Skip to main content
Materials Today Bio logoLink to Materials Today Bio
. 2023 Jun 1;20:100673. doi: 10.1016/j.mtbio.2023.100673

Challenges and advances for glioma therapy based on inorganic nanoparticles

Die Hu a,1, Miao Xia a,1, Linxuan Wu a,1, Hanmeng Liu a, Zhigang Chen a, Hefeng Xu a, Chuan He c,∗∗∗, Jian Wen b,∗∗, Xiaoqian Xu a,
PMCID: PMC10333687  PMID: 37441136

Abstract

Glioma is one of the most serious central nervous system diseases, with high mortality and poor prognosis. Despite the continuous development of existing treatment methods, the median survival time of glioma patients is still only 15 months. The main treatment difficulties are the invasive growth of glioma and the obstruction of the blood-brain barrier (BBB) to drugs. With rapid advancements in nanotechnology, inorganic nanoparticles (INPs) have shown favourable application prospects in the diagnosis and treatment of glioma. Due to their extraordinary intrinsic features, INPs can be easily fabricated, while doping with other elements and surface modification by biological ligands can be used to enhance BBB penetration, targeted delivery and biocompatibility. Guided glioma theranostics with INPs can improve and enhance the efficacy of traditional methods such as chemotherapy, radiotherapy and gene therapy. New strategies, such as immunotherapy, photothermal and photodynamic therapy, magnetic hyperthermia therapy, and multifunctional inorganic nanoplatforms, have also been facilitated by INPs. This review emphasizes the current state of research and clinical applications of INPs, including glioma targeting and BBB penetration enhancement methods, in vivo and in vitro biocompatibility, and diagnostic and treatment strategies. As such, it provides insights for the development of novel glioma treatment strategies.

Keywords: Glioma, Inorganic nanoparticles, BBB penetration, Biocompatibility, Diagnosis and treatment

Graphical abstract

Glioma treatment is facing to many challenges. Theranostic inorganic nanoparticles is becoming popular in recent years, as they can provide opportunities to combat glioma through different diagnosis and treatment strategies with satisfied biocompatibilities. Summarize of the current progresses would provide exceptional insights for a new generation of nanomaterials to the management of glioma.

Image 1

1. Introduction

Glioma is a type of brain tumor that originates from glial stem cells or progenitor cells. It accounts for 30% of all primary brain tumors and 80% of all malignant brain tumors. Glioma is also the most common malignant tumor in the central nervous system (CNS) and its patients have a poor prognosis and a high risk of relapse and death. Currently, the main clinical treatment modes are surgical resection, radiotherapy, chemotherapy [1]. However, because of the particularity of the intracranial structure and the pathological characteristics of glioma, current treatments often lead to high recurrence rates, susceptibility to drug resistance, large side effects, and poor prognosis. Hence, new technological strategies are urgently needed to overcome these difficulties and provide improved glioma treatment strategies.

With the development of nanotechnologies, nanomaterials have been extensively explored in the field of biomedical applications, including the diagnosis and treatment of glioma. Research on the application of nanoparticles (NPs) in the diagnosis and treatment of glioma has been ongoing since the late 20th century. The increasing number of articles over the past 20 years were summarized in Fig. 1, which indicates that approaches brought about by nanotechnology have attracted much attention. The reviews in the past five years mainly focus on the following aspects, including NPs based targeted drug delivery systems to overcome BBB, advances of nanomaterials in brain diseases but not limited to glioma, the prospect of NPs as carrier-loaded Temozolomide (TMZ) targeting therapy for glioblastoma (GBM), and the application of nanomedicine in glioma immunotherapy or gene therapy [[2], [3], [4], [5], [6], [7]].

Fig. 1.

Fig. 1

The bibliometric analysis of nanomaterials used in the diagnosis and treatment of glioma. Data source from the Web of Science Core Collection (2000–2022).

Among them, INPs have also experienced considerable development. As shown in Fig. 1, the application research of INPs in the diagnosis and treatment of glioma also shows a trend of development year by year, with an annual circulation of about 100 articles in the past five years. Initially, the research mainly focused on the application of magnetic nanoparticles (MNPs) for magnetic resonance imaging (MRI) and gradually expanded the research scope of inorganic nanomaterials, including but not limited to metal nanomaterials, graphene-based nanomaterials, mesoporous silica nanoparticles (MSNs), upconversion nanoparticles, and quantum dots [8]. For instance, Norouzi et al. [9] reviewed and discussed the multiple applications of Au NPs in the diagnosis and treatment of gliomas from 2004 to 2020, including chemotherapy, phototherapy, gene therapy, MRI, Photoacoustic imaging (PAI), surface-enhanced Raman scattering (SERS). They discussed the cellular uptake and pharmacokinetic differences of gold-based NPs (GNPs) of different shapes, and crossing BBB capabilities of GNPs of different types, as well as the possible impact of varying tumor models on NPs studies in vivo, paving the way for future clinical studies. In another work, Choi et al. [10] reviewed previous studies using high atomic number (high-Z) metal NPs in GBM radiotherapy between 2005 and 2020 to showcase the potential of high-Z metal NPs application in the radiosensitization of GBM. Verma et al. [11] introduced studies in which MNPs, gold nanorods (GNRs), and carbon nanotubes (CNTs) have been applied for hyperthermia in GBM from 1996 to 2014. They introduced the diverse synthesis methods of various INPs and discussed their application in hyperthermia ablation and enhanced drug delivery in GBM. To date, research progress of glioma diagnosis and treatment based on INPs has been updated rapidly.

Generally, the physical and chemical properties of inorganic elements, such as the quantum size effect, surface effect, small size effect, macroscopic quantum tunnelling effect and dielectric domain effect, can be obtained by nanoscale construction [12,13]. Consequently, a series of properties propitious to biological applications can be generated, especially for tumor diagnosis and treatment. These include efficient endocytosis with low cytotoxicity, an enhanced tumor permeability and retention effect, high optical imaging efficiency, and tumor ablation effects that can be excited by physical means, such as light and nuclear magnetism. The increasing biomedical research into INPs is providing new strategies and opportunities for the diagnosis and intervention of gliomas [14,15]. The main inorganic elements currently employed for nanoparticle fabrication related to glioma research are gold (Au), silver (Ag), copper (Cu), cadmium (Cd), zinc (Zn), titanium (Ti), molybdenum (Mo), lanthanum (La), gadolinium (Gd), superparamagnetic iron oxides, graphene (GPE), mesoporous silicon, and carbon (C). INPs fabricated by these elements have been widely used in tumor therapies, such as for tumor imaging, improved drug target delivery, synergistic therapy, prolonged drug cycling, reduced cytotoxicity, and the utilization of photosensitizers (PSs) in Photothermal Therapy (PTT) to obtain safer and more effective therapeutic effects [[16], [17], [18]].

Nevertheless, there are few comprehensive and systematic reviews of glioma diagnosis and treatment based on all types of INPs mentioned above. The advantages of our work lie in the selection of representative research results in the past five years, as well as the unique and systematic summary of the latest progress and new glioma treatment strategies of INPs. First, the staging and typing of gliomas are introduced according to the World Health Organization's (WHO) clinical guidelines. Then, the characteristics of BBB structures at different stages of glioma and the commonly used targeting ligands are summarized. Furthermore, the biocompatibility of different INPs is compared and analyzed. Finally, nanotechnology applications specific to glioma, such as various imaging modalities, new strategies and improvements in glioma treatment based on INPs, are comprehensively compared and discussed. Fig. 2 is a schematic illustration of the BBB-crossing inorganic nanoplatforms used for highly efficient glioma theranostics.

Fig. 2.

Fig. 2

Schematic illustration of BBB-crossing inorganic nanoplatforms used for efficient glioma theranostics.

2. Classification of glioma and limitations of treatment

Gliomas can be classified into many types according to malignancy grades. WHO classification of CNS tumors is always the gold standard in clinical work. According to the disease progression, they are mainly categorized into four grades (1–4) [19]. WHO grade 1 includes the least aggressive tumors, which are curable if they can be surgically removed. In contrast, WHO grade 4 tumors are highly malignant and hold the highest mortality rate and shortest survival time. In the newest 5th edition of the guidelines, gliomas are classified into five different families according to their grade of malignancy: adult-type diffuse gliomas, paediatric-type diffuse low-grade gliomas, paediatric-type diffuse high-grade gliomas, circumscribed astrocytic gliomas, and ependymomas [20]. Their gene alterations are summarized in Table 1. Among them, the more aggressive gliomas are the diffuse tumors, which are characterised by rapid growth. Diffuse gliomas, which account for the majority of glioma patients, are characterized by tumor cells invading the nervous system individually or in groups, and present diffuse infiltrating growth in the parenchyma of the CNS [21]. Moreover, this particular pattern of diffuse cell invasion is rare in other tumors [22]. In the traditional classification, diffuse gliomas are histologically classified as astrocytomas (astrocytic), oligodendrocytomas (oligodendrocytic), and mixed gliomas. In the 2021 WHO classification of CNS tumors, diffuse gliomas are divided into adult and paediatric types for the first time. This classification is mainly based on the clinical characteristics of the tumor distributions in different age groups. This does not mean that adult-type gliomas do not occur in children. Paediatric gliomas are classified into two types: low-grade with a good prognosis and high-grade with a very poor prognosis. The name of circumscribed astrocytic gliomas is relative to the diffuse gliomas, whose growth pattern is more limited; however, there is still a risk of invasion. The pathogenesis of gliomas is very complex; however, the recognized root causes are mainly gene mutations in vivo, as well as abnormal cell signalling pathways and functional disorders. The common mutations mainly include IDH, H3, BRAF. A detailed summary of the latest 2021 WHO glioma grading its associated genes markers/genetic alterations is provided in Table 1. Among the primary brain tumors, the GBM type undoubtedly has the highest degree of malignancy. GBM is considered to be the most aggressive and highly invasive type since it can quickly spread to any part of the brain. Thus, it is classified as grade 4, which has the worst prognosis [23]. Currently, obstacles to the treatment of glioma include the intricacy of brain tissue, acquired resistance owing to chemotherapy, the heterogeneity and aggressiveness of the tumor, and difficulty in detecting malignant tissue margins. Surgical excision is always the first choice. However, most gliomas possess crab-like infiltrative growth properties and some tumors are always located in brain areas that dominate important functions; therefore, it is difficult to determine malignancies from normal brain tissue [24]. Consequently, total surgical excision of gliomas is almost impossible due to their aggressive behaviour. Even after substantial resection, gliomas can still recur locally within a few centimetres of the resection margin [25]. TMZ as the first-line chemotherapy drug for glioma, is prone to drug resistance [26]. Treatment by radiotherapy must be considered carefully, since severe or even irreversible damage to the CNS can be generated [27]. In addition, the presence of the BBB structure limits the implementation of most treatments. One of the greatest challenges to the development of effective drugs for the treatment of glioma disease is their capacity to effectively penetrate the BBB.

Table 1.

Classification and grading of gliomas [20].

Glioma Classification Name Genes markers/Genetic alterations WHO grade
Diffuse gliomas Adult-type diffuse gliomas Astrocytoma IDH1, IDH2, ATRX, TP53, CDKN2A/B 2/3/4
Oligodendroglioma IDH1, IDH2, 1p/19q, TERT promoter, CIC, FUBP1, NOTCH1 2/3
Glioblastoma IDH-wildtype, TERT promoter, chromosomes 7/10, EGFR 4
Paediatric-type diffuse low-grade gliomas Diffuse astrocytoma∗ MYB, MYBL1 1
Angiocentric glioma MYB 1
Polymorphous low-grade neuroepithelial tumor of the young∗ BRAF, FGFR family 1
Diffuse low-grade glioma FGFR1, BRAF NA
Paediatric-type diffuse high-grade gliomas Diffuse midline glioma H3 K27, TP53, ACVR1, PDGFRA, EGFR, EZHIP 4
Diffuse hemispheric glioma∗ H3 G34, TP53, ATRX 4
Diffuse paediatric-type high-grade glioma∗ IDH-wildtype, H3-wildtype, PDGFRA, MYCN, EGFR (methylome) 4
Infant-type hemispheric glioma∗ NTRK family, ALK, ROS, MET NA
Non-diffuse gliomas Circumscribed astrocytic gliomas Pilocytic astrocytoma KIAA1549-BRAF, BRAF, NF1
High-grade astrocytoma with piloid features∗ BRAF, NF1, ATRX, CDKN2A/B (methylome) NA
Pleomorphic xanthoastrocytoma BRAF, CDKN2A/B
Subependymal giant cell astrocytoma TSC1, TSC2
Chordoid glioma PRKCA
Astroblastoma∗ MN1 NA
Ependymal tumors Supratentorial ependymoma∗ ZFTA, RELA, YAP1, MAML2 2/3
Posterior fossa ependymoma∗ H3 K27me3, EZHIP (methylome) 2/3
Spinal ependymoma∗ NF2, MYCN NA
Myxopapillary ependymoma 2
Subependymoma

The Arabic numerals 1–4 are used to represent the tumoral grades. Those ∗ labelled are newly added tumor types, some ∗ labelled tumors have not been clearly graded, and the rest can refer to the 2016 fourth-edition.

3. INPs-based solutions for efficient glioma treatment

A high drug utilization ratio is critical for glioma treatment, as low drug utilization often leads to the drug's side effects outweighing its efficacy. Drug delivery to glioma can be categorized into two main areas: bypassing BBB and crossing BBB. Denzil et al. [28] provide a concise overview of how therapeutic drugs can be delivered to the CNS without establishing BBB-crossing pathways. Therapeutic methods that bypass BBB have problems, including slow drug distribution, rapid drug elimination through active transport, and extremely low penetration into the brain parenchyma [29]. There are few studies on the biological application of INPs in the bypass BBB pathways. This paper mainly discusses glioma treatment strategies through the BBB-crossing pathways based on INPs. In general, the efficiency of drugs used to treat glioma by BBB-crossing methods is dependent on both their BBB permeability ability and their targeting ratio to glioma [30]. Nanomedicine holds promise for the development of successful strategies to penetrate the BBB and target glioma. In this section, we discuss the INPs-based solutions to the main challenges for guaranteeing the drug utilization ratio.

In addition, it is also worth mentioning that, benefit from the responsive properties to some physical factors, including laser, optoacoustic signal and magnetic field, INPs can also enhance the penetration of BBB and combat glioma. In this article, INPs used as photoacoustic/magnetic/fluorescent responsive nanoprobes for molecular imaging, see Sections 5 and 6.3. On the other hand, INPs used for phototherapy/magnetothermal therapy due to the characteristics of the optical/magnetic response of inorganic elements were reviewed in Section 6.2.

3.1. Structure of BBB and challenges during the progression of glioma

The BBB is constituted of brain capillary endothelial cells (BCECs), pericytes, astrocytes, tight junctions, neurons, and the basement membrane [31], which is a diffusion barrier that prevents the entry of substances from the blood into the brain [32]. The transendothelial electrical resistance of the BBB is estimated to be 8000 ​W/cm2 [33], which permits the BBB to separate blood from brain tissue while also selectively inhibiting the entry of certain toxic macromolecules into the brain circulation [34,35].

During the progression of glioma, the integrity of BBB structure also changes. In the early pathological period, BBB structure is typically intact, while it is gradually disrupted with tumor progression. However, some tumor lesions are encapsulated by endothelial cells, which can still prevent drugs from arriving at the tumor area [36,37]. Therefore, BBB penetration efficiency should be considered during all stages of glioma development. In addition, a drug's ability to target glioma cells is also important in improving the drug utilization ratio.

The BBB penetration ratio is the first challenge for glioma treatment. Many therapies have proven efficient in killing glioma cells in vitro; however, due to their inability to penetrate the BBB [38], systemic dosing fails to stop tumor progression in vivo, while excessive doses can cause general toxicity to the main organs. In addition, the poor solubility and short half-life of many therapeutic medicines in blood circulation limit their penetration of the BBB. In the aspect of molecular permeability [39], molecules greater than 400 ​Da cannot pass through the BBB, especially in highly water-soluble settings, unless a suitable specific transporter is present. Fat-soluble small molecules tend to be more BBB-permeable. For instance, the DNA-alkylating drug TMZ, which has poor solubility in water, crosses the BBB easily and is one of the few available treatments for GBM [40,41]. There is evidence that albumin barely crosses the intact BBB at all [42], so drugs that bind to plasma proteins may only cross the BBB through specific transporters. Some non-nanomedicine methods have been used clinically to solve the challenges of crossing the BBB and targeting gliomas. For example, arterial mannitol perfusion has been used to enhance the delivery of chemotherapy drugs to gliomas, since hypertonic solutions cause rapid contraction of BCECs, temporarily increase the endocytosis rate, and destroy tight junctions [43]. Another very promising and less-invasive approach to improve drug delivery is transcranial-focused ultrasound [[44], [45], [46]]; however, repeated stimulations are needed. Despite this, the survival rate for glioma is still very low and more efficient therapeutic methods are urgently required.

3.2. Fabrication strategies of theranostic INPs for combating glioma

A large number of INPs have been extensively studied for glioma treatment, due to their excellent physical properties and different targeting strategies. According to the functions, they can be used for drug delivery, imaging, phototherapy, radiotherapy, gene therapy, immunotherapy and magnetic hyperthermia.

3.2.1. BBB-crossing strategies based on INPs

Some INPs exhibit the ability to freely penetrate the BBB and passively target gliomas. For example, Sun et al. [47] synthesized a novel fluorescent carbon dot (CD-Asp) using d-glucose and l-aspartate via a straightforward thermolysis route. In vivo imaging of C6 glioma-bearing mice injected with CD-Asp via the tail vein was performed. The results indicated that CD-Asp could freely cross the BBB and accumulate at glioma tissue because of its high enhanced permeability and retention efficiency (EPR). Therefore, CD-Asp is an ideal sensitizer to guide non-invasive glioma fluorescent imaging in vivo.

Size, charge and shape are crucial influences on the uptake and BBB penetration abilities of INPs. In Chen's work [48], the size of MSNs was controlled by adjusting the hydrolysis rate and degree of condensation of the reactants, then loaded with doxorubicin (DOX) and modified with (polyethyleneimine) PEI-cRGD tumor-targeting polymers to enhance the anti-glioma effect (Fig. 3A). Human brain microvascular endothelial cells (HBMVEC) and human brain astrocytoma (U87) glioma cells co-culture were established as an in vitro BBB model to assess the BBB-penetrating ability of DOX@MSNs. Fluorescence imaging revealed the uptake of permeable drugs. DOX@MSNs showed higher BBB permeability and cellular uptake than free DOX. Furthermore, among DOX@MSNs with three particle sizes (20 ​nm, 40 ​nm, and 80 ​nm), DOX@MSNs of 40 ​nm particle size exhibited the maximum BBB transport rate and drug toxicity toward glioma cells. In view of this, the wrong particle size may be detrimental to BBB permeability. So, carefully altering the particle size of MSNs may be an effective strategy for combating GBM and overcoming the BBB barrier effect. Unsurprisingly, the surface electricity of INPs plays a crucial role in their brain delivery. Generally, positively charged nanoparticles are more accessible to cells, as adsorption-mediated transcytosis is triggered by electrostatic interactions between cationic targeting components and the negatively charged membranes of BCECs [49]. As described above, Chen [48] changed the charge of DOX@MSN INPs from negative to positive by PEI-cRGD modification. These changes enhanced the cellular uptake and stability of these nanoparticles in comparison with free MSNs. Shape is also an important influence on INPs uptake and penetration into the brain. Eswaramoorthy et al. [50] synthesized iron oxide (Fe3O4) nanoparticles with different morphologies (spheres, spindles, biconcaves, and nanotubes) and coated them with a fluorescent carbon layer derived from glucose (Fe3O4@C). The biconcave-shaped nanoparticles (Fe3O4@CBC) exhibited the highest U87MG cellular uptake ratio. In vivo studies further showed that the biconcave nanostructures localized predominantly in the nuclei of the brain cells of mice, whereas nanotube (Fe3O4@CNT) localized mostly in the cytoplasm. Particle shape affects cellular uptake due to shape-dependent cell surface adhesion [51] and endocytic pathways [52,53]. However, as the nuclear entry pathways of INPs are generally different from those of their cellular entry [54], it cannot be concluded that there is a direct relationship between shape and entry pathways. In a word, the biconcave morphology of Fe3O4@C enables nuclear-specific drug delivery to brain cells, thereby enhancing their enzymatic activity and corresponding gene expression. Nanotubes can be utilized to deliver molecules specifically into the cytoplasm. Hence, INPs with different morphologies can be selected for drug loading at different brain sites, or have different mechanisms of action and exert different functions.

Fig. 3.

Fig. 3

BBB-crossing strategies based on INPs used for efficient glioma treatment.A.DOX@MSNs; B. TAT-Au NP-Dox/TAT-Au NP-Gd; C. O-MWNTs-PEG-ANG. Panel A is adapted with permission from Ref. [48], copyright 2016 ACS Applied Materials & Interfaces; Panel B is adapted with permission from Ref. [58], copyright 2014 Small. Panel C is adapted with permission from Ref. [59], copyright 2019 Dalton Transactions.

Moreover, therapeutic efficiency is highly correlated with surface modification ligands, which can determine how nanoparticles cross the BBB and target glioma tissues. A classic example is polyethylene glycol (PEG)-coated Fe3O4 NPs [55]. A PEG coating on the surface can reduce non-specific protein adsorption and decrease reticuloendothelial system (RES) uptake of Fe3O4 NPs, thereby obtaining a longer circulation time and increasing the potential for BBB penetration.

In addition, studies have confirmed that surface modification by cell-penetrating peptides (CPP) can be an effective strategy for enhancing BBB penetration efficiency [56]. For instance, the trans-activating transcriptional activator (TAT) derived from HIV is used to deliver a wide range of cargoes across lipophilic barriers and can increase the permeability of brain endothelial cells by inhibiting occludin expression and cleaving occludin via matrix metalloproteinase-9 [57]. As described in the literature [58,59], by modification of the TAT peptide on the surface of Au and Carbon-based INPs, BBB permeability can be increased from 9.4% to 40.4% and from 0.6% to 2.9%, respectively (Fig. 3B and C).

3.2.2. Active targeting strategies based on INPs

More importantly, a series of active targeting ligands have been selected based on specific receptors and BBB transporters, which can also be used to enhance the targeting ability of the INP delivery system [[60], [61], [62], [63]]. At present, studies have reported INP delivery systems modified with Angiopep-2 (ANG), lactoferrin (Lf), transferrin (Tf), asparagine-glycine-arginine (NGR) peptide, and RGD peptide dimer (RGD2). Common INP delivery systems and their related target groups for glioma diagnosis and treatment are briefly described in Table 2.

Table 2.

Summary of the BBB-penetrating strategies of INPs.

Inorganic nanoplatform Connection methods BBB-crossing ligand Proposed transporter(s) Target-ing glioma BBB permeability Main cargo (es) Ref.
Drug loading Au TAT-Au NP-Dox/TAT-Au NP-Gd Hydrazone TAT peptide None NO 0.6% → 2.9% DOX [58]
An-PEG-DOX-AuNPs Hydrazone ANG LRP1 mono-targeting Not mentioned DOX [64]
Fe FeGd-HN@LF/RGD2 EDC/NHS Lf/RGD2 Lf receptor/integrin αvβ3 dual-targeting 13.6 ​± ​1.9% → 41.0 ​± ​0.9% CDDP [65]
Lf-Cur-PDNCs EDC/NHS Lf Lf receptor mono-targeting fourfold Cur [66]
Si MNP-MSN-PLGA-Tf NPs EDC/NHS Tf Tf receptor dual-targeting 8 ​→ ​33 DOX and PTX [67]
DOX-MNP-MSN-PF-127-Tf EDC/NHS Tf Tf receptor mono-targeting 4296 ​→ ​56,747 DOX [68]
ANG-PTX-PLGA-DOX-MSNs-75 EDC/NHS ANG LRP1 mono-targeting 2627 ​→ ​7497 DOX and PTX [69]
DOX@MSNs EDC cRGD peptide ανβ3 integrin mono-targeting 32.8% → 59.2% DOX [48]
DOX@MSN–SS–iRGD&1 ​MT PEG iRGD peptide None NO Not Mentioned DOX [70]
C O-MWNTs-PEG-ANG PEG ANG LRP1 dual-targeting Not Mentioned DOX [71]
TBCNT@OXA None TAT-PEI-Biotin copolymer None No 9.4% → 40.4% OXA [59]
Diagnosis/Treatment Au 131I–Au PENPs-CTX PEG、EDC Chlorotoxin (CTX) MMP2 mono-targeting Not Mentioned None [72]
Cd QD-MPA NHS T7 peptide Tf receptor mono-targeting Not Mentioned None [73]
NGR-PEG-QDs PEG NGR peptide CD13 mono-targeting Not Mentioned None [74]
QD-Apt PEG aptamer 32 EGFRvIII mono-targeting Not Mentioned None [75]
La (Er) NaErF4:Ce@NaAF4@NaLuF4 DCNPs SH ANG LRP1 mono-targeting 2.2-fold None [76]
Si RVG-PEG-AuNR@SiO2 MAL-PEG-NHS RVG29 peptide Nicotinic AchR mono-targeting Not Mentioned None [77]

∗EDC represents for 1- ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride; NHS represents for N-hydroxysuccinimide; SH represents for SH seal; MAL represents for maleimide; CDDP represents for cisplatin; Cur represents for curcumin; PTX represents for paclitaxel; OXA represents for oxaliplatin; MMP2 represents for matrix metalloprote-inase 2; EGFRvIII represents for epidermal growth factor receptor variant III.

Based on the active targeting strategies and the alternative ligands, numerous mono- and dual-targeting nanoplatforms were fabricated for the integrated diagnosis and treatment of glioma at different stages.

3.2.2.1. Mono-targeting INPs

ANG targeting ligands can enhance glioma treatment efficacy through interactions with low-density lipoprotein (LRP) receptor-related proteins. Gao and colleagues [76] investigated NaErF4:Ce@NaAF4@NaLuF4 down-conversion nanoparticles with Dye-brush polymer and angiopep-2 peptide (Er-DCNPs-Dye-BP-ANG) for use in imaging-guided surgery of orthotopic glioma (Fig. 4A). They employed focused ultrasound sonication, which can temporarily control the opening of the BBB and bind microvesicles to enhance the delivery of nanoprobes to gliomas [[78], [79], [80]]. The study demonstrated that U87 ​cells incubated with Er-DCNPs-Dye-BP-ANG had stronger fluorescence than cells from the Er-DCNPs-Dye-BP group. Then, Er-DCNPs-Dye-BP-ANG was injected intravenously for fluorescence imaging of subcutaneously transplanted gliomas. The tumor brightness of the ANG (+) group was much higher than that of the ANG (−) group. Moreover, the tumor-liver fluorescence ratio increased 2.2-fold in the presence of ANG. In conclusion, these results demonstrate the significant potential of Er-DCNPs-Dye-BP-ANG for fluorescence imaging-guided glioma surgery due to the excellent targeting ability of the ANG peptide to the LRP receptor.

Fig. 4.

Fig. 4

Mono-targeting INPs used for efficient glioma treatment. A. Er-DCNPs-Dye-BP-ANG; B. Lf-Cur-PDNCs; C. MNP-MSN-PLGA-Tf NPs; D. NGR-PEG-QDs. Panel A is adapted with permission from Ref. [76], copyright 2020 Nano Today. Panel B is adapted with permission from Ref. [66], copyright ​2016 Advanced Healthcare Materials. Panel C is adapted with permission from Ref. [67], copyright 2013 Biomaterials. Panel D is adapted with permission from Ref. [74], copyright 2016 Nanomedicine: Nanotechnology, Biology, and Medicine.

Many studies have found that lactoferrin receptor (LfR) is not only present on the BBB, but also on the cell surface of GBM [81]. Therefore, INPs coated by Lf-targeting ligands can enhance transport of nanodrugs across the BBB and accumulation at the glioma site via receptor-mediated transcytosis and endocytosis [82,83]. A magnetic polydiacetylene nanocarrier (PDNC) delivery system modified with Lf ligands was used to confirm that Lf ligands can increase BBB transport rates (4-fold) while targeting gliomas (Fig. 4B) [66]. Similarly, Tf has also been widely applied to enhance the cellular uptake of INPs, since Tf-receptors are overexpressed in the brain capillary endothelium and glioma cells, involved in receptor-mediated transcytosis [84,85]. Tf-conjugated magnetic silica PLGA nanoparticles (MNP-MSN-PLGA-Tf NPs) loaded with DOX and PTX are a promising carrier of dual drugs for brain glioma treatment, as they increase transport across the BBB and blood-tumor barrier (BTB; Fig. 4C) [67].

Another interesting study synthesized a unique nanoprobe by combining biotinylated NGR peptide with avidin PEG-coated CdSe/ZnS quantum dots (QDs) [74]. Although the heavy metal Cd2+ in the CdSe core is toxic to organisms, biocompatibility can be enhanced by coating the surfaces of CdSe/ZnS core/shell QDs with ZnS shells and PEG [86,87]. CD13 is a biomarker expressed in tumor blood vessels [88]. Compared with other tumor types, it has higher expression in gliomas and the tumor-associated neovascularization stage [89]. So, functional NGR sequences could target CD13-overexpressing tumors [90,91]. Based on this, a back-to-back BBB/BTB system was developed by seeding BCEC on the upper side of a filter membrane and astrocytes or C6 glioma cells on the lower side (Fig. 4D). Rapid infiltration occurred mainly in the first hour and then stabilized. After 5 ​h, more than 80% of the cells were labelled by the NGR-PEG-QDs in the BTB model. In summary, this work revealed that NGR-PEG-QDs could cross the BBB and target CD13-overexpressing gliomas. Hence, NGR-PEG-QDs can be used for fluorescence imaging to target glioma and tumor vascular systems.

3.2.2.2. Dual-targeting INPs

Dual-targeting inorganic nanosystems have more efficient BBB penetration and glioma targeting abilities, providing more efficient diagnosis and treatment for early-stage glioma patients with a relatively intact BBB structure. As a case in point, Jiang et al. [71] developed a dual-targeting drug delivery system based on PEGylated oxidized multi-walled carbon nanotubes (O-MWNTs) connected with Angiopep-2 (O-MWNTs-PEG-ANG) for the treatment of brain gliomas (Fig. 5A). O-MWNTs could accumulate in tumors and loaded DOX as a drug carrier. Brain fluorescence images showed that DOX-O-MWNTs-PEG (b) could be distributed in the brain more efficiently than DOX (c) and that DOX-O-MWNTs-PEG-ANG (a) could also enhance the distribution of drugs in the brain. In addition, fluorescence imaging of glioma-bearing brains indicated that DOX-O-MWNTs-PEG could obviously accumulate on gliomas and that DOX-O-MWNTs-PEG-ANG could further enhance glioma targeting. Accordingly, the combination of O-MWNTs-PEG and ANG constitutes an ideal dual-targeting drug delivery system for the diagnosis and treatment of early glioma patients with an intact BBB.

Fig. 5.

Fig. 5

Dual-targeting INPs used for efficient glioma treatment. A. TBCNT@OXA; B. FeGd-HN@Pt@LF/RGD2. Panel A is adapted with permission from Ref. [71], copyright 2012 Biomaterials. Panel B is adapted with permission from Ref. [65], copyright 2018 ACS Nano.

Similarly, CDDP-loaded Fe3O4/Gd2O3 hybrid nanoparticles with a conjugation of LF and RGD2 (FeGd-HN@Pt@LF/RGD2) were exploited by Chen's group [65]. FeGd-HN@Pt@LF/RGD2 were designed to cross the BBB by LF receptor-mediated transcytosis and be internalized into cancer cells via integrin αvβ3 (RGD2 receptor) binding, where Fe2+, Fe3+ and CDDP were released after endosomal uptake and degradation. Fe2+ and Fe3+ can directly participate in the Fenton reaction, while CDDP can indirectly produce H2O2 to accelerate it. The Fenton reaction accelerates the generation of reactive oxygen species (ROS), which induces cancer cell death. As shown in Fig. 5B, LF-mediated transcytosis and the use of RGD2 were crucial for the transportability of nanoparticles across the BBB, increasing it from 13.6 ​± ​1.9% to 41.0 ​± ​0.9%. The results demonstrate that FeGd-HN@Pt2@LF/RGD2 nanoparticles have a high Fenton efficacy on orthotopic gliomas.

In summary, a variety of factors (particle size, electrical properties, shape, surface modification, etc.) can influence the BBB-penetrating and tumor-accumulation abilities of INPs. In addition, anchoring antibodies and peptides to nanoparticles can enhance these abilities by interacting with receptors or transporter proteins on endothelial cells. All these factors must be carefully tuned to yield the most efficient inorganic nanoplatforms for the diagnosis and treatment of glioma.

4. Biocompatibility evaluations of INPs used for glioma treatment

Toxicity assessment of INPs used for glioma treatment is indispensable before biological application. Unlike other cancer treatments, due to structural and functional complexity of the brain, drugs used for glioma should be considered the avoidance of both neurotoxicity and systemic toxicity. The conventional method is to determine the in vitro neurotoxicity of INPs by measuring their biocompatibility with brain microvascular endothelial cells (BMECs), astrocytes and so on. In vivo neurotoxicity is evaluated by studying the BBB's integrity or the cognitive and behavioural levels of experimental animals. This section systematically reviews the cell-dependent toxicity, neurotoxicity, major organ damage, biodistribution and clearance of INPs involved in glioma diagnosis and treatment from both in vivo and in vitro perspectives.

4.1. In vitro biocompatibility

On cellular level, it is required to consider the influence of INPs on the integrity of the BBB; at least, the toxicity to BMECs, astrocytes and neurons. As summarized in Table 3, various INPs have been extensively studied in both normal nerve cell lines, as well as different glioma cell lines at different conditions to mimick physiological environment.

Table 3.

Cytotoxicity of INPs used for glioma treatment.

Glioma cell type Inorganic nanoplatform Glioma cell safety concentration Normal cell line (safety concentration) Incubation time Ref.
U87 (human brain astrocytoma) MNP-MSN-PLGA-Tf NPs 100 ​μg/mL bEnd.3 (100 ​μg/mL) 96 ​h [67]
GNS-ICG-BSA 0.5 ​mg/mL bEnd.3 (0.5 ​mg/mL) 48 ​h [95]
U251 (human glioblastoma) L-/D-Cys-CdSe/CdS DRs 50 ​μg/mL NHA (75 ​μg/mL) 24 ​h [99]
BV2 (mouse microglia) GQDs@CCM 200 ​μg/mL GMI-R1 (200 ​μg/mL)/Rat astrocytes cells (200 ​μg/mL) 24 ​h [100]
C6 (rat glioma cells) O-MWNTs-PEG-ANG 100 ​μg/mL BCEC (100 ​μg/mL) 24 ​h [71]
NGR-PEG-QDs 20 ​nM BCECs, astrocytes (20 ​nM) 96 ​h [74]
GL261 (mouse glioma cells) Au–OMV 200 ​μg/mL bEnd.3 (200 ​μg/mL)/C8D1A (200 ​μg/mL) 24 ​h [101]

The cytotoxicity of INPs after entering the cell can be mainly determined according to particle size and surface charges. The efficiency of endocytosis uptake is highly dependent on the size of the INPs. Generally, smaller particles enter cells more easily. Therefore, smaller INPs have better biocompatibility, which is beneficial to glioma diagnosis and treatment. Taking the widely employed iron oxide magnetic nanoparticles (IONPs) as an example, Chekhonin et al. [92] investigated the relationship between particle size and toxicity to human glioma cells (U251). They found that within 24 ​h incubation, there was no significant cytotoxicity and no difference between IONPs with different hydrodynamic diameters, even at 10−3 ​mol/L concentration. However, after 48 ​h, 40-nm IONPs showed better biocompatibility than 80-nm ones. As a result, it is crucial to tailor the size of INPs to optimise biocompatibility and therapeutic effects on glioma.

In addition to size, the surface charge of nanoparticles affects biocompatibility. It has been demonstrated that positively charged nanoparticles are more easily internalized into cells than neutral and negatively charged ones [93]. This can subsequently affect the endocytosis, passive targeting ratio and cytotoxicity, as described in Section 3.2. Lu's group [94] prepared cetuximab (C225)-encapsulated core-shell Fe3O4@Au magnetic nanoparticles (Fe3O4@Au–C225 composite-targeted MNPs). C225 is a monoclonal antibody targeting the EGFR, which is overexpressed in cancer cells. The adsorption of C225 shifted the zeta potential of INPs to a positive charge, enhancing their interactions with glioma cells and improving the internalization efficiency. Of all the concentrations tested, 0.5 ​mg Fe/mL was considered the optimal cell-safe concentration that ensures the killing of glioma cells without damaging normal cells. Ultimately, the dual magnetic and photothermal action mediated by Fe3O4 @Au–C225 composite-targeted MNPs showed remarkably improved synergistic therapeutic effects in glioma treatment.

Currently, coating and surface modification are the strategies most commonly employed for increasing the biocompatibility of INPs used in the diagnosis and treatment of glioma. Widely used surface modification molecules include bovine serum albumin (BSA), PEG, chitosan, chiral cysteine (Cys) and mercaptopropionic acid (MPA). For example, Wang et al. [95] prepared a non-toxic nanoprobe from gold nanostar-indocyanine green (ICG) incorporated in the BSA (GNS-ICG-BSA) to achieve SERS imaging-guided in vitro PTT in U87 glioma cells (Fig. 6A). BSA surface modification can prevent ICG-encoded GNS aggregation under physiological conditions, ensuring excellent stability and biocompatibility of the nanoprobe. Cytotoxicity evaluation of U87 glioma cells and bEnd.3 normal microvascular endothelial cells revealed that GNS-ICG-BSA SERS nanoprobes have satisfactory biocompatibility and low cytotoxicity, even at concentrations of up to 0.5 ​mg/mL. Furthermore, Xu et al. [96] synthesized hydrophilic PEG-chlorin e6 (Ce6)-chelated gadolinium-ion (Gd3+) nanoparticles (PEG-Ce6-Gd NPs) via a chelation and self-assembly process (Fig. 6B). PEG-modified nanoparticles present favourable biocompatibility and water solubility, prolonged circulation times and enhanced accumulation in tumor sites via the EPR effect. Biocompatibility results illustrate that proper cytotoxicity make PEG-Ce6-coated Gd-based NPs to be promising glioma nano-agents for photodynamic therapy (PDT) and contrast-enhanced MRI diagnosis. Another classic example is the chitosan-coated Fe3O4 nanoparticles with high heating efficiency performance developed by Kuanr and colleagues [97] (Fig. 6C). The colloidal stability of Fe3O4 NPs in a physiological medium can be improved by surface modification using biocompatible chitosan polymers [98]. Notably, the C6 glioma cells treated with chitosan-coated Fe3O4 NPs displayed significantly lower overall cytotoxicity than that of bare Fe3O4 NPs, ranging from 0 to 0.5 ​mg/mL over a 24-h testing period. Finally, complete glioma ablation was achieved at a relatively low alternating magnetic field and frequency that was within the clinical safety range. Similarly, Xu et al. [99] synthesized L-Cys and D-Cys modified CdSe/CdS DRs (dot-in-rods) via the post-ligand-exchange method, which exhibited fluorescence imaging-guided photodynamic/chemodynamic therapeutic effects in brain glioma (Fig. 6D). The cellular viability of U251 and U87 human glioma cell lines was maintained at > 90% after 24 ​h incubation with 50 ​μg/mL D-/L-Cys-CdSe/CdS DRs. In addition, chiral CdSe/CdS DRs showed lower toxicity to NHA (normal human astrocytes) than glioma cells in the low concentration range (0–75 ​μg/mL). After that, the cytotoxicity of MPA-modified DRs and QDs (sphere), as well as of L-Cys- and D-Cys-modified DRs, was assessed in U251 ​cells. The results illustrate that the chiral-cysteine coverage strategy can further improve the biocompatibility of CdSe-based QRs and QDs.

Fig. 6.

Fig. 6

Cytotoxicity of INPs used for glioma treatment. A. GNS-ICG-BSA; B. PEG-Ce6-Gd NPs; C. Chitosan-coated Fe3O4 NPs; D. D-and L-Cys-CdSe/CdS DRs; Panel A is adapted with permission from Ref. [95], copyright 2017 Nanoscale. Panel B is adapted with permission from Ref. [96], copyright 2021 Oncology Reports. Panel C is adapted with permission from Ref. [97], copyright 2021 Biomaterials Science. Panel D is adapted with permission from Ref. [99], copyright 2023 Materials and Design.

According to the needs of glioma diagnosis and treatment (imaging, optical therapy, drug delivery, etc.), we should not only select materials based on the characteristics of their inorganic elements but also consider improving their potential for targeting, reducing toxicity and being used in different modification strategies. This will promote the preparation of corresponding nanocomplexes with diagnostic and treatment functions.

4.2. In vivo compatibility

In biomedical applications against glioma, INPs are usually delivered into glioma animal models by intravenous or orthotopic injection. INPs generally circulate in the bloodstream and enter the major organs before being cleared by the metabolic organs.

Firstly, in terms of evaluating the effects of INPs on neurotoxicity, Cheon et al. [102] synthesized CNS-permeable INPs composed of a magnetic metal ferrite nanoparticle core and a cross-linked serum albumin (SA) surface coating (SA-MNPs; Fig. 7A). To evaluate the effects of SA-MNPs on BBB integrity, the BBB leakage was tested using Evans Blue, which does not penetrate the BBB under normal conditions. No brains from SA-MNP-treated animals showed any blue colour, indicating that BBB integrity is not influenced by SA-MNPs. This also confirms their biocompatibility with the BBB structure, which indicates that there are prospects to apply SA-MNPs in nervous system disease treatment.

Fig. 7.

Fig. 7

In vivo compatibility evaluation of INPs used for glioma treatment. A. SA-MNPs; B. QD-Apt; C. BSA-AuNPs; Panel A is adapted with permission from Ref. [102], copyright 2012 Chemical Communications. Panel B is adapted with permission from Ref. [75], copyright 2017 International Journal of Nanomedicine. Panel C is adapted with permission from Ref. [106], copyright 2015 RSC Advances.

Immunohistochemical methods such as haemoxylinoglobulin (HE) are generally used to evaluate damage by INPs to major organs in vivo. For example, Cheng et al. [75] constructed a novel quantum dot (QD)-labelled aptamer (QD-Apt) nanoprobe by coupling aptamer 32 (A32) to the QD surface (Fig. 7B). This increases the ability to target glioma cells. A32 is a single-stranded DNA capable of binding to the EGFRvIII, which is especially distributed on the surface of glioma cells. The toxicity of QD-Apt was examined in vivo through tail vein injection of 5 ​nM QD-Apt in male C57BL/6 mice. The main organs were subjected to histological analysis (HE staining) after treatment. No obvious tissue damage, necrosis or inflammation was observed in the brain or other main organs. The results indicate that the safe doses of Cd-based fluorescent probes for the diagnosis and treatment of glioma are 100 ​nM in vitro and 5 ​nM in vivo. At a safe concentration, the QD-Apt nanoprobe provides a novel and effective strategy for obtaining preoperative diagnosis, intraoperative resection and postoperative examination of gliomas.

The distribution and excretion of INPs in the main organs of the body are vital indicators of in vivo toxicity. Generally, the contents of the main inorganic elements in the blood and major organs are detected and analyzed by mass spectrum, including inductively coupled plasma optical emission spectrometry (ICP-OES) [103], ICP atomic emission spectrometry (ICP-AES) [104] or ICP mass spectrometry (ICP-MS) [105]. Tu's group [106] prepared BSA-capped Au NPs (BSA-AuNPs) as an efficient sensitizer for glioma radiation therapy (Fig. 7C). ICP-AES was used to investigate the level of Au in the blood and organs of mice after intravenous injection of 1.3 ​mg/mL BSA-AuNPs. Time-dependent analysis of blood Au levels first showed that rapid clearance of BSA-AuNPs occurred within the first 2 ​h. Then, the mice were continuously irradiated for 2 ​h and 24 ​h after intravenous injection of BSA-AuNPs. The ICP analysis of the various organs showed that there was no significant difference in Au levels between 2 ​h and 24 ​h. Finally, the long-term (20 day) biocompatibility of BSA-AuNPs was also evaluated by recording the weight of mice and performing HE staining of their main organs, which indicated no obvious damage. In brief, these results confirm that BSA-AuNPs have satisfactory biocompatibility as promising sensitizer candidates for glioma radiation therapy.

In vitro and in vivo studies have found numerous inorganic nanomedicines with great application prospects for glioma treatment. However, there is still very limited research work explored the bio-safety of INPs at pathological level, such as BBB integrity, brain structure damage and inflammation; as well as the long-term toxicity to brain functions, including cognitive function, memory, behaviour, etc. However, with the further development of nanomedicine and the need for transformation, brain damage should be evaluated more deeply before the use of nanomedicines.

5. Diagnostic strategies for glioma based on INPs

Due to the infiltrative growth characteristics of gliomas, they have no clear boundary with normal tissue; thus, gliomas cannot be cured by surgery alone. This means that we urgently need to develop clear imaging tools and contrast agents to surmount the current treatments and diagnosis limitations. Inorganic nanomaterials are excellent PSs and contrast agents and can realize early tumor diagnosis, image mediation and real-time treatment monitoring. Hence, they have been widely used in the biomedical imaging and diagnosis of gliomas. A variety of physical, chemical and biological diagnostic methods, including MRI, computed tomography (CT), positron emission tomography (PET), PAI and fluorescence imaging, have substantially benefited the diagnosis of tumor outcomes.

5.1. Magnetic resonance imaging

MRI is a universal standard for clinical diagnosis and preoperative localization of gliomas. Because of its high spatiotemporal resolution (0.2–1 ​mm), high sensitivity (10−3–10−5 ​mol/L), non-invasiveness, and multi-directional, multi-sequential and multi-parametric imaging capabilities, it is particularly suitable for the diagnosis of intracranial diseases [107]. The main obstacles to glioma diagnosis are the unclear boundaries of malignant gliomas and the efficiency of the magnetic material in penetrating the BBB and targeting glioma cells [65,108,109]. The aggressive growth of gliomas often results in enlarged blood vessels and abnormal leakage of endothelial cells, as well as impaired lymphatic drainage [110]. However, based on their high-EPR effects, INPs coupled with active targeting molecules can greatly improve the effect of nuclear magnetic imaging of glioma lesions [108]. Among the nanomaterials used as magnetic sensitizers, IONPs have been extensively studied because of their high longitudinal and transverse relaxation values, superior imaging performance, and low cytotoxicity [111]. Numerous works have confirmed that iron oxide nanoparticles have promising applications as MRI T1/T2/T2∗ contrast agents [[112], [113], [114]]. IONPs can enhance the cellular internalization of tumor cells and active phagocytes to improve the visualization of intracranial tumors within healthy brain tissue and slow down the clearance of tumor sites [115] so as to better depict glioma boundaries [116]. Yang et al. [117] reported that PEG-coated IONPs have a long circulating lifetime in the brain of 9 ​L glioma rats, and their enhanced ability to target glioma tissues was confirmed by MRI and histological analysis (Fig. 8A). In Yang and his colleagues’ work, PEG-coated IONPs exhibited high colloidal stability, which improved magnetic capture and retention time at the tumor site. In subsequent biodistribution studies, they also demonstrated that PEG-MNPs selectively enhance brain tumor targeting by improving the circulation lifetime of MNPs [118].

Fig. 8.

Fig. 8

Diagnostic strategies for glioma based on INPs. A. PEG-IONPs; B. Au PENPs; C. MSNs; D. S–MoS2 Nanosheets; E. PEGNIO/Qds/MIONS/Tf; Panel A is adapted with permission from Ref. [118], copyright 2011 Elsevier. Panel B is adapted with permission from Ref. [72], copyright 2019 Springer Nature. Panel C is adapted with permission from Ref. [139], copyright 2014 American Chemical Society. Panel D is adapted with permission from Ref. [149], copyright 2016 Wiley-VCH. Panel E is adapted with permission from Ref. [162], copyright 2019 Elsevier.

In addition to the iron mentioned above, Marc-Andre’ Fortin et al. [119] developed ultra-small gadolinium oxide nanoparticles to label GL-261 GBM polymorphic cells and subsequently localize them in vivo for tumor visualization by using MRI. A relaxometric study showed that the relaxation ratio (r2/r1) of DEG-US-Gd2O3 is suitable for T1-weighted imaging [120,121]. The designed nanostructure contains about 400 Gd ions per 3-nm diameter nanoparticle, producing a signal increase at least 200 times more specific per unit of contrast than that of Gd-DTPA chelates. Conjugates such as DTPA, DOTA and DTPA-BMEA are used to coat Gd, since direct overexposure to gadolinium is toxic [122]. Not only that, the tumor retention time of this material is longer than that of Gd-DTPA. Compared with the Gd-DTPA used by de Stasio [123], Fortin's procedure required only 0.6 ​mg of Gd/mL DEG-US-Gd2O3 and 8 ​h incubation to reach the higher residual concentration of Gd, which was at least six times that of Gd-DTPA. During in vivo experiments, they used the chorioallantoic membrane of chick embryos as a tumor growth model and studied it using a 1.5 ​T clinical MRI scanner. At 14, 38, and 84 ​h after inoculation, the labelled cancer cells were clearly seen as they showed obvious contrast with the unlabelled normal tissue cells. GL-261 glioma cells appeared bright on the T1-weighted MRI images, indicating that the very-high-concentration Gd was effectively internalized and retained in the cells. This also laid a foundation for the development and use of Gd-based NMR contrast agents. Other elements commonly used for MRI of glioma include Fe, Au, Mn, Ce, Gd and Dy, which also have high imaging sensitivity and are good choices for use as contrast agents [58,119,124,125].

5.2. Computed tomography and positron emission tomography

CT and PET are widely employed modalities for intracranial tumor diagnosis. They can also be used to observe tumor images inside the substantial organs. However, the diagnosis of glioma still lacks sufficient accuracy. This may be related to early neuroinflammation, late radiation necrosis, or tumor recurrence [[126], [127], [128]]. Inorganic nanomaterials, especially magnetic nanoparticles, semiconductor materials and metal elements with high atomic numbers such as Au [72], Ag [129], Yb [130] and Bi [131], are reported to act as enhancers for CT and PET scanning and diagnosis of glioma due to their high X-ray absorption capacity, favourable biocompatibility and excellent optoelectronic properties [132,133].

Zhao et al. [72] designed a multifunctional targeting nanoprobe from Ref. [131]I-labelled CTX-functionalized Au PENPs ([131]I–Au PENPs-CTX). The surface of polyethylenimine (PEI) was modified with 3-(4-hydroxyphenyl) propionic acid-OSu and CTX to entrap AuNPs. CTX, a 36-amino-acid peptide with strong affinity for matrix metalloproteinase-2 (MMP-2) on glioma cells, was used as a targeting ligand and the radionuclide [131]I was used to label CTX [134,135]. In vitro experiments conducted on the C6 cells found that, compared with Omnipaque (a small-molecule CT contrast agent used clinically), the CT images of Au PENPs-CTX were brighter. With increases in gold and iodine concentration, the CT values were further enhanced. The X-ray attenuation characteristic of this nanoparticle was stronger than Omnipaque, while its high X-ray attenuation was associated with favourable CT imaging capability, demonstrating its excellent application potential [136]. In vivo experiments were consistent with in vitro evaluations, and quantitative measurements 8 ​h after injection showed that tumor CT values were 1.7-fold higher in mice treated with Au PENPS-CTX than in those treated with Au PENPs (Fig. 8B). To sum up, AuNPs were used as a sensitizer to link CTX with enhanced glioma targeting. This provides better in vitro and in vivo imaging performance in C6 glioma mouse models, highlighting the superior application prospects of such nanoparticles in glioma-targeted diagnosis.

PET is a powerful biomedical imaging technology with high sensitivity and quantitative accuracy. It can provide additional information about tumor metabolic processes that can aid in brain tumor differential diagnosis, grading, molecular typing and the differentiation of therapeutic effects from disease recurrence [137,138]. Cai et al. [139] developed a new multifunctional nano-platform, VEGF121-conjugated mesoporous silicon dioxide (MSN), as a theranostic nanomedicine for glioma. In the composite nanostructure, VEGF121 (as a radiolabelling ligand) can target VEGF and provides early and sensitive lesion detection by PET [140]. MSN was used as a platform for drug loading, facilitating targeted PET detection to enhance the imaging of glioma lesions and achieve a better drug-release effect [64].Cu is also a radio tracer suitable for PET imaging. Cai et al. performed serial whole-body PET on U87 tumor-bearing mice at multiple time points (0.6, 3, 6 and 22 ​h) after injection of targeted and untargeted groups (Fig. 8C). In vivo PET imaging showed that tumor accumulation in the targeted MSN group was 3-fold higher than in the non-targeted MSN group, while the specific accumulation occurred in the targeted MSN group as early as 0.5 ​h. The targeted inorganic nanomaterial system described above for radionuclide-based PET imaging provides high sensitivity and the ability to quantitatively analyse whole-body images, which are essential to treatment optimisation.

As different inorganic elements have different physicochemical properties, it is very important to select appropriate materials and construction methods for the preparation of functional nanoplatforms.

5.3. Photoacoustic imaging

PAI is a promising non-invasive optical imaging method that uses light pulses to excite exogenous molecular contrast agents to produce ultrasonic waves that can be detected by transducers, generating 3D images of a region of interest [141,142]. The excellent optical properties and surface functionalization features of inorganic nanomaterials have been used to develop exogenous contrast agents for enhanced PAI. Among them, the most commonly used nanomaterials are gold derivatives with favourable SPR absorption mechanisms [141]. Examples include silicon-coated gold NPs for PAI of surface gliomas, hollow gold nanospheres with glioma-targeting properties for PA molecular imaging, and hollow gold nanospheres that generate intense photoacoustic signals and induce effective photothermal ablation of gliomas using hybrid imaging modalities [[143], [144], [145]]. Transition metals also have potential PAI applications. Research shows that molybdenum disulfide (MoS2) nanosheets, as a representation, have better near-infrared (NIR) absorbance, stability and biocompatibility than gold and graphene derivatives [[146], [147], [148]]. Based on this property, Zheng et al. designed single-layer (S–MoS2) nanosheets and used them to visualize glioma cell boundaries and tumor locations via PAI [149]. A custom-built acoustic-resolution PA microscopy (AR-PAM) system was used to study the PA effect of MoS2 nanosheets [150]. Compared with few-layer (F–MoS2) and multi-layer (M ​− ​MoS2), S–MoS2 nanosheets produced longer amplitudes and higher signal intensities. This is because the elastic modulus of MoS2 nanosheets increases with decreases in the number of layers and is related to the PA intensity, suggesting high application capability [[151], [152], [153]]. As shown in Fig. 8D, effective tumor retention and PA contrast sensitivity of S–MoS2 were demonstrated in a U87 glioma-cell-bearing mice model. The PA signal at tumor sites increased with time after intravenous injection of S–MoS2. Moreover, in vitro PAI after tumor resection also confirmed that the PA signal at 24 ​h after injection came from tumor tissue. Quantitative experiments showed that, compared with the endogenous PA signal before injection, the signal intensity increased about 54-fold 24 ​h after injection of S–MoS2. Overall, S–MoS2 nanosheets are a potential contrast agent for highly sensitive PAI of subcutaneous and orthotopic brain tumors in mice. Other inorganic nanomaterials, such as bismuth, cuprum, as well as transition metal-based oxides, all have PAI properties [[154], [155], [156], [157]]. However, there are few reports on their application to glioma medicine and further research is expected.

5.4. Fluorescence imaging

Some inorganic QDs with fluorescence imaging properties have promising prospects in the diagnosis and treatment of glioma by fluorescence imaging. Because of the infiltrative nature of gliomas, it is difficult to minimize the number of cancer cells and preserve normal brain tissue intraoperatively. Fluorescence imaging materials enable lesions to fluoresce, helping the surgeon to detect their edges and operate on them [158]. However, current reported fluorescein materials, such as 5-aminoketoglutarate and sodium fluorescein, suffer from low quantum fluorescence rates and poor photostability, which limit their clinical applications. Semiconductor QDs, usually composed of CdSe cores and ZnS shells [159], are widely studied in glioma fluorescence imaging [160,161]. According to this property, Mansur et al. prepared ZnS@CMC nanohybrids for GBM fluorescence imaging. In this work, zinc fluorescent quantum dots (ZnS-Qds; a fluorescent nanoprobe) were chemically modified by carboxymethylcellulose (CMC, a water-soluble capping ligand and bio-functional layer) and coupled with the anticancer drug adriamycin (DOX) to modify drug release. This provided a dual function of bio-imaging and killing of U87 glioma cells [162]. The results show that blue light released by QDs was concentrated in the cytoplasm 2 ​h after injection, while DOX-related intrinsic fluorescence was mainly concentrated in the nucleus (Fig. 8E). Likewise, Zhang et al. reported the use of high-quality CuInS2/ZnS QDs—with CuInS2 as the core and ZnS as the shell—for CD133+ glioma fluorescence imaging. The anti-CD133 monoclonal antibody pQDs-CD133 ​mAb was attached to QDs for tumor cell targeting to achieve precise targeted fluorescence imaging of glioma stem cells [160]. In short, the combination of appropriate targeting ligands with fluorescent nanoparticles has promising application prospects in the accurate fluorescence imaging of gliomas.

Various imaging methods for glioma diagnosis using nanoprobes have advantages and disadvantages. MRI is a non-invasive technique that can demonstrate the macroscopic structure of a tumor without causing X-ray damage. This makes it an ideal imaging test that avoids the damage associated with other tests. However, it has some inherent problems, such as poor spatial resolution and long acquisition and scan processing times. On the other hand, CT imaging has high spatial resolution but is not ideal for assessing brain tumors due to its poor soft tissue contrast. PET imaging is highly sensitive but requires a radioactive tracer, which increases the risk of cancer in humans. As a hybrid imaging modality, PAI combines the high contrast and spectral-based specificity of optical imaging with the high penetration depth of ultrasound imaging. This makes it ideal for deep tissue resolution and provides multi-scale, multi-dimensional photoacoustic image information. However, it does not enable better functional imaging and has lower image contrast. Fluorescence imaging has high detection sensitivity, an extremely wide range of applications, strong signal intensity, and the ability to label samples with multiple colors. However, it has the disadvantage of limited penetration and low spatial resolution.

Nanoscale platforms have now been developed to assist in the identification and diagnosis of tumors in clinical research. The Food and Drug Administration (FDA)-approved 30 ​nm MNPs have been used to predict the co-localization of therapeutic NPs with tumors and to quantify MNPs levels using the amount of change in the mean T2 profile [163]. Gadolinium-based contrast agents and other low molecular weight gadolinium chelates, such as Gd-DTPA and Gd-DOPA, have also been used in clinical practice [164,165]. In addition, nanoprobes can be tracked by linking peptides and antibodies that interact with glioma-specific molecules. Nanoprobe tracers that are capable of specifically binding to integrin αvβ3, vascular endothelial growth factor receptor, or IDH-mutations (glioma-specific molecules) have been successfully studied in clinical glioma patients in preliminary studies, paving the way for more nanoprobes to enter the clinic [166]. Overall, the development of nanoscale platforms and the use of imaging modalities can aid in the identification and diagnosis of glioma in clinical research. However, further efforts are needed for the full clinical translation of INPs-based imaging for glioma diagnosis. It is worth noting that INPs-based systems, due to the convenience of nanoscale platform preparation, the ability of different inorganic elements to be doped, the advantages of surface modification, and the addition of biological ligands, can simultaneously obtain multi-modality imaging technologies. The complementary advantages between different imaging modalities can provide more accurate information for the diagnosis of glioma, as detailed in 6.3.1.

6. Therapeutic strategies for glioma based on INPs

In clinical practice, gliomas are mainly treated by surgical resection. For unresectable gliomas and postoperative adjuvant treatment, chemotherapy, radiotherapy and gene therapy are mainly adopted. Traditional therapies involving INPs have been proven to enhance therapeutic targeting as well as therapeutic efficacy. On the other hand, the development of INPs has also created new strategies for glioma therapy. The next section introduces traditional methods that are enhanced by INPs, new glioma treatment strategies guided by INPs, and multi-functional nanoplatforms for guided synergistic diagnosis or treatment.

6.1. INPs-enhanced strategies in conventional glioma therapy

INPs can cross the BBB and have excellent targeting properties after connecting with targeting groups, which can greatly improve traditional glioma treatment methods. For example, chemotherapy and gene therapy can be effectively carried out by encapsulating chemotherapeutic drugs or gene interference sequences, while INPs can also be used as radiosensitizing agents to improve the efficacy of radiotherapy.

6.1.1. INPs-enhanced chemotherapy

Chemotherapy is still one of the standard treatments applied after maximal safe surgical resection. Despite some breakthroughs [167], glioma therapeutic efficacy remains fairly poor due to inadequate medication at lesion sites. INPs can be employed as drug carriers for glioma chemotherapeutics, enabling more precise and superior therapeutic effects. This is because nanomedicine can be used to transport unsuitable molecules into the brain by binding them or wrapping them in an appropriate vehicle. Therefore, chemotherapy mediated by nanomaterials can provide targeted drug delivery that overcomes the shortcomings of traditional chemotherapy. The combination of chemotherapy medications with INPs [59,124] that can penetrate the BBB can increase biocompatibility while maintaining or promoting the efficacy of chemotherapeutic drugs. Common targeting molecules that can help cross the BBB are shown in Table 2. There are many drugs that can be used for nanomaterial delivery, such as PTX [168], tetrandrinev [169], oleanolic acid [170], DOX [171], CDDP [172], methotrexate [173] and Cur [174]. Among these, DOX—a typical chemotherapeutic agent—is an anthracycline with broad-spectrum anticancer activity that is widely employed in clinical use, including glioma treatment [175,176]. Some of the many carriers suitable for DOX, and their effects, are described next.

Mansurand and colleagues developed innovative supramolecular complexes (ZnS@CMC-DOX) composed of an inorganic ZnS quantum dot core (ZnS-QD) biofunctionalized with a CMC polymer shell (ZnS@CMC) and conjugated with the DOX anticancer drug (Fig. 9A) [162]. Non-toxic binary ZnS semiconductor QDs have been designed to construct drug delivery systems [177]. An aqueous medium at pH ​= ​5.5 ​± ​0.2 (CE ​> ​99%) was selected to simulate the acidic microenvironment of cancer cells [178]. ZnS@CMC-DOX nanoconjugates exhibited a drug release rate of approximately 45% after incubation for 2 days, while the ‘free DOX’ release rate exceeded 80%. As a result, the drug release rate was slower after loading into the nanocomposites, which could increase the duration and action time of the drug and avoid the toxic side effects caused by rapid release. Other evidence illustrates that ZnS@CMC nanoconjugates without an anticancer drug load were not cytotoxic to normal cells and brain cancer cells. After loading with DOX, ZnS@CMC-DOX hybrid complexes retained a lower DOX cytotoxicity than free DOX, which was reflected in the following results. The cell viability responses of normal and tumor cells exceeded 70% after incubation of ZnS@CMC-DOX for 6 ​h. Cell viability was further reduced to 38% with the extension of culture time (24 ​h). Meanwhile, the cell viability of the free DOX group decreased to about 35–40% after 6 ​h. Consequently, the above studies prove that the ZnS@CMC-DOX nanocarrier can decrease the cytotoxicity in the short term and improve the sustained release effect of DOX. This reduces damage to normal tissue but does not affect the simultaneous cytokilling efficacy of DOX, making it an ideal carrier for drug delivery.

Fig. 9.

Fig. 9

Applications of INPs in glioma chemotherapy. A. ZnS@CMC-DOX; B. PST-GNP-DOX NPs; C. TMZ/MNPs-FA. Panel A is adapted with permission from Ref. [162], copyright 2019 Colloids and Surfaces B: Biointerfaces. Panel B is adapted with permission from Ref. [179], copyright 2019 Nanomaterials. Panel C is adapted with permission from Ref. [182], copyright 2021Nanomedicine.

GNPs are an attractive drug carrier candidate due to their reasonably low cytotoxicity, adjustable surface properties and stability under in vivo conditions. Komeri et al. [179] synthesized GNPs capped with galactoxyloglucan PST001 isolated from the seed kernels of ripened Tamarindus indica fruit. This was employed as a DOX nanocarrier and named PST-GNP-DOX NPs (Fig. 9B). PST001 has innate antitumor and immune-potentiating functions and was used as a GNPs reducing and stabilizing agent without introducing any environmental toxicity or biohazard [180]. The efficacy of PST-GNP-DOX NPs to glioma cells was confirmed using apoptotic staining. It is worth highlighting that the cell density of glioma cells treated with PST-GNP-DOX NPs was significantly lower than in cells treated with free DOX, especially at low doses. More importantly, PST-GNP-DOX NPs also had a longer mean residence time in plasma after intravenous and intraperitoneal administration. Meanwhile, PST-GNP-DOX NPs show less accumulation in the heart, lungs and spleen and a more uniform distribution in the brain, which suggests that PST-GNP-DOX NPs could escape from the RES. Neutral nanoparticles with an average diameter of 100 ​nm have been found to be able to evade the RES and mononuclear phagocyte system [181]. PST-GNP-DOX NPs might successfully escape from the RES when they have a mean hydrodynamic diameter of 133 ​nm and a surface charge that is nearly neutral in this circumstance. Besides, PST-GNP-DOX can greatly inhibit the growth of DOX-resistant U87 tumor cells compared with free DOX. Not only that, PST-GNP-DOX NPs also decreased the expression of BCRP efflux protein, which was overexpressed in drug-resistant tumor cells. At the same time, PST-GNP-DOX NPs decreased the expression of NF-κB associated with tumor progression, which may explain the effect of PST-GNP-DOX NPs to overcome drug-resistant gliomas. However, the specific molecular mechanism has not been explored. In brief, PST-GNP-DOX NPs are promising candidates for delivering DOX to glioma tumors.

Afzalipour et al. [182] developed TMZ-loaded magnetic nanoparticles conjugated with folic acid (TMZ/MNPs-FA) to create thermosensitive magnetic NPs (Fig. 9C). To find out how MNPs-FA affected drug release under magnetic and thermal circumstances, they first demonstrated that it could slow down the TMZ release rate compared with free TMZ [183]. The drug release rate increased with increasing temperature from 37 ​°C to 43 ​°C. Within the first 8 ​min of exposure to an alternating magnetic field (AMF), the drug release rate rose dramatically to 52.1 ​± ​2.3% but fell sharply once the AMF was turned off. In the absence of AMF, the release rates were 30% at 43 ​°C and <2% at 37 ​°C. Nanoparticles exposed to a magnetic field generate heat via Neal relaxation phenomenon [184]. Thus, MNPs-FA are heat-sensitive, which can regulate drug release. The results above show that TMZ/MNPs-FA nanoparticles are relatively stable at physiological temperatures (37 ​°C), while under the influence of AMF, the heat generated increases the drug release rate by affecting the polymer packing. This research extensively studied the biomedical application prospects of MNPs-FA in glioma treatment. MNPs-FA demonstrate potential for loading TMZ and transporting it across the BBB, while also showing excellent potential for remote-controlled drug release through heat induction.

Overall, INPs-based delivery carriers were proven that can enhance glioma chemotherapeutic efficacy by improving both the BBB-penetration efficiency and glioma-targeting ability. In addition, some INPs-base carriers can also control the release rates of chemotherapeutic drugs to prolong their action time. However, the long-term biological toxicity of various nano-delivery carriers, as well as the specific mechanisms of clearance and metabolism from the brain are worthy of further investigation.

6.1.2. Radiotherapy

Radiation therapy is a major strategy and common adjunct in the treatment of GBM. One of the major challenges in radiotherapy is minimizing radiation damage to healthy tissue without sacrificing the therapeutic effect on tumor cells. Clinical and experimental studies have revealed many treatment failures and unsatisfactory outcomes due to the radiation resistance of GBM [185]. The introduction of nanomedicine, such as by utilizing INPs as a radiosensitizer, offers an opportunity to enhance the therapeutic index of radiation therapy [186]. In addition, elevated concentrations of INPs in cancer cells might result in substantial accumulation of radiation energy at tumor regions compared with surrounding healthy tissue [187].

Liu et al. [188] compared the radiation-sensitizing effectiveness of AuNPs and AgNPs on glioma (Fig. 10A). The effects of AuNPs and AgNPs in combination with radiation on colony formation by U251 ​cells showed that the survival rates of the cells decreased sharply with increasing doses of 6 ​MV X-rays. The colony-forming assay is the gold standard for detecting radiosensitivity [189]. Notably, AgNPs showed the highest radiosensitizing activity, followed by AuNPs. This might be on account of AgNPs having stronger proapoptotic activity and upregulated autophagy levels. The results also showed that treatment of INPs with or without radiation therapy had little effect on noncancerous cell lines; thus, the prepared nanomaterial-mediated radiosensitization was tumor cell-specific. Next, radiation therapy was performed after the injection of AgNPs or AuNPs intratumorally in a glioma mice model. Kaplan-Meier survival plots confirmed the in vivo anticancer effects of AgNPs. Clearly, this work demonstrates the application prospects of AgNPs as highly effective nano-radiosensitizers for the treatment of glioma.

Fig. 10.

Fig. 10

INPs-enhanced strategies in conventional glioma therapy. (A–B). The application in radiotherapy; A. Ag/AuNPs; B. La2O3 NPs. (C–D). The application in gene therapy: C. QD-PEI; D. pSiNPs. Panel A is adapted with permission from Ref. [188], copyright 2016 International Journal of Nanomedicine. Panel B is adapted with permission from Ref. [190], copyright 2020 Scientific Reports. Panel C is adapted with permission from Ref. [194], copyright 2017 Frontiers in Pharmacology. Panel D is adapted with permission from Ref. [196], copyright 2018 Journal of Nanobiotechnology.

Lu and colleagues [190] investigated the augmented radiation effects of lanthanum oxide (La2O3) NPs on glioma via cellular ROS mechanisms (Fig. 10B). Due to its special chemical characteristics, lanthanum is cytotoxic to cancer cells and is able to enhance existing anti-cancer therapies. After intravenous injection, La2O3 NPs can reach the brain and target GBM cells through endocytosis, then dissociate to produce cytotoxicity, enhancing the therapeutic effect of radiation therapy. Radiation therapy with La2O3 NPs significantly decreased the amounts of colony-forming units (CFUs) and increased the production of ROS in GBM patient-derived cell lines (RN1, G53 and WK1). There was evidence that the amount of CFUs recovered after pre-treatment with a ROS scavenging agent (ROSSc) for 1 ​h. Notably, La2O3 NPs were associated with markers of ROS-dependent endogenous and exogenous apoptotic pathways, as well as involving direct DNA damage and autophagy pathways. These may be the mechanisms by which La2O3 NPs sensitize GBM to radiotherapy. These findings imply that cytotoxic La2O3 NPs can sensitize GBM cells to radiation therapy.

The acceptable intensity of radiation for an organism is limited. It is important to note that all of the mentioned INPs have achieved a radiotherapy-enhanced effect, thereby reducing the radiation hazard. Studies have reported that radiation therapy can cause irreversible CNS damage, so radiation therapy still needs to be chosen carefully in clinical practice. Therefore, safer and more effective radiosensitizers need to be urgently developed and INPs will be one of the options.

6.1.3. Gene therapy

Gene therapy has attracted extensive attention over the past few years due to its novel and effective cancer treatment. Among gene therapy techniques, gene silencing based on RNA interference (RNAi) has emerged as a potential strategy for biological therapy [191]. In addition to therapeutic targeting, another crucial factor for successful gene therapy is obtaining an ideal delivery vehicle with safe and high-efficiency gene knockdown. Since small interfering RNA (siRNA) molecules have difficulty traversing cell membranes and are susceptible to serum nucleases during blood circulation, they can be encapsulated in nanocarriers. This obtains excellent biocompatibility and easy intracellular uptake, enabling efficient knockdown of oncogenes or other related genes in vivo [192,193]. In recent years, the application of INPs greatly improve the effectiveness of glioma gene therapy via enhanced tumor targeting (connecting targeted groups) and BBB permeability.

Yong's group [194] fabricated cadmium sulphoselenide/zinc sulfide quantum dot (CdSSe/ZnS QD)-based nanocarriers conjugated with PEI to form a stable nanocomplex (QD-PEI). This was used to load siRNA specifically to GBM cells by targeting human telomerase reverse transcriptase (TERT). TERT is the catalytic protein subunit of telomerase, which is crucial for maintaining telomerase activity in cancer cells [195]. The excellent optical properties of QDs enable real-time monitoring of the siRNA delivery process. The expression of TERT mRNA and protein in human glioma cell lines (U87/U251) transfected with QD-PEI-TERT siRNA nanocomplex was found to be significantly inhibited, as shown in Fig. 10C. This demonstrates the efficiency of gene silencing and provides a stable platform for gene therapy applications. Next, the author investigated the targeted gene therapy ability of a QD-PEI-TERT siRNA nanocomplex. Comparing with controls, tumor cells were considerably fewer in a QD-PEI-TERT siRNA treatment group. Hence, the QD-PEI developed in this study can successfully deliver siRNA interference sequences, which is anticipated to be a candidate delivery vector for glioma gene therapy.

Voelcker et al. [196] established porous silicon NPs with a PEI capping (PEI-pSiNP/siRNA), which enables high-capacity loading and delivery of siRNA for GBM gene therapy. Overexpression of multidrug resistance-associated protein 1 (MRP1) plays an important role in chemotherapy resistance in GBM, leading to lethal consequences [197]. As shown in Fig. 10D, the INPs successfully downregulated MRP1 expression by 30%. Compared with untreated controls, PEI-pSiNP/siRNA treatment significantly depressed the expression level of Ki67 (a marker of cell proliferation) [198] in U87 ​cells, which might lead to the inhibition of GBM proliferation. Furthermore, their results indicated that in vitro observations are translatable in vivo. In summary, efficient MRP1-siRNA delivery using PEI-pSiNP/siRNA might achieve a dual-GBM therapeutic role of chemotherapeutic sensitivity and tumor suppression.

As a safe and efficient gene knockout vector, INPs can be used for adjuvant therapy of glioma in many aspects, such as the knockout of the mentioned above genes that maintain telomerase activity of cancer cells or play an important role in GBM chemotherapy resistance. Therefore, such targeted gene therapy to promote glioma treatment will be an interesting endeavor.

6.2. Novel glioma therapeutic strategies guided by INPs

INPs have provided new treatment methods for glioma due to their unique properties. On the one hand, INPs can be used as immunogenic cell death (ICD) inducers to improve the efficacy of immunotherapy. On the other hand, they can be used as PSs or photothermal agents (PTAs) for the PDT or PTT of gliomas. It is also possible to treat gliomas with magnetic hyperthermia in the presence of AMF. However, these novel therapies are still mainly the subject of research and have not been clinically applied.

6.2.1. Phototherapy

Tumor phototherapy has attracted particular attention because it has greater efficacy and less invasiveness than conventional cancer therapies [[199], [200], [201]]. PDT and PTT are the most prevalent phototherapy strategies, both of which are effective in destroying tumor cells by producing ROS or providing selective local heat via laser absorption [109]. PDT is a minimally invasive procedure that combines PSs, a specific wavelength of laser and molecular oxygen, which might be selectively retained by pathological tissue and not normal tissue when administered to patients [202,203]. PDT-mediated tumor death is generally an active death process, occurring mainly by autophagy and apoptosis [204]. The temperature range of PTT is 41–47 ​°C, which can achieve the purpose of tumor ablation by inducing the denaturation of proteins in tumor cells and causing their irreversible death [[205], [206], [207]]. However, due to the complexity of the brain structure, high temperatures are difficult to tolerate. Therefore, many scholars have proposed using low-temperature phototherapy at about 41–43 ​°C for the PTT of gliomas. In addition, if the nanocomposites have a drug-delivery function, the photothermal effect will also lead to drug release. Then, the nanocomposites can be combined with phototherapy to achieve photothermal and chemotherapy-enhanced therapeutic effects and regulate the drug release rate [208].

INPs have become essential materials in the field of glioma phototherapy due to their physical properties, and high physiological stability and biocompatibility. Furthermore, NIR lasers possess high bio-penetration and retention properties [209], while the rapid development of nanotechnology has made it possible to apply NIR-excited lasers in both in vitro and in vivo therapies. The skull and scalp can severely attenuate laser transmissions so that only a small proportion of the energy reaches the glioma site. Therefore, during glioma phototherapy, INPs excited by long wavelengths (>900 ​nm) are better for PDT/PTT treatment of glioma than traditional PSs/PTAs [210].

Currently, various INPs, such as noble metal nanoparticles [77,211,212] and graphene QDs [100] have been employed for NIR laser-mediated PDT/PTT of glioma. Many NIR-I-activated INPs with multiple optical functions can be used to treat gliomas. Kim and colleagues [213] developed anatase titania-coated plasmonic gold nanorods decorated with upconversion nanoparticles (Au NR@aTiO2@UCNPs; Fig. 11A). This novel architecture incorporates an anatase titanium dioxide (aTiO2) photosensitizer as a spacer and takes advantage of the localized surface plasmon resonance (LSPR) properties of the Au core. A superior ROS-producing ability was found with Au NR@aTiO2@UCNPs under laser irradiation (808 ​nm, 2 ​W/cm2) compared with controls. Meanwhile, the temperature of Au NR@aTiO2 and Au NR@aTiO2@UCNPs could reach 42 ​°C after 10 ​min under NIR irradiation. This excellent low-temperature photothermal efficiency is expected to play a key role in assisting glioma PDT. The heating is primarily attributed to the plasmonic effect of the Au NR cores. In vitro tumor cell ablation research demonstrated that Au NR@aTiO2@UCNPs facilitated a remarkable reduction in cell viability once activated. An in vivo anticancer experiment also showed that complete glioma ablation could only be achieved by NIR laser irradiation in the presence of Au NR@aTiO2@UCNPs. Overall, Au NR@aTiO2@UCNPs are excellent multifunctional phototherapeutic agents that can provide a valuable complement to laser-triggered deep-tissue glioma PTT/PDT.

Fig. 11.

Fig. 11

INPs used in glioma phototherapy. A. Au NR@aTiO2@UCNPs; B. Cu2-xSe NPs; C. Fe@Au NPs. Panel A is adapted with permission from Ref. [213], copyright 2021 ACS Applied Materials & Interfaces. Panel B is adapted with permission from Ref. [80], copyright 2019 Nanoscale. Panel C is adapted with permission from Ref. [215], copyright 2021 pharmaceutics.

Compared with NIR-I lasers (750–900 ​nm), NIR-II region lasers (1000–1700 ​nm) have deeper tissue penetration ability, a higher radiation ceiling and better tissue tolerance. They also have greater potential for application in the phototherapy of brain tumors [214]. However, there are few reports on the treatment of deep-seated gliomas by NIR–II–mediated phototherapy. Gao et al. [80] reported the treatment of in situ GBM with imaging-guided NIR-II PDT and chemotherapy using Cu2-xSe (CS) NPs (Fig. 11B). These drug-loaded ultra-small theranostic NPs can cross the BBB with the assistance of ultrasound and act as both a photosensitizer and therapeutic agent. After a CS NPs/H2O2 mixture was irradiated with a 1064 ​nm laser, H2O2 was efficiently degraded at a rate higher than that attained at 808 ​nm. Meanwhile, stronger ROS fluorescence was also observed under 1064 ​nm irradiation. Subsequently, the authors demonstrated that CS NPs effectively degraded H2O2 and oxygen within the tumor via electron transfer and energy transfer mechanisms. Next, a low concentration of CS NPs (25 ​μg/mL) was incubated with U87 ​cells and irradiated for 5 ​min to detect intracellular ·OH and 1O2 radical production. The cells showed intense red (·OH) and green (1O2 radicals) fluorescence, which was 3.3- and 4.3-fold higher than that of unirradiated CS NPs cells, respectively. To sum up, the advantages of ultra-small Cu2-xSe NPs combined with a deeply penetrating 1064 ​nm laser show great potential in NIR-II PDT combined therapy for glioma. Additionally, García-Martín et al. [215] investigated the photothermal properties of gold-coated iron oxide NPs coated with polyvinylpyrrolidone (Fe@Au NPs; Fig. 11C). Fe@Au NPs exhibit an exploitable “photothermal” conversion ability in the NIR-II region due to the existence of LSPR in gold. As expected, under a NIR laser (1064 ​nm, 1.22 ​W/cm2), Fe@Au NPs demonstrated a rapid increase in temperature under laser heating and cease after laser shutdown. The heating rate of the NP suspension was 3.3 ​°C/min, which was 10 times that of the control. In addition, during four cycles of NIR irradiation, the maximum temperature remained similar, proving the excellent stability of Fe@Au NPs under NIR irradiation. Notably, the calculated photothermal conversion efficiency was 42.6%. In vitro irradiation experiments showed that >90% of the glioma cells were killed in the target region within 7 ​min. Finally, Fe@Au NPs were successfully used for PTT against GBM in a mouse model. These results imply that Fe@Au NPs have the potential to effectively eliminate GBM as thermal-coupling agents in vitro and in vivo.

Currently, the majority of INPs applied in glioma phototherapy are those that respond to NIR-I, while exploration of phototherapy materials responding to NIR-II with better penetration ability is urgently needed. Moreover, some phototherapy materials used in vitro may not have satisfactory anti-glioma effects in vivo due to their failure to penetrate the skull under short-wavelength excitation or the high temperatures of PTT, which can result in damage to other brain tissues. Hence, INPs used for phototherapy need to be further optimized.

6.2.2. Immunotherapy

Immunotherapy is mainly used for advanced tumors. Current immunotherapy techniques include immune checkpoint blocking, overwhelming immune tolerance, modified T-cell therapy, and the identification of novel tumor antigens using next-generation sequencing. Passive cancer immunotherapy involves the use of drugs such as lymphocytes, cytokines, and monoclonal antibodies to promote an existing antitumor response. Active cancer immunotherapy, on the other hand, involves stimulating an autoimmune response to invading tumor cells through non-specific immune regulation, vaccination, and explicit targeting of antigen receptors. In fact, only a small proportion of patients benefit from the immunotherapy described above. One of the leading reasons is that tumor cells have a high mutation rate and produce new antigens, resulting in low immunogenicity and poor recognition ability in dendritic cells (DCs), which allows tumor cells to evade the immune response.

In 2022, Vemana [216] outlined NPs-based nanomedicine in targeted delivery of immunotherapeutic medications and immunomodulatory chemicals, and their role in regulating the tumor microenvironment to promote tumor immunotherapy. Several in vitro and in vivo studies have shown promising effects of NPs in cancer immunotherapy, including significant anti-degradation drug protection, intracellular delivery, regulated and sustained release, and the prevention of multi-drug resistance in various types of NPs. Compared with other immunotherapy, NPs-based immunotherapy can provide durable vaccine efficacy and broad immune responses. Firstly, NPs are the most typical delivery vehicles used to ensure that tumor antigens reach lymph nodes and achieve effective immunotherapy. Secondly, NPs with specific functionalization can induce, inhibit or alter the innate immune system; for example, by inducing cytokine production or by activating downregulation mechanisms or immunosuppressive immune cells [[217], [218], [219], [220]]. In addition to tumor antigens, NPs can efficiently deliver adjuvants to antigen-presenting cells (APCs; e.g., macrophages and DC) located in lymph nodes, thus allowing antigen presentation.

At present, a novel use of NPs in cancer immunotherapy is to promote the maturation of DCs, activate NK cells or T cells, and jointly release immune factors (e.g., TNF-α, IL-6 and IL-1β) to achieve ICD [[221], [222], [223], [224]]. In this part, the ICD therapy for glioma based on INPs was mainly reviewed.

Zhang and colleagues [225] reported hybrid ‘clusterbomb’ nanovaccines (MPSDP-ZnO/Ag) that can activate the in vivo immune response via the lymphatic vessels. MPSDP-ZnO/Ag with high antigen loading and well-defined hybrid nanostructures can trigger a cluster to ‘bomb’ in response to APC for simultaneous release of antigen and adjuvant. Hence, the nanosystem can be applied for the combined immunotherapy of GBM without crossing the BBB. As shown in Fig. 12A, MPSDP-ZnO/Ag can significantly increase the cellular uptake of Ag in DCs and generate EGFRvIII-specific CD8+ T cells from vaccinated rats,thus trigger the strongest Ag-specific cytotoxicity. Subsequently, it was also shown that MPSDP-ZnO/Ag might effectively stimulate the proliferation of CD4+ T and CD8+ T cells in the spleen of tumor-bearing rats, which may be attributed to the high cell-uptake efficiency and intracellular transfer of APC. Furthermore, the expression of anti-inflammatory cytokine IL-10 was significantly reduced, while pro-inflammatory cytokines IFN-γ and TNF-α were significantly increased in the glioma microenvironment after treatment with MPSDP-ZnO/Ag nanovaccine. Notably, the survival rate of the tumor-bearing rats treated with MPSDP-ZnO/Ag was distinctly prolonged. The designed INPs clearly boosted the maturation of DCs and cytokine secretion, which ultimately enhanced immunotherapy against GBM.

Fig. 12.

Fig. 12

Novel glioma therapeutic strategies guided by INPs. (A–B). The application in immunotherapy; A. MPSDP-ZnO/Ag; B. Nano-DOX. (C–D). The application in MHT: C. SPANs; D. IONPs. Panel A is adapted with permission from Ref. [225], copyright 2019 Materials Horizons. Panel B is adapted with permission from Ref. [227], copyright 2018 Biomaterials. Panel C is adapted with permission from Ref. [239], copyright 2018 Materials Science & Engineering C. Panel D is adapted with permission from Ref. [240], copyright 2017 Theranostics.

In the carbon nanomaterials family, nanodiamonds are a special class of INPs due to their unique physicochemical properties [226]. DOX-polyglycerol-nanodiamond composites (Nano-DOX) [227] can serve as a DOX-based ICD inducer (more potent than free DOX) that can be delivered into GBM via DCs and, subsequently, initiate an anti-GBM immune response (Fig. 12B). The immunogenicity includes an adjuvantability conferred by the release of damage-associated molecular patterns (DAMPs) and antigenicity involved in the production and supply of antigens. The emissions of DAMPs (CRT, ATP, HMGB-1 and HSP90) were significantly higher in Nano-DOX-treated GBM cells (GCs) than in free DOX-treated GCs. The MHC-I upregulation stimulated by Nano-DOX caused more cytoplasmic antigens to be displayed on the GC surface. Moreover, DC must be mature/activated in order to present antigens to the lymphocyte (LC). The authors found that nano-DOX can enhance DC maturation and activation. DC shows a mature morphology and significantly upregulated surface expression of MHC-II, CD80 and CD86, indicative of enhanced antigen presentation. Next, in the presence of nano-DOX, mouse LC (CD4+ and CD8+) showed marked proliferation when co-cultured with GC and mouse bone marrow-derived DC (mDC), and the release of IFN-γ, a characteristic cytokine of LC activation, was sharply increased. Release of IL-2, another hallmark of LC activation, was increased in all culture environments compared to controls. Lastly, enhanced infiltration and activation of mDC and LC, as well as increased GC apoptosis, were also detected in GBM xenografts treated by nano-DOX. Taken together, the DC-mediated delivery of nano-DOX appears to elicit robust anti-GBM immunity by stimulating immunogenicity in GCs, providing a novel tool for anti-GBM immunotherapy that has great therapeutic potential.

INPs can be passively or actively targeted to gliomas by the EPR effect or ligand surface modification, respectively. Thus, nanotechnology offers a variety of options for efficient delivery of ICD inducers to gliomas at optimal doses due to the development of various construction processes [[228], [229], [230]]. INPs can be used as a local source of ICD induction therapy or to enhance therapy by external energy fields to reduce damage to healthy tissues [231]. INPs with imaging capabilities can be used to monitor their localization and therapeutic effects in real time, while INPs with immunomodulatory effects can be used as adjuvants or immune enhancers [231]. However, based on immunotherapy and nanotechnology, improving the abilities of INPs to cross the BBB and target gliomas, and promoting the clinical translation of glioma immunotherapy, are challenges that remain to be addressed.

6.2.3. Magnetic hyperthermia therapy

Other therapies, such as magnetic hyperthermia therapy (MHT) using MNPs composed of inorganic elements [94,232], have been explored for glioma treatment. In general, MNPs can be exited in the presence of AMF and generate heat through hysteresis loss [233]. The heating capacity depends on the properties of the MNPs and the AMF parameters. MHT was approved as an adjuvant therapy for recurrent GBM in combination with radiotherapy in Europe since 2012 [234]. In clinical trials using MHT to treat GBM patients, an AMF with an intensity of 18 ​kA/m and a frequency of 100 ​kHz was safely applied to the brain in multiple stages [235]. The advantage of MHT for glioma treatment is that it can guide treatment more precisely due to the tumor tissue specificity of INPs. It can also enhance treatment safety due to the lower destruction of intracranial glioma at lower temperatures [236,237]. More importantly, it has the non-invasive feature with unrestricted tissue penetration. For instance, virtually, all of the AMF energy at 400 ​kHz can be targeted to sites of MNPs accumulation, even penetrating up to 15 ​cm [238].

Sawant and colleagues [239] synthesized pH-sensitive alloy-drug nanoconjugates (SPANs) based on a composite of Fe and Pt that have excellent heating ability and can be excited by both AMF and NIR lasers (Fig. 12C). The pH-responsive drug release can help overcome drug and metal damage to normal cells. The glioma cell temperature was increased by 8–11 ​°C under the dual heating mode at 335.2 ​Oe AMF and 1064 ​nm laser irradiation. On account of the higher absorbance in the NIR-II window deriving from the Fe2+–Fe3+ transition in SPANs, the heating under 1064 ​nm irradiation was higher than under 808 ​nm irradiation. Furthermore, Fe-based alloy nanoparticles catalyzed by the Fenton and Haber-Weiss reactions may generate ROS to induce tumor death. Summarily, the work demonstrate that SPANs excited by both AMF and NIR lasers enhance the suppression of U87MG cell viability.

A magnetosome chain is an iron oxide nanoparticle (IONP) synthesized by magnetotactic bacteria, which can replace chemically synthesized IONP and further improve the therapeutic effect. Alphandéry and colleagues [240] fabricated magnetosomes-poly-l-lysine (M-PLL) by removing potentially toxic organic bacterial residues and coating magnetosome mineral core surfaces with poly-l-lysine to achieve complete biocompatibility. The anti-glioma efficacy of M-PLL was compared with that of chemically synthesized IONPs currently used in magnetic hyperthermia (Fig. 12D). When GL-261 (murine glioma cell line) subcutaneous glioma tumors were exposed to AMF at a frequency of 198 ​kHz and 11–31 ​mT, their temperature could reach 43–46 ​°C within 30 ​min. This shows that M-PLL can achieve better therapeutic effects at a very low magnetic field and energy consumption. More importantly, MHT induced by M-PLL ​+ ​AMF can completely ablate tumors or inhibit tumor growth. In conclusion, MHT for GBM has been improved by the use of biocompatible magnetosomes.

MNPs for glioma MHT have tumor specificity and long-lasting therapeutic effects. Compared with PTT, the penetration depth of MHT is deeper. Moreover, due to the optical catalytic activity of some magnetic metal NPs, the combination of MHT with phototherapy can increase the therapeutic outcome of glioma. Although MHT has been shown to be effective in vitro, in animal models, and in human experiments, there are still some challenges in becoming a standard of treatment for gliomas, such as accurate thermometry within the tumor mass and precise tumor heating.

6.3. Multifunctional inorganic nanoplatforms for glioma diagnosis or treatment

With the development of nanotechnology, increasing numbers of multifunctional INPs have arisen for tumor diagnosis and therapy. This presents a new perspective for establishing efficiency and precision-guided multifunctional glioma treatment platforms. Representative examples of INPs applied to the integrated diagnosis and treatment of glioma in the past decade are summarized in Table 4, which can provide optimal preparation strategies for multifunctional nanoplatforms as required. In the applications described above, many INPs have been modified or doped according to the functions of different dominant elements. This enables nanocomposites to obtain three or more kinds of multi-functional diagnosis or treatment characteristics, thereby improving the efficiency of integrated glioma treatment.

Table 4.

Summary of glioma diagnosis and treatment based on inorganic nanoplatforms.

Inorganic nanoplatform Size (nm) Cell type Cell safety concentration (Incubation time) Animal tumor model Application(s) Ref.
Au TAT-Au NP 5.9 ​± ​2.1 U87 None Intracranial U87 mice model MRI [58]
DOX@CMXG
@AuNPs
8–10 LN-299 10 ​μg/mL (24 ​h) None FLI [247]
AuNPs 20 C6 None None FLI [248]
PEI-based Au NPs 151.0 ​± ​25.4 C6 None Subcutaneous C6 rats model CT, RT [72]
RGD-[64Cu]
Au NR
L: 25.1 ​± ​2
W:8.0 ​± ​0.5
U87 None Subcutaneous U87 mice model PET, PTT (808 ​nm) [249]
AuNDIPs 282 ​± ​15 C6 500 ​μg/mL (24 ​h) None PTT/PDT (808 ​nm)
/CHEMO
[211]
GC-pep@SiNC
-AuNC
160 U87 5 ​μg/mL (6 ​h) Intracranial U87 mice model PDT (785 ​nm) [250]
SNAs 13 U87 5 ​nM (24 ​h) Intracranial U87 mice model GT [251]
SNAs 13 ​± ​1 U87 and GIC-20 None Intracranial U87/GIC-20 mice model GT [252]
RDGNR/shRNA L: 50
W: 10
U87 25 ​μg/mL (48 ​h) Subcutaneous U87 mice model GT [253]
Au PENPs 24.2 U87 3000 ​nM (24 ​h) Subcutaneous U87 mice model GT [254]
D&H-A-A&C 38.1 ​± ​1.6 C6 None Intracranial C6 mice model IMT/CHEMO [255]
Au–OMV 42 G261, bEND.3 and C8D1A 2 ​μg/mL (24 ​h) Subcutaneous/Intracranial G261 mice model IMT/RT [101]
cRGD–HN–DOX 90 U87 10 ​μg/mL (4 ​h) Subcutaneous U87 mice model CHEMO [256]
An-PEG-DOX-AuNPs 39.9 C6 None Intracranial C6 mice model CHEMO [64]
AuNPs-DC-RRGD 55.9 C6 None Intracranial C6 mice model CHEMO [171]
GNP-UP-Cis 7 U87, U251, T98G and U138 None Intracranial U251 mice model CHEMO/RT [124]
Au PENPs-CTX 4.4 ​± ​1.3 C6 200 ​μM (24 ​h) Subcutaneous/Intracranial C6 mice/rats model CT, RT [72]
BSA-AuNPs 18 U87 0.3 ​mg/mL (24 ​h) Subcutaneous U87 mice model RT [106]
FA-AuNCs 5.5 ​± ​0.4 C6 and L929 160 ​μg/mL (24 ​h) Intracranial C6 rats model RT [257]
DNA-AuNPs 4.5 U251 None None RT [258]
Ag Ag2S QDs 5.4 ​± ​0.3 MDA-MB-468 and U87 None None FLI [259]
AgNPs 21.2 ​± ​0.31 C6 None Intracranial C6 rats model RT [260]
Ag-PNP-CTX 114 ​± ​2 U87 and T98 None Intracranial U87 mice model RT [261]
Cu Cu2-xSe NPs 3 ​± ​0.3 None None None PAI/PET [79]
Cd CdTe/CdS QDs 4.05 ​± ​0.3 U251 None None FLI [262]
QD-MPA None 293 ​T, HUVEC, MCF7, U87 and SW620 None Subcutaneous HT1080/MCF-7 mice model FLI [73]
CdSe QDs/CdSeNPls 22 ​± ​2/33 ​± ​8 C6 None None FLI [263]
NGR-PEG-QDs 10 BCECs, astrocytes and C6 20 ​nM (96 ​h) Intracranial C6 rats model FLI [74]
QD-Apt 20 U87 and HUVEC None Intracranial U87 mice model FLI [75]
Zn ZnS-QDs 3.6 U-87 MG, HEK293T None None FLI, CHEMO [162]
CdSSe/ZnS QDs 12.97 ​± ​4.62 U87 and U251 20 ​μg/mL (48 ​h) None GT [194]
ZnS@CMC-DOX 3.6 HEK 293 ​T and U87 None None FLI, CHEMO [162]
Ti Au NR@aTiO2
@UCNPs
∼200 U87 120 ​μg/mL (24 ​h) Subcutaneous U87 mice model PTT/PDT (808 ​nm) [213]
Mo MoS2 Nanosheets 181 ​± ​3 U87 1 ​mg/mL (1.5 ​h) Subcutaneous U87 mice model PAI [149]
La (Er) NaErF4:Ce@NaAF4@NaLuF4 17.9 ​± ​0.9 U87 800 ​μg/mL (24 ​h) Subcutaneous/Intracranial U87 mice model FLI [76]
La2O3 NPs <100 RN1, GBML1, WK1, PDCL G53 and NHA None None RT [190]
Gd PFG-Ce6-Gd NPs 130 C6 1.25 ​μg/mL (12 ​h) Subcutaneous
C6 rats model
MRI [96]
ANG/PEG-UCNPs 17.2 ​± ​0.5 U87 1000 ​μg/mL (24 ​h) Intracranial U87/BCECs mice model MRI/FLI [264]
Fe Pac-MNPs 202 ​± ​3.7 U87 None Intracranial U87 rats model MRI [265]
IONPs 33.24 U87 2 ​mg/mL (2 ​h) Subcutaneous U87 mice model MRI [125]
PDNCs 100–120 RG2 1000 ​μg/mL (24 ​h) Intracranial RG2 rats model MRI/FLI [66]
GPEI-Fe NPs 110 9 ​L 45 ​μg/mL (3 ​h) Intracranial
9 ​L rats model
MRI [266]
CS-DX-SPIONs 55 U87 and C6 10 ​μg/mL (24 ​h) Intracranial C6 rats model MRI [267]
SPIO@AuNP 82 U87 4 ​μg/mL (48 ​h) Intracranial U87 mice model MRI/PAI [268]
Fe@Au NPs 164 ​± ​44 HFF-1 and C6 100 ​μg/mL (24 ​h) Subcutaneous C6 mice model MRI/CT, PTT (1064 ​nm) [215]
SPIO@DSPE-PEG/DOX/ICG
NPs
22.9 U251, C6 and HUVECs 100 ​μg/mL (72 ​h) Intracranial C6 rats model MRI/FLI, CHEMO [246]
IONPs L: 20 ​± ​5
W: 17 ​± ​4
GL-261 None Subcutaneous GL-261 mice model MHT [240]
Fe3O4 @Au–C225 MNPs 46 U251 None Subcutaneous U251 mice model MHT [94]
M-PLL and IONP 17–20 3T3 and U87 None Intracranial U87 mice model MHT [269]
MCNP-ATAP 46.8 ​± ​2.3 U87 and MDA-MB-231 50 ​μg/mL (48 ​h) None MHT [270]
Si MSNs 2.9 U87 400 ​μg/mL (48 ​h) None CT [271]
SiNPs 45 None 3 ​mg/mL (24 ​h) None FLI [272]
MSNs 88.5 ​± ​2.2 U87MG None Subcutaneous U87 mice model PET [139]
pSiNPs L: 7.5
W: 2.5
U87 and 4T1 None Subcutaneous U87 mice model GT [196]
DOX@MSN–SS–iRGD&1 ​MT 50 GL261 None Intracranial GL261 mice model IMT/CHEMO [70]
DOX@MSN 20/40/80 U87, U251, C6 and CHEM-5 None None CHEMO [48]
CPT@MSN-hyd-DOX 2.9 U87 400 ​μg/mL (48 ​h) None CHEMO [271]
MNP-MSN-PLGA-Tf NPs ∼150 bEND.3 and U87 100 ​μg/mL (96 ​h) Intracranial U87 mice model CHEMO [67]
DOX-MNP-MSN-PF-127-Tf 110 U87 1 ​μg/mL (96 ​h) None CHEMO [68]
VPA-MSNs L: 180
W: 120
C6 and U87 100 ​mg/mL (24 ​h) None RT [273]
GPE GSPID 50–250 U251 50 ​μg/mL (2 ​h) None PTT (808 ​nm)
/CHEMO
[274]
GQDs/DOX
@CCM
130 ​± ​10 BV2, GMI-R1 and rat astrocyte 200 ​μg/mL (24 ​h) None PTT (808 ​nm)
/CHEMO
[100]
C NCDDG 30 U87 256 ​μg/mL (24 ​h) Intravenous U87 mice model MRI/FLI [275]
O-MWNTs-PEG-ANG 10 C6 100 ​μg/mL (24 ​h) Intracranial
C6 mice model
FLI [71]
CD-Asp 2.28 ​± ​0.42 C6 None Intracranial C6 mice model FLI [47]
HCCDs 6–8 U87 800 ​μg/mL (24 ​h) Intracranial U87 mice model FLI/PAI, PTT (808 ​nm) [242]
CNT None U251, U87, LN229 and T98G 3 ​μg/mL (72 ​h) Subcutaneous U251 mice model PTT (NIR) [276]
CNT 8–15 U87, U373, NHA and D54 100 ​μg/mL (24 ​h) None PTT (970 ​nm) [277]

∗L represents for Length; W represents for Width; CHEMO represents for chemotherapy; RT represents for radiotherapy; GT represents for gene therapy; IMT represents for immunotherapy; FLI represents for fluorescence imaging.

6.3.1. The glioma multimodality molecular imaging based on INPs

Multimodality molecular imaging combines two or more imaging techniques, integrating their benefits while overcoming the limitations of a single imaging modality, so as to offer sufficient information for accurate diagnosis. Advances in the design and synthesis of inorganic nanoplatforms will offer the ability to produce novel multimodality imaging probes, which can target glioma and reflect more pathophysiological details. Multimodality molecular imaging based on INPs can be tailored as necessary, such as through the assembly of inorganic elements with diverse properties and the encapsulation of molecules with varying functions.

Xu and colleagues [241] constructed multifunctional hierarchical nanocomposites (pDNA/Au@PDM/Fe3O4) for trimodal PAI/MRI/CT diagnostic methods through a consecutive electrostatic assembly technique (Fig. 13A). Au@pDM/Fe3O4 is based on ternary assemblies of negatively charged Fe3O4 cores (Fe3O4-PDA), polycation-modified Au nanorods (Au NR-pDM) and polycations. In terms of optical imaging, this nanocomposite can be used for PAI because of the combined NIR absorbance properties of Fe3O4-PDA and Au NR-pDM, which can greatly enhance optical absorption by tissue and reflect the characteristics of tumor tissue. Besides, Fe3O4 NPs and Au NPs have been demonstrated to be promising T2-weighted MRI and CT contrast agents. In vitro and in vivo assay results suggest that the pDNA/Au@PDM/Fe3O4 should be one promising trimodal contrast agent for PAI/MRI/CT of tumors.

Fig. 13.

Fig. 13

Multifunctional inorganic nanoplatforms for combined glioma diagnosis or treatment. A. Au@PDM/Fe3O4; B. HCCDs; C. Cur-MMCN; D. SPIO@DSPE-PEG/DOX/ICG NPs. Panel A is adapted with permission from Ref. [241], copyright 2016 Materials Views. Panel B is adapted with permission from Ref. [242], copyright 2018 ACS Applied Materials & Interfaces. Panel C is adapted with permission from Ref. [245], copyright 2019 ACS biomaterials science & engineering. Panel D is adapted with permission from Ref. [246], copyright 2019 International Journal of Nanomedicine.

In addition to the hierarchical nanocomposites mentioned above, Huang et al. [242] developed multicolour highly crystalline carbon nanodots (HCCDs) for glioma dual-modality molecular imaging (Fig. 13B). HCCDs were synthesized using F127 and phenol resin as a carbon source to form a mesoporous carbon template via an in-situ solid-state transformation technique. The HCCDs contain a highly crystalline carbon nanocore and a hydrophilic surface, thereby providing both intense photoacoustic as well as adjustable fluorescence emissions. The potential to use HCCDs for dual-modal-imaging-guided PTT of gliomas in vitro and in vivo was verified. According to these observations, a visualized PTT for gliomas can be guided by deep space-resolved PAI and sensitive fluorescence imaging, where the timing and location of NIR irradiation can be adjusted to minimize side effects to normal tissue following the injection of HCCDs.

In brief, the combination of multiple diagnostic approaches can be achieved as needed by using appropriate inorganic components and construction strategies. Although this field is still in its infancy, multimodality molecular imaging based on INPs has exciting features in glioma diagnosis.

6.3.2. The glioma synergistic therapy based on INPs

As described above, Xu [241] constructed pDNA/Au@PDM/Fe3O4 for synergistic gene therapy and PTT (Fig. 13A). Positively charged Au@pDM/Fe3O4 is also able to compound plasmid DNA into pDNA/Au@pDM/Fe3O4 and efficiently mediate gene therapy. It is well known that the p53 gene is a tumor-suppressor and could be applied for clinical gene therapy [243,244]. Au@PDM/Fe3O4 can be employed to deliver plasmid p53 (cloned with p53 genes) for gene therapy. Furthermore, Au@PDM/Fe3O4 produced a great temperature change of 50.1 ​°C after laser irradiation for 5 ​min, indicative of excellent photothermal conversion performance. Glioma C6 cells incubated with Au@PDM/Fe3O4 were irradiated with an 808 ​nm laser, displaying substantially lower cell viability. In vitro molecular mechanism experiments demonstrated that cells incubated with p53/Au@PDM/Fe3O4 exhibited distinctly lower viability than pRL/Au@PDM/Fe3O4 (control), and can be further eliminated by laser irradiation. Likewise, a p53/Au@PDM/Fe3O4 nanoplatform exhibited strong tumor growth inhibition ability in vivo antitumor therapy combining PTT and p53 gene therapy. These results demonstrate that the combination of gene therapy and PTT based on p53/Au@PDM/Fe3O4 nanoplatform can obviously increase glioma therapeutic effects.

Another example of glioma synergistic therapy based on INPs is a mesoporous magnetic nanohybrid as a combination platform for chemotherapy, PDT and MHT (Fig. 13C) [245]. The mesoporous magnetic nanohybrid was functionalized with 14 ​wt% carbon nitride (CN) and loaded with Cur (Cur-MMCN). Cur-MMCN is formed by means of physical adsorption between the extended planar structure of Cur and the 2D architecture of CN. The combination of acidic pH and AMF could trigger the release of Cur from MMCN. CN based nanomaterials have emerged as a new class of photocatalysts for PDT. Cur-MMCN increased ROS by 350% at a 7.5 ​μM concentration under blue LED light irradiation compared to control groups. In addition to the photocatalytic anti-glioma effect of Cur-MMCN, the intrinsic presence of H2O2 in glioma cells could also assist in the generation of high reactive hydroxyl and superoxide radicals catalyzed by Fe2+ ions through the Fenton reaction. Cur-MMCN activated with blue light effectively killed ∼60% of cells, while cell viability was further reduced to ∼15% when co-exposed to AMF. The caspase 3/7 apoptosis assay showed that a combination of mild MHT and high oxidative stress is capable of triggering the apoptotic pathway leading to the highest level of glioma cell death. Summarily, the combination of Cur-MMCN, blue light and AMF achieves better glioma therapeutic outcomes.

6.3.3. Combined glioma diagnosis and treatment based on INPs

One of the classic works on a multifunctional inorganic nanoplatform for synergistic glioma diagnosis and treatment is imaging-mediated chemotherapy. Shen et al. [246] constructed a novel multifunctional drug delivery system (SPIO@DSPE-PEG/DOX/ICG NPs), which combines real-time MR/fluorescence imaging and chemotherapy (Fig. 13D). Hydrophobic superparamagnetic iron oxide nanoparticles (SPIONPs) stabilized with oleylamine and oleic acid were prepared by a thermal decomposition method. DOX-loaded NPs were coated with DSPE-PEG 2000 using a classic thin-film hydration method. The amphiphilic structure of ICG enabled it to be readily loaded onto the phospholipid layer of the NPs. SPIO can be used as an MRI contrast agent that has excellent spatial resolution but poor sensitivity. Therefore, the combination of SPIO with the highly sensitive fluorescent imaging agent ICG could achieve great dual-modality imaging results. The fluorescence signal of free ICG was detectable in mice only 2 ​h after SPIO@DSPE-PEG/DOX/ICG NPs injection, after which the signal gradually faded. After 24 ​h, the majority of the signals was concentrated in the tumor tissue, which indicated that the NPs could be blocked mainly in tumor tissues due to strong EPR effects. Furthermore, the MR imaging data showed that the multifunctional NPs remained in the tumor for at least 60 ​h, while the non-tumor brain areas returned to the pre-injection level after 4 ​h. Subsequently, they investigated the effect of chemotherapy using SPIO@DSPE-PEG/DOX/ICG NPs in orthotopic glioma-bearing rats, with tumor sizes monitored in real-time using MRI. A rapid rate of tumor growth was observed in brain MR images of control rats, whereas rats treated with SPIO@DSPE-PEG/DOX/ICG NPs showed inhibited tumor growth. The current results provide useful information for the design and development of multifunctional nanocomposites for the diagnosis and treatment of glioma based on appropriate assemblies of various INPs and polycations.

7. Conclusion and future perspectives

In this review, we extensively discussed the last decade's advances in developing inorganic nanomaterials for the treatment and diagnosis of glioma. As the treatment of intracranial glioma is more difficult than treatment in other organs, there are also few suitable nanomaterials. The design and preparation of theranostic nanomaterials are based on the disease characteristics of glioma. It is also necessary to consider the characteristics of the BBB in the different stages of glioma and to enhance “sequential targeting” across the BBB and glioma tissues by connecting targeting groups. The choice of targeting groups is also important. As damage to the BBB occurs in different stages of glioma, different targeting groups could be selected. In the early stage when the BBB is relatively complete, bipolar targeting could be considered, while in the late stage when the BBB is completely destroyed, suitable ligand molecules for targeting gliomas should be considered (specific molecules are summarized in Table 2).

As proliferating gliomas have unclear boundaries, clear imaging techniques are extremely important. Many inorganic nanomaterials can be constructed by doping and other methods to achieve multimodal imaging ability, which can improve diagnostic accuracy, identification of glioma margins and surgical resection [278]. Also, targeting efficiency can be improved by connecting different active targeting groups. The choice of imaging tool depends on how different materials respond to excitation conditions. In terms of treatment, the main functions of INPs are as carriers to deliver chemotherapy drug, immune adjuvants, and siRNAs that go into the brain. The intrinsic physical properties of inorganic nanomaterials, such as their laser and nuclear magnetic response characteristics, can also be used to improve imaging sensitization and chemotherapy efficacy, or to regulate drug duration and release efficiency. At the same time, because of the increasing number of active targeting groups, much work has been repeatedly verified to obtain some efficient targeting groups, such as low-density lipoprotein receptor-associated protein-1,64 hyaluronic acid [279], Tf [280] and rituximab [281].

Gd-based contrast agents have been used in the clinical diagnosis of glioma. Gd-based AGuIX nanoparticles have also been reported to have achieved promising results in Phase I clinical trials. In one trial, 13 out of 14 patients with evaluable brain metastases showed clinical benefit [282]. Moreover, a non-randomized clinical trial has confirmed that fluorescence-guided sentinel lymph node (SLN) biopsy based on ultrasmall, molecularly targeted core-shell silica nanoparticles is feasible and safe for the treatment of head and neck melanoma [283]. Research into nanomaterials is also accelerating the pace of clinical applications; however, there are still several important issues that need to be addressed. For instance, the type, size, shape, surface charge, modification and other characteristics of nanoparticles may affect their efficacy in vivo and influence the structure of the BBB. These factors may not only affect cytotoxicity but also the distribution and metabolism of nanoparticles, which are associated with long-term toxicity to the brain and other parts of the organism. Furthermore, systematic evaluation of the effects on brain function, such as changes in learning, cognitive ability, BBB integrity and permeability, is needed.

In addition to biocompatibility, the dosage of medications required to achieve the desired effects must be considered due to the high permeability of nanoparticles. Surface functional modifications can reduce cytotoxicity and improve biocompatibility. Different surface charges and surface modifications can generate different consequences. In general, depending on electrostatic interactions, positively charged INPs are more easily absorbed by cell membranes than negatively or neutrally charged INPs, which have longer blood circulation times because they have difficulty binding to proteins or cells in the blood [[284], [285], [286], [287]]. For gold nanomaterials, chemical modification of the surface may produce different degrees of immune stimulation; this property can be applied in immunotherapy [288]. Thus, we should consider their intrinsic properties to achieve the desired effect when designing nanomaterials. Additionally, the mechanism of cell death induced by some INPs is increased intracellular ROS levels, as well as involving apoptosis and autophagy pathways. The other potential mechanism of treating glioma based on INPs is still unclear and needs further exploration.

Moreover, extrinsic physical techniques, especially lasers, can be used to enhance chemotherapy and phototherapy and improve the quality and sensitivity of in vivo PAI and fluorescence imaging. In addition, the excellent photostability and long half-life of INPs make them ideal for phototherapy. To improve skull penetration efficiency, research on how inorganic nanocomplexes respond to NIR-II lasers is providing new strategies for the optical diagnosis and treatment of glioma. Considering intracranial safety, the commonly used methods are low-injury hypothermia PTT and PDT. However, whether the ROS over-produced by PDT has side effects on normal tissues still needs to be studied, so as to determine the optimal ROS production rate and achieve the best therapeutic effect. Besides, evaluation of the long-term toxicity of degraded nanocomposites in vivo should not be ignored.

In conclusion, the current progress and issues in glioma theranostics have created demands for a new generation of nanomaterials. The neurotoxicity of multifunctional nanomaterials still warrants further exploration. INPs based new generation nanoplatform is also expected to provide more effective diagnosis and treatment of glioma and extend human life.

Ethics approval

This study does not contain any studies with human or animal subjects performed by any of the authors.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

This research was supported by grants from “The Central Guided Science and Technology Development Foundation of Liaoning Province” (No. 2022 JH6/100100036).

Contributor Information

Chuan He, Email: hec@cmu.edu.cn.

Jian Wen, Email: wenjian_163@hotmail.com.

Xiaoqian Xu, Email: xqxu@cmu.edu.cn.

Data availability

Data will be made available on request.

References

  • 1.Sim H.W., Morgan E.R., Mason W.P. Contemporary management of high-grade gliomas. CNS Oncol. 2018;7:51–65. doi: 10.2217/cns-2017-0026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Tang W., Fan W., Lau J., Deng L., Shen Z., Chen X. Emerging blood-brain-barrier-crossing nanotechnology for brain cancer theranostics. Chem. Soc. Rev. 2019;48:2967–3014. doi: 10.1039/c8cs00805a. [DOI] [PubMed] [Google Scholar]
  • 3.Zhang X., Zhou J., Gu Z., Zhang H., Gong Q., Luo K. Advances in nanomedicines for diagnosis of central nervous system disorders. Biomaterials. 2021;269 doi: 10.1016/j.biomaterials.2020.120492. [DOI] [PubMed] [Google Scholar]
  • 4.Yasaswi P.S., Shetty K., Yadav K.S. Temozolomide nano enabled medicine: promises made by the nanocarriers in glioblastoma therapy. J. Contr. Release. 2021;336:549–571. doi: 10.1016/j.jconrel.2021.07.003. [DOI] [PubMed] [Google Scholar]
  • 5.Wu X., Yang H., Yang W., Chen X., Gao J., Gong X., Wang H., Duan Y., Wei D., Chang J. Nanoparticle-based diagnostic and therapeutic systems for brain tumors. J. Mater. Chem. B. 2019;7:4734–4750. doi: 10.1039/c9tb00860h. [DOI] [PubMed] [Google Scholar]
  • 6.Garcia-Fabiani M.B., Ventosa M., Comba A., Candolfi M., Nicola Candia A.J., Alghamri M.S., Kadiyala P., Carney S., Faisal S.M., Schwendeman A., Moon J.J., Scheetz L., Lahann J., Mauser A., Lowenstein P.R., Castro M.G. Immunotherapy for gliomas: shedding light on progress in preclinical and clinical development. Expet Opin. Invest. Drugs. 2020;29:659–684. doi: 10.1080/13543784.2020.1768528. [DOI] [PubMed] [Google Scholar]
  • 7.Luly K.M., Choi J., Rui Y., Green J.J., Jackson E.M. Safety considerations for nanoparticle gene delivery in pediatric brain tumors. Nanomedicine. 2020;15:1805–1815. doi: 10.2217/nnm-2020-0110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Zimmer C., Weissleder R., Poss K., Bogdanova A., Wright S.C., Enochs W.S. MR imaging of phagocytosis in experimental gliomas. Radiology. 1995;197:533–538. doi: 10.1148/radiology.197.2.7480707. [DOI] [PubMed] [Google Scholar]
  • 9.Norouzi M. Gold nanoparticles in glioma theranostics. Pharmacol. Res. 2020;156 doi: 10.1016/j.phrs.2020.104753. [DOI] [PubMed] [Google Scholar]
  • 10.Choi J., Kim G., Cho S.B., Im H.J. Radiosensitizing high-Z metal nanoparticles for enhanced radiotherapy of glioblastoma multiforme. J. Nanobiotechnol. 2020;18:122–123. doi: 10.1186/s12951-020-00684-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Verma J., Lal S., Van Noorden C.J. Nanoparticles for hyperthermic therapy: synthesis strategies and applications in glioblastoma. Int. J. Nanomed. 2014;9:2863–2877. doi: 10.2147/IJN.S57501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Liu Q., Zhan C., Kohane D.S. Phototriggered drug delivery using inorganic nanomaterials. Bioconjug Chem. 2017;28:98–104. doi: 10.1021/acs.bioconjchem.6b00448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Darr J.A., Zhang J., Makwana N.M., Weng X. Continuous hydrothermal synthesis of inorganic nanoparticles: applications and future directions. Chem. Rev. 2017;117:11125–11238. doi: 10.1021/acs.chemrev.6b00417. [DOI] [PubMed] [Google Scholar]
  • 14.Meng J., Agrahari V., Youm I. Advances in targeted drug delivery approaches for the central nervous system tumors: the inspiration of nanobiotechnology. J. Neuroimmune Pharmacol. 2017;12:84–98. doi: 10.1007/s11481-016-9698-1. [DOI] [PubMed] [Google Scholar]
  • 15.Reynolds J.L., Mahato R.I. Nanomedicines for the treatment of CNS diseases. J. Neuroimmune Pharmacol. 2017;12:1–5. doi: 10.1007/s11481-017-9725-x. [DOI] [PubMed] [Google Scholar]
  • 16.Deng G., Peng X., Sun Z., Zheng W., Yu J., Du L., Chen H., Gong P., Zhang P., Cai L., Tang B.Z. Natural-killer-cell-inspired nanorobots with aggregation-induced emission characteristics for Near-Infrared-II fluorescence-guided glioma theranostics. ACS Nano. 2020;14:11452–11462. doi: 10.1021/acsnano.0c03824. [DOI] [PubMed] [Google Scholar]
  • 17.Xiao D., Jia H.Z., Zhang J., Liu C.W., Zhuo R.X., Zhang X.Z. A dual-responsive mesoporous silica nanoparticle for tumor-triggered targeting drug delivery. Small. 2014;10:591–598. doi: 10.1002/smll.201301926. [DOI] [PubMed] [Google Scholar]
  • 18.Zhang P., Tang M., Huang Q., Zhao G., Huang N., Zhang X., Tan Y., Cheng Y. Combination of 3-Methyladenine therapy and Asn-Gly-Arg (NGR)-modified mesoporous silica nanoparticles loaded with temozolomide for glioma therapy in vitro. Biochem. Biophys. Res. Commun. 2019;509:549–556. doi: 10.1016/j.bbrc.2018.12.158. [DOI] [PubMed] [Google Scholar]
  • 19.Louis D.N., Perry A., Reifenberger G., von Deimling A., Figarella-Branger D., Cavenee W.K., Ohgaki H., Wiestler O.D., Kleihues P., Ellison D.W. The 2016 world health organization classification of tumors of the central nervous system: a summary. Acta Neuropathol. 2016;131:803–820. doi: 10.1007/s00401-016-1545-1. [DOI] [PubMed] [Google Scholar]
  • 20.Louis D.N., Perry A., Wesseling P., Brat D.J., Cree I.A., Figarella-Branger D., Hawkins C., Ng H.K., Pfister S.M., Reifenberger G., Soffietti R., von Deimling A., Ellison D.W. The 2021 WHO classification of tumors of the central nervous system: a summary. Neuro Oncol. 2021;23:1231–1251. doi: 10.1093/neuonc/noab106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Osswald M., Jung E., Sahm F., Solecki G., Venkataramani V., Blaes J., Weil S., Horstmann H., Wiestler B., Syed M., Huang L., Ratliff M., Karimian Jazi K., Kurz F.T., Schmenger T., Lemke D., Gömmel M., Pauli M., Liao Y., Häring P., Pusch S., Herl V., Steinhäuser C., Krunic D., Jarahian M., Miletic H., Berghoff A.S., Griesbeck O., Kalamakis G., Garaschuk O., Preusser M., Weiss S., Liu H., Heiland S., Platten M., Huber P.E., Kuner T., von Deimling A., Wick W., Winkler F. Brain tumor cells interconnect to a functional and resistant network. Nature. 2015;528:93–98. doi: 10.1038/nature16071. [DOI] [PubMed] [Google Scholar]
  • 22.Claes A., Idema A.J., Wesseling P. Diffuse glioma growth: a guerilla war. Acta Neuropathol. 2007;114:443–458. doi: 10.1007/s00401-007-0293-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Frosina G. Limited advances in therapy of glioblastoma trigger re-consideration of research policy. Crit. Rev. Oncol. Hematol. 2015;96:257–261. doi: 10.1016/j.critrevonc.2015.05.013. [DOI] [PubMed] [Google Scholar]
  • 24.Manrique-Guzmán S., Herrada-Pineda T., Revilla-Pacheco F. In: Glioblastoma. Codon Publications. De Vleeschouwer S., editor. Vol. 12. 2017. Surgical management of glioblastoma; pp. 243–261. [PubMed] [Google Scholar]
  • 25.Sneed P.K., Gutin P.H., Larson D.A., Malec M.K., Phillips T.L., Prados M.D., Scharfen C.O., Weaver K.A., Wara W.M. Patterns of recurrence of glioblastoma multiforme after external irradiation followed by implant boost. Int. J. Radiat. Oncol. Biol .Phys. 1994;29:719–727. doi: 10.1016/0360-3016(94)90559-2. [DOI] [PubMed] [Google Scholar]
  • 26.Friedman H.S., Kerby T., Calvert H. Temozolomide and treatment of malignant glioma. Clin Cancer Res. 2000;6:2585–2597. [PubMed] [Google Scholar]
  • 27.Kumar G., Dutta P., Parihar V.K., Chamallamudi M.R., Kumar N. Radiotherapy and its impact on the nervous system of cancer survivors. CNS Neurol Disord Drug Targets. 2020;19:374–385. doi: 10.2174/1871527319666200708125741. [DOI] [PubMed] [Google Scholar]
  • 28.Furtado D., Björnmalm M., Ayton S., Bush A.I., Kempe K., Caruso F. Overcoming the blood-brain barrier: the role of nanomaterials in treating neurological diseases. Adv. Mater. 2018;30 doi: 10.1002/adma.201801362. [DOI] [PubMed] [Google Scholar]
  • 29.Krol S., Macrez R., Docagne F., Defer G., Laurent S., Rahman M., Hajipour M.J., Kehoe P.G., Mahmoudi M. Therapeutic benefits from nanoparticles: the potential significance of nanoscience in diseases with compromise to the blood brain barrier. Chem. Rev. 2013;113:1877–1903. doi: 10.1021/cr200472g. [DOI] [PubMed] [Google Scholar]
  • 30.Xie J., Shen Z., Anraku Y., Kataoka K., Chen X. Nanomaterial-based blood-brain-barrier (BBB) crossing strategies. Biomater. 2019;224 doi: 10.1016/j.biomaterials.2019.119491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Persano F., Batasheva S., Fakhrullina G., Gigli G., Leporatti S., Fakhrullin R. Recent advances in the design of inorganic and nano-clay particles for the treatment of brain disorders. J. Mater. Chem. B. 2021;9:2756–2784. doi: 10.1039/d0tb02957b. [DOI] [PubMed] [Google Scholar]
  • 32.Alyautdin R., Khalin I., Nafeeza M.I., Haron M.H., Kuznetsov D. Nanoscale drug delivery systems and the blood-brain barrier. Int. J. Nanomed. 2014;9:795–811. doi: 10.2147/IJN.S52236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Abbott N.J., Patabendige A.A., Dolman D.E., Yusof S.R., Begley D.J. Structure and function of the blood-brain barrier. Neurobiol. Dis. 2010;37:13–25. doi: 10.1016/j.nbd.2009.07.030. [DOI] [PubMed] [Google Scholar]
  • 34.Li X.Y., Zhao Y., Sun M.G., Shi J.F., Ju R.J., Zhang C.X., Li X.T., Zhao W.Y., Mu L.M., Zeng F., Lou J.N., Lu W.L. Multifunctional liposomes loaded with paclitaxel and artemether for treatment of invasive brain glioma. Biomaterials. 2014;35:5591–5604. doi: 10.1016/j.biomaterials.2014.03.049. [DOI] [PubMed] [Google Scholar]
  • 35.Gao X., Qian J., Zheng S., Changyi Y., Zhang J., Ju S., Zhu J., Li C. Overcoming the blood-brain barrier for delivering drugs into the brain by using adenosine receptor nanoagonist. ACS Nano. 2014;8:3678–3689. doi: 10.1021/nn5003375. [DOI] [PubMed] [Google Scholar]
  • 36.Oberoi R.K., Parrish K.E., Sio T.T., Mittapalli R.K., Elmquist W.F., Sarkaria J.N. Strategies to improve delivery of anticancer drugs across the blood-brain barrier to treat glioblastoma. Neuro Oncol. 2016;18:27–36. doi: 10.1093/neuonc/nov164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Seo J.W., Ang J., Mahakian L.M., Tam S., Fite B., Ingham E.S., Beyer J., Forsayeth J., Bankiewicz K.S., Xu T., Ferrara K.W. Self-assembled 20-nm 64Cu-Micelles Enhance Accumulation in Rat Glioblastoma. J. Controlled Release. 2015;220:51–60. doi: 10.1016/j.jconrel.2015.09.057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lesniak M.S., Brem H. Targeted therapy for brain tumors. Nat Rev Drug Discov. 2004;3:499–508. doi: 10.1038/nrd1414. [DOI] [PubMed] [Google Scholar]
  • 39.Banks W.A. Characteristics of compounds that cross the blood-brain barrier. BMC Neurol. 2009;9(Suppl 1):S3. doi: 10.1186/1471-2377-9-S1-S3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Arvanitis C.D., Ferraro G.B., Jain R.K. The blood-brain barrier and blood-tumor barrier in brain tumors and metastases. Nat. Rev. Cancer. 2020;20:26–41. doi: 10.1038/s41568-019-0205-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Portnow J., Badie B., Chen M., Liu A., Blanchard S., Synold T.W. The neuropharmacokinetics of temozolomide in patients with resectable brain tumors: potential implications for the current approach to chemoradiation. Clin. Cancer Res. 2009;15:7092–7098. doi: 10.1158/1078-0432.CCR-09-1349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Radu M., Chernoff J. An in vivo assay to test blood vessel permeability. J. Vis. Exp. 2013;16 doi: 10.3791/50062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Burks S.R., Kersch C.N., Witko J.A., Pagel M.A., Sundby M., Muldoon L.L., Neuwelt E.A., Frank J.A. Blood-brain barrier opening by intracarotid artery hyperosmolar mannitol induces sterile inflammatory and innate immune responses. Proc. Natl. Acad. Sci. USA. 2021;118 doi: 10.1073/pnas.2021915118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Arvanitis C.D., Askoxylakis V., Guo Y., Datta M., Kloepper J., Ferraro G.B., Bernabeu M.O., Fukumura D., McDannold N., Jain R.K. Mechanisms of enhanced drug delivery in brain metastases with focused ultrasound-induced blood-tumor barrier disruption. Proc. Natl. Acad. Sci. USA. 2018;115:8717–8726. doi: 10.1073/pnas.1807105115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Mainprize T., Lipsman N., Huang Y., Meng Y., Bethune A., Ironside S., Heyn C., Alkins R., Trudeau M., Sahgal A., Perry J., Hynynen K. Blood-brain barrier opening in primary brain tumors with non-invasive mr-guided focused ultrasound: a clinical safety and feasibility study. Sci. Rep. 2019;9:321–322. doi: 10.1038/s41598-018-36340-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Meng Y., Hynynen K., Lipsman N. Applications of focused ultrasound in the brain: from thermoablation to drug delivery. Nat. Rev. Neurosci. 2021;17:7–22. doi: 10.1038/s41582-020-00418-z. [DOI] [PubMed] [Google Scholar]
  • 47.Zheng M., Ruan S., Liu S., Sun T., Qu D., Zhao H., Xie Z., Gao H., Jing X., Sun Z. Self-targeting fluorescent carbon dots for diagnosis of brain cancer cells. ACS Nano. 2015;9:11455–11461. doi: 10.1021/acsnano.5b05575. [DOI] [PubMed] [Google Scholar]
  • 48.Mo J., He L., Ma B., Chen T. Tailoring particle size of mesoporous silica nanosystem to antagonize glioblastoma and overcome blood-brain barrier. ACS Appl. Mater. Interfaces. 2016;8:6811–6825. doi: 10.1021/acsami.5b11730. [DOI] [PubMed] [Google Scholar]
  • 49.Béduneau A., Saulnier P., Benoit J.P. Active targeting of brain tumors using nanocarriers. Biomaterials. 2007;28:4947–4967. doi: 10.1016/j.biomaterials.2007.06.011. [DOI] [PubMed] [Google Scholar]
  • 50.Chaturbedy P., Kumar M., Salikolimi K., Das S., Sinha S.H., Chatterjee S., Suma B.S., Kundu T.K., Eswaramoorthy M. Shape-directed compartmentalized delivery of a nanoparticle-conjugated small-molecule activator of an epigenetic enzyme in the brain. J. Contr. Release. 2015;217:151–159. doi: 10.1016/j.jconrel.2015.08.043. [DOI] [PubMed] [Google Scholar]
  • 51.Barua S., Yoo J.W., Kolhar P., Wakankar A., Gokarn Y.R., Mitragotri S. Particle shape enhances specificity of antibody-displaying nanoparticles. Proc. Natl. Acad. Sci. USA. 2013;110:3270–3275. doi: 10.1073/pnas.1216893110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Herd H., Daum N., Jones A.T., Huwer H., Ghandehari H., Lehr C.M. Nanoparticle geometry and surface orientation influence mode of cellular uptake. ACS Nano. 2013;7:1961–1973. doi: 10.1021/nn304439f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Agarwal R., Singh V., Jurney P., Shi L., Sreenivasan S.V., Roy K. Mammalian cells preferentially internalize hydrogel nanodiscs over nanorods and use shape-specific uptake mechanisms. Proc. Natl. Acad. Sci. USA. 2013;110:17247–17252. doi: 10.1073/pnas.1305000110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Pérez-Martínez F.C., Guerra J., Posadas I., Ceña V. Barriers to non-viral vector-mediated gene delivery in the nervous system. Pharm. Res. 2011;28:1843–1858. doi: 10.1007/s11095-010-0364-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Chen F.H., Zhang L.M., Chen Q.T., Zhang Y., Zhang Z.J. Synthesis of a novel magnetic drug delivery system composed of doxorubicin-conjugated Fe3O4 nanoparticle cores and a peg-functionalized porous silica shell. Chem. 2010;46:8633–8635. doi: 10.1039/c0cc02577a. [DOI] [PubMed] [Google Scholar]
  • 56.Cho C.F., Wolfe J.M., Fadzen C.M., Calligaris D., Hornburg K., Chiocca E.A., Agar N.Y.R., Pentelute B.L., Lawler S.E. Blood-brain-barrier spheroids as an in vitro screening platform for brain-penetrating agents. Nat. Commun. 2017;8 doi: 10.1038/ncomms15623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Zhao X., Shang T., Zhang X., Ye T., Wang D., Rei L. Passage of magnetic tat-conjugated Fe (3)O(4) @SiO(2) nanoparticles across in vitro blood-brain barrier. Nanoscale Res. Lett. 2016;11:451. doi: 10.1186/s11671-016-1676-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Cheng Y., Dai Q., Morshed R.A., Fan X., Wegscheid M.L., Wainwright D.A., Han Y., Zhang L., Auffinger B., Tobias A.L., Rincón E., Thaci B., Ahmed A.U., Warnke P.C., He C., Lesniak M.S. Blood-brain barrier permeable gold nanoparticles: an efficient delivery platform for enhanced malignant glioma therapy and imaging. Small. 2014;10:5137–5150. doi: 10.1002/smll.201400654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.You Y., Wang N., He L., Shi C., Zhang D., Liu Y., Luo L., Chen T. Designing dual-functionalized carbon nanotubes with high blood-brain-barrier permeability for precise orthotopic glioma therapy. Dalton Trans. 2019;48:1569–1573. doi: 10.1039/c8dt03948h. [DOI] [PubMed] [Google Scholar]
  • 60.Chen Y., Liu L. Modern methods for delivery of drugs across the blood-brain barrier. Adv. Drug Deliv. Rev. 2012;64:640–665. doi: 10.1016/j.addr.2011.11.010. [DOI] [PubMed] [Google Scholar]
  • 61.Ulbrich K., Hekmatara T., Herbert E., Kreuter J. Transferrin- and transferrin-receptor-antibody-modified nanoparticles enable drug delivery across the blood-brain barrier (BBB) Eur. J. Pharm. Biopharm. 2009;71:251–256. doi: 10.1016/j.ejpb.2008.08.021. [DOI] [PubMed] [Google Scholar]
  • 62.Lucarelli M., Gennarelli M., Cardelli P., Novelli G., Scarpa S., Dallapiccola B., Strom R. Expression of receptors for native and chemically modified low-density lipoproteins in brain microvessels. FEBS Lett. 1997;401:53–58. doi: 10.1016/s0014-5793(96)01435-4. [DOI] [PubMed] [Google Scholar]
  • 63.Lu Y., Luo Q., Jia X., Tam J.P., Yang H., Shen Y., Li X. Multidisciplinary strategies to enhance therapeutic effects of flavonoids from Epimedii Folium: integration of herbal medicine, enzyme engineering, and nanotechnology. J Pharm Anal. 2023;13:239–254. doi: 10.1016/j.jpha.2022.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Ruan S., Yuan M., Zhang L., Hu G., Chen J., Cun X., Zhang Q., Yang Y., He Q., Gao H. Tumor microenvironment sensitive doxorubicin delivery and release to glioma using angiopep-2 decorated gold nanoparticles. Biomaterials. 2015;37:425–435. doi: 10.1016/j.biomaterials.2014.10.007. [DOI] [PubMed] [Google Scholar]
  • 65.Shen Z., Liu T., Li Y., Lau J., Yang Z., Fan W., Zhou Z., Shi C., Ke C., Bregadze V.I., Mandal S.K., Liu Y., Li Z., Xue T., Zhu G., Munasinghe J., Niu G., Wu A., Chen X. Fenton-reaction-acceleratable magnetic nanoparticles for ferroptosis therapy of orthotopic brain tumors. ACS Nano. 2018;12:11355–11365. doi: 10.1021/acsnano.8b06201. [DOI] [PubMed] [Google Scholar]
  • 66.Fang J.H., Chiu T.L., Huang W.C., Lai Y.H., Hu S.H., Chen Y.Y., Chen S.Y. Dual-targeting lactoferrin-conjugated polymerized magnetic polydiacetylene-assembled nanocarriers with self-responsive fluorescence/magnetic resonance imaging for in vivo brain tumor therapy. Adv Healthc Mater. 2016;5:688–695. doi: 10.1002/adhm.201500750. [DOI] [PubMed] [Google Scholar]
  • 67.Cui Y., Xu Q., Chow P.K., Wang D., Wang C.H. Transferrin-conjugated magnetic silica PLGA nanoparticles loaded with doxorubicin and paclitaxel for brain glioma treatment. Biomaterials. 2013;34:8511–8520. doi: 10.1016/j.biomaterials.2013.07.075. [DOI] [PubMed] [Google Scholar]
  • 68.Heggannavar G.B., Hiremath C.G., Achari D.D., Pangarkar V.G., Kariduraganavar M.Y. Development of doxorubicin-loaded magnetic silica-pluronic f-127 nanocarriers conjugated with transferrin for treating glioblastoma across the blood-brain barrier using an in vitro model. ACS Omega. 2018;3:8017–8026. doi: 10.1021/acsomega.8b00152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Heggannavar G.B., Vijeth S., Kariduraganavar M.Y. Development of dual drug loaded PLGA based mesoporous silica nanoparticles and their conjugation with angiopep-2 to treat glioma. J. Drug Deliv. Sci. Technol. 2019;53 [Google Scholar]
  • 70.Kuang J., Song W., Yin J., Zeng X., Han S., Zhao Y., Tao J., Liu C.-J., He X., Zhang X. IRGD modified chemo-immunotherapeutic nanoparticles for enhanced immunotherapy against glioblastoma. Adv. Funct. Mater. 2018;28:1–18. [Google Scholar]
  • 71.Ren J., Shen S., Wang D., Xi Z., Guo L., Pang Z., Qian Y., Sun X., Jiang X. The targeted delivery of anticancer drugs to brain glioma by PEGylated oxidized multi-walled carbon nanotubes modified with angiopep-2. Biomaterials. 2012;33:3324–3333. doi: 10.1016/j.biomaterials.2012.01.025. [DOI] [PubMed] [Google Scholar]
  • 72.Zhao L., Li Y., Zhu J., Sun N., Song N., Xing Y., Huang H., Zhao J. Chlorotoxin peptide-functionalized polyethylenimine-entrapped gold nanoparticles for glioma SPECT/CT imaging and radionuclide therapy. J. Nanobiotechnol. 2019;17:30–31. doi: 10.1186/s12951-019-0462-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Wang Y., Lin T., Zhang W., Jiang Y., Jin H., He H., Yang V.C., Chen Y., Huang Y. A prodrug-type, mmp-2-targeting nanoprobe for tumor detection and imaging. Theranostics. 2015;5:787–795. doi: 10.7150/thno.11139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Huang N., Cheng S., Zhang X., Tian Q., Pi J., Tang J., Huang Q., Wang F., Chen J., Xie Z., Xu Z., Chen W., Zheng H., Cheng Y. Efficacy of ngr peptide-modified PEGylated quantum dots for crossing the blood-brain barrier and targeted fluorescence imaging of glioma and tumor vasculature. Nanomedicine. 2017;13:83–93. doi: 10.1016/j.nano.2016.08.029. [DOI] [PubMed] [Google Scholar]
  • 75.Tang J., Huang N., Zhang X., Zhou T., Tan Y., Pi J., Pi L., Cheng S., Zheng H., Cheng Y. Aptamer-conjugated PEGylated quantum dots targeting epidermal growth factor receptor variant III for fluorescence imaging of glioma. Int. J. Nanomed. 2017;12:3899–3911. doi: 10.2147/IJN.S133166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Ren F., Liu H., Zhang H., Jiang Z., Xia B., Genevois C., He T., Allix M., Sun Q., Li Z., Gao M. Engineering NIR-IIb fluorescence of Er-based lanthanide nanoparticles for through-skull targeted imaging and imaging-guided surgery of orthotopic glioma. Nano Today. 2020;34 [Google Scholar]
  • 77.Lee C., Hwang H.S., Lee S., Kim B., Kim J.O., Oh K.T., Lee E.S., Choi H.G., Youn Y.S. Rabies virus-inspired silica-coated gold nanorods as a photothermal therapeutic platform for treating brain tumors. Adv. Mater. 2017;29:1605563. doi: 10.1002/adma.201605563. [DOI] [PubMed] [Google Scholar]
  • 78.Liu Z., Ren F., Zhang H., Yuan Q., Jiang Z., Liu H., Sun Q., Li Z. Boosting often overlooked long wavelength emissions of rare-earth nanoparticles for NIR-II fluorescence imaging of orthotopic glioblastoma. Biomaterials. 2019;219 doi: 10.1016/j.biomaterials.2019.119364. [DOI] [PubMed] [Google Scholar]
  • 79.Zhang H., Wang T., Qiu W., Han Y., Sun Q., Zeng J., Yan F., Zheng H., Li Z., Gao M. Monitoring the opening and recovery of the blood-brain barrier with noninvasive molecular imaging by biodegradable ultrasmall Cu (2-x) Se nanoparticles. Nano Lett. 2018;18:4985–4992. doi: 10.1021/acs.nanolett.8b01818. [DOI] [PubMed] [Google Scholar]
  • 80.Zhang H., Wang T., Liu H., Ren F., Qiu W., Sun Q., Yan F., Zheng H., Li Z., Gao M. Second near-infrared photodynamic therapy and chemotherapy of orthotopic malignant glioblastoma with ultra-small Cu(2-x) Se nanoparticles. Nanoscale. 2019;11:7600–7608. doi: 10.1039/c9nr01789e. [DOI] [PubMed] [Google Scholar]
  • 81.Recht L., Torres C.O., Smith T.W., Raso V., Griffin T.W. Transferrin receptor in normal and neoplastic brain tissue: implications for brain-tumor immunotherapy. J. Neurosurg. 1990;72:941–945. doi: 10.3171/jns.1990.72.6.0941. [DOI] [PubMed] [Google Scholar]
  • 82.Fang J., Nakamura H., Maeda H. The EPR effect: unique features of tumor blood vessels for drug delivery, factors involved, and limitations and augmentation of the effect. Adv. Drug Deliv. Rev. 2011;63:136–151. doi: 10.1016/j.addr.2010.04.009. [DOI] [PubMed] [Google Scholar]
  • 83.Kobayashi H., Watanabe R., Choyke P.L. Improving conventional enhanced permeability and retention (EPR) effects; what is the appropriate target? Theranostics. 2013;4:81–89. doi: 10.7150/thno.7193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Jones A.R., Shusta E.V. Blood-brain barrier transport of therapeutics via receptor-mediation. Pharm. Res. 2007;24:1759–1771. doi: 10.1007/s11095-007-9379-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Martell L.A., Agrawal A., Ross D.A., Muraszko K.M. Efficacy of transferrin receptor-targeted immunotoxins in brain tumor cell lines and pediatric brain tumors. Cancer Res. 1993;53:1348–1353. [PubMed] [Google Scholar]
  • 86.Bruneau A., Fortier M., Gagne F., Gagnon C., Turcotte P., Tayabali A., Davis T.A., Auffret M., Fournier M. In vitro immunotoxicology of quantum dots and comparison with dissolved cadmium and tellurium. Environ Toxicol. 2015;30:9–25. doi: 10.1002/tox.21890. [DOI] [PubMed] [Google Scholar]
  • 87.Tang Y., Han S., Liu H., Chen X., Huang L., Li X., Zhang J. The role of surface chemistry in determining in vivo biodistribution and toxicity of CdSe/ZnS core-shell quantum dots. Biomaterials. 2013;34:8741–8755. doi: 10.1016/j.biomaterials.2013.07.087. [DOI] [PubMed] [Google Scholar]
  • 88.Pasqualini R., Koivunen E., Kain R., Lahdenranta J., Sakamoto M., Stryhn A., Ashmun R.A., Shapiro L.H., Arap W., Ruoslahti E. Aminopeptidase is a receptor for tumor-homing peptides and a target for inhibiting angiogenesis. Cancer Res. 2000;60:722–727. [PMC free article] [PubMed] [Google Scholar]
  • 89.Di Matteo P., Arrigoni G.L., Alberici L., Corti A., Gallo-Stampino C., Traversari C., Doglioni C., Rizzardi G.P. Enhanced expression of CD13 in vessels of inflammatory and neoplastic tissues. J. Histochem. Cytochem. 2011;59:47–59. doi: 10.1369/jhc.2010.956644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Parmiani G., Pilla L., Corti A., Doglioni C., Cimminiello C., Bellone M., Parolini D., Russo V., Capocefalo F., Maccalli C. A pilot Phase I study combining peptide-based vaccination and NGR-hTNF vessel targeting therapy in metastatic melanoma. OncoImmunology. 2014;3 doi: 10.4161/21624011.2014.963406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Yang Y., Yang Y., Xie X., Cai X., Zhang H., Gong W., Wang Z., Mei X. PEGylated liposomes with NGR ligand and heat-activable cell-penetrating peptide-doxorubicin conjugate for tumor-specific therapy. Biomaterials. 2014;35:4368–4381. doi: 10.1016/j.biomaterials.2014.01.076. [DOI] [PubMed] [Google Scholar]
  • 92.Abakumov M.A., Semkina A.S., Skorikov A.S., Vishnevskiy D.A., Ivanova A.V., Mironova E., Davydova G.A., Majouga A.G., Chekhonin V.P. Toxicity of iron oxide nanoparticles: size and coating effects. J. Biochem. Mol. Toxicol. 2018;32 doi: 10.1002/jbt.22225. [DOI] [PubMed] [Google Scholar]
  • 93.Jo D.H., Kim J.H., Lee T.G., Kim J.H. Size, surface charge, and shape determine therapeutic effects of nanoparticles on brain and retinal diseases. Nanomedicine. 2015;11:1603–1611. doi: 10.1016/j.nano.2015.04.015. [DOI] [PubMed] [Google Scholar]
  • 94.Lu Q., Dai X., Zhang P., Tan X., Zhong Y., Yao C., Song M., Song G., Zhang Z., Peng G., Guo Z., Ge Y., Zhang K., Li Y. Fe3O4@Au composite magnetic nanoparticles modified with cetuximab for targeted magneto-photothermal therapy of glioma cells. Int. J. Nanomed. 2018;13:2491–2505. doi: 10.2147/IJN.S157935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Chen J., Sheng Z., Li P., Wu M., Zhang N., Yu X.F., Wang Y., Hu D., Zheng H., Wang G.P. Indocyanine green-loaded gold nanostars for sensitive SERS imaging and subcellular monitoring of photothermal therapy. Nanoscale. 2017;9:11888–11901. doi: 10.1039/c7nr02798b. [DOI] [PubMed] [Google Scholar]
  • 96.Xu D., Baidya A., Deng K., Li Y.S., Wu B., Xu H.B. Multifunctional nanoparticle PEG-Ce6-Gd for MRI-guided photodynamic therapy. Oncol. Rep. 2021;45:547–556. doi: 10.3892/or.2020.7871. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Chauhan A., Midha S., Kumar R., Meena R., Singh P., Jha S.K., Kuanr B.K. Rapid tumor inhibition via magnetic hyperthermia regulated by caspase 3 with time-dependent clearance of iron oxide nanoparticles. Biomater. Sci. 2021;9:2972–2990. doi: 10.1039/d0bm01705a. [DOI] [PubMed] [Google Scholar]
  • 98.Javid A., Ahmadian S., Saboury A.A., Kalantar S.M., Rezaei-Zarchi S. Chitosan-coated superparamagnetic iron oxide nanoparticles for doxorubicin delivery: synthesis and anticancer effect against human ovarian cancer cells. Chem. Biol. Drug Des. 2013;82:296–306. doi: 10.1111/cbdd.12145. [DOI] [PubMed] [Google Scholar]
  • 99.Hao J.J., Hu D., Chen P.X., Ma L., Xia M., Chen M.J., Xu H.F., Liu H.M., Li Y.W., Cheng J.J., Jin X., Duan P.F., Xu X.Q. Enantioselective theranostics of brain glioma using chiral quantum structures. Mater. Des. 2023;226 [Google Scholar]
  • 100.Ren Y., Miao C., Tang L., Liu Y., Ni P., Gong Y., Li H., Chen F., Feng S. Homotypic cancer cell membranes camouflaged nanoparticles for targeting drug delivery and enhanced chemo-photothermal therapy of glioma. Pharmaceutics. 2022;15:157. doi: 10.3390/ph15020157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Chen M.H., Liu T.Y., Chen Y.C., Chen M.H. Combining augmented radiotherapy and immunotherapy through a nano-gold and bacterial outer-membrane vesicle complex for the treatment of glioblastoma. Nanomaterials. 2021;11:1661. doi: 10.3390/nano11071661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Yim Y.S., Choi J.S., Kim G.T., Kim C.H., Shin T.H., Kim D.G., Cheon J. A facile approach for the delivery of inorganic nanoparticles into the brain by passing through the blood-brain barrier (BBB) Chem. 2012;48:61–63. doi: 10.1039/c1cc15113d. [DOI] [PubMed] [Google Scholar]
  • 103.Taheri A., Cheniany M., Ganjeali A., Arefi-Oskouie A. ICP-OES assessment of trace and toxic elements in Ziziphora clinopodioides Lam. from Iran by chemometric approaches. Biometals:an international journal on the role of metal ions in biology, biochemistry, and medicine. 2022;35:1169–1186. doi: 10.1007/s10534-022-00433-1. [DOI] [PubMed] [Google Scholar]
  • 104.Vievard J., Amoikon T.L., Coulibaly N.A., Devouge-Boyer C., Arellano-Sánchez M.G., Aké M.F.D., Djeni N.T., Mignot M. Extraction and quantification of pesticides and metals in palm wines by HS-SPME/GC-MS and ICP-AES/MS. Food Chem. 2022;393 doi: 10.1016/j.foodchem.2022.133352. [DOI] [PubMed] [Google Scholar]
  • 105.Mittal M., Kumar K., Anghore D., Rawal R.K. ICP-MS: analytical method for identification and detection of elemental impurities. Curr. Drug Discovery Technol. 2017;14:106–120. doi: 10.2174/1570163813666161221141402. [DOI] [PubMed] [Google Scholar]
  • 106.Tu Y., Chen N., Yang W., Bao Y., Xu H., Qin S. BSA capped Au nanoparticle as an efficient sensitizer for glioblastoma tumor radiation therapy. RSC Adv. 2015;5:40514–40520. [Google Scholar]
  • 107.Rees J. Advances in magnetic resonance imaging of brain tumors. Curr. Opin. Neurobiol. 2003;16:643–650. doi: 10.1097/01.wco.0000102626.38669.b9. [DOI] [PubMed] [Google Scholar]
  • 108.Maeda H., Wu J., Sawa T., Matsumura Y., Hori K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J. Contr. Release. 2000;65:271–284. doi: 10.1016/s0168-3659(99)00248-5. [DOI] [PubMed] [Google Scholar]
  • 109.Shao Y., Yang G., Lin J., Fan X., Guo Y., Zhu W., Cai Y., Huang H., Hu D., Pang W., Liu Y., Li Y., Cheng J., Xu X. Shining light on chiral inorganic nanomaterials for biological issues. Theranostics. 2021;11:9262–9295. doi: 10.7150/thno.64511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Marcucci F., Corti A. How to improve exposure of tumor cells to drugs: promoter drugs increase tumor uptake and penetration of effector drugs. Adv. Drug Deliv. Rev. 2012;64:53–68. doi: 10.1016/j.addr.2011.09.007. [DOI] [PubMed] [Google Scholar]
  • 111.Miri A., Najafzadeh H., Darroudi M., Miri M.J., Kouhbanani M.A.J., Sarani M. Iron oxide nanoparticles: biosynthesis, magnetic behavior, cytotoxic effect. Chemistry. 2021;10:327–333. doi: 10.1002/open.202000186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Jung H., Park B., Lee C., Cho J., Suh J., Park J., Kim Y., Kim J., Cho G., Cho H. Dual MRI T1 and T2(∗) contrast with size-controlled iron oxide nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2014;10:1679–1689. doi: 10.1016/j.nano.2014.05.003. [DOI] [PubMed] [Google Scholar]
  • 113.Wang G., Gao W., Zhang X., Mei X. Au nanocage functionalized with ultra-small Fe3O4 nanoparticles for targeting T1-T2dual MRI and CT imaging of tumor. Sci. Rep. 2016;6 doi: 10.1038/srep28258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.So Y.H., Lee W., Park E.A., Kim P.K. Investigation of the characteristics of new, uniform, extremely small iron-based nanoparticles as T1 contrast agents for MRI. Korean J. Radiol. 2021;22:1708–1718. doi: 10.3348/kjr.2020.1455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Neuwelt E.A., Várallyay P., Bagó A.G., Muldoon L.L., Nesbit G., Nixon R. Imaging of iron oxide nanoparticles by MR and light microscopy in patients with malignant brain tumors. Neuropathol. Appl. Neurobiol. 2004;30:456–471. doi: 10.1111/j.1365-2990.2004.00557.x. [DOI] [PubMed] [Google Scholar]
  • 116.Varallyay P., Nesbit G., Muldoon L.L., Nixon R.R., Delashaw J., Cohen J.I., Petrillo A., Rink D., Neuwelt E.A. Comparison of two superparamagnetic viral-sized iron oxide particles ferumoxides and ferumoxtran-10 with a gadolinium chelate in imaging intracranial tumors. AJNR Am J Neuroradiol. 2002;23:510–519. [PMC free article] [PubMed] [Google Scholar]
  • 117.Cole A.J., David A.E., Wang J., Galbán C.J., Hill H.L., Yang V.C. Polyethylene glycol modified, cross-linked starch-coated iron oxide nanoparticles for enhanced magnetic tumor targeting. Biomaterials. 2011;32:2183–2193. doi: 10.1016/j.biomaterials.2010.11.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Cole A.J., David A.E., Wang J., Galbán C.J., Yang V.C. Magnetic brain tumor targeting and biodistribution of long-circulating PEG-modified, cross-linked starch-coated iron oxide nanoparticles. Biomaterials. 2011;32:6291–6301. doi: 10.1016/j.biomaterials.2011.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Faucher L., Guay-Bégin A.A., Lagueux J., Côté M.F., Petitclerc E., Fortin M.A. Ultra-small gadolinium oxide nanoparticles to image brain cancer cells in vivo with MRI. Contrast Media Mol. Imaging. 2011;6:209–218. doi: 10.1002/cmmi.420. [DOI] [PubMed] [Google Scholar]
  • 120.Shen Z., Song J., Zhou Z., Yung B.C., Aronova M.A., Li Y., Dai Y., Fan W., Liu Y., Li Z., Ruan H., Leapman R.D., Lin L., Niu G., Chen X., Wu A. Dotted core-shell nanoparticles for T1 -weighted MRI of tumors. Adv. Mater. 2018;30 doi: 10.1002/adma.201803163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Park J.Y., Baek M.J., Choi E.S., Woo S., Kim J.H., Kim T.J., Jung J.C., Chae K.S., Chang Y., Lee G.H. Paramagnetic ultrasmall gadolinium oxide nanoparticles as advanced T1 MRI contrast agent: account for large longitudinal relaxivity, optimal particle diameter, and in vivo T1 MR images. ACS Nano. 2009;3:3663–3669. doi: 10.1021/nn900761s. [DOI] [PubMed] [Google Scholar]
  • 122.Idée J.M., Port M., Raynal I., Schaefer M., Le Greneur S., Corot C. Clinical and biological consequences of transmetallation induced by contrast agents for magnetic resonance imaging: a review. Fundam Clin Pharmacol. 2006;20:563–576. doi: 10.1111/j.1472-8206.2006.00447.x. [DOI] [PubMed] [Google Scholar]
  • 123.De Stasio G., Casalbore P., Pallini R., Gilbert B., Sanità F., Ciotti M.T., Rosi G., Festinesi A., Larocca L.M., Rinelli A., Perret D., Mogk D.W., Perfetti P., Mehta M.P., Mercanti D. Gadolinium in human glioblastoma cells for gadolinium neutron capture therapy. Cancer Res. 2001;61:4272–4277. [PubMed] [Google Scholar]
  • 124.Coluccia D., Figueiredo C.A., Wu M.Y., Riemenschneider A.N., Diaz R., Luck A., Smith C., Das S., Ackerley C., O’Reilly M., Hynynen K., Rutka J.T. Enhancing glioblastoma treatment using cisplatin-gold-nanoparticle conjugates and targeted delivery with magnetic resonance-guided focused ultrasound. Nanomedicine. 2018;14:1137–1148. doi: 10.1016/j.nano.2018.01.021. [DOI] [PubMed] [Google Scholar]
  • 125.Tomanek B., Iqbal U., Blasiak B., Abulrob A., Albaghdadi H., Matyas J.R., Ponjevic D., Sutherland G.R. Evaluation of brain tumor vessels specific contrast agents for glioblastoma imaging. Neuro Oncol. 2012;14:53–63. doi: 10.1093/neuonc/nor183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Pasi F., Persico M.G., Buroni F.E., Aprile C., Hodolic M., Corbella F., Nano R., Facoetti A., Lodola L. Uptake of 18F-FET and 18F-FCH in human glioblastoma T98G cell line after irradiation with photons or carbon ions. Contrast Media Mol. Imaging. 2017. 2017:6491674. doi: 10.1155/2017/6491674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Yoshii Y. Pathological review of late cerebral radionecrosis. Brain Tumor Pathol. 2008;25:51–58. doi: 10.1007/s10014-008-0233-9. [DOI] [PubMed] [Google Scholar]
  • 128.Herrera-Rivero M., Heneka M.T., Papadopoulos V. Translocator protein and new targets for neuroinflammation. Clin. Transl. Med. 2015;3:391–402. [Google Scholar]
  • 129.Cui Y., Yang J., Zhou Q., Liang P., Wang Y., Gao X., Wang Y. Renal clearable ag nanodots for in vivo computer tomography imaging and photothermal therapy. ACS Appl. Mater. Interfaces. 2017;9:5900–5906. doi: 10.1021/acsami.6b16133. [DOI] [PubMed] [Google Scholar]
  • 130.Liu Z., Ju E., Liu J., Du Y., Li Z., Yuan Q., Ren J., Qu X. Direct visualization of gastrointestinal tract with lanthanide-doped BaYbF5 upconversion nanoprobes. Biomaterials. 2013;34:7444–7452. doi: 10.1016/j.biomaterials.2013.06.060. [DOI] [PubMed] [Google Scholar]
  • 131.Wei B., Zhang X., Zhang C., Jiang Y., Fu Y.Y., Yu C., Sun S.K., Yan X.P. Facile synthesis of uniform-sized bismuth nanoparticles for CT visualization of gastrointestinal tract in vivo. ACS Appl. Mater. Interfaces. 2016;8:12720–12726. doi: 10.1021/acsami.6b03640. [DOI] [PubMed] [Google Scholar]
  • 132.Cheng Y., Lu H., Yang F., Zhang Y., Dong H. Biodegradable FeWO(x) nanoparticles for CT/MR imaging-guided synergistic photothermal, photodynamic, and chemodynamic therapy. Nanoscale. 2021;13:3049–3060. doi: 10.1039/d0nr07215j. [DOI] [PubMed] [Google Scholar]
  • 133.Attia M.F., Wallyn J., Anton N., Vandamme T.F. Inorganic nanoparticles for X-ray computed tomography imaging. Crit. Rev. Ther. Drug Carrier Syst. 2018;35:391–431. doi: 10.1615/CritRevTherDrugCarrierSyst.2018020974. [DOI] [PubMed] [Google Scholar]
  • 134.Cohen-Inbar O., Zaaroor M. Glioblastoma multiforme targeted therapy: the chlorotoxin story. J. Clin. Neurosci.: official journal of the Neurosurgical Society of Australasia. 2016;33:52–58. doi: 10.1016/j.jocn.2016.04.012. [DOI] [PubMed] [Google Scholar]
  • 135.Wu X.S., Jian X.C., Yin B., He Z.J. Development of the research on the application of chlorotoxin in imaging diagnostics and targeted therapies for tumors. Chin. J. Cancer. 2010;29:626–630. doi: 10.5732/cjc.009.10359. [DOI] [PubMed] [Google Scholar]
  • 136.Curry T., Kopelman R., Shilo M., Popovtzer R. Multifunctional theranostic gold nanoparticles for targeted CT imaging and photothermal therapy. Contrast Media Mol. Imaging. 2014;9:53–61. doi: 10.1002/cmmi.1563. [DOI] [PubMed] [Google Scholar]
  • 137.Mansoor N.M., Thust S., Militano V., Fraioli F. PET imaging in glioma: techniques and current evidence. J. Nucl. Med. 2018;39:1064–1080. doi: 10.1097/MNM.0000000000000914. [DOI] [PubMed] [Google Scholar]
  • 138.Weissleder R., Mahmood U. Molecular imaging. Radiology. 2001;219:316–333. doi: 10.1148/radiology.219.2.r01ma19316. [DOI] [PubMed] [Google Scholar]
  • 139.Goel S., Chen F., Hong H., Valdovinos H.F., Hernandez R., Shi S., Barnhart T.E., Cai W. VEGF₁₂₁-conjugated mesoporous silica nanoparticle: a tumor targeted drug delivery system. ACS Appl. Mater. Interfaces. 2014;6:21677–21685. doi: 10.1021/am506849p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Wang H., Chen K., Niu G., Chen X. Site-specifically biotinylated VEGF121 for near-infrared fluorescence imaging of tumor angiogenesis. Mol. Pharm. 2009;6:285–294. doi: 10.1021/mp800185h. [DOI] [PubMed] [Google Scholar]
  • 141.Yang X., Stein E.W., Ashkenazi S., Wang L.V. Nanoparticles for photoacoustic imaging. Wiley interdisciplinary reviews. Artif Cells Nanomed Biotechnol. 2009;1:360–368. doi: 10.1002/wnan.42. [DOI] [PubMed] [Google Scholar]
  • 142.Nie L., Chen X. Structural and functional photoacoustic molecular tomography aided by emerging contrast agents. Chem. Soc. Rev. 2014;43:7132–7170. doi: 10.1039/c4cs00086b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Fan Q., Cheng K., Yang Z., Zhang R., Yang M., Hu X., Ma X., Bu L., Lu X., Xiong X., Huang W., Zhao H., Cheng Z. Perylene-diimide-based nanoparticles as highly efficient photoacoustic agents for deep brain tumor imaging in living mice. Adv. Mater. 2015;27:843–847. doi: 10.1002/adma.201402972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Kircher M.F., de la Zerda A., Jokerst J.V., Zavaleta C.L., Kempen P.J., Mittra E., Pitter K., Huang R., Campos C., Habte F., Sinclair R., Brennan C.W., Mellinghoff I.K., Holland E.C., Gambhir S.S. A brain tumor molecular imaging strategy using a new triple-modality MRI-photoacoustic-Raman nanoparticle. Nat. Med. 2012;18:829–834. doi: 10.1038/nm.2721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Lu W., Melancon M.P., Xiong C., Huang Q., Elliott A., Song S., Zhang R., Flores L.G., Gelovani J.G., Wang L.V., Ku G., Stafford R.J., Li C. Effects of photoacoustic imaging and photothermal ablation therapy mediated by targeted hollow gold nanospheres in an orthotopic mouse xenograft model of glioma. Cancer Res. 2011;71:6116–6121. doi: 10.1158/0008-5472.CAN-10-4557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Wang S., Li X., Chen Y., Cai X., Yao H., Gao W., Zheng Y., An X., Shi J., Chen H. A facile one-pot synthesis of a two-dimensional MoS2/Bi2S3 composite theranostic nanosystem for multi-modality tumor imaging and therapy. Adv. Mater. 2015;27:2775–2782. doi: 10.1002/adma.201500870. [DOI] [PubMed] [Google Scholar]
  • 147.Hao J., Song G., Liu T., Yi X., Yang K., Cheng L., Liu Z. In vivo long-term biodistribution, excretion, and toxicology of PEGylated transition-metal dichalcogenides MS2(M = Mo, W, Ti) nanosheets. Adv. Sci. 2017;4 doi: 10.1002/advs.201600160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Chen Y., Tan C., Zhang H., Wang L. Two-dimensional graphene analogues for biomedical applications. Chem. Soc. Rev. 2015;44:2681–2701. doi: 10.1039/c4cs00300d. [DOI] [PubMed] [Google Scholar]
  • 149.Chen J., Liu C., Hu D., Wang F., Haiwei W., Gong X., Liu X., Song L., Sheng Z., Zheng H. Single-layer MoS2 nanosheets with amplified photoacoustic effect for highly sensitive photoacoustic imaging of orthotopic brain tumors. Adv. Funct. Mater. 2016;26:26–27. [Google Scholar]
  • 150.Wang H., Liu C., Gong X., Hu D., Lin R., Sheng Z., Zheng C., Yan M., Chen J., Cai L., Song L. In vivo photoacoustic molecular imaging of breast carcinoma with folate receptor-targeted indocyanine green nanoprobes. Nanoscale. 2014;6:14270–14279. doi: 10.1039/c4nr03949a. [DOI] [PubMed] [Google Scholar]
  • 151.Castellanos-Gomez A., Poot M., Steele G.A., van der Zant H.S., Agraït N., Rubio-Bollinger G. Elastic properties of freely suspended MoS2 nanosheets. Adv. Mater. 2012;24:772–775. doi: 10.1002/adma.201103965. [DOI] [PubMed] [Google Scholar]
  • 152.Yang F., Chen Z., Xing D. All-optical noncontact phase-domain photoacoustic elastography. Opt Lett. 2021;46:5063–5066. doi: 10.1364/OL.438599. [DOI] [PubMed] [Google Scholar]
  • 153.Singh M.S., Jiang H. Elastic property attributes to photoacoustic signals: an experimental phantom study. Opt Lett. 2014;39:3970–3973. doi: 10.1364/OL.39.003970. [DOI] [PubMed] [Google Scholar]
  • 154.Cheng L., Shen S., Shi S., Yi Y., Wang X., Song G., Yang K., Liu G., Barnhart T.E., Cai W., Liu Z. FeSe2-decorated Bi2Se3 nanosheets fabricated via cation exchange for chelator-free 64cu-labeling and multimodal image-guided photothermal-radiation therapy. Adv. Funct. Mater. 2016;26:2185–2197. doi: 10.1002/adfm.201504810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Ain N.U., Abdul Nasir J., Khan Z., Butler I.S., Rehman Z. Copper sulfide nanostructures: synthesis and biological applications. RSC Adv. 2022;12:7550–7567. doi: 10.1039/d1ra08414c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Chen Z., Wang Q., Wang H., Zhang L., Song G., Song L., Hu J., Wang H., Liu J., Zhu M., Zhao D. Ultrathin PEGylated W18O49 nanowires as a new 980 nm-laser-driven photothermal agent for efficient ablation of cancer cells in vivo. Adv. Mater. 2013;25:2095–2100. doi: 10.1002/adma.201204616. [DOI] [PubMed] [Google Scholar]
  • 157.Pillarisetti S., Uthaman S., Huh K.M., Koh Y.S., Lee S., Park I.K. Multimodal composite iron oxide nanoparticles for biomedical applications. Tissue Eng Regen Med. 2019;16:451–465. doi: 10.1007/s13770-019-00218-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Stummer W., Suero Molina E. Fluorescence imaging/agents in tumor resection. Neurosurg. Clin. 2017;28:569–583. doi: 10.1016/j.nec.2017.05.009. [DOI] [PubMed] [Google Scholar]
  • 159.Rhyner M.N., Smith A.M., Gao X., Mao H., Yang L., Nie S. Quantum dots and multifunctional nanoparticles: new contrast agents for tumor imaging. Nanomedicine. 2006;1:209–217. doi: 10.2217/17435889.1.2.209. [DOI] [PubMed] [Google Scholar]
  • 160.Zhang J., Hao G., Yao C., Hu S., Hu C., Zhang B. Paramagnetic albumin decorated CuInS2/ZnS QDs for CD133(+) glioma bimodal MR/fluorescence targeted imaging. J. Mater. Chem. B. 2016;4:4110–4118. doi: 10.1039/c6tb00834h. [DOI] [PubMed] [Google Scholar]
  • 161.Ag Seleci D., Maurer V., Barlas F.B., Porsiel J.C., Temel B., Ceylan E., Timur S., Stahl F., Scheper T., Garnweitner G. Transferrin-decorated niosomes with integrated InP/ZnS quantum dots and magnetic iron oxide nanoparticles: dual targeting and imaging of glioma. Int. J. Mol. Sci. 2021;22:4556. doi: 10.3390/ijms22094556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Mansur A.A.P., Caires A.J., Carvalho S.M., Capanema N.S.V., Carvalho I.C., Mansur H.S. Dual-functional supramolecular nanohybrids of quantum dot/biopolymer/chemotherapeutic drug for bioimaging and killing brain cancer cells. in vitro. Colloids Surf. B. 2019;184 doi: 10.1016/j.colsurfb.2019.110507. [DOI] [PubMed] [Google Scholar]
  • 163.Miller M.A., Arlauckas S., Weissleder R. Prediction of anti-cancer nanotherapy efficacy by imaging. Nanotheranostics. 2017;6:296–312. doi: 10.7150/ntno.20564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Mura S., Couvreur P. Nanotheranostics for personalized medicine. Adv. Drug Deliv. Rev. 2012;64:1394–1416. doi: 10.1016/j.addr.2012.06.006. [DOI] [PubMed] [Google Scholar]
  • 165.Luo K., Liu G., Zhang X., She W., He B., Nie Y., Li L., Wu Y., Zhang Z., Gong Q., Gao F., Song B., Ai H., Gu Z. Functional L-lysine dendritic macromolecules as liver-imaging probes. Macromol. Biosci. 2009;9:1227–1236. doi: 10.1002/mabi.200900231. [DOI] [PubMed] [Google Scholar]
  • 166.Slack R.J., Macdonald S.J.F., Roper J.A., Jenkins R.G., Hatley R.J.D. Emerging therapeutic opportunities for integrin inhibitors. Nat. Rev. Drug Discov. 2022;21:60–78. doi: 10.1038/s41573-021-00284-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Sorensen A.G., Batchelor T.T., Wen P.Y., Zhang W.T., Jain R.K. Response criteria for glioma. Nat. Clin. Pract. Oncol. 2008;5:634–644. doi: 10.1038/ncponc1204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Zhang H., Tian Y., Zhu Z., Xu H., Li X., Zheng D., Sun W. Efficient antitumor effect of co-drug-loaded nanoparticles with gelatin hydrogel by local implantation. Sci. Rep. 2016;6 doi: 10.1038/srep26546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Zhang J., Wang L., Fai Chan H., Xie W., Chen S., He C., Wang Y., Chen M. Co-delivery of paclitaxel and tetrandrine via iRGD peptide conjugated lipid-polymer hybrid nanoparticles overcome multidrug resistance in cancer cells. Sci. Rep. 2017;7 doi: 10.1038/srep46057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Chen X., Niu S., Bremner D.H., Zhang X., Zhang H., Zhang Y., Li S., Zhu L.M. Co-delivery of doxorubicin and oleanolic acid by triple-sensitive nanocomposite based on chitosan for effective promoting tumor apoptosis. Carbohydr. Polym. 2020;247 doi: 10.1016/j.carbpol.2020.116672. [DOI] [PubMed] [Google Scholar]
  • 171.Ruan S., He Q., Gao H. Matrix metalloproteinase triggered size-shrinkable gelatin-gold fabricated nanoparticles for tumor microenvironment sensitive penetration and diagnosis of glioma. Nanoscale. 2015;7:9487–9496. doi: 10.1039/c5nr01408e. [DOI] [PubMed] [Google Scholar]
  • 172.Nguyen N.T., Nguyen N.N.T., Tran N.T.N., Le P.N., Nguyen T.B.T., Nguyen N.H., Bach L.G., Doan V.N., Tran H.L.B., Le V.T., Tran N.Q. Synergic activity against MCF-7 breast cancer cell growth of nanocurcumin-encapsulated and cisplatin-complexed nanogels. Molecules. 2018;23:3347. doi: 10.3390/molecules23123347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Rudnik L.A.C., Farago P.V., Manfron Budel J., Lyra A., Barboza F.M., Klein T., Kanunfre C.C., Nadal J.M., Bandéca M.C., Raman V., Novatski A., Loguércio A.D., Zanin S.M.W. Co-loaded curcumin and methotrexate nanocapsules enhance cytotoxicity against non-small-cell lung cancer cells. Molecules. 2020;25:1913. doi: 10.3390/molecules25081913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Fang J.H., Lai Y.H., Chiu T.L., Chen Y.Y., Hu S.H., Chen S.Y. Magnetic core-shell nanocapsules with dual-targeting capabilities and co-delivery of multiple drugs to treat brain gliomas. Adv Healthc Mater. 2014;3:1250–1260. doi: 10.1002/adhm.201300598. [DOI] [PubMed] [Google Scholar]
  • 175.Mussi S.V., Sawant R., Perche F., Oliveira M.C., Azevedo R.B., Ferreira L.A., Torchilin V.P. Novel nanostructured lipid carrier co-loaded with doxorubicin and docosahexaenoic acid demonstrates enhanced in vitro activity and overcomes drug resistance in MCF-7/ADR cells. Pharm. Res. 2014;31:1882–1892. doi: 10.1007/s11095-013-1290-2. [DOI] [PubMed] [Google Scholar]
  • 176.Liu Y., Fang J., Kim Y.J., Wong M.K., Wang P. Codelivery of doxorubicin and paclitaxel by cross-linked multilamellar liposome enables synergistic antitumor activity. Mol. Pharm. 2014;11:1651–1661. doi: 10.1021/mp5000373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Mansur H.S., Mansur A.A.P., Soriano-Araújo A., Lobato Z.I.P. Beyond biocompatibility: an approach for the synthesis of ZnS quantum dot-chitosan nano-immunoconjugates for cancer diagnosis. Green Chem. 2015;17:1820–1830. [Google Scholar]
  • 178.Wang J., Li W., Wang H., Ma Q., Li S., Chang H.M., Jameel H. Liquefaction of Kraft Lignin by Hydrocracking with Simultaneous Use of a Novel Dual Acid-Base Catalyst and a Hydrogenation Catalyst. Bioresour Technol. 2017;243:100–106. doi: 10.1016/j.biortech.2017.06.024. [DOI] [PubMed] [Google Scholar]
  • 179.Komeri R., S M., B S U., Sreekutty J., G U P., Maiti K., Tt S. Enhancing therapeutic efficacy of doxorubicin via efflux-pump modulating, galactoxyloglucan endowed gold nanocarrier to confront drug resistant brain tumor. Sreelekha ACS Applied Nano Materials. 2019;2:6287–6299. [Google Scholar]
  • 180.Joseph M.M., Aravind S.R., Varghese S., Mini S., Sreelekha T.T. PST-gold nanoparticle as an effective anticancer agent with immunomodulatory properties. Colloids Surf. B Biointerfaces. 2013;104:32–39. doi: 10.1016/j.colsurfb.2012.11.046. [DOI] [PubMed] [Google Scholar]
  • 181.Li S.D., Huang L. Pharmacokinetics and biodistribution of nanoparticles. Mol. Pharm. 2008;5:496–504. doi: 10.1021/mp800049w. [DOI] [PubMed] [Google Scholar]
  • 182.Afzalipour R., Khoei S., Khoee S., Shirvalilou S., Raoufi N.J., Motevalian M., Karimi M.Y. Thermosensitive magnetic nanoparticles exposed to alternating magnetic field and heat-mediated chemotherapy for an effective dual therapy in rat glioma model. Nanomedicine. 2021;31 doi: 10.1016/j.nano.2020.102319. [DOI] [PubMed] [Google Scholar]
  • 183.Afzalipour R., Khoei S., Khoee S., Shirvalilou S., Jamali, Raoufi N., Motevalian M., Karimi M.R. Dual-targeting temozolomide loaded in folate-conjugated magnetic triblock copolymer nanoparticles to improve the therapeutic efficiency of rat brain gliomas. ACS Biomater. Sci. Eng. 2019;5:6000–6011. doi: 10.1021/acsbiomaterials.9b00856. [DOI] [PubMed] [Google Scholar]
  • 184.Cheng Y., Muroski M.E., Petit D., Mansell R., Vemulkar T., Morshed R.A., Han Y., Balyasnikova I.V., Horbinski C.M., Huang X., Zhang L., Cowburn R.P., Lesniak M.S. Rotating magnetic field induced oscillation of magnetic particles for in vivo mechanical destruction of malignant glioma. J. Contr. Release. 2016;223:75–84. doi: 10.1016/j.jconrel.2015.12.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Koul D., Shen R., Bergh S., Sheng X., Shishodia S., Lafortune T.A., Lu Y., de Groot J.F., Mills G.B., Yung W.K. Inhibition of Akt survival pathway by a small-molecule inhibitor in human glioblastoma. Mol. Cancer Therapeut. 2006;5:637–644. doi: 10.1158/1535-7163.MCT-05-0453. [DOI] [PubMed] [Google Scholar]
  • 186.Babaei M., Ganjalikhani M. The potential effectiveness of nanoparticles as radio sensitizers for radiotherapy. Bioimpacts. 2014;4:15–20. doi: 10.5681/bi.2014.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Shilo M., Sharon A., Baranes K., Motiei M., Lellouche J.P., Popovtzer R. The effect of nanoparticle size on the probability to cross the blood-brain barrier: an in-vitro endothelial cell model. J. Nanobiotechnol. 2015;13:19. doi: 10.1186/s12951-015-0075-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Liu P., Jin H., Guo Z., Ma J., Zhao J., Li D., Wu H., Gu N. Silver nanoparticles outperform gold nanoparticles in radiosensitizing U251 cells in vitro and in an intracranial mouse model of glioma. Int. J. Nanomed. 2016;11:5003–5014. doi: 10.2147/IJN.S115473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Nahas S.A., Gatti R.A. DNA double strand break repair defects, primary immunodeficiency disorders, and 'radiosensitivity'. Curr. Opin. Allergy Clin. Immunol. 2009;9:510–516. doi: 10.1097/ACI.0b013e328332be17. [DOI] [PubMed] [Google Scholar]
  • 190.Lu V.M., Jue T.R., McDonald K.L. Cytotoxic lanthanum oxide nanoparticles sensitize glioblastoma cells to radiation therapy and temozolomide: an in vitro rationale for translational studies. Sci. Rep. 2020;10 doi: 10.1038/s41598-020-75372-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Han H. RNA interference to knock down gene expression. Methods Mol. Biol. 2018;1706:293–302. doi: 10.1007/978-1-4939-7471-9_16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Kanasty R., Dorkin J.R., Vegas A., Anderson D. Delivery materials for siRNA therapeutics. Nat. Mater. 2013;12:967–977. doi: 10.1038/nmat3765. [DOI] [PubMed] [Google Scholar]
  • 193.Yang X.Z., Du J.Z., Dou S., Mao C.Q., Long H.Y., Wang J. Sheddable ternary nanoparticles for tumor acidity-targeted siRNA delivery. ACS Nano. 2012;6:771–781. doi: 10.1021/nn204240b. [DOI] [PubMed] [Google Scholar]
  • 194.Lin G., Chen T., Zou J., Wang Y., Wang X., Li J., Huang Q., Fu Z., Zhao Y., Lin M.C., Xu G., Yong K.T. Quantum dots-siRNA nanoplexes for gene silencing in central nervous system tumor cells. Front. Pharmacol. 2017;8:182. doi: 10.3389/fphar.2017.00182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Osterhage J.L., Friedman K.L. Chromosome end maintenance by telomerase. J. Biol. Chem. 2009;284:16061–16065. doi: 10.1074/jbc.R900011200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Tong W.Y., Alnakhli M., Bhardwaj R., Apostolou S., Sinha S., Fraser C., Kuchel T., Kuss B., Voelcker N.H. Delivery of siRNA in vitro and in vivo using PEI-capped porous silicon nanoparticles to silence MRP1 and inhibit proliferation in glioblastoma. J. Nanobiotechnol. 2018;16:38. doi: 10.1186/s12951-018-0365-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Munoz M., Henderson M., Haber M., Norris M. Role of the MRP1/ABCC1 multidrug transporter protein in cancer. IUBMB Life. 2007;59:752–757. doi: 10.1080/15216540701736285. [DOI] [PubMed] [Google Scholar]
  • 198.Cuylen S., Blaukopf C., Politi A.Z., Müller-Reichert T., Neumann B., Poser I., Ellenberg J., Hyman A.A., Gerlich D.W. Ki-67 acts as a biological surfactant to disperse mitotic chromosomes. Nature. 2016;535:308–312. doi: 10.1038/nature18610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Day E.S., Thompson P.A., Zhang L., Lewinski N.A., Ahmed N., Drezek R.A., Blaney S.M., West J.L. Nanoshell-mediated photothermal therapy improves survival in a murine glioma model. J. Neuro Oncol. 2011;104:55–63. doi: 10.1007/s11060-010-0470-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Tang X.L., Jing F., Lin B.L., Cui S., Yu R.T., Shen X.D., Wang T.W. PH-responsive magnetic mesoporous silica-based nanoplatform for synergistic photodynamic therapy/chemotherapy. ACS Appl. Mater. Interfaces. 2018;10:15001–15011. doi: 10.1021/acsami.7b19797. [DOI] [PubMed] [Google Scholar]
  • 201.Henderson B.W., Dougherty T. J. Dougherty. How does photodynamic therapy work? Photochem. Photobiol. 1992;55:145–157. doi: 10.1111/j.1751-1097.1992.tb04222.x. [DOI] [PubMed] [Google Scholar]
  • 202.Mang T.S. Lasers and light sources for PDT: past, present and future. Photodiagnosis and photodynamic therapy. Photodiagnosis Photodyn. Ther. 2004;1:43–48. doi: 10.1016/S1572-1000(04)00012-2. [DOI] [PubMed] [Google Scholar]
  • 203.Agostinis P., Berg K., Cengel K.A., Foster T.H., Girotti A.W., Gollnick S.O., Hahn S.M., Hamblin M.R., Juzeniene A., Kessel D., Korbelik M., Moan J., Mroz P., Nowis D., Piette J., Wilson B.C., Golab J. Photodynamic therapy of cancer: an update. CA Cancer J. Clin. 2011;61:250–281. doi: 10.3322/caac.20114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Hou Y.J., Yang X.X., Liu R.Q., Zhao D., Guo C.X., Zhu A.C., Wen M.N., Liu Z., Qu G.F., Meng H.X. Pathological mechanism of photodynamic therapy and photothermal therapy based on nanoparticles. Int. J. Nanomed. 2020;15:6827–6838. doi: 10.2147/IJN.S269321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Norouzi M., Nazari B., Miller D.W. Injectable hydrogel-based drug delivery systems for local cancer therapy. Drug Discov. Today. 2016;21:1835–1849. doi: 10.1016/j.drudis.2016.07.006. [DOI] [PubMed] [Google Scholar]
  • 206.Cardoso V.F., Francesko A., Ribeiro C., Bañobre-López M., Martins P., Lanceros-Mendez S. Advances in magnetic nanoparticles for biomedical applications. Adv Healthc Mater. 2018;7:1700845. doi: 10.1002/adhm.201700845. [DOI] [PubMed] [Google Scholar]
  • 207.S M M., Veeranarayanan S., Maekawa T., D S.K. External stimulus responsive inorganic nanomaterials for cancer theranostics. Adv. Drug Deliv. Rev. 2019;138:18–40. doi: 10.1016/j.addr.2018.10.007. [DOI] [PubMed] [Google Scholar]
  • 208.Yang J., Zhai S., Qin H., Yan H., Xing D., Hu X. NIR-controlled morphology transformation and pulsatile drug delivery based on multifunctional phototheranostic nanoparticles for photoacoustic imaging-guided photothermal-chemotherapy. Biomaterials. 2018;176:1–12. doi: 10.1016/j.biomaterials.2018.05.033. [DOI] [PubMed] [Google Scholar]
  • 209.Pansare V., Hejazi S., Faenza W., Prud’homme R.K. Review of long-wavelength optical and NIR imaging materials: contrast agents, fluorophores, and multifunctional nano carriers. Chem. Mater. 2012;24:812–827. doi: 10.1021/cm2028367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Rai P., Mallidi S., Zheng X., Rahmanzadeh R., Mir Y., Elrington S., Khurshid A., Hasan T. Development and applications of photo-triggered theranostic agents. Adv. Drug Deliv. Rev. 2010;62:1094–1124. doi: 10.1016/j.addr.2010.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Dube T., Kompella U.B., Panda J.J. Near infrared triggered chemo-PTT-PDT effect mediated by glioma directed twin functional-chimeric peptide-decorated gold nanoroses. J. Photochem. Photobiol B. 2022;228 doi: 10.1016/j.jphotobiol.2022.112407. [DOI] [PubMed] [Google Scholar]
  • 212.Seo B., Lim K., Kim S.S., Oh K.T., Lee E.S., Choi H.G., Shin B.S., Youn Y.S. Small gold nanorods-loaded hybrid albumin nanoparticles with high photothermal efficacy for tumor ablation. Colloids Surf. B Biointerfaces. 2019;179:340–351. doi: 10.1016/j.colsurfb.2019.03.068. [DOI] [PubMed] [Google Scholar]
  • 213.Yu S., Jang D., Yuan H., Huang W.T., Kim M., Marques Mota F., Liu R.S., Lee H., Kim S., Kim D.H. Plasmon-triggered upconversion emissions and hot carrier injection for combinatorial photothermal and photodynamic cancer therapy. ACS Appl. Mater. Interfaces. 2021;13:58422–58433. doi: 10.1021/acsami.1c21949. [DOI] [PubMed] [Google Scholar]
  • 214.Zhao J., Zhong D., Zhou S. NIR-I-to-NIR-II fluorescent nanomaterials for biomedical imaging and cancer therapy. J. Mater. Chem. B. 2018;6:349–365. doi: 10.1039/c7tb02573d. [DOI] [PubMed] [Google Scholar]
  • 215.Caro C., Gámez F., Quaresma P., Páez-Muñoz J.M., Domínguez A., Pearson J.R., Pernía Leal M., Beltrán A.M., Fernandez-Afonso Y., De la Fuente J.M., Franco R., Pereira E., García-Martín M.L. Fe3O4-Au core-shell nanoparticles as a multimodal platform for in vivo imaging and focused photothermal therapy. Pharmaceutics. 2021;13:416. doi: 10.3390/pharmaceutics13030416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Alphandéry E., Idbaih A., Adam C., Delattre J.Y., Schmitt C., Guyot F., Chebbi I. Chains of magnetosomes with controlled endotoxin release and partial tumor occupation induce full destruction of intracranial U87-Luc glioma in mice under the application of an alternating magnetic field. J. Contr. Release. 2017;262:259–272. doi: 10.1016/j.jconrel.2017.07.020. [DOI] [PubMed] [Google Scholar]
  • 217.Gowd V., Ahmad A., Tarique M., Suhail M., Zughaibi T.A., Tabrez S., Khan R. Advancement of cancer immunotherapy using nanoparticles-based nanomedicine. Semin. Cancer Biol. 2022;86:624–644. doi: 10.1016/j.semcancer.2022.03.026. [DOI] [PubMed] [Google Scholar]
  • 218.Manke A., Wang L., Rojanasakul Y., Rojanasakul Mechanisms of nanoparticle-induced oxidative stress and toxicity. Bio Med Res. Int. 2013;2013 doi: 10.1155/2013/942916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Blank F., Gerber P., Rothen-Rutishauser B., Sakulkhu U., Salaklang J., De Peyer K., Gehr P., Nicod L.P., Hofmann H., Geiser T., Petri-Fink A., Von Garnier C. Biomedical nanoparticles modulate specific CD4+ T cell stimulation by inhibition of antigen processing in dendritic cells. Nanotoxicology. 2011;5:606–621. doi: 10.3109/17435390.2010.541293. [DOI] [PubMed] [Google Scholar]
  • 220.Ngobili T.A., Daniele M.A. Nanoparticles and direct immunosuppression. Exp. Biol. Med. 2016;241:1064–1073. doi: 10.1177/1535370216650053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Zhou Z., Liu Y., Song W., Jiang X., Deng Z., Xiong W., Shen J. Metabolic reprogramming mediated PD-L1 depression and hypoxia reversion to reactivate tumor therapy. J. Contr. Release. 2022;352:793–812. doi: 10.1016/j.jconrel.2022.11.004. [DOI] [PubMed] [Google Scholar]
  • 222.Han J.H., Lee Y.Y., Shin H.E., Han J., Kang J.M., James Wang C.P., Park J.H., Kim S.N., Yoon J.H., Kwon H.K., Park D.H., Park T.E., Choy Y.B., Kim D.H., Kim T.H., Min J., Kim I.H., Park C.G., Han D.K., Park W. Image-guided in situ cancer vaccination with combination of multi-functional nano-adjuvant and an irreversible electroporation technique. Biomaterials. 2022;289 doi: 10.1016/j.biomaterials.2022.121762. [DOI] [PubMed] [Google Scholar]
  • 223.Krysko D.V., Garg A.D., Kaczmarek A., Krysko O., Agostinis P., Vandenabeele P. Immunogenic cell death and DAMPs in cancer therapy. Nat. Rev. Cancer. 2012;12:860–875. doi: 10.1038/nrc3380. [DOI] [PubMed] [Google Scholar]
  • 224.Kono H., Rock K.L. How dying cells alert the immune system to danger. Nat. Rev. Immunol. 2008;8:279–289. doi: 10.1038/nri2215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Shen Q., Yang J., Liu R., Liu L., Zhang J., Shen S., Zhang X. Hybrid ‘clusterbombs’ as multifunctional nanoplatforms potentiate brain tumor immunotherapy. Mater. Horiz. 2019;6:810–816. [Google Scholar]
  • 226.Mochalin V.N., Shenderova O., Ho D., Gogotsi Y. The properties and applications of nanodiamonds. Nat. Nanotechnol. 2011;7:11–23. doi: 10.1038/nnano.2011.209. [DOI] [PubMed] [Google Scholar]
  • 227.Li T.F., Li K., Zhang Q., Wang C., Yue Y., Chen Z., Yuan S.J., Liu X., Wen Y., Han M., Komatsu N., Xu Y.H., Zhao L., Chen X. Dendritic cell-mediated delivery of doxorubicin-polyglycerol-nanodiamond composites elicits enhanced anti-cancer immune response in glioblastoma. Biomaterials. 2018;181:35–52. doi: 10.1016/j.biomaterials.2018.07.035. [DOI] [PubMed] [Google Scholar]
  • 228.Irvine D.J., Hanson M.C., Rakhra K., Tokatlian T. Synthetic nanoparticles for vaccines and immunotherapy. Chem. Rev. 2015;115:11109–11146. doi: 10.1021/acs.chemrev.5b00109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Ng K.K., Lovell J.F., Zheng G. Lipoprotein-inspired nanoparticles for cancer theranostics. Acc. Chem. Res. 2011;44:1105–1113. doi: 10.1021/ar200017e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Zhang Z., Cao W., Jin H., Lovell J.F., Yang M., Ding L., Chen J., Corbin I., Luo Q., Zheng G. Biomimetic nanocarrier for direct cytosolic drug delivery. Angew Chem. Int. Ed. Engl. 2009;48:9171–9175. doi: 10.1002/anie.200903112. [DOI] [PubMed] [Google Scholar]
  • 231.Duan X., Chan C., Lin W. Nanoparticle-mediated immunogenic cell death enables and potentiates cancer immunotherapy. Angew Chem. Int. Ed. Engl. 2019;58:670–680. doi: 10.1002/anie.201804882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Alphandéry E., Idbaih A., Adam C., Delattre J.Y., Schmitt C., Guyot F., Chebbi I. Chains of magnetosomes with controlled endotoxin release and partial tumor occupation induce full destruction of intracranial U87-Luc glioma in mice under the application of an alternating magnetic field. J. Contr. Release. 2017;262:259–272. doi: 10.1016/j.jconrel.2017.07.020. [DOI] [PubMed] [Google Scholar]
  • 233.Mahmoudi K., Hadjipanayis C.G. The application of magnetic nanoparticles for the treatment of brain tumors. Front. Chem. 2014;2:109. doi: 10.3389/fchem.2014.00109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Maier-Hauff K., Ulrich F., Nestler D., Niehoff H., Wust P., Thiesen B., Orawa H., Budach V., Jordan A. Efficacy and safety of intratumoral thermotherapy using magnetic iron-oxide nanoparticles combined with external beam radiotherapy on patients with recurrent glioblastoma multiforme. J. Neuro Oncol. 2011;103:317–324. doi: 10.1007/s11060-010-0389-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.P. O'Brien, A. Panneerselvam, M. Green, R.A. Pattrick, V.S. Coker, C.I. Pearce, N.D. Telling, V.D.L. Gerrit, J.R. Lloyd, S. Haigh, Nanoscience: Volume 1: Nanostructures through Chemistry, J. Comput. Theor. Nanosci. 10 (2013) 2478–2483.
  • 236.Hanini A., Lartigue L., Gavard J., Schmitt A., Kacem K., Wilhelm C., Gazeau F., Chau F., Ammar S. Thermosensitivity profile of malignant glioma U87-MG cells and human endothelial cells following gamma-Fe2O3 NPs internalization and magnetic field application. RSC Adv. 2016;19:15415–15423. [Google Scholar]
  • 237.Hanini A., Lartigue L., Gavard J., Kacem K., Wilhelm C., Gazeau F., Chau F., Ammar S. Zinc substituted ferrite nanoparticles with Zn0.9Fe2.1O4 formula used as heating agents for in vitro hyperthermia assay on glioma cells. J. Magn. Magn. Mater. 2016;416:315–320. [Google Scholar]
  • 238.Derfus A.M., von Maltzahn G., Harris T.J., Duza T., Vecchio K.S., Ruoslahti E., Bhatia S.N. Remotely triggered release from magnetic nanoparticles. Adv. Mater. 2010;19:3932–3936. [Google Scholar]
  • 239.Pandey A., Singh K., Patel S., Singh R., Patel K., Sawant K. Hyaluronic acid tethered pH-responsive alloy-drug nanoconjugates for multimodal therapy of glioblastoma: an intranasal route approach. Mater Sci Eng C Mater Biol Appl. 2019;98:419–436. doi: 10.1016/j.msec.2018.12.139. [DOI] [PubMed] [Google Scholar]
  • 240.R. Le, Fèvre, M. Durand-Dubief, I. Chebbi, C. Mandawala, F. Lagroix, J.P. Valet, A. Idbaih, C. Adam, J.Y. Delattre, C. Schmitt, C. Maake, F. Guyot, E. Alphandéry, Enhanced antitumor efficacy of biocompatible magnetosomes for the magnetic hyperthermia treatment of glioblastoma, Theranostics. 7 (2017) 4618-4631. [DOI] [PMC free article] [PubMed]
  • 241.Hu Y., Zhou Y., Zhao N., Liu F., Xu F.J. Multifunctional pDNA-conjugated polycationic Au nanorod-coated Fe3O4 hierarchical nanocomposites for trimodal imaging and combined photothermal/gene therapy. Small. 2016;12:2459–2468. doi: 10.1002/smll.201600271. [DOI] [PubMed] [Google Scholar]
  • 242.Qian M., Du Y., Wang S., Li C., Jiang H., Shi W., Chen J., Wang Y., Wagner E., Huang R. Highly crystalline multicolor carbon nanodots for dual-modal imaging-guided photothermal therapy of glioma. ACS Appl. Mater. Interfaces. 2018;10:4031–4040. doi: 10.1021/acsami.7b19716. [DOI] [PubMed] [Google Scholar]
  • 243.Misra S.K., Naz S., Kondaiah P., Bhattacharya S. A cationic cholesterol based nanocarrier for the delivery of p53-EGFP-C3 plasmid to cancer cells. Biomaterials. 2014;35:1334–1346. doi: 10.1016/j.biomaterials.2013.10.062. [DOI] [PubMed] [Google Scholar]
  • 244.Mitha A.T., Rekha M.R. Multifunctional polymeric nanoplexes for anticancer co-delivery of p53 and mitoxantrone. J. Mater. Chem. B. 2014;2:8005–8016. doi: 10.1039/c4tb01298d. [DOI] [PubMed] [Google Scholar]
  • 245.Yadav P., Zhang C., Whittaker A.K., Kailasam K., Shanavas A. Magnetic and photocatalytic curcumin bound carbon nitride nanohybrids for enhanced glioma cell death. ACS Biomater. Sci. Eng. 2019;5:6590–6601. doi: 10.1021/acsbiomaterials.9b01224. [DOI] [PubMed] [Google Scholar]
  • 246.Shen C., Wang X., Zheng Z., Gao C., Chen X., Zhao S., Dai Z. Doxorubicin and indocyanine green loaded superparamagnetic iron oxide nanoparticles with PEGylated phospholipid coating for magnetic resonance with fluorescence imaging and chemotherapy of glioma. Int. J. Nanomed. 2019;14:101–117. doi: 10.2147/IJN.S173954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Alle M., B R G., Kim T.H., Park S.H., Lee S.H., Kim J.C. Doxorubicin-carboxymethyl xanthan gum capped gold nanoparticles: microwave synthesis, characterization, and anti-cancer activity. Carbohydr. Polym. 2020;229 doi: 10.1016/j.carbpol.2019.115511. [DOI] [PubMed] [Google Scholar]
  • 248.Ruan S., Hu C., Tang X., Cun X., Xiao W., Shi K., He Q., Gao H. Increased gold nanoparticle retention in brain tumors by in situ enzyme-induced aggregation. ACS Nano. 2016;10:10086–10098. doi: 10.1021/acsnano.6b05070. [DOI] [PubMed] [Google Scholar]
  • 249.Sun X., Huang X., Yan X., Wang Y., Guo J., Jacobson O., Liu D., Szajek L.P., Zhu W., Niu G., Kiesewetter D.O., Sun S., Chen X. Chelator-free 64Cu-integrated gold nanomaterials for positron emission tomography imaging guided photothermal cancer therapy. ACS Nano. 2014;8:8438–8446. doi: 10.1021/nn502950t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Battogtokh G., Gotov O., Kang J.H., Hong E.J., Shim M.S., Shin D., Ko Y.T. Glycol chitosan-coated near-infrared photosensitizer-encapsulated gold nanocages for glioblastoma phototherapy. Nanomedicine. 2019;18:315–325. doi: 10.1016/j.nano.2018.10.007. [DOI] [PubMed] [Google Scholar]
  • 251.Jensen S.A., Day E.S., Ko C.H., Hurley L.A., Luciano J.P., Kouri F.M., Merkel T.J., Luthi A.J., Patel P.C., Cutler J.I., Daniel W.L., Scott A.W., Rotz M.W., Meade T.J., Giljohann D.A., Mirkin C.A., Stegh A.H. Spherical Nucleic Acid Nanoparticle Conjugates as an RNAi-Based Therapy for Glioblastoma. Sci Transl Med. 2013;5:209ra152. doi: 10.1126/scitranslmed.3006839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Kouri F.M., Hurley L.A., Daniel W.L., Day E.S., Hua Y., Hao L., Peng C.Y., Merkel T.J., Queisser M.A., Ritner C., Zhang H., James C.D., Sznajder J.I., Chin L., Giljohann D.A., Kessler J.A., Peter M.E., Mirkin C.A., Stegh A.H. miR-182 integrates apoptosis, growth, and differentiation programs in glioblastoma. Genes Dev. 2015;29:732–745. doi: 10.1101/gad.257394.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Wang F., Zhang W., Shen Y., Huang Q., Zhou D., Guo S. Efficient RNA delivery by integrin-targeted glutathione responsive polyethyleneimine capped gold nanorods. Acta Biomater. 2015;23:136–146. doi: 10.1016/j.actbio.2015.05.028. [DOI] [PubMed] [Google Scholar]
  • 254.Kong L., Qiu J., Sun W., Yang J., Shen M., Wang L., Shi X. Multifunctional PEI-entrapped gold nanoparticles enable efficient delivery of therapeutic siRNA into glioblastoma cells. Biomater. Sci. 2017;5:258–266. doi: 10.1039/c6bm00708b. [DOI] [PubMed] [Google Scholar]
  • 255.Ruan S., Xie R., Qin L., Yu M., Xiao W., Hu C., Yu W., Qian Z., Ouyang L., He Q., Gao H. Aggregable nanoparticles-enabled chemotherapy and autophagy inhibition combined with anti-PD-L1 antibody for improved glioma treatment. Nano Lett. 2019;19:8318–8332. doi: 10.1021/acs.nanolett.9b03968. [DOI] [PubMed] [Google Scholar]
  • 256.Zhong Y., Wang C., Cheng R., Cheng L., Meng F., Liu Z., Zhong Z. cRGD-directed, NIR-responsive and robust AuNR/PEG–PCL hybrid nanoparticles for targeted chemotherapy of glioblastoma in vivo. J. Contr. Release. 2014;195:63–71. doi: 10.1016/j.jconrel.2014.07.054. [DOI] [PubMed] [Google Scholar]
  • 257.Kefayat A., Ghahremani F., Motaghi H., Amouheidari A. Ultra-small but ultra-effective: folic acid-targeted gold nanoclusters for enhancement of intracranial glioma tumors’ radiation therapy efficacy. Nanomedicine. 2019;16:173–184. doi: 10.1016/j.nano.2018.12.007. [DOI] [PubMed] [Google Scholar]
  • 258.Kunoh T., Shimura T., Kasai T., Matsumoto S., Mahmud H., Khayrani A.C., Seno M., Kunoh H., Takada J. Use of DNA-generated gold nanoparticles to radiosensitize and eradicate radioresistant glioma stem cells. Nanotechnology. 2019;30 doi: 10.1088/1361-6528/aaedd5. [DOI] [PubMed] [Google Scholar]
  • 259.Zhang Y., Hong G., Zhang Y., Chen G., Li F., Dai H., Wang Q. Ag2S quantum dot: a bright and biocompatible fluorescent nanoprobe in the second near-infrared window. ACS Nano. 2012;6:3695–3702. doi: 10.1021/nn301218z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Liu P., Huang Z., Chen Z., Xu R., Wu H., Zang F., Wang C., Gu N. Silver nanoparticles: a novel radiation sensitizer for glioma? Nanoscale. 2013;5:11829–11836. doi: 10.1039/c3nr01351k. [DOI] [PubMed] [Google Scholar]
  • 261.Tamborini M., Locatelli E., Rasile M., Monaco I., Rodighiero S., Corradini I., Franchini M.C., Passoni L., Matteoli M. A combined approach employing chlorotoxin-nanovectors and low dose radiation to reach infiltrating tumor niches in glioblastoma. ACS Nano. 2016;10:2509–2520. doi: 10.1021/acsnano.5b07375. [DOI] [PubMed] [Google Scholar]
  • 262.Bai Q., Zhao Z., Sui H., Chen J., Xie X., Wen F. The preparation and application of dendrimer modified CdTe/CdS near infrared quantum dots for brain cancer cells imaging. Appl. Sci. 2015;5:1076–1085. [Google Scholar]
  • 263.Radchanka A., Iodchik A., Terpinskaya T., Balashevich T., Yanchanka T., Palukoshka A., Sizova S., Oleinikov V., Feofanov A., Artemyev M. Emitters with different dimensionality: 2D cadmium chalcogenide nanoplatelets and 0D quantum dots in non-specific cell labeling and two-photon imaging. Nanotechnology. 2020;31 doi: 10.1088/1361-6528/aba5b5. [DOI] [PubMed] [Google Scholar]
  • 264.Ni D., Zhang J., Bu W., Xing H., Han F., Xiao Q., Yao Z., Chen F., He Q., Liu J., Zhang S., Fan W., Zhou L., Peng W., Shi J. Dual-targeting upconversion nanoprobes across the blood-brain barrier for magnetic resonance/fluorescence imaging of intracranial glioblastoma. ACS Nano. 2014;8:1231–1242. doi: 10.1021/nn406197c. [DOI] [PubMed] [Google Scholar]
  • 265.Dilnawaz F., Singh A., Mewar S., Sharma U., Jagannathan N.R., Sahoo S.K. The transport of non-surfactant based paclitaxel loaded magnetic nanoparticles across the blood brain barrier in a rat model. Biomaterials. 2012;33:2936–2951. doi: 10.1016/j.biomaterials.2011.12.046. [DOI] [PubMed] [Google Scholar]
  • 266.Chertok B., David A.E., Yang V.C. Polyethyleneimine-modified iron oxide nanoparticles for brain tumor drug delivery using magnetic targeting and intra-carotid administration. Biomaterials. 2010;31:6317–6324. doi: 10.1016/j.biomaterials.2010.04.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Shevtsov M., Nikolaev B., Marchenko Y., Yakovleva L., Skvortsov N., Mazur A., Tolstoy P., Ryzhov V., Multhoff G. Targeting experimental orthotopic glioblastoma with chitosan-based superparamagnetic iron oxide nanoparticles (CS-DX-SPIONs) Int. J. Nanomed. 2018;13:1471–1482. doi: 10.2147/IJN.S152461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Qiao Y., Gumin J., MacLellan C.J., Gao F., Bouchard R., Lang F.F., Stafford R.J., Melancon M.P. Magnetic resonance and photoacoustic imaging of brain tumor mediated by mesenchymal stem cell labeled with multifunctional nanoparticle introduced via carotid artery injection. Nanotechnology. 2018;29 doi: 10.1088/1361-6528/aaaf16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Alphandéry E., Idbaih A., Adam C., Delattre J.Y., Schmitt C., Guyot F., Chebbi I. Development of non-pyrogenic magnetosome minerals coated with poly-l-lysine leading to full disappearance of intracranial U87-Luc glioblastoma in 100% of treated mice using magnetic hyperthermia. Biomaterials. 2017;141:210–222. doi: 10.1016/j.biomaterials.2017.06.026. [DOI] [PubMed] [Google Scholar]
  • 270.Shah B.P., Pasquale N., De G., Tan T., Ma J., Lee K.B. Core-shell nanoparticle-based peptide therapeutics and combined hyperthermia for enhanced cancer cell apoptosis. ACS Nano. 2014;8:9379–9387. doi: 10.1021/nn503431x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Li Z.Y., Liu Y., Wang X.Q., Liu L.H., Hu J.J., Luo G.F., Chen W.H., Rong L., Zhang X.Z. One-pot construction of functional mesoporous silica nanoparticles for the tumor-acidity-activated synergistic chemotherapy of glioblastoma. ACS Appl. Mater. Interfaces. 2013;5:7995–8001. doi: 10.1021/am402082d. [DOI] [PubMed] [Google Scholar]
  • 272.He X., Nie H., Wang K., Tan W., Wu X., Zhang P. In vivo study of biodistribution and urinary excretion of surface-modified silica nanoparticles. Anal. Chem. 2008;80:9597–9603. doi: 10.1021/ac801882g. [DOI] [PubMed] [Google Scholar]
  • 273.Zhang H., Zhang W., Zhou Y., Jiang Y., Li S. Dual functional mesoporous silicon nanoparticles enhance the radiosensitivity of VPA in glioblastoma. Transl. Oncol. 2017;10:229–240. doi: 10.1016/j.tranon.2016.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Wang Y., Wang K., Zhao J., Liu X., Bu J., Yan X., Huang R. Multifunctional mesoporous silica-coated graphene nanosheet used for chemo-photothermal synergistic targeted therapy of glioma. J. Am. Chem. Soc. 2013;135:4799–4804. doi: 10.1021/ja312221g. [DOI] [PubMed] [Google Scholar]
  • 275.Du Y., Qian M., Li C., Jiang H., Yang Y., Huang R. Facile marriage of Gd3+ to polymer-coated carbon nanodots with enhanced biocompatibility for targeted MR/fluorescence imaging of glioma. Int. J. Pharm. 2018;552:84–90. doi: 10.1016/j.ijpharm.2018.09.010. [DOI] [PubMed] [Google Scholar]
  • 276.Santos T., Fang X., Chen M.T., Wang W., Ferreira R., Jhaveri N., Gundersen M., Zhou C., Pagnini P., Hofman F.M., Chen T.C. Sequential administration of carbon nanotubes and near-infrared radiation for the treatment of gliomas. Front. Oncol. 2014;4:180. doi: 10.3389/fonc.2014.00180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Eldridge B.N., Bernish B.W., Fahrenholtz C.D., Singh R. Photothermal therapy of glioblastoma multiforme using multiwalled carbon nanotubes optimized for diffusion in extracellular space. ACS Biomater. Sci. Eng. 2016;2:963–976. doi: 10.1021/acsbiomaterials.6b00052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Xie S., Fei X., Wang J., Zhu Y.C., Liu J., Du X., Liu X., Dong L., Zhu Y., Pan J., Dong B., Sha J., Luo Y., Sun W., Xue W. Engineering the MoS2/MXene heterostructure for precise and noninvasive diagnosis of prostate cancer with clinical specimens. Adv. Sci. 2023;10 doi: 10.1002/advs.202206494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Liu Y., Dai R., Wei Q., Li W., Zhu G., Chi H., Guo Z., Wang L., Cui C., Xu J., Ma K. Dual-functionalized janus mesoporous silica nanoparticles with active targeting and charge reversal for synergistic tumor-targeting therapy. ACS Appl. Mater. Interfaces. 2019;11:44582–44592. doi: 10.1021/acsami.9b15434. [DOI] [PubMed] [Google Scholar]
  • 280.Jin X., Yu J., Yin M., Sinha A., Jin G. Combined ultrasound treatment with transferrin-coupled nanoparticles improves active targeting of 4T1 mammary carcinoma cells. Technol. Cancer Res. Treat. 2021;20 doi: 10.1177/15330338211062325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Hamaly M.A., Abulateefeh S.R., Al-Qaoud K.M., Alkilany A.M. Freeze-drying of monoclonal antibody-conjugated gold nanorods: colloidal stability and biological activity. Int. J. Pharm. 2018;550:269–277. doi: 10.1016/j.ijpharm.2018.08.045. [DOI] [PubMed] [Google Scholar]
  • 282.Verry C., Dufort S., Villa J., Gavard M., Iriart C., Grand S., Charles J., Chovelon B., Cracowski J.L., Quesada J.L., Mendoza C., Sancey L., Lehmann A., Jover F., Giraud J.Y., Lux F., Crémillieux Y., McMahon S., Pauwels P.J., Cagney D., Berbeco R., Aizer A., Deutsch E., Loeffler M., Le Duc G., Tillement O., Balosso J. Theranostic AGuIX nanoparticles as radiosensitizer: a phase I, dose-escalation study in patients with multiple brain metastases (NANO-RAD trial) Radiother. Oncol. 2021;160:159–165. doi: 10.1016/j.radonc.2021.04.021. [DOI] [PubMed] [Google Scholar]
  • 283.Zanoni D.K., Stambuk H.E., Madajewski B., Montero P.H., Matsuura D., Busam K.J., Ma K., Turker M.Z., Sequeira S., Gonen M., Zanzonico P., Wiesner U., Bradbury M.S., Patel S.G. Use of ultrasmall core-shell fluorescent silica nanoparticles for image-guided sentinel lymph node biopsy in head and neck melanoma: a nonrandomized clinical trial. JAMA Netw Open. 2021;4 doi: 10.1001/jamanetworkopen.2021.1936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Reitsma S., Slaaf D.W., Vink H., van Zandvoort M.A., oude Egbrink M.G. The endothelial glycocalyx: composition, functions, and visualization. Pflugers Arch. 2007;454:345–359. doi: 10.1007/s00424-007-0212-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 285.Zhao X., Wei Z., Zhao Z., Miao Y., Qiu Y., Yang W., Jia X., Liu Z., Hou H. Design and development of graphene oxide nanoparticle/chitosan hybrids showing pH-sensitive surface charge-reversible ability for efficient intracellular doxorubicin delivery. ACS Appl. Mater. Interfaces. 2018;10:6608–6617. doi: 10.1021/acsami.7b16910. [DOI] [PubMed] [Google Scholar]
  • 286.Mailänder V., Landfester K. Interaction of nanoparticles with cells. Biomacromolecules. 2009;10:2379–2400. doi: 10.1021/bm900266r. [DOI] [PubMed] [Google Scholar]
  • 287.Li X., Hetjens L., Wolter N., Li H., Shi X., Pich A. Charge-reversible and biodegradable chitosan-based microgels for lysozyme-triggered release of vancomycin. J. Adv. Res. 2023;43:87–96. doi: 10.1016/j.jare.2022.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Moyano D.F., Liu Y., Ayaz F., Hou S., Puangploy P., Duncan B., Osborne B.A., Rotello V.M. Immunomodulatory effects of coated gold nanoparticles in LPS-stimulated in vitro and in vivo murine model systems. Chem. 2016;1:320–327. doi: 10.1016/j.chempr.2016.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data will be made available on request.


Articles from Materials Today Bio are provided here courtesy of Elsevier

RESOURCES