Next Article in Journal
Enhancing/Improving Forming Limit Curve and Fracture Height Predictions in the Single-Point Incremental Forming of Al1050 Sheet Material
Next Article in Special Issue
Gas-Phase Fabrication and Photocatalytic Activity of TiO2 and TiO2–CuO Nanoparticulate Thin Films
Previous Article in Journal
Graphene-Based Composites for Thermoelectric Applications at Room Temperature
Previous Article in Special Issue
Influence of Potassium-Based Activation on Adsorptive Properties of Carbon Spheres Modified with Iron(III) Citrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Adverse Effects of Non-Metallic Nanoparticles in the Central Nervous System

by
Katarzyna Sikorska
1,*,
Krzysztof Sawicki
2,
Magdalena Czajka
2,
Lucyna Kapka-Skrzypczak
2,3,
Marcin Kruszewski
1,2 and
Kamil Brzóska
1
1
Centre for Radiobiology and Biological Dosimetry, Institute of Nuclear Chemistry and Technology, Dorodna 16, 03-195 Warsaw, Poland
2
Department of Molecular Biology and Translational Research, Institute of Rural Health, Jaczewskiego 2, 20-090 Lublin, Poland
3
World Institute for Family Health, Calisia University, Nowy Swiat 4, 62-800 Kalisz, Poland
*
Author to whom correspondence should be addressed.
Materials 2023, 16(23), 7264; https://doi.org/10.3390/ma16237264
Submission received: 25 October 2023 / Revised: 17 November 2023 / Accepted: 17 November 2023 / Published: 21 November 2023
(This article belongs to the Special Issue Functional Nanoparticle Materials: From Synthesis to Application)

Abstract

:
The interest in nanoparticles (NPs) and their effects on living organisms has been continuously growing in the last decades. A special interest is focused on the effects of NPs on the central nervous system (CNS), which seems to be the most vulnerable to their adverse effects. Non-metallic NPs seem to be less toxic than metallic ones; thus, the application of non-metallic NPs in medicine and industry is growing very fast. Hence, a closer look at the impact of non-metallic NPs on neural tissue is necessary, especially in the context of the increasing prevalence of neurodegenerative diseases. In this review, we summarize the current knowledge of the in vitro and in vivo neurotoxicity of non-metallic NPs, as well as the mechanisms associated with negative or positive effects of non-metallic NPs on the CNS.

1. Introduction

The unique features of nanoparticles (NPs), e.g., size and strong absorption properties, make them more reactive in the biological environments, due to their exceptional chemical properties and ability to get different sites of organisms, as compared to a bulk material with the same chemical composition. To exploit these new capabilities, NPs were introduced to everyday life as industry and medicine products, e.g., biosensors, biomaterials, tissue engineering systems, drugs and drug-delivery systems for diagnosis and treatment of many diseases, including that of the central nervous system (CNS) [1,2,3,4,5,6,7].
The CNS can be reached by NPs through three different pathways. The first and the most likely is through systemic blood circulation, as NPs can cross blood–brain and blood–spinal cord barriers. The second path is a nose-to-brain route along nerve bundles that cross the cribriform plate to the olfactory bulb. Finally, NPs can diffuse through the nasal cavity mucosa to reach the branches of the trigeminal nerve in the olfactory and respiratory regions and then reach the brain stem via axonal transport [8,9,10].
The presence of nanomaterials in the CNS may have unexpected consequences, including acute or chronic neurological complications [2,8]. Though numerous studies clearly indicate in vitro and in vivo oxidative stress-associated adverse effects of metallic NPs in the CNS (for recent review, see Sawicki et al. [11]), the toxicity of non-metallic NPs (nmNPs) in the CNS remains obscure. Here, we review the in vitro and in vivo studies regarding the neurotoxicity of nmNPs and summarize the current state of knowledge in this area (Figure 1).

2. Non-Metallic NPs

Pure carbon NPs constitute the largest and most important group of nmNPs, which include nanotubes, nanohorns, nano-onions, nanodiamonds, fullerenes and graphene, and exhibit large diversity in structure, morphology, physical properties and chemical reactivity [1,12,13,14,15]. Another carbon-containing, large group of nmNPs are dendrimers. These NPs have a highly branched three-dimensional structure, consisting of an initial core, several internal layers, repetitive units and several terminal active surface groups. The presence of a hydrophobic core and multiple surface groups makes them a good candidate for high-load drug carriers [1,16,17]. Other nmNPs discussed later in this review include polymer NPs, silica NPs, apatite, and hydroxyapatite NPs, also widely described in scientific literature in connection with CNS.
These and other types of NPs discussed in the present review are summarized in Figure 2.

3. Uptake of Non-Metallic NPs

3.1. Uptake of Non-Metallic NPs by Normal CNS-Derived Cells In Vitro

Various cells responsible for maintaining brain homeostasis can accumulate nmNPs. The most important seems to be microglia, brain macrophages responsible for clearance from damaged or dead cells and other dangerous particles [18]. Microglia are also primary cells responsible for the clearance of multi-walled carbon nanotubes (MWNTs), an allotropic form of pure carbon structured in a shape of empty nested fibers (concentric cylinders). MWTNTs distinguish from other types of nanomaterials, as their length can be thousands of times larger than their diameter [19]. Carboxylic acid group modified MWNTs were observed inside microglial cells in ex vivo mixed cultures isolated from the rat’s brain [20]. In vitro, an average phagocytosis time for MWNTs by the human microglia (BV-2 cells) was approximately 2 h, and MWNTs were completely internalized after 6 h [21]. In line, the presence of acid-oxidized or pristine MWNTs in N9 microglial cells was observed after several hours of treatment [22]. The MWNTs accumulated in various cellular compartments, such as cytoplasm, phagolysosomes, endosomes and lysosomes, with the exception of the nucleus, which seemed to be free from MWNTs [20,22]. With regard to other types of nmNPs internalized by microglial cells, the following can be listed: silica nanoparticles (SiNPs) [23], indocyanine green (ICG)-coated polycaprolactone (PCL)-rhodaminedoped NPs, Si-ICG/PCL-polylactic acid (PLLA)-rhodamine-doped NPs [24] and G4 and G4-C12-modified polyamidoamine (PAMAM) dendrimers [25]. The internalization of cyanine5-labeled hydroxyl terminated 4th generation (G4) PAMAM dendrimers by microglial cells was more efficient than by other types of brain cells, such as astrocytes, oligodendrocytes or neurons [26,27,28,29,30].
Besides microglia, the ability to internalize nmNPs has been shown for astrocytes, which are involved in many processes, including the biochemical support of endothelial cells, the formation of blood–brain barrier (BBB), the provision of nutrients to the nervous tissue, maintenance of extracellular ion balance and repair and scarring process of the brain [31]. The astrocytes internalized MWNTs, less efficiently than microglial cells [19]; however, the process was facilitated by the functionalization of MWNTs with amine groups [32].
The internalization of nmNPs by neurons, paramount cells in the CNS, is less efficient when compared to the other cell types. Resveratrol-loaded NPs based on poly(N-vinylpyrrolidone)-b-poly(ε-caprolactone) (PVP-b-PCL), were mainly localized in the cytoplasm, dendrites and axons of cortical neurons [33]. In line, neurons internalized also polylactide-co-glycolic-acid (PLGA) NPs loaded with curcumin and modified with g7 ligand [34], poly(lactide-co-glycolide)-block-poly(ethylene glycol) (PLGA-PEG) conjugated with B6 peptide and loaded with curcumin [35] or lactose myristoyl carboxymethyl chitosan and algal polysaccharide myristoyl carboxymethyl chitosan NPs [36]. In addition, mouse hippocampal neurons and mouse dorsal root ganglion neurons internalized nanodiamonds and accumulated them inside the cell bodies of cortical neurons and inside membrane-surrounded organelles [37]. It was also shown that PAMAM dendrimers were able to penetrate into human neural progenitor cells cultured as a 3D neurosphere model [38] and that amino-functionized G5 PAMAM dendrimers accumulate in the plasma membrane of neurons, as soon as 30 min after treatment [39].
Finally, the presence of nmNPs was also reported for brain endothelial cells. This type of cell maintains the delicate balance of ions, nutrients and other molecules essential for proper brain function and is responsible for removing toxins from the CNS [40]. Liu et al. [41] showed that SiNPs modified with PEG were taken up by mouse cerebral endothelial cells (bEnd.3) with the efficiency dependent on the size of NPs. This was confirmed by Ye et al. [42], who reported that the uptake of SiNPs by human capillary microvascular endothelial cells (hCMEC/D3) was more efficient for 50 nm NPs than for 200 nm NPs. SiNPs were localized inside intracellular vesicles along the endo–lysosomal pathway, inside membrane-bound vesicles and in late endosomes. Other kinds of nmNPs taken up by brain endothelial cells include PEG and polyethylenimine nanogel [43], amino-functionalized MWNTs [32] and fullerenes [44]. The interaction of MWNTs with the plasma membrane of endothelial cells was observed after 4 h of incubation, while their accumulation in endoplasmic vesicles and multi-vesicular bodies was observed after 24 h of treatment and was depended on the NPs concentration [32].

3.2. Uptake of Non-Metallic NPs by CNS-Derived Cancer Cells In Vitro

Though NP uptake by normal brain cells may have adverse effects, the internalization of NPs by brain cancer cells opens the possibility for the development of new NP-based theranostic therapies. Among several types of brain cancers, neuroblastoma and astrocytoma are the most frequent ones [45,46] and are often used for testing their interactions with NPs. Various neuroblastoma and astrocytoma cells were able to take up nmNPs, including SiNPs [47], rhodamine doped Si-ICG/PCL/PLLA NPs [24], curcumin-loaded lactoferrin NPs [48], curcumin-loaded PLGA NPs [49], carbon nanotubes (CNTs) [50], and polymeric NPs and liposomes [51]. These particles accumulated in a membrane region, cytoplasm and a region over the nucleus [47,49,50]. According to Listik et al. [51], the efficiency of internalization depends on the type of NPs and cancer cells. The authors studied an uptake of polysorbate-80 coated polymeric NPs and liposomes coupled with G-protein estrogen receptor selective agonist, in N2a and SHSY5Y neuroblastoma cells. Liposomes were identified inside N2a cells after 6 and 24 h of incubation, while polymeric NPs were detected only after 24 h. In the case of SHSY5Y cells, liposomes were internalized in less than 30 min, whereas the internalization of polymeric NPs is characterized by a slower kinetic profile (internalization was observed after more than 6 h).
Gliomas, which arise from glial cells of the brain or spine and are one of the most invincible cancers, constitute another important group of brain cancers [52]. Many studies indicated that nmNPs were easily taken up by glioma cells, e.g., C6 cells. The NPs tested in this in vitro model include G7 PAMAM dendrimers with amine, acetamide and carboxylate end groups [53], coumarin-6 loaded D-α-tocopheryl PEG1000 succinate (TPGS) coated liposomes [54], PEG-PCL NPs loaded with resveratrol [55], curcumin-loaded polysaccharide nanoformulations based on hyaluronic acid and chitosan hydrochloride NPs [56], temozolomide-loaded PLGA NPs functionalized with anti-EPHA3 [57], and transferrin-conjugated polylactide (PLA)-D-α-tocopheryl-PEG-succinate diblock copolymer NPs [58]. The uptake of NPs occurs very fast; the NPs were observed inside glioma C6 cells after 2 h of incubation [59]. Moreover, it was shown that modifications of NPs and/or targeting with receptor-specific antibody (ephrin type-A receptor 3 tyrosine kinase antibody-modified PLGA NPs and enhancer-modified albumin NPs) change internalization efficacy [57,60].

3.3. Uptake of Non-Metallic NPs by CNS In Vivo

In order to reach the brain, NPs must cross the BBB, a decisive barrier between the brain and systemic circulation. This structure is formed by a complex system of endothelial cells, astroglia, pericytes and perivascular mast cells and is stabilized by a multiprotein complex that seals gaps between endothelial cells and prevents leakage of the barrier, known as a tight junction. The BBB plays an important metabolic role by disposing of waste products, metabolizing different chemical compounds, both drugs and toxins, and protecting the CNS from changes in the ionic composition of cerebrospinal fluid [61,62].
NPs can penetrate through the tight junctions of BBB and their presence in brain cells and tissue was described for single-walled carbon nanotubes (SWNTs) [63], cationic albumin conjugated PEG–PLLA NPs [64], SiNPs [65] and many others. Translocation from systemic circulation to brain tissue occurred very fast. Labeled, amino-functionalized MWNTs were present in the brain of mice 5 min after intravenous injection [32]. Similarly, poly(n-butylcyanoacrylate) dextran polymer NPs coated with polysorbate 80 were detected in mice brains 18 min after injection [66]. The mechanisms of NP uptake by brain cells include phagocytosis, macropinocytosis, and clathrin and caveolin-mediated endocytosis [25,67]. Different receptors may also be involved, as was shown for folic acid [68], apolipoprotein E, low-density lipoprotein receptors and GLUT transporter [67,69].
Initial localization of NPs in the brain tissue depends mainly on a mode of administration. Intravenously administered amino-functionalized MWNTs localized mainly in brain capillaries and parenchyma fraction [32], whereas radiolabeled SiNPs modified with aminopropyltriethoxysilane applied to mice through intranasal instillation localized mostly in the striatum. From brain microvasculature, NPs can be transferred to other regions of the brain, such as the hippocampus (CA1 and CA3 regions), brain stem, cerebellum, and frontal cortex, while intranasal administration resulted in the presence of NPs also in the olfactory bulb [65]. The final localization of NPs depends mostly on their modifications and mode of administration. Bardi et al. [70] tested the internalization of oxidized and non-oxidized amino-functionalized MWNTs (oxMWNT-NH3+ and MWNT-NH3+, respectively) and reported that, after stereotactic administration, MWNTs-NH3+ were abundantly and evenly dispersed along the injection site and within the brain parenchyma, while oxMWNTs-NH3+ were observed mainly as small clusters in intracellular vesicles in microglia, astrocytes and neurons, with the minority in extravesicular cytoplasmic or brain parenchymal areas. After intracortical administration, MWNTs-NH3+ were dispersed throughout the brain parenchyma, forming small aggregates inside membranous intracellular vesicles or as a single nanotube residing in the cytoplasm. In contrast, oxMWNTs-NH3+ were present in clusters and were rarely seen in the cytoplasm, as a single nanotube.
Despite NP modifications and mode of administration, NP uptake and localization in the brain tissue depend also on their size, coating and cargo. Size-dependent differences in the uptake and biodistribution of nmNPs were reported for fluorescent polystyrene latex nanospheres. 20 nm fluorescent polystyrene latex nanospheres were not detected in the brain of rats after intravenous injection or oral pharyngeal aspiration, whereas 100 nm and 1000 nm spheres were present in the CNS 24 h after administration [71]. Similarly, the study with hydroxyl-PAMAM dendrimers in a dog model of a hypothermic circulatory arrest revealed that G6 dendrimers (approx. diameter 6.7 nm) showed extended blood circulation time and increased accumulation in cerebrospinal fluid (CSF), hippocampus, cerebellum and cortex of the injured brain, whereas smaller, G4 dendrimers (approx. diameter 4.3 nm) were undetectable in the brain even 48 h after the final administration [30].
It was also shown that coating with a ligand for the low-density lipoprotein receptor-related protein-1 peptide (angiopep-2), antibodies, bovine serum albumin, apolipoprotein E, and transferrin improve the NP uptake by CNS. MWNTs coated with angiopep-2 were more intensively taken up by glioma-bearing brains than NPs without the ligand [72]. Enhanced accumulation in brain tissue was also observed for transferrin-loaded solid lipid NPs [73], liposomes with bovine serum albumin [74], biodegradable polymersomes with cationic bovine serum albumin [75] and antibody-coated polymer-based NPs [76]. This effect was attributed to enhanced phagocytosis by macrophages [76]. Enhanced uptake of NPs into the brain was also observed after the coating of NPs with a nonionic surfactant—polysorbate 80, nonionic block copolymer—poloxamer 188 and chitosan polysaccharide [77]. In line, Calvo et al. [78] reported that PEGylated poly(cyanoacrylate) NPs can penetrate the brain of mice and rats. The concentration of PEGylated NPs was significantly higher than uncoated ones in the majority of brain structures, especially those in the deeper regions of the brain (striatum, hippocampus, hypothalamus and thalamus). Further, fluorescent COOH-modified polystyrene particles covalently modified with methoxyPEG(NH2) rapidly diffused within normal rat brain tissue, but only if coated with an exceptionally dense layer of PEG [79].
An Important factor affecting the penetration of NPs is also the brain intactness. The ability of G6 dendrimers to cross the BBB correlated with the extent of CNS inflammation, while their accumulation was more efficient in injured places [30]. Similarly, in a rabbit cerebral palsy model, ethylenediamine-core PAMAM dendrimers were more abundantly present in the brain parenchyma of regions characterized with significant inflammation, as compared to the healthy regions [26].

4. Toxicity of Non-Metallic NPs in Mammals

4.1. In Vitro and Ex Vivo Toxicity

Many reports indicate that metallic NPs are toxic to CNS. Their neurotoxicity is associated with the induction of oxidative stress in the brain tissue through the release of metal ions [11]. Similarly, numerous reports also show detrimental effects of nmNPs (Table 1).
The tables included in this publication present information on the toxic and non-toxic effects of various types of NPs on the CNS (in vitro and in vivo studies).
In vitro research on various types of brain cells clearly show that toxicity of nmNPs is dependent on cell type and uptake of NPs. In accordance, Du et al. [81] observed an intense uptake of SiNPs by microglial N9 and endothelial bEnd.3 cells, while neuronal HT22 cells barely internalized the NPs. Consequently, toxicity of SiNPs described as morphological changes, such as swelling and cell membrane blebbing, was higher in BV-2 and N9 cell lines than in HT22 cells. A similar result was reported for CNTs, as sensitivity of internalization prone microglia to CNTs was higher than internalization tardy neurons [19].On the other hand, many publications reported also the lack of detrimental effects of nmNPs. In the study conducted by Ducray et al. [24] exposure of primary hippocampal cultures to SiNPs coated ICG/PCL-rhodamine-doped NPs or Si-ICG/PLLA-rhodamine-doped NPs did not affect cell viability, however as noticed earlier, these NPs impaired cell differentiation. The lack of impact of SiNPs on cellular viability was also observed for endothelial cells (bEnd.3) [96] and for A-172 brain cells [105]. In addition, carbon NPs [19,21,22,32,44,97], modified PAMAM dendrimers [28,89,106], PEG–based dendrimers [107], polymer NPs [102], PEG/polyethylenimine (“nanogel”) NPs modified with oligonucleotides [43], PLGA NPs modified with PEG and phage-displayed peptides [101], solid lipid NPs, as well as albumin and chitosan NPs [67,75,91,108] were also reported to be non-toxic to glial cells, neurons and endothelial cells.
Toxicity of nmNPs to CNS-derived normal cells is definitely an unfavorable phenomenon. On the other hand, toxicity to cancer cells may be exploited for development of new anticancer drugs provided that selective accumulation of NPs in cancer cells can be achieved. Numerous reports indicate the toxic effects of nmNPs on glioma and glioblastoma cells in vitro. Transferrin conjugated PLA-D-α-tocopheryl PEG succinate diblock copolymer NPs [58] inhibited growth of glioma C6 cells. In line, exposure to PAMAM dendrimers and dendrons resulted in reduced cell size, rounded shape and loss of neurites of cells belonging to several glioma lines [109].
In contrast to the above results, several publications showed the lack of toxicity of nmNPs to CNS-derived cancer cells, e.g., MWNTs and SWNTs shown no toxicity to glioma GL261 [21,110], also no adverse effects were observed after treatment of glioblastoma A-172 cell line with SiNPs [105].
More detailed information about nmNPs effects on CNS-derived non-cancer and cancer cells in vitro is presented in Table 1 and Table 2.

4.2. Toxicity In Vivo

Many in vivo studies conducted using mammals describe neurotoxic effects associated with the presence of nmNPs in the brain (Table 3).
Interestingly, exposure to nmNPs might exert different effects in various brain regions, e.g., polysorbate 80-modified chitosan NPs were deposited mainly in the frontal cortex and cerebellum of the rats’ brain after systemic injection. Although no signs of oxidative stress were observed, apoptosis and necrosis of neurons and mild inflammatory response were observed in the frontal cortex, whereas in the cerebellum a decrease in the expression of glial fibrillary acidic protein was the only detected sign of degenerative changes [143]. In line, though reductions in the number and dispersion of Nissl bodies were observed in neurons of mice exposed to FITC-tagged SiNPs, the effect was spatially limited to the frontal cortex and was not present in the other brain regions, such as the hippocampus [137]. The toxicity of nmNPs partially depends on the disruption of BBB integrity, as acute pulmonary exposure to MWNTs caused a neuroinflammatory response in rodents that was dependent on the disruption of BBB integrity [141].
On the other hand, many in vivo studies show also that nmNPs are low- or non-toxic to CNS. Bardi et al. [97] demonstrated that the injection of MWNTs coated with pluronic F127 surfactant into mice brain resulted only in a small injury area. No damage to the overall brain structure and tissue was observed. No pathological changes were also observed in mice brain after 28 days of exposure to carboxylated MWNTs [147]. No adverse effects associated with CNS in rodents, despite the penetration of BBB, were also detected for other kinds of carbon NPs, such as fullerenes [148], dextran-coated graphene oxide nanoplatelets [163], water soluble fluorescent carbon nano onions [151] and nanodiamonds [37]. In line, similar results were published for other types of nmNPs, such as SiNPs [105], dendrimers, dendriplexes [90] and poly(n-butyl cyanoacrylate)dextran polymers [66]. See Table 3 for details.

5. Mechanism of nmNPs Toxicity. Other Adverse Effects of Non-Metallic NPs in Mammals

In spite of direct interactions with cells or cellular components leading to non-metallic nanomaterials toxicity, several indirect mechanisms were proposed to have a role in this process.
Non-toxic doses of SiNPs caused Ca2+ flux into neuronal cells in a size and surface charge dependent manner, stimulating long-lasting but reversible calcium signaling. Voltage-dependent and transient receptor potential-vanilloid 4 channels were involved in this process. Interestingly, NP internalization was not necessary to induce the calcium flux [83]. The enhancement of Ca2+ concentration after interaction with plasma membrane was also observed in pyramidal neurons and astroglial cells of rat hippocampal slices treated with 5th generation dendrimers (PAMAM G5). The increase in Ca2+ concentration was followed by the mitochondria depolarization of astroglial cells [39].
Another indirect mechanism of toxicity of non-metallic NPs involves alterations in gene and protein expression. Inhibition of the expression of mitochondrial deacetylase SIRT3 that plays an important role in regulation of cellular metabolism was proposed as a cause of increased oxidative stress in mitochondria and reprogramming of cellular metabolism in LPS-stimulated mice microglia treated with single-walled carbon nanohorns. This resulted in G1 arrest and increased the apoptosis of treated cells [87].
The treatment of rat brain capillary endothelial cells with gold and polymer-coated Si-ICG/PCL and Si-ICG/PCL/PLLA NPs resulted in a time- and concentration-dependent decrease in the phosphorylation of MAPKs, which participate in cellular response to a diverse array of stimuli, such as mitogens, osmotic stress and pro-inflammatory cytokines, and regulate cell proliferation, gene expression, differentiation, mitosis and many others [82]. Yet, another alteration in gene expression was shown in a 3D model of human neural progenitor cells incubated with PAMAM-NH2 dendrimers. In this model, NPs inhibited cell proliferation and migration, which was accompanied by the down-regulation of several genes, including early growth response gene 1 (EGR1), insulin-like growth factor-binding protein 3 (IGFBP3), tissue factor pathway inhibitor (TFPI2) and adrenomedullin (ADM) [38].
Incubation of neuroblastoma N2a cells with MWNTs promoted nuclear translocation and acetylation of NFκB transcription factor in a dose-dependent manner, followed by up-regulation of nNOS and an increase in NO production [111]. The NFκB-dependent signaling pathway is involved in the regulation of immunity, inflammation, cell differentiation, proliferation and apoptosis [164].
Another important cellular signaling pathway affected by nmNPs is the p53-mediated pathway. The incubation of PC12 cells, often used as a model of dopaminergic neurons, with SiNPs, caused an upregulation of p21 and GADD45A proteins that resulted in G2/M arrest and induction of apoptosis [65]. The upregulation of p21 and GADD45A proteins, accumulation of cells in the G2/M phase of cell cycle and induction of apoptosis were also observed in primary cultures of rats’ cortical neurons and N9 microglia cells after incubation with SiNPs [81].
Another mechanism that has been proposed to be responsible for adverse effects of nmNPs is alteration in redox balance. Apoptosis triggered by ROS production was proposed as a mechanism of SWNTs and graphene toxicities. ROS generation in PC12 cells treated with graphene layers [113] and SWNTs [114] resulted in the upregulation of caspase 3 apoptosis and was dependent on the PEG coating of nanomaterials. MWNTs were also able to induce alterations in the microenvironment and microstructure of brain tissue associated with NOS and ROS production. After treatment with MWNTs, high expression of nNOS and increased ROS production were observed in rostral ventrolateral medulla and nucleus tractus solitaries—two regions of the brain, which play important roles in the regulation of sympathetic nerve activity [111]. This result was confirmed by Zheng et al. [88] who proved that inhalation of MWNTs significantly alters the balance between the sympathetic and parasympathetic nervous system in rats. An enhanced ROS production and lipid peroxidation were also observed in rat hippocampus after exposure to SWNTs functionalized with PEG [142] and SiNPs [138]. The latter was confirmed by an increase in the expression of superoxide dismutase (SOD) and catalase (CAT) activity, two enzymes responsible for the removal of ROS [138].
Incubation of microglial BV-2 and N9 cells with MWNTs and single-wall carbon nanohorns resulted in a dose-dependent cell division arrest and apoptosis [86,87]. Similarly, the increase in apoptosis level was reported for neuronal NeuroScreen-1 cells (NS-1) incubated with MWNTs [85] or primary neuronal cells treated with G4-C12 PAMAM dendrimers [25].
Yet, another indirect effect mechanism leading to nmNPs toxicity in CNS-derived cells involves the release of pro-inflammatory cytokines that may led to inflammation in the brain. N9 microglial cells treated with SiNPs produced pro-inflammatory interleukin 1β (IL-1β) and N-terminal fragment of gasdermin D, a marker protein for pyroptosis [81], whereas nanoplastic treatment induced pro-inflammatory response in astrocytes, including the up-regulation of tumor necrosis factor alpha and IL-1β [165]. MWNTs caused an increase in TNF-α and IL-1β expression, indicating the pro-inflammatory action of nmNPs inside the brain (in vivo studies). The release of pro-inflammatory cytokines was also reported for shortened-by-oxidation, amino-functionalized MWNT (oxMWNTs-NH3+). Furthermore, oxMWNTs-NH3+ induced the higher-than-long, non-oxidized analog expression of GFAP and CD11b, which points to the more intense glial cell activity and degenerative changes [70]. An increase in the expression of TNF-α and IL-1β was also observed in the striatum of rats treated with SiNPs [65].
nmNPs also affected the functionality of neural cells, such as the differentiation and formation of neurite and dendrites. Treatment of SH-SY5Y cells with ICG/PCL-rhodamine-doped NPs and Si-ICG/PLLA-rhodamine-doped NPs resulted in a significant down-regulation of the expression of differentiation marker MAP-2 [24]. Accordingly, a study by Hollinger et al. [27] showed that incubation of rabbits primary mixed glial cells with 2-PMPA dendrimers triggered up-regulation of transforming growth factor beta (TGFβ), which plays an important role in the regulation and differentiation of immune and stem cells. Treatment of primary cultured cortical neurons with increased concentrations of SiNPs resulted in a statistically significant decrease in the number of dendrites. Furthermore, subsequent analysis showed that soma of the neurons collapsed and dendrites disappeared [80]. In line, reduction in the ability to form neurites after NGF stimulation was observed in PC12 cells treated with SiNPs modified with aminopropyltriethoxysilane, likely due to disorder in cytoskeletal structure [65].
The impact of nmNPs on neurotransmitter secretion was described by Wu et al. [65], who showed that exposure of rats to SiNPs caused a decrease in dopamine level in the striatum and in hippocampus. In line tissue, decreased levels of epinephrine, norepinephrine and dopamine was reported in the blood of mice exposed to MWNTs [111]. The negative impact of SiNPs on the functionality of dopaminergic neurons shows that these nmNPs can be a risk factor of neurodegenerative diseases. Further evidence that SiNPs might be at risk of neurodegenerative diseases was delivered by You et al. [137], who reported that phosphorylation Tau protein was significantly increased in the frontal cortex of SiNP exposed mice. It was accompanied by increased phosphorylation of ERK in the frontal cortex and hippocampus, and c-Jun N-terminal kinases (JNK) in the frontal cortex. The impairment of Tau protein phosphorylation is associated with neurodegeneration and contributes to the development of Alzheimer disease. The above-mentioned changes were associated with microglia activation and upregulation of pro-inflammatory markers, suggesting that exposure to SiNPs can lead to neuroinflammation that underlies many neurodegenerative disorders. The same work showed the significant impact of SiNP exposure on synapses function and structure. Exocytosis and endocytosis are essential processes in synapse firing, allowing communication to occur between neurons. It was found that exocytosis was significantly impaired in the SiNP-exposed mice in the frontal cortex. However, neither endocytosis nor exocytosis was affected in the hippocampus. Furthermore, after ex vivo exposure of primary cortical neurons to SiNPs, a decrease in the expression of synapsin I and a parallel increase in the expression of synaptophysin were observed, both proteins playing a pivotal role in the proper functioning of synapses.
Finally, nmNPs also affect the behavior of exposed animals. Treatment of mice with SiNPs via intranasal instillation resulted in mood dysfunction and cognitive impairment. Short-term memory and spatial learning, estimated by using a Morris water maze test, were impaired. Furthermore, the mice social interaction activity was decreased after 2 months of NP exposure; however, any symptoms of depression were not detected [137]. On the contrary, Wu et al. [65] reported that intranasal administration of SiNPs for 1 or 7 days did not result in any changes in animals’ behavior or cause histological changes in the brain tissue; however, it should be noted that dose used in this study was relatively low (20 μg/day) and exposition time was shorter, as compared to the study mentioned before. Adverse effects of nmNPs on rodent behavior was confirmed by the study of Dal Bosco et al. [142], who showed that the treatment of rats with SWNTs-PEG caused a significant deficit in the retrieval of fear memory.

6. Adverse Effects of Non-Metallic NPs in Non-Mammalian Organisms

Many types of nmNPs are released into the environment, as a result of their use, and may impact animals living there. Therefore, testing the toxicity of nmNPs on the CNS of higher non-mammalian organisms seems to be crucial for estimating of the environmental risk associated with their use.
Exposure of adult Japanese rice fish (Oryzias latipes) to fluorescent polystyrene NPs for 7 days revealed the presence of particles in the brain, indicating that nanoplastics have the innate capacity to cross the BBB [166]. In line, amino-modified polystyrene nanobeads were more prevalent than similar microparticles in the brain of exposed Crucian carp (Carassius carassius). The presence of polystyrene nano- and microparticles in the brain coincided with alterations in behavioral patterns, decreased brain mass and morphological changes in the cerebral gyri [167]. In line, 15 days of exposure to polyethylene nano- and microplastic caused varying degrees of necrosis, fibrosis, changes in blood capillaries, tissue detachment, edema, degenerated connective tissues, and necrosis in large cerebellar neurons and ganglion cells in the tectum of juvenile common carp (Cyprinus carpio). The changes were accompanied by a decrease in the activity of acetylcholinesterase (aChE) [168]. The gut–brain axis related toxicity of nanoplastic was recently reviewed, and it is clear that these particles may induce oxidative stress, disturb neurodevelopment, and impact behaviour and immune system activation [165].
The studies performed on bivalves and crustaceans confirmed the neurotoxic effect of nanoplastic. It was shown that nanoplastic can inhibit cholinesterase in the hemolymph of Mediterranean mussel (Mytilus galloprovincialis), an enzyme responsible for the breakdown of neurotransmitters [169]. The same effect was observed in brine shrimp (Artemia fransiscana) [170].
Despite nanoplastic, significant lipid peroxidation and GSH depletion were found in the brains of largemouth bass (Micropterus salmoides) after 48 h of exposure to uncoated fullerenes [171]. In line, 21 days exposure of zebrafish (Danio rerio) to fullerenes, short and long MWNTs and SWNTs, caused significant disturbances in lipid, revealed as an elevation of the lipid to protein ratio and in the brain and gills. In addition, a decrease in the level of unsaturated lipids was reported in the brain of fullerene exposed males. In contrast to the result obtained for male zebrafish, the level of unsaturated lipids in the brain in female fish exposed to fullerenes increased [172].
The zebrafish model was also used to study embryonic developmental toxicity of SiNPs. The results revealed persistent changes in larval behavior [173]. In agreement, transcriptomic analysis suggests neurodegeneration and motor dysfunction in larval zebrafish after polystyrene NP treatment. The authors clearly indicate the changes in behavior and physiology, potentially decreasing organismal fitness in contaminated ecosystems [174].
Studies on in vivo toxicity of different types of nmNPs in mammals and non-mammalian organisms are summarized in detail in Table 3 and Table 4.

7. Conclusions

In the past decade, rapid growth of interest in nanotechnology and increasing use of NPs in commercial applications have been widely observed. In general, despite being similar in shape and size, nmNPs seem to be less toxic than metal nanomaterials. The exact cause of the lesser toxicity of nmNPs over similar metallic ones is not yet explained. It might be speculated that the smaller density of nmNPs, as compared with the metallic ones, affects their interactions with cellular components, such as cell membranes, they might differ in composition and/or behavior to protein corona, or a smaller amount of metal ions is released from the NPs that might contribute to the generation of oxidative stress. Whatever it is needs further investigation; however, the impact of nmNPs on human health should not be disregarded.
The summary of current knowledge about the toxicity of nmNPs, presented in this study, demonstrates that nmNPs can easily penetrate the animal body, both of terrestrial and aquatic organisms and can be toxic to the CNS in vivo and to CNS-derived cells in vitro. Though a large and diverse group of nmNPs has been engineered and studied, they seem to share common mechanisms of toxicity. Adverse effects of nmNPs are usually associated with the generation of oxidative stress that leads to the malfunction of mitochondria, the activation of different signaling pathways, and subsequent activation of autophagy or apoptosis. Some nmNPs can, however, directly interact with cell membrane proteins, which leads to the activation of ion channel and flux of ion, e.g., calcium, from cellular milieu. The potential effects of nmNPs in the CNS are summarized in Figure 3.

8. Limitation of the Study and Future Directions

A general impression of non-metallic nanomaterials’ effects on the CNS after reading this review might be biased by the fact that the review is focused on the adverse effects of nmNPs. Thus, it must be understood that there is also a large number of publications describing little or no effects of nmNPs on the CNS. These publications have been omitted as being beyond the scope of this review. However, while limited toxicity might be a potential benefit of non-metallic nanomaterials, predisposing them to biomedical applications, it should be kept in mind that it can be harmful and there is a need to estimate the risk of its use, especially in medicine and diagnostics. While this review clearly shows a negative impact of nmNPs on the CNS, including mammals and non-mammalian vertebrates, their impact on human health in terms of long-term daily exposure is still unclear. Available data in regard to human exposure are very limited, and it is not possible to draw reliable conclusions. Thus, similar to the recently published report by the European Commission Directorate General for Environment on nanoplastic health impact [184], reporting standards should be developed for various nmNPs, which can penetrate different biological barriers and slip through filters more than other particles. This requires enormous work by nanotechnology scientists to conduct numerous experiments associated with the estimation of toxicity-tested materials. The report should include current quantification and assessment methods, the occurrence of NPs in the environment, their ecotoxicity, and what is most importantly their impacts on human health.

Author Contributions

All authors contributed to the study’s conception and design. Data collection and analysis were performed by K.S. (Katarzyna Sikorska), K.B., M.K., L.K.-S., K.S. and M.C. (Magdalena Czajka). The first draft of the manuscript was written by K.S. (Katarzyna Sikorska) and all authors commented on subsequent versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Centre project No. 2019/35/B/NZ7/04133 “Nanoplastic toxicity: effects on the gut-brain axis” and statutory funding for the Institute of Nuclear Chemistry and Technology (K.S. (Katarzyna Sikorska), K.B. and M.K.), National Science Centre Project No. 2020/39/B/NZ7/03197 “Nanoplastic and silver nanoparticles as a factor modulating the activity of estrogen-dependent intracellular signaling pathways in in vitro and in vivo systems” (M.K. and K.B.), and statutory funding from the Institute of Nuclear Chemistry and Technology (K.S. (Katarzyna Sikorska), K.B. and M.K.)

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saeedi, M.; Eslamifar, M.; Khezri, K.; Dizaj, S.M. Applications of nanotechnology in drug delivery to the central nervous system. Biomed. Pharmacother. 2019, 111, 666–675. [Google Scholar] [CrossRef]
  2. Yang, Z.; Liu, Z.W.; Allaker, R.P.; Reip, P.; Oxford, J.; Ahmad, Z.; Ren, G. A review of nanoparticle functionality and toxicity on the central nervous system. J. R. Soc. Interface 2010, 7, 411–422. [Google Scholar] [CrossRef]
  3. Li, W.; Cheng, J.; He, F.; Zhang, P.; Zhang, N.; Wang, J.; Song, Q.; Hou, Y.; Gan, Z. Cell membrane-based nanomaterials for theranostics of central nervous system diseases. J. Nanobiotechnol. 2023, 21, 276. [Google Scholar] [CrossRef]
  4. Ekhator, C.; Qureshi, M.Q.; Zuberi, A.W.; Hussain, M.; Sangroula, N.; Yerra, S.; Devi, M.; Arsal Naseem, M.; Bellegarde, S.B.; Pendyala, P.R. Advances and Opportunities in Nanoparticle Drug Delivery for Central Nervous System Disorders: A Review of Current Advances. Cureus 2023, 15, e44302. [Google Scholar] [CrossRef]
  5. Yang, W.; Zhang, M.; He, J.; Gong, M.; Sun, J.; Yang, X. Central nervous system injury meets nanoceria: Opportunities and challenges. Regen. Biomater. 2022, 9, rbac037. [Google Scholar] [CrossRef]
  6. Annu Sartaj, A.; Qamar, Z.; Md, S.; Alhakamy, N.A.; Baboota, S.; Ali, J. An Insight to Brain Targeting Utilizing Polymeric Nanoparticles: Effective Treatment Modalities for Neurological Disorders and Brain Tumor. Front. Bioeng. Biotechnol. 2022, 10, 788128. [Google Scholar]
  7. Montegiove, N.; Calzoni, E.; Emiliani, C.; Cesaretti, A. Biopolymer Nanoparticles for Nose-to-Brain Drug Delivery: A New Promising Approach for the Treatment of Neurological Diseases. J. Funct. Biomater. 2022, 13, 125. [Google Scholar] [CrossRef]
  8. Athira, S.S.; Prajitha, N.; Mohanan, P.V. Interaction of nanoparticles with central nervous system and its consequences. Am. J Res. Med. Sci. 2018, 4, 12–32. [Google Scholar] [CrossRef]
  9. Facciolà, A.; Visalli, G.; La Maestra, S.; Ceccarelli, M.; D’Aleo, F.; Nunnari, G.; Pellicanò, G.F.; Di Pietro, A. Carbon nanotubes and central nervous system: Environmental risks, toxicological aspects and future perspectives. Environ. Toxicol. Pharmacol. 2018, 65, 23–30. [Google Scholar] [CrossRef]
  10. Zia, S.; Islam Aqib, A.; Muneer, A.; Fatima, M.; Atta, K.; Kausar, T.; Zaheer, C.-N.F.; Ahmad, I.; Saeed, M.; Shafique, A. Insights into nanoparticles-induced neurotoxicity and cope up strategies. Front. Neurosci. 2023, 17, 1127460. [Google Scholar] [CrossRef]
  11. 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]
  12. Baldrighi, M.; Trusel, M.; Tonini, R.; Giordani, S. Carbon Nanomaterials Interfacing with Neurons: An In vivo Perspective. Front. Neurosci. 2016, 10, 250. [Google Scholar] [CrossRef]
  13. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and Toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  14. Zhang, W.; Sigdel, G.; Mintz, K.J.; Seven, E.S.; Zhou, Y.; Wang, C.; Leblanc, R.M. Carbon Dots: A Future Blood–Brain Barrier Penetrating Nanomedicine and Drug Nanocarrier. Int. J. Nanomed. 2021, 16, 5003–5016. [Google Scholar] [CrossRef]
  15. Zielinski, A.; Majkowska-Marzec, B. Whether Carbon Nanotubes Are Capable, Promising, and Safe for Their Application in Nervous System Regeneration. Some Critical Remarks and Research Strategies. Coatings 2022, 12, 1643. [Google Scholar] [CrossRef]
  16. Gauro, R.; Nandave, M.; Jain, V.K.; Jain, K. Advances in dendrimer-mediated targeted drug delivery to the brain. J. Nanoparticle Res. 2021, 23, 76. [Google Scholar] [CrossRef]
  17. Masserini, M. Nanoparticles for brain drug delivery. ISRN Biochem. 2013, 2013, 1–18. [Google Scholar] [CrossRef]
  18. Li, Q.; Barres, B.A. Microglia and macrophages in brain homeostasis and disease. Nat. Rev. Immunol. 2018, 18, 225–242. [Google Scholar] [CrossRef]
  19. Bussy, C.; Al-Jamal, K.T.; Boczkowski, J.; Lanone, S.; Prato, M.; Bianco, A.; Kostarelos, K. Microglia Determine Brain Region-Specific Neurotoxic Responses to Chemically Functionalized Carbon Nanotubes. ACS Nano. 2015, 9, 7815–7830. [Google Scholar] [CrossRef]
  20. Bussy, C.; Hadad, C.; Prato, M.; Bianco, A.; Kostarelos, K. Intracellular degradation of chemically functionalized carbon nanotubes using a long-term primary microglial culture model. Nanoscale 2016, 8, 590–601. [Google Scholar] [CrossRef]
  21. Kateb, B.; Van Handel, M.; Zhang, L.; Bronikowski, M.J.; Manohara, H.; Badiea, B. Internalization of MWCNTs by microglia: Possible application in immunotherapy of brain tumors. NeuroImage 2007, 37, 9–17. [Google Scholar] [CrossRef]
  22. Goode, A.E.; Gonzalez Carter, D.A.; Motskin, M.; Pienaar, I.S.; Chen, S.; Hu, S.; Ruenraroengsak, P.; Ryan, M.P.; Shaffer, M.S.P.; Dexter, D.T.; et al. High resolution and dynamic imaging of biopersistence and bioreactivity of extra and intracellular MWNTs exposed to microglial cells. Biomaterials 2015, 70, 57–70. [Google Scholar] [CrossRef]
  23. Choi, J.; Zheng, Q.; Katz, H.E.; Guilarte, T.R. Silica-Based Nanoparticle Uptake and Cellular Response by Primary Microglia. Environ. Health Perspect. 2010, 118, 589–595. [Google Scholar] [CrossRef]
  24. Ducray, A.D.; Stojiljkovic, A.; Möller, A.; Stoffel, M.H.; Widmer, H.R.; Frenz, M.; Mevissen, M. Uptake of silica nanoparticles in the brain and effects on neuronal differentiation using different in vitro models. Nanomedicine 2017, 13, 1195–1204. [Google Scholar] [CrossRef]
  25. Albertazzi, L.; Gherardini, L.; Brondi, M.; Sato, S.S.; Bifone, A.; Pizzorusso, T.; Ratto, G.M.; Bardi, G. In Vivo Distribution and Toxicity of PAMAM Dendrimers in the Central Nervous System Depend on Their Surface Chemistry. Mol. Pharm. 2013, 10, 249–260. [Google Scholar] [CrossRef]
  26. Dai, H.; Navath, R.S.; Balakrishnan, B.; Guru, B.R.; Mishra, M.K.; Romero, R.; Kannan, R.M.; Kannan, S. Intrinsic targeting of inflammatory cells in the brain by polyamidoamine dendrimers upon subarachnoid administration. Nanomedicine 2010, 5, 1317–1329. [Google Scholar] [CrossRef]
  27. Hollinger, K.; Sharma, A.; Tallon, C.; Lovell, L.; Thomas, A.G.; Zhu, X.; Kambhampati, S.P.; Liaw, K.; Sharma, R.; Rojas, C.; et al. Microglia-targeted dendrimer-2PMPA therapy robustly inhibits GCPII and improves cognition in a mouse model of multiple sclerosis. bioRxiv 2020. preprint. [Google Scholar] [CrossRef]
  28. Salgado, A.J.; Oliveira, J.M.; Pirraco, R.P.; Pereira, V.H.; Fraga, J.S.; Marques, A.P.; Neves, N.M.; Mano, J.F.; Reis, R.L.; Sousa, N. Carboxymethylchitosan/Poly(amidoamine) Dendrimer Nanoparticles in Central Nervous Systems-Regenerative Medicine: Effects on Neuron/Glial Cell Viability and Internalization Efficiency. Macromol. Biosci. 2010, 10, 1130–1140. [Google Scholar] [CrossRef]
  29. Zhang, F.; Nance, E.; Alnasser, Y.; Kannan, R.; Kannan, S. Microglial migration and interactions with dendrimer nanoparticles are altered in the presence of neuroinflammation. J. Neuroinflamm. 2016, 13, 65. [Google Scholar] [CrossRef]
  30. Zhang, F.; Magruder, J.T.; Lin, Y.A.; Crawford, T.C.; Grimm, J.C.; Sciortino, C.M.; Wilson, M.A.; Blue, M.E.; Kannan, S.; Johnston, M.V.; et al. Generation-6 hydroxyl PAMAM dendrimers improve CNS penetration from intravenous administration in a large animal brain injury model. J. Control. Release 2017, 10, 173–182. [Google Scholar] [CrossRef]
  31. Siracusa, R.; Fusco, R.; Cuzzocrea, S. Astrocytes: Role and Functions in Brain Pathologies. Front. Pharmacol. 2019, 10, 1114. [Google Scholar] [CrossRef]
  32. Kafa, H.; Wang, J.T.W.; Rubio, N.; Venner, K.; Anderson, G.; Pach, E.; Ballesteros, B.; Preston, J.E.; Abbott, N.J.; Al-Jamal, K.T. The interaction of carbon nanotubes with an in vitro blood-brain barrier model and mouse brain in vivo. Biomaterials 2015, 53, 437–452. [Google Scholar] [CrossRef]
  33. Lu, X.; Xu, H.; Sun, B.; Zhu, Z.; Zheng, D.; Li, X. Enhanced Neuroprotective Effects of Resveratrol Delivered by Nanoparticles on Hydrogen Peroxide-Induced Oxidative Stress in Rat Cortical Cell Culture. Mol. Pharm. 2013, 10, 2045–2053. [Google Scholar] [CrossRef]
  34. Barbara, R.; Belletti, D.; Pederzoli, F.; Masoni, M.; Keller, J.; Ballestrazzi, A.; Vandelli, M.A.; Tosi, G.; Grabrucker, A.M. Novel Curcumin loaded nanoparticles engineered for Blood-Brain Barrier crossing and able to disrupt A-beta aggregates. Int. J. Pharm. 2017, 30, 413–424. [Google Scholar] [CrossRef]
  35. Fan, S.; Zheng, Y.; Liu, X.; Fang, W.; Chen, X.; Liao, W.; Jing, X.; Lei, M.; Tao, E.; Ma, Q.; et al. Curcumin-loaded PLGA-PEG nanoparticles conjugated with B6 peptide for potential use in Alzheimer’s disease. Drug Deliv. 2018, 25, 1044–1055. [Google Scholar] [CrossRef]
  36. Kou, C.H.; Han, J.; Han, X.L.; Zhuang, H.J.; Zhao, Z.M. Preparation and characterization of the Adriamycin loaded amphiphilic chitosan nanoparticles and their application in the treatment of liver cancer. Oncol. Lett. 2017, 17, 7833–7841. [Google Scholar] [CrossRef]
  37. Huang, Y.A.; Kao, C.W.; Liu, K.K.; Huang, H.S.; Chiang, M.H.; Soo, C.R.; Chang, H.C.; Chiu, T.W.; Chao, J.I.; Hwang, E. The effect of fluorescent nanodiamonds on neuronal survival and morphogenesis. Sci. Rep. 2014, 4, 6919. [Google Scholar] [CrossRef]
  38. Zeng, Y.; Kurokawa, Y.; Zeng, Q.; Win-Shwe, T.T.; Nansai, H.; Zhang, Z.; Sone, H. Effects of Polyamidoamine Dendrimers on a 3-D Neurosphere System Using Human Neural Progenitor Cells. Toxicol. Sci. 2016, 152, 128–144. [Google Scholar] [CrossRef]
  39. Nyitrai, G.; Héja, L.; Jablonkai, I.; Pál, I.; Visy, J.; Kardos, J. Polyamidoamine dendrimer impairs mitochondrial oxidation in brain tissue. J. Nanobiotechnol. 2013, 11, 9. [Google Scholar] [CrossRef]
  40. Stamatovic, S.M.; Keep, R.F.; Andjelkovic, A.V. Brain Endothelial Cell-Cell Junctions: How to “Open” the Blood Brain Barrier. Curr. Neuropharmacol. 2008, 6, 179–192. [Google Scholar] [CrossRef]
  41. Liu, D.; Lin, B.; Shao, W.; Zhu, Z.; Ji, T.; Yang, C. In Vitro and in Vivo Studies on the Transport of PEGylated Silica Nanoparticles across the Blood–Brain Barrier. ACS Appl. Mater. Interfaces 2014, 6, 2131–2136. [Google Scholar] [CrossRef]
  42. Ye, D.; Raghnaill, M.N.; Bramini, M.; Mahon, E.; Aberg, C.; Salvati, A.; Dawson, K.A. Nanoparticle accumulation and transcytosis in brain endothelial cell layers. Nanoscale 2013, 5, 11153–11165. [Google Scholar] [CrossRef]
  43. Vinogradov, S.V.; Batrakova, E.V.; Kabanov, A.V. Nanogels for Oligonucleotide Delivery to the Brain. Bioconjug. Chem. 2004, 15, 50–60. [Google Scholar] [CrossRef]
  44. Lao, F.; Chen, L.; Li, W.; Ge, C.; Qu, Y.; Sun, Q.; Zhao, Y.; Han, D.; Chen, C. Fullerene Nanoparticles Selectively Enter Oxidation-Damaged Cerebral Microvessel Endothelial Cells and Inhibit JNK Related Apoptosis. ACS Nano 2009, 3, 3358–3368. [Google Scholar] [CrossRef]
  45. Van Arendonk, K.J.; Chung, D.H. Neuroblastoma: Tumor Biology and Its Implications for Staging and Treatment. Children 2019, 6, 12. [Google Scholar] [CrossRef]
  46. Wang, J.; Bettegowda, C. Genomic discoveries in adult astrocytoma. Curr. Opin. Genet. Dev. 2015, 30, 17–24. [Google Scholar] [CrossRef]
  47. Lungare, S.; Hallam, K.; Badhan, R.K.S. Phytochemical-loaded mesoporous silica nanoparticles for nose-to-brain olfactory drug delivery. Int. J. Pharm. 2016, 20, 280–293. [Google Scholar] [CrossRef]
  48. Bollimpelli, V.S.; Kumar, P.; Kumari, S.; Kondapi, A.K. Neuroprotective effect of curcumin-loaded lactoferrin nano particles against rotenone induced neurotoxicity. Neurochem. Int. 2016, 95, 37–45. [Google Scholar] [CrossRef]
  49. Paka, G.D.; Doggui, S.; Zaghmi, A.; Safar, R.; Dao, L.; Reisch, A.; Klymchenko, A.; Roullin, V.G.; Joubert, O.; Ramassamy, C. Neuronal Uptake and Neuroprotective Properties of Curcumin- Loaded Nanoparticles on SK-N-SH Cell Line: Role of Poly(lactide-coglycolide) Polymeric Matrix Composition. Mol. Pharm. 2016, 13, 391–403. [Google Scholar] [CrossRef]
  50. Dong, L.; Witkowski, C.M.; Craig, M.M.; Greenwade, M.M.; Joseph, K.L. Cytotoxicity Effects of Different Surfactant Molecules Conjugated to Carbon Nanotubes on Human Astrocytoma Cells. Nanoscale Res. Lett. 2009, 4, 1517–1523. [Google Scholar] [CrossRef]
  51. Listik, E. Development and optimization of G-1 polymeric nanoparticulated and liposomal systems for central nervous system applications. Neurol. Disord. Therap. 2018, 2, 1–11. [Google Scholar]
  52. Feng, C.; Wu, Y.; Gao, L.; Guo, X.; Wang, Z.; Xing, B. Publication Landscape Analysis on Gliomas: How Much Has Been Done in the Past 25 Years? Front. Oncol. 2020, 9, 1463. [Google Scholar] [CrossRef]
  53. Hong, S.; Rattan, R.; Majoros, I.J.; Mullen, D.G.; Peters, J.L.; Shi, X.; Bielinska, A.U.; Blanco, L.; Orr, B.O.; Baker, J.R., Jr.; et al. The Role of Ganglioside GM1 in Cellular Internalization Mechanisms of Poly(amidoamine) Dendrimers. Bioconjug. Chem. 2009, 19, 1503–1513. [Google Scholar] [CrossRef]
  54. Vijayakumar, M.R.; Vajanthri, K.Y.; Balavigneswaran, C.K.; Mahto, S.K.; Mishra, N.; Muthu, M.S.; Singh, S. Pharmacokinetics, biodistribution, in vitro cytotoxicity andbiocompatibility of Vitamin E TPGS coated trans resveratrol liposomes. Colloids Surf. B Biointerfaces 2016, 145, 479–491. [Google Scholar] [CrossRef]
  55. Shao, J.; Li, X.; Lu, X.; Jiang, C.; Hu, Y.; Li, Q.; You, Y.; Fu, Z. Enhanced growth inhibition effect of Resveratrol incorporated into biodegradable nanoparticles against glioma cells is mediated by the induction of intracellular reactive oxygen species levels. Colloids Surf. B Biointerfaces 2009, 72, 40–47. [Google Scholar] [CrossRef]
  56. Xu, Y.; Asghar, S.; Yang, L.; Li, H.; Wang, Z.; Ping, Q.; Xiao, Y. Lactoferrin-coated polysaccharide nanoparticles based on chitosanhydrochloride/hyaluronic acid/PEG for treating brain glioma. Carbohydr. Polym. 2017, 157, 419–428. [Google Scholar] [CrossRef]
  57. Chu, L.; Wang, A.; Ni, L.; Yan, X.; Song, Y.; Zhao, M.; Sun, K.; Mu, H.; Liu, S.; Wu, Z.; et al. Nose-to-brain delivery of temozolomide-loaded PLGA nanoparticles functionalized with anti-EPHA3 for glioblastoma targeting. Drug Deliv. 2018, 25, 1634–1641. [Google Scholar] [CrossRef]
  58. Gan, C.W.; Feng, S.S. Transferrin-conjugated nanoparticles of Poly(lactide)-D-a-Tocopheryl polyethylene glycol succinate diblock copolymer for targeted drug delivery across the blood-brain barrier. Biomaterials 2010, 31, 7748–7757. [Google Scholar] [CrossRef]
  59. Vijayakumar, M.R.; Kosuru, R.; Singh, S.K.; Prasad, C.B.; Narayan, G.; Muthua, M.S.; Singh, S. Resveratrol loaded PLGA:D-a-tocopheryl polyethylene glycol 1000 succinate blend nanoparticles for brain cancer therapy. RSC Adv. 2016, 6, 74254–74268. [Google Scholar] [CrossRef]
  60. Liang, J.; Gao, C.; Zhu, Y.; Ling, C.; Wang, Q.; Huang, Y.; Qin, J.; Wang, J.; Lu, W.; Wang, J. Natural Brain Penetration Enhancer-Modified Albumin Nanoparticles for Glioma Targeting Delivery. ACS Appl. Mater. Interfaces 2018, 10, 30201–30213. [Google Scholar] [CrossRef]
  61. Ceña, V.; Játiva, P. Nanoparticle crossing of blood–brain barrier: A road to new therapeutic approaches to central nervous system diseases. Nanomedicine 2018, 13, 1513–1516. [Google Scholar] [CrossRef]
  62. Chen, S.; Auriat, A.; Koudrina, A.; DeRosa, M.; Cao, X.; Tsai, E.C. Nano-engineering Nanoparticles for Clinical Use in the Central Nervous System: Clinically Applicable Nanoparticles and Their Potential Uses in the Diagnosis and Treatment of CNS Aliments. In Nanoengineering Materials for Biomedical Uses; Springer: Berlin/Heidelberg, Germany, 2019; pp. 125–145. [Google Scholar]
  63. Yang, S.T.; Guo, W.; Lin, Y.; Deng, X.Y.; Wang, H.F.; Sun, H.F.; Liu, Y.J.; Wang, X.; Wang, W.; Chen, M.; et al. Biodistribution of Pristine Single-Walled Carbon Nanotubes In Vivo. J. Phys. Chem. 2007, 111, 17761–17764. [Google Scholar] [CrossRef]
  64. Lu, W.; Zhang, Y.; Tan, Y.Z.; Hu, K.L.; Jiang, X.G.; Fu, S.K. Cationic albumin conjugated pegylated nanoparticles as novel drug carrier for brain delivery. J. Control. Release 2005, 107, 428–448. [Google Scholar] [CrossRef]
  65. Wu, J.; Wang, C.; Sun, J.; Xue, J. Neurotoxicity of Silica Nanoparticles: Brain Localization and Dopaminergic Neurons Damage Pathways. ACS Nano 2011, 5, 4476–4489. [Google Scholar] [CrossRef]
  66. Koffie, R.M.; Farrar, C.T.; Saidi, L.J.; William, C.M.; Hyman, B.T.; Spires-Jones, T.L. Nanoparticles enhance brain delivery of blood–brain barrier-impermeable probes for in vivo optical and magnetic resonance imaging. Proc. Natl. Acad. Sci. USA 2011, 108, 18837–18842. [Google Scholar] [CrossRef]
  67. Zensi, A.; Begley, D.; Pontikis, C.; Legros, C.; Mihoreanu, L.; Wagner, S.; Büchel, C.; von Briesen, H.; Kreuter, J. Albumin nanoparticles targeted with Apo E enter the CNS by transcytosis and are delivered to neurons. J. Control. Release 2009, 137, 78–86. [Google Scholar] [CrossRef]
  68. Xin, Y.; Liu, T.; Yang, C. Development of PLGA-lipid nanoparticles with covalently conjugated indocyanine green as a versatile nanoplatform for tumor-targeted imaging and drug delivery. Int. J. Nanomed. 2016, 11, 5807–5821. [Google Scholar] [CrossRef]
  69. Sun, W.; Xie, C.; Wang, H.; Hu, Y. Specific role of polysorbate 80 coating on the targeting of nanoparticles to the brain. Biomaterials 2004, 25, 3065–3071. [Google Scholar] [CrossRef]
  70. Bardi, G.; Nunes, A.; Gherardini, L.; Bates, K.; Al-Jamal, K.T.; Gaillard, C.; Kostarelos, K. Functionalized Carbon Nanotubes in the Brain: Cellular Internalization and Neuroinflammatory Responses. PLoS ONE 2013, 8, e80964. [Google Scholar] [CrossRef]
  71. Sarlo, K.; Blackburn, K.L.; Clark, E.D.; Grothaus, J.; Chaney, J.; Neu, S.; Flood, J.; Abbott, D.; Bohne, C.; Casey, K.; et al. Tissue distribution of 20nm, 100 nm and 1000 nm fluorescent polystyrene latex nanospheres following acute systemic or acute and repeat airway exposure in the rat. Toxicology 2009, 263, 117–126. [Google Scholar] [CrossRef]
  72. Kafa, H.; Wang, J.T.W.; Rubio, N.; Klippstein, R.; Costa, P.M.; Hassan, H.A.F.M.; Sosabowski, J.K.; Bansal, S.S.; Preston, J.E.; Abbott, N.J.; et al. Translocation of LRP1 targeted carbon nanotubes of different diameters across the blood–brain barrier in vitro and in vivo. J. Control. Release 2016, 225, 217–229. [Google Scholar] [CrossRef]
  73. Gupta, Y.; Jain, A.; Jain, S.K. Transferrin-conjugated solid lipid nanoparticles for enhanced delivery of quinine dihydrochloride to the brain. J. Pharm. Pharmacol. 2007, 59, 935–940. [Google Scholar] [CrossRef]
  74. Helm, F.; Fricker, G. Liposomal Conjugates for Drug Delivery to the Central Nervous System. Pharmaceutics 2015, 7, 27–42. [Google Scholar] [CrossRef]
  75. Pang, Z.; Gao, H.; Chen, J.; Shen, S.; Zhang, B.; Ren, J.; Guo, L.; Qian, Y.; Jiang, X.; Mei, H. Intracellular delivery mechanism and brain delivery kinetics of biodegradable cationic bovine serum albumin-conjugated polymersomes. Int. J. Nanomed. 2012, 7, 3421–3432. [Google Scholar]
  76. Lozić, I.; Hartz, R.V.; Bartlett, C.A.; Shaw, J.A.; Archer, M.; Naidu, P.S.R.; Smith, N.M.; Dunlop, S.A.; Swaminathan Iyer, K.; Kilburn, M.R.; et al. Enabling dual cellular destinations of polymeric nanoparticles for treatment following partial injury to the central nervous system. Biomaterials 2016, 74, 200–216. [Google Scholar] [CrossRef]
  77. Tahara, K.; Miyazaki, Y.; Kawashima, Y.; Kreuter, J.; Yamamoto, H. Brain targeting with surface-modified poly(D,L-lactic-co-glycolic acid) nanoparticles delivered via carotid artery administration. Eur. J. Pharm. Biopharm. 2011, 77, 84–88. [Google Scholar] [CrossRef]
  78. Calvo, P.; Gouritin, B.; Chacun, H.; Desmaele, D.; D’Angelo, J.; Noel, J.P.; Georgin, D.; Fattal, E.; Andreux, J.P.; Couvreur, P. Long-Circulating PEGylated Polycyanoacrylate Nanoparticles as New Drug Carrier for Brain Delivery. Pharm. Res. 2001, 18, 1157–1166. [Google Scholar] [CrossRef]
  79. Nance, E.; Zhang, C.; Shih, T.Y.; Xu, Q.; Schuster, B.S.; Hanes, J. Brain-Penetrating Nanoparticles Improve Paclitaxel Efficacy in Malignant Glioma Following Local Administration. ACS Nano 2014, 8, 10655–10664. [Google Scholar] [CrossRef]
  80. Inoue, Y.; Ezure, H.; Ito, J.; Sawa, C.; Yamamoto, M.; Hata, H.; Moriyama, H.; Manome, Y.; Otsuka, N. Effect of Silica Nanoparticles on Cultured Central Nervous System Cells. J. Neurosci. 2018, 8, 146–156. [Google Scholar] [CrossRef]
  81. Du, Q.; Ge, D.; Mirshafiee, V.; Chen, C.; Li, M.; Xue, C.; Ma, X.; Sun, B. Assessment of neurotoxicity induced by differentsized Stöber silica nanoparticles: Induction of pyroptosis in microglia. Nanoscale 2019, 11, 12965–12972. [Google Scholar] [CrossRef]
  82. Bittner, A.; Ducray, A.D.; Widmer, H.R.; Stoffel, M.H.; Mevissen, M. Effects of gold and PCL- or PLLA-coated silica nanoparticles on brain endothelial cells and the blood–brain barrier. Beilstein J. Nanotechnol. 2019, 10, 941–954. [Google Scholar] [CrossRef]
  83. Gilardino, A.; Catalano, F.; Ruffinatti, F.A.; Alberto, G.; Nilius, B.; Antoniotti, S.; Martra, G.; Lovisolo, D. Interaction of SiO2 nanoparticles with neuronal cells: Ionic mechanisms involved in the perturbation of calcium homeostasis. Int. J. Biochem. Cell Biol. 2015, 66, 101–111. [Google Scholar] [CrossRef]
  84. Yuan, X.; Yang, Y.; Xia, D.; Meng, L.; He, M.; Liu, C.; Zhang, Z. Silica Nanoparticles Promote α-Synuclein Aggregation and Parkinson’s Disease Pathology. Front. Neurosci. 2022, 15, 807988. [Google Scholar] [CrossRef]
  85. Bang, J.; Yeyeodu, S.; Gilyazova, N.; Witherspoon, S.; Ibeanu, G. Effects of Carbon Nanotubes on a Neuronal Cell Model In Vitro. J. Biol. 2011, 1, 70–77. [Google Scholar] [CrossRef]
  86. Villegas, J.C.; Álvarez-Montes, L.; Rodríguez-Fernández, L.; González, J.; Valiente, R.; Fanarraga, M.L. Multiwalled Carbon Nanotubes Hinder Microglia Function Interfering with Cell Migration and Phagocytosis. Adv. Healthcare Mater. 2014, 3, 424–432. [Google Scholar] [CrossRef]
  87. Li, L.; Zhang, J.; Yang, Y.; Wang, Q.; Gao, L.; Yang, Y.; Chang, T.; Zhang, X.; Xiang, G.; Cao, Y.; et al. Single-wall carbon nanohorns inhibited activation of microglia induced by lipopolysaccharide through blocking of Sirt3. Nanoscale Res. Lett. 2013, 8, 100. [Google Scholar] [CrossRef]
  88. Zheng, W.; McKinney, W.; Kashon, M.; Salmen, R.; Castranova, V.; Kan, H. The influence of inhaled multi-walled carbon nanotubes on the autonomic nervous system. Part. Fibre Toxicol. 2016, 13, 8. [Google Scholar] [CrossRef]
  89. Janaszewska, A.; Ciolkowski, M.; Wróbel, D.; Petersen, J.F.; Ficker, M.; Christensen, J.B.; Bryszewska, M.; Klajnert, B. Modified PAMAM dendrimer with 4-carbomethoxypyrrolidone surface groups reveals negligible toxicity against three rodent cell-lines. Nanomedicine 2013, 9, 461–464. [Google Scholar] [CrossRef]
  90. Serramía, M.J.; Álvarez, S.; Fuentes-Paniagua, E.; Clemente, M.I.; Sánchez-Nieves, J.; Gómez, R.; de la Mata, J.; Muñoz-Fernández, M.A. In vivo delivery of siRNA to the brain by carbosilane dendrimer. J. Control. Release 2015, 200, 60–70. [Google Scholar] [CrossRef]
  91. Gomes, M.J.; Fernandes, C.; Martins, S.; Borges, F.; Sarmento, B. Tailoring Lipid and Polymeric Nanoparticles as siRNA Carriers towards the Blood-Brain Barrier—From Targeting to Safe Administration. J. Neuroimmune Pharmacol. 2017, 12, 107–119. [Google Scholar] [CrossRef]
  92. Mastorakos, P.; Zhang, C.; Berry, S.; Oh, Y.; Lee, S.; Eberhart, C.H.; Woodworth, G.F.; Suk, J.S.; Hanes, J. Highly PEGylated DNA Nanoparticles Provide Uniform and Widespread Gene Transfer in the Brain. Adv. Healthc. Mater. 2015, 4, 1023–1033. [Google Scholar] [CrossRef]
  93. Murali, K.; Kenesei, K.; Li, Y.; Demeter, K.; Környei, Z.; Madarász, E. Uptake and bioreactivity of polystyrene nanoparticles is affected by surface modifications, ageing and LPS adsorption: In vitro studies on neural tissue cells. Nanoscale 2015, 7, 4199–4210. [Google Scholar] [CrossRef]
  94. Jung, B.K.; Han, S.W.; Park, S.H.; Bae, J.S.; Choi, J.; Ryu, K.Y. Neurotoxic potential of polystyrene nanoplastics in primary cells originating from mouse brain. NeuroToxicology 2020, 81, 189–196. [Google Scholar] [CrossRef]
  95. Song, Y.; Du, D.; Li, L.; Xu, J.; Dutta, P.; Lin, Y. In Vitro Study of Receptor-Mediated Silica Nanoparticles Delivery across Blood–Brain Barrier. ACS Appl. Mater. Interfaces 2017, 9, 20410–20416. [Google Scholar] [CrossRef]
  96. Baghirov, H.; Karaman, D.; Viitala, T.; Duchanoy, A.; Lou, Y.R.; Mamaeva, V.; Rosenholm, J.M. Feasibility Study of the Permeability and Uptake of Mesoporous Silica Nanoparticles across the Blood- Brain Barrier. PLoS ONE 2016, 11, e0160705. [Google Scholar] [CrossRef]
  97. Bardi, G.; Tognini, P.; Ciofani, G.; Raffa, V.; Costa, M.; Pizzorusso, T. Pluronic-coated carbon nanotubes do not induce degeneration of cortical neurons in vivo and in vitro. Nanomedicine 2009, 5, 96–104. [Google Scholar] [CrossRef]
  98. Li, N.; Zhang, X.; Song, Q.; Su, R.; Zhang, Q.; Kong, T.; Liu, L.; Jin, G.; Tang, M.; Cheng, G. The promotion of neurite sprouting and outgrowth of mouse hippocampal cells in culture by graphene substrates. Biomaterials 2011, 32, 9374–9382. [Google Scholar] [CrossRef]
  99. Pojo, M.; Cerqueira, S.R.; Mota, T.; Xavier-Magalhães, A.; Ribeiro-Samy, S.; Mano, J.F.; Oliveira, J.M.; Reis, R.L.; Sousa, N.; Costa, B.M.; et al. In vitro evaluation of the cytotoxicity and cellular uptake of CMCht/PAMAM dendrimer nanoparticles by glioblastoma cell models. J. Nanopart. Res. 2013, 15, 1621. [Google Scholar] [CrossRef]
  100. Tiwari, S.K.; Agarwal, S.; Seth, B.; Yadav, A.; Nair, S.; Bhatnagar, P.; Karmakar, M.; Kumari, M.; Chauhan, L.K.S.; Patel, D.K.; et al. Curcumin-Loaded Nanoparticles Potently Induce Adult Neurogenesis and Reverse Cognitive Deficits in Alzheimer’s Disease Model via Canonical Wnt/β-Catenin Pathway. ACS Nano 2014, 8, 76–103. [Google Scholar] [CrossRef]
  101. Li, J.; Feng, L.; Fan, L.; Zha, H.; Guo, L.; Zhang, Q.; Chen, J.; Pang, Z.; Wang, Y.; Jiang, X.; et al. Targeting the brain with PEG-PLGA nanoparticles modified with phage-displayed peptides. Biomaterials 2011, 32, 4943–4950. [Google Scholar] [CrossRef]
  102. Lu, Q.; Cai, X.; Zhang, X.; Li, S.; Song, Y.; Du, D.; Dutta, P.; Lin, Y. Synthetic Polymer Nanoparticles Functionalized with Different Ligands for Receptor-Mediated Transcytosis across the Blood–Brain Barrier. ACS Appl. Bio Mater. 2018, 1, 1687–1694. [Google Scholar] [CrossRef]
  103. Wong, L.R.; Ho, P.C. Role of serum albumin as a nanoparticulate carrier for nose-to-brain delivery of R-flurbiprofen: Implications for the treatment of Alzheimer’s disease. J. Pharm. Pharmacol. 2018, 70, 59–69. [Google Scholar] [CrossRef]
  104. Tian, X.; Wei, F.; Wang, T.; Wang, P.; Lin, X.; Wang, J.; Wang, D.; Ren, L. In vitro and in vivo studies on gelatin-siloxane nanoparticles conjugated with SynB peptide to increase drug delivery to the brain. Int. J. Nanomed. 2012, 7, 1031–1041. [Google Scholar]
  105. Cha, K.E.; Heejoon, M. Cytotoxic Effects of Nanoparticles Assessed In Vitro and In Vivo. J. Microbiol. Biotechnol. 2007, 17, 1573–1578. [Google Scholar]
  106. He, H.; Li, Y.; Jia, X.R.; Du, J.; Ying, X.; Lu, W.L.; Lou, J.N.; Wei, Y. PEGylated Poly(amidoamine) dendrimer-based dual-targeting carrier for treating brain tumors. Biomaterials 2011, 32, 478–487. [Google Scholar] [CrossRef]
  107. Sharma, A.; Sharma, R.; Zhang, Z.; Liaw, K.; Kambhampati, S.P.; Porterfield, J.E.; Lin, K.C.; DeRidder, L.B.; Kannan, S.; Kannan, R.M. Dense hydroxyl polyethylene glycol dendrimer targets activated glia in multiple CNS disorders. Sci. Adv. 2020, 6, eaay8514. [Google Scholar] [CrossRef]
  108. Ganesan, P.; Kim, B.; Ramalaingam, P.; Karthivashan, G.; Revuri, V.; Park, S.; Kim, J.S.; Ko, Y.T.; Choi, D.K. Antineuroinflammatory Activities and Neurotoxicological Assessment of Curcumin Loaded Solid Lipid Nanoparticles on LPS-Stimulated BV-2 Microglia Cell Models. Molecules 2019, 24, 1170. [Google Scholar] [CrossRef]
  109. Wang, S.; Li, Y.; Fan, J.; Wang, Z.; Zeng, X.; Sun, Y.; Song, P.; Ju, D. The role of autophagy in the neurotoxicity of cationic PAMAM Dendrimers. Biomaterials 2014, 35, 7588–7597. [Google Scholar] [CrossRef]
  110. Zhao, D.; Alizadeh, D.; Zhang, L.; Liu, W.; Farrukh, O.; Manuel, E.; Diamond, D.J.; Badie, B. Carbon Nanotubes Enhance CpG Uptake and Potentiate Anti-Glioma Immunity. Clin. Cancer Res. 2011, 17, 771–782. [Google Scholar] [CrossRef]
  111. Ma, X.; Zhong, L.; Guo, H.; Wang, Y.; Gong, N.; Wang, Y.; Cai, J.; Liang, X.J. Multiwalled Carbon Nanotubes Induced Hypotension by Regulating the Central Nervous System. Adv. Funct. Mater. 2018, 28, 1705479. [Google Scholar] [CrossRef]
  112. Vittorio, O.; Raffa, V.; Cuschieri, A. Influence of purity and surface oxidation on cytotoxicity of multiwalled carbon nanotubes with human neuroblastoma cells. Nanomedicine 2009, 5, 424–431. [Google Scholar] [CrossRef]
  113. Zhang, Y.; Ali, S.F.; Dervishi, E.; Xu, Y.; Li, Z.; Casciano, D.; Biris, A.S. Cytotoxicity Effects of Graphene and Single-Wall Carbon Nanotubes in Neural Phaeochromocytoma-Derived PC12 Cells. ACS Nano 2010, 4, 3181–3186. [Google Scholar] [CrossRef]
  114. Zhang, Y.; Xu, Y.; Li, Z.; Chen, T.; Lantz, S.M.; Howard, P.C.; Paule, M.G.; Slikker, W.; Watanabe, F.; Mustafa, T.; et al. Mechanistic Toxicity Evaluation of Uncoated and PEGylated Single-Walled Carbon Nanotubes in Neuronal PC12 Cells. ACS Nano 2011, 5, 7020–7033. [Google Scholar] [CrossRef]
  115. Hong, S.W.; Lee, J.; Kang, S.H.; Hwang, E.J.; Hwang, Y.S.; Lee, M.H.; Han, D.W.; Park, J.C. Enhanced Neural Cell Adhesion and Neurite Outgrowth on Graphene-Based Biomimetic Substrates. BioMed Res. Int. 2014, 2014, 212149. [Google Scholar] [CrossRef]
  116. Athira, S.S.; Biby, T.E.; Mohanan, P.V. Effect of polymer functionalized fullerene soot on C6 glial cells. Eur. Polym. J. 2020, 127, 109572. [Google Scholar] [CrossRef]
  117. Waite, C.L.; Sparks, S.M.; Uhrich, K.E.; Roth, C.M. Acetylation of PAMAM dendrimers for cellular delivery of siRNA. BMC Biotechnol. 2009, 9, 38. [Google Scholar] [CrossRef]
  118. Janiszewska, J.; Posadas, I.; Játiva, P.; Bugaj- Zarebska, M.; Urbanczyk-Lipkowska, Z.; Ceña, V. Second Generation Amphiphilic Poly-Lysine Dendrons Inhibit Glioblastoma Cell Proliferation without Toxicity for Neurons or Astrocytes. PLoS ONE 2016, 11, e0165704. [Google Scholar] [CrossRef]
  119. Vijayakumar, M.R.; Kumari, L.; Patel, K.K.; Vuddanda, P.R.; Vajanthri, K.Y.; Mahtoc, S.K.; Singh, S. Intravenous administration of trans-resveratrol loaded TPGS-coated solid lipid nanoparticles for prolonged systemic circulation, passive brain targeting and improved in vitro cytotoxicity against C6 glioma cell lines. RSC Adv. 2016, 6, 50336–50348. [Google Scholar] [CrossRef]
  120. Liu, J.L.; Li, J.; Zhang, L.Y.; Zhang, P.L.; Zhou, J.L.; Liu, B. Preparation of N, N, N-trimethyl chitosan-functionalized retinoic acid-loaded lipid nanoparticles for enhanced drug delivery to glioblastoma. Trop. J. Pharm. Res. 2017, 16, 1765–1772. [Google Scholar] [CrossRef]
  121. Jhaveri, A.; Deshpande, P.; Pattni, B.; Torchilin, V. Transferrin-targeted, resveratrol-loaded liposomes for the treatment of glioblastoma. J. Control Release. 2018, 10, 89–101. [Google Scholar] [CrossRef]
  122. Delalat, B.; Sheppard, V.C.; Ghaemi, S.R.; Rao, S.; Prestidge, C.A.; McPhee, G.; Rogers, M.L.; Donoghue, J.F.; Pillay, V.; Johns, T.G.; et al. Targeted drug delivery using genetically engineered diatom biosilica. Nat. Commun. 2015, 6, 8791. [Google Scholar] [CrossRef]
  123. Wang, Z.H.; Wang, Z.Y.; Sun, C.S.; Wang, C.Y.; Jiang, T.Y.; Wang, S.L. Trimethylated chitosan-conjugated PLGA nanoparticles for the delivery of drugs to the brain. Biomaterials 2010, 31, 908–915. [Google Scholar] [CrossRef]
  124. Guo, J.; Gao, X.; Su, L.; Xia, H.; Gu, G.; Pang, Z.; Jiang, X.; Yao, L.; Chen, J.; Chen, H. Aptamer-functionalized PEG-PLGA nanoparticles for enhanced anti-glioma drug delivery. Biomaterials 2011, 32, 8010–8020. [Google Scholar] [CrossRef]
  125. Lim, K.J.; Bisht, S.; Bar, E.E.; Maitra, A.; Eberhart, C.G. A polymeric nanoparticle formulation of curcumin inhibits growth, clonogenicity and stem-like fraction in malignant brain tumors. Cancer Biol. Ther. 2011, 11, 464–473. [Google Scholar] [CrossRef]
  126. Tammam, S.N.; Azzazy, H.M.E.; Lamprecht, A. Nuclear and cytoplasmic delivery of lactoferrin in glioma using chitosan nanoparticles: Cellular location dependent-action of lactoferrin. Eur. J. Pharm. Biopharm. 2018, 129, 74–79. [Google Scholar] [CrossRef]
  127. Nejat, H.; Rabiee, M.; Varshochian, R.; Tahriri, M.; Jazayeri, H.E.; Rajadas, J.; Ye, H.; Cui, Z.; Tayebi, L. Preparation and characterization of cardamom extract-loaded gelatin nanoparticles as effective targeted drug delivery system to treat glioblastoma. React. Funct. Polym. 2017, 120, 46–56. [Google Scholar] [CrossRef]
  128. Xu, M.; Zhou, H.; Liu, Y.; Sun, J.; Xie, W.; Zhao, P.; Liu, J. Ultrasound-Excited Protoporphyrin IX-Modified Multifunctional Nanoparticles as a Strong Inhibitor of Tau Phosphorylation and β-Amyloid Aggregation. ACS Appl. Mater. Interfaces 2018, 10, 32965–32980. [Google Scholar] [CrossRef]
  129. Patel, D.A.; Henry, J.E.; Good, T.A. Attenuation of β-amyloid induced toxicity by sialic acidconjugated dendrimers: Role of sialic acid attachment. Brain Res. 2007, 3, 95–105. [Google Scholar] [CrossRef]
  130. Kuo, Y.C.; Tsai, H.C. Rosmarinic acid- and curcumin-loaded polyacrylamide-cardiolipin-poly (lactide-co-glycolide) nanoparticles with conjugated 83-14 monoclonal antibody to protect β-amyloid-insulted neurons. Mater. Sci. Eng. 2018, 91, 445–457. [Google Scholar] [CrossRef]
  131. Saraiva, A.M.; Cardoso, I.; Saraiva, M.J.; Tauer, K.; Pereira, M.K.; Coelho, M.A.N.; Möhwald, H.; Brezesinski, G. Randomization of Amyloid-b-Peptide(1-42)Conformation by Sulfonated and SulfatedNanoparticles Reduces Aggregation and Cytotoxicity. Macromol. Biosci. 2010, 10, 1152–1163. [Google Scholar] [CrossRef]
  132. Shyam, R.; Ren, Y.; Lee, J.; Braunstein, K.E.; Mao, H.Q.; Wong, P.C. Intraventricular Delivery of siRNA Nanoparticles to the Central Nervous System. Mol. Ther. Nucleic Acids. 2015, 4, e242. [Google Scholar] [CrossRef]
  133. Rakotoarisoa, M.; Angelov, B.; Garamus, V.M.; Angelova, A. Curcumin- and Fish Oil-Loaded Spongosome and Cubosome Nanoparticles with Neuroprotective Potential against H2O2-Induced Oxidative Stress in Differentiated Human SH-SY5Y Cells. ACS Omega 2019, 4, 3061–3073. [Google Scholar] [CrossRef]
  134. Jha, A.; Ghormade, V.; Kolge, H.; Paknikar, K.M. Dual effect of chitosan-based nanoparticles on the inhibition of b-amyloid peptide aggregation and disintegration of the preformed fibrils. J. Mater. Chem. B 2019, 7, 3362–3373. [Google Scholar] [CrossRef]
  135. Sookhaklari, R.; Geramizadeh, B.; Abkar, M.; Moosavi, M. The neuroprotective effect of BSA-based nanocurcumin against 6-OHDA-induced cell death in SH-SY5Y cells. Avicenna J. Phytomed. 2019, 9, 92–100. [Google Scholar]
  136. Mulik, R.S.; Mönkkönen, J.; Juvonen, R.O.; Mahadik, K.R.; Paradkar, A.R. ApoE3 Mediated Poly(butyl) Cyanoacrylate Nanoparticles Containing Curcumin: Study of Enhanced Activity of Curcumin against Beta Amyloid Induced Cytotoxicity Using In Vitro Cell Culture Model. Mol. Pharm. 2010, 7, 815–825. [Google Scholar] [CrossRef]
  137. You, R.; Ho, Y.S.; Hung, C.H.L.; Liu, Y.; Huang, C.X.; Chan, H.N.; Ho, S.L.; Lui, S.Y.; Li, H.W.; Chang, R.C.C. Silica nanoparticles induce neurodegeneration-like changes in behavior, neuropathology, and affect synapse through MAPK activation. Part. Fibre Toxicol. 2018, 15, 28. [Google Scholar] [CrossRef]
  138. Çömelekoğlu, Ü.; Ball, E.; Yalın, S.; Eroğlu, P.; Bayrak, G.; Yaman, S.; Söğüt, F. Effects of different sizes silica nanoparticle on the liver, kidney and brain in rats: Biochemical and histopathological evaluation. J. Res. Pharm. 2019, 23, 344–353. [Google Scholar] [CrossRef]
  139. Wei, W.; Yan, Z.; Liu, X.; Qin, Z.; Tao, X.; Zhu, X.; Song, E.; Chen, C.; Chun Ke, P.; Leong, D.T.; et al. Brain Accumulation and Toxicity Profiles of Silica Nanoparticles: The Influence of Size and Exposure Route. Environ. Sci. Technol. 2022, 56, 8319–8325. [Google Scholar] [CrossRef]
  140. Maurya, M.K.; Langeh, R.; Anuradha, J.; Sanjeevani, R.; Sanjeevi, R.; Tripathi, S.; Chauhan, D.S. Silica Nanoparticles Induced Oxidative Stress in Different Brain Regions of Male Albino Rats. Sch. Acad. J. Biosci. 2021, 9, 139–144. [Google Scholar] [CrossRef]
  141. Aragon, M.J.; Topper, L.; Tyler, C.R.; Sanchez, B.; Zychowski, K.; Young, T.; Herbert, G.; Hall, P.; Erdely, A.; Eye, T.; et al. Serum-borne bioactivity caused by pulmonary multiwalled carbon nanotubes induces neuroinflammation via blood–brain barrier impairment. Proc. Natl. Acad. Sci. USA 2017, 21, E1968–E1976. [Google Scholar] [CrossRef]
  142. Dal Bosco, L.; Weber, G.E.B.; Parfitt, G.M.; Paese, K.; Gonçalves, C.O.F.; Serodre, T.M.; Furtado, C.A.; Santos, A.P.; Monserrat, J.M.; Barros, D.M. PEGylated Carbon Nanotubes Impair Retrieval of Contextual Fear Memory and Alter Oxidative Stress Parameters in the Rat Hippocampus. BioMed Res. Int. 2015, 2015, 104135. [Google Scholar] [CrossRef]
  143. Yuan, Z.Y.; Hu, Y.L.; Gao, J.Q. Brain Localization and Neurotoxicity Evaluation of Polysorbate 80-Modified Chitosan Nanoparticles in Rats. PLoS ONE 2015, 10, e0134722. [Google Scholar] [CrossRef]
  144. Olivier, J.C.; Fenart, L.; Chauvet, R.; Pariat, C.; Cecchelli, R.; Couet, W. Indirect evidence that drug-brain targeting using Polysorbate 80-coated Polybutylcyanoacrylate nanoparticles is related to toxicity. Pharm. Res. 1999, 16, 1836–1842. [Google Scholar] [CrossRef]
  145. Liu, X.; Zhao, Y.; Dou, J.; Hou, Q.; Cheng, J.; Jiang, X. Bioeffects of Inhaled Nanoplastics on Neurons and Alteration of Animal Behaviors through Deposition in the Brain. Nano Lett. 2022, 22, 1091–1099. [Google Scholar] [CrossRef]
  146. Van Handel, M.; Alizadeh, D.; Zhang, L.; Kateb, B.; Bronikowski, M.; Manohara, H.; Badie, B. Selective uptake of multi-walled carbon nanotubes by tumor macrophages in a murine glioma model. J. Neuroimmunol. 2009, 208, 3–9. [Google Scholar] [CrossRef]
  147. Qu, G.; Bai, Y.; Zhang, Y.; Jia, Q.; Zhang, W.; Yan, B. The effect of multiwalled carbon nanotube agglomeration on their accumulation in and damage to organs in mice. Carbon 2009, 47, 2060–2069. [Google Scholar] [CrossRef]
  148. Yamago, S.; Tokuyama, H.; Nakamuralr, E.; Kikuchi, K.; Kananishl, S.; Sueki, K.; Nakahara, H.; Enomoto, S.; Ambe, F. In vivo biological behavior of a water-miscible fullerene:14C labeling, absorption, distribution, excretion and acute toxicity. Chem. Biol. 1995, 2, 385–389. [Google Scholar] [CrossRef]
  149. Shytikov, D.; Shytikova, I.; Rohila, D.; Kulaga, A.; Dubiley, T.; Pishel, I. Effect of Long-Term Treatment with C60 Fullerenes on the Lifespan and Health Status of CBA/Ca Mice. Rejuv. Res. 2021, 24, 345–353. [Google Scholar] [CrossRef]
  150. Yuan, Y.; Chen, Y.; Liu, J.H.; Wang, H.; Liu, Y. Biodistribution and fate of nanodiamonds in vivo. Diam. Relat. Mater. 2009, 18, 95–100. [Google Scholar] [CrossRef]
  151. Pakhira, B.; Ghosh, M.; Allam, A.; Sarkar, S. Carbon nano onions cross the blood brain barrier. RSC Adv. 2016, 6, 29779–29782. [Google Scholar] [CrossRef]
  152. Huang, R.; Ke, W.; Liu, Y.; Jiang, C.; Pei, Y. The use of lactoferrin as a ligand for targeting the polyamidoamine-based gene delivery system to the brain. Biomaterials 2008, 29, 238–246. [Google Scholar] [CrossRef]
  153. Mishra, M.K.; Beaty, C.A.; Lesniak, W.G.; Kambhampati, S.P.; Zhang, F.; Wilson, M.A.; Blue, M.E.; Troncoso, J.C.; Kannan, S.; Johnston, M.V.; et al. Dendrimer Brain Uptake and Targeted Therapy for Brain Injury in a Large Animal Model of Hypothermic Circulatory Arrest. ACS Nano 2014, 8, 2134–2147. [Google Scholar] [CrossRef]
  154. MacKay, J.A.; Deen, D.F.; Szoka, F.C., Jr. Distribution in brain of liposomes after convection enhanced delivery; modulation by particle charge, particle diameter, and presence of steric coating. Brain Res. 2005, 1035, 139–153. [Google Scholar] [CrossRef]
  155. Lee, S.; Ashizawa, A.T.; Kim, K.S.; Falk, D.J.; Notterpek, L. Liposomes to Target Peripheral Neurons and Schwann Cells. PLoS ONE 2013, 8, e78724. [Google Scholar] [CrossRef]
  156. Salvalaio, M.; Rigon, L.; Belletti, D.; D’Avanzo, F.; Pederzoli, F.; Ruozi, B.; Tosi, G. Targeted Polymeric Nanoparticles for Brain Delivery of High Molecular Weight Molecules in Lysosomal Storage Disorders. PLoS ONE 2016, 11, e0156452. [Google Scholar] [CrossRef]
  157. Tosi, G.; Bondioli, L.; Ruozi, B.; Badiali, L.; Severini, G.M.; Biffi, S.; De Vita, A.; Bortot, B.; Dolcetta, D.; Forni, F.; et al. NIR-labeled nanoparticles engineered for brain targeting: In vivo optical imaging application and fluorescent microscopy evidences. J. Neural Transm. 2011, 118, 145–153. [Google Scholar] [CrossRef]
  158. Orunoğlu, M.; Kaffashi, A.; Pehlivan, S.B.; Şahin, S.; Söylemezoğlu, F.; Oğuz, K.K.; Mut, M. Effects of curcumin-loaded PLGA nanoparticles on the RG2 rat glioma model. Mater. Sci. Eng. 2017, 78, 32–38. [Google Scholar] [CrossRef]
  159. Lu, W.; Wan, J.; She, Z.; Jiang, X. Brain delivery property and accelerated blood clearance of cationic albumin conjugated pegylated nanoparticle. J. Control. Release 2007, 118, 38–53. [Google Scholar] [CrossRef]
  160. Bergonzi, M.C.; Guccione, C.; Grossi, C.; Piazzini, V.; Torracchi, A.; Luccarini, I.; Casamenti, F.; Bilia, A.R. Albumin Nanoparticles for Brain Delivery: A Comparison of Chemical versus Thermal Methods and in vivo Behavior. ChemMedChem 2016, 11, 1840–1849. [Google Scholar] [CrossRef]
  161. Alam, S.; Khan, Z.I.; Mustafa, M.; Kumar, M.; Islam, F.; Bhatnagar, A.; Ahmad, F.J. Development and evaluation of thymoquinone-encapsulated chitosan nanoparticles for nose-to-brain targeting: A pharmacoscintigraphic study. Int. J. Nanomed. 2012, 7, 5705–5718. [Google Scholar] [CrossRef]
  162. Sun, M.; Gao, Y.; Guo, C.; Cao, F.; Song, Z.; Xi, Y.; Yu, A.; Li, A.; Zhai, G. Enhancement of transport of curcumin to brain in mice by poly(n-butylcyanoacrylate) nanoparticle. J. Nanopart. Res. 2010, 12, 3111–3122. [Google Scholar] [CrossRef]
  163. Kanakia, S.; Toussaint, J.; Chowdhury, S.M.; Tembulkar, T.; Lee, S.; Jiang, Y.P.; Lin, R.Z.; Shroyer, K.R.; Moore, W.; Sitharaman, B. Dose Ranging, Expanded Acute Toxicity and Safety Pharmacology Studies for Intravenously Administered Functionalized Graphene Nanoparticle Formulations. Biomaterials 2014, 35, 7022–7031. [Google Scholar] [CrossRef]
  164. Xia, L.; Tan, S.; Zhou, Y.; Lin, J.; Wang, H.; Oyang, L.; Tian, Y.; Liu, L.; Su, M.; Wang, H.; et al. Role of the NFκB-signaling pathway in cancer. OncoTargets Ther. 2018, 11, 2063–2073. [Google Scholar] [CrossRef]
  165. Grodzicki, W.; Dziendzikowska, K.; Gromadzka-Ostrowska, J.; Kruszewski, M. Nanoplastic Impact on the Gut-Brain Axis: Current Knowledge and Future Directions. Int. J. Mol. Sci. 2021, 22, 12795. [Google Scholar] [CrossRef]
  166. Kashiwada, S. Distribution of nanoparticles in the see-through medaka (Oryzias latipes). Environ. Health Perspect. 2006, 114, 1697–1702. [Google Scholar] [CrossRef]
  167. Mattsson, K.; Johnson, E.V.; Malmendal, A.; Linse, S.; Hansson, L.A.; Cedervall, T. Brain damage and behavioural disorders in fish induced by plastic nanoparticles delivered through the food chain. Sci. Rep. 2017, 7, 11452. [Google Scholar] [CrossRef]
  168. Hamed, M.; Martyniuk, C.J.; Naguib, M.; Lee, J.S.; Sayed, A.E.H. Neurotoxic effects of different sizes of plastics (nano, micro, and macro) on juvenile common carp (Cyprinus carpio). Front. Mol. Neurosci. 2022, 15, 1028364. [Google Scholar] [CrossRef]
  169. Brandts, I.; Teles, M.; Gonçalves, A.P.; Barreto, A.; Franco-Martinez, L.; Tvarijonaviciute, A.; Oliveira, M. Effects of nanoplastics on Mytilus galloprovincialis after individual and combined exposure with carbamazepine. Sci. Total Environ. 2018, 643, 775–784. [Google Scholar] [CrossRef]
  170. Varó, I.; Perini, A.; Torreblanca, A.; Garcia, Y.; Bergami, E.; Vannuccini, M.L.; Corsi, I. Time-dependent effects of polystyrene nanoparticles in brine shrimp Artemia franciscana at physiological, biochemical and molecular levels. Sci. Total Environ. 2019, 675, 570–580. [Google Scholar] [CrossRef]
  171. Oberdörster, E. Manufactured Nanomaterials (Fullerenes, C60) Induce Oxidative Stress in the Brain of Juvenile Largemouth Bass. Environ. Health Perspect. 2004, 112, 1058–1062. [Google Scholar] [CrossRef]
  172. Li, J.; Ying, G.G.; Jones, K.C.; Martin, F.L. Real-world carbon nanoparticle exposures induce brain and gonadal alterations in zebrafish (Danio rerio) as determined by biospectroscopy techniques. Analyst 2015, 140, 2687–2695. [Google Scholar] [CrossRef]
  173. Duan, J.; Yu, Y.; Shi, H.; Tian, L.; Guo, C.; Huang, P.; Sun, Z. Toxic Effects of Silica Nanoparticles on Zebrafish Embryos and Larvae. PLoS ONE 2013, 8, e74606. [Google Scholar] [CrossRef]
  174. Pedersen, A.F.; Meyer, D.N.; Petriv, A.M.V.; Soto, A.L.; Shields, J.N.; Akemann, C.; Baker, B.B.; Tsou, W.L.; Zhang, Y.; Baker, T.R. Nanoplastics impact the zebrafish (Danio rerio) transcriptome: Associated developmental and neurobehavioral consequences. Environ. Pollut. 2020, 266, 115090. [Google Scholar] [CrossRef]
  175. Shang, Y.; Wang, S.; Jin, Y.; Xue, W.; Zhong, Y.; Wang, H.; An, J.; Li, H. Polystyrene nanoparticles induced neurodevelopmental toxicity in Caenorhabditis elegans through regulation of dpy-5 and rol-6. Ecotoxicol. Environ. Saf. 2021, 222, 112523. [Google Scholar] [CrossRef]
  176. Chen, Q.; Gundlach, M.; Yang, S.; Jiang, J.; Velki, M.; Yin, D.; Hollert, H. Quantitative investigation of the mechanisms of microplastics and nanoplastics toward zebrafish larvae locomotor activity. Sci. Total Environ. 2017, 584–585, 1022–1031. [Google Scholar] [CrossRef]
  177. Pitt, J.A.; Kozal, J.S.; Jayasundara, N.; Massarsky, A.; Trevisan, R.; Geitner, N.; Wiesner, M.; Levin, E.D.; Di Giulio, R.T. Uptake, tissue distribution, and toxicity of polystyrene nanoparticles in developing zebrafish (Danio rerio). Aquat. Toxicol. 2018, 194, 185–194. [Google Scholar] [CrossRef]
  178. Sarasamma, S.; Audira, G.; Siregar, P.; Malhotra, N.; Lai, Y.H.; Liang, S.T.; Chen, J.R.; Chen, K.H.C.; Hsiao, C.D. Nanoplastics Cause Neurobehavioral Impairments, Reproductive and Oxidative Damages, and Biomarker Responses in Zebrafish: Throwing up Alarms of Wide Spread Health Risk of Exposure. Int. J. Mol. Sci. 2020, 21, 1410. [Google Scholar] [CrossRef]
  179. Sokmen, T.O.; Sulukan, E.; Turkoğlu, M.; Baran, A.; Ozkaraca, M.; Ceyhun, S.B. Polystyrene nanoplastics (20 nm) are able to bioaccumulate and cause oxidative DNA damages in the brain tissue of zebrafish embryo (Danio rerio). NeuroToxicology 2020, 77, 51. [Google Scholar] [CrossRef]
  180. Qu, M.; Kong, Y.; Yuan, Y.; Wang, D. Neuronal damage induced by nanopolystyrene particles in nematode Caenorhabditis elegans. Environ. Sci. Nano 2019, 6, 2591. [Google Scholar] [CrossRef]
  181. Lei, L.; Liu, M.; Song, Y.; Lu, S.; Hu, J.; Cao, C.; Xie, B.; Shi, H.; Defu, H. Polystyrene (nano)microplastics cause sizedependent neurotoxicity, oxidative damage and other adverse effects in Caenorhabditis elegans. Environ. Sci. Nano 2018, 5, 2009. [Google Scholar] [CrossRef]
  182. Ding, J.; Zhang, S.; Razanajatovo, R.M.; Zou, H.; Zhu, W. Accumulation, tissue distribution, and biochemical effects of polystyrene microplastics in the freshwater fish red tilapia (Oreochromis niloticus). Environ. Pollut. 2018, 238, 1–9. [Google Scholar] [CrossRef]
  183. Chen, Q.; Yin, D.; Jia, Y.; Schiwy, S.; Legradi, J.; Yang, S.; Hollert, H. Enhanced uptake of BPA in the presence of nanoplastics can lead to neurotoxic effects in adult zebrafish. Sci. Total Environ. 2017, 609, 1312–1321. [Google Scholar] [CrossRef]
  184. European Commission; Directorate-General for Environment. Nanoplastics: State of Knowledge and Environmental and Human Health Impacts; Publications Office of the European Union: Luxembourg, 2023. [Google Scholar] [CrossRef]
Figure 1. The scheme of direct and indirect action of nmNPs in CNS.
Figure 1. The scheme of direct and indirect action of nmNPs in CNS.
Materials 16 07264 g001
Figure 2. Types of the nmNPs discussed in the present review.
Figure 2. Types of the nmNPs discussed in the present review.
Materials 16 07264 g002
Figure 3. Effects of nmNPs in CNS.
Figure 3. Effects of nmNPs in CNS.
Materials 16 07264 g003
Table 1. The effect of nmNPs (toxic or non-toxic) on CNS-derived normal cells in vitro.
Table 1. The effect of nmNPs (toxic or non-toxic) on CNS-derived normal cells in vitro.
Toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
SILICA NPs
None30 nm,
0.01, 0.1 and
1 mg/mL,
10 min, 30 min and 1 h
primary cultures of cortical neurons isolated from ratsDecreased neuron dendrite and induction of ROS production. Early stage apoptosis process and necrosis.[80]
None50, 100, 300 nm,
25–200 μg/mL,
24 h
microglia (N9 and BV-2), endothelial cells (bEnd.3), neuronal cells (HT22 cells)Size- and cell type dependent cytotoxicity. Increased ROS production in microglia, decreased GSH level, lysosomal damage. Induction of release of IL-1β and N-terminal GSDMD, a marker of pyroptosis.[81]
ICG/PCL-
(PCL-SiNPs) or ICG/PLLA-coated SiNPs (PLLA-SiNPs)
90 nm (PCL-SiNPs)
95 nm (PLLA-SiNPs),
2.49 × 10−7 μg/mL to
24.9 × 10−7μg/mL,
2 and 24 h
brain endothelial cells and rat brain capillary endothelial cells (rBCEC4)Time- and concentration-dependent decrease in cell viability, but not proliferation, differentiation. No inflammation process.[82]
None150–200 nm,
4 × 1010 NPs/mL to 7 × 1010 NPs/mL,
24 h
primary rat microgliaIncreased production of ROS and RNS. Decrease in TNFα gene expression and cyclooxygenase-2 gene. Increased IL-1β cytokine release.[23]
None47; 49; 55; 500; 2000 nm,
20 µg/mL,
30 min and 4 h
neuronal cell lines (GT1–7 and GN11 cells)Long lasting but reversible calcium signaling, independ on NP internalization.[83]
None509.3, 356.0, 469.0 nm,
100 and 200 µg/mL,
24 and 48 h
7 days
HEK293 cells and
primary mouse cortical neurons
α-Synucleinopathy in HEK293 cells and neurons. [84]
CARBON NPs
Carboxylated
and aminated MWNTs
200–300 nm,
5–100 μg/mL,
24 h
primary neuronal and glial cells isolated from fetal rat frontal cortex and striatumDecreased microglia viability at concentration of 20 μg/mL or higher. Release of NO.[19]
MWNTs85–115 nm,
190 and 295 ppm,
24 and 48 h
NeuroScreen-1 (NS-1) cellsReduced viability, apoptosis. Mitochondria depolarization and disruption of membranes.[85]
MWNTs5 to 15 nm,
0.6 μg/mL,
70 h
microglial cells (BV-2)Dose-dependent cell division arrest and apoptosis. Perturbation of cell migration and phagocytosis.[86]
Single-wall carbon nanohorns80 to 100 nm,
16 h
microglial cells (N9 and BV-2)Proliferation inhibition, promotion of apoptosis. Perturbation of cell cycle.[87]
DENDRIMERS
G4 and G4-C12 PAMAM1 to 100 nM,
24 h
primary neuronal culturesApoptosis and cell death.[25]
PAMAM-NH2 G4
(PAMAM-NH2) or (PAMAM-NH2) modified with sodium carboxylate (PAMAM-S.C.)
26.1, 38.2, 104.8 nm
(PAMAM-NH2)
45.3, 173.4, 267.7 nm
(PAMAM-SC),
0.3, 1, 3, and 10 µg/mL,
72 h
human neural progenitor cells cultured as a 3D neurosphere modelInhibition of neurosphere growth, cell proliferation and neuronal migration. Down-regulation of early growth response gene 1, insulin-like growth factor-binding protein 3 and tissue factor pathway inhibitor (TFPI2).[88]
PAMAM G4, bare or modified with 4-carbomethoxypyrrolidone surface groups10, 40, 80, 120, 200 µM,
24 h
embryonic mouse hippocampal cells (mHippoE-18)Minor toxicity. No effect on ROS production nor mitochondrial membrane potential.[89]
Cationic carbosilane dendrimer G2 loaded with FITC100 nM,
24 and 48 h
human primary astrocytesCell death.[90]
None0.1 mg/mL,
30 min
neurons and glial cells from ratsIncrease in intratracellular Ca2+ flux and mitochonria depolarization and impaired oxidative metabolism in pyramidal neurons and astrocytes.[39]
SOLID LIPID NPs
PLGA NPs with a peptide-binding transferrin receptor, loaded with siRNA115 and 150 nm,
0.1; 1; 10; 100; 1000 µg/mL,
24 h
immortalized human cerebral microvascular endothelial cell line (hCMEC/D3 cell line)Decreased metabolic activity of cells at the highest concentration.[91]
POLY(ETHYLENEGLYCOL)–POLY(LACTIDE) NPs
PLGA NPs coated with BSA80 nm,
0.025 to 8.0 mg/mL,
4 h
capillary endothelial cell (BCEC)Concentration dependent cytotoxicity.[64]
DNA NPs
DNA NPs modified with PEG<60 nm,
1, 5, 10 μg/mL,
24, 48 and 72 h
rabbit and rat primary astrocytes Toxicity.[92]
NANOGEL
PEG or PEG-PEI nanogel<100 nm,
0.001; 0.01; 0.1; 1; 10 mg/mL,
2 h
bovine brain microvessel endothelial cells (BBMEC)Neglectable cytotoxicity.[43]
NANOPLASTIC
Carboxylated polystyrene NPs55 nm,
7.8–250 mg/L,
24 h
NE-4C neuroectodermal stem cells; primary brain cell cultures from mouse; microglia; brain vascular endothelial cell culturesLDH leakage of neuronal cells. Microglial cells were able to internalize carboxylated polystyrene nanoparticles by phagocytosis.[93]
Polystyrene NPs100 nm,
100–200 μg/mL
GES-1,
different primary brain cells
Reduced cell viability and defective neuronal development, reactive astrocytosis in astrocytes[94]
Non-toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
SILICA NPs
ICG/PCL-rhodaminedoped SiNPs (PCL-NPs) and ICG/PLLA-rhodamine-doped SiNPs (PLLA-NPs)90 nm (PCL-NPs)
95 nm (PLLA-NPs),
2.6 × 109, 5.2 × 109 and 2.6 × 1010 (PCL-NPs/mL)
[2.9 × 109, 5.8 × 109 and 2.9 × 1010 (PLLA-NPs/mL),
6 and 24 h
rat primary hippocampal cultureNo change in the release of IL-1β and TNFα.[24]
PEG and lactoferrin26; 53 and 105 nm,
1 to 10 nM,
24 h
in vitro blood−brain barrier (BBB) model consisting of three distinct types of cells: endocytes, pericytes, and astrocytesLactoferrin enhanced efficiency of NPs transport across the BBB. The cell viability at a level over 93%.[95]
Mesoporous silica NPs bare or coated with a PEG-PEI block copolymer 50 to 240 nm,
10, 20, 50 μg/mL,
36 h
rat brain endothelial cells (RBE4)No toxicity up to 50 μg/mL. No damage to BBB.[96]
CARBON NPs
Carboxylated
and aminated
MWNTs
20–30 nm,
5–100 μg/mL,
24 h
primary neuronal and glial cell populations isolated from fetal rat
frontal cortex and striatum
Neurons from both brain regions were generally not affected by exposure to NPs. The viability of mixed glia was not reduced in frontal cortex.[19]
Prestine and pluronic F127 coated MWNTs20–30 nm,
3.5, 17.5 μg/mL,
24 h
neurons and gliaNo toxicity and no apoptosis after co-treatment with Pluronic F127 and MWNTs. No toxicity after treatment with prestine MWNTs.[97]
Amino-functionalized MWNTs18.9, 20 and 50 µg/mL,
24 and 72 h
primary porcine
brain endothelial cells (PBEC) and primary rat astrocytes
No statistically significant toxicity.[32]
MWNTs modified with plasmid DNA20 nm,
2.5 μg of MWNTs,
15, 24 and 48 h
microglial cells (BV-2)No detrimental effect on proliferation or cytokine production.[21]
Acid-oxidized MWNTs5, 7 and 8 µm,
1–10 µg/mL,
24 h
microglial cells (N9)Confirmed NP internalization without any effect on viability, pro-inflammatory cytokine release or NO production.[22]
Graphene modified with poly-L-lysine7 daysmouse hippocampal culture modelNo effect on cell viability and morphology. No effect on neuron growth.[98]
Nanodiamonds114.7 nm,
1, 5, 10, 25, 50, 100, 250 µg/mL,
2–3 days
mouse hippocampal neurons and mouse dorsal root ganglion neuronsLow neuronal toxicity but disturbances in neuronal morphogenesis.[37]
DENDRIMERS
CMCht/PAMAM 22.0 to 30.7 nm,
200 µg/mL,
1, 2, 6, 12, 15, 18, 24 and 48 h
hippocampal neuron cultures and cortical glial cellsNo cytotoxicity.[28]
CMCht/PAMAM dendrimer NPs45 and 250 nm,
200–400 µg/mL,
1, 6, 12, 24, and 48 h, 7 days
human immortalized astrocytes
(hTERT/E6/E7)
Low level of cytotoxicity (20% of decrease in metabolic activity) after long-term exposures (7 days).[99]
POLY(LACTIDE-CO-GLYCOLIC ACID) NPs
PLGA loaded with curcumin, and modified with g7 peptide200–250 nm,
10, 20, 40 μM,
24 h
primary hippocampal neurons from ratsLack of cytotoxicity. No significant increase in apoptotic nor necrosis.[34]
PLGA loaded with curcumin200 nm,
0.001, 0.01, 0.1, 0.2, 0.5, 5, and 50 μM,
24 h
neurospheres in culture and neural stem cellsEenhanced proliferation of the neural stem cells at doses as low as 0.001 μM, with the highest proliferation at 0.5 μM. 0.5 μM of NPs were non-cytotoxic.[100]
PLGA curcumin loaded, modified with PEG and conjugated with B6 peptideless than 150 nm,
50, 100, 200, and 500 µg/mL,
24 h
HT22 cellsNo effect on cell viability.[35]
PLGA modified with PEG and phage-displayed peptides121.46 nm,
0.1; 0.5, 1; 2.5 mg/mL,
4 h
endothelial cells (bEnd.3)No cytotoxicity.[101]
POLYMER NPs
TEB NPs, bare25 nm,
50, 100, 250, 500, 800 ng/mL,
24 h
endothelial cells (bEnd.3)No obvious effect on cell viability.[102]
SOLID LIPID NPs
PLGA functionalized with a transferrin receptor or peptide mimicking transferrin receptor and loaded with siRNA115, 150 nm,
0.1; 1; 10; 100; 1000 µg/mL,
24 h
immortalized human cerebral microvascular endothelial cell line (hCMEC/D3 cell line) brain capillary endothelial cell line (BCEC)No toxicity.[91]
SERUM ALBUMIN NPs
R-flurbiprofen284.4 nm,
25, 50, 100 and 200 μM,
48 h
chinese hamster ovary (CHO) cells stably transfected with mouse Ab precursor protein 695No effect on cell viability.[103]
Polysorbate 80 or with attached Apolipoprotein E249 nm,
0.1, 1 and 2 mg/mL,
24 h
endothelial cells (b.End3)Stimulation of cell viability.[67]
OTHERS
Gelatin-siloxane NPs modified with SynB peptides-cell penetrating peptides194.55 nm,
100–600 µg/mL,
4 and 12 h
primary cultures of rat brain capillary endothelial cellsGood biocompatibility with brain capillary endothelial cells.[104]
Biodegradable polymersomes conjugated with cationic albumin95 nm,
0.25, 0.5, 1, 2, 4, 8 mg/mL,
60 min
endothelial cells (bEnd.3)Only little toxicity (viability above 85%).[75]
ICG—indocyanine green (dye), PLGA, PLG—poly(lactic-co-glicolic acid), PCL—poly(caprolatone), PLLA—poly(l-lactic acid), PEG/PLA—poly(ethylene glycol)/poly(lactide) copolymer, PAMAM—polyamidoamine, PEG-PEI—poly(ethylene glycol)/poly(ethylenimine), TEB—poly[Triphenylamine-4-vinyl-(P-methoxy-benzene)], CMCht—carboxymethylchitosan, MWNTs—multi-walled carbon nanotubes, SWCNTs—single-walled carbon nanotubes.
Table 2. The effect of nmNPs (toxic or non-toxic) in CNS-derived cancer cell in vitro.
Table 2. The effect of nmNPs (toxic or non-toxic) in CNS-derived cancer cell in vitro.
Toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
SILICA NPs
(ICG)/(PCL)-rhodamine-doped
NPs and Si-ICG/PCL/
PLLA rhodamine-doped NPs
90 and 95 nm,
2.6 × 109, 5.2 × 109, 2.6 × 1010 PCL-NPs/mL and 2.9 × 109, 5.8 × 109, 2.9 × 1010 PLLA-NPs/mL,
6 and 24 h
neuroblastoma
(SH-SY5Y)
Neurite outgrowth was not significantly altered. Reduction in neuronal differentiation.[24]
Aminopropyltriethoxysilane92.6 nm,
25–200 μg/mL,
24 h
neuronal cells
(PC12)
The accumulation of cells in the G2/M phase at a concentration of 100 and 200 μg/mL. Activation of the p53-mediated signaling pathway and apoptosis.[65]
None509.3, 356.0, 469.0 nm,
100 and 200 µg/mL,
24 and 48 h,
7 days
Neuroblastoma
(SH-SY5Y)
Mitochondrial dysfunction, pathological autophagy and cell apoptosis. Oxidative stress.[84]
CARBON NPs
MWNTs modified with carboxyl groups50 and 500 nm,
10 and 40 μg/mL,
24 h
neuroblastoma (N2a)Up-regulation of nNOS level in cells via promoting nuclear translocation and acetylation of NF-κB. Increase in NO level.[111]
MWNTs with pluronic F127 solution5 μg/mL or 10 μg/mL,
3 days, 1 week, 2 weeks
Neuroblastoma
(SH-SY5Y)
Higher concentrations of NPs (50 and 500 μg/mL) and longer incubation times (1 and 2 weeks) caused a decrease in the viability of cells.[112]
SWNTs with DNA conjugated with different surfactants sodium dodecyl sulfate, sodium dodecylbenzene sulfonate and sodium cholate0.5 mg/mL,
72 h
astrocytoma (1321N1)NPs with SC did not affect cell morphology, proliferation, or growth. NPs with SDS and SDBS surfactants demonstrated irregular cell morphology, and they were toxic to cells.[50]
Graphene layers and SWNTs0.8–1.2; 0.01; 0.1; 1; 10; 100 μg/mL,
24 h
neuronal cells (PC12)The generation of ROS after exposure of graphene layers. The upregulation of caspase 3 indicates an apoptosis of PC12 cells.[113]
PEG coated CNTs2.5–4.5 nm,
0.1; 1; 10; 100 μg/mL,
24 h
neuronal cells (PC12)The decrease in metabolic activity and generation of ROS. NPs with PEG exhibited less cytotoxic potency than uncoated NPs.[114]
Graphene, SWNTs and MWNTs coated with an ultrathin layer of gold/platinum50 nm (graphene),
∼100 nm (SWCNTs),
and 2–5; 10–15 nm (MWNTs),
0.5–500 ppm,
24 h
neuronal cells (PC12)The significant dose-dependent decreases in the viability of cells. Graphene exerted adverse effects on the neural cells at a concentration over 62.5 ppm.[115]
Fullerenes C60 and C70 functionalized with dextran polymer30.29 nm glioma (C6)Dose-dependent reduction in cell viability. NPs affect the growth, proliferation, functional and phenotypic aspects of cells. The slight degeneration of actin filaments and cytoskeletal destruction. Lysosomal integrity as well as mitochondrial membrane potential was significantly affected. ROS level inside cells was slight.[116]
DENDRIMERS
(G7) PAMAM with amine, acetamide, and carboxylate end groups, with GM1-pyrene4–8 nm,
100 nM, 200 nM, and 400 nM and 1 μM,
1 h
glioma (C6)The cells exhibited a greater sensitivity to G5-NH2 and G7-NH2 exposure but no differential effect was observed as a function of the presence of GM1 in the membrane.[53]
PAMAM 0–100 μg/mL
24 h
human glioma cell lines (87MG, U251MG, U118 and A172)Toxic. Akt/mTOR pathway was involved in the initiation of dendrimers-induced autophagy. The autophagy process induced by dendrimers is mediated by intracellular ROS generation.[109]
Complexed with siRNA200 nm,
0.005 and 0.01 mM,
4 h
glioblastoma (U87)Primary amine acetylation of dendrimers reduced their cytotoxicity.[117]
Second generation amphiphilic polylysine dendrons with siRNA3 μM
72 h
glioblastoma
(U87 and C6)
Inhibition of the proliferation of two glioblastoma cell lines. Non-toxic for non-tumoural CNS cells. Mitochondrial depolarization and the increase in ROS production.[118]
CMCht/PAMAM dendrimer 45 and 250 nm,
200–400 µg/mL,
1, 6, 12, 24, and 48 h, 7 days
GBM cel line (U87MG)Only long-term exposures (7 days) induced low levels of cytotoxicity (20% decrease in metabolic activity).[99]
LIPID NPs
D-α-tocopheryl PEG 1000 succinate -coated solid lipid NPs with resveratrol128.6 to 429.1 nm,
0–200 µg/mL,
72 h
glioma (C6)High toxicity in cancer cells. The metabolic activity significantly decreased depending on the concentration of NPs.[119]
Solid lipid NPs with retinoic acid and functionalized electrostatically with trimethyl chitosan214 nm, to 500 µg/mL,
24 h
glioblastoma (U87MG)Anti-tumor inhibitory effect by decrease in cell viability and the presence of cells in early and late apoptotic and necrotic phases.[120]
D-α-tocopheryl PEG 1000 succinate (TPGS) coated liposomes with resveratrol61–262 nm,
0–200 µg/mL,
72 h
glioma (C6)High toxicity in cancer cells. The metabolic activity significantly decreased depending on the concentration of NPs.[54]
Liposomes loaded with resveratrol PEGylated, modified with transferrin230 nm,
12.5 μM to 200 μM,
24 and 48 h
glioblastoma (U87MG)Toxic. Higher levels of apoptosis accompanied by activation of caspases 3/7. NPs with transferrin were more effective in inducing toxicity.[121]
Liposomes and micelles with diatom microalgae-derived nanoporous biosilica90.5 and 115 nm,
1.70 mg /mL,
1 h
neuroblastoma
(SH-SY5Y)
Toxic. The viability was 9–10%.[122]
PLGA lipid NPs conjugated with folic acid and ICG with resveratrol104.5–121.1 nm,
5, 10, 50 and 100 μg/mL,
24 h
glioblastoma (U87)No changes in viability rate. Apoptosis process.[68]
POLY(LACTIDE-CO-GLYCOLIDE) NPs, POLIMERIC NPs
PLGA, PLG -tocopheryl PEG NPs with resveratrol135–222 nm,
0–200 μg/mL,
72 h
glioma (C6)High toxicity in cancer cells. The metabolic activity decreased depending on the concentration of NPs.[59]
TMC surface- PLGA, PLG NPs150 nm,
0.025–8.0 mg/mL,
24 h
neuroblastoma
(SH-SY5Y)
Cell viability was slightly reduced at higher concentrations.[123]
Ephrin type-A receptor 3 (EPHA3) tyrosine kinase antibody-modified PLGA NPs145.9 nm,
4 ng/mL,
48 h
glioma (C6)NPs inhibited of cell growth.[57]
Aptamer was conjugated to the surface of PEG-PLGA NPs156 nm,
0.019, 0.038, 0.38, 0.76, 3.8, 12 and 24 mg/mL,
24, 48 and 96 h
glioma (C6)The IC50 value was detected at a concentration of 1.5 µg/mL (24 h incubation time). The cytotoxicity is dependent on incubation time.[124]
Polymeric NPs with curcumin5 or 10 μM,
24 h
medulloblastoma, glioblastoma
(DAOY, D283Med)
A dose-dependent decrease in cell growth via programmed cell death and cell cycle arrest. Dose-dependent reduction in expression of both the IGF ligands and IGF-1R. Reduction in total STAT3α protein levels and increase phosphorylation of STAT3 at Tyr 705.[125]
Tf conjugated NPs, poly(lactide)-D-a-Tocopheryl PEG PLA-TPGS diblock copolymer137.6 nm,
0.05, 0.50 and 2.50 µg/mL,
24, 48 and 72 h
glioma (C6)Toxic. 50% of death cells was observed at a concentration of 5 µg/mL.[58]
mPEG–PCL NPs with resveratrol87.5 nm,
0–31 μM
48 h
glioma (C6)A dose-dependent cytotoxicity against cells.[55]
DNA NPs
Highly PEGylated 43, 47, 59 nm,
1, 5, 10 µg/mL,
24 h
rat gliosarcoma cells (9L)Toxic. The viability was significant decreased.[92]
CHITOSAN NPs
Lactoferrin300, 800, 1200 µg/mL,
24 h
glioma 261Results suggest that the cytotoxicity of lactoferrin and NPs on glioma is attributable to its cytoplasmic allocation.[126]
OTHERS
Cardamom extract-loaded gelatin
NPs
<200 nmglioblastoma (U87MG)Apoptosis process and inhibition of the viability.[127]
Lactoferrin-curcuminoid coated polysaccharide NPs based on chitosan hydrochloride/hyaluronic acid/PEG210–240 nm,
0–10 µg/mL
Glioma (C6)Loaded NPs showed 50% cytotoxicity at concentration of 8–10 µg/mL.[56]
Non-toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
SILICA NPs
None10–20, 40–50 nm, and 90–110 nm,
0.24; 2.4; 24; 240; 2400 ppb, 24 h
glioblastoma (A-172)NPs were not toxic but NPs alter the membrane permeability.[105]
CARBON NPs
PKH26-labeled MWNTs200–400 μm length,
80 μg/mL
0–48 h
glioma (GL261) expressing eGFPNPs did not affect to cells proliferation.[21]
SWNTs with CpG oligodeoxynucleotides5 µg/mL,
24 h
glioma (GL261) expressing eGFP and luciferase-No toxicity in gliomas.[110]
Protoporphyrin IX (PX)-modified oxidized mesoporous carbon nanospheres90 nm,
0–50 µg/mL,
24 h
Neuroblastoma
(SH-SY5Y)
No toxicity during a 24 h treatment. NPs with ultrasounds were protective through the decrease in Aβ-mediated cellular toxicity.[128]
DENDRIMERS
Sialic acid0–15 µM
24 h
Neuroblastoma
(SH-SY5Y)
Attenuation of Aβ induced neurotoxicity.[129]
POLY(LACTIDE-CO-GLYCOLIC ACID) NPs, POLIMERIC NPs
G-protein estrogen receptor (GPER-1) selective agonist
was encapsulated in polymeric NPs and liposomes,
coated with polyssorbate-80
84.36 nm,
1, 10, 100 mg/mL
15 h and 24 h for N2a; 12 h and 24 h for SHSY5Y
neuroblastoma
(N2a and SHSY5Y)
NPs were not toxic to cells at 324 µg/mL, but were taken up by cells.[51]
PLGA, PLG NPs with curcumin101 nm,
0.5 µM,
1 h
human neuroblastoma (SK-N-SH)Prevention of the phosphorylation of Akt and Tau proteins in cells after induction by H2O2. anti-inflammatory and antioxidant activities of NPs.[49]
Rosmarinic acid- and curcumin-loaded polyacrylamide-cardiolipin- PLGA, PLG NPs with conjugated 83-14 monoclonal antibody40 μg/mL,
24, 36 h
neuroblastoma
(SK-N-MC)
The protective role of NPs in cells induced with Aβ peptide. NPs caused the recovery of pp38 and p-S202 expressions to normal levels.[130]
Polymeric nanostructures sulfonated and sulfated NPs40 nm,
1.6 µg/mL
24 h
neuroblastoma
(SH-SY5Y)
The protective role of NPs in cells induced by toxic Aβ peptide through decreasing caspase-3 activity and increasing cell viability.[131]
LIPID NPs
Linear polyethyleneimine (LPEI)-g-PEG copolymer-based micellar nanoparticle with siRNA<100 nm,
24 h
neuroblastoma
(N2a)
No toxicity. The viability was not decreased.[132]
Spongosome and cubosome lipid NPs co-encapsulate curcumin and fish oil, rich in ω-3 polyunsaturated fatty acids100 and 400 nm,
300 and 500 nM,
24 h
neuroblastoma
(SH-SY5Y)
The cytotoxicity of the blank and antioxidant-loaded nanocarriers was negligible. The protective effect in cells induced by H2O2.[133]
OTHERS
Chitosan NPs copolymerized with PLGA, PLG110 nm,
40 µg/mL
24 h, 96 h
neuroblastoma
(SH-SY5Y)
The protective effect in cells against toxicity induced by Aβ peptide.[134]
Lactoferrin
NPs with curcumin
43–60 nm,
2 µM
24 h
neuroblastoma
(SK-N-SH)
The cells were rescued from rotenone-induced neurotoxicity after NPs treatment. Antioxidant activity of NPs.[48]
Nanocurcumin with BSA 153 nm,
0–500 nM,
24 h
neuroblastoma
(SH-SY5Y)
NPs prevented cell death induced by 6-hydroxydopamine. The reversion of decrement p-Akt/t-Akt ratio in cells.[135]
Apolipoprotein E3 mediated poly(butyl) cyanoacrylate NPs containing curcumin178 nm,
10, 100, 1000 nM,
24 h
neuroblastoma
(SH-SY5Y)
The protective role in cells induced by Aβ peptide (antioxidant effect of NPs). The decrease in apoptotic cell population.[136]
Natural brain penetration enhancer-modified 100–200 nm,
50 ng/mL,
4 h
glioma (C6)NPs showed good biocompatibility and negligible cytotoxicity.[60]
ICG—indocyanine green (dye), PLGA, PLG—poly(lactic-co-glicolic acid), PCL—poly(caprolatone), PLLA—poly(l-lactic acid), PEG/PLA—poly(ethylene glycol)/poly(lactide) copolymer, PAMAM—polyamidoamine, PEG-PEI—poly(ethylene glycol)/poly(ethylenimine), TEB—poly[triphenylamine-4-vinyl-(P-methoxy-benzene)], CMCht—carboxymethylchitosan, MWNTs—multi-walled carbon nanotubes, SWNTs—single-walled carbon nanotubes, TMC—trimethylated chitosan, Tf—transferrin, BSA—bovine serum albumin.
Table 3. The effect of nmNPs (toxic or non-toxic) in vivo. Mammalian species.
Table 3. The effect of nmNPs (toxic or non-toxic) in vivo. Mammalian species.
Toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
SILICA NPs
SiO2NPs modified with aminopropyl-triethoxysilane 15 nm,
20 μg/rat,
1 day, 7 days
adult ratsInduction of the oxidative stress and an increased inflammatory response in the striatum. The decrease in neurotransmitter dopamine and the downregulation of tyrosine hydroxylase protein in the brain.[65]
None 115 nm,
8 mg/kg,
1 and 2 months
male C57BL/6 N miceMood dysfunction and cognitive impairment and neurodegeneration-like pathology and synaptic changes via ERK activation.[137]
None6, 20 and 50 nm,
150 μg/mL,
28 days
Wistar male ratsSOD and CAT activity was increased in the brain. The MDA level was increased in the brain and degenerative changes in the nerve fibers.[138]
None509.3, 356.0, 469.0 nm,
5 μg/mL,
3 months
male transgenic mice expressing A53T human a-Syn (a-Syn
A53T Tg mice)
NPs promoted PD-like pathology including α-Synuclein aggregation and dopaminergic neuronal degeneration. Mitochondria impairment, oxidative stress, autophagy dysfunction, and neuronal apoptosis.
[84]
None50 and 500 nm,
20 mg/kg, 1 h, 2 h, 24 h and 28 days
miceDegeneration of neurons. The regulation of apoptosis by regulating Bax and Bcl-XL expression. Elevation of the autophagic responses. NPs administered via the intranasal instillation route resulted in more severe brain lesions compared to the intravenous injection route.[139]
None40 and 80 mg/kg,
14 days
ratsNPs passed from the BBB into the brain. The reduction in the activity of SOD and CAT. Cellular morphological modifications, mitochondrial dysfunction, and oxidative stress.[140]
CARBON NPs
MWNTs modified with carboxyl groups5–40 µg/mL,
2, 4, 7 days
C57BL/6J miceThe neurotransmitter level was decreased. The increase in the NOS release. High levels of nNOS expression in regions associated with regulation of sympathetic nerve activity.[111]
MWNTs5 mg/m3,
5 h
male Sprague–Dawley ratsInhalation of NPs significantly changes the balance between sympathetic and parasympathetic nervous system.[88]
MWNTs shortened by oxidation and functionalized with amino groups (oxMWNT-NH3+) or functionalized with amino groups (MWNT-NH3+)20–30 nm,
500 ng/mouse,
30 days post-injection
female C57/Bl6 miceThe increase in pro-inflammatory cytokine gene expression. oxMWNT-NH3+ induced a higher expression of pro-inflammatory cytokines compared to MWNT-NH3+. In addition, oxMWNT-NH3+ induced higher expression of GFAP and CD11b.[70]
MWNTs49 nm,
10 or 40 μg/mouse,
4 h
male C57BL/6J miceNeuroinflammatory responses depend on the disruption of BBB integrity.[141]
SWNTs modified with PEG10 to 1000; or 1000 to 10,000 nm,
0.5, 1.0, and 2.1 mg/mL, 24 h and 30 min
Wistar male ratsImpairment of fear memory retrieval. Lipid peroxidation in the hippocampus.[142]
DENDRIMERS
G4 and G4-C12 modified PAMAM1 μM,
24 h
C57/BL6-j miceHigher concentrations of G4-C12 PAMAM dendrimer were toxic and caused the apoptosis process, while G4 PAMAM accumulation did not show any sign of apoptosis. Low level of glial activation.[25]
OTHERS
Chitosan NP modified with Polysorbate 80251 nm,
3, 10, 30 mg/kg,
0.5, 2, 4, 8, and 24 h, 7 days
Sprague-Dawley male ratsNPs can enter the brain and induce the apoptosis and necrosis of neurons, slight inflammatory response in the frontal cortex. The decrease in GFAP expression in the cerebellum.[143]
Polibuthylcyanoacrylate NPs modified with Polysorbate 80 and polystyrene NPs modified with dalargin200 nm,
13.5 mg/kg,
5 min
miceThe locomotor activity decreased.[144]
Polystyrene NPs, COOH-modified80, 100, 200 nm,
7 days
mice after the aerosol inhalationNPs with a size of 80 nm can deposit in the brain of mice via aerosol inhalation triggering neuron toxicity and altering the animal behavior. Inhibition of aChE activities.[145]
Non-toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
CARBON NPs
MWNTs, NH2-functionalized18.9 nm,
[111In]- MWNTs (50 µg, 0.5 MBq) in 100 mL PBS,5 min, 30 min, 1 h, 4 h and 24 h
C57/Bl6 miceNPs were present in both brain capillaries and parenchyma fractions. NPs are potential nanocarriers to use for the delivery of drugs.[32]
MWNTs, coated with Pluronic F12710–30 nm,
0.5 mg/mL,
3 days
miceNPs provoked no damage to the overall organization of the brain.[97]
MWNTs~20 nm,
5 µg,
72 h
GL261-bearing miceNP uptake occurred by tumor-associated macrophages. No toxicity to mice with glioma.[146]
MWNTs,
COOH-modified
40 nm,
1 mg/mL,
7, 28 days
BALB/C miceThe lack of pathological changes in the brain.[147]
Fullerene C60200–500 mg/kg,
1, 3, 6, 16, 30 and 160 h
ratsNPs were able to penetrate the BBB, but the toxicity was found to be quite low.[148]
Fullerene C603.4 mg/kg,
10–12 month
miceNo health deterioration in mice. No significant effect on body weight, spontaneous locomotor activity, and grip strength. Fasting blood glucose and glucose tolerance are not affected. No changes in the blood parameters in mice. No significant influence of NPs on organ weight except for a higher kidney weight in males compared with females.[149]
Nanodiamond50 nm,
20 mg/kg,
28 days
miceVery low NPs concentration was detected in the brain.[150]
Nanodiamond114.7 nm,
10 μL of NPs at a concentration of 100 mg/mL,
10 min
ratsNPs did not induce cytotoxicity in primary neurons from either central (CNS) or peripheral nervous system (PNS) and did not affect animal behavior.[37]
Carbon nano-onion15 nm,
10 mL per g body weight of 1.0 mg/mL of NPs,
4, 12, 24 h
FVB/N miceCarbon nano-onion crosses not only through the BBB into the brain of leukoencephalopathy mice but also through the glioblastoma multiforme-induced mice.[151]
DENDRIMERS
PEG1.9 nm,
55 mg/kg,
1, 4 and 24 h
cerebral palsy rabbits model, murine orthotopic model of glioblastomaNPs were taken up by the brain and accumulated in the corpus callosum (white matter), hippocampus, and cortex. They fully penetrate and distribute throughout the solid tumor. [107]
Modified with PEG, lactoferrin and DNA50 µg/mouse,
2 h
Balb/c miceLactoferrin improved the NP uptake. NPs can be a potential non-viral gene vector to the brain via noninvasive administration. [152]
Dendrimers and dendriplexes loaded with siRNA15 mg/kg in 200μL PBS,
1 and 24 h
BALB/c miceNPs were present inside the brain, but there was no specific brain histology alterations.[90]
N-acetyl cysteine and valproic acid∼5 mg/kg dogs, 48 hdogsNPs were not toxic but were present in the brain. Dendrimers with drugs improved neurological outcome in injured brain.[153]
None~6.7 and ~4.3 nm,
6 mg/kg,
48 h
dogs (brain injury model)Generation 6 dendrimers showed extended blood circulation times and increased accumulation in the injured brain compared to generation 4 dendrimers, which were undetectable in the brain by 48 h after final administration.[30]
Ethylenediamine-core4 nm,
2.5 µg of dendrimer in 5 µL of PBS,
24 h
white rabbitsThe high uptake of NPs into astrocytes and microglia cells.[26]
LIPID NPs
Liposomes modified with BSA118.2–185.8 nm,
11 and 25 mg lipid /kg,
1, 3, 6 and 24 h
Wistar ratsNPs were able to move into brain tissue. NP uptake is dependent on the presence of BSA. NPs are a promising tool in drug delivery to the CNS.[74]
Liposomes modified with PEG
40, 80, 200 nm,
1 mM,
1, 6, 24 and 48 h
ratsLiposomes accumulated in a subpopulation of perivascular cells within the brain dependent on charge and PEG coated. NPs are a promising tool in drug delivery to the CNS.[154]
D-α-tocopheryl polyethylene glycol 1000 succinate (TPGS) coated liposomes with resveratrol 107.8, 212.5 and 262.3 nm,
2 mg/kg of resveratrol,
0, 0.083, 0.25, 1, 2, 4, 8, 12, 24, 36 and 48 h
Charles Foster ratsNPs showed high accumulation in the brain but are biocompatible and safe.[54]
Liposomes with phospholipid, a polymer surfactant and cholesterol<200 nm,
10 and 150 µL form 20 nM solution,
4 and 24 h
rats and C57/BL6 miceNPs were delivered to myelinated peripheral nerves. The liposomes were presented in choroid epithelium, but not in myelinated white matter regions or in brain parenchyma. NPs are a promising tool in drug delivery to the CNS.[155]
Micellar linear polyethyleneimine (LPEI)-g-polyethylene glycol (PEG) copolymer NPs
below 100 nm,
0.6%, 0.8% and 1.2% PEG grafting density in LPEI-g-PEG/Bace NPs,
from 2 to 7 days
C57BL/6J miceNPs did not impact on the activation of astrocytes and microglial cells in ipsilateral hippocampus of mice. No inflammation and no cytotoxicity in the brain. Normal morphology of astrocytes and microglia.[132]
POLY(LACTIC-CO-GLYCOLIC ACID) AND POLYMERIC NPs
PLGA, PLG NPs modified with a 7-aminoacid glycopeptide and with albumin234 nm,
2 mg NPs,
2 h
C57BL/6 mouseNPs were able to deliver low molecular weight molecules-albumin across the BBB in two murine models of lysosomal storage disorders. NPs are a promising tool in drug delivery to the CNS.[156]
PEG/PLA NPs, modified with cationic BSA80 nm,
60 mg/kg,
30 min
miceNPs were observed in the lateral ventricle, third ventricle and periventricular region of the brain. NPs are a promising tool in drug delivery to the CNS with low toxicity.[64]
PLGA, PLG NPs modified with glycopeptide (g7)200 nm,
1 mL of NPs (8 mg) suspension,
3 h
Balb/c miceNPs can enter to the brain. NPs are a promising tool in drug delivery to the CNS.[157]
PLGA, PLG -tocopheryl PEG 1000 succinate blend NPs with resveratrol175.5, 199.7 and 222.5 nm,
2 mg/kg of resveratrol,
0.25, 0.5, 1, 2, 4, 8, 12, 24, 36 and 48 h
Charles Foster ratsNPs showed high accumulation in the brain. NPs are a promising tool in cancer therapy, and they are safe.[59]
PLGA, PLG tocopheryl PEG succinate diblock copolymer modified with transferrin161.5, 121.6,
137.6 nm,
5 mL/kg,
4 h
Sprague-Dawley ratsNPs were present in the brain. NPs could be able to deliver imaging/therapeutic agents across the BBB.[58]
12-amino-acid-peptide conjugated onto the surface of PLGA, PLG PEG NPs121 nm,
30 µg/kg,
0.083 h, 0.25 h, 0.5 h,1 h, 2 h, 4 h, 8 h,12 h and 24 h
nude
mice
Enhanced brain accumulation efficiency together with lower accumulation in the liver and spleen was observed in the mice intravenously. Injection with peptide- conjugated NPs compared with plain NPs, shows powerful brain selectivity. NPs could be able to deliver imaging/therapeutic agents across the BBB.[101]
PLGA, PLG NPs modified with ephrin type-A receptor 3 tyrosine kinase antibody145 nm,
0.5 mg/kg,
1, 2, 4, and 8 h
Sprague-Dawley ratsAnti-EPHA3-modified NPs showed high fluorescence intensity in the brain and effectively accumulated in glioma tissues. A significantly higher tumor cell apoptosis.[57]
PLGA, PLG 1,2-distearoyl-glycerol-3-phospho-ethanolamine-N-[methoxy(PEG)-2000] ammonium salt hybrid NPs modified with curcumin169 nm,
25 µM,
5 days
Wistar ratsTumor size decreased significantly after 5 days.[158]
PLGA, PLG NPs modified with polysorbate 80, poloxamer 188 and chitosan250 to 400 nm,
100 mg/mL,
0, 15, 60, 120, 240 min
ratsNPs concentrations in the brain were increased by NP surface modification. NPs could be able to deliver therapeutic agents across the BBB.[77]
PLGA, PLG NPs modified with cationic BSA97, 98, 104 nm,
30 mg/ kg,
0.033 h and 1 h
BALB/c miceThe increase in surface density of the NPs enhanced the BBB permeability. NPs are a promising tool in drug delivery to the CNS.[159]
PEG-poly(ε-caprolactone) polymersomes modified with cationic BSA95 nm,
10 mg/kg,
0.5, 1, 2, and 4 h
Sprague-Dawley ratsNPs can enter to the brain and might be a promising carrier for drug delivery to the brain.[75]
ALBUMIN NPs
None143 and 151 nm,
1 µL from 10 µg/mL rats and 100 mg/ kg mice,
1,7, 14 days
adult rats and miceNPs located in different brain tissues but did not induce an inflammatory response. No locomotor, explorative, or cognitive function impairment.[160]
Polysorbate 80 or with attached apolipoprotein E249 nm,
200 μg of NPs/g body weight,
15 and 30 min
female SV 129 miceThe uptake was dependent on the presence of apolipoprotein E. NPs with apolipoprotein E were taken up less by animal brain cells. NPs are a promising tool in drug delivery to the CNS.[67]
borneol, muscone and menthol100–200 nm,
40 mg/kg,
1, 2, 4, 8, 12, and 24 h
BALB/c miceAccumulate in glioma cells with a much higher targeting efficiency than that of transferrin-modified NPs.[60]
CHITOSAN NPs
Thymoquinone-encapsulated between 150 and 200 nm,
500 μg of TQ/25 μL,
0.25, 0.5, 2, 4, 6, 24 h
Wistar ratsTaken up by brain tissue. NPs are a promising tool in drug delivery to the CNS.[161]
OTHERS
Poly(n-butyl cyanoacrylate) dextran polymers coated with polysorbate 8048 nm,
200 µL from 60 mg/mL of NPs,
2 days
APP/PS1 mice (AD model)No induction of non-specific BBB disruption. NPs collaborate with plasma apolipoprotein E to facilitate BBB crossing.[66]
Cyanoacrylate NPs, PEGylated and coated with polysorbate 80 or poloxamine 908137 nm,
30 mg/kg, 60 mg/kg,
30 min; 1 and 4 h
OF1 mice and rats dark agouti DA/RNPs can penetrate into the brain. Localization in the ependymal cells of the choroid plexuses, in the epithelial cells of pia mater and ventricles and to a lower extent in the capillary endothelial cells of BBB. NPs are a promising tool in drug delivery to the CNS.[78]
Polybutylcyanoacrylate NPs coated with polysorbate 80, loaded with curcumin 152 nm,
5 mg/kg,
0.033; 0.083; 0.25; 0.5; 1.25; 0.75; 1; 2; 3; 4; 8; h
miceNPs inside the brain tissue. NPs are a promising tool in drug delivery to the CNS.[162]
Polysaccharide NPs based on hyaluronic acid and chitosan hydrochloride, functionalized with lactoferrin and loaded with curcumin>200 nm,
1.25 mg/kg,
24 h
ICR miceAccumulated in the brain. The presence of lactoferrin improved the NP uptake. NPs can be potentially used for cancer therapy.[56]
PLA NPs coated with T-80162 nm,
8 mg/kg,
45 min
micePresent in the brain. The specific role of T-80 coating on NPs in brain targeting. NPs are a promising tool in drug delivery to the CNS.[69]
Polystyrene latex nanospheres20, 100 and 1000 nm,
90–120 days
Fischer F344 ratsLow numbers of particles in the brain.[71]
Polystyrene NPs COOH-modified and methoxy (MeO)-PEG amine (NH2)60 nm,
2.5 mg/kg,
1 and 24 h
Fischer F344 ratsDiffusion in normal brain tissue. Delayed tumor growth following local administration, this effect was dependent on the presence of PEG.[79]
PEG and polyethylenimine (complexes of oligonucleotides with 3H-labeled nanogel)<100 nm,
200 µL,
1 h
miceAccumulation in the brain. Nanogel is a promising system for the delivery of oligonucleotides to the brain.[43]
Gelatin-siloxane NPs modified with SynB peptide194 nm,
1 mL, 60 mg/kg,
0.5, 1, 2, and 4 h
nude mice and Sprague-Dawley ratsThe presence of peptide improved the NP uptake. No toxicity.[104]
PLGA, PLG—poly(lactic-co-glicolic acid), PCL—poly(caprolatone), PAMAM—polyamidoamine, MWNTs—multi-walled carbon nanotubes, SWNTs—single-walled carbon nanotubes.
Table 4. The effect of nmNPs (toxic) in vivo. Non-mammalian species.
Table 4. The effect of nmNPs (toxic) in vivo. Non-mammalian species.
Toxic Effect
Surface Coating and/or NPsSize, Concentration and Exposure TimeModelResultsRef.
SILICA NPs
None62 nm,
25, 50, 100, 200 µg/mL,
4–96 h
zebrafish embryos
(D. rerio)
Embryonic developmental toxicity, resulted in persistent effects on larval behavior.[173]
CARBON NPs
Fullerenes aqueous suspended colloids (nC60)0.5 ppm and 1 ppm,
48 h
juvenile largemouth bassLipid peroxidation in brain after 48 h of exposure.[171]
Fullerenes (C60), long or shortMWNTs, SWNTs and 0.001 mg/L,
21 days
D. rerioHigh level of alterations in the DNA/ RNA region, especially in the brain.[172]
POLYSTYRENE NPs
None50 nm,
0, 1, 10, 100 and 1000 μg/L,
72 h
Caenorhabditis elegans used as a model organism to evaluate the neurodevelopmental toxicitySignificant inhibition in body length, survival rate, head thrashes, and body bending. The increase in ROS production. The lipofuscin accumulation, apoptosis and decrease in dopamine contents. pink-1 gene was involved in the polystyrene NPs-induced neurotoxicity. NPs at concentration of 100 μg/L caused up-regulation of 89 genes and down-regulation of 56 genes regulated differently expressed genes correlated with biological function of cuticle development and molting cycle. The neurodevelopmental toxicity and oxidative stress responses induced by NPs were regulated by dpy-5 and rol-6.[175]
NH2-modified, positively charged25–330 nm,
0.005 g/L to 0.150 g/L,
24 h
Daphnia magnaThe alterations in behavioral patterns, decreased brain mass and morphological changes in the cerebral gyri.[167]
None<100 nm,
100 mg/L NPs,
15 days
Cyprinus carpioDegrees of necrosis, fibrosis, changes in blood capillaries, tissue detachment, edema, degenerated connective tissues, and necrosis in large cerebellar neurons and ganglion cells. The activity of acetylcholinesterase (aChE) significantly decreased after exposure to 100 mg/L.[168]
None50 and 200 nm,
10–10,000 parts per billion,
5 days
larval zebrafish
(D. rerio)
Accumulation in the tissues of larval zebrafish, alteration of their transcriptome, and changes in behavior and physiology. Potentially decreasing organismal fitness in contaminated ecosystems.[174]
None50 nm,
1 mg/L,
till 120 h post fertilization
D. rerioNeurotoxicity. Suppression of locomotion activity.[176]
Fluorescent NPs51 nm,
0.1, 1, 10 ppm,
until 120 h post fertilization
D. rerioAccumulation in the brain and alteration of larval behavior.[177]
Fluorescent NPs70 nm,
0.5, 1.5, 5 ppm,
7 days, 30 days and 7 week
D. rerioAlteration of neuro-behavior and neurotransmitter regulation.
[178]
Fluorescent NPs20 nm,
270 ppm,
120 h
D. rerioNPs can reach brain, oxidative damage and apoptosis.[179]
None100 nm,
1, 10, 100, 1000 µg/L,
Prolonged exposure:from L1-larvae to adult day-1
C. elegansInduction of neurodegeneration of D-type GABAergic motor neurons.
Alteration of forward and backward movement.
[180]
None0.1, 0.5,
1, 2 and 5 μm,
1 mg/L,
3 days
C. elegansExcitatory toxicity on locomotive behavior. Damage to cholinergic and GABAergic neurons.[181]
None0,11 µm,
0.005, 0.05, 0.5, 5 and 50 mg/L,
96 h
Mediterranean
mussel (Mytilus
galloprovincialis)
Changes in gene expression. Cholinesterase inhibition in hemolymph.[169]
NH2-modified, positively charged50 nm,
0.1, 1.0, 3.0 and 10.0 mg/L,
48 h up to 14 days
Brine shrimp (Artemia fransiscana)Cholinesterase inhibition. Glutathione S-Transferase and catalase decreased.[170]
None40 nm,
10 mg/L,
7 days
Japanese rice fish
(Oryzias latipes)
Particle presence in the brain, suggesting penetration of BBB.[166]
None0,1 µm,
1, 10 and 100 μg/L,
1–14 days
Red tilapia
(Oreochromis
niloticus)
Particle presence in brain tissue. Inhibition of aChE activity in the brain.[182]
NH2-modified, positively charged100 mg/L,
64 days
Crucian carp
(Carassius carassius)
Particle were present in the brain and caused brain weight loss. Behavioural changes and enlarged cerebral gyri.[167]
None50 nm,
1 mg/L,
3 days
Zebrafish, larvae
(D. rerio)
Particle were present in head, gills and muscle. Decrease aChE activity.[183]
MWNTs—multi-walled carbon nanotubes, SWNTs—single-walled carbon nanotubes.
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

Sikorska, K.; Sawicki, K.; Czajka, M.; Kapka-Skrzypczak, L.; Kruszewski, M.; Brzóska, K. Adverse Effects of Non-Metallic Nanoparticles in the Central Nervous System. Materials 2023, 16, 7264. https://doi.org/10.3390/ma16237264

AMA Style

Sikorska K, Sawicki K, Czajka M, Kapka-Skrzypczak L, Kruszewski M, Brzóska K. Adverse Effects of Non-Metallic Nanoparticles in the Central Nervous System. Materials. 2023; 16(23):7264. https://doi.org/10.3390/ma16237264

Chicago/Turabian Style

Sikorska, Katarzyna, Krzysztof Sawicki, Magdalena Czajka, Lucyna Kapka-Skrzypczak, Marcin Kruszewski, and Kamil Brzóska. 2023. "Adverse Effects of Non-Metallic Nanoparticles in the Central Nervous System" Materials 16, no. 23: 7264. https://doi.org/10.3390/ma16237264

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