z.; Bao, X.; Wang, R. Clinical PET imaging of microglial activation: Implications for microglial therapeutics in Alzheimer’s disease. Front. Aging Neurosci. 2018, 10, 314. [Google Scholar] [CrossRef] [PubMed]
  24. Yao, K.; Zu, H.B. Microglial polarization: Novel therapeutic mechanism against Alzheimer’s disease. Inflammopharmacology 2020, 28, 95–110. [Google Scholar] [CrossRef] [PubMed]
  25. Tondo, G.; Iaccarino, L.; Caminiti, S.P.; Presotto, L.; Santangelo, R.; Iannaccone, S.; Magnani, G.; Perani, D. The combined effects of microglia activation and brain glucose hypometabolism in early-onset Alzheimer’s disease. Alzheimer’s Res. Ther. 2020, 12, 50. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, M.-M.; Miao, D.; Cao, X.-P.; Tan, L.; Tan, L. Innate immune activation in Alzheimer’s disease. Ann. Transl. Med. 2018, 6, 177. [Google Scholar] [CrossRef] [PubMed]
  27. Fernández-Arjona, M.d.M.; Grondona, J.M.; Fernández-Llebrez, P.; López-Ávalos, M.D. Microglial Morphometric Parameters Correlate with the Expression Level of IL-1β, and Allow Identifying Different Activated Morphotypes. Front. Cell. Neurosci. 2019, 13, 472. [Google Scholar] [CrossRef]
  28. Chakraborty, B.; Mukerjee, N.; Maitra, S.; Zehravi, M.; Mukherjee, D.; Ghosh, A.; Massoud, E.E.S.; Rahman, M.H. Therapeutic Potential of Different Natural Products for the Treatment of Alzheimer’s Disease. Oxidative Med. Cell. Longev. 2022, 2022, 6873874. [Google Scholar] [CrossRef]
  29. Marucci, G.; Buccioni, M.; Ben, D.D.; Lambertucci, C.; Volpini, R.; Amenta, F. Efficacy of acetylcholinesterase inhibitors in Alzheimer’s disease. Neuropharmacology 2021, 190, 108352. [Google Scholar] [CrossRef]
  30. Tayeb, H.O.; Yang, H.D.; Price, B.H.; Tarazi, F.I. Pharmacotherapies for Alzheimer’s disease: Beyond cholinesterase inhibitors. Pharmacol. Ther. 2012, 134, 8–25. [Google Scholar] [CrossRef]
  31. Dominik, G.; Glinz, D.; Gloy, V.L.; Monsch, A.U.; Kressig, R.W.; Patel, C.; McCord, K.A.; Ademy, Z.; Tomonaga, Y.; Schwenkglenks, M.; et al. Acetylcholinesterase inhibitors combined with memantine for moderate to severe Alzheimer’s disease: A meta-analysis. Swiss Med. Wkly. 2019, 149, w20093. [Google Scholar] [CrossRef]
  32. Kuns, B.; Rosani, A.; Varghese, D.; Kuns, B.; Rosani, A.; Varghese, D. Memantine; StatPearls Publishing: St. Petersburg, FA, USA, 2022. [Google Scholar]
  33. Rosini, M.; Simoni, E.; Caporaso, R.; Minarini, A.A. Multitarget strategies in Alzheimer’s disease: Benefits and challenges on the road to therapeutics. Future Med. Chem. 2016, 8, 697–711. [Google Scholar] [CrossRef] [PubMed]
  34. Deardorff, W.J.; Grossberg, G.T. A fixed-dose combination of memantine extended-release and donepezil in the treatment of moderate-to-severe Alzheimer’s disease. Drug Des. Devel. Ther. 2016, 10, 3267–3279. [Google Scholar] [CrossRef] [PubMed]
  35. Tang, B.; Wang, Y.; Ren, J. Basic information about memantine and its treatment of Alzheimer’s disease and other clinical applications. Ibrain 2023, 9, 340–348. [Google Scholar] [CrossRef] [PubMed]
  36. Loureiro, J.C.; Silva, L.F.A.L.; Pais, M.V.; Forlenza, O.V. Anti-amyloid agents for treating incipient Alzheimer’s disease: A new hope? Braz. J. Psychiatry 2022, 44, 368–369. [Google Scholar] [CrossRef]
  37. Bespalov, A.; Courade, J.P.; Khiroug, L.; Terstappen, G.C.; Wang, Y. A call for better understanding of target engagement in Tau antibody development. Drug Discov. Today 2022, 27, 103338. [Google Scholar] [CrossRef] [PubMed]
  38. Padda, I.S.; Parmar, M. Aducanumab; StatPearls Publishing: St. Petersburg, FA, USA, 2023. [Google Scholar]
  39. Vaz, M.; Silva, V.; Monteiro, C.; Silvestre, S. Role of Aducanumab in the Treatment of Alzheimer’s Disease: Challenges and Opportunities. Clin. Interv. Aging 2022, 17, 797–810. [Google Scholar] [CrossRef]
  40. Wojtunik-Kulesza, K.; Rudkowska, M.; Orzeł-Sajdłowska, A. Aducanumab—Hope or Disappointment for Alzheimer’s Disease. Int. J. Mol. Sci. 2023, 24, 4367. [Google Scholar] [CrossRef]
  41. Haeberlein, S.B.; Aisen, P.S.; Barkhof, F.; Chalkias, S.; Chen, T.; Cohen, S.; Dent, G.; Hansson, O.; Harrison, K.; Hehn, C.; et al. Two Randomized Phase 3 Studies of Aducanumab in Early Alzheimer’s Disease. J. Prev. Alzheimer’s Dis. 2022, 9, 197–210. [Google Scholar] [CrossRef]
  42. Rahman, A.; Hossen, M.A.; Chowdhury, M.F.I.; Bari, S.; Tamanna, N.; Sultana, S.S.; Haque, S.N.; Al Masud, A.; Saif-Ur-Rahman, K.M. Aducanumab for the treatment of Alzheimer’s disease: A systematic review. Psychogeriatrics 2023, 23, 512–522. [Google Scholar] [CrossRef]
  43. Management, P. Letter to Editor Accelerated Approval of Highly Expensive Disease-modifying Agents: Lessons Learned from the Aducanumab Approval. J. Pharmacoecon. Pharm. Manag. 2022, 8, 2–5. [Google Scholar]
  44. Brockmann, R.; Nixon, J.; Love, B.L.; Yunusa, I. Impacts of FDA approval and Medicare restriction on antiamyloid therapies for Alzheimer’s disease: Patient outcomes, healthcare costs, and drug development. Lancet Reg. Health-Am. 2023, 20, 100467. [Google Scholar] [CrossRef] [PubMed]
  45. Hogan, D.; Frank, C. Challenges with new treatments for Alzheimer disease: Accelerated approval of aducanumab in the United States raises questions. Can. Fam. Physician 2023, 69, 160–161. [Google Scholar] [CrossRef] [PubMed]
  46. Levy, H.B. Accelerated Approval of Aducanumab: Where Do We Stand Now? Ann. Pharmacother. 2021, 56, 736–739. [Google Scholar] [CrossRef]
  47. Qin, Q.; Tang, Y. Lecanemab: The game changer in the ongoing fight to treat Alzheimer’s disease? Human Brain 2022, 2, 1–4. [Google Scholar] [CrossRef]
  48. Hardy, J.; Mummery, C. An anti-amyloid therapy works for Alzheimer’s disease: Why has it taken so long and what is next? Brain 2023, 146, 1240–1242. [Google Scholar] [CrossRef]
  49. McDade, E.; Cummings, J.L.; Dhadda, S.; Swanson, C.J.; Reyderman, L.; Kanekiyo, M.; Koyama, A.; Irizarry, M.; Kramer, L.D.; Bateman, R.J. Lecanemab in patients with early Alzheimer’s disease: Detailed results on biomarker, cognitive, and clinical effects from the randomized and open-label extension of the phase 2 proof-of-concept study. Alzheimer’s Res. Ther. 2022, 14, 191. [Google Scholar] [CrossRef]
  50. Honig, L.S.; Barakos, J.; Dhadda, S.; Kanekiyo, M.; Reyderman, L.; Irizarry, M.; Kramer, L.D.; Swanson, C.J.; Sabbagh, M. ARIA in Alzheimer’s disease background. Alzheimer’s Dement. 2023, 9, e12377. [Google Scholar] [CrossRef]
  51. Gautam, A.S.; Pandey, S.K.; Lasure, V.; Dubey, S. Monoclonal antibodies for the management of central nervous system diseases: Clinical success and future strategies. Expert Opin. Biol. Ther. 2023, 23, 603–618. [Google Scholar] [CrossRef]
  52. Revheim, M.; Carlsen, P.F.H.; Costa, T.; Alavi, A.; Kepp, K.P.; Sensi, S.L.; Perry, G.; Robakis, N.K.; Barrio, J.R.; Vissel, B. Passive Alzheimer’s immunotherapy: A promising or uncertain option? Ageing Res. Rev. 2023, 90, 101996. [Google Scholar]
  53. Lois, F.; Lavand, P.; Leonard, D.; Remue, C.; Bellemans, V.; First, A.K. Background Connect with Wiley. Photodermatol. Photoimmunol. Photomed. 2019, 29, 4–6. [Google Scholar]
  54. Bateman, R.J.; Cummings, J.; Schobel, S.; Salloway, S.; Vellas, B.; Boada, M.; Black, S.E.; Blennow, K.; Fontoura, P.; Klein, G.; et al. An anti-amyloid monoclonal antibody with potential disease-modifying effects in early Alzheimer’s disease. Alzheimer’s Res. Ther. 2022, 14, 178. [Google Scholar] [CrossRef] [PubMed]
  55. Valiukas, Z.; Ephraim, R.; Tangalakis, K.; Davidson, M.; Apostolopoulos, V.; Feehan, J. Immunotherapies for Alzheimer’s Disease—A Review. Vaccines 2022, 10, 1527. [Google Scholar] [CrossRef] [PubMed]
  56. Song, C.; Shi, J.; Zhang, P.; Zhang, Y.; Xu, J.; Zhao, L.; Zhang, R.; Wang, H.; Chen, H. I Immunotherapy for Alzheimer’s disease: Targeting β-amyloid and beyond. Transl. Neurodegener. 2022, 11, 18. [Google Scholar] [CrossRef]
  57. Hoque, M.; Samanta, A.; Sahajada, S.; Alam, M.; Zughaibi, T.A. Neuroscience & Biobehavioral Reviews Nanomedicine-based immunotherapy for Alzheimer’s disease. Neurosci. Biobehav. Rev. 2023, 144, 104973. [Google Scholar]
  58. Ramanan, V.K.; Day, G.S. Molecular Neurodegeneration Anti-amyloid therapies for Alzheimer disease: Finally, good news for patients. Mol. Neurodegener. 2023, 18, 42. [Google Scholar] [CrossRef]
  59. Abushouk, A.I.; Elmaraezy, A.; Aglan, A.; Salama, R.; Fouda, S.; Fouda, R.; AlSafadi, A.M. Bapineuzumab for mild to moderate Alzheimer’s disease: A meta-analysis of randomized controlled trials. BMC Neurol. 2017, 17, 66. [Google Scholar] [CrossRef]
  60. Godyń, J.; Jończyk, J.; Panek, D.; Malawska, B. Therapeutic strategies for Alzheimer’s disease in clinical trials. Pharmacol. Rep. 2016, 68, 127–138. [Google Scholar] [CrossRef]
  61. Ostrowitzki, S.; Bittner, T.; Sink, K.M.; Mackey, H.; Rabe, C.; Honig, L.S.; Cassetta, E.; Woodward, M.; Boada, M.; Van Dyck, C.H.; et al. Evaluating the Safety and Efficacy of Crenezumab vs Placebo in Adults with Early Alzheimer Disease: Two Phase 3 Randomized Placebo-Controlled Trials. JAMA Neurol. 2022, 79, 1113–1121. [Google Scholar] [CrossRef]
  62. Landen, J.W.; Andreasen, N.; Cronenberger, C.L.; Schwartz, P.F.; Hanson, A.B.; Östlund, H.; Sattler, C.A.; Binneman, B.; Bednar, M.M. Ponezumab in mild-to-moderate Alzheimer’s disease: Randomized phase II PET-PIB study. Alzheimer’s Dement. Transl. Res. Clin. Interv. 2017, 3, 393–401. [Google Scholar] [CrossRef]
  63. Teng, E.; Manser, P.T.; Pickthorn, K.; Brunstein, F.; Blendstrup, M.; Bohorquez, S.S.; Wildsmith, K.R.; Toth, B.; Dolton, M.; Ramakrishnan, V.; et al. Safety and Efficacy of Semorinemab in Individuals with Prodromal to Mild Alzheimer Disease: A Randomized Clinical Trial. JAMA Neurol. 2022, 79, 758–767. [Google Scholar] [CrossRef] [PubMed]
  64. Dam, T.; Boxer, A.L.; Golbe, L.I.; Höglinger, G.U.; Morris, H.R.; Litvan, I.; Lang, A.E.; Corvol, J.-C.; Aiba, I.; Grundman, M.; et al. Safety and efficacy of anti-tau monoclonal antibody gosuranemab in progressive supranuclear palsy: A phase 2, randomized, placebo-controlled trial. Nat. Med. 2021, 27, 1451–1457. [Google Scholar] [CrossRef] [PubMed]
  65. Florian, H.; Wang, D.; Arnold, S.E.; Boada, M.; Guo, Q.; Jin, Z.; Zheng, H.; Fisseha, N.; Kalluri, H.V.; Rendenbach-Mueller, B.; et al. Tilavonemab in early Alzheimer’s disease: Results from a phase 2, randomized, double-blind study. Brain 2023, 146, 2275–2284. [Google Scholar] [CrossRef] [PubMed]
  66. Unnisa, A.; Greig, N.; Kamal, M. Nanotechnology-based gene therapy as a credible tool in the treatment of Alzheimer’s disease. Neural Regen. Res. 2023, 18, 2127–2133. [Google Scholar] [CrossRef] [PubMed]
  67. Harilal, S.; Jose, J.; Parambi, D.G.T.; Kumar, R.; Mathew, G.E.; Uddin, M.S.; Kim, H.; Mathew, B. Advancements in nanotherapeutics for Alzheimer’s disease: Current perspectives. J. Pharm. Pharmacol. 2019, 71, 1370–1383. [Google Scholar] [CrossRef] [PubMed]
  68. Medicinal, F.; Hassan, N.A.; Alshamari, A.K.; Hassan, A.A.; Elharrif, M.G. Advances on Therapeutic Strategies for Alzheimer’s Disease: From Medicinal Plant to Nanotechnologyg. Molecules 2022, 27, 4839. [Google Scholar]
  69. Ming, 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]
  70. Ferreira, M.D.; Duarte, J.; Veiga, F.; Paiva-Santos, A.C.; Pires, P.C. Nanosystems for Brain Targeting of Antipsychotic Drugs: An Update on the Most Promising Nanocarriers for Increased Bioavailability and Therapeutic Efficacy. Pharmaceutics 2023, 15, 678. [Google Scholar] [CrossRef]
  71. Tiwari, V.; Tiwari, A.; Sharma, A.; Kumar, M.; Kaushik, D.; Sagadevan, S. An optimistic approach to nanotechnology in Alzheimer’s disease management: An overview. J. Drug Deliv. Sci. Technol. 2023, 86, 104722. [Google Scholar] [CrossRef]
  72. 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. [Google Scholar] [CrossRef]
  73. 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] [PubMed]
  74. Di Filippo, L.D.; De Carvalho, S.G.; Duarte, J.L.; Luiz, M.T.; Paes Dutra, J.A.; De Paula, G.A.; Chorilli, M.; Conde, J. A receptor-mediated landscape of druggable and targeted nanomaterials for gliomas. Mater. Today Bio 2023, 20, 100671. [Google Scholar] [CrossRef] [PubMed]
  75. Sharma, S.; Dang, S. Nanocarrier-Based Drug Delivery to Brain: Interventions of Surface Modification. Curr. Neuropharmacol. 2022, 21, 517–535. [Google Scholar] [CrossRef] [PubMed]
  76. Ramalho, M.J.; Andrade, S.; Loureiro, J.A.; Pereira, M.D.C. Nanotechnology to improve the Alzheimer’s disease therapy with natural compounds. Drug Deliv. Transl. Res. 2020, 10, 380–402. [Google Scholar] [CrossRef]
  77. Puranik, N.; Yadav, D.; Song, M. Advancements in the Application of Nanomedicine in Alzheimer’s Disease: A Therapeutic Perspective. Int. J. Mol. Sci. 2023, 24, 14044. [Google Scholar] [CrossRef] [PubMed]
  78. Rajput, A.; Pingale, P.; Dhapte-Pawar, V. Nasal delivery of neurotherapeutics via nanocarriers: Facets, aspects, and prospects. Front. Pharmacol. 2022, 13, 979682. [Google Scholar] [CrossRef]
  79. 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, 433–468. [Google Scholar] [CrossRef]
  80. Kou, D.; Gao, Y.; Li, C.; Zhou, D.; Lu, K.; Wang, N.; Zhang, R.; Yang, Z.; Zhou, Y.; Chen, L.; et al. Intranasal Pathway for Nanoparticles to Enter the Central Nervous System. Nano Lett. 2023, 23, 5381–5390. [Google Scholar] [CrossRef]
  81. Chu, J.; Zhang, W.; Liu, Y.; Gong, B.; Ji, W.; Yin, T.; Gao, C.; Liangwen, D.; Hao, M.; Chen, C.; et al. Biomaterials-based anti-inflammatory treatment strategies for Alzheimer’s disease. Neural Regen. Res. 2024, 19, 100–115. [Google Scholar] [CrossRef]
  82. Saucier-Sawyer, J.K.; Deng, Y.; Seo, Y.-E.; Cheng, C.J.; Zhang, J.; Quijano, E.; Saltzman, W.M. Systemic delivery of blood-brain barrier-targeted polymeric nanoparticles enhances delivery to brain tissue. J. Drug Target. 2015, 23, 736–749. [Google Scholar] [CrossRef]
  83. Zhang, W.; Mehta, A.; Tong, Z.; Esser, L.; Voelcker, N.H. Development of Polymeric Nanoparticles for Blood–Brain Barrier Transfer—Strategies and Challenges. Adv. Sci. 2021, 8, 2003937. [Google Scholar] [CrossRef] [PubMed]
  84. Colson, Y.L.; Grinstaff, M.W. Biologically responsive polymeric nanoparticles for drug delivery. Adv. Mater. 2012, 24, 3878–3886. [Google Scholar] [CrossRef] [PubMed]
  85. Hoyos-Ceballos, G.P.; Ruozi, B.; Ottonelli, I.; Da Ros, F.; Vandelli, M.A.; Forni, F.; Daini, E.; Vilella, A.; Zoli, G.; Tosi, G.; et al. PLGA-PEG-Ang–2 nanoparticles for blood–brain barrier crossing: Proof-of-concept study. Pharmaceutics 2020, 12, 72. [Google Scholar] [CrossRef] [PubMed]
  86. Li, H.; Tong, Y.; Bai, L.; Ye, L.; Zhong, L.; Duan, X.; Zhu, Y. Lactoferrin functionalized PEG-PLGA nanoparticles of shikonin for brain targeting therapy of glioma. Int. J. Biol. Macromol. 2018, 107, 204–211. [Google Scholar] [CrossRef] [PubMed]
  87. Martin, J.; Desfoux, A.; Martinez, J.; Amblard, M.; Mehdi, A. Bottom-up strategies for the synthesis of peptide-based polymers. Prog. Polym. Sci. 2021, 115, 101377. [Google Scholar] [CrossRef]
  88. La Barbera, L.; Mauri, E.; D’Amelio, M.; Gori, M. Functionalization strategies of polymeric nanoparticles for drug delivery in Alzheimer’s disease: Current trends and future perspectives. Front. Neurosci. 2022, 16, 939855. [Google Scholar] [CrossRef]
  89. Caprifico, A.E.; Foot, P.J.S.; Polycarpou, E.; Calabrese, G. Overcoming the blood-brain barrier: Functionalised chitosan nanocarriers. Pharmaceutics 2020, 12, 1013. [Google Scholar] [CrossRef]
  90. Zhu, X.; Jin, K.; Huang, Y.; Pang, Z. Brain drug delivery by adsorption-mediated transcytosis. In Brain Targeted Drug Delivery Systems: A Focus on Nanotechnology and Nanoparticulates; Academic Press: Cambridge, MA, USA, 2018; pp. 159–183. [Google Scholar] [CrossRef]
  91. Cano, A.; Sánchez-López, E.; Ettcheto, M.; López-Machado, A.; Espina, M.; Souto, E.B.; Galindo, R.; Camins, A.; García, M.L.; Turowski, P. Current advances in the development of novel polymeric nanoparticles for the treatment of neurodegenerative diseases. Nanomedicine 2020, 15, 1239–1261. [Google Scholar] [CrossRef]
  92. Kamaly, N.; Yameen, B.; Wu, J.; Farokhzad, O.C. Degradable controlled-release polymers and polymeric nanoparticles: Mechanisms of controlling drug release. Chem. Rev. 2016, 116, 2602–2663. [Google Scholar] [CrossRef]
  93. Elmowafy, M.; Shalaby, K.; Elkomy, M.H.; Alsaidan, O.A.; Gomaa, H.A.M.; Abdelgawad, M.A.; Mostafa, E.M. Polymeric Nanoparticles for Delivery of Natural Bioactive Agents: Recent Advances and Challenges. Polymers 2023, 15, 1123. [Google Scholar] [CrossRef]
  94. Mittal, G.; Carswell, H.; Brett, R.; Currie, S.; Kumar, M.N.V.R. Development and evaluation of polymer nanoparticles for oral delivery of estradiol to rat brain in a model of Alzheimer’s pathology. J. Control. Release 2011, 150, 220–228. [Google Scholar] [CrossRef] [PubMed]
  95. 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] [PubMed]
  96. Bhavna, B.; Shadab; Ali, M.; Baboota, S.; Sahni, J.K.; Bhatnagar, A.; Ali, J. Preparation, characterization, in vivo biodistribution and pharmacokinetic studies of donepezil-loaded PLGA nanoparticles for brain targeting. Drug Dev. Ind. Pharm. 2014, 40, 278–287. [Google Scholar] [CrossRef]
  97. Wilson, B.; Samanta, M.K.; Muthu, M.S.; Vinothapooshan, G. Design and evaluation of chitosan nanoparticles as novel drug carrier for the delivery of rivastigmine to treat Alzheimer’s disease. Ther. Deliv. 2011, 2, 599–609. [Google Scholar] [CrossRef] [PubMed]
  98. Jagaran, K.; Singh, M. Lipid Nanoparticles: Promising Treatment Approach for Parkinson’s Disease. Int. J. Mol. Sci. 2022, 23, 9361. [Google Scholar] [CrossRef] [PubMed]
  99. 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]
  100. Marques, A.C.; Costa, P.C.; Velho, S.; Amaral, M.H. Lipid Nanoparticles Functionalized with Antibodies for Anticancer Drug Therapy. Pharmaceutics 2023, 15, 216. [Google Scholar] [CrossRef]
  101. Gugleva, V.; Andonova, V. Drug delivery to the brain—Lipid nanoparticles-based approach. Pharmacia 2023, 70, 113–120. [Google Scholar] [CrossRef]
  102. Yan, D.; Qu, X.; Chen, M.; Wang, J.; Li, X.; Zhang, Z.; Liu, Y.; Kong, L.; Yu, Y.; Ju, R.; et al. Functionalized curcumin/ginsenoside Rb1 dual-loaded liposomes: Targeting the blood-brain barrier and improving pathological features associated in APP/PS-1 mice. J. Drug Deliv. Sci. Technol. 2023, 86, 104633. [Google Scholar] [CrossRef]
  103. Sokolik, V.V.; Berchenko, O.G. The cumulative effect of the combined action of miR-101 and curcumin in a liposome on a model of Alzheimer’s disease in mononuclear cells. Front. Cell. Neurosci. 2023, 17, 1169980. [Google Scholar] [CrossRef]
  104. Andrade, S.; Pereira, M.C.; Loureiro, J.A. Caffeic acid loaded into engineered lipid nanoparticles for Alzheimer’s disease therapy. Colloids Surf. B Biointerfaces 2023, 225, 113270. [Google Scholar] [CrossRef] [PubMed]
  105. Shivananjegowda, M.G.; Hani, U.; Osmani, R.A.M.; Alamri, A.H.; Ghazwani, M.; Alhamhoom, Y.; Rahamathulla, M.; Paranthaman, S.; Gowda, D.V.; Siddiqua, A. Development and Evaluation of Solid Lipid Nanoparticles for the Clearance of Aβ in Alzheimer’s Disease. Pharmaceutics 2023, 15, 221. [Google Scholar] [CrossRef] [PubMed]
  106. 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] [PubMed]
  107. Raju, M.; Kunde, S.S.; Auti, S.T.; Kulkarni, Y.A.; Wairkar, S. Berberine loaded nanostructured lipid carrier for Alzheimer’s disease: Design, statistical optimization and enhanced in vivo performance. Life Sci. 2021, 285, 2021–2023. [Google Scholar] [CrossRef] [PubMed]
  108. Ismail, N.; Ismail, M.; Azmi, N.H.; Abu Bakar, M.F.; Yida, Z.; Abdullah, M.A.; Basri, H. Thymoquinone-rich fraction nanoemulsion (TQRFNE) decreases Aβ40 and Aβ42 levels by modulating APP processing, up-regulating IDE and LRP1, and down-regulating BACE1 and RAGE in response to high fat/cholesterol diet-induced rats. Biomed. Pharmacother. 2017, 95, 780–788. [Google Scholar] [CrossRef] [PubMed]
  109. Duong, V.A.; Nguyen, T.T.L.; Maeng, H.J. Recent Advances in Intranasal Liposomes for Drug, Gene, and Vaccine Delivery. Pharmaceutics 2023, 15, 207. [Google Scholar] [CrossRef] [PubMed]
  110. Gyanani, V.; Goswami, R. Key Design Features of Lipid Nanoparticles and Electrostatic Charge-Based Lipid Nanoparticle Targeting. Pharmaceutics 2023, 15, 1184. [Google Scholar] [CrossRef]
  111. Hernandez, C.; Shukla, S. Liposome based drug delivery as a potential treatment option for Alzheimer’s disease. Neural Regen. Res. 2022, 17, 1190–1198. [Google Scholar] [CrossRef]
  112. Satapathy, M.K.; Yen, T.-L.; Jan, J.-S.; Tang, R.-D.; Wang, J.-Y.; Taliyan, R.; Yang, C.-H. Solid lipid nanoparticles (Slns): An advanced drug delivery system targeting brain through bbb. Pharmaceutics 2021, 13, 1183. [Google Scholar] [CrossRef]
  113. Mosallaei, N.; Jaafari, M.R.; Hanafi-Bojd, M.Y.; Golmohammadzadeh, S.; Malaekeh-Nikouei, B. Docetaxel-loaded solid lipid nanoparticles: Preparation, characterization, in vitro, and in vivo evaluations. J. Pharm. Sci. 2013, 102, 1994–2004. [Google Scholar] [CrossRef]
  114. Souto, E.B.; Fangueiro, J.F.; Fernandes, A.R.; Cano, A.; Sanchez-Lopez, E.; Garcia, M.L.; Severino, P.; Paganelli, M.O.; Chaud, M.V.; Silva, A.M. Physicochemical and biopharmaceutical aspects influencing skin permeation and role of SLN and NLC for skin drug delivery. Heliyon 2022, 8, e08938. [Google Scholar] [CrossRef] [PubMed]
  115. Patel, M.; Souto, E.B.; Singh, K.K. Advances in brain drug targeting and delivery: Limitations and challenges of solid lipid nanoparticles. Expert Opin. Drug Deliv. 2013, 10, 889–905. [Google Scholar] [CrossRef] [PubMed]
  116. El-Nashar, H.A.S.; Abbas, H.; Zewail, M.; Noureldin, M.H.; Ali, M.M.; Shamaa, M.M.; Khattab, M.A. Neuroprotective Effect of Artichoke-Based Nanoformulation in Sporadic Alzheimer’s Disease Mouse Model: Focus on Antioxidant, Anti-Inflammatory, and Amyloidogenic Pathways. Pharmaceuticals 2022, 15, 1202. [Google Scholar] [CrossRef] [PubMed]
  117. 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. 2017, 45, 364–371. [Google Scholar] [CrossRef] [PubMed]
  118. Patr, A.B.; Prata, M.; Nadhman, A.; Chintamaneni, P.K.; Fonte, P. Solid Lipid Nanoparticles vs. Nanostructured Lipid Carriers: A Comparative Review. Pharmaceutics 2023, 15, 1593. [Google Scholar]
  119. Nasar, S.; Afzal, O.; Altamimi, A.S.A.; Ather, H.; Sultana, S.; Almalki, W.H.; Bharti, P.; Sahoo, A.; Dwivedi, K.; Khan, G.; et al. Nanomedicine in the Management of Alzheimer’s Disease: State-of-the-Art Biomedicines. Biomedicines 2023, 11, 1752. [Google Scholar] [CrossRef]
  120. Garg, J.; Pathania, K.; Sah, S.P.; Pawar, S.V. Nanostructured lipid carriers: A promising drug carrier for targeting brain tumours. Future J. Pharm. Sci. 2022, 8, 25. [Google Scholar] [CrossRef]
  121. Mendes, I.T.; Carvalho, F.C.; Bonfilio, R.; Pereira, G.R. Colloids and Surfaces B: Biointerfaces Development and characterization of nanostructured lipid carrier-based gels for the transdermal delivery of donepezil. Colloids Surf. B Biointerfaces 2019, 177, 274–281. [Google Scholar] [CrossRef]
  122. Chauhan, M.K. Optimization and characterization of rivastigmine nanolipid carrier loaded transdermal patches for the treatment of dementia. Chem. Phys. Lipids 2019, 224, 104794. [Google Scholar] [CrossRef]
  123. Hamano, N.; Li, S.; Chougule, M.; Shoyele, S.A.; Alexander, A. 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]
  124. Battaglia, L.; Gallarate, M. Lipid nanoparticles: State of the art, new preparation methods and challenges in drug delivery. Expert Opin. Drug Deliv. 2012, 9, 497–508. [Google Scholar] [CrossRef] [PubMed]
  125. Nirale, P.; Paul, A.; Yadav, K.S. Nanoemulsions for targeting the neurodegenerative diseases: Alzheimer’s, Parkinson’s and Prion’s. Life Sci. 2020, 245, 117394. [Google Scholar] [CrossRef] [PubMed]
  126. Haider, F.; Khan, S.; Gaba, B.; Alam, T. Optimization of rivastigmine nanoemulsion for enhanced brain delivery: In-vivo and toxicity evaluation. J. Mol. Liq. 2018, 255, 384–396. [Google Scholar] [CrossRef]
  127. Yukuyama, M.N.; Ishida, K.; de Araujo, G.L.B.; de Castro Spadari, C.; de Souza, A.; Löbenberg, R.; Henostroza, M.A.B.; Folchini, B.R.; Peroni, C.M.; Peters, M.C.C.; et al. Rational design of oral flubendazole-loaded nanoemulsion for brain delivery in cryptococcosis. Colloids Surf. A Physicochem. Eng. Asp. 2021, 630, 127631. [Google Scholar] [CrossRef]
  128. Atinderpal, K.; Kuldeep, N.; Sukriti, S.; Amit, T.; Shweta, D. Memantine nanoemulsion: A new approach to treat Alzheimer’s disease. J. Microencapsul. 2020, 37, 355–365. [Google Scholar] [CrossRef]
  129. Line, S.; Guyon, L.; Maurel, M.; Verdié, P.; Davis, A.; Corvaisier, S.; Lisowski, V.; Dallemagne, P.; Groo, A.-C.; Malzert-Fréon, A. Active Targeted Nanoemulsions for Repurposing of Tegaserod in Alzheimer’s Disease Treatment. Pharmaceutics 2021, 13, 1626. [Google Scholar]
  130. Valmiki, V.C.; Gangadhara, A. Review on metal nanoparticles as nanocarriers: Current challenges and perspectives in drug delivery systems. Emergent Mater. 2022, 5, 1593–1615. [Google Scholar] [CrossRef]
  131. Sharma, B.; Pervushin, K. Magnetic nanoparticles as in vivo tracers for Alzheimer’s disease. Magnetochemistry 2020, 6, 13. [Google Scholar] [CrossRef]
  132. Sawicki, K.; Czajka, M.; Matysiak-Kucharek, M.; Fal, B.; Drop, B.; Męczyńska-Wielgosz, S.; Sikorska, K.; Kruszewski, M.; Kapka-Skrzypczak, L. Toxicity of metallic nanoparticles in the central nervous system. Nanotechnol. Rev. 2019, 8, 175–200. [Google Scholar] [CrossRef]
  133. Zhao, J.; Xu, N.; Yang, X.; Ling, G.; Zhang, P. The roles of gold nanoparticles in the detection of amyloid-β peptide for Alzheimer’s disease. Colloid Interface Sci. Commun. 2022, 46, 100579. [Google Scholar] [CrossRef]
  134. Barrier, B.; Wong, K.H.; Riaz, M.K.; Xie, Y.; Zhang, X.; Liu, Q. Review of Current Strategies for Delivering Alzheimer’s Disease Drugs across the Blood-Brain Barrier. Int. J. Mol. Sci. 2019, 20, 381. [Google Scholar] [CrossRef]
  135. Kim, Y.; Park, J.; Lee, H.; Nam, J. How Do the Size, Charge and Shape of Nanoparticles Affect Amyloid β Aggregation on Brain Lipid Bilayer? Sci. Rep. 2016, 6, 19548. [Google Scholar] [CrossRef] [PubMed]
  136. Khalil, A.T.; Ullah, F. Biosynthesized Metal Nanoparticles as Potential Alzheimer’s Disease Therapeutics; Elsevier Inc.: Amsterdam, The Netherlands, 2020. [Google Scholar] [CrossRef]
  137. Carro, C.E.; Pilozzi, A.R.; Huang, X. Nanoneurotoxicity and Potential Nanotheranostics for Alzheimer’s Disease. EC Pharmacol. Toxicol. 2019, 7, 1–7. [Google Scholar] [PubMed]
  138. Medici, S.; Peana, M.; Pelucelli, A.; Zoroddu, M.A. An updated overview on metal nanoparticles toxicity. Semin. Cancer Biol. 2021, 76, 17–26. [Google Scholar] [CrossRef] [PubMed]
  139. Kulkarni, N.; Muddapur, U. Biosynthesis of metal nanoparticles: A review. J. Nanotechnol. 2014, 2014, 510246. [Google Scholar] [CrossRef]
  140. Tajahmadi, S.; Molavi, H.; Ahmadijokani, F.; Shamloo, A.; Shojaei, A.; Sharifzadeh, M.; Rezakazemi, M.; Fatehizadeh, A.; Aminabhavi, T.M.; Arjmand, M. Metal-organic frameworks: A promising option for the diagnosis and treatment of Alzheimer’s disease. J. Control. Release 2023, 353, 1–29. [Google Scholar] [CrossRef] [PubMed]
  141. Elmonem, H.A.A.; Morsi, R.M.; Mansour, D.S.; El-Sayed, E.S.R. Myco-fabricated ZnO nanoparticles ameliorate neurotoxicity in mice model of Alzheimer’s disease via acetylcholinesterase inhibition and oxidative stress reduction. BioMetals 2023, 36, 1391–1404. [Google Scholar] [CrossRef] [PubMed]
  142. Yang, L.; Chen, Y.; Jia, Z.; Yuan, X.; Liu, J. Electrostatic assembly of gold nanoparticle and metal-organic framework nanoparticles attenuates amyloid β aggregate-mediated neurotoxicity. J. Mater. Chem. B 2023, 11, 4453–4463. [Google Scholar] [CrossRef]
  143. 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]
  144. Yin, Z.; Zhang, Z.; Gao, D.; Luo, G.; Ma, T.; Wang, Y.; Lu, L.; Gao, X. Stepwise Coordination-Driven Metal-Phenolic Nanoparticle as a Neuroprotection Enhancer for Alzheimer’s Disease Therapy. ACS Appl. Mater. Interfaces 2022, 15, 524–540. [Google Scholar] [CrossRef]
  145. Tang, R.; Yuan, X.; Jia, Z.; Yang, F.; Ye, G.; Liu, J. Ruthenium Dioxide Nanoparticles Treat Alzheimer’s Disease by Inhibiting Oxidative Stress and Alleviating Neuroinflammation. ACS Appl. Nano Mater. 2023, 6, 11661–11678. [Google Scholar] [CrossRef]
  146. Ling, T.S.; Chandrasegaran, S.; Xuan, L.Z.; Suan, T.L.; Elaine, E.; Visva Nathan, D.; Chai, Y.H.; Gunasekaran, B.; Salvamani, S. Review Article the Potential Benefits of Nanotechnology in Treating Alzheimer’s Disease. BioMed Res. Int. 2021, 2021, 5550938. [Google Scholar] [CrossRef] [PubMed]
  147. Di Stefano, A.; Iannitelli, A.; Laserra, S.; Sozio, P. Drug delivery strategies for Alzheimer’s disease treatment. Expert Opin. Drug Deliv. 2011, 8, 581–603. [Google Scholar] [CrossRef]
  148. Bellettato, C.M.; Scarpa, M. Possible strategies to cross the blood–brain barrier. Ital. J. Pediatr. 2018, 44, 127–133. [Google Scholar] [CrossRef]
  149. Sánchez-Dengra, B.; González-Álvarez, I.; Bermejo, M.; González-Álvarez, M. Access to the CNS: Strategies to overcome the BBB. Int. J. Pharm. 2023, 636, 122759. [Google Scholar] [CrossRef] [PubMed]
  150. Khan, A.R.; Liu, M.; Khan, M.W.; Zhai, G. Progress in brain targeting drug delivery system by nasal route. J. Control. Release 2017, 268, 364–389. [Google Scholar] [CrossRef] [PubMed]
  151. Correia, A.C.; Monteiro, A.R.; Silva, R.; Moreira, J.N.; Lobo, J.M.S.; Silva, A.C. Lipid nanoparticles strategies to modify pharmacokinetics of central nervous system targeting drugs: Crossing or circumventing the blood—Brain barrier (BBB) to manage neurological disorders. Adv. Drug Deliv. Rev. 2022, 189, 114485. [Google Scholar] [CrossRef] [PubMed]
  152. Gyimesi, G. Transporter-Mediated Drug Delivery. Molecules 2023, 28, 1151. [Google Scholar] [CrossRef]
  153. Chen, R.; Zhao, X.; Hu, K. Efflux Pump Inhibition to Enhance Brain Targeting Delivery; Elsevier Ltd.: Amsterdam, The Netherlands, 2018. [Google Scholar] [CrossRef]
  154. Botti, G.; Dalpiaz, A.; Pavan, B. Targeting systems to the brain obtained by merging prodrugs, nanoparticles, and nasal administration. Pharmaceutics 2021, 13, 1144. [Google Scholar] [CrossRef]
  155. Saxena, S.; Bhardwaj, S.; Aggarwal, A. Brain Targeted Drug Delivery System: A Review. Res. Anal. J. 2023, 6, 16–29. [Google Scholar] [CrossRef]
  156. Pardridge, W.M. Brain Delivery of Nanomedicines: Trojan Horse Liposomes for Plasmid DNA Gene Therapy of the Brain. Front. Med. Technol. 2020, 2, 602236. [Google Scholar] [CrossRef] [PubMed]
  157. Van den Broek, S.L.; Shalgunov, V.; Herth, M.M. Transport of nanomedicines across the blood-brain barrier: Challenges and opportunities for imaging and therapy. Biomater. Adv. 2022, 141, 213125. [Google Scholar] [CrossRef] [PubMed]
  158. Formica, M.L.; Real, D.A.; Picchio, M.L.; Catlin, E.; Donnelly, R.F.; Paredes, A.J. On a highway to the brain: A review on nose-to-brain drug delivery using nanoparticles. Appl. Mater. Today 2022, 29, 101631. [Google Scholar] [CrossRef]
  159. Crowe, T.P.; Hsu, W.H. Evaluation of Recent Intranasal Drug Delivery Systems to the Central Nervous System. Pharmaceutics 2022, 14, 629. [Google Scholar] [CrossRef] [PubMed]
  160. Fortuna, A.; Schindowski, K.; Sonvico, F. Editorial: Intranasal Drug Delivery: Challenges and Opportunities. Front. Pharmacol. 2022, 13, 868986. [Google Scholar] [CrossRef]
  161. Lochhead, J.J.; Kumar, N.N.; Nehra, G.; Stenslik, M.J.; Bradley, L.H.; Thorne, R.G. Intranasal Drug Delivery to the Brain; Springer: Berlin/Heidelberg, Germany, 2022; Volume 33. [Google Scholar] [CrossRef]
  162. Thakur, A.; Singh, P.; Biswal, S.S.; Kumar, N.; Jha, C.B.; Singh, G.; Kaur, C.; Wadhwa, S.; Kumar, R. Drug delivery through nose: A noninvasive technique for brain targeting. J. Rep. Pharm. Sci. 2020, 9, 168–175. [Google Scholar] [CrossRef]
  163. Patel, D.; Thakkar, H. Formulation considerations for improving intranasal delivery of CNS acting therapeutics. Ther. Deliv. 2022, 13, 371–381. [Google Scholar] [CrossRef]
  164. Kumar, H.; Mishra, G.; Sharma, A.K.; Gothwal, A.; Kesharwani, P.; Gupta, U. Intranasal Drug Delivery: A Non-Invasive Approach for the Better Delivery of Neurotherapeutics. Pharm. Nanotechnol. 2017, 5, 203–214. [Google Scholar] [CrossRef]
  165. Bahadur, S.; Jha, M.K. Emerging nanoformulations for drug targeting to brain through intranasal delivery: A comprehensive review. J. Drug Deliv. Sci. Technol. 2022, 78, 103932. [Google Scholar] [CrossRef]
  166. Marcello, E.; Chiono, V. Biomaterials-Enhanced Intranasal Delivery of Drugs as a Direct Route for Brain Targeting. Int. J. Mol. Sci. 2023, 24, 3390. [Google Scholar] [CrossRef]
  167. Wang, Z.; Xiong, G.; Tsang, W.C.; Schätzlein, A.G.; Uchegbu, I.F. Nose-to-Brain Delivery. J. Pharmacol. Exp. Ther. 2019, 370, 593–601. [Google Scholar] [CrossRef] [PubMed]
  168. Govender, M.; Indermun, S.; Kumar, P.; Choonara, Y.E. Potential Targeting Sites to the Brain Through Nasal Passage. In Nasal Drug Delivery; Springer: Berlin/Heidelberg, Germany, 2023; pp. 83–99. [Google Scholar] [CrossRef]
  169. Misra, S.K.; Pathak, K. Nose-to-Brain Targeting via Nanoemulsion: Significance and Evidence. Colloids Interfaces 2023, 7, 23. [Google Scholar] [CrossRef]
  170. Rai, G.; Gauba, P.; Dang, S. Recent advances in nanotechnology for Intra-nasal drug delivery and clinical applications. J. Drug Deliv. Sci. Technol. 2023, 86, 104726. [Google Scholar] [CrossRef]
  171. Journal, A.I.; Selvaraj, K.; Gowthamarajan, K.; Venkata, V. Nose to brain transport pathways an overview: Potential of nanostructured lipid carriers in nose to brain targeting. Artif. Cells Nanomed. Biotechnol. 2018, 46, 2088–2095. [Google Scholar] [CrossRef]
  172. Erdő, F.; Bors, L.A.; Farkas, D.; Bajza, Á.; Gizurarson, S. Evaluation of intranasal delivery route of drug administration for brain targeting. Brain Res. Bull. 2018, 143, 155–170. [Google Scholar] [CrossRef] [PubMed]
  173. Borkar, S.P.; Raizaday, A. Different Strategies for Nose-to-Brain Delivery of Small Molecules. In Nasal Drug Delivery; Springer: Berlin/Heidelberg, Germany, 2023; pp. 361–379. [Google Scholar] [CrossRef]
  174. Alabsi, W.; Eedara, B.B.; Encinas-Basurto, D.; Polt, R.; Mansour, H.M. Nose-to-Brain Delivery of Therapeutic Peptides as Nasal Aerosols. Pharmaceutics 2022, 14, 1870. [Google Scholar] [CrossRef] [PubMed]
  175. Kushwaha, S.K.S.; Keshari, R.K.; Rai, A.K. Advances in nasal trans-mucosal drug delivery. J. Appl. Pharm. Sci. 2011, 1, 21–28. [Google Scholar]
  176. Hemalatha, B.; Kalpana, M.; Rekha, B.S.; Varalakshmi, A.; Padmalatha, K. An Overview on Nasal Drug Delivery System. Asian J. Pharm. Res. 2022, 12, 249–258. [Google Scholar] [CrossRef]
  177. Dhuria, S.V.; Hanson, L.R.; Frey, W.H. Intranasal delivery to the central nervous system: Mechanisms and experimental considerations. J. Pharm. Sci. 2010, 99, 1654–1673. [Google Scholar] [CrossRef]
  178. Kotha, A.K.; Ghosh, S.; Komanduri, N.; Wang, R.; Bhowmick, S.; Chougule, M.B. Approaches in barriers, modifications, route of administrations, and formulations of therapeutic agents for brain delivery. In Novel Drug Delivery Technologies: Innovative Strategies for Drug Re-Positioning; Springer: Berlin/Heidelberg, Germany, 2020; pp. 383–401. [Google Scholar] [CrossRef]
  179. Duan, X.; Mao, S. New strategies to improve the intranasal absorption of insulin. Drug Discov. Today 2010, 15, 416–427. [Google Scholar] [CrossRef]
  180. De Ponti, R.; Lardini, E. Use of chemical enhancers for nasal drug delivery. Drug Dev. Ind. Pharm. 1991, 17, 1419–1436. [Google Scholar] [CrossRef]
  181. Chavanpatil, M.D.; Vavia, P.R. The influence of absorption enhancers on nasal absorption of acyclovir. Eur. J. Pharm. Biopharm. 2004, 57, 483–487. [Google Scholar] [CrossRef] [PubMed]
  182. Davis, S.S.; Illum, L. Absorption Enhancers for Nasal Drug Delivery. Clin. Pharmacokinet. 2003, 42, 1107–1128. [Google Scholar] [CrossRef] [PubMed]
  183. Baldassi, D.; Ambike, S.; Feuerherd, M.; Cheng, C.-C.; Peeler, D.J.; Feldmann, D.P.; Porras-Gonzalez, D.L.; Wei, X.; Keller, L.-A.; Kneidinger, N.; et al. Inhibition of SARS-CoV-2 replication in the lung with siRNA/VIPER polyplexes. J. Control. Release 2022, 345, 661–674. [Google Scholar] [CrossRef] [PubMed]
  184. Naqvi, S.; Panghal, A.; Flora, S.J.S. Nanotechnology: A Promising Approach for Delivery of Neuroprotective Drugs. Front. Neurosci. 2020, 14, 494. [Google Scholar] [CrossRef] [PubMed]
  185. Nazem, A.; Mansoori, G.A. Nanotechnology Solutions for Alzheimer’s Disease: Advances in Research Tools, Diagnostic Methods and Therapeutic Agents. J. Alzheimer’s Dis. 2020, 13, 199–223. [Google Scholar] [CrossRef] [PubMed]
  186. Rabiee, N.; Ahmadi, S.; Afshari, R.; Khalaji, S.; Rabiee, M.; Bagherzadeh, M.; Fatahi, Y.; Dinarvand, R.; Tahriri, M.; Tayebi, L.; et al. Polymeric Nanoparticles for Nasal Drug Delivery to the Brain: Relevance to Alzheimer’s Disease. Adv. Ther. 2020, 4, 2000076. [Google Scholar] [CrossRef]
  187. Brambilla, D.; Droumaguet, B.L.; Nicolas, J.; Hashemi, S.H.; Wu, L.-P.; Moghimi, S.M.; Couvreur, P.; Andrieux, K. Nanotechnologies for Alzheimer’s disease: Diagnosis, therapy, and safety issues. Nanomed. Nanotechnol. Biol. Med. 2011, 7, 521–540. [Google Scholar] [CrossRef]
  188. Petschauer, J.S.; Madden, A.J.; Kirschbrown, W.P.; Song, G.; Zamboni, W.C. The effects of nanoparticle drug loading on the pharmacokinetics of anticancer agents. Nanomedicine 2015, 10, 447–463. [Google Scholar] [CrossRef]
  189. Modi, G.; Pillay, V.; Choonara, Y.E. Advances in the treatment of neurodegenerative disorders employing nanotechnology. Ann. N. Y. Acad. Sci. 2010, 1184, 154–172. [Google Scholar] [CrossRef]
  190. Raj, R.; Wairkar, S.; Sridhar, V.; Gaud, R. International Journal of Biological Macromolecules Pramipexole dihydrochloride loaded chitosan nanoparticles for nose to brain delivery: Development, characterization and in vivo anti-Parkinson activity. Int. J. Biol. Macromol. 2018, 109, 27–35. [Google Scholar] [CrossRef]
  191. Wilson, B.; Nasralla, B.; Alobaid, M.; Mukundan, K.; Leno, J. Chitosan nanoparticles to enhance nasal absorption and brain targeting of sitagliptin to treat Alzheimer’s disease. J. Drug Deliv. Sci. Technol. 2020, 61, 102176. [Google Scholar] [CrossRef]
  192. Kandil, L.S.; Farid, R.M.; Elgamal, S.S.; Hanafy, A.S. Intranasal galantamine/chitosan complex nanoparticles elicit neuroprotection potentials in rat brains via antioxidant effect. Drug Dev. Ind. Pharm. 2021, 47, 735–740. [Google Scholar] [CrossRef] [PubMed]
  193. Zhang, L.; Yang, S.; Wong, L.R.; Xie, H.; Ho, P.C. In Vitro and In Vivo Comparison of Curcumin-Encapsulated Chitosan-Coated Poly (lactic-co-glycolic acid) Nanoparticles and Curcumin/Hydroxypropyl-β-Cyclodextrin Inclusion Complexes Administered Intranasally as Therapeutic Strategies for Alzheimer ’s Disease. Mol. Pharm. 2020, 17, 4256–4269. [Google Scholar] [CrossRef] [PubMed]
  194. Pawar, D.; Mangal, S.; Goswami, R.; Jaganathan, K.S. European Journal of Pharmaceutics and Biopharmaceutics Development and characterization of surface modified PLGA nanoparticles for nasal vaccine delivery: Effect of mucoadhesive coating on antigen uptake and immune adjuvant activity. Eur. J. Pharm. Biopharm. 2013, 85, 550–559. [Google Scholar] [CrossRef] [PubMed]
  195. Lee, D.; Minko, T. Nanotherapeutics for Nose-to-Brain Drug Delivery: An Approach to Bypass the Blood Brain Barrier. Pharmaceutics 2021, 13, 2049. [Google Scholar] [CrossRef] [PubMed]
  196. Elnaggar, Y.S.R.; Etman, S.M.; Abdelmonsif, D.A.; Abdallah, O.Y. Intranasal Piperine-Loaded Chitosan Nanoparticles as Brain-Targeted Therapy in Alzheimer’s Disease: Optimization, Biological Efficacy, and Potential Toxicity. J. Pharm. Sci. 2015, 104, 3544–3556. [Google Scholar] [CrossRef]
  197. Fazil, M.; Haque, S.; Kumar, M.; Baboota, S. European Journal of Pharmaceutical Sciences Development and evaluation of rivastigmine loaded chitosan nanoparticles for brain targeting. Eur. J. Pharm. Sci. 2012, 47, 6–15. [Google Scholar] [CrossRef]
  198. Musumeci, T.; Di Benedetto, G.; Carbone, C.; Bonaccorso, A.; Amato, G.; Faro, M.J.O.; Burgaletto, C.; Puglisi, G.; Bernardini, R.; Cantarella, G. Intranasal Administration of a TRAIL Neutralizing Monoclonal Antibody Adsorbed in PLGA Nanoparticles and NLC Nanosystems: An In Vivo Study on a Mouse Model of Alzheimer’s Disease. Biomedicines 2022, 10, 985. [Google Scholar] [CrossRef]
  199. Su, Y.; Sun, B.; Gao, X.; Dong, X.; Fu, L.; Zhang, Y.; Li, Z.; Wang, Y.; Jiang, H.; Han, B. Intranasal Delivery of Targeted Nanoparticles Loaded With miR-132 to Brain for the Treatment of Neurodegenerative Diseases. Front. Pharmacol. 2020, 11, 1165. [Google Scholar] [CrossRef]
  200. Nanaki, S.G.; Spyrou, K.; Bekiari, C.; Veneti, P.; Baroud, T.N.; Karouta, N.; Grivas, I.; Papadopoulos, G.C.; Gournis, D.; Bikiaris, D.N. Hierarchical porous Carbon—PLLA and PLGA hybrid nanoparticles for intranasal delivery of galantamine for Alzheimer’s disease therapy. Pharmaceutics 2020, 12, 227. [Google Scholar] [CrossRef] [PubMed]
  201. Shamarekh, K.S.; Gad, H.A.; Soliman, M.E.; Sammour, O.A. Development and evaluation of protamine-coated PLGA nanoparticles for nose-to-brain delivery of tacrine: In-vitro and in-vivo assessment. J. Drug Deliv. Sci. Technol. 2020, 57, 101724. [Google Scholar] [CrossRef]
  202. Meng, Q.; Hua, H.; Jiang, Y.; Wang, Y.; Mu, H.; Wu, Z. 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]
  203. Araya, E.; Olmedo, I.; Bastus, N.G.; Guerrero, S.; Puntes, V.F.; Giralt, E.; Kogan, M.J. Gold nanoparticles and microwave irradiation inhibit beta-amyloid amyloidogenesis. Nanoscale Res. Lett. 2008, 3, 435–443. [Google Scholar] [CrossRef]
  204. Nazem, A.; Mansoori, G.A. Nanotechnology for Alzheimer’s disease detection and treatment. Insciences J. 2011, 1, 169–193. [Google Scholar] [CrossRef]
  205. Salem, H.F.; Kharshoum, R.M.; Abou-Taleb, H.A.; Naguib, D.M. Brain targeting of resveratrol through intranasal lipid vesicles labelled with gold nanoparticles: In vivo evaluation and bioaccumulation investigation using computed tomography and histopathological examination. J. Drug Target. 2019, 27, 1127–1134. [Google Scholar] [CrossRef]
  206. Zhang, L.; Dong, W.F.; Sun, H.B. Multifunctional superparamagnetic iron oxide nanoparticles: Design, synthesis and biomedical photonic applications. Nanoscale 2013, 5, 7664–7684. [Google Scholar] [CrossRef]
  207. Mahmoudi, M.; Quinlan-Pluck, F.; Monopoli, M.P.; Sheibani, S.; Vali, H.; Dawson, K.A.; Lynch, I. Influence of the physiochemical properties of superparamagnetic iron oxide nanoparticles on amyloid β protein fibrillation in solution. ACS Chem. Neurosci. 2013, 4, 475–485. [Google Scholar] [CrossRef]
  208. Fernandes, A.P.; Gandin, V. Selenium compounds as therapeutic agents in cancer. Biochim. Biophys. Acta—Gen. Subj. 2015, 1850, 1642–1660. [Google Scholar] [CrossRef]
  209. Akel, H.; Ismail, R.; Csóka, I. Progress and perspectives of brain-targeting lipid-based nanosystems via the nasal route in Alzheimer’s disease. Eur. J. Pharm. Biopharm. 2020, 148, 38–53. [Google Scholar] [CrossRef]
  210. Patel, S.; Chavhan, S.; Soni, H.; Babbar, A.K.; Mathur, R.; Mishra, A.K.; Sawant, K. Brain targeting of risperidone-loaded solid lipid nanoparticles by intranasal route. J. Drug Target. 2011, 19, 468–474. [Google Scholar] [CrossRef] [PubMed]
  211. Arora, D.; Bhatt, S.; Kumar, M.; Verma, R.; Taneja, Y.; Kaushal, N.; Tiwari, A.; Tiwari, V.; Alexiou, A.; Albogami, S.; et al. QbD-based rivastigmine tartrate-loaded solid lipid nanoparticles for enhanced intranasal delivery to the brain for Alzheimer’s therapeutics. Front. Aging Neurosci. 2022, 14, 960246. [Google Scholar] [CrossRef] [PubMed]
  212. Yasir, M.; Sara, U.V.S.; Chauhan, I.; Gaur, P.K.; Singh, A.P.; Puri, D. Ameeduzzafar, Solid lipid nanoparticles for nose to brain delivery of donepezil: Formulation, optimization by Box–Behnken design, in vitro and in vivo evaluation. Artif. Cells Nanomed. Biotechnol. 2018, 46, 1838–1851. [Google Scholar] [CrossRef]
  213. Yusuf, M.; Khan, M.; Khan, R.A.; Ahmed, B. Preparation, characterization, in vivo and biochemical evaluation of brain targeted Piperine solid lipid nanoparticles in an experimentally induced Alzheimer’s disease model. J. Drug Target. 2013, 21, 300–311. [Google Scholar] [CrossRef]
  214. Saini, S.; Sharma, T.; Jain, A.; Kaur, H.; Katare, O.P.; Singh, B. Systematically designed chitosan-coated solid lipid nanoparticles of ferulic acid for effective management of Alzheimer’s disease: A preclinical evidence. Colloids Surf. B Biointerfaces 2021, 205, 111838. [Google Scholar] [CrossRef]
  215. Anand, A.; Arya, M.; Kaithwas, G.; Singh, G.; Saraf, S.A. Sucrose stearate as a biosurfactant for development of rivastigmine containing nanostructured lipid carriers and assessment of its activity against dementia in C. elegans model. J. Drug Deliv. Sci. Technol. 2019, 49, 219–226. [Google Scholar] [CrossRef]
  216. 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] [PubMed]
  217. Jojo, G.M.; Kuppusamy, G.; De, A.; Narayan, V.V.S. 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]
  218. Rompicherla, S.K.L.; Arumugam, K.; Bojja, S.L.; Kumar, N.; Rao, C.M. Pharmacokinetic and pharmacodynamic evaluation of nasal liposome and nanoparticle based rivastigmine formulations in acute and chronic models of Alzheimer’s disease. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2021, 394, 1737–1755. [Google Scholar] [CrossRef]
  219. Sokolik, V.; Berchenko, O.G.; Shulga, S. Comparative Analysis of Nasal Therapy with Soluble and Liposomal Forms of Curcumin on Rats with Alzheimer’s Disease Model. J. Alzheimer’s Dis. Park. 2017, 7, 2161–0460. [Google Scholar] [CrossRef]
  220. Li, W.; Zhou, Y.; 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]
  221. Migliore, M.M.; Vyas, T.K.; Campbell, R.B.; Amiji, M.M.; Waszczak, B.L. Brain delivery of proteins by the intranasal route of administration: A comparison of cationic liposomes versus aqueous solution formulations. J. Pharm. Sci. 2010, 99, 1745–1761. [Google Scholar] [CrossRef] [PubMed]
  222. 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] [PubMed]
  223. 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] [PubMed]
  224. El-Helaly, S.N.; Elbary, A.A.; Kassem, M.A.; El-Nabarawi, M.A. Electrosteric stealth rivastigmine loaded liposomes for brain targeting: Preparation, characterization, ex vivo, bio-distribution and in vivo pharmacokinetic studies. Drug Deliv. 2017, 24, 692–700. [Google Scholar] [CrossRef]
  225. 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] [PubMed]
  226. Mourtas, S.; Lazar, A.N.; Markoutsa, E.; Duyckaerts, C.; Sophia, G. Multifunctional nanoliposomes with curcumin–lipid derivative and brain targeting functionality with potential applications for Alzheimer disease. Eur. J. Med. Chem. 2014, 80, 175–183. [Google Scholar] [CrossRef]
  227. Fonseca-santos, B. Nanotechnology-based drug delivery systems for the treatment of Alzheimer’s disease. Int. J. Nanomed. 2015, 10, 4981–5003. [Google Scholar] [CrossRef]
  228. Hanafy, A.S.; Farid, R.W.; 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 Pharmacological, toxicological and neuronal localization asse. Drug Deliv. 2016, 7544, 3111–3122. [Google Scholar] [CrossRef]
  229. Wang, X.; Chi, N.; Tang, X. Preparation of estradiol chitosan nanoparticles for improving nasal absorption and brain targeting. Eur. J. Pharm. Biopharm. 2008, 70, 735–740. [Google Scholar] [CrossRef]
  230. Ke, W.; Shao, K.; Huang, R.; Han, L.; Liu, Y.; Li, J.; Kuang, Y.; Ye, L.; Lou, J.; Jiang, C. Biomaterials Gene delivery targeted to the brain using an Angiopep-conjugated polyethyleneglycol-modified polyamidoamine dendrimer. Biomaterials 2009, 30, 6976–6985. [Google Scholar] [CrossRef] [PubMed]
  231. Pal, I.; Bhandari, R.; Bhandari, S.; Kakkar, V. Potential of solid lipid nanoparticles in brain targeting. J. Control. Release 2008, 127, 97–109. [Google Scholar] [CrossRef]
  232. Rassu, G.; Soddu, E.; Posadino, A.M.; Pintus, G.; Sarmento, B.; Giuchedi, P.; Gavini, E. Colloids and Surfaces B: Biointerfaces 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]
  233. Vaz, G.R.; Hädrich, G.; Bidone, J.; Rodrigues, J.L.; Falkembach, M.C.; Putaux, J.-L.; Hort, M.A.; Monserrat, J.M.; Varela Junior, A.S.; Teixeira, H.F.; et al. Development of Nasal Lipid Nanocarriers Containing Curcumin for Brain Targeting. J. Alzheimer’s Dis. 2017, 59, 961–974. [Google Scholar] [CrossRef] [PubMed]
  234. Gao, X.; Wu, B.; Zhang, Q.; Chen, J.; Zhu, J.; Zhang, W. Brain delivery of vasoactive intestinal peptide enhanced with the nanoparticles conjugated with wheat germ agglutinin following intranasal administration. J. Control. Release 2007, 121, 156–167. [Google Scholar] [CrossRef]
  235. Liu, Z.; Jiang, M.; Kang, T.; Miao, D.; Gu, G.; Song, Q.; Yao, L.; Hu, Q.; Tu, Y.; Chen, Z.; et al. Biomaterials Lactoferrin-modi fi ed PEG-co-PCL nanoparticles for enhanced brain delivery of NAP peptide following intranasal administration. Biomaterials 2013, 34, 3870–3881. [Google Scholar] [CrossRef]
  236. Bana, L.; Minniti, S.; Salvati, E.; Sesana, S.; Zambelli, V.; Cagnotto, A.; Orlando, A.; Cazzaniga, E.; Zwart, R.; Scheper, W.; et al. Liposomes bi-functionalized with phosphatidic acid and an ApoE-derived peptide affect A β aggregation features and cross the blood—Brain-barrier: Implications for therapy of Alzheimer disease. Nanomed. Nanotechnol. Biol. Med. 2013, 10, 1583–1590. [Google Scholar] [CrossRef]
  237. Chen, Z.; Huang, M.; Wang, X.-R.; Fu, J.; Han, M.; Shen, Y.-Q.; Xia, Z.; Gao, J.-Q. Transferrin-modified liposome promotes α-Mangostin to penetrate the blood-brain barrier. Nanomed. Nanotechnol. Biol. Med. 2015, 12, 421–430. [Google Scholar] [CrossRef]
  238. Salvati, E.; Sesana, S.; Sancini, G. Liposomes functionalized to overcome the blood—Brain barrier and to target amyloid-β peptide: The chemical design affects the permeability across an in vitro model. Int. J. Nanomed. 2013, 8, 1749–1758. [Google Scholar]
  239. Papadia, K.; Giannou, A.D.; Markoutsa, E.; Bigot, C.; Vanhoute, G.; Mourtas, S.; Van der Linded, A.; Stathopoulos, G.T.; Antimisiaris, S.G. European Journal of Pharmaceutical Sciences Multifunctional LUV liposomes decorated for BBB and amyloid targeting—B. In vivo brain targeting potential in wild-type and APP/PS1 mice. Eur. J. Pharm. Sci. 2017, 102, 180–187. [Google Scholar] [CrossRef]
  240. Qiang, F.; Shin, H.J.; Lee, B.J.; Han, H.K. Enhanced systemic exposure of fexofenadine via the intranasal administration of chitosan-coated liposome. Int. J. Pharm. 2012, 430, 161–166. [Google Scholar] [CrossRef] [PubMed]
  241. Bonferoni, M.C.; Rossi, S.; Sandri, G.; Ferrari, F.; Gavini, E.; Rassu, S.; Giunchedi, P. Nanoemulsions for ‘nose-to-brain’ drug delivery. Pharmaceutics 2019, 11, 84. [Google Scholar] [CrossRef] [PubMed]
  242. Shinde, R.L.; Jindal, A.B.; Devarajan, P.V. Microemulsions and Nanoemulsions for Targeted Drug Delivery to the Brain. Curr. Nanosci. 2011, 7, 119–133. [Google Scholar] [CrossRef]
  243. Kaur, A.; Nigam, K.; Bhatnagar, I.; Sukhpal, H.; Awasthy, S.; Shankar, S.; Tyagi, A.; Dang, S. Treatment of Alzheimer’s diseases using donepezil nanoemulsion: An intranasal approach. Drug Deliv. Transl. Res. 2020, 10, 1862–1875. [Google Scholar] [CrossRef] [PubMed]
  244. Song, Y.; Wang, X.; Wang, X.; Wang, J.; Hao, Q.; Hao, J. Osthole- Loaded Nanoemulsion Enhances Brain Target in the Treatment of Alzheimer’s Disease via Intranasal Administration. Oxidative Med. Cell. Longev. 2021, 2021, 8844455. [Google Scholar] [CrossRef] [PubMed]
  245. Vasdev, N.; Handa, M. Rosemary oil low energy nanoemulsion: Optimization, µrheology, in silico, in vitro, and ex vivo characterization. J. Biomater. Sci. Polym. Ed. 2022, 33, 1901–1923. [Google Scholar] [CrossRef] [PubMed]
  246. Kotta, S.; Aldawsari, H.M.; Badr-Eldin, S.M.; Alhakamy, N.A.; Md, S. Coconut oil-based resveratrol nanoemulsion: Optimization using response surface methodology, stability assessment and pharmacokinetic evaluation. Food Chem. 2021, 357, 129721. [Google Scholar] [CrossRef] [PubMed]
  247. 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]
  248. Wen, M.M.; Ismail, N.I.K.; Nasra, M.M.A.; El-Kamel, A.H. Repurposing ibuprofen-loaded microemulsion for the management of Alzheimer’s disease: Evidence of potential intranasal brain targeting. Drug Deliv. 2021, 28, 1188–1203. [Google Scholar] [CrossRef]
  249. Zussy, C.; John, R.; Urgin, T.; Otaegui, L.; Vigor, C.; Acar, N.; Canet, G.; Vitalis, M.; Morin, F.; Planel, E.; et al. Intranasal Administration of Nanovectorized Docosahexaenoic Acid (DHA) Improves Cognitive Function in Two Complementary Mouse Models of Alzheimer’s Disease. Antioxidants 2022, 11, 838. [Google Scholar] [CrossRef]
  250. Chen, Y.; Cheng, G.; Hu, R.; Chen, S.; Lu, W.; Gao, S.; Xia, H.; Wang, B.; Sun, C.; Nie, X.; et al. A Nasal Temperature and pH Dual-Responsive In Situ Gel Delivery System Based on Microemulsion of Huperzine A: Formulation, Evaluation, and In Vivo Pharmacokinetic Study. AAPS PharmSciTech 2019, 20, 301. [Google Scholar] [CrossRef] [PubMed]
  251. Khunt, D.; Shrivas, M.; Polaka, S.; Gondaliya, P.; Misra, M. Role of Omega-3 Fatty Acids and Butter Oil in Targeting Delivery of Donepezil Hydrochloride Microemulsion to Brain via the Intranasal Route: A Comparative Study. AAPS PharmSciTech 2020, 21, 45. [Google Scholar] [CrossRef] [PubMed]
  252. Shah, B.; Khunt, D.; Misra, M.; Padh, H. Formulation and In-vivo Pharmacokinetic Consideration of Intranasal Microemulsion and Mucoadhesive Microemulsion of Rivastigmine for Brain Targeting. Pharm. Res. 2017, 35, 8. [Google Scholar] [CrossRef] [PubMed]
  253. Pathak, R.; Prasad, R.; Misra, M. Role of mucoadhesive polymers in enhancing delivery of nimodipine microemulsion to brain via intranasal route. Acta Pharm. Sin. B 2014, 4, 151–160. [Google Scholar] [CrossRef] [PubMed]
  254. Kumar, M.; Misra, A.; Babbar, A.K.; Mishra, A.K.; Mishra, P.; Pathak, K. Intranasal nanoemulsion based brain targeting drug delivery system of risperidone. Int. J. Pharm. 2008, 358, 285–291. [Google Scholar] [CrossRef] [PubMed]
  255. Mahajan, H.S.; Mahajan, M.S.; Nerkar, P.P.; Agrawal, A. Nanoemulsion-based intranasal drug delivery system of saquinavir mesylate for brain targeting. Drug Deliv. 2014, 7544, 148–154. [Google Scholar] [CrossRef]
  256. Mizrahi, M.; Friedman-Levi, Y.; Larush, L.; Frid, K.; Binyamin, O.; Dori, D.; Fainstein, N.; Ovadia, H.; Ben-Hur, T.; Magdassi, S.; et al. Pomegranate seed oil nanoemulsions for the prevention and treatment of neurodegenerative diseases: The case of genetic CJD. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 1353–1363. [Google Scholar] [CrossRef]
  257. Wang, X.; Jiang, Y.; Wang, Y.; Huang, M. Food Chemistry Enhancing anti-inflammation activity of curcumin through O/W nanoemulsions. Food Chem. 2008, 108, 419–424. [Google Scholar] [CrossRef]
  258. Yadav, S.; Gandham, S.K.; Panicucci, R.; Amiji, M.M. Intranasal brain delivery of cationic nanoemulsion-encapsulated TNFα siRNA in prevention of experimental neuroinflammation. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 987–1002. [Google Scholar] [CrossRef]
  259. Jacob, S.; Nair, A.B.; Shah, J. Emerging role of nanosuspensions in drug delivery systems. Biomater. Res. 2020, 24, 1–16. [Google Scholar] [CrossRef]
  260. Dibaei, M.; Rouini, M.R.; Sheikholeslami, B.; Gholami, M.; Dinarvand, R. The effect of surface treatment on the brain delivery of curcumin nanosuspension: In vitro and in vivo studies. Int. J. Nanomed. 2019, 14, 5477–5490. [Google Scholar] [CrossRef] [PubMed]
  261. Bhavna; Shadab, M.; 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] [PubMed]
  262. Müller, R.; Junghanns, J.-U.A.H. Nanocrystal technology, drug delivery and clinical applications. Int. J. Nanomed. 2008, 3, 295. [Google Scholar] [CrossRef] [PubMed]
  263. Pawar, V.K.; Singh, Y.; Meher, J.G.; Gupta, S.; Chourasia, M.K. Engineered nanocrystal technology: In-vivo fate, targeting and applications in drug delivery. J. Control. Release 2014, 183, 51–66. [Google Scholar] [CrossRef] [PubMed]
  264. Wu, C.; Benyue, L.; Yi, Z.; Tingting, C.; Chuangrong, C.; Wei, D.; Qi, W.; Tongkai, C. Intranasal delivery of paeoniflorin nanocrystals for brain targeting. Asian J. Pharm. Sci. 2020, 15, 326–335. [Google Scholar] [CrossRef] [PubMed]
  265. Babylon, L.; Grewal, R.; Stahr, P.L.; Eckert, R.W.; Keck, C.M.; Eckert, G. P Hesperetin nanocrystals improve mitochondrial function in a cell model of early Alzheimer disease. Antioxidants 2021, 10, 1003. [Google Scholar] [CrossRef] [PubMed]
  266. Zhu, S.; Zhang, S.; Pang, L.; Ou, G.; Zhu, L.; Ma, J.; Li, R.; Liu, Y.; Wang, L.; Wang, L.; et al. Effects of armodafinil nanocrystal nasal hydrogel on recovery of cognitive function in sleep-deprived rats. Int. J. Pharm. 2021, 597, 120343. [Google Scholar] [CrossRef]
  267. Fomicheva, A. Signal Enhancement in Antibody Microarrays Using Quantum Dots Nanocrystals: Application to Potential Alzheimer’s Disease Biomarker Screening. Anal. Chem. 2012, 84, 6821–6827. [Google Scholar] [CrossRef]
  268. Thakur, G.; Micic, M.; Yang, Y.; Li, W.; Movia, D.; Giordani, S.; Zhang, H.; Leblanc, R.M. Conjugated Quantum Dots Inhibit the Amyloid β (1–42) Fibrillation Process. Int. J. Alzheimer’s Dis. 2011, 2011, 502386. [Google Scholar] [CrossRef]
  269. Quan, L.; Wu, J.; Lane, L.A.; Wang, J.; Lu, Q.; Gu, Z.; Wang, Y. Enhanced Detection Specificity and Sensitivity of Alzheimer’s Disease Using Amyloid-beta Targeted Quantum Dots. Bioconjugate Chem. 2016, 27, 809–814. [Google Scholar] [CrossRef]
  270. Liu, Y.; Xu, L.; Dai, W.; Dong, H.; Wen, Y.; Zhang, X. Graphene quantum dots for the inhibition of β amyloid aggregation. Nanoscale 2015, 7, 19060–19065. [Google Scholar] [CrossRef] [PubMed]
  271. Mars, A.; Hamami, M.; Bechnak, L.; Patra, D.; Raouafi, N. Curcumin-graphene quantum dots for dual mode sensing platform: Electrochemical and fluorescence detection of APOe4, responsible of Alzheimer’s disease. Anal. Chim. Acta 2018, 1036, 141–146. [Google Scholar] [CrossRef] [PubMed]
  272. Sharma, S.; Singh, N.; Nepovimova, E.; Korabecny, J.; Satnami, M.L.; Ghosh, K.K. Interaction of synthesized nitrogen enriched graphene quantum dots with novel anti-Alzheimer’s drugs: Spectroscopic insights. J. Biomol. Struct. Dyn. 2019, 38, 1822–1837. [Google Scholar] [CrossRef] [PubMed]
  273. Tang, M.; Pi, J.; Long, Y.; Huang, N.; Cheng, Y.; Zheng, H. Quantum dots-based sandwich immunoassay for sensitive detection of Alzheimer’s disease-related Aβ1–42. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 201, 82–87. [Google Scholar] [CrossRef] [PubMed]
  274. Xiao, L.; Zhao, D.; Chan, W.; Choi, M.M.F.; Li, H. Biomaterials Inhibition of beta 1–40 amy loid fibrillation with N-acetyl-L-cysteine capped quantum dots. Biomaterials 2010, 31, 91–98. [Google Scholar] [CrossRef]
  275. Xiao, S.; Zhou, D.; Luan, P.; Gu, B.; Feng, L.; Fan, S.; Liao, W.; Fang, W.; Yang, L.; Tao, E.; et al. Department of Neurology and Outpatient Department of Internal Medicine, Guangdong. Biomaterials 2016, 106, 98–110. [Google Scholar] [CrossRef] [PubMed]
  276. Jain, K.K. Nanobiotechnology-based strategies for crossing the blood—Brain barrier. Nanomedicine 2012, 7, 1225–1233. [Google Scholar] [CrossRef] [PubMed]
  277. Al-azzawi, S.; Masheta, D.; Guildford, A.L.; Phillips, G. 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]
  278. Al-azzawi, S.K. Improving Flurbiprofen Brain—Permeability and Targeting in Alzheimer’s Disease by Using a Novel Dendronised ApoE—Derived Peptide Carrier System. Ph.D. Thesis, University of Brighton, Brighton, UK, 2017. [Google Scholar]
  279. Agrawal, M.; Saraf, S.; Safar, S.; Dubey, S.K.; Puri, A.; Gupta, U.; Kesharwani, P.; Ravichandiran, V.; Kumar, P.; Naidu, V.G.M.; et al. Stimuli-responsive in situ gelling system for nose-to-brain drug delivery. J. Control. Release 2020, 327, 235–265. [Google Scholar] [CrossRef]
  280. Patil, R.P.; Pawara, D.D.; Gudewar, C.S.; Tekade, A.R. Nanostructured cubosomes in an in situ nasal gel system: An alternative approach for the controlled delivery of donepezil HCl to brain. J. Liposome Res. 2019, 29, 264–273. [Google Scholar] [CrossRef]
  281. Cunha, S.; Swedroska, M.; Bellahnid, Y.; Xu, Z.; Sousa Lobo, J.M.; Forbes, B.; Silva, A.C. Thermosensitive in situ hydrogels of rivastigmine-loaded lipid-based nanosystems for nose-to-brain delivery: Characterisation, biocompatibility, and drug deposition studies. Int. J. Pharm. 2022, 620, 121720. [Google Scholar] [CrossRef] [PubMed]
  282. Picone, P.; Sabatino, M.A.; Ditta, L.A.; Amato, A.; Biagio, P.L.S.; Mule’, F.; Giacomazza, D.; Dispenze, C.; Carlo, M.D. Nose-to-brain delivery of insulin enhanced by a nanogel carrier. J. Control. Release 2017, 270, 23–36. [Google Scholar] [CrossRef] [PubMed]
  283. Picone, P.; Ditta, L.A.; Sabatino, M.A.; Militello, V.; Biagio, P.L.S.; Giacinto, M.L.D.; Cristaldi, L.; Nuzzo, D.; Dispenza, C.; Giacomazza, D.; et al. Biomaterials Ionizing radiation-engineered nanogels as insulin nanocarriers for the development of a new strategy for the treatment of Alzheimer’s disease. Biomaterials 2016, 80, 179–194. [Google Scholar] [CrossRef] [PubMed]
  284. Kalaiarasi, S.; Arjun, P.; Nandhagopal, S.; Brijitta, J.; Iniyan, A.M.; Vincent, S.G.P.; Kannan, R.R. ScienceDirect Development of biocompatible nanogel for sustained drug release by overcoming the blood brain barrier in zebrafish model. J. Appl. Biomed. 2016, 14, 157–169. [Google Scholar] [CrossRef]
  285. Azadi, A.; Hamidi, M.; Khoshayand, M.; Amini, M. Preparation and optimization of surface-treated methotrexate-loaded nanogels intended for brain delivery. Carbohydr. Polym. 2012, 90, 462–471. [Google Scholar] [CrossRef] [PubMed]
  286. Boridy, S.; Takahashi, H.; Akiyoshi, K.; Maysinger, D. Biomaterials The binding of pullulan modified cholesteryl nanogels to A b oligomers and their suppression of cytotoxicity. Biomaterials 2009, 30, 5583–5591. [Google Scholar] [CrossRef] [PubMed]
  287. Vinogradov, S.V.; Batrakova, E.V.; Kabanov, A.V. Nanogels for Oligonucleotide Delivery to the Brain. Bioconjugate Chem. 2003, 15, 50–60. [Google Scholar] [CrossRef] [PubMed]
  288. Maurer-jones, M.A.; Gunsolus, I.L.; Murphy, C.J.; Haynes, C.L. Toxicity of Engineered Nanoparticles in the Environment. Anal. Chem. 2013, 85, 3036–3049. [Google Scholar] [CrossRef]
  289. Buzea, C.; Pacheco, I.I.; Robbie, K. Nanomaterials and nanoparticles: Sources and toxicity Nanomaterials and nanoparticles: Sources and toxicity. Biointerphases 2007, 2, MR17–MR71. [Google Scholar] [CrossRef]
  290. Auffan, M.; Rose, J.; Wiesner, M.R.; Bottero, J. C Chemical stability of metallic nanoparticles: A parameter controlling their potential cellular toxicity in vitro. Environ. Pollut. 2009, 157, 1127–1133. [Google Scholar] [CrossRef]
  291. Singh, N.; Jenkins, G.J.S.; Asadi, R.; Doak, S.H. Potential toxicity of superparamagnetic iron oxide nanoparticles (SPION). Nano Rev. 2010, 1, 5358. [Google Scholar] [CrossRef]
  292. Murphy, C.J.; Gole, A.M.; Stone, J.W.; Sisco, P.N.; Alkilany, A.M.; Goldsmith, E.C.; Baxter, S.C. Gold Nanoparticles in Biology: Beyond Toxicity to Cellular Imaging. Accounts Chem. Res. 2008, 41, 1721–1730. [Google Scholar] [CrossRef] [PubMed]
  293. Hussain, S.M. In vitro toxicity of nanoparticles in BRL 3A rat liver cells. Toxicol. Vitr. 2005, 19, 975–983. [Google Scholar] [CrossRef]
  294. Khlebtsov, N.; Dykman, L. Biodistribution and toxicity of engineered gold nanoparticles: A review of in vitro and in vivo studies. Chem. Soc. Rev. 2011, 40, 1647–1671. [Google Scholar] [CrossRef] [PubMed]
  295. Taylor, P.; Jeng, H.A.; Swanson, J. Toxicity of Metal Oxide Nanoparticles in Mammalian Cells Toxicity of Metal Oxide Nanoparticles in Mammalian Cells. J. Environ. Sci. Health Part A 2006, 41, 2699–2711. [Google Scholar] [CrossRef]
  296. Badawy, A.M.E.L.; Silva, R.G.; Morris, B.; Scheckel, K.G.; Suidan, M.T. Surface Charge-Dependent Toxicity of Silver Nanoparticles. Environ. Sci. Technol. 2010, 45, 283–287. [Google Scholar] [CrossRef] [PubMed]
  297. Asharani, P.V.; Gong, Z.; Valiyaveettil, S. Comparison of the toxicity of silver, gold and platinum nanoparticles in developing zebrafish embryos. Nanotoxicology 2011, 5, 43–54. [Google Scholar] [CrossRef]
  298. Gaur, N.; Sharma, N.; Dahiya, A.; Yadav, P.; Ojha, H. Toxicity and Regulatory Concerns for Nanoformulations in Medicine. In The ELSI Handbook of Nanotechnology: Risk, Safety, ELSI and Commercialization; ELSI: Ōta, Tokyo, 2020; pp. 333–357. [Google Scholar]
  299. Hejmady, S.; Singhvi, G.; Saha, R.N.; Dubey, S.K. Regulatory aspects in process development and scale-up of nanopharmaceuticals. Ther. Deliv. 2020, 11, 341–343. [Google Scholar] [CrossRef]
  300. Ehmann, F.; Sakai-Kato, K.; Duncan, R.; Pérez de la Ossa, D.H.; Pita, R.; Vidal, J.M.; Kohli, A.; Tothfalusi, L.; Sanh, A.; Tinton, S.; et al. Next-generation nanomedicines and nanosimilars: EU regulators’ initiatives relating to the development and evaluation of nanomedicines. Nanomedicine 2013, 8, 849–856. [Google Scholar] [CrossRef]
  301. Tinkle, S.; McNeil, S.; Mühlebach, S.; Bawa, R.; Borchard, G.; Barenholz, Y.; Tamarkin, L.; Desai, N. Nanomedicines: Addressing the scientific and regulatory gap. Ann. N. Y. Acad. Sci. 2014, 1313, 35–56. [Google Scholar] [CrossRef]
  302. Desai, N. Challenges in development of nanoparticle-based therapeutics. AAPS J. 2012, 14, 282–295. [Google Scholar] [CrossRef] [PubMed]
  303. Trows, S.; Wuchner, K.; Spycher, R.; Steckel, H. Analytical Challenges and Regulatory Requirements for Nasal Drug Products in Europe and the U.S. Pharmaceutics 2014, 6, 195–219. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The graphical representation delineates the primary hallmarks of Alzheimer’s disease, elucidates the mechanisms underlying current treatment strategies, and outlines various intranasal treatment approaches for Alzheimer’s disease management based on nanocarriers.
Figure 1. The graphical representation delineates the primary hallmarks of Alzheimer’s disease, elucidates the mechanisms underlying current treatment strategies, and outlines various intranasal treatment approaches for Alzheimer’s disease management based on nanocarriers.
Pharmaceutics 16 00058 g001
Figure 2. Visual representation illustrating the trajectory of the delivery system after transport through distinct intranasal pathways, namely: olfactory pathway, trigeminal pathway, and respiratory pathway.
Figure 2. Visual representation illustrating the trajectory of the delivery system after transport through distinct intranasal pathways, namely: olfactory pathway, trigeminal pathway, and respiratory pathway.
Pharmaceutics 16 00058 g002
Table 1. List of monoclonal antibody-based immunotherapies explored for targeting different hallmarks of AD (e.g., Aβ and tau proteins, etc.).
Table 1. List of monoclonal antibody-based immunotherapies explored for targeting different hallmarks of AD (e.g., Aβ and tau proteins, etc.).
NameStatusOutcomeRef
DonanemabPhase 3Donanemab showed maximum affinity towards Aβplaques, resulting in a deceleration of disease progression. Further, PET images revealed the absence of Aβ plaques in patients after12 months of treatment.[58]
BapineuzumabFailed in Phase 3Had an anti-Alzheimer’s effect by targeting tau phosphorylation, thereby decreasing the tau concentration in CSF. However, bapineuzumab failed to show clinical efficacy and its clinical use was associated with a high risk of ARIA and TEAE.[59]
SolanezumabTerminatedSolanezumab acts by identifying and targeting soluble monomer Aβ except for fibrillary Aβ. However, the trial was terminated due to negligible benefits to mild AD patients and not meeting clinical endpoints.[60]
CrenezumabCompletedThis antibody was well tolerated with no prominent side effects even when increasing the dosage. However, no commercial translation occurred as it failed to show clinical efficacy.[61]
PonezumabPhase 2Treatment with ponezumab led to increased Aβ level in plasma. On completion of treatment, no alterations in CSF biomarkers, Aβ burden, and cognition were reported, which could be due to its low penetrability.[62]
SemorinemabPhase 2Semorinemab had a well-tolerated safety profile. The 73-week treatment did not reduce disease progression and no clinical outcome was reported.[63]
GosuranemabPhase 2Considerable concentration of gosuranemab in serum and CSF was noted. Further, ~98% of unbound tau was reduced in CSF, yet no benefits were observed in a population at risk for PSP.[64]
TilavonemabPhase 2Tilavonemab showed no effect on disease progression. Intriguingly, a reduced level of free tau in CSF (38.0–46.3%) was reported, which reached a plateau when the dosage was increased.[65]
Table 2. Overview of various polymers that have been investigated for delivering therapeutics targeting Alzheimer’s disease (AD).
Table 2. Overview of various polymers that have been investigated for delivering therapeutics targeting Alzheimer’s disease (AD).
DrugPolymerTargeting Route of AdministrationResultsRef
EstradiolPolylactide-co-glycolide (PLGA)Tween 80 (mimics LDL particles by adsorbing apolipoprotein and achieves targeting via LDL receptors)Oral Route
  • Higher brain uptake (1.969 ± 0.197 ng/g) was seen in coated NPs compared to that of uncoated PLGA nanoparticles (1.105 ± 0.136 ng/g)
  • Enhanced drug fraction reached brain following oral administration.
[94]
DonepezilPLGA-b-PEG-NA
  • Donepezil nanoparticles demonstrated destabilization action against Aβ fibril.
  • Improved transport across in vitro BBB model as compared to free drug
[95]
DonepezilPolylactide-co-glycolide (PLGA)Tween-80 (internalize via LDL receptors)Intravenous
  • Biphasic release pattern with sustained release (87.42 ± 0.06%) for up to 25 days
  • Higher Cmax in brain homogenate (121.68 ± 13.23 ng/mL) as compared to that of drug solution (6.66 ± 1.13 ng/mL)
[96]
RivastigmineChitosanTween-80 (internalize through LDL receptors)Intravenous
  • Chitosan NPs demonstrated high drug loading (11.51 ± 0.32%) with particle size of 47± 4 nm
  • Biphasic release pattern with sustained release (97.25 ± 0.83%) for up to 12 h.
  • Reduced accumulation in liver, spleen, and heart.
[97]
Table 3. List of various lipid carriers, including liposomes, solid lipid nanoparticles (SLNs), nanostructured lipid carriers (NLCs), and nanoemulsions, that have been investigated for delivering therapeutics targeting Alzheimer’s disease (AD).
Table 3. List of various lipid carriers, including liposomes, solid lipid nanoparticles (SLNs), nanostructured lipid carriers (NLCs), and nanoemulsions, that have been investigated for delivering therapeutics targeting Alzheimer’s disease (AD).
DrugCarrierTargetRouteDescriptionRef
Curcumin/Ginsenoside Rb1LiposomeMannosei.v.
  • Demonstrated high encapsulation efficiency for mannose–curcumin (94.23 ± 2.886) % and Rb1 (90.56 ± 1.307) % with particle size of approx. 100 nm
  • Dual-loaded liposomes exhibited increased cell uptake and accumulation in N2a cells
  • In vivo studies in APP/PS-1 mice showed reduced oxidative stress and inflammation
[102]
miR-101 and CurcuminLiposome-NA
  • miR-101 liposomes demonstrated lowering of Aβ for up to 3 h. Prolong effect was seen for 12 h when using a combination approach
  • Curcumin demonstrated delayed and indirect effects on mRNAAPP transcription
  • Meanwhile, miR-101 shows a direct and rapid effect on transcription
  • Further prepared dual liposomes demonstrated an anti-inflammatory effect
[103]
Caffeic AcidLiposomeTransferrinNA
  • The Tf-CA-liposomes had size of 139 ± 9 nm with PdI of 0.20 ± 0.03 and %EE of 23 ± 4%
  • The modification with Tf was confirmed with ATR-FTIR
  • The sustained release was observed with 11 ± 3% at 24 h
  • ThT fluorescence showed disaggregation capacity of Tf-CA liposome against Aβ42 peptide in which 13% reduction in fibril was observed after 1 h of incubation
[104]
Memantine HCl and TramiprosateSLN-oral
  • Tramiprosate demonstrated higher inhibition (16.56%) in ThT studies as compared to memantine HCl (3.22%)
  • PK studies showed delayed clearance (>4 h) of SLNs as compared to drug solutions (1 h). Also, SLNs showed higher conc. in the brain (177.959 ± 18.366 and 30.294± 2.012 µg/mL) as compared to a solution with a lower concentration in other organs, e.g., liver and kidney
  • PD and behavioral studies indicated a neuroprotective role of SLNs
[105]
ErythropoietinSLN-i.p.
  • EPO-SLN had an optimum particle size (219.9 ± 15.6 nm), PDI (0.18 ± 0.03), and drug loading (41.4 ± 3.6 IU/mg)
  • In MWM test, EPO-SLNs demonstrated improvement in spatial and learning memory.
  • Histopathological examination showed SLNs’ potential ability to hinder Aβ effects
  • Further reduction in lipid peroxidation was observed in EPO-SLN group
[106]
BerberineNLC-oral
  • Berb-NLCs were optimized by 32 full factorial model in which final batch had size of 186 nm and 88% EE
  • The Berb-NLCs exhibited sustained release (86%) for up to 24 h
  • Pharmacodynamics studies involving behavioural evaluation showed improved cognition as compared to that when using pure berberine
[107]
ThymoquinoneNE-oral
  • Exhibited decrease Aβ40 and Aβ42 levels in HFCD-induced rats
  • Attenuation in IDE and LRP1 levels was observed which could lead to Aβ degradation
[108]
Table 4. A list of research on metal-nanoparticle-based delivery for the treatment of Alzheimer’s disease.
Table 4. A list of research on metal-nanoparticle-based delivery for the treatment of Alzheimer’s disease.
DrugCarrierTargetRoute of AdministrationDescriptionRef
-Myco-fabricated ZnO nanoparticles-i.p.
  • Myco-fabricated ZnO-NPs exhibited substantial anti-inflammatory and anti-acetylcholinesterase properties
  • A therapeutic dose of 5 mg/kg improves learning and memory activity
[141]
PEG-MIL-101 (MOF)AuNPs
  • The PEG-MIL-101 conjugated anionic AuNPs exhibited uniform binding with Aβ monomers and Aβ42 fibrils
  • Developed PEG-MIL 101-AuNPs demonstrated a marked decline in fibrillation by disrupting Aβ42 fibrils, thereby decreasing the aggregation
[142]
-Cadmium sulfide and Iron oxide nanoparticlesProtein cappedNA
  • The PC-CdS (≤20 nm) and Fe3O4 NPs (~40–50 nm) had a nanometric size
  • Concentration-dependent and time-dependent tau inhibitory action was exhibited by protein-capped CdS (63%) and Fe3O4 (49%) NPs
  • Upon treatment with NPs, a significant decrease in fibrillary aggregation was observed
[143]
Rhein and PolydopamineFe–Rh/Pda NPs(KLVFFAED)/K8 peptidei.v.
  • The 7T MRI images showed efficient transit of NPs across BBB
  • Also, SDA-PAGE analysis following treatment revealed considerable Aβ42 targeting ability of developed NPs, which was further verified by in vivo studies in APP/PS1 mice
  • Prepared NPs remarkably improved brain bioavailability (~11.2-fold) of rhein as compared to that of rhein solution
  • Improved antioxidant and anti-Aβ effects were reported
[144]
Ruthenium dioxideBorneoli.v.
  • RuO2-Bor showed concentration-dependent enzymatic activity including CAT, SOD, and POD
  • Significant decrease in ROS level was observed indicating ROS scavenging action
  • RuO2-Bor NPs exhibited inhibitory action on Aβ42 aggregation (~18.8%), maintained mitochondrial homeostasis, and restored cognition function in Aβ42 mice
[145]
Table 5. Overview of research on nanoparticle-aided intranasal delivery of anti-Alzheimer’s drugs.
Table 5. Overview of research on nanoparticle-aided intranasal delivery of anti-Alzheimer’s drugs.
DrugNanoparticleTargeting AgentMethod of PreparationPharmacological DataRef
Tacrinepoly (n-butyl cyanoacrylate)polysorbate 80Emulsion polymerization technique
  • In comparison to uncoated nanoparticles and free tacrine, a substantially increased tacrine concentration (170 ng/mL) was observed in the brain upon coating poly(n-butylcyanoacrylate) nanoparticles with 1% polysorbate 80.
[227]
GalantamineHydrobromide Chitosan complex NP
(GH–chitosan NP)
-Ionic interaction method
  • Prolonged release was obtained (58.07% ± 6.67 after 72 h) with delayed mucociliary clearance
  • GH-chitosan NPs showed improved cholinergic activity with reduced AchE levels
  • No significant toxicity to the brain was observed
[228]
EstradiolChitosan NP Ionic interaction method
  • NPs loaded with estradiol showed significantly lower concentration in plasma i.n. (32.7+/− 10.1 ng mL−1; t(max) 28 +/− 4.5 min) as compared to i.v. (151.4 +/− 28.2 ng mL−1)
  • Higher concentration(76.4 +/− 14.0 ng mL−1 and t(max) 28 +/− 17.9 min) in CSF were observed for i.n. delivery as compared to those with i.v. delivery
[229]
-Gene (DNA)Polyamidoamine dendrimers-Polyethlene glycol (PAMAM-PEG-) NPAngiopepFirst, PEG PAMAM modification of angiopep was performed followed by complexation with DNA
  • Higher efficiency to penetrate and accumulate in the brain was observed with angiopep-modified NPs as compared to non-modified NPs with higher gene expression
[230]
DoxorubicinStealth(PEG2000) and non-stealth SLN High-pressure homogenization
  • An increased accumulation of doxorubicin was observed in the brain upon increasing the level of stealthing agent PEG2000
  • Amount of doxorubicin in the brain after 30 min was found to be 27.5 ng/g in case of nonstealth SLNs while it was 242.0 ng/g for stealth SLNs loaded with 0.45% PEG; this pattern persisted for 2 h
[231]
RVG-9R -BACE1 siRNA AChitosan-coated and uncoated SLN-High-pressure homogenization
  • For siRNA, a 15 min lag time was reported whereas it took 30 min using NPs coated with chitosan
[232]
CurcuminLipid NP-Hot solvent diffusion method
  • Curcumin lipid NPs showed sustained release upto 72 h
  • The DPPH assay demonstrated 95% scavenging activity
  • It also showed enhanced permeation as compared to the free curcumin
  • Cytotoxicity studies demonstrated no toxicity with GI50 >80g/mL
[233]
Vasoactive intestinal peptide (VIP)PEG-PLA NPWheat germ agglutininDouble-emulsion solvent evaporation
  • AUC of WGA-VIP NP depicted a more than five-fold increase in the brain uptake upon i.n. administration than plain VIP solution
  • Improved brain delivery (30–50%) was observed for targeted NPs
[234]
Neuroprotective peptidePEG-co-PCL NPLactoferrinEmulsion solvent evaporation
  • Enhanced cellular accumulation was observed for lactoferrin-modified NPs as compared to unmodified NPs
  • The AUC of Coumarin-6-incorporated lactoferrinNPs was 1.56 fold higher in olfactory tract than the Coumarin-6-incorporated unmodified NPs
[235]
Table 6. List of different liposome-based intranasal drug delivery systems explored for Alzheimer’s disease treatment.
Table 6. List of different liposome-based intranasal drug delivery systems explored for Alzheimer’s disease treatment.
Liposome FormulationProblem to EncounterPharmacological DataRef
Bifunctionalized liposome mApoE-PA-LIPEffective targeting of AβThe mApoE-PA-LIP showed temporal and dose-dependent inhibition of Aβ42 aggregates, while destabilization of preformed aggregates was found to be time- and lipid-dose-dependent. Also, five-fold increased radioactivity of brain/blood ratio was seen for mApoE-PA-LIP compared to PA-LIP. [236]
Transferrin-modified alpha-M liposomesPoor penetrationThe alpha-M demonstrated an entrapment efficiency greater than 88% with improved bioavailability[237]
Fluorescent liposomes functionalized with Antibody R17217Effective binding to AβFunctionalization improved cellular uptake and permeation. The functionalized liposomes also demonstrated higher EP (7.24 ± 0.39 ×10−6 cm/min) as compared to that of biotin/streptavidin-RI-A-LIP (4.97 ± 0.51 × 10−6 cm/min).[238]
Multifunctionalized liposomes attached with two BBB-specific ligands and curcumin–lipid ligandTo locate and target formulationIn vivo study in mice demonstrated efficacy of liposomes to traverse across BBB. Addition of TREG–lipid curcumin derivative in liposome did not influence the functionality of ligands [239]
Liposome coated with chitosan and encapsulated with fexofenadineEffective brain targetingIncreased stability and retention time. Chitosan-coated liposomes showed enhanced bioavailability (34.7 ± 6.3%) as compared to non-liposomes (25.0 ± 8.0%) and uncoated liposomes (24.5 ± 7.5%). Sustained release was obtained for a period of 12 h[240]
Table 7. A list of various research studies that explored the potential of microemulsion and nanoemulsion delivery systems for Alzheimer’s disease treatment.
Table 7. A list of various research studies that explored the potential of microemulsion and nanoemulsion delivery systems for Alzheimer’s disease treatment.
Drug and DDSPharmacological EvidenceRef
Risperidone-loaded chitosan-based nanoemulsionThe mucoadhesive nanoemulsion was most effective with higher drug targeting efficiency (476 ± 2.14%) and rapid transport as compared to the drug solution[254]
Saquinavir mesylate-loaded nanoemulsionA higher concentration of drug (7290.46± 143.15 ng/g) was found at a faster rate with the NE with no toxicity and higher targeting efficiency (2919.261 ± 5.68%)[255]
Pomegranate seed oil (PSO) nanoemulsionPSO contains phytoconstituents such as polyunsaturated fatty acids and punicic acid, which reduced lipid oxidation and loss of neuronal functionality, suggesting the formulation to be neuro-protective and safe[256]
Curcumin-based o/w nanoemulsionCurcumin has low solubility and poor bioavailability. To improve its bioavailability, a curcumin-loaded NE was formulated. The prepared NE had a droplet size in the range of 618.6 nm to 79.5 nm. Anti-inflammatory action was shown using mouse ear inflammation model induced by TPA. The inhibition percentages observed were43% (in case of 618.6 nm droplets) and 85% (79.5 nm droplets), respectively.[257]
anti-TNFα siRNA-encapsulated flaxseed nanoemulsionSiRNA-loaded nanoemulsion showed 70 ± 10% encapsulation efficiency. Higher cellular uptake was observed at 15 min end point (10-fold greater) and after 2.5 h (25-fold greater), respectively. Nanoemulsion loaded with SiRNA showed improved brain targeting (two-fold greater) than SiRNA solution at the end point of 6 hr. Nanoemulsion-based delivery was found to be effective in gene knockdown and preventing neuroinflammation[258]
Table 8. A list of various studies that investigated quantum dots as a suitable carrier for delivery of different anti-AD therapeutics via intranasal route.
Table 8. A list of various studies that investigated quantum dots as a suitable carrier for delivery of different anti-AD therapeutics via intranasal route.
Drug and CarrierInvestigationResultsRef
PEG-BTA quantum dotsSpecificity and sensitivity of disease detectionEffective binding to amyloid beta peptide[269]
Graphene QDsInhibitory effect on AβSuppressed formation of fibrils. The inhibitory effect increased when surface negative charge decreased[270]
Curcumin–graphene QD coated with Indium-TO electrodeFor detection of ApoE4Reproducibility, repeatability, and high efficiency of curcumin platform for sensing even in a complex matrix[271]
High-fluorescence NGQDsTo sense enzymatic action and efficacyDecreased activity of AChE[272]
Biotinylated N-Ab and streptavidinquantum dotsTo detect AβSuccessful for detecting Aβ in CSF[273]
N-acetyl-L-cysteine-capped quantum dotsFor inhibition of amyloid fibrillationInhibitory effect with an AUC that was 100 times increased [274]
Grapheme quantum dots (GQDs) conjugated with peptide glycine–proline–glutamateNeuroprotective effectInhibition of fibril with enhanced memory and reduced inflammation[275]
Table 9. Summarizes various attempts by researchers to improve brain targeting using an insitu gelling system.
Table 9. Summarizes various attempts by researchers to improve brain targeting using an insitu gelling system.
DrugPharmacological DataRef
Poly (N-vinyl pyrrolidone) functionalized insulin nanogelReceptor binding with protection from degradation and effective transport[282]
E-beam-irradiation-based nanogel of poly(N-vinyl pyrrolidone) attached to insulinIntranasal delivery was enhanced based on activated level of AKT with increased insulin delivery[283]
Donepezil nanogel functionalized with Poly(N-isopropylacrylamide) (PNIPAM)Biocompatible with sustained release pattern and enhanced entrapment efficiency of 87.5%[284]
Methotrexate nanogels coated with polysorbate 80Effective brain targeting was achieved by coating with polysorbate 80[285]
Cholesterol-modified pullulan (CHP)- loaded hydrogel nanoparticlesInteracted with oligomeric Aβ and reduced its toxicity[286]
Oligonucleotide-based nanogelLess degradation with 15-fold enhanced biodistribution and two times less accumulation in the liver as compared to naked ODN[287]
Table 10. Lists various toxicity studies conducted using diverse types of nanomaterials and their pharmacological inferences.
Table 10. Lists various toxicity studies conducted using diverse types of nanomaterials and their pharmacological inferences.
NanomaterialPharmacological DataRef
Surface-modified gold NPs of various sizesThe concentration of gold atoms up to ~100µM does not cause any toxicity to leukemia cells. Cell viability studies demonstrated no cytotoxicity[294]
Engineered gold NPsNPs with a diameter of1–2 nm showed toxicity due to irreversible binding. No toxicity was observed in the case of NPs of the range of3–100 nm[142]
Metal oxide NPs (TiO2, ZnO, FeSO4, Al2O3, and CrO) with a size range of 30–45 nmFeSO4, Al2O3, and TiO2(concentration > 200 µg/mL) demonstrated no toxicity and at high doses, they showed LDH leakage. ZnO with a concentration range of 50–100µg/mL reduced mitochondrial function [295]
Silver NPsNPs showed toxicity via oxidative stress and a concentration of 5–50µg/mL reduced mitochondrial function along with enhanced LDH leakage[293]
Silver NPs with surface chargesAgNPs exhibited toxicity depending upon their surface charge [296]
Gold, silver, and platinum NPsExhibited toxicity via accumulation. Out of all three, the silver NPs were the most toxic whereas gold NPs were non-toxic[297]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dighe, S.; Jog, S.; Momin, M.; Sawarkar, S.; Omri, A. Intranasal Drug Delivery by Nanotechnology: Advances in and Challenges for Alzheimer’s Disease Management. Pharmaceutics 2024, 16, 58. https://doi.org/10.3390/pharmaceutics16010058

AMA Style

Dighe S, Jog S, Momin M, Sawarkar S, Omri A. Intranasal Drug Delivery by Nanotechnology: Advances in and Challenges for Alzheimer’s Disease Management. Pharmaceutics. 2024; 16(1):58. https://doi.org/10.3390/pharmaceutics16010058

Chicago/Turabian Style

Dighe, Sayali, Sunil Jog, Munira Momin, Sujata Sawarkar, and Abdelwahab Omri. 2024. "Intranasal Drug Delivery by Nanotechnology: Advances in and Challenges for Alzheimer’s Disease Management" Pharmaceutics 16, no. 1: 58. https://doi.org/10.3390/pharmaceutics16010058

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

Article Metrics

Back to TopTop