Next Article in Journal
Non-Woven Filters Made of PLA via Solution Blowing Process for Effective Aerosol Nanoparticles Filtration
Previous Article in Journal
Theoretical Analysis of the Movement Law of Top Coal and Overburden in a Fully Mechanized Top-Coal Caving Face with a Large Mining Height
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Synthetic Route of Silica Nanoparticles to Theranostic Applications

1
Medical Bionanotechnology, Faculty of Allied Health Sciences, Chettinad Hospital & Research Institute (CHRI), Chettinad Academy of Research and Education (CARE), Kelambakkam, Chennai 603103, India
2
Department of Pharmaceutical Sciences, College of Pharmacy, AlMaarefa University, Ad Diriyah 13713, Saudi Arabia
3
Department of Pharmacy Practice, College of Pharmacy, AlMaarefa University, Ad Diriyah 13713, Saudi Arabia
4
Department of Pharmaceutics, College of Pharmacy, King Saud University, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Processes 2022, 10(12), 2595; https://doi.org/10.3390/pr10122595
Submission received: 1 April 2022 / Revised: 17 November 2022 / Accepted: 28 November 2022 / Published: 5 December 2022

Abstract

:
The advancements in nanotechnology have quickly developed a new subject with vast applications of nanostructured materials in medicine and pharmaceuticals. The enormous surface-to-volume ratio, ease of surface modification, outstanding biocompatibility, and, in the case of mesoporous nanoparticles, the tunable pore size make the silica nanoparticles (SNPs) a promising candidate for nano-based medical applications. The preparation of SNPs and their contemporary usage as drug carriers, contrast agents for imaging, carrier of photosensitizers (PS) in photodynamic, as well as photothermal treatments are intensely discussed in this review. Furthermore, the potential harmful responses of silica nanoparticles are reviewed using data obtained from in vitro and in vivo experiments conducted by several studies. Moreover, we showcase the engineering of SNPs for the theranostic applications that can address several intrinsic limitations of conventional therapeutics and diagnostics. In the end, a personal perspective was outlined to state SNPs’ current status and future directions, focusing on SNPs’ significant potentiality and opportunities.

1. Introduction

Nanomedicine is defined as the use of nanotechnology in medical science, which provides nanoscale materials for the diagnosis, treatment, and screening of diseases [1,2,3]. The use of nanosized drug moieties can achieve targeted delivery with increased bioavailability and therapeutic efficacy [4,5]. Various nanomedicine-based drug delivery systems have been utilized in the development and targeted drug delivery systems of different biomolecules [6,7,8,9,10,11]. It is clear that biomedical applications of a large number of nanomaterials, such as silica nanoparticles (SNPs), graphene, gold nanoparticles, superparamagnetic iron oxide nanoparticles (SPIONs), and other metal and metal oxide nanoparticles, are receiving more attention [12,13,14,15,16]. Among all other nanomaterials, the use of silica nanoparticles in medicine has been increased due to their excellent properties [12]. Silica nanoparticles are usually spherical amorphous materials, and they can be fabricated in a wide range of sizes from 1 to 100 nm [17,18]. Their straightforward synthesis techniques distinguish SNPs for reliable delivery mechanisms and the chemistry of their surfaces can be adjusted to suit required purposes.
In the past few decades, silica nanoparticles have become a key component of biological imaging and the delivery of drugs and genetic material, due to their chemical and physical stability, well-defined hydrophilic surfaces, and their ability to protect drugs from active immune responses [19,20]. SNPs are available in many forms, including core-shell silica nanoparticles, nonporous SNPs, hollow mesoporous silica nanoparticles (HMSN), and mesoporous silica nanoparticles (MSN) [21]. There are several physicochemical parameters associated with SNPs, such as size, shape, porosity, and surface area, that play a crucial role in the delivery of the payloads to the target site and their subsequent elimination from the body [22,23]. The size of the silica nanoparticles is one of the most important factors that determine their uptake by cells, as well as their biocompatibility, and it is controlled by varying the reaction parameters [24,25]. The shape of silica nanoparticles has also been found to be a key player in modulating their physiological behavior and activity [26,27]. A wide variety of therapeutic and diagnostic applications use nanospheres and nanorods with different aspect ratios as SNPs [28]. Nanoparticle uptake and biodistribution are primarily influenced by size and shape, but payload delivery was more affected by porosity.
Silica nanoparticles have received more interest in medicinal applications than silicon, popularly known as silicon quantum dots, because of their wide range of commercial approaches in pharmaceuticals, pigments, humidity sensors, and medical insulators [29,30]. The structure of mesoporous silica nanoparticles (MSNs) is highly porous and can load many drugs and other desired agents. Fluorescence molecules loaded on the MSNs are kept safe from severe external environments, which improve the stability of the fluorophore [31]. Light focusing is of higher resolution, which can overcome the limitations of biomedical imaging and other monitoring applications. The biggest safety concerns regarding SNPs are their toxic behaviors leading to immunotoxicity, genotoxicity, and cytotoxicity, which affect their size, hemolysis, aggregation, and bioaccumulation [32].
This review discusses the preparation of SNPs and their applications as drug carriers, contrast agents for imaging, and carriers of photosensitizers (PS) in photodynamic as well as photothermal treatments. In addition, the potential harmful responses of SNPs and the present scenario in medicines in terms of clinical trials are reviewed using data obtained from in vitro and in vivo experiments conducted by several studies.

2. Different Methods to Prepare SNPs

2.1. Microemulsion Method

The microemulsion is a synthetic “nonreactor” made of two incompatible liquids: isotropic, surfactant stable, and thermodynamically stable [33]. The standard composition consists of organic solvent such as cyclohexane, surfactants such as alkylphenol ethoxylate, water, and silica alkoxides such as tetraethylorthosilicates or tetramethylorthosilicates. In aqueous phase, reactants often produce a microemulsion, which is subsequently combined with another reactant to generate nanoparticles (NPs). The NPs produced by this technology have a narrow particle shape. Varying particle sizes and shapes are achieved with the alteration in the properties of the microemulsion system (Figure 1). The microemulsion process may be used to synthesize SNPs with a size less than 100 nm. The approach could be improved if the organic compounds created in the synthesis procedure were removed, resulting in higher SNP purity [34].

2.2. Gas-Phase Method

The raw material used in the gas-phase process involves silicon tetrachloride, which is hydrolyzed with oxygen and hydrogen at a high temperature to produce pyrogenic silica. Heat is then applied to the droplets to form the required NPs with less density and a distinct highly porous structure. The gas-phase method provides advantages of high dispersion, increased purity, and low number of hydroxyl groups on the surface [35]. The associated drawbacks are typically the cost of precursors, energy consumed during the process, sophisticated technology, and special instruments. Because of their great strengthening qualities, the nanoparticles synthesized by the gas-phase technique are primarily employed to improve the rubber characteristics [12,36].

2.3. Stöber’s Method

In 1968, Stöber, Fink, and Bohn proposed the Stöber technique, one of the sol-gel procedures [37,38]. They are widely utilized in the production of silica nanoparticles (size ranging from 5 to 2000 nm) because they can generate monodispersed particles, even without any template. A mixture of tetraethyl orthosilicate (TEOS) with alcohol, ammonia, and water can produce a large percentage of silicic acids (Figure 2). Si–O–Si condensate can be generated either by dealcoholization or dehydration of silicic acid [39]. The main SNPs immediately combine to form primary silica particles once the concentration approaches supersaturation, and then develop into more stable silica particles. SNPs are formed when soluble silicic acid condensates continuously react and have a more substantial influence on the development on their surface. The Stöber’s technique can be used to make solid and mesoporous SNPs [40]. SNPs have a great degree of homogeneity and purity, as well as a clear crystal structure. On the other hand, the precursors necessary in the synthesis process are time-consuming, expensive, and pollute the environment. As a result, additional in-depth research with the development of the Stöber’s technique will necessitate more cost-effective raw materials and more efficient production.
Si OR + H 2 O   Hydrolysis   Si OH + ROH   ( R   =   alkyl   group )
Si OH + OH Si   Polycondensation   Si O Si + H 2 O
Si OH + RO Si   Polycondensation   Si O Si + ROH
Si ( OR ) 4 + 2   H 2 O   H +   or   OH   SiO 2 + 4 ROH

2.4. Precipitation Method

Refluxing silica is one precipitation technique that includes combining multiple chemicals in the solution and then mixing the precursor (sodium silicate, Na2SiO3) to the priorly prepared solution to produce the precipitate [41]. SNPs are made by drying or calcining the precipitate. To obtain excellent SNP dispersion in the silicate solution, it is necessary to ensure the solution contains monodispersed organic compounds. The use of specific solvents in the precipitation method results in excellent silica particles. Compared to the sol-gel approach, the precipitation method is a simple and efficient way to prepare silica nanoparticles but, sometimes, impurities and aggregation can quickly occur. Highly pure and amorphous silica was more acceptable in many applications. Methods for the extraction of silica from rice husk ash (RHA) were also developed [42].

2.5. Green Synthesis

Green and sustainable routes are used for the synthesis of nanoparticles with desired physicochemical properties [43]. SNP produced by green routes is useful in medicine, cosmetics, pharmaceuticals, and many other fields. In green synthesis, eco-friendly reagents replace harmful chemicals to use the synthesized nanoparticles for biomedical applications. Several natural precursors are used as silica sources to react with sustainable biological bases. Many clays, agricultural wastes, industrial wastes, and some common plants are prime sources of silica with different forms for synthesizing SNPs [44]. The aqueous extract of different parts of plants, fruit pulps, and microbial extracts are used as an alkaline catalyst for the green synthesis of SNPs. Table 1 shows the advantages and disadvantages of different synthetic routes of SNPs synthesis.

3. Mesoporous SNPs

Mesoporous SNPs (MSNs) are porous, spherical nanomaterials with unique features that include easy surface functionalization, mesoporosity, and high biocompatibility [45]. MSNs provide the potential ability to load a large number of drugs onto the nanocarrier due to their unique structural features. Nucleic acids, such as DNA and RNA, are frequently bound on the exterior surface of MSNs due to their tiny pore diameters. They exhibit greater drug encapsulation capacity and stimulus-responsive drug-releasing profile when it comes to drug delivery systems [46,47]. MSNs are very accessible to multifunctionalize, and their magnetic, fluorescent, and photothermal characteristics enable nanotheranostics to perform simultaneous bioimaging and drug administration. Moreover, the release of the drug of MSNs can be regulated through a nanovalve structure that responds to internal and external stimuli. The electrostatic interaction between MSNs and negatively charged RNA is facilitated by the positive surface charge and, also, allows MSNs to escape from endosomes through the proton sponge effect [48].

4. Core-Shell SNPs

A research group has developed a method for synthesizing core-shell SNPs that are now widely utilized. They used poly-L-lysine to cover polystyrene templates before performing a polycondensation with tetramethyl orthosilicate in hydrolyzed form [49]. The thickness of the hollow nanoshells is about 6–10 nm after calcination, and the diameter of the prepared nanoparticles was reduced by 10–20 percent. Metals might be used in the core for templating, resulting in core-shell SNPs. The magnetite–silica core-shell nanoparticles were synthesized by oxidation of silicon dust [50]. The above technique would be used to generate core-shell NPs for biological applications. Some researchers created a novel polymeric covering for gold–core SNPs. By mixing varying ratios of two polymers, β-cyclodextrin (β-CD) and poly-2-ethyl-2-oxazoline, biological characteristics of the silica-coated gold nanoparticles were enhanced, and novel cancer therapy concepts were generated [12].

5. Biomedical Applications

5.1. Drug Delivery System

Due to the advancement of nanotechnology, materials generated at the nanoscale level have received growing attention in disciplines such as drug delivery [3,51,52,53]. Because of its unique qualities, such as a vast surface-to-volume ratio, controlled particle size, pore volume, and high biocompatibility, porous silica nanoparticles (NPs) have been studied extensively among all known nanomaterials [54,55,56,57]. Mesoporous silica nanoparticles (MSNs) are good prospects for drug delivery and biological applications, with pore sizes ranging from 2 nm to 50 nm. MSNs are synthesized in the presence of a supramolecular assembled surfactant that functions as a targeting moiety [25]. MSN-based biological applications have shown to be highly promising (Figure 3). The advantages of MSNs include: (i) pore channels with a vast surface area and pore volume have a lot of potential for drug adsorption and loading; (ii) compared to other nanostructures, the mesoporous silica has a large drug-loading capacity and release kinetics because of its porous nature and tunable pore size; (iii) the therapeutic effectiveness of drugs is improved, and toxicity is reduced when delivered via a surface that can be easily adjusted for the regulation and targeted drug administration [58,59]. In vivo biosafety studies for cytotoxicity, biodegradation, biodistribution, and excretion have given positive results. (iv) Drug delivery and bioimaging can be performed simultaneously when magnetic and/or luminous substances are used; (v) bioactive materials with good surface characteristics and porosity have proven to be good options for bone regeneration [60]. The number of studies on mesoporous silica materials has risen considerably as a result of these distinct properties. Additional research on MSNs, which are covalently attached to dipalmitoyl molecules with the support of phosphorylated lipids, were effectively produced into a controlled release mechanism. Through a chemically induced disulfide reduction, the system may release fluorescein moieties from the porous structure of MSNs coated with the lipid bilayer (LB-MSNs) [61]. According to the system, LB-MSNs can be employed to create a controlled release of drug delivery system.

5.2. Photodynamic Therapy

PDT is an emerging noninvasive therapeutic strategy based on the activation of a photosensitizer (PS) with the help of light in a specific wavelength. In the process of illumination, the excited PS transfers its energy to the molecular oxygen, leading to the generation of reactive oxygen species (ROS). The dioxygen, otherwise called singlet oxygen produced, can efficiently oxidize the main cellular macromolecules that result in vascular closure and tumor cell death [52]. PDT does not pose harmful risks to the biological system, as in the case of ionizing radiation therapy that causes damage to the surrounding normal cells.
It is clear that, compared with traditional treatment modalities, PDT has its advantages due to its limited invasiveness and negligible cumulative toxicity. Therefore, PDT aims to boost the quality of life of cancer patients. PDT is highly successful in treating lung, skin, head, and neck cancers [62]. PDT has not yet been accepted as a first-line cancer treatment process due to the lack of perfect PS and difficulties experienced during the PS formulation. Nanostructured micelles, liposomes, polymers, proteins, or metals are currently available for pharmaceutical purposes in various forms. PS-loaded silica nanoparticles are shown to be potential singlet oxygen-producing platforms for increasing therapeutic-possessing photoactive activity and improves selectivity and solubility [63,64,65]. The photosensitizers can be either tagged or encapsulated onto silica nanoparticles in their inner or outer surfaces. As a result, loading the photosensitizer into the nanomaterial provides outstanding photostability while restraining oxygen species diffusion.
The surface functionalized silica nanoparticles exhibit a greater therapeutic profile in photodynamic therapy against cancer. The porphyrin dye used in PDT suffers from low solubility and selectivity. SNPs are widely used to overcome the issues faced by the PS. The anticancer effect of core-shell nanoparticles where the core was made up of silica and coated with xylan bearing 5-(4-hydroxyphenyl)-10,15,20-triphenylporphyrin (TPPOH) dye was investigated. The prepared hybrid particle was then characterized to assess the pharmacokinetic profile. The in vitro analysis showed better anticancer activity in colorectal cancer compared to the dye alone. The photodynamic activity of mesoporous silica nanoparticle coated with two different photosensitizers, namely, indocyanine Green and Chlorin e6, against prostate cancer was also reported. The formulations have enhanced cancer killing capacity compared to the nude dye. The dual combination of dyes exhibits better therapeutic efficiency when nanoformulated using SNPs.

5.3. Photothermal Therapy

Photothermal therapy (PTT) utilizes a photothermal agent (PA) to convert heat from light energy. It is a noninvasive therapeutic modality with a minimum side effect and is widely used for the treatment of drug-resistant tumors. Mesoporous silica nanoparticles are the most efficient drug delivery agent used for the codelivery of PA and chemotherapeutics in combination therapy [66]. Wang et al. developed NIR-triggered photothermal nanoparticles incorporating indocyanine green into amino-group-modified silica nanoparticles [67]. They have successfully demonstrated that the synthesized photothermal nanoparticles have the potential for enhanced cancer cell killings. Wang et al. formulated multifunctional Janus nanoparticles combining gold triangle and mesoporous silica nanoparticles as drug delivery vehicles to deliver hypoxia-active prodrugs for topical PTT [68]. The folate functionalization on the pegylated Janus structure helps in targeting cancer cells. Therefore, MSN is useful in cancer therapy and a widely used system for theranostics.

5.4. Imaging and Monitoring

Recent studies have explored the usage of SNPs in imaging modalities and found that contrast agents encapsulated in SNPs produce high-resolution images compared to conventional imaging techniques [69]. These investigations revealed that SNPs could deliver contrast agents with high specificity and sensitivity for targeting and better imaging contrast. In addition, SNPs have a controlled release in the body because they have a substantial delay. In biomedical imaging, tumor imaging is an essential factor. Other findings indicate that intravenous administrations of core-shell nanoparticles (silica core and metallic shells) result in an improvement in computed tomography (CT) and magnetic resonance imaging (MRI). By using MSNs as a multimodal strategy, researchers successfully detected tumor blood capillaries in live rat model organisms according to near-infrared (NIR) imaging and positron emission tomography (PET) and other techniques [70]. The results showed that this combination of devices improves optical and PET imaging resolutions. Furthermore, fluorescently tagged mesoporous silica nanoparticles were utilized as an endoscopic contrast medium to diagnose abnormal cells of the gut. The researchers developed a new type of Magnus nano-bullet (Mn-DTPA-F-MSNs), which consists of a head with magnetic property (Fe3O4-NPs) and mesoporous silica (SiO2) body [71]. Researchers redesigned the mesoporous SiO2 before loading it with Mn2+. Both in vitro and in vivo studies demonstrated that the Magnus nano-bullet accelerated the T1-weighted MR images. The magnus nano-bullets showed enhancement in the detection of GSH, a biomarker responsible for redox responsive T1 MRI. A macrophage-driven PET imaging tracker was engineered using aza-dibenzocyclooctyne-tethered PEGylated MSNs (DBCO-MSNs) combining short-lived 18F labeled radiotracer. The in vivo cancer cell tracking and imaging were performed by the PET imaging protocol with the movement of macrophage cells into the tumor site by the bio-orthogonal SPAAC reactions to form 18F-labeled aza-dibenzocycloocta-triazolic MSNs inside RAW 264.7 cells [12]. In response, the macrophage was more capable of moving towards the tumor effectively. Finally, the tissue radioactive dispersion measurements were consistent with the PET imaging data. A novel cup-shaped SNP, which was effective for ultrasonic imaging, was also discovered.
Using an emulsification slicing process, researchers developed exosomes-like silica nanoparticles (ELS), or cup-shaped drops were constructed [72]. Researchers discovered that ELS was more effective at enhancing particle echogenicity and cell tagging when compared to SNPs shaped like cup-shaped drops. The study demonstrated that stem cells can be used in ultrasound imaging in the future and that SNPs can improve the echogenicity of stem cells in vitro and in vivo [73]. Furthermore, they described a biodegradable ultrasound-capable material. These SNPs did not witness cytotoxic effects at the 250 µg/mL concentration necessary for tagging and then were removed from cells within three weeks [73]. SNP-based imaging techniques have both advantages and disadvantages. Simple optical imaging with a modest imaging resolution is limited, whereas high-quality CT imaging may contaminate the body with radiation. Researchers show that multimodal imaging is capable of enhancing individual imaging methods. Long-term imaging of sentinel lymph nodes draining tumor utilizing mesoporous nanosilica-based triple-modal probes was demonstrated in a study [74]. Due to the excellent stability and resilience of nanoprobes, imaging results from various modalities are coherent and comparable. Furthermore, biocompatibility is a crucial aspect of the use of SNPs in nanomedicine and imaging techniques. Nanomaterials tagged with multimodal imaging nanoprobes were successfully identified using optical imaging, as well as MRI, and also provide more outstanding biocompatibility [74].

5.5. Protein Recognition and Isolation

The versatile properties of silica set it apart from other materials. The ease of synthesis, high surface area, low cost, and surface functionalization are the major reasons for their biomedical applications. Other than drug delivery and diagnostic applications, they can be used to recognize and separate proteins from the sample. Researchers are keen on the application of silica, since the toxicity inside the biological system is negligible. In order to afford simple and quick separation of proteins, iron oxide nanoparticles (IONP) can be coated with silica [75]. The magnetic property of IONP is used to achieve the isolation process. The silica nanoparticle allows surface imprinting, which is an efficient strategy to obtain superior adsorption affinity as the initial step for separation. Typically, the nanostructures are fabricated employing iron oxide nanoparticles as a core, which is then laminated with silica to accomplish quick biomolecule separation. The core responds to external magnetic fields, whereas the shell serves as a spot for protein entrapment, providing biocompatibility and stability [76]. Core-shell nanoparticles of this sort are implemented in bioseparation, biosensors, immunoassay analysis, nucleic acid detection, and other applications. Mesoporous silica has a higher porous architecture, which offers greater adsorption of the adsorbates. The surface charge of silica nanoparticles is a significant issue. Pure mesoporous silica is said to have a neutral charge. This neutral structure of silica may be susceptible to leaching risks. As a result, chemical modification with tagging of the organic group is used to achieve stability [77].

5.6. Nucleic Acid Detection and Purification

The nucleic acid detection technique is performed in wide areas, including clinical diagnosis, food technology, and environmental safety monitoring [78]. Since nucleic acid is a fundamental component for storing and transmitting genetic information, scientists are focusing their efforts on constructing a facile method for detecting and purifying it in preparation for future research. Three main factors that facilitate DNA adsorption are hydrogen bond formation, electrostatic repulsion, and dehydration. Currently, existing approaches have many implications [79]. The main concern with the conventional approach is that they require pipetting procedures, which apparently reduce sensitivity. To achieve the desired sensitivity, additional treatment, such as amplification, is required. However, the amplification stage frequently produces aerosol, which is detrimental to the environment. The technique is quite complicated and necessitates a high level of expertise. Researchers must exercise extra precautions when working with extremely contagious viruses. In order to upgrade the method, silica-coated magnetic nanoparticles-driven separation was established [80,81,82]. Several investigations suggest that integrating magnetic nanoparticles (MNPs) as a core with a silica shell could be very effective. Such modified core-shell-structured silica-coated MNPs, were used to extract nucleic acid and amplify them. Therefore, silica shells can be employed to perform simultaneous, automated, and precise nucleic acid sequence detection.

5.7. Gene Therapy

Nanotechnology-based gene therapy is now emerging as a possible method of delivering genes to treat a wide variety of genetic complications, including cancer. Nanoparticles are expected to be vital vehicles that could carry genetic molecules into cells. Highly porous nanostructures exhibit high loading volume for incorporation of target moiety. Among all the other nanoparticles, the most preferred carrier vehicle is mesoporous silica nanoparticles (MSNs), since they possess a large surface area and high pore volume [83]. Silica is considered a nonviral vector that could effectively deliver genetic molecules to the target site. To achieve targeted release, receptor-mediated endocytosis predominates to deliver the cargo or payloads to the diseased site. Different bioconjugation and chemical modification can further upregulate the function of silica inside the biological system. Polymers are generally utilized for surface modification to improve performance while avoiding cytotoxicity [84]. In vivo evaluation of chemical modification of silica with sodium chloride or sodium iodide has recently proven to be a promising method for gene transfer without causing any pathological alterations in the cells [85]. Since the DNA is negatively charged, the carrier molecule has to be prepared with a positive surface charge. However, the bonding strength should be weak between the carrier molecules and the targeting moiety as the gene involved in breakage and release is weak [86]. Strong adhesion might lead to release in undesired sites, or may undergo digestion with no release. Hence, the particle with payload has to be tailored to achieve satisfactory results. The control over the morphology of nanomaterial during fabrication serves the purpose [87].

5.8. Vaccine Delivery

Vaccines were at the frontline to maintain human wellbeing and have been a cost-effective solution to prevent serious infections. Concerns rising regarding conventional viral systems are their immunogenicity and toxicity, which are the primary reasons for gaining research interest for designing novel vaccination methods. The present focus of research seems to be on the development and implementation of effective carriers for the delivery of the genetic payload to the target tissue [88]. Three major things to be considered during the formulation of vaccines are effective adjuvant, appropriate antigen, and route of administration. The antigen is used to activate the immunological response, whereas the adjuvant is used to support the antigen to elicit the immune response. Vaccination can be performed either by introducing a protein antigen or by using RNA/DNA that encodes an antigenic protein. Developing a vehicle that can carry the genetic material to the site free of toxicity was a major hurdle for researchers. MSNs are being used in vaccinations that serve as both a carrier and an adjuvant (Table 2). MSNs as a carrier can transport RNA/DNA to the target tissue, promoting protein antigen synthesis, and acts as an adjuvant to induce both cell-mediated and humoral immune responses [89]. Several in vitro studies have shown that the vaccines can be delivered based on the acid–base triggered system. Since lymph nodes are the major site for vaccine targets, they need to undergo lymphatic drainage to be taken up by dendritic cells for further action. The surface modification with negative groups makes them a complete lymph-nodes-targeted vehicle. The current pandemic situation due to COVID-19 has evidenced the importance of vaccines. Oligonucleotides-based vaccines are the primary and only therapeutic option available for now. Researchers working on COVID-19 vaccinations are now concentrating on the development of vaccines with nanocarrier for RNA/DNA delivery that encodes specific spike protein [90]. The stability, biocompatibility, and release kinetic profile of MSNs makes them a potent tool for antigen carrier and adjuvant.

5.9. Other Applications

Besides the applications listed above, SNPs are also essential to prevent bacterial infections, speed up wound healing, and promote bone regeneration [100,101]. To enhance the benefits of antibacterial infections, nanocarrier-encapsulated antibiotics have been used to enhance the efficiency of antibacterial treatment with the ability to penetrate bacterial walls [12]. SNP characteristics, such as huge surface-to-volume ratio, pore size, and controlled loading, and drug release of antimicrobial agents are remarkable in this regard. When antibacterial peptides such as LL-37, non-steroidal anti-inflammatory medications, and antibiotics using a combination of levofloxacin-loaded MSNs and polycationic dendrimers are added to the mix, the antibiotics are effective at combating Gram-negative bacterial biofilm. The results of the study show Gram-negative E. coli and Gram-positive Staphylococcus aureus do not grow in the presence of SNPs loaded with captured bacteria [12]. Another study examined Gram-negative bacteria that were driven by targeted QS into virulence. The unit of study was to test the bacterial communication or quorum quenching (QQ) capacity of metals and metal oxide nanoparticles. Miller created bifunctional SNPs with the quenching molecule cyclodextrin [12,102]. These results suggest that using functionalized cyclodextrin SNPs, remain in the bacterial microenvironment and inhibit extracellular bacterial communication proteins from seeping out, resulting in antibiotic activity. The study investigated that SNPs were shown to influence migration and invasion of human skin fibroblasts in a local wound healing model (CCD-25SK). The silicic acid is generated by the fibroblasts when the silica nanoparticles become dissolved after absorption in the cells and stimulates wound healing [103,104]. According to the study, researchers found the biocomposite treatment shortened wound epithelialization time in Wistar rats compared with the control group. Furthermore, the composite has antibacterial properties, good drug availability, and a high absorption rate, which could make SNP biocomposite a potential wound healing material in the future [105]. Moreover, silica at the nanoscale has a crucial role in bone regeneration.
The findings also reveal that SNPs are an essential component of bone grafts in this study. SNPs were found to inhibit osteoclasts and activate osteoblasts in vitro. In mice, SNPs improved bone mineral density significantly [60]. In mice in vivo models, it was also found that the 50 nm SNPs also seemed to promote osteoblast proliferation while inhibiting osteoclast differentiation, thereby enhancing bone density as well as biochemical indicators of bone formation. SNPs play a key role in bone regeneration. Researchers discovered that magnetic nanoparticle (Fe3O4)-coated mesoporous silica nanoparticles (M-MSNs) could stimulate bone regeneration in rodents during distraction osteogenesis (DO) in mice [106]. A system was constructed using MSCs encapsulated with polymer to stimulate osteogenic differentiation and bone regeneration. SOST siRNA suppressed the osteoporosis-subjecting gene SOST, which lowers osteoblastic development, resulting in increased bone formation. A team of researchers synthesized bioactive glass nanoparticles with monodispersed strontium using improved Stöber’s method. SNPs with diameter 90 ± 10 nm were formed prior to incorporation of calcium and strontium [107]. MC3T3-E1 cells from mice were labeled with antibodies against colla1, osteocalcin (OSC), and osteopontin (OSP) for immunohistochemistry to determine the protein expression in response to nanoparticles after three weeks of growth. Furthermore, 14 percent Sr-BGNPs also notably demonstrate a high expression of markers, OSC and OSP, which is representative for osteogenic differentiation compared to zero percent Sr-BGNPs. These results indicate that the Sr-BGNPs serve as an effective agent for bone regeneration [107]. All the biomedical applications are summarized in Table 3.

6. Biocompatibility

The application and study of SNPs for further research are simplified due to their characteristics, which include huge surface area, facile surface functionalization, and strong biocompatibility. However, there are some safety issues concerning SNPs, the conflicting evidence reported on both in vitro and in vivo research. A wide number of toxicological studies demonstrate that SNPs’ toxicity depends on their surface characteristics, size, concentration, and other parameters. The processes underlying toxicity of SNP must be better understood before their implementation in nanomedicine. The countermeasures regarding the toxicity must be taken to minimize health consequences.

Toxicological Studies

SNPs induce cellular damage in a variety of different ways. Autophagy, oxidative stress, apoptosis, and inflammatory reactions are some of the mechanisms reported to be involved in cytotoxicity. Among those, ROS-mediated effects contribute highly to cellular damage. SNPs were shown to have deleterious effects on cells, even when the level of reactive oxygen species is not elevated [108,109,110]. Various investigations were conducted, which proves that the SNPs can induce cytotoxic effects in other ways by hindering cellular respiration, proliferation, and apoptosis. SNPs, for example, can trigger autophagy by altering the configuration of lysosomes, promoting membrane permeability, cathepsin B, and limiting the secretion of lysosomal proteases, impairing the lysosome function. SNP-induced autophagy is well documented, despite the unclear understanding of the precise mechanisms involved. Other adverse consequences of SNPs include immunotoxicity, genotoxicity, and neurotoxicity. Among these, genotoxicity and immunotoxicity are the two most noticeable toxic effects. SNPs may lead to the development of micronuclei at high cytotoxic concentrations [111]. These toxicity levels are consistent across cell lines and SNPs. SNPs have been found to cause genotoxicity in cancer cell lines, which is thought to be linked to the implementation of cellular damage in recent research. The size of the NPs and the extent of this effect are inversely proportional to one another [112]. Negatively charged SNPs have been shown to have the most immunotoxic activities in lymphocytes when examined in vivo [111]. They suppress the generation of proinflammatory cytokines and nitric oxide, while inhibiting the lymphocyte proliferation and NK cell killing capability.
In vivo investigations typically comprise animal toxicity trials where SNPs are supplied mainly through inhalation, topically, orally, or intravenously [12,113]. SNPs can be found across many regions of the body, including the liver, lungs, spleen, kidney, skin, and brain tissue, which solely depends on the exposure site [114,115,116,117]. NPs can be excreted from the body either through the kidneys, where it is filtered out by urine, or by the liver, through bile secretion in the feces. Researchers used ultrastructural localization and fluorescence methods and discovered that SNPs have distributed across the spleen and liver of mice after one shot as an intravenous injection produced substantial lymphocytic infiltration, degeneration, as well as necrosis [118]. The ease of excretion is decided by the smaller size of the nanoparticle injected. As per the outcomes of numerous scientific experiments on mice, there seems to be no buildup of SNPs on the brain tissue and no adverse consequences have been detected. A number of in vivo and in vitro studies demonstrate that SNPs have high toxicity on neurons. The central nervous system under specific circumstances was used as a reference for potential safety application and development [119]. The summary of toxicological effects of silica nanoparticles and their probable reasons are tabulated in Table 4. Comprehensive and systematic studies are required to clarify the deleterious effects of SNPs. However, further investigations on the role of SNP in the biochemical characteristics in vivo and in vitro are highly needed. Chronic toxicity and health impacts of persistent exposure, including co-exposure issues with nano-silica and its structural composition, must be assessed regularly. Nanotoxicity-related norms for detection techniques of toxicity and their potential implementation also have to be considered. Another technique to minimize possible toxicity is to analyze and synthesize biodegradable nanoparticle materials.
Synthesized SNPs contain unreacted chemicals, such as surfactants, organic solvents, etc., that cause cyto-, immuno-, and genotoxicity. In order to resolve the issue, the end product can be further purified before their use in biological systems. In addition to it, their surface can be functionalized with various functional groups, such as phosphate and amino groups. This functionalization can increase the selectivity and can interact with the diseased site, thereby reducing the dendritic effects caused. The alteration of the surface properties of SNP can also reduce the production of free radicals. Despite the surface chemistry, the shapes also determine the level of toxicity. Modification of the shapes can also be performed to reduce the toxic effects of SNPs.
Despite the significant interest in biomedical applications, the fate of the silica nanoparticles is one of the important aspects to be considered. The chemical transformation, degradation, excretion, and their ecotoxicity needs to be well understood before their potential use. The size, shape, and their surface charge have a great impact on their fate. Several studies indicate that the smaller particles will be subsequently eliminated through urine or feces, whereas the larger particles can be degraded by the liver. However, the ultra-fine particles in the range of approximately 6 nm become immediately eliminated by renal clearance. These types of particles cannot be used as drug carriers, but they have a preferential role in imaging and diagnostics. Research proved that the degradation and elimination rate can be modified according to the application by altering the size distribution, morphology, and surface chemistry.

7. Clinical Trials of Silica Nanoparticles

The US Food and Drug Administration (FDA) has recognized silica as generally recognized as safe (GRAS) under section 201(s) of the FD&C act after completing the clinical trial (NCT03667430) [120]. The European Food Safety Authority has recognized amorphous forms of silica and silicates up to 1500 mg per day as safe ingredients for oral delivery [120]. Several human trials were conducted for silica nanoparticles, as per ClinicalTrails.gov (accessed on 31 March 2022), though there is a lack of significant human trials on mesoporous silica nanoparticles for cancer therapy and diagnosis. Janjua et al. recently reported about 11 clinical trials and 2 clinical studies on humans to evaluate the safety profile of silica nanoparticles [121]. Silica nanoparticles improve the solubility of hydrophobic drugs and alter the pharmacokinetics profile. They show good tolerability with minimum side effects when used as an ingredient for oral delivery agents. Tan et al. reported an almost two-fold increase in the bioavailability of ibuprofen when it was formulated using silica–lipid hybrid nanoparticles [122]. A human trial was conducted on 12 adults to demonstrate the tolerability and pharmacokinetics of novel simvastatin encapsulated in silica–lipid hybrid nanoformulations (ACTRN12618001929291) [123]. The formulations show 3.5-fold increased bioavailability of simvastatin compared to the free drugs. Kharlamov et al. reported silica modified SPION and gold nanoparticles for localized plasmonic photothermal therapy to treat atherosclerosis [124]. A phase I clinical trial successfully established that the silica–metal core-shell nanostructure with a diameter of 90–150 nm could reduce coronary atherosclerosis (NCT01270139) [125]. Similarly, focal therapy, diagnosis, and imaging (MRI/Ultrasound) using silica–gold nanoshell showed better efficacy in malignant prostate cancer (NCT04240639, NCT02680535) [126]. The reference number NCT04656678 shows the preliminary effectiveness of prostate cancer focal US-guided theranostics. The main advantage includes the accumulation of nanoparticles or nanoshells at tumor sites that helps in accurate and predictable ablation therapy for cancer with minimal side effects compared to the conventional therapy. Ultrasmall silica nanoparticles, Cornel dots with a 6–10 nm diameter, can be used as tracers for melanoma or malignant brain tumors (NCT03465618) [127]. In a clinical trial (NCT01266096), PET imaging of a patient with melanoma and a malignant brain tumor was demonstrated using a 124I-labeled cRGDY silica nanomolecular particle tracer to establish the efficacy of the nanosilica [121]. Targeted silica nanoparticles can be used effectively for real-time image-guided mapping of nodal metastases (NCT02106598) [128]. The clinical trial results demonstrated that ultrasmall silica dots are stable, readily taken up by the tumor cells, and well-tolerated, with minimum side effects. Wiesner group has been working on silica-based nanoparticles since 2000 for cancer theranostics. They developed ultrasmall (<10 nm) fluorescent ‘C dots’ (Cornell dots) comprised a silica shell encapsulating Cy5 fluorescent dye coated with polyethylene glycol [129,130]. The particles were decorated with radiolabeled RGDY (124I-cRGDY) for integrin targeting. The first-generation C dots have been tested in first-in-human FDA-approved clinical trials. This was the first inorganic–polymer hybrid diagnostic probe used in clinical trials [131]. One can easily conclude that the silica nanoparticles are very promising in theranostics considering these many clinical trials.

8. Conclusions and Future Perspectives

Nanotechnology in medicine has advanced significantly due to the advances in science and technology in a variety of research methodologies. Silica nanoparticles have become one of the inevitable research hotspots in nanomedical applications, owing to a huge surface–volume ratio, simplicity in surface modification, easy fabrication, and amplification with excellent biocompatible properties. MSNs use their intrinsic mesoporous surface and porosity characteristics by altering the size of the pore and adjusting the characteristics of the surface to accomplish the drug delivery. Furthermore, in contrast with traditional SNPs, the refined SNPs being created now offer more significant potential for simultaneous molecular imaging and drug administration. Bioimaging can be used to track drug distribution in real time and enhance therapeutic strategies. SNPs are also useful in a variety of scientific domains, including wound healing, photodynamic therapy, photothermal therapy, antimicrobial infection, and bone regeneration. SNPs are widely employed; however, there are concerns about their biosafety. SNPs have been shown to induce cytotoxic effects, hepatosplenic toxicity, carcinogenicity, and hypersensitivity in both in vitro and in vivo investigations. However, the outcomes among many toxicology studies are uncertain, and the research path is unclear. Therefore, it is predicted that the use of SNPs in therapeutic applications would provide additional health advantages to humans in the future. Besides all these advantages, SNP-based nanomedicines face several challenges in biomedical applications. The fabrication of SNP-based nanomedicines is a multistep procedure that starts with synthesis and ends with surface modifications to serve specific purposes in drug delivery and imaging. These complicated syntheses and fabrications are multistep procedures that can be an issue in reproducibility and scaling-up in a cost-effective manner. The biosafety of SNPs is another challenge that biomedical scientists must address to bring SNP-based theranostics from bench to bedside, where chemical composition and physicochemical properties are critical parameters to maintain. Our intention is not to discourage researchers from stating these limitations and challenges. This field is up and coming, and researchers can address these limitations and challenges to develop new nanoformulations that we need to fight against several life-threatening diseases.

Author Contributions

Conceptualization, P.P. and A.G.; methodology, P.P., K.H., P.G. and K.G.; software, F.S.; validation, F.S., S.A. and M.M.G.; formal analysis, A.A.; investigation, F.S., S.A. and A.G.; resources, S.A.; data curation, A.A.; writing—original draft preparation, P.P.; writing—review and editing, S.A., A.G. and F.S.; visualization, S.A.; supervision, A.G.; project administration, A.G.; funding acquisition, S.A., M.M.G. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This article did not report any data.

Acknowledgments

The authors are thankful to AlMaarefa University for their financial support. CARE is also acknowledged for infrastructural support. P.P., K.H., and P.G. acknowledge CARE for fellowship and contingency.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Decuzzi, P.; Peer, D.; Di Mascolo, D.; Palange, A.L.; Manghnani, P.N.; Moghimi, S.M.; Farhangrazi, Z.S.; Howard, K.A.; Rosenblum, D.; Liang, T. Roadmap on nanomedicine. Nanotechnology 2020, 32, 12001. [Google Scholar] [CrossRef] [PubMed]
  2. Ghosh, S.; Girigoswami, K.; Girigoswami, A. Membrane-encapsulated camouflaged nanomedicines in drug delivery. Nanomedicine 2019, 14, 2067–2082. [Google Scholar] [CrossRef] [PubMed]
  3. Sharmiladevi, P.; Girigoswami, K.; Haribabu, V.; Girigoswami, A. Nano-enabled theranostics for cancer. Mater. Adv. 2021, 2, 2876–2891. [Google Scholar] [CrossRef]
  4. Haribabu, V.; Farook, A.S.; Goswami, N.; Murugesan, R.; Girigoswami, A. Optimized Mn-doped iron oxide nanoparticles entrapped in dendrimer for dual contrasting role in MRI. J. Biomed. Mater. Res. Part B Appl. Biomater. 2016, 104, 817–824. [Google Scholar] [CrossRef] [PubMed]
  5. Haribabu, V.; Sharmiladevi, P.; Akhtar, N.; Farook, A.S.; Girigoswami, K.; Girigoswami, A. Label free ultrasmall fluoromagnetic ferrite-clusters for targeted cancer imaging and drug delivery. Curr. Drug Del. 2019, 16, 233–241. [Google Scholar] [CrossRef]
  6. Kazi, M.; Alhajri, A.; Alshehri, S.M.; Elzayat, E.M.; Al Meanazel, O.T.; Shakeel, F.; Noman, O.; Altamimi, M.A.; Alanazi, F.K. Enhancing oral bioavailability of apigenin using a bioactive self-nanoemulsifying drug delivery system (Bio-SNEDDS): In vitro, in vivo and stability evaluations. Pharmaceutics 2020, 12, 749. [Google Scholar] [CrossRef]
  7. Hussain, A.; Altamimi, M.A.; Alshehri, S.; Imam, S.S.; Shakeel, F.; Singh, S.K. Novel approach for transdermal delivery of rifampicin to induce synergistic antimycobacterial effects against cutaneous and systemic tuberculosis using a cationic nanoemulsion gel. Int. J. Nanomed. 2020, 15, 1073. [Google Scholar] [CrossRef] [Green Version]
  8. Shakeel, F.; Salem-Bekhit, M.M.; Haq, N.; Alshehri, S. Nanoemulsification Improves the Pharmaceutical Properties and Bioactivities of Niaouli Essential Oil (Melaleuca quinquenervia L.). Molecules 2021, 26, 4750. [Google Scholar] [CrossRef]
  9. Rahamathulla, M.; Bhosale, R.R.; Osmani, R.A.; Mahima, K.C.; Johnson, A.P.; Hani, U.; Ghazwani, M.; Begum, M.Y.; Alshehri, S.; Ghoneim, M.M. Carbon Nanotubes: Current Perspectives on Diverse Applications in Targeted Drug Delivery and Therapies. Materials 2021, 14, 6707. [Google Scholar] [CrossRef]
  10. Perween, N.; Alshehri, S.; Easwari, T.; Verma, V.; Faiyazuddin, M.; Alanazi, A.; Shakeel, F. Investigating the feasibility of mefenamic acid nanosuspension for pediatric delivery: Preparation, characterization, and role of excipients. Processes 2021, 9, 574. [Google Scholar] [CrossRef]
  11. Javed, S.; Alshehri, S.; Shoaib, A.; Ahsan, W.; Sultan, M.H.; Alqahtani, S.S.; Kazi, M.; Shakeel, F. Chronicles of Nanoerythrosomes: An Erythrocyte-Based Biomimetic Smart Drug Delivery System as a Therapeutic and Diagnostic Tool in Cancer Therapy. Pharmaceutics 2021, 13, 368. [Google Scholar] [CrossRef] [PubMed]
  12. Li, Z.; Mu, Y.; Peng, C.; Lavin, M.F.; Shao, H.; Du, Z. Understanding the mechanisms of silica nanoparticles for nanomedicine. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnology 2021, 13, e1658. [Google Scholar] [CrossRef] [PubMed]
  13. Amsaveni, G.; Farook, A.S.; Haribabu, V.; Murugesan, R.; Girigoswami, A. Engineered multifunctional nanoparticles for DLA cancer cells targeting, sorting, MR imaging and drug delivery. Adv. Sci. Eng. Med. 2013, 5, 1340–1348. [Google Scholar] [CrossRef]
  14. Sharmiladevi, P.; Haribabu, V.; Girigoswami, K.; Farook, A.S.; Girigoswami, A. Effect of mesoporous nano water reservoir on MR relaxivity. Sci. Rep. 2017, 7, 11179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Sharmiladevi, P.; Akhtar, N.; Haribabu, V.; Girigoswami, K.; Chattopadhyay, S.; Girigoswami, A. Excitation wavelength independent carbon-decorated ferrite nanodots for multimodal diagnosis and stimuli responsive therapy. ACS Appl. Bio. Mater. 2019, 2, 1634–1642. [Google Scholar] [CrossRef] [PubMed]
  16. Haribabu, V.; Girigoswami, K.; Sharmiladevi, P.; Girigoswami, A. Water–Nanomaterial Interaction to Escalate Twin-Mode Magnetic Resonance Imaging. ACS Biomater. Sci. Eng. 2020, 6, 4377–4389. [Google Scholar] [CrossRef] [PubMed]
  17. Croissant, J.G.; Butler, K.S.; Zink, J.I.; Brinker, C.J. Synthetic amorphous silica nanoparticles: Toxicity, biomedical and environmental implications. Nat. Rev. Mater. 2020, 5, 886–909. [Google Scholar] [CrossRef]
  18. Fu, J.; Jiao, J.; Song, H.; Gu, Z.; Liu, Y.; Geng, J.; Jack, K.S.; Du, A.; Tang, J.; Yu, C. Fractal-in-a-sphere: Confined self-assembly of fractal silica nanoparticles. Chem. Mater. 2019, 32, 341–347. [Google Scholar] [CrossRef]
  19. De Jong, W.H.; Borm, P.J. Drug delivery and nanoparticles: Applications and hazards. Int. J. Nanomed. 2008, 3, 133. [Google Scholar] [CrossRef] [Green Version]
  20. He, Q.; Shi, J. Mesoporous silica nanoparticle based nano drug delivery systems: Synthesis, controlled drug release and delivery, pharmacokinetics and biocompatibility. J. Mater. Chem. 2011, 21, 5845–5855. [Google Scholar] [CrossRef]
  21. Girija, A.R.; Balasubramanian, S. Theragnostic potentials of core/shell mesoporous silica nanostructures. Nanotheranostics 2019, 3, 1–40. [Google Scholar] [CrossRef] [PubMed]
  22. Wilczewska, A.Z.; Niemirowicz, K.; Markiewicz, K.H.; Car, H. Nanoparticles as drug delivery systems. Pharmacol. Rep. 2012, 64, 1020–1037. [Google Scholar] [CrossRef] [PubMed]
  23. Kankala, R.K.; Han, Y.H.; Na, J.; Lee, C.H.; Sun, Z.; Wang, S.B.; Kimura, T.; Ok, Y.S.; Yamauchi, Y.; Chen, A.Z. Nanoarchitectured structure and surface biofunctionality of mesoporous silica nanoparticles. Adv. Mater. 2020, 32, 1907035. [Google Scholar] [CrossRef] [PubMed]
  24. Bimbo, L.M.; Sarparanta, M.; Santos, H.A.; Airaksinen, A.J.; Makila, E.; Laaksonen, T.; Peltonen, L.; Lehto, V.-P.; Hirvonen, J.; Salonen, J. Biocompatibility of thermally hydrocarbonized porous silicon nanoparticles and their biodistribution in rats. ACS Nano 2010, 4, 3023–3032. [Google Scholar] [CrossRef] [PubMed]
  25. Rosenholm, J.M.; Sahlgren, C.; Lindén, M. Towards multifunctional, targeted drug delivery systems using mesoporous silica nanoparticles—Opportunities & challenges. Nanoscale 2010, 2, 1870–1883. [Google Scholar]
  26. Kumar, R.; Roy, I.; Ohulchanskky, T.Y.; Vathy, L.A.; Bergey, E.J.; Sajjad, M.; Prasad, P.N. In vivo biodistribution and clearance studies using multimodal organically modified silica nanoparticles. ACS Nano 2010, 4, 699–708. [Google Scholar] [CrossRef] [Green Version]
  27. Hu, S.-H.; Liu, T.-Y.; Huang, H.-Y.; Liu, D.-M.; Chen, S.-Y. Magnetic-sensitive silica nanospheres for controlled drug release. Langmuir 2008, 24, 239–244. [Google Scholar] [CrossRef]
  28. Knopp, D.; Tang, D.; Niessner, R. Bioanalytical applications of biomolecule-functionalized nanometer-sized doped silica particles. Anal. Chim. Acta 2009, 647, 14–30. [Google Scholar] [CrossRef]
  29. Na, M.; Chen, Y.; Han, Y.; Ma, S.; Liu, J.; Chen, X. Determination of potassium ferrocyanide in table salt and salted food using a water-soluble fluorescent silicon quantum dots. Food Chem. 2019, 288, 248–255. [Google Scholar] [CrossRef]
  30. Maity, A.; Polshettiwar, V. Dendritic fibrous nanosilica for catalysis, energy harvesting, carbon dioxide mitigation, drug delivery, and sensing. ChemSusChem 2017, 10, 3866. [Google Scholar] [CrossRef] [Green Version]
  31. Chen, W.; Glackin, C.A.; Horwitz, M.A.; Zink, J.I. Nanomachines and other caps on mesoporous silica nanoparticles for drug delivery. Acc. Chem. Res. 2019, 52, 1531–1542. [Google Scholar] [CrossRef] [PubMed]
  32. Paris, J.L.; Baeza, A.; Vallet-Regí, M. Overcoming the stability, toxicity, and biodegradation challenges of tumor stimuli-responsive inorganic nanoparticles for delivery of cancer therapeutics. Expert Opin. Drug Deliv. 2019, 16, 1095–1112. [Google Scholar] [CrossRef]
  33. Ren, D.; Xu, J.; Chen, N.; Ye, Z.; Li, X.; Chen, Q.; Ma, S. Controlled synthesis of mesoporous silica nanoparticles with tunable architectures via oil-water microemulsion assembly process. Colloids Surf. Physicochem. Eng. Asp. 2021, 611, 125773. [Google Scholar] [CrossRef]
  34. Rosenberg, D.J.; Alayoglu, S.; Kostecki, R.; Ahmed, M. Synthesis of microporous silica nanoparticles to study water phase transitions by vibrational spectroscopy. Nanoscale Adv. 2019, 1, 4878–4887. [Google Scholar] [CrossRef] [Green Version]
  35. Lu, Y.; Fan, H.; Stump, A.; Ward, T.L.; Rieker, T.; Brinker, C.J. Aerosol-assisted self-assembly of mesostructured spherical nanoparticles. Nature 1999, 398, 223–226. [Google Scholar] [CrossRef] [Green Version]
  36. Mahtabani, A.; La Zara, D.; Anyszka, R.; He, X.; Paajanen, M.; Van Ommen, J.R.; Dierkes, W.; Blume, A. Gas Phase Modification of Silica Nanoparticles in a Fluidized Bed: Tailored Deposition of Aminopropylsiloxane. Langmuir 2021, 37, 4481–4492. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, S.-H.; Mou, C.-Y.; Lin, H.-P. Synthesis of mesoporous silica nanoparticles. Chem. Soc. Rev. 2013, 42, 3862–3875. [Google Scholar] [CrossRef] [PubMed]
  38. Stöber, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci. 1968, 26, 62–69. [Google Scholar] [CrossRef]
  39. Liu, X.; Lu, X.; Wen, P.; Shu, X.; Chi, F. Synthesis of ultrasmall silica nanoparticles for application as deep-ultraviolet antireflection coatings. Appl. Surf. Sci. 2017, 420, 180–185. [Google Scholar] [CrossRef]
  40. Fernandes, R.S.; Raimundo, I.M., Jr.; Pimentel, M.F. Revising the synthesis of Stöber silica nanoparticles: A multivariate assessment study on the effects of reaction parameters on the particle size. Colloids Surf. Physicochem. Eng. Aspects 2019, 577, 1–7. [Google Scholar] [CrossRef]
  41. Nele, M.; Vidal, A.; Bhering, D.L.; Pinto, J.C.; Salim, V.M.M. Preparation of high loading silica supported nickel catalyst: Simultaneous analysis of the precipitation and aging steps. Appl. Catal. A Gen. 1999, 178, 177–189. [Google Scholar] [CrossRef]
  42. Alvarez-Toral, A.; Fernández, B.; Malherbe, J.; Claverie, F.; Pecheyran, C.; Pereiro, R. Synthesis of amino-functionalized silica nanoparticles for preparation of new laboratory standards. Spectrochim. Acta Part B At. Spectrosc. 2017, 138, 1–7. [Google Scholar] [CrossRef] [Green Version]
  43. Ghosh, U.; Nagalakshmi, M.; Sadhukhan, R.; Girigoswami, A. Biocapped nanoparticles are more bioactive to HeLa cells than its chemical counterpart. Adv. Sci. Eng. Med. 2013, 5, 783–788. [Google Scholar] [CrossRef]
  44. Karande, S.D.; Jadhav, S.A.; Garud, H.B.; Kalantre, V.A.; Burungale, S.H.; Patil, P.S. Green and sustainable synthesis of silica nanoparticles. Nanotechnol. Environ. Eng. 2021, 6, 1–14. [Google Scholar] [CrossRef]
  45. Huang, R.; Shen, Y.-W.; Guan, Y.-Y.; Jiang, Y.-X.; Wu, Y.; Rahman, K.; Zhang, L.-J.; Liu, H.-J.; Luan, X. Mesoporous silica nanoparticles: Facile surface functionalization and versatile biomedical applications in oncology. Acta Biomater. 2020, 116, 1–15. [Google Scholar] [CrossRef]
  46. Yin, Q.; Shen, J.; Zhang, Z.; Yu, H.; Li, Y. Reversal of multidrug resistance by stimuli-responsive drug delivery systems for therapy of tumor. Adv. Drug Del. Rev. 2013, 65, 1699–1715. [Google Scholar] [CrossRef]
  47. Su, S.; Kang, P.M. Recent advances in nanocarrier-assisted therapeutics delivery systems. Pharmaceutics 2020, 12, 837. [Google Scholar] [CrossRef]
  48. Juneja, R.; Vadarevu, H.; Halman, J.; Tarannum, M.; Rackley, L.; Dobbs, J.; Marquez, J.; Chandler, M.; Afonin, K.; Vivero-Escoto, J.L. Combination of nucleic acid and mesoporous silica nanoparticles: Optimization and therapeutic performance in vitro. ACS Appl. Mater. Interfaces 2020, 12, 38873–38886. [Google Scholar] [CrossRef]
  49. Karaman, D.Ş.; Pamukçu, A.; Karakaplan, M.B.; Kocaoglu, O.; Rosenholm, J.M. Recent Advances in the Use of Mesoporous Silica Nanoparticles for the Diagnosis of Bacterial Infections. Int. J. Nanomed. 2021, 16, 6575. [Google Scholar] [CrossRef]
  50. Ianăşi, C.; Picioruş, E.-M.; Nicola, R.; Putz, A.-M.; Negrea, A.; Ciopec, M.; Len, A.; Almásy, L. Synthesis and characterization of magnetic iron oxide—silica nanocomposites used for adsorptive recovery of palladium (II). Soft Mater. 2021, 1–8. [Google Scholar] [CrossRef]
  51. Deepika, R.; Girigoswami, K.; Murugesan, R.; Girigoswami, A. Influence of divalent cation on morphology and drug delivery efficiency of mixed polymer nanoparticles. Curr. Drug Del. 2018, 15, 652–657. [Google Scholar] [CrossRef] [PubMed]
  52. Vimaladevi, M.; Divya, K.C.; Girigoswami, A. Liposomal nanoformulations of rhodamine for targeted photodynamic inactivation of multidrug resistant gram negative bacteria in sewage treatment plant. J. Photochem. Photobiol. B Biol. 2016, 162, 146–152. [Google Scholar] [CrossRef]
  53. Niu, Y.; Yu, M.; Hartono, S.B.; Yang, J.; Xu, H.; Zhang, H.; Zhang, J.; Zou, J.; Dexter, A.; Gu, W. Nanoparticles mimicking viral surface topography for enhanced cellular delivery. Adv. Mater. 2013, 25, 6233–6237. [Google Scholar] [CrossRef] [PubMed]
  54. Harun, S.N.; Ahmad, H.; Lim, H.N.; Chia, S.L.; Gill, M.R. Synthesis and optimization of mesoporous silica nanoparticles for ruthenium polypyridyl drug delivery. Pharmaceutics 2021, 13, 150. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, W.; Wang, H.; Wang, H.; Han, Y.; Zheng, Z.; Liu, X.; Feng, B.; Zhang, H. Light-responsive dual-functional biodegradable mesoporous silica nanoparticles with drug delivery and lubrication enhancement for the treatment of osteoarthritis. Nanoscale 2021, 13, 6394–6399. [Google Scholar] [CrossRef] [PubMed]
  56. Gou, K.; Wang, Y.; Guo, X.; Wang, Y.; Bian, Y.; Zhao, H.; Guo, Y.; Pang, Y.; Xie, L.; Li, S. Carboxyl-functionalized mesoporous silica nanoparticles for the controlled delivery of poorly water-soluble non-steroidal anti-inflammatory drugs. Acta Biomater. 2021, 134, 576–592. [Google Scholar] [CrossRef]
  57. Xu, D.; Song, X.; Zhou, J.; Ouyang, X.; Li, J.; Deng, D. Virus-like hollow mesoporous silica nanoparticles for cancer combination therapy. Colloids Surf. B Biointerfaces 2021, 197, 111452. [Google Scholar] [CrossRef]
  58. Wang, Y.; Zhao, Q.; Han, N.; Bai, L.; Li, J.; Liu, J.; Che, E.; Hu, L.; Zhang, Q.; Jiang, T. Mesoporous silica nanoparticles in drug delivery and biomedical applications. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 313–327. [Google Scholar] [CrossRef]
  59. Firmansyah, A.; Nugrahani, I.; Wirasutisna, K.; Ibrahim, S. Formation of boron-silica based mesoporous and studies of its adsorption ability for curcuminoids. Biointerface Res. Appl. Chem. 2020, 10, 7977–7981. [Google Scholar]
  60. Shadjou, N.; Hasanzadeh, M. Bone tissue engineering using silica-based mesoporous nanobiomaterials: Recent progress. Mater. Sci. Eng. C 2015, 55, 401–409. [Google Scholar] [CrossRef]
  61. Roggers, R.A.; Lin, V.S.-Y.; Trewyn, B.G. Chemically reducible lipid bilayer coated mesoporous silica nanoparticles demonstrating controlled release and HeLa and normal mouse liver cell biocompatibility and cellular internalization. Mol. Pharm. 2012, 9, 2770–2777. [Google Scholar] [CrossRef] [PubMed]
  62. Dolmans, D.E.; Fukumura, D.; Jain, R.K. Photodynamic therapy for cancer. Nat. Rev. Cancer 2003, 3, 380–387. [Google Scholar] [CrossRef]
  63. Bouramtane, S.; Bretin, L.; Pinon, A.; Leger, D.; Liagre, B.; Richard, L.; Brégier, F.; Sol, V.; Chaleix, V. Porphyrin-xylan-coated silica nanoparticles for anticancer photodynamic therapy. Carbohydr. Polym. 2019, 213, 168–175. [Google Scholar] [CrossRef]
  64. Zhou, Y.; Chang, C.; Liu, Z.; Zhao, Q.; Xu, Q.; Li, C.; Chen, Y.; Zhang, Y.; Lu, B. Hyaluronic acid-functionalized hollow mesoporous silica nanoparticles as pH-sensitive nanocarriers for cancer chemo-photodynamic therapy. Langmuir 2021, 37, 2619–2628. [Google Scholar] [CrossRef]
  65. Prieto-Montero, R.; Prieto-Castañeda, A.; Katsumiti, A.; Cajaraville, M.P.; Agarrabeitia, A.R.; Ortiz, M.J.; Martínez-Martínez, V. Functionalization of Photosensitized Silica Nanoparticles for Advanced Photodynamic Therapy of Cancer. Int. J. Mol. Sci. 2021, 22, 6618. [Google Scholar] [CrossRef] [PubMed]
  66. Gao, Y.; Gao, D.; Shen, J.; Wang, Q. A review of mesoporous silica nanoparticle delivery systems in chemo-based combination cancer therapies. Front. Chem. 2020, 8, 1086. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, Y.; Niu, C.; Fan, S.; Li, Y.; Li, X.; Dai, Y.; Shi, J.; Wang, X. Indocyanine Green Loaded Modified Mesoporous Silica Nanoparticles as an Effective Photothermal Nanoplatform. Int. J. Mol. Sci. 2020, 21, 4789. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, Z.; Chang, Z.-M.; Shao, D.; Zhang, F.; Chen, F.; Li, L.; Ge, M.-F.; Hu, R.; Zheng, X.; Wang, Y. Janus gold triangle-mesoporous silica nanoplatforms for hypoxia-activated radio-chemo-photothermal therapy of liver cancer. ACS Appl. Mater. Interfaces 2019, 11, 34755–34765. [Google Scholar] [CrossRef]
  69. Cha, B.G.; Kim, J. Functional mesoporous silica nanoparticles for bio-imaging applications. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnology 2019, 11, e1515. [Google Scholar] [CrossRef] [Green Version]
  70. Song, B.; Liu, Q.; Ma, H.; Tang, Z.; Liu, C.; Zou, J.; Tan, M.; Yuan, J. Tumor-targetable magnetoluminescent silica nanoparticles for bimodal time-gated luminescence/magnetic resonance imaging of cancer cells in vitro and in vivo. Talanta 2020, 220, 121378. [Google Scholar] [CrossRef]
  71. Khatik, R.; Wang, Z.; Li, F.; Zhi, D.; Kiran, S.; Dwivedi, P.; Xu, R.X.; Liang, G.; Qiu, B.; Yang, Q. “Magnus nano-bullets” as T1/T2 based dual-modal for in vitro and in vivo MRI visualization. Nanomed. Nanotechnol. Biol. Med. 2019, 15, 264–273. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, F.; Ma, M.; Wang, J.; Wang, F.; Chern, S.-X.; Zhao, E.R.; Jhunjhunwala, A.; Darmadi, S.; Chen, H.; Jokerst, J.V. Exosome-like silica nanoparticles: A novel ultrasound contrast agent for stem cell imaging. Nanoscale 2017, 9, 402–411. [Google Scholar] [CrossRef] [PubMed]
  73. Kempen, P.J.; Greasley, S.; Parker, K.A.; Campbell, J.L.; Chang, H.-Y.; Jones, J.R.; Sinclair, R.; Gambhir, S.S.; Jokerst, J.V. Theranostic mesoporous silica nanoparticles biodegrade after pro-survival drug delivery and ultrasound/magnetic resonance imaging of stem cells. Theranostics 2015, 5, 631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Huang, X.; Zhang, F.; Lee, S.; Swierczewska, M.; Kiesewetter, D.O.; Lang, L.; Zhang, G.; Zhu, L.; Gao, H.; Choi, H.S. Long-term multimodal imaging of tumor draining sentinel lymph nodes using mesoporous silica-based nanoprobes. Biomaterials 2012, 33, 4370–4378. [Google Scholar] [CrossRef] [Green Version]
  75. Lee, H.; Shin, T.-H.; Cheon, J.; Weissleder, R. Recent developments in magnetic diagnostic systems. Chem. Rev. 2015, 115, 10690–10724. [Google Scholar] [CrossRef] [Green Version]
  76. Pham, X.-H.; Hahm, E.; Kim, H.-M.; Son, B.S.; Jo, A.; An, J.; Tran Thi, T.A.; Nguyen, D.Q.; Jun, B.-H. Silica-coated magnetic iron oxide nanoparticles grafted onto graphene oxide for protein isolation. Nanomaterials 2020, 10, 117. [Google Scholar] [CrossRef] [Green Version]
  77. Mostafaei, M.; Hosseini, S.N.; Khatami, M.; Javidanbardan, A.; Sepahy, A.A.; Asadi, E. Isolation of recombinant Hepatitis B surface antigen with antibody-conjugated superparamagnetic Fe3O4/SiO2 core-shell nanoparticles. Protein Expr. Purif. 2018, 145, 1–6. [Google Scholar] [CrossRef]
  78. Choi, J.R.; Hu, J.; Tang, R.; Gong, Y.; Feng, S.; Ren, H.; Wen, T.; Li, X.; Abas, W.A.B.W.; Pingguan-Murphy, B. An integrated paper-based sample-to-answer biosensor for nucleic acid testing at the point of care. Lab. Chip 2016, 16, 611–621. [Google Scholar] [CrossRef]
  79. Melzak, K.A.; Sherwood, C.S.; Turner, R.F.; Haynes, C.A. Driving forces for DNA adsorption to silica in perchlorate solutions. J. Colloid Interface Sci. 1996, 181, 635–644. [Google Scholar] [CrossRef]
  80. Min, J.H.; Woo, M.-K.; Yoon, H.Y.; Jang, J.W.; Wu, J.H.; Lim, C.-S.; Kim, Y.K. Isolation of DNA using magnetic nanoparticles coated with dimercaptosuccinic acid. Anal. Biochem. 2014, 447, 114–118. [Google Scholar] [CrossRef]
  81. Ma, C.; Li, C.; He, N.; Wang, F.; Ma, N.; Zhang, L.; Lu, Z.; Ali, Z.; Xi, Z.; Li, X. Preparation and characterization of monodisperse core–shell Fe3O4@ SiO2 microspheres and its application for magnetic separation of nucleic acids from E. coli BL21. J. Biomed. Nanotechnol. 2012, 8, 1000–1005. [Google Scholar] [CrossRef] [PubMed]
  82. Yue, H.; Shin, J.M.; Tegafaw, T.; Han, H.S.; Chae, K.-S.; Chang, Y.; Lee, G.H. Magnetic separation of nucleic acids from various biological samples using silica-coated iron oxide nanobeads. J. Nanopart. Res. 2020, 22, 1–12. [Google Scholar] [CrossRef]
  83. Zhou, Y.; Quan, G.; Wu, Q.; Zhang, X.; Niu, B.; Wu, B.; Huang, Y.; Pan, X.; Wu, C. Mesoporous silica nanoparticles for drug and gene delivery. Acta Pharm. Sin. B 2018, 8, 165–177. [Google Scholar] [CrossRef] [PubMed]
  84. Bharali, D.J.; Klejbor, I.; Stachowiak, E.K.; Dutta, P.; Roy, I.; Kaur, N.; Bergey, E.J.; Prasad, P.N.; Stachowiak, M.K. Organically modified silica nanoparticles: A nonviral vector for in vivo gene delivery and expression in the brain. Proc. Natl. Acad. Sci. USA 2005, 102, 11539–11544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Xia, T.; Kovochich, M.; Liong, M.; Meng, H.; Kabehie, S.; George, S.; Zink, J.I.; Nel, A.E. Polyethyleneimine coating enhances the cellular uptake of mesoporous silica nanoparticles and allows safe delivery of siRNA and DNA constructs. ACS Nano 2009, 3, 3273–3286. [Google Scholar] [CrossRef]
  86. Roy, I.; Ohulchanskyy, T.Y.; Bharali, D.J.; Pudavar, H.E.; Mistretta, R.A.; Kaur, N.; Prasad, P.N. Optical tracking of organically modified silica nanoparticles as DNA carriers: A nonviral, nanomedicine approach for gene delivery. Proc. Natl. Acad. Sci. USA 2005, 102, 279–284. [Google Scholar] [CrossRef] [Green Version]
  87. Hartono, S.B.; Phuoc, N.T.; Yu, M.; Jia, Z.; Monteiro, M.J.; Qiao, S.; Yu, C. Functionalized large pore mesoporous silica nanoparticles for gene delivery featuring controlled release and co-delivery. J. Mater. Chem. B 2014, 2, 718–726. [Google Scholar] [CrossRef]
  88. Mody, K.T.; Popat, A.; Mahony, D.; Cavallaro, A.S.; Yu, C.; Mitter, N. Mesoporous silica nanoparticles as antigen carriers and adjuvants for vaccine delivery. Nanoscale 2013, 5, 5167–5179. [Google Scholar] [CrossRef]
  89. Gordon, S.; Teichmann, E.; Young, K.; Finnie, K.; Rades, T.; Hook, S. In vitro and in vivo investigation of thermosensitive chitosan hydrogels containing silica nanoparticles for vaccine delivery. Eur. J. Pharm. Sci. 2010, 41, 360–368. [Google Scholar] [CrossRef]
  90. Qiao, L.; Chen, M.; Li, S.; Hu, J.; Gong, C.; Zhang, Z.; Cao, X. A peptide-based subunit candidate vaccine against SARS-CoV-2 delivered by biodegradable mesoporous silica nanoparticles induced high humoral and cellular immunity in mice. Biomater. Sci. 2021, 9, 7287–7296. [Google Scholar] [CrossRef]
  91. Lee, J.Y.; Kim, M.K.; Nguyen, T.L.; Kim, J. Hollow mesoporous silica nanoparticles with extra-large mesopores for enhanced cancer vaccine. ACS Appl. Mater. Interfaces 2020, 12, 34658–34666. [Google Scholar] [CrossRef] [PubMed]
  92. Montalvo-Quirós, S.; Vallet-Regí, M.; Palacios, A.; Anguita, J.; Prados-Rosales, R.C.; González, B.; Luque-Garcia, J.L. Mesoporous Silica Nanoparticles as a Potential Platform for Vaccine Development against Tuberculosis. Pharmaceutics 2020, 12, 1218. [Google Scholar] [CrossRef] [PubMed]
  93. Abdelwahab, W.M.; Riffey, A.; Buhl, C.; Johnson, C.; Ryter, K.; Evans, J.T.; Burkhart, D.J. Co-adsorption of synthetic Mincle agonists and antigen to silica nanoparticles for enhanced vaccine activity: A formulation approach to co-delivery. Int. J. Pharm. 2021, 593, 120119. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Z.; Ru, J.; Sun, S.; Teng, Z.; Dong, H.; Song, P.; Yang, Y.; Guo, H. Uniform dendrimer-like mesoporous silica nanoparticles as a nano-adjuvant for foot-and-mouth disease virus-like particle vaccine. J. Mater. Chem. B 2019, 7, 3446–3454. [Google Scholar] [CrossRef]
  95. Song, H.; Yang, Y.; Tang, J.; Gu, Z.; Wang, Y.; Zhang, M.; Yu, C. DNA vaccine mediated by rambutan-like mesoporous silica nanoparticles. Adv. Ther. 2020, 3, 1900154. [Google Scholar] [CrossRef]
  96. Cha, B.G.; Jeong, J.H.; Kim, J. Extra-large pore mesoporous silica nanoparticles enabling co-delivery of high amounts of protein antigen and toll-like receptor 9 agonist for enhanced cancer vaccine efficacy. ACS Cent. Sci. 2018, 4, 484–492. [Google Scholar] [CrossRef] [Green Version]
  97. Soares, D.C.F.; Soares, L.M.; de Goes, A.M.; Melo, E.M.; de Barros, A.L.B.; Bicalho, T.C.A.S.; Leao, N.M.; Tebaldi, M.L. Mesoporous SBA-16 silica nanoparticles as a potential vaccine adjuvant against Paracoccidioides brasiliensis. Microporous Mesoporous Mater. 2020, 291, 109676. [Google Scholar] [CrossRef]
  98. Amin, M.K.; Boateng, J.S. Surface modification of mobile composition of matter (MCM)-41 type silica nanoparticles for potential oral mucosa vaccine delivery. J. Pharm. Sci. 2020, 109, 2271–2283. [Google Scholar] [CrossRef]
  99. Hajizade, A.; Salmanian, A.H.; Amani, J.; Ebrahimi, F.; Arpanaei, A. EspA-loaded mesoporous silica nanoparticles can efficiently protect animal model against enterohaemorrhagic E. coli O157: H7. Artif. Cells Nanomed. Biotechnol. 2018, 46, S1067–S1075. [Google Scholar] [CrossRef] [Green Version]
  100. Injorhor, P.; Ruksakulpiwat, Y.; Ruksakulpiwat, C. Effect of shrimp shell chitosan loading on antimicrobial, absorption and morphological properties of natural rubber composites reinforced with silica-chitosan hybrid filler. Biointerface Res. Appl. Chem 2020, 10, 5656–5659. [Google Scholar]
  101. Alghuthaymi, M. Magnetic-silica nanoshell for extraction of fungal genomic DNA from Rhizopus oryzae. Biointerface Res. Appl. Chem. 2020, 10, 4972–4976. [Google Scholar]
  102. Miller, K.P. Bacterial Communication and Its Role As a Target for Nanoparticlebased Antimicrobial Therapy. Doctoral Dissertation, University of South Carolina, Columbia, SC, USA, 2015. [Google Scholar]
  103. Masood, A.; Maheen, S.; Khan, H.U.; Shafqat, S.S.; Irshad, M.; Aslam, I.; Rasul, A.; Bashir, S.; Zafar, M.N. Pharmaco-Technical Evaluation of Statistically Formulated and Optimized Dual Drug-Loaded Silica Nanoparticles for Improved Antifungal Efficacy and Wound Healing. ACS Omega 2021, 6, 8210–8225. [Google Scholar] [CrossRef] [PubMed]
  104. Abolghasemzade, S.; Pourmadadi, M.; Rashedi, H.; Yazdian, F.; Kianbakht, S.; Navaei-Nigjeh, M. PVA based nanofiber containing CQDs modified with silica NPs and silk fibroin accelerates wound healing in a rat model. J. Mater. Chem. B 2021, 9, 658–676. [Google Scholar] [CrossRef]
  105. Quignard, S.; Coradin, T.; Powell, J.J.; Jugdaohsingh, R. Silica nanoparticles as sources of silicic acid favoring wound healing in vitro. Colloids Surf. B Biointerfaces 2017, 155, 530–537. [Google Scholar] [CrossRef] [PubMed]
  106. Shi, Y.; Li, Y.; Coradin, T. Magnetically-oriented type I collagen-SiO2@ Fe3O4 rods composite hydrogels tuning skin cell growth. Colloids Surf. B Biointerfaces 2020, 185, 110597. [Google Scholar] [CrossRef] [PubMed]
  107. Naruphontjirakul, P.; Porter, A.E.; Jones, J.R. In vitro osteogenesis by intracellular uptake of strontium containing bioactive glass nanoparticles. Acta Biomater. 2018, 66, 67–80. [Google Scholar] [CrossRef] [PubMed]
  108. Guo, C.; Yang, M.; Jing, L.; Wang, J.; Yu, Y.; Li, Y.; Duan, J.; Zhou, X.; Li, Y.; Sun, Z. Amorphous silica nanoparticles trigger vascular endothelial cell injury through apoptosis and autophagy via reactive oxygen species-mediated MAPK/Bcl-2 and PI3K/Akt/mTOR signaling. Int. J. Nanomedicine 2016, 11, 5257. [Google Scholar] [CrossRef] [Green Version]
  109. Kim, I.-Y.; Joachim, E.; Choi, H.; Kim, K. Toxicity of silica nanoparticles depends on size, dose, and cell type. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 1407–1416. [Google Scholar] [CrossRef]
  110. Duan, J.; Yu, Y.; Li, Y.; Yu, Y.; Li, Y.; Zhou, X.; Huang, P.; Sun, Z. Toxic effect of silica nanoparticles on endothelial cells through DNA damage response via Chk1-dependent G2/M checkpoint. PLoS ONE 2013, 8, e62087. [Google Scholar] [CrossRef] [Green Version]
  111. Decan, N.; Wu, D.; Williams, A.; Bernatchez, S.; Johnston, M.; Hill, M.; Halappanavar, S. Characterization of in vitro genotoxic, cytotoxic and transcriptomic responses following exposures to amorphous silica of different sizes. Mutat. Res. /Genet. Toxicol. Environ. Mutagen. 2016, 796, 8–22. [Google Scholar] [CrossRef]
  112. Guo, C.; Wang, J.; Jing, L.; Ma, R.; Liu, X.; Gao, L.; Cao, L.; Duan, J.; Zhou, X.; Li, Y. Mitochondrial dysfunction, perturbations of mitochondrial dynamics and biogenesis involved in endothelial injury induced by silica nanoparticles. Environ. Pollut. 2018, 236, 926–936. [Google Scholar] [CrossRef] [PubMed]
  113. Huang, X.; Li, L.; Liu, T.; Hao, N.; Liu, H.; Chen, D.; Tang, F. The shape effect of mesoporous silica nanoparticles on biodistribution, clearance, and biocompatibility in vivo. ACS Nano 2011, 5, 5390–5399. [Google Scholar] [CrossRef] [PubMed]
  114. Mohammadpour, R.; Cheney, D.L.; Grunberger, J.W.; Yazdimamaghani, M.; Jedrzkiewicz, J.; Isaacson, K.J.; Dobrovolskaia, M.A.; Ghandehari, H. One-year chronic toxicity evaluation of single dose intravenously administered silica nanoparticles in mice and their Ex vivo human hemocompatibility. J. Control. Release 2020, 324, 471–481. [Google Scholar] [CrossRef] [PubMed]
  115. Azouz, R.A.; Korany, R. Toxic impacts of amorphous silica nanoparticles on liver and kidney of male adult rats: An in vivo study. Biol. Trace Elem. Res. 2021, 199, 2653–2662. [Google Scholar] [CrossRef]
  116. Deng, Y.-D.; Zhang, X.-D.; Yang, X.-S.; Huang, Z.-L.; Wei, X.; Yang, X.-F.; Liao, W.-Z. Subacute toxicity of mesoporous silica nanoparticles to the intestinal tract and the underlying mechanism. J. Hazard. Mater. 2021, 409, 124502. [Google Scholar] [CrossRef]
  117. Yu, Y.; Duan, J.; Li, Y.; Li, Y.; Jing, L.; Yang, M.; Wang, J.; Sun, Z. Silica nanoparticles induce liver fibrosis via TGF-β1/Smad3 pathway in ICR mice. Int. J. Nanomed. 2017, 12, 6045. [Google Scholar] [CrossRef] [Green Version]
  118. Mohammadpour, R.; Dobrovolskaia, M.A.; Cheney, D.L.; Greish, K.F.; Ghandehari, H. Subchronic and chronic toxicity evaluation of inorganic nanoparticles for delivery applications. Adv. Drug Del. Rev. 2019, 144, 112–132. [Google Scholar] [CrossRef]
  119. Wu, J.; Wang, C.; Sun, J.; Xue, Y. Neurotoxicity of silica nanoparticles: Brain localization and dopaminergic neurons damage pathways. ACS Nano 2011, 5, 4476–4489. [Google Scholar] [CrossRef]
  120. Diab, R.; Canilho, N.; Pavel, I.A.; Haffner, F.B.; Girardon, M.; Pasc, A. Silica-based systems for oral delivery of drugs, macromolecules and cells. Adv. Colloid. Interface Sci. 2017, 249, 346–362. [Google Scholar] [CrossRef]
  121. Janjua, T.I.; Cao, Y.; Yu, C.; Popat, A. Clinical translation of silica nanoparticles. Nat. Rev. Mater. 2021, 6, 1072–1074. [Google Scholar] [CrossRef]
  122. Tan, A.; Eskandar, N.G.; Rao, S.; Prestidge, C.A. First in man bioavailability and tolerability studies of a silica–lipid hybrid (Lipoceramic) formulation: A Phase I study with ibuprofen. Drug Deliv. Transl. Res. 2014, 4, 212–221. [Google Scholar] [CrossRef]
  123. Meola, T.R.; Abuhelwa, A.Y.; Joyce, P.; Clifton, P.; Prestidge, C.A. A safety, tolerability, and pharmacokinetic study of a novel simvastatin silica-lipid hybrid formulation in healthy male participants. Drug Deliv. Transl. Res. 2021, 11, 1261–1272. [Google Scholar] [CrossRef] [PubMed]
  124. Kharlamov, A.N.; Feinstein, J.A.; Cramer, J.A.; Boothroyd, J.A.; Shishkina, E.V.; Shur, V. Plasmonic photothermal therapy of atherosclerosis with nanoparticles: Long-term outcomes and safety in NANOM-FIM trial. Future Cardiol. 2017, 13, 345–363. [Google Scholar] [CrossRef] [PubMed]
  125. Kharlamov, A. Plasmonic Photothermal Therapy of Atherosclerosis And Preparation Of Target Lesion In Patients With Arterial Remodeling: Subanalysis Of Nanom-Fim Trial. Atherosclerosis 2019, 287, e34. [Google Scholar] [CrossRef]
  126. Han, H.; Choi, K. Advances in Nanomaterial-Mediated Photothermal Cancer Therapies: Toward Clinical Applications. Biomedicines 2021, 9, 305. [Google Scholar] [CrossRef]
  127. Tyagi, P. Harnessing the nano-bio interface: Application of membrane coating to long acting silica particles. Eur. J. Pharm. Biopharm. 2021, 158, 382–389. [Google Scholar] [CrossRef]
  128. Carvalho, M.; Reis, R.; Oliveira, J.M. Dendrimer nanoparticles for colorectal cancer applications. J. Mater. Chem. B 2020, 8, 1128–1138. [Google Scholar] [CrossRef]
  129. Phillips, E.; Penate-Medina, O.; Zanzonico, P.B.; Carvajal, R.D.; Mohan, P.; Ye, Y.; Humm, J.; Gönen, M.; Kalaigian, H.; Schöder, H. Clinical translation of an ultrasmall inorganic optical-PET imaging nanoparticle probe. Sci. Transl. Med. 2014, 6, 260ra149. [Google Scholar] [CrossRef] [Green Version]
  130. Ma, K.; Mendoza, C.; Hanson, M.; Werner-Zwanziger, U.; Zwanziger, J.; Wiesner, U. Control of ultrasmall sub-10 nm ligand-functionalized fluorescent core—Shell silica nanoparticle growth in water. Chem. Mater. 2015, 27, 4119–4133. [Google Scholar] [CrossRef]
  131. Gidwani, B.; Sahu, V.; Shukla, S.S.; Pandey, R.; Joshi, V.; Jain, V.K.; Vyas, A. Quantum dots: Prospectives, toxicity, advances and applications. J. Drug Deliv. Sci. Technol. 2021, 61, 102308. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram representing microemulsion method for the synthesis of silica nanoparticles (SNPs).
Figure 1. Schematic diagram representing microemulsion method for the synthesis of silica nanoparticles (SNPs).
Processes 10 02595 g001
Figure 2. Schematic representation of Stöber’s method.
Figure 2. Schematic representation of Stöber’s method.
Processes 10 02595 g002
Figure 3. Biomedical applications of SNPs.
Figure 3. Biomedical applications of SNPs.
Processes 10 02595 g003
Table 1. Comparison of different synthetic routes in terms of advantages and disadvantages.
Table 1. Comparison of different synthetic routes in terms of advantages and disadvantages.
Sl.No.MethodsAdvantagesDisadvantages
1.Microemulsion
  • Thermodynamically stable
  • Exhibit low viscosity compared to emulsions
  • Do not require any energy contribution
  • Limited solubility for compounds that possess high melting points
  • Requires large amount of surfactants
2.Gas-phase
  • Can form binary and ternary NPs
  • Wide variety of substrates can be used
  • Uniform size distribution
  • Low production rate
  • Temperature-sensitive
3.Stöber’s
  • Simple, cost-effective, and efficient
  • Functionalization with thick coating can be conducted to avoid corrosion
  • High possibility of Sinteration at low temperature
  • Prolonged processing time
  • Organic solvents usage may lead to cytotoxicity
  • Toxic byproducts
4.Precipitation
  • Simple and effective
  • Homogenous yield
  • Requires further purification
  • Less reproducible
5.Green Synthesis
  • Cost-effective
  • Nontoxic and eco-friendly
  • High reproducibility
  • Size distribution
  • Requires more processing time
  • Low yield
Table 2. Summary of relevant theranostic applications of silica nanoparticles.
Table 2. Summary of relevant theranostic applications of silica nanoparticles.
Sl.NoType of SilicaModificationDisease/DisorderRef.
1.Biodegradable mesoporous silica nanoparticlesEncapsulated vaccinesSARS-CoV-2[90]
2.Hollow mesoporous silica nanoparticles with extra-large mesoporesCore-shell with Poly(ethylenimine) coating (PEI)Malignancy[91]
3.Mesoporous Silica NanoparticlesCarboxylic acidsTuberculosis [92]
4.Silica nanoparticlesMincle agonists tuberculosis[93]
5.Dendrimer-like mesoporous silica nanoparticlesFoot-and-mouth disease VLPs (virus-like particles)Foot-and-mouth disease[94]
6.Rambutan-Like Mesoporous Silica NanoparticlesDNA vaccine with PEI coatingChronic infections and cancers.[95]
7.Extra-large pore MSNs (XL-MSNs)Cancer antigen with Amine modificationMalignancy[96]
8.Mesoporous SBA-16 and silanized SBA-16 (APTES-SBA-16) nanoparticles(3-Aminopropyl) triethoxysilane (APTES)Paracoccidioidomycosis[97]
9.MCM-41 type silica nanoparticlesPolymer and amineOral protein-based vaccine[98]
10.Spherical MSNsRecombinant EspA loaded Enterohemorrhagic Escherichia coli[99]
Table 3. Summary of important silica nanoparticles (SNPs) in theranostics.
Table 3. Summary of important silica nanoparticles (SNPs) in theranostics.
Types of SNPs/Synthetic RoutesSurface Chemistry/ModificationsBiomedical ApplicationsRef.
Fractal SNP/surfactant-free Stöber method Fractal silica nanoparticles Sharp edge and rough surface with enhanced adhesion towards enzyme[18]
Solid SNP/Stöber methodAmine group/ polyethylenimine Mimicking virus surface topology and surface roughness enhances the binding of biomolecules [53]
MSN/co-condensation -Ruthenium polypyridyl delivery for cancer treatment [54]
MSNAzobenzene-modified Lubrication enhancement and drug release for osteoarthritis [55]
MSNCarboxyl-functionalized Controlled delivery of NSAIDs[56]
Virus like hollow MSN/self-consuming perovskite Electrostatic adsorption of doxorubicin Combination of chemotherapy and immunotherapy[57]
MSN/co-condensation Covalently bound dipalmitoylControlled release of fluorescein [61]
Core-shell hybrid SNPs/sol-gel process following modified Stöber methodSilica core with xylan linked 5-(4-hydroxyphenyl)-10,15,20-triphenylporphyrin (TPPOH)PDT against colorectal cancer cells [63]
Hollow-MSN/Stöber methodHyaluronic acidCancer chemo-PDT[64]
MSN/ modified Stöber methodhaloBODIPYs, PEG and FATargeted cancer PDT [65]
Silica nanospheres/sol-gel methodIndocyanine green to amino modified surfacePTT against drug resistant tumors[67]
Janus-structured gold triangle-MSN/ sol-gel methodAmino functionalization and attachment of tirapazamine (TPZ)Extrinsic radiosensitization, local PTT and hypoxia-specific chemotherapy[68]
FA-Gd-Tb@SiO2/covalent conjugationCovalent conjugation of luminescent Tb3+ to Si-O-Si framework followed by attachment of Gd and FA (folic acid)Targeted cellular time-gated luminescence (TGL) and cancer cell MR imaging[70]
Janus magnus nano-bullets (Mn-DTPA-F-MSNs)/sol-gel process Mn-DTPA-functionalized Fe3O4-MSNsGSH (glutathione) responsive T1/T2 MRI[71]
Exosomes-like cup-shaped SNP/emulsion template method-Ultra sound contrast for stem cell imaging and drug delivery[72]
Theranostic MSNAmino-silane conjugation of fluorescein followed by Gd-DOTAUltra sound and MR imaging of stem cells and slow-release reservoir of insulin-like growth factor (IGF) [73]
MSN/sol-gel methodSalicylic acid and ketoconazoleAntifungal and wound healing[103]
SNPCarbon quantum dots/silica nanoparticles/silk fibroin nanocomposites Wound repair [104]
Rod-like silica nanoparticles/one pot method in presence of PVPType I collagen-SiO2@Fe3O4Cell guidance and drug delivery[106]
Table 4. Observed toxicological effects of SNPs.
Table 4. Observed toxicological effects of SNPs.
Types of ParticlesMode of StudyProbable ReasonsRef.
MSN (20–200 nm) In vitroSize and dose dependency, ROS generation, and changes in membrane integrity induced by cellular uptake. [109]
SNP (62 nm)In vitroInduced ROS formation and DNA damage response and caused toxicity to endothelial cells through Chk1-dependant G2/M DNA damage checkpoint. [110]
Fluorescent MSNIn vivoMore than 80% intravenously administered MSNs deposited in liver, spleen, and lung and shape of the particles play a major role. [113]
SNPs (46 ± 4.9 & 432.0 ± 18.7 nm) and MSNs (466.0 ± 86.0 nm)In vivo & ex vivoMicroscopic lesions in liver, kidney, spleen, and lungs. Physicochemical properties, dose, and frequency of administration is responsible. [114]
SNPIn vivoDose-dependent hepatotoxicity and nephrotoxicity in male rats due to oxidative stress and apoptosis. [115]
MSN (75 nm)In vivoInduced intestinal oxidative stress and colonic epithelial cell apoptosis in mice[116]
SNP (57.66 ± 7.30 nm)In vitroImpairment in mitochondrial dynamics and biogenesis leading mitochondrial dysfunction, oxidative stress, and, finally, cardiovascular disease. [112]
SNP (64.43 ± 10.50)In vivoInduce hepatic dysfunction and granuloma formation in liver.[117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pallavi, P.; Harini, K.; Alshehri, S.; Ghoneim, M.M.; Alshlowi, A.; Gowtham, P.; Girigoswami, K.; Shakeel, F.; Girigoswami, A. From Synthetic Route of Silica Nanoparticles to Theranostic Applications. Processes 2022, 10, 2595. https://doi.org/10.3390/pr10122595

AMA Style

Pallavi P, Harini K, Alshehri S, Ghoneim MM, Alshlowi A, Gowtham P, Girigoswami K, Shakeel F, Girigoswami A. From Synthetic Route of Silica Nanoparticles to Theranostic Applications. Processes. 2022; 10(12):2595. https://doi.org/10.3390/pr10122595

Chicago/Turabian Style

Pallavi, Pragya, Karthick Harini, Sultan Alshehri, Mohammed M. Ghoneim, Areej Alshlowi, Pemula Gowtham, Koyeli Girigoswami, Faiyaz Shakeel, and Agnishwar Girigoswami. 2022. "From Synthetic Route of Silica Nanoparticles to Theranostic Applications" Processes 10, no. 12: 2595. https://doi.org/10.3390/pr10122595

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

Article Metrics

Back to TopTop