Next Article in Journal
Synthesis of Spherical Nanoparticle Hybrids via Aerosol Thiol-Ene Photopolymerization and Their Bioconjugation
Next Article in Special Issue
Nanomaterial-Based Therapy for Wound Healing
Previous Article in Journal
Field Emission of Multi-Walled Carbon Nanotubes from Pt-Assisted Chemical Vapor Deposition
Previous Article in Special Issue
Bacterial Cellulose (BC) and BC Composites: Production and Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring Various Techniques for the Chemical and Biological Synthesis of Polymeric Nanoparticles

Ecobiomaterial Research Laboratory, School of Biological Sciences, Universiti Sains Malaysia, Gelugor 11800, Penang, Malaysia
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(3), 576; https://doi.org/10.3390/nano12030576
Submission received: 30 December 2021 / Revised: 31 January 2022 / Accepted: 6 February 2022 / Published: 8 February 2022
(This article belongs to the Special Issue Nanomaterials for Advanced Biological Applications)

Abstract

:
Nanoparticles (NPs) have remarkable properties for delivering therapeutic drugs to the body’s targeted cells. NPs have shown to be significantly more efficient as drug delivery carriers than micron-sized particles, which are quickly eliminated by the immune system. Biopolymer-based polymeric nanoparticles (PNPs) are colloidal systems composed of either natural or synthetic polymers and can be synthesized by the direct polymerization of monomers (e.g., emulsion polymerization, surfactant-free emulsion polymerization, mini-emulsion polymerization, micro-emulsion polymerization, and microbial polymerization) or by the dispersion of preformed polymers (e.g., nanoprecipitation, emulsification solvent evaporation, emulsification solvent diffusion, and salting-out). The desired characteristics of NPs and their target applications are determining factors in the choice of method used for their production. This review article aims to shed light on the different methods employed for the production of PNPs and to discuss the effect of experimental parameters on the physicochemical properties of PNPs. Thus, this review highlights specific properties of PNPs that can be tailored to be employed as drug carriers, especially in hospitals for point-of-care diagnostics for targeted therapies.

1. Introduction

Nanoparticles (NPs) are defined as particles with all three dimensions confined within the range of 1 to 100 nm [1,2,3,4]. The growing attention towards NPs stems from the fact that their mechanical, chemical, optical, electrical, and magnetic properties differ from those of their bulk counterparts, and these properties can be altered by varying the size of NPs [5,6]. NPs are of great interest in a variety of sectors, including physics, agriculture, chemistry, engineering, electronics, biology, food technology, medicine, and bioengineering, due to their small size and ability to tailor their properties for specific requirements [7,8,9,10,11,12,13,14,15,16,17].
NPs offer the perfect characteristics for delivering therapeutic medications to the body’s target sites [18]. In contrast to micron-sized particles that are rapidly eliminated by the immune system, NPs demonstrated much higher efficiency as drug delivery carriers [19,20,21]. Because of their larger surface area, NPs can effectively penetrate cells and traverse the blood–brain barrier and they are easily destroyed [22,23,24]. NPs can be produced using a variety of natural and synthetic materials, which are biodegradable or non-biodegradable [25]. Examples of NPs include solid–lipid nanoparticles, silver nanoparticles, gold nanoparticles, magnetic nanoparticles, mesoporous silica nanoparticles, nanocrystals, carbon nanotubes, albumin nanoparticles, fullerene nanoparticles, and polymeric nanoparticles (PNPs).
Many types of NPs have been investigated for clinical use but have not been accepted widely due to their toxicity to some extent [26]. Biopolymers are employed in the manufacturing of NPs for biomedical applications to avoid cytotoxicity concerns [27,28]. Biopolymers are well-known for being non-toxic, biodegradable, and biocompatible [29,30]. Depending on the intended uses, PNPs can be simply and cost-effectively generated on a wide scale using a variety of technologies. PNPs have applications in different fields such as electronics [31], photonics [32], environmental technology [33], medicine [34], bio-imaging [35], diagnostics [36], biotechnology [37], biomedical drug delivery [38,39,40], and energy harvesting [41].
Due to their subcellular size, biodegradability, biocompatibility with tissue and cells, and controlled and sustained-release capabilities, PNPs are attractive candidates for the delivery of vaccinations, antibiotics, and cancer treatments [42,43,44,45,46]. PNPs can enhance the bioavailability, solubility, and retention time of drugs. Moreover, PNPs do not cause any toxic, inflammatory, or immunogenic side effects [47,48]. Different polymers such as polyhydroxyalkanoate (PHA) [49,50,51,52], polylactic acid (PLA) [53,54,55], poly(lactic-co-glycolic acid) (PLGA) [56,57], polycaprolactone (PCL) [58,59,60], polyglycolide (PGA) [61], polyanhydride [62], polycyanoacrylate [63], poly glutamic acid [64], polymalic acid [65,66], poly(N-vinyl pyrrolidone) [66,67], poly(methyl methacrylate) (PMMA) [68,69], poly(vinyl alcohol) [70,71], poly(acrylic acid) [72,73], poly acrylamide [74,75], and poly(methacrylic acid) [76,77] have been used for the synthesis of PNPs.
This review paper describes the different methods used for producing PNPs and how variation in experimental parameters can enable the control of NP properties. As PNPs are colloidal systems made up of natural or synthetic polymers, their synthesis methods are generally categorized into two groups. They are (1) the direct polymerization of monomers (emulsion polymerization, surfactant-free emulsion polymerization, mini-emulsion polymerization, micro-emulsion polymerization, and microbial polymerization) and (2) the dispersion of preformed polymers (e.g., nanoprecipitation, emulsification solvent evaporation, emulsification solvent diffusion, and salting-out). Table 1 describes the advantages and limitations of these two types of polymer synthesis methods.

2. Methods for Producing PNPs

PNP preparation can be divided into two categories: monomer polymerization and preformed polymer dispersion [96,97,98]. Emulsion polymerization, surfactant-free emulsion polymerization, mini-emulsion polymerization, and micro-emulsion polymerization are all processes that can be used to polymerize monomers [99,100]. Likewise, nanoprecipitation, emulsification solvent evaporation, emulsification solvent diffusion, and salting-out can all be utilized to make PNPs from preformed polymers [101,102,103]. The type of polymer, size requirement, and application region all influence the method of preparation [104,105]. The technique of preparation is crucial to achieving the desired qualities. PNPs made for biological applications, for example, should be free of additives and reactants [106].
The type of polymer used determines the features of the produced NPs that are designed for a certain purpose [107,108]. The drug delivery capabilities of PLGA and poly(3-hydroxybutyrate) P(3HB) were studied by employing docetaxel (DTXL). Although the toxicity profiles of P(3HB) and PLGA were similar, P(3HB) had a nearly two-fold higher loading efficacy and poorer retention rates than PLGA [109]. Dissolution, solubility, cellular uptake, release of drugs, bio-distribution, and circulatory half-life are all influenced by the size of PNPs [110,111,112]. The challenge in the preparation of PNPs is the ability to produce uniform particles to have consistent performance [113,114]. NPs with a broad size distribution result in difficulty in establishing their applications [115].

2.1. Formation of NPs from Preformed Polymers

This section discusses the many ways to make PNPs from pre-formed polymers, including nanoprecipitation, emulsification solvent evaporation, emulsification solvent diffusion, and salting-out [101,102,103]. The initial stage in all of these approaches is to prepare an emulsification system, which is the same for all of them. The second step is the formation of PNP, which is different for each method. The name of the method is conferred by the principles of the second step, which can occur either by precipitation or by the evaporation of the organic solvent [116,117].

2.1.1. Nanoprecipitation

Fessi et al. devised the nanoprecipitation approach, often known as the solvent displacement, antisolvent precipitation, solvent shifting, and desolvation methods, for the creation of PNPs in 1989 [118]. Nanoprecipitation is a simple, easy, fast, and reproducible single-step method. This approach does not demand a lot of energy and can be scaled up simply [119]. Nanoprecipitation is time-efficient, inexpensive, and does not need a precursor emulsion like other methods [120]. The size of the NPs generated by this approach is changed by altering the parameters, and they are small with a limited size distribution [121]. Nanoprecipitation is based on interfacial deposition, in which the transport of a solvent into a non-solvent causes the polymer to dissolve, leading to nuclei growth, crystal growth, and nanoprecipitation [122,123,124].
An organic phase is introduced to the aqueous phase during nanoprecipitation. The polymer and water-miscible organic solvent, which must be miscible in the aqueous medium, make up the organic phase, which has a diffusion effect [125,126,127,128,129,130,131]. To slow aggregation, the polymer must be insoluble in the aqueous solution, which might contain a stabilizer like a surfactant [132,133,134,135]. Dropwise addition of the organic phase to the aqueous phase with moderate agitation produces NPs [136,137]. Ultracentrifugation is used to collect the NPs, which are subsequently rinsed with water to remove the surfactant. The organic solvent evaporates, hardening the NPs, which are subsequently recovered by filtering, spinning, or freeze-drying [138,139]. Organic solvents that evaporate easily such as ethanol, acetone, hexane, or methylene chloride should be chosen as a polymer solvent. Binary solvent blends such as combinations of acetone with either ethanol or methanol can also be used. Likewise, a mixture of non-solvents can be used to form NPs in this method [140,141,142]. Figure 1 shows a schematic illustration of the nanoprecipitation process. According to Quintanar et al., the difference in surface tensions induces intrafacial turbulence and thermal disparities in the system, resulting in the production of continuous solvent eddies at the interface of both liquids. When the polymer aggregates on the hydrophobic drug surface as the solvent runs away (solvent diffusion) from low surface tension regions, nanocapsules are generated. [97].
The polymer content, variety of solvent and non-solvent, proportion of solvent to non-solvent, rate of the addition of solvent to non-solvent, the effect of the stabilizer, and stirring speed all influence the size of NPs [143,144]. Due to the increase in viscosity that hinders polymer diffusion from the solvent to the non-solvent, an increase in polymer concentration leads to the creation of bigger nanoparticles [144]. Smaller NPs in a narrow size range are produced by solvents with high diffusion coefficients, such as acetone and acetonitrile [145]. It has also been established that a decrease in the solvent-to-non-solvent-volume ratio results in smaller NP sizes [146]. The nature of the stabilizer and its concentration has been shown to influence the size of NPs [147,148].
A study found that increasing the amount of surfactant (e.g., Pluronic) reduced the size of NPs by lowering interfacial tension [149]. In addition to that, employing a surfactant in the nanoprecipitation method is not necessary, enabling the production of surfactant-free particles [150]. Meanwhile, higher stirring rates have been found to produce smaller NPs [151,152]. Zhang and colleagues demonstrated that raising the stirring speed from 300 to 1200 rpm reduces particle diameter from 800 to 300 nm [153]. Specifically, low external energy input is sufficient for the nanoprecipitation method, hence a moderate stirring speed is required instead of a high stirring speed that raises the temperature [154,155,156].
NP formation using the nanoprecipitation method occurs through three different steps; particle nucleation, molecular growth and particle agglomeration [157]. Nucleation takes place when the polymer concentration reaches the saturation level, i.e., when the polymer solute in the solution is more than the amount that the solvent can dissolve [158]. The mean particle size increased significantly as the polymer concentration was increased [159,160]. Molecular growth and particle agglomeration occur with a release of energy [161,162].
Chorny and coworkers used the nanoprecipitation approach to make PLA NPs loaded with tyrphostin. Particle size increases from 70 nm to 140 nm when the polymer concentration is increased from 100 mg (5 mg/mL) to 300 mg (15 mg/mL) [163]. NPs of poly(lactide)-poly(ethylene glycol)-poly(lactide) (PLA-PEG-PLA) were synthesized by the nanoprecipitation method under different conditions. It was discovered that increasing the agitation rate resulted in a reduction in particle size [164]. Meanwhile, in another study comparing two methods for NP preparation, the nanoprecipitation method was found to be more efficient for preparing PLGA NPs encapsulating cucurbitacin compared to using the emulsion solvent evaporation method [165].
In the preparation of cellulose NPs loaded with mefenamic acid [166] and PLGA NPs loaded with N-acetylcysteine (NAC), the solvent/nonsolvent ratio, the concentration of polymer and the choice of solvent as well as nonsolvent were found to affect the size of NPs. The nanoprecipitation method, in addition to efficiently entrapping hydrophobic molecules, also has a great potential as an alternate entrapment method for hydrophilic chemicals, according to the findings [167]. Chidambaram et al. proposed changes to the traditional nanoprecipitation process in order to reduce NP size and create NPs with a narrow size distribution. They used sonication to prepare both the organic and aqueous phases, yielding Eudragit E100 NPs with a particles size of 114 nm and a uniformity of 0.259 [168].
Three distinct proteins (tetanus toxoid, lysozyme, and insulin) were entrapped in poly(d,l-lactic acid) and poly(d,l-lactic-co-glycolic acid) NPs using modified nanoprecipitation and double emulsion (w1/o/w2) techniques in a separate investigation. The use of miscible organic solvents like dimethylsulfoxide (DMSO) rather than conventional organic solvents like acetone or ethanol, as well as non-solvents like methanol or ethanol rather than water, have all been added to the nanoprecipitation process. Nanoprecipitation proved to be a suitable option to the extensively employed double emulsion approach. Nanoprecipitation was found to be the best approach for protein trapping in small, densely loaded NPs [169]. Luo et al. applied a combination of electrospraying and nanoprecipitation to produce multifunctional superhydrophobic polymethylsilsesquioxane (PMSQ) NPs with sizes smaller than 100 nm [170].
Additionally, continuous flow microfluidics is a great solution for nanoprecipitation operations, enhancing product controllability, homogeneity, and reproducibility. Nanoprecipitation through a hydrodynamic flow-focusing microchannel was used to synthesize PLGA-poly(ethylene glycol) nanoparticles (PLGA-PEG NPs). Variations in flow rates, polymer concentration, and polymer composition can be used to obtain the preferred size, drug loading, and polydispersity of the synthesized product [171]. Polycaprolactone (PCL) nanoparticles, which are biodegradable and have a tremendous potential for controlled drug delivery, were synthesized through a similar nanoprecipitation process [172]. Moreover, this technique may be used to assemble other polymers like chitosan, heparin, and hyaluronic acid in microfluidic devices, especially to produce PNPs for controlled release as well as drug delivery [173].
Meanwhile, P(3HB) NPs were prepared by nanoprecipitation with a variety of solvent/non-solvent combinations such as ethyl acetate:DMSO, chloroform:water, chloroform:DMSO, and ethyl acetate:water. In the reported study, spherically shaped P(3HB) NPs with sizes ranging from 40 to 100 nm were successfully formed while the size of loaded PNPs were typically between 200 to 600 nm as shown in Figure 2 [174]. In another attempt, P(3HB) NPs were prepared by nanoprecipitation with a low concentration of Tween 80 as a surfactant. The size and size distribution of NPs decreased as the amount of Tween 80 in water increased to 1% (v/v) [175].
In all the examples mentioned above, P(3HB) is initially biosynthesized by microorganisms and stored in the microbial cell cytoplasm. The produced and accumulated natural polyester is then removed from the bacterial cells using suitable solvents like chloroform and purified by reprecipitation in a non-solvent like methanol. The purified P(3HB) can be mixed in solvents and used in the nanoprecipitation process to make NPs.

2.1.2. Emulsification-Solvent Evaporation

The first and most extensively used method for the synthesis of PNPs is emulsification-solvent evaporation. The first step involves emulsifying the polymer solution into an aqueous phase, and the second step entails the evaporation of the solvent, which results in polymer precipitation, resulting in the production of NP [176,177,178]. The first step is to form the emulsions, which can occur by either of two main strategies. The first strategy is to produce single-emulsions, i.e., oil-in-water (o/w) and the second one is to produce double-emulsions, i.e., water-oil-water (w/o/w) or oil-water-oil (o/w/o) [179,180]. In a double emulsion, the primary emulsion (w1/o) is first prepared by dispersing the aqueous phase in an immiscible organic solvent containing the polymer. Subsequently, the primary emulsion is homogenized in an outer aqueous phase containing the emulsifier using a high-shear homogenizer to form the organic phase and then emulsified in the aqueous phase containing a surfactant [181,182,183,184,185,186].
The solvent is then continuously evaporated while the NPs are recovered by ultracentrifugation [187,188]. The NPs are thoroughly rinsed with water and then lyophilized to remove the surfactants [189,190,191]. Figure 3 depicts a schematic illustration of the solvent evaporation procedure. The diameter of NPs can be controlled by adjusting the stirring speed, the viscosity of the aqueous and organic phases, and the type and concentration of the dispersing agent [192]. The solvent evaporation approach was used by Musyanovych et al. to make poly(l-lactide) (PLLA), PLGA, and poly(caprolactone) (PCL) NPs. The size of the NPs produced is affected by the type of polymer used. The smallest particle size was found in PLGA NPs, whereas the highest particle size was found in PCL NPs [192].
Another study generated haloperidol-loaded PLGA/PLA NPs and found that raising the polymer concentration from 5 to 66.6 mg/mL improved the NP size from 200 to 300 nm while retaining a unimodal particle-size distribution. It was discovered that lowering the solvent/non-solvent volume ratio reduced the size of PLGA/PLA NPs [193]. Bilati et al. examined at how the sonication procedure affected the properties of poly(lactide-co-glycolide) nanocapsules made by the water-in-oil/water solvent evaporation method. The second mixing step’s sonication time (for w/o/w emulsion) has a bigger impact on the final NP size than the first step’s sonication duration (for water-in-oil emulsion). [194].
Poly(d,l-lactide-co-glycolide) NPs containing praziquantel were produced by employing methylene chloride or ethyl acetate, separately, as an organic solvent in the dispersion phase. The size of methylene chloride-prepared NPs was larger than that of ethyl acetate-prepared NPs [195]. When ethanol was used as a solvent and Pluronic F-108 was used as a stabilizer, the average size of poly(ethylene oxide) (PEO) NPs generated using the single-emulsion approach was 100 to 150 nm. It was evident that the polymer concentration influenced the characteristics of the PEO NPs [196].
By modifying experimental conditions such as homogenization rate, surfactant concentration, and polymer/solvent ratio, poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) [P(3HB-co-3HHx)] NPs could be produced in the size range of 180 nm to 1.5 µm. The size of P(3HB-co-3HHx) NPs decreased as the surfactant concentration and homogenization rate increased, whereas P(3HB-co-3HHx) NP size increased by increasing the polymer to solvent ratio [197]. When the ultrasound exposure period, amplitude, and exterior aqueous phase volume were increased, PCL NPs generated using the double emulsion solvent evaporation method showed a decrease in particle size. The size of NPs grew from 235 to 748 nm when the concentration of PCL was raised from 1 to 5 g. Meanwhile, the size of PCL NPs decreased with increasing surfactant (e.g., PVA) concentration from 0.05 to 0.2% [198].
Folate-targeted poly(3-hydroxybutyrate-co-3-hydroxyoctanoate) P(3HB-co-3HO) NPs were prepared by the w1/o/w2 solvent evaporation method. These NPs were loaded with doxorubicin (DOX), a chemotherapeutic drug in cancer treatment. An in vivo antitumor study of the NPs revealed a great potential of these NPs to improve the sustained release profile of doxorubicin [199]. This approach produced PEG end-capped P(3HB-co-3HHx) with a particle size of roughly 200 nm, which showed promise as a nanocarrier for sustained rapamycin delivery with increased cellular absorption and kinase inhibitory efficacy [200]. In addition to the P(3HB-co-3HO) NPs, poly(3-hydroxyvalerate-co-4-hydroxybutyrate) P(3HV-co-4HB) NPs could also be synthesized by the emulsification–solvent evaporation method. It was found that the cisplatin-loaded NPs accumulated more efficiently in the tumor cells and had a higher tumor regression effect than freely administered cisplatin, indicating that this nanocarrier was suitable for drug delivery applications [201]. Curcumin was loaded into the P(3HB-co-3HHx) NPs for use in breast cancer treatment. Higher drug release and better decline in tumor cell activity were observed in curcumin-loaded P(3HB-co-3HHx) NPs than curcumin alone, indicating that the P(3HB-co-3HHx) NPs are a promising tool to enable the sustained and controlled release of some drugs [202].
As a nanocarrier for ellipticine, poly(3-hydroxybutyrate-co-3-hydroxyvalerate) P(3HB-co-3HV) NPs were produced by solvent evaporation (EPT). In an in vitro test, the percentage of inhibition for EPT-PHBV NPs was around two times that of free EPT, showing that P(3HB-co-3HV) NPs are a viable vehicle for the administration of hydrophobic medicines for cancer treatment [203]. For cisplatin delivery, poly(4-hydroxybutyrate)-mPEG (P(4HB)-mPEG nanocarriers were developed. The cisplatin-loaded P(4HB)-mPEG NPs were shown to be more effective than free cisplatin, demonstrating that the P(4HB)-mPEG) nanocarriers are effective in delivering cisplatin to cancer cells [204].

2.1.3. Emulsification Solvent Diffusion

Leroux et al. were the first to propose the emulsification-solvent diffusion approach. To start, the polymer is dissolved in an organic solvent that is saturated with water, generating an organic phase. The organic phase is then emulsified in the aqueous solution, resulting in solvent diffusion and NP production [78,135]. To precipitate the polymer, it is necessary to dilute the solvent with extra water to improve its diffusion. Lastly, the solvent is eliminated by distillation or crossflow filtration [205,206,207,208]. The aqueous phase contains a stabilizer, and the dilution phase is often water. This process has the benefit of not necessitating a homogenizer, having excellent reproducibility, and being simple to scale up [209,210]. The drawback of this procedure is that it requires a large amount of water to be eliminated from the suspension [211]. Figure 4 shows a schematic illustration of the emulsification solvent diffusion process.
Quintanar et al. proposed a mechanism for NP formation in which each droplet forms several NPs [209]. Perez et al. and Ma et al. then proceeded to modify the method suggested by Quintanar et al. for the nanoencapsulation of hydrophilic active substances. In their proposed method, the aqueous inner phase includes an active substance as well as a stabilizing agent such as PVA or poly(vinylpyrrolidone) (PVP), while the external phase comprises the polymer and organic solvent. The emulsion was initially diluted with the solvent (ethanol), resulting in organic solvent migration. Then, water was added to facilitate the collection of NPs [212,213]. Hassou and Moinard-Chécot et al. used a step-by-step diffusion analysis using the stopped-flow methodology to represent different states that occur in the emulsification solvent diffusion method during the dilution stage. It was discovered that the solvent diffuses quickly from the droplets, taking less than 20 ms [214,215]. Pramual et al. formulated 5,10,15,20-Tetrakis(4-hydroxy-phenyl)-21H, 23H-porphine pTHPP (hydrophobic photosensitizer) loaded P(3HB-co-3HV) NPs for photodynamic therapy (PDT) by the emulsification-diffusion method. The size distribution of P(3HB-co-3HV) NPs was narrow, ranging from 169.0 to 211.2 nm. The pTHPP-loaded P(3HB-co-3HV) NPs exhibited high photocytotoxicity towards HT-29 human colon cancer cells compared to pTHPP alone. These results indicated that the P(3HB-co-3HV) NPs are potential vehicles for the delivery of hydrophobic photosensitizer drugs in photodynamic therapy [216]. PHA NPs encapsulating TGX221 anti-cancer drugs were also developed. TGX221 was slowly liberated from the PHA-based NP and proliferation in NP-TGX221-treated cells was considerably slower than in cells receiving free TGX221 [217].
Using a modified emulsification solvent diffusion process, Chen et al. developed curcumin-loaded PLGA (PLGA-Cur) NPs with a mean range of 190 nm. Anti-tumor activity was successfully detected following the delivery of PLGA-Cur NPs into cells, and in comparison, with free curcumin, PLGA-Cur NPs demonstrated the increased inhibition of HL60 and HepG2 cancer cells with lower IC50 values. Moreover, confocal microscopy analysis showed that the curcumin-loaded PLGA NPs increased apoptosis in cancer cells when compared with free curcumin [218]. PCL NPs were made using ethyl acetate as the solvent and PVA as the stabilizing agent, respectively. The polymer concentration, solvent volume, type and amount of the surfactant, as well as the concentration of oil in the organic phase, were all observed to affect the size of PCL NPs [219]. The solvent and stabilizing agents utilized to make PLA NPs were ethyl acetate and Pluronic F68, respectively. Particle size increased from 260 to 530 nm as PLA content increased [220]. In another investigation, as the amount of surfactant was raised, the size of PCL NPs shrank [221]. A comparison made using different stabilizers, di-dodecyl dimethylammonium bromide (DMAB) and PVA, for the production of PLGA NPs revealed that DMAB produced smaller PLGA NPs [222]. The influence of homogenization and sonication on the size of PLGA NPs was explored by Jain et al., who discovered that sonication resulted in smaller particles with an average size of 165 nm, whereas homogenization resulted in particles with an average size of 225 nm [223].

2.1.4. Salting-Out Technique

The salting-out method is a variation of the emulsification-solvent diffusion method that makes use of the salting-out effect [224,225]. The difference between the salting-out method and the emulsion diffusion method is that the former method does not require a solvent diffusion step, due to the existence of salts [226,227]. The aqueous phase consists of water, stabilizer, and salting-out agents. Electrolytes including sodium chloride, magnesium chloride, calcium chloride, and magnesium acetate, as well as non-electrolytes like sucrose, are salting-out agents [228,229,230]. The salting-out agents should be insoluble in the organic solvent. The kind of salting-out agent used has a big impact on how well drugs are encapsulated [231,232].
The organic phase, which contains the polymer in a water-miscible organic solvent, is introduced to the aqueous phase in this process. The emulsion is then diluted with adequate amounts of water while constantly swirling to reduce the electrolyte’s ionic strength and improve solvent diffusion [233,234]. The generation of NPs is caused by the migration of the solvent from the organic phase to the aqueous phase during dilution. Lastly, crossflow filtering is used to remove the salting-out agent, and the produced NPs are collected [85,235,236]. The lack of chlorinated solvents, which are hazardous to the physiological system, is an advantage of the salting-out procedure. The use of salt in the preparation process necessitates purifying processes, which is a downside of this method [84,85]. A schematic representation of the salting-out technique is shown in Figure 5.
Poly(trimethylene carbonate) (PTMC) NPs were produced by single-emulsion and salting-out methods. The PTMC NPs formed by the salting-out method were smaller than those formed by the single-emulsion method. In the single-emulsion approach, the effect of polymer concentration and stirring speed on NP size was more pronounced. Another difference between these two processes is the type of organic solvents utilized; in the salting-out approach, water-miscible THF was used, whereas in the single-emulsion method, water-immiscible dichloromethane (DCM) was used [237]. Salting-out, emulsification-diffusion, and nanoprecipitation procedures were used to make methacrylic acid copolymer NPs. The size range of the methacrylic acid copolymer NPs was broader for the salting-out method (123–710 nm) compared to the emulsification-diffusion method (108–715 nm) and the nanoprecipitation method (147–245 nm) [238]. Ethylcellulose (EC) and Eudragit-S-100 (ED) NPs produced by the salting-out technique had a mean particle size and zeta potential value of 211 nm and −43.7 mV, respectively [239].
In another work, sodium chloride was used as the salting-out agent rather than magnesium chloride or magnesium acetate to make PLGA NPs. The NPs generated were spherical, measuring 111.4 ± 2.35 nm in diameter and having a modest polydispersity (0.062 ± 0.023) [240]. Meanwhile, tetrahydrofuran (THF) was used to make PLA NPs with a diameter of less than 200 nm [241], while paracetamol-loaded Eudragit S100 NPs were produced using ethanol as solvent, sodium carboxymethylcellulose as a stabilizer, and zinc sulfate heptahydrate (ZnSO4·7H2O) as the salting-out agent [242]. Zweers et al. used acetone and magnesium chloride hexahydrate (MgCl2·6H2O) as the organic solvent and salting-out agent, respectively, to make PEO-PLGA NPs with a size of about 200 nm [243,244]. Similarly, PLA end-capped with 1- pyrenebutanol (PLAP) NPs were synthesized using MgCl2·6H2O as the salting-out agent [245].

2.2. Formation of Nanoparticles by Polymerization of Monomers

The methods that were explained in the previous sections are used to produce PNPs from preformed polymers. PNPs can also be produced by the polymerization of monomers. This section explores the methods employed for the polymerization of monomers such as emulsion polymerization, surfactant-free emulsion polymerization, mini-emulsion polymerization and micro-emulsion polymerization.

2.2.1. Emulsion Polymerization

Emulsion polymerization is one of the most commonly used, fastest, and readily scalable method for producing PNPs. The method of emulsion polymerization can be divided into two groups, depending on whether the continuous phase is organic or aqueous. The dispersion of monomer into an emulsion or a substance in which the monomer is not soluble is part of the continuous organic phase technique. The monomer is dissolved in a continuous phase, which is commonly an aqueous solution, without the use of surfactants or emulsifiers in the aqueous continuous phase [211,246]. Surfactant-based emulsion polymerization can be divided into two types: conventional and surfactant-free emulsion polymerization [86,247]. A water-soluble initiator, water, a somewhat water-soluble monomer, and surfactant are utilized in the traditional approach. Water is an environmentally friendly dispersion medium that also helps to dissipate heat during polymerization [107,108,248]. When a monomer dissolves in the continuous phase, initiation takes place with an initiator molecule. The initiating agent forms monomeric radicals that interact with the monomer and initiate the reaction. The radicals propagate until they reach a critical chain length, at which point their aqueous solubility is reduced. At this stage most of the surfactants are engaged in stabilization process and have formed polymer particles. However, the polymerization reaction continues until no new particles are nucleated. The termination process occurs at the end stage when there is a decrease in the polymerization rate [249,250,251]. Figure 6 depicts a schematic illustration of the emulsion polymerization technique.
Other PNPs successfully produced by the same method include polystyrene-b-poly[poly(ethylene glycol) methyl ether methacrylate] (PS-b-P(PEGMA300), and PS-b-P(PEGMA1100)) PNPs [252]. Using SDS as a surfactant, Garay-Jimenez et al. synthesised polyacrylate NPs by the emulsion polymerization of acrylate compounds in a mixture of butyl acrylate and styrene. Anionic, cationic, zwitterionic, and noncharged (amphiphilic) surfactants were used to create poly(butyl acrylate-styrene) emulsions. The emulsions’ cytotoxicity and microbiological activity were compared before and after purification. The findings showed that attaching a polymerizable surface to the nanoparticle matrix has no effect on the emulsion’s cytotoxic or antibacterial effects, irrespective of whether the emulsion is purified or not and that the perfect properties are associated with using non-ionic surfactants rather than those with zwitterionic, cationic, or anionic [253].
To encapsulate magnetite particles and improve particle-size distribution, emulsion polymerization was used to create magnetic polymer matrix composite nanoparticles (MPCNPs) (PSD). Transmission electron microscopy (TEM) and vibrating sample magnetometry were used to characterize MPCNPs (VSM). The results showed that the emulsion polymerization approach was successful in encapsulating magnetite particles [254]. Under microwave radiation, styrene emulsion polymerization was carried out at 70 °C using sodium dodecyl sulphate (SDS) as a surfactant and potassium persulfate (KPS) as an initiator [255]. Another study used microwave irradiation to accomplish the emulsion polymerization of methyl methacrylate (MMA) and butyl acrylate (BuA) using potassium persulfate (K2S2O8) as an initiator and Disponil A3065 as an emulsifier [256]. The size of polystyrene NPs produced by the ultrasonic irradiation emulsion polymerization of styrene using polymeric carboxymethyl cellulose and alkyl poly(etheroxy) acrylate (CMCA9) as surfactant was 30 to 60 nm [257]. The size of PVK NPs is regulated by the concentration of VCz. Polyvinylcarbazole (PVK) NPs were generated via emulsion polymerization of N-vinylcarbazole (VCz) [258].

2.2.2. Surfactant-Free Emulsion Polymerization

Surfactants are utilized in the traditional emulsion polymerization procedure and should be eliminated from the finished product. Surfactant removal is a time-consuming operation that raises manufacturing costs [89,259]. An emulsion polymerization process without surfactants, i.e., a surfactant-free emulsion polymerization method, was devised to alleviate this limitation [260,261]. To generate PNPs, this method offers a straightforward, green approach that does not require the inclusion and subsequent removal of stabilizing chemicals. A water-soluble initiator (KPS, potassium persulfate), monomers, and water are the reagents utilized in this process. The stabilization of PNPs is achieved using ionizable initiators or ionic co-monomers in this technique [262,263,264,265].
The surfactant-free emulsion polymerization procedure using microwave irradiation produced PMMA NPs with a narrow size distribution. When the monomer concentration was increased from 0 to 0.3 mol/L, the size of the NPs rose from 103 to 215 nm [266]. The Cu2+/HSO3 redox initiation system was used to commence the surfactant-free emulsion polymerization of MMA and PMMA NPs with a negative charge in the size range of 165 to 223 nm were produced [267]. PMMA NPs in the size range of 200 to 600 nm were successfully developed using hydrophilic laponite clay to stabilize methyl methacrylate emulsions dispersed in distilled water [268], while NPs with a dimension less than 100 nm and high solid content were accomplishment of this project using KPS and acetone as initiator and co-solvent, respectively [269]. Using NaSS as a stabilizing agent and water as the reaction medium, poly-acrylate NPs with fluorine and silicon in the shell with a mean range of 172.5 nm were produced [270]. By the surfactant-free emulsion polymerization of styrene utilizing ultrasonic irradiation in the presence of potassium persulfate (KPS) as an anionic initiator and cetyl alcohol as a co-stabilizer, Faridi Majidi et al. produced polystyrene NPs in the size range of 200–250 nm [271]. Surfactant-free emulsion polymerization produced poly(hydroxyethyl methacrylate) (PHEMA) NPs with a mean size of 150 nm, a polydispersity index of 1.171, and a surface area of 17,779 m2/g [272]. Lee et al. used Fe3+ catalyzed emulsion polymerization to produce poly(styrene/thiophene) NPs with particle sizes ranging from 300 to 800 nm [273]. Polyimide NPs were synthesized in a continuous phase by heterophase polycondensation of various aromatic tetra-carboxylic acids and diamines in imidazolium-based ionic liquids (IL) [274]. Kim et al. synthesized polypyrrole NPs utilizing benzene octanol and ethyl acetate as continuous phases. Changing the water and octanol volume ratios led to fewer particles with an average size of 60 nm [275].
Colloidal NPs with a PMMA or poly(butyl methacrylate) core and a cationic polymer stabilizing shell were produced using reversible addition fragmentation chain transfer-mediated surfactant-free emulsion polymerization and had hydrodynamic diameters ranging from 32 to 96 nm. The wetting behaviour of such core-shell NPs, which can be fine-tuned depending on the internal nanostructure (soft or rigid core) and external temperature, allows for the creation of controllable functional hybrid colloidal arrays [276].

2.2.3. Mini-Emulsion Polymerization

The co-stabilizer, initiator, surfactant, monomer mixture, and water are all required components for the mini-emulsion process. The utilization of a low-molecular-mass co-stabilizer as well as a high-shear device such as ultrasound in this approach [91,277,278,279,280] is the fundamental distinction between mini-emulsion polymerization and emulsion polymerization. The sort of co-stabilizer and initiator used has a big impact on how the NPs develop and what they look like. Figure 7 shows a diagram illustration of the mini-emulsion polymerization process.
The mini-emulsion polymerization process was used to make a variety of PNPs. This approach produced polyacrylonitrile NPs in the size range of 100 to 180 nm using HD and SDS as the co-stabilizer and surfactant, respectively [281]. Similarly, PHEMA NPs were made with the surfactant Span 80 or KLE3729 and the co-stabilizer CH or HD. The nanoparticles produced were reported to be between 50 and 200 nm in size [282]. SDS/DMA and SDS/SMA were used, in another study, to stabilize mini-emulsion polymerizations of styrene [283]. Other examples include the composite colloidal NPs, made of magnetite as magnetic core and poly(ethyl-2-cyanoacrylate) as a polymeric shell [284]; polystyrene-single wall carbon nanotube (PS-SWNT) with SDS as surfactant and 1-pentanol as co-stabilizer [285]; and phosphonated polystyrene, as well as PMMA NPs (size range of 102 to 312 nm) produced by the free-radical copolymerization of vinylphosphonic acid (VPA) [286].

2.2.4. Micro-Emulsion Polymerization

A new method for manufacturing nanosized PNPs is micro-emulsion polymerization. Despite the fact that emulsion polymerization and micro-emulsion polymerization are similar processes that form polymers with high molar mass, their kinetics differ, resulting in micro-emulsion polymerization having smaller particle sizes and fewer chains per particle [287,288,289,290]. A water-soluble initiator is introduced to the aqueous phase, which contains a lot of surfactants, in the microemulsion polymerization technique. Because initiation cannot occur simultaneously in all microdroplets, polymer chains begin to form only in some of them. Due to osmotic and elastic forces, microdroplets will collapse later, resulting in larger particles and the development of empty micelles [291,292,293,294]. In a microemulsion, polymerization kinetics, PNP properties, and the concentration and type of initiator, surfactant and monomer are determining factors [293]. Some researchers have been carried out to see how these parameters affect the characteristics of NPs.
Micro-emulsion polymerization was used to create poly(vinyl acetate) lattices with a high total solid [93]. On the micro-emulsion polymerization of vinyl acetate stabilized with Aerosol OT (AOT), the effects of temperature, concentration and type of initiator (V-50 and KPS) were investigated. It was found that the reaction rates increased with the concentration of V-50 and temperature. Furthermore, the differences in electrostatic attraction between KPS and V-50 free radicals, as well as charged micro-emulsion droplets, resulted in quicker polymerization rates for KPS [295]. A cationic surfactant (e.g., CTAB) and a non-ionic surfactant were used to make poly(dimethylsiloxane) (PDMS) NPs in the range of sizes of 12–80 nm [296]. Furthermore, stabilizers of dodecyltrimethylammonium bromide (DTAB) and didodecyldimethylammonium bromide (DDAB) were used to make polyhexylmethacrylate NPs with a size range of 38 to 53 nm [297]. SABS-8 and SABS-10, two polymerizable anionic surfactants, were employed successfully in microemulsion polymerization of butyl methacrylate (BMA) at room temperature utilizing the redox initiator ammonium persulfate (APS)/tetramethylethylenediamine (TMEDA) [298]. Some other work used the cationic surfactant decyltrimethylammonium bromide DeTAB to make polypyrrole NPs with a particle size of 2 nm [299]. The polymerization of butyl acrylate with a sodium dodecyl sulfate/Aerosol OT surfactant combination and potassium peroxodisulfate as an initiator yielded particles smaller than 40 nm [300].

3. Biologically Synthesized Biodegradable Polyhydroxyalkanoate-Based Nanoparticles

A non-toxic, reliable, and eco-friendly experimental protocol for the synthesis of NPs is highly in demand. Natural entities such as secondary metabolites, enzymes, polysaccharides, biodegradable polymers, vitamins, and microorganisms can be utilized for the synthesis of NPs [95,301]. One such promising approach is the biosynthesis of NPs using bacteria. To date, a large variety of bacterial species have already been studied in the hopes of developing alternate NP synthesis techniques. For the time being, scientists are producing NPs using bacteria’s biomass or cell extracts [302]. In comparison to other biological entities, bacteria are thought to be a promising biofactory for the synthesis of NPs. Bacterial biosynthesis of NPs is a fast-growing study area in the field of science and nanotechnology, with many species of bacteria being used to synthesize NPs all around the world [303].
PHA is one of the PNPs generated spontaneously in the bacterial cytoplasm. PHA belongs to the aliphatic polyesters family of biodegradable and biocompatible polymers [304,305,306]. PHA is produced spontaneously by some bacteria in the form of nanosized granules under unbalanced growth conditions, such as an excess of carbon source and nutritional limitations, such as nitrogen, oxygen, and phosphorus [307,308,309]. Figure 8 depicts TEM images of nanosized PHA granules inside bacterial cells. Numerous parameters can affect the size of PHA granules such as PHA granule-associated proteins (phasins), bacterial species or strains, cultivation conditions, and time [310,311,312,313,314]. PHA granules made this way are tunable (for chemical composition, size, or other key qualities) to levels not possible with chemical synthesis, genetically altering the bacterial strain or adjusting the production circumstances such as the culture media composition. The crux of interest that lies in PHA NPs is their self-assembling properties, their production using easily cultivated bacterial species, and the different morphological types of particles [315].
P(3HB-co-3HV) nanocarriers were used by Williams et al. for the controlled release of tetracycline [316]. In the hepatocellular carcinoma cell BEL7402, the PHA granules binding protein PhaP has been used in a receptor-mediated drug carrier using RBITC as a drug delivery model created using the modified emulsification/solvent diffusion approach [317]. Meanwhile, P(3HB) NPs with 55 nm average diameter were used for encapsulating retinoic acid through the dialysis method [318]. Shishatskaya et al. studied P(3HB) incorporated with rubomycin in vivo, and it was found that rubomycin-loaded PHB fabricated using the solvent evaporation method was effective in arresting carcinoma proliferation and thus, increased mice survival [319]. P(3HB-co-3HO) was employed as the drug carrier in targeted drug delivery. In another study, a new nanocarrier was formulated with folic acid (FA) and doxorubicin (DOX) as the targeting ligand and anticancer drug, respectively. This nanocarrier was found to be a potential candidate for the targeted delivery of anticancer drugs to the folate receptor-overexpressed cancer cells [320]. Sulperazone-loaded P(3HB-co-3HV) was employed for an in vitro antibiotic release [321]. Rossi et al. studied the release profile of gentamycin incorporated into P(3HB-co-3HV) and found that the copolymer with higher HV content released more gentamycin [322]. Additionally, P(3HHx-co-3HO) NPs were found to facilitate the permeation of tamsulosin drugs into the skin [323].

4. Conclusions and Future Perspectives

Different methods for the formation of PNPs were discussed in this review, including the dispersion of preformed polymers (emulsification solvent evaporation, nanoprecipitation, emulsification solvent diffusion, and salting-out) and the polymerization of monomers (emulsion polymerization, surfactant-free emulsion polymerization, mini-emulsion polymerization, and micro-emulsion polymerization). Furthermore, the effects of experimental variables on the characteristics of the formed PNPs were discussed. PNPs have a wide variety of applications; however, there are many challenges associated with the synthesis of PNPs that need to be addressed before PNPs can be fully utilized and integrated into applications, for example, the ability to reproduce PNPs with a good size distribution. The size distribution of PNPs obtained using currently existing methods is usually very broad and is not suited for most applications. In addition, there are not many reports on the scaled-up production of PNPs. Many scientific reports have also claimed that the synthesis of PNPs at a lab scale exists only as a proof of concept of the technology. Many produced and evaluated NPs never reach clinical trials due to their non-biocompatible physiochemical properties. For this reason, PHA NPs are ideal for use as drug carriers as they abide by the present regulatory requirements in terms of biodegradability, stability, and non-toxicity. Furthermore, because the size distribution is extremely large, the particle size loses certainty due to the wide range of size distribution. This situation poses a great challenge in using PNPs for drug delivery applications. Furthermore, PHA NPs can be produced using bacteria, which allows for green synthesis to produce nanocarriers that can be used extensively in the field of nanomedicine. With regards to application, PNPs can be used as drug carriers to target specific sites within cells or organs for more advanced treatment due to their unique properties and size. This would greatly improve the performance of targeted therapies. The PNPs can also be used for diagnostic purposes, either in the lab or in hospitals (point-of-care diagnostics).

Author Contributions

T.P., P.F., J.-A.C. and K.S. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Authors would like to thank Ministry of Education Malaysia (203/PBIOLOGI/67811001), titled “Soil Analysis and Value-Addition to Oil Palm Trunk (OPT) and Sap Through Biotechnology” as well as Science and Technology Research Partnership for Sustainable Development (SATREPS) (grant number JPMJSA1801) of the Japan Science and Technology Agency (JST)/Japan International Cooperation Agency (JICA) for their financial support.

Data Availability Statement

Data sharing is not applicable for this review.

Conflicts of Interest

The authors declare that they have no competing interests.

References

  1. Bhatia, S. Nanoparticles Types, Classification, Characterization, Fabrication Methods and Drug Delivery Applications. In Natural Polymer Drug Delivery Systems; Springer: Cham, Switzerland, 2016; pp. 33–93. [Google Scholar] [CrossRef]
  2. Albanese, A.; Tang, P.S.; Chan, W.C. The Effect of Nanoparticle Size, Shape, and Surface Chemistry on Biological Systems. Annu. Rev. Biomed. Eng. 2012, 14, 1–16. [Google Scholar] [CrossRef] [Green Version]
  3. Docter, D.; Strieth, S.; Westmeier, D.; Hayden, O.; Gao, M.; Knauer, S.K.; Stauber, R.H. No king without a crown—impact of the nanomaterial-protein corona on nanobiomedicine. Nanomedicine 2015, 10, 503–519. [Google Scholar] [CrossRef] [Green Version]
  4. Joye, I.J.; McClements, D.J. Production of Nanoparticles by Anti-Solvent Precipitation for Use in Food Systems. Trends Food Sci. Technol. 2013, 34, 109–123. [Google Scholar] [CrossRef]
  5. Rahman, M.; Laurent, S.; Tawil, N.; Yahia, L.H.; Mahmoudi, M. Nanoparticle and Protein Corona. In Protein-nanoparticle interactions; Springer: Berlin/Heidelberg, Germany, 2013; pp. 21–44. [Google Scholar] [CrossRef]
  6. Mahmoudi, M.; Lynch, I.; Ejtehadi, M.R.; Monopoli, M.P.; Bombelli, F.B.; Laurent, S. Protein−Nanoparticle Interactions: Opportunities and Challenges. Chem. Rev. 2011, 111, 5610–5637. [Google Scholar] [CrossRef]
  7. Ezhilarasi, P.N.; Karthik, P.; Chhanwal, N.; Anandharamakrishnan, C. Nanoencapsulation Techniques for Food Bioactive Components: A Review. Food Bioprocess Technol. 2013, 6, 628–647. [Google Scholar] [CrossRef]
  8. Sripriyalakshmi, S.; Jose, P.; Ravindran, A.; Anjali, C.H. Recent Trends in Drug Delivery System Using Protein Nanoparticles. Cell Biophys. 2014, 70, 17–26. [Google Scholar] [CrossRef]
  9. Bonifácio, B.V.; Silva, P.B.; Ramos, M.A.; Negri, K.M.; Bauab, T.M.; Chorilli, M. Nanotechnology-based drug delivery systems and herbal medicines: A review. Int. J. Nanomed. 2013, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  10. Weiss, J.; Takhistov, P.; McClements, D.J. Functional Materials in Food Nanotechnology. J. Food Sci. 2006, 71, R107–R116. [Google Scholar] [CrossRef] [Green Version]
  11. Kuhlbusch, T.A.; Asbach, C.; Fissan, H.; Göhler, D.; Stintz, M. Nanoparticle exposure at nanotechnology workplaces: A review. Part. Fibre Toxicol. 2011, 8, 22. [Google Scholar] [CrossRef] [Green Version]
  12. Sanguansri, P.; Augustin, M.A. Nanoscale Materials Development–A Food Industry Perspective. Trends Food Sci. Technol. 2006, 17, 547–556. [Google Scholar] [CrossRef]
  13. Rai, M.; Ingle, A. Role of nanotechnology in agriculture with special reference to management of insect pests. Appl. Microbiol. Biotechnol. 2012, 94, 287–293. [Google Scholar] [CrossRef]
  14. Justin, C.; Philip, S.A.; Samrot, A.V. Synthesis and characterization of superparamagnetic iron-oxide nanoparticles (SPIONs) and utilization of SPIONs in X-ray imaging. Appl. Nanosci. 2017, 7, 463–475. [Google Scholar] [CrossRef]
  15. Aggarwal, P.; Hall, J.B.; McLeland, C.B.; Dobrovolskaia, M.A.; McNeil, S.E. Nanoparticle interaction with plasma proteins as it relates to particle biodistribution, biocompatibility and therapeutic efficacy. Adv. Drug Deliv. Rev. 2009, 61, 428–437. [Google Scholar] [CrossRef] [Green Version]
  16. Kharazian, B.; Hadipour, N.; Ejtehadi, M. Understanding the nanoparticle–protein corona complexes using computational and experimental methods. Int. J. Biochem. Cell Biol. 2016, 75, 162–174. [Google Scholar] [CrossRef]
  17. Schöttler, S.; Landfester, K.; Mailänder, V. Controlling the Stealth Effect of Nanocarriers through Understanding the Protein Corona. Angew. Chem. Int. Ed. 2016, 55, 8806–8815. [Google Scholar] [CrossRef]
  18. Mirshafiee, V.; Mahmoudi, M.; Lou, K.; Cheng, J.; Kraft, M.L. Protein corona significantly reduces active targeting yield. Chem. Commun. 2013, 49, 2557–2559. [Google Scholar] [CrossRef]
  19. Lee, Y.K.; Choi, E.-J.; Webster, T.J.; Kim, S.-H.; Khang, D. Effect of the protein corona on nanoparticles for modulating cytotoxicity and immunotoxicity. Int. J. Nanomed. 2014, 10, 97–113. [Google Scholar] [CrossRef] [Green Version]
  20. Rak, J. Microparticles in Cancer. Semin. Thromb. Hemost. 2010, 36, 888–906. [Google Scholar] [CrossRef] [Green Version]
  21. Mause, S.F.; Weber, C. Microparticles. Circ. Res. 2010, 107, 1047–1057. [Google Scholar] [CrossRef]
  22. Rizvi, S.A.; Saleh, A.M. Applications of nanoparticle systems in drug delivery technology. Saudi Pharm. J. 2017, 26, 64–70. [Google Scholar] [CrossRef]
  23. Chenthamara, D.; Subramaniam, S.; Ramakrishnan, S.G.; Krishnaswamy, S.; Essa, M.M.; Lin, F.H.; Qoronfleh, M.W. Therapeutic efficacy of nanoparticles and routes of administration. Biomater. Res. 2019, 23, 20. [Google Scholar] [CrossRef]
  24. Zhou, Y.; Peng, Z.; Seven, E.S.; Leblanc, R.M. Crossing the blood-brain barrier with nanoparticles. J. Control. Release 2017, 270, 290–303. [Google Scholar] [CrossRef]
  25. Khan, R.; Ahmad, E.; Zaman, M.; Qadeer, A.; Rabbani, G. Nanoparticles in relation to peptide and protein aggregation. Int. J. Nanomed. 2014, 9, 899–912. [Google Scholar] [CrossRef] [Green Version]
  26. Magazù, S.; Migliardo, F.; Telling, M. Structural and dynamical properties of water in sugar mixtures. Food Chem. 2008, 106, 1460–1466. [Google Scholar] [CrossRef]
  27. Samrot, A.; Burman, U.; Philip, S.A.; Shobana, N.; Chandrasekaran, K. Synthesis of curcumin loaded polymeric nanoparticles from crab shell derived chitosan for drug delivery. Inform. Med. Unlocked 2018, 10, 159–182. [Google Scholar] [CrossRef]
  28. Panhwar, A.H.; Tuzen, M.; Hazer, B.; Kazi, T.G. Solid phase microextraction method using a novel polystyrene oleic acid imidazole polymer in micropipette tip of syringe system for speciation and determination of antimony in environmental and food samples. Talanta 2018, 184, 115–121. [Google Scholar] [CrossRef]
  29. Honarkar, H.; Barikani, M. Applications of Biopolymers I: Chitosan. Monatsh. Chem. 2009, 140, 1403. [Google Scholar] [CrossRef]
  30. Saleh, T.A.; Tuzen, M.; Sarı, A. Polyamide magnetic palygorskite for the simultaneous removal of Hg(II) and methyl mercury; with factorial design analysis. J. Environ. Manag. 2018, 211, 323–333. [Google Scholar] [CrossRef]
  31. Nasir, A.; Kausar, A.; Younus, A. A Review on Preparation, Properties and Applications of Polymeric Nanoparticle-Based Materials. Polym. Technol. Eng. 2014, 54, 325–341. [Google Scholar] [CrossRef]
  32. Geckeler, K.E.; Nishide, H. Advanced Nanomaterials; Wiley Online Library: Weinheim, Germany, 2010; Volume 2. [Google Scholar] [CrossRef]
  33. Derman, S.; Kizilbey, K.; Akdeste, Z.M. Polymeric Nanoparticles. Sigma J. Eng. Nat. Sci. 2013, 31, 107–120. [Google Scholar]
  34. Öztürk, K. Serbest Radikal Temizleyici Madde İçeren Nanopartiküler Taşıyıcı Sistemlerin Tasarımı Ve Değerlendirilmesi. Master’s Thesis, Hacettepe University, Ankara, Turkey, 2010. Available online: http://nek.istanbul.edu.tr:4444/ekos/TEZ/47041.pdf (accessed on 20 December 2021).
  35. Wang, Y.-J.; Larsson, M.; Huang, W.-T.; Chiou, S.-H.; Nicholls, S.J.; Chao, J.-I.; Liu, D.-M. The use of polymer-based nanoparticles and nanostructured materials in treatment and diagnosis of cardiovascular diseases: Recent advances and emerging designs. Prog. Polym. Sci. 2016, 57, 153–178. [Google Scholar] [CrossRef]
  36. Chang, M.-W.; Edirisinghe, M.; Stride, E. Ultrasound mediated release from stimuli-responsive core–shell capsules. J. Mater. Chem. B 2013, 1, 3962–3971. [Google Scholar] [CrossRef]
  37. Muller, R.H.; Keck, C.M. Challenges and Solutions for the Delivery of Biotech Drugs–A Review of Drug Nanocrystal Technology and Lipid Nanoparticles. J. Biotechnol. 2004, 113, 151–170. [Google Scholar] [CrossRef]
  38. Reis, C.; Neufeld, R.J.; Ribeiro, A.; Veiga, F. Nanoencapsulation, I. Methods for preparation of drug-loaded polymeric nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2006, 2, 8–21. [Google Scholar] [CrossRef] [Green Version]
  39. Anderson, D.G.; Burdick, J.A.; Langer, R. Smart Biomaterials. Science 2004, 305, 1923–1924. [Google Scholar] [CrossRef]
  40. Klodzinska, S.N.; Wan, F.; Jumaa, H.; Sternberg, C.; Rades, T.; Nielsen, H.M. Improved drug loading and antibacterial activity of minocycline-loaded PLGA nanoparticles prepared by solid/oil/water ion pairing method. Int. J. Nanomed. 2012, 7, 221–234. [Google Scholar] [CrossRef] [Green Version]
  41. Hornig, S.; Heinze, T.; Becer, C.R.; Schubert, U.S. Synthetic polymeric nanoparticles by nanoprecipitation. J. Mater. Chem. 2009, 19, 3838–3840. [Google Scholar] [CrossRef]
  42. Nagarwal, R.C.; Kant, S.; Singh, P.N.; Maiti, P.; Pandit, J.K. Polymeric nanoparticulate system: A potential approach for ocular drug delivery. J. Control. Release 2009, 136, 2–13. [Google Scholar] [CrossRef]
  43. Moreno-Vega, A.-I.; Gómez-Quintero, T.; Nuñez-Anita, R.E.; Acosta-Torres, L.-S.; Castaño, V. Polymeric and Ceramic Nanoparticles in Biomedical Applications. J. Nanotechnol. 2012, 2012, 1–10. [Google Scholar] [CrossRef] [Green Version]
  44. Yadav, H.K.; Almokdad, A.A.; Shaluf, S.I.; Debe, M.S. Polymer-Based Nanomaterials for Drug-Delivery Carriers. In Nanocarriers for Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2018; pp. 531–556. [Google Scholar] [CrossRef]
  45. Sailaja, A.K. A Review on Biomedical Applications of Polymeric Nanoparticles. Drug Des. Intellect. Prop. Int. J. 2018, 2, 216–220. [Google Scholar] [CrossRef]
  46. Parveen, S.; Misra, R.; Sahoo, S.K. Nanoparticles: A Boon to Drug Delivery, Therapeutics, Diagnostics and Imaging. Nanomed. Nanotechnol. Biol. Med. 2012, 8, 147–166. [Google Scholar] [CrossRef]
  47. De Jong, W.H.; Borm, P.J.A. Drug delivery and Nanoparticles: Applications and Hazards. Int. J. Nanomed. 2008, 3, 133–149. [Google Scholar] [CrossRef] [Green Version]
  48. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; del Pilar Rodriguez-Torres, M.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar] [CrossRef] [Green Version]
  49. Pignatello, R.; Impallomeni, G.; Cupri, S.; Puzzo, G.; Curcio, C.; Rizzo, M.G.; Guglielmino, S.; Ballistreri, A. Unsaturated Poly(Hydroxyalkanoates) for the Production of Nanoparticles and the Effect of Cross-Linking on Nanoparticle Features. Materials 2019, 12, 868. [Google Scholar] [CrossRef] [Green Version]
  50. Umesh, M.; Priyanka, K.; Thazeem, B.; Preethi, K. Biogenic PHA nanoparticle synthesis and characterization from Bacillus subtilis NCDC0671 using orange peel medium. Int. J. Polym. Mater. Polym. Biomater. 2017, 67, 996–1004. [Google Scholar] [CrossRef]
  51. Koosha, F.; Muller, R.; Washington, C. Production of Polyhydroxybutyrate (PHB) Nanoparticles for Drug Targeting. J. Pharm. Pharmacol. 1987, 39, 136P. [Google Scholar]
  52. Koosha, F.; Muller, R.; Davis, S.S.; Davies, M.C. The surface chemical structure of poly(β-hydroxybutyrate) microparticles produced by solvent evaporation process. J. Control. Release 1989, 9, 149–157. [Google Scholar] [CrossRef]
  53. Seyler, I.; Appel, M.; Devissaguet, J.-P.; Legrand, P.; Barratt, G. Macrophage Activation by a Lipophilic Derivative of Muramyldipeptide within Nanocapsules: Investigation of the Mechanism of Drug Delivery. J. Nanopart. Res. 1999, 1, 91–97. [Google Scholar] [CrossRef]
  54. Legrand, P.; Lesieur, S.; Bochot, A.; Gref, R.; Raatjes, W.; Barratt, G.; Vauthier, C. Influence of polymer behaviour in organic solution on the production of polylactide nanoparticles by nanoprecipitation. Int. J. Pharm. 2007, 344, 33–43. [Google Scholar] [CrossRef]
  55. Ueda, M.; Kreuter, J. Optimization of the preparation of loperamide-loaded poly (L-lactide) nanoparticles by high pressure emulsification-solvent evaporation. J. Microencapsul. 1997, 14, 593–605. [Google Scholar] [CrossRef]
  56. Nehilla, B.J.; Bergkvist, M.; Popat, K.C.; Desai, T.A. Purified and surfactant-free coenzyme Q10-loaded biodegradable nanoparticles. Int. J. Pharm. 2008, 348, 107–114. [Google Scholar] [CrossRef]
  57. Yallapu, M.M.; Gupta, B.K.; Jaggi, M.; Chauhan, S.C. Fabrication of curcumin encapsulated PLGA nanoparticles for improved therapeutic effects in metastatic cancer cells. J. Colloid Interface Sci. 2010, 351, 19–29. [Google Scholar] [CrossRef]
  58. Prado, L.B.; Huber, S.C.; Barnabé, A.; Bassora, F.D.S.; Paixão, D.S.; Durán, N.; Annichino-Bizzacchi, J.M. Characterization of PCL and Chitosan Nanoparticles as Carriers of Enoxaparin and Its Antithrombotic Effect in Animal Models of Venous Thrombosis. J. Nanotechnol. 2017, 2017, 1–7. [Google Scholar] [CrossRef] [Green Version]
  59. Ajiboye, A.L.; Trivedi, V.; Mitchell, J. Preparation of polycaprolactone nanoparticles via supercritical carbon dioxide extraction of emulsions. Drug Deliv. Transl. Res. 2017, 8, 1790–1796. [Google Scholar] [CrossRef] [Green Version]
  60. Shokri, N.; Javar, H.A.; Fouladdel, S.; Khalaj, A.; Khoshayand, M.R.; Dinarvand, R.; Atyabi, F.; Nomani, A.; Azizi, E. Preparation and Evaluation of Poly (Caprolactone Fumarate) Nanoparticles Containing Doxorubicin HCl. Daru J. Fac. Pharm. Tehran Univ. Med. Sci. 2011, 19, 12–22. [Google Scholar]
  61. Kateb, B.; Chiu, K.; Black, K.L.; Yamamoto, V.; Khalsa, B.; Ljubimova, J.Y.; Ding, H.; Patil, R.; Portilla-Arias, J.A.; Modo, M.; et al. Nanoplatforms for constructing new approaches to cancer treatment, imaging, and drug delivery: What should be the policy? NeuroImage 2011, 54, S106–S124. [Google Scholar] [CrossRef] [Green Version]
  62. Mansour, H.M.; Sohn, M.; Al-Ghananeem, A.; DeLuca, P.P. Materials for Pharmaceutical Dosage Forms: Molecular Pharmaceutics and Controlled Release Drug Delivery Aspects. Int. J. Mol. Sci. 2010, 11, 3298–3322. [Google Scholar] [CrossRef] [Green Version]
  63. Sundar, S.; Kundu, J.; Kundu, S.C. Biopolymeric Nanoparticles. Sci. Technol. Adv. Mater. 2010, 11, 014104. [Google Scholar] [CrossRef]
  64. Ludwig, A. The use of mucoadhesive polymers in ocular drug delivery. Adv. Drug Deliv. Rev. 2005, 57, 1595–1639. [Google Scholar] [CrossRef]
  65. Lanz-Landázuri, A.; Portilla-Arias, J.; De Ilarduya, A.M.; Álvarez, M.G.; Holler, E.; Ljubimova, J.; Muñoz-Guerra, S. Nanoparticles of Esterified Polymalic Acid for Controlled Anticancer Drug Release. Macromol. Biosci. 2014, 14, 1325–1336. [Google Scholar] [CrossRef] [Green Version]
  66. Loyer, P.; Cammas-Marion, S. Natural and synthetic poly(malic acid)-based derivates: A family of versatile biopolymers for the design of drug nanocarriers. J. Drug Target. 2014, 22, 556–575. [Google Scholar] [CrossRef] [PubMed]
  67. Gutul, T.; Rusu, E.; Condur, N.; Ursaki, V.; Goncearenco, E.; Vlazan, P. Preparation of poly(N-vinylpyrrolidone)-stabilized ZnO colloid nanoparticles. Beilstein J. Nanotechnol. 2014, 5, 402–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Sahu, A.; Solanki, P.; Mitra, S. Curcuminoid-loaded poly(methyl methacrylate) nanoparticles for cancer therapy. Int. J. Nanomed. 2018, 13, 101–105. [Google Scholar] [CrossRef] [Green Version]
  69. Mendes, A.N.; Hubber, I.; Siqueira, M.; Barbosa, G.M.; Moreira, D.D.L.; Holandino, C.; Pinto, J.C.; Nele, M. Preparation and Cytotoxicity of Poly(Methyl Methacrylate) Nanoparticles for Drug Encapsulation. Macromol. Symp. 2012, 319, 34–40. [Google Scholar] [CrossRef]
  70. Andreasen, S.; Chong, S.-F.; Wohl, B.M.; Goldie, K.N.; Zelikin, A.N. Poly(vinyl alcohol) Physical Hydrogel Nanoparticles, Not Polymer Solutions, Exert Inhibition of Nitric Oxide Synthesis in Cultured Macrophages. Biomacromolecules 2013, 14, 1687–1695. [Google Scholar] [CrossRef]
  71. Madlova, M.; Jones, S.; Zwerschke, I.; Ma, Y.; Hider, R.; Forbes, B. Poly(vinyl alcohol) nanoparticle stability in biological media and uptake in respiratory epithelial cell layers in vitro. Eur. J. Pharm. Biopharm. 2009, 72, 438–443. [Google Scholar] [CrossRef]
  72. Hu, Y.; Jiang, X.; Ding, Y.; Ge, H.; Yuan, Y.; Yang, C. Synthesis and characterization of chitosan–poly(acrylic acid) nanoparticles. Biomaterials 2002, 23, 3193–3201. [Google Scholar] [CrossRef]
  73. Molnar, R.M.; Bodnar, M.; Hartmann, J.F.; Borbely, J. Preparation and characterization of poly(acrylic acid)-based nanoparticles. Colloid Polym. Sci. 2009, 287, 739–744. [Google Scholar] [CrossRef]
  74. Gualdesi, M.; Igarzabal, C.A.; Vara, J.; Ortiz, C. Synthesis and physicochemical properties of polyacrylamide nanoparticles as photosensitizer carriers. Int. J. Pharm. 2016, 512, 213–218. [Google Scholar] [CrossRef]
  75. Giuntini, F.; Dumoulin, F.; Daly, R.; Ahsen, V.; Scanlan, E.M.; Lavado, A.S.P.; Aylott, J.W.; Rosser, G.A.; Beeby, A.; Boyle, R.W. Orthogonally bifunctionalised polyacrylamide nanoparticles: A support for the assembly of multifunctional nanodevices. Nanoscale 2012, 4, 2034–2045. [Google Scholar] [CrossRef]
  76. López-Muñoz, R.; Treviño, M.E.; Morales, G.; Valdez-Garza, J.A.; de León, R.D.; Saade, H.; Enríquez-Medrano, F.J.; López, R.G. Ultrafine Nanoparticles of Poly(Methyl Methacrylate-co-Methacrylic Acid) Loaded with Aspirin. J. Nanomater. 2019, 2019, 1–9. [Google Scholar] [CrossRef] [Green Version]
  77. Zhong, J.X.; Clegg, J.R.; Ander, E.W.; Peppas, N.A. Tunable poly(methacrylic acid-co-acrylamide) nanoparticles through inverse emulsion polymerization. J. Biomed. Mater. Res. Part A 2018, 106, 1677–1686. [Google Scholar] [CrossRef]
  78. Wang, Y.; Li, P.; Tran, T.T.-D.; Zhang, J.; Kong, L. Manufacturing Techniques and Surface Engineering of Polymer Based Nanoparticles for Targeted Drug Delivery to Cancer. Nanomaterials 2016, 6, 26. [Google Scholar] [CrossRef] [Green Version]
  79. Zielińska, A.; Carreiró, F.; Oliveira, A.; Neves, A.; Pires, B.; Venkatesh, D.; Durazzo, A.; Lucarini, M.; Eder, P.; Silva, A.; et al. Polymeric Nanoparticles: Production, Characterization, Toxicology and Ecotoxicology. Molecules 2020, 25, 3731. [Google Scholar] [CrossRef]
  80. Crucho, C.I.C.; Barros, M.T. Polymeric nanoparticles: A study on the preparation variables and characterization methods. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 80, 771–784. [Google Scholar] [CrossRef]
  81. Ibraheem, D.; Fessi, H.; Elaissari, A. DNA Encapsulation via Double Emulsion Like Process. J. Colloid Sci. Biotechnol. 2013, 2, 328–333. [Google Scholar] [CrossRef]
  82. Hao, S.; Wang, B.; Wang, Y.; Zhu, L.; Wang, B.; Guo, T. Preparation of Eudragit L 100-55 enteric nanoparticles by a novel emulsion diffusion method. Colloids Surf. B Biointerfaces 2013, 108, 127–133. [Google Scholar] [CrossRef]
  83. Saranya, S.; Radha, K.V. Review of Nanobiopolymers for Controlled Drug Delivery. Polym. Technol. Eng. 2014, 53, 1636–1646. [Google Scholar] [CrossRef]
  84. Valente, I.M.; Moreira, M.M.; Neves, P.; Da Fé, T.; Gonçalves, L.M.; Almeida, P.J.; Rodrigues, J.A. An Insight on Salting-out Assisted Liquid-Liquid Extraction for Phytoanalysis. Phytochem. Anal. 2017, 28, 297–304. [Google Scholar] [CrossRef]
  85. Hyde, A.M.; Zultanski, S.L.; Waldman, J.H.; Zhong, Y.-L.; Shevlin, M.; Peng, F. General Principles and Strategies for Salting-Out Informed by the Hofmeister Series. Org. Process Res. Dev. 2017, 21, 1355–1370. [Google Scholar] [CrossRef] [Green Version]
  86. Distler, D.; Neto, W.S.; Machado, F. Emulsion Polymerization. Ref. Modul. Mater. Sci. Mater. Eng. 2017. [Google Scholar] [CrossRef]
  87. Yamak, H.B. Emulsion Polymerization: Effects of Polymerization Variables on The Properties of Vinyl Acetate Based Emulsion Polymers. In Polymer Science; Yilmaz, F., Ed.; IntechOpen: London, UK, 2013; pp. 35–72. [Google Scholar] [CrossRef] [Green Version]
  88. Zhou, J.; Yao, H.; Ma, J. Recent advances in RAFT-mediated surfactant-free emulsion polymerization. Polym. Chem. 2018, 9, 2532–2561. [Google Scholar] [CrossRef]
  89. Bilgin, S.; Tomovska, R.; Asua, J.M. Surfactant-free high solids content polymer dispersions. Polymer 2017, 117, 64–75. [Google Scholar] [CrossRef]
  90. Faucheu, J.; Gauthier, C.; Chazeau, L.; Cavaillé, J.-Y.; Mellon, V.; Lami, E.B. Miniemulsion polymerization for synthesis of structured clay/polymer nanocomposites: Short review and recent advances. Polymer 2010, 51, 6–17. [Google Scholar] [CrossRef]
  91. Crespy, D.; Landfester, K. Miniemulsion polymerization as a versatile tool for the synthesis of functionalized polymers. Beilstein J. Org. Chem. 2010, 6, 1132–1148. [Google Scholar] [CrossRef]
  92. Iqbal, M.; Zafar, N.; Fessi, H.; Elaissari, A. Double emulsion solvent evaporation techniques used for drug encapsulation. Int. J. Pharm. 2015, 496, 173–190. [Google Scholar] [CrossRef]
  93. Sosa, N.; Peralta, R.; López, R.; Ramos, L.; Katime, I.; Cesteros, C.; Mendizábal, E.; Puig, J. A comparison of the characteristics of poly(vinyl acetate) latex with high solid content made by emulsion and semi-continuous microemulsion polymerization. Polymer 2001, 42, 6923–6928. [Google Scholar] [CrossRef]
  94. Rehm, B.H.A. Bacterial polymers: Biosynthesis, modifications and applications. Nat. Rev. Microbiol. 2010, 8, 578–592. [Google Scholar] [CrossRef]
  95. Moradali, M.F.; Rehm, B.H.A. Bacterial biopolymers: From pathogenesis to advanced materials. Nat. Rev. Microbiol. 2020, 18, 195–210. [Google Scholar] [CrossRef]
  96. Zhang, G.; Niu, A.; Peng, S.; Jiang, M.; Tu, Y.; Li, M.; Wu, C. Formation of Novel Polymeric Nanoparticles. Accounts Chem. Res. 2001, 34, 249–256. [Google Scholar] [CrossRef]
  97. Quintanar-Guerrero, D.; Allémann, E.; Fessi, H.; Doelker, E. Preparation Techniques and Mechanisms of Formation of Biodegradable Nanoparticles from Preformed Polymers. Drug Dev. Ind. Pharm. 1998, 24, 1113–1128. [Google Scholar] [CrossRef] [PubMed]
  98. Couvreur, P.; Dubernet, C.; Puisieux, F. Controlled Drug Delivery with Nanoparticles: Current Possibilities and Future Trends. Eur. J. Pharm. Biopharm. 1995, 41, 2–13. [Google Scholar]
  99. Wang, Q.; Wu, P.; Ren, W.; Xin, K.; Yang, Y.; Xie, C.; Yang, C.; Liu, Q.; Yu, L.; Jiang, X.; et al. Comparative studies of salinomycin-loaded nanoparticles prepared by nanoprecipitation and single emulsion method. Nanoscale Res. Lett. 2014, 9, 351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Kumar, A.; Sawant, K. Encapsulation of exemestane in polycaprolactone nanoparticles: Optimization, characterization, and release kinetics. Cancer Nanotechnol. 2013, 4, 57–71. [Google Scholar] [CrossRef]
  101. Moorthi, C.; Kathiresan, K. Fabrication of Dual Drug Loaded Polymeric Nanosuspension: Incorporating Analytical Hierarchy Process and Data Envelopment Analysis in the Selection of A Suitable Method. Int. J. Pharm. Pharm. Sci. 2013, 5, 499–504. [Google Scholar]
  102. Mora-Huertas, C.E.; Fessi, H.; Elaissari, A. Polymer-based nanocapsules for drug delivery. Int. J. Pharm. 2010, 385, 113–142. [Google Scholar] [CrossRef]
  103. Noriega-Peláez, E.K.; Mendoza-Muñoz, N.; Ganem-Quintanar, A.; Quintanar-Guerrero, D. Optimization of the emulsification and solvent displacement method for the preparation of solid lipid nanoparticles. Drug Dev. Ind. Pharm. 2010, 37, 160–166. [Google Scholar] [CrossRef]
  104. Müller, K.; Bugnicourt, E.; Latorre, M.; Jorda, M.; Echegoyen Sanz, Y.E.; Lagaron, J.M.; Miesbauer, O.; Bianchin, A.; Hankin, S.; Bölz, U.; et al. Review on the Processing and Properties of Polymer Nanocomposites and Nanocoatings and Their Applications in the Packaging, Automotive and Solar Energy Fields. Nanomaterials 2017, 7, 74. [Google Scholar] [CrossRef] [Green Version]
  105. de Oliveira, A.D.; Beatrice, C.A.G. Polymer Nanocomposites with Different Types of Nanofiller. In Nanocomposites-Recent Evolutions; IntechOpen: London, UK, 2018; pp. 103–128. [Google Scholar] [CrossRef] [Green Version]
  106. Parambath, A. Engineering of Biomaterials for Drug Delivery Systems: Beyond Polyethylene Glycol; Woodhead Publishing: Kidlington, UK, 2018. [Google Scholar]
  107. Sailaja, A.K.; Shreya, M. Preparation and Characterization of Naproxen Loaded Niosomes by Ether Injection Method. Nano Biomed. Eng. 2018, 10, 174–180. [Google Scholar] [CrossRef]
  108. Chuah, J.-A.; Kaplan, D.L.; Numata, K. Engineering Peptide-based Carriers for Drug and Gene Delivery. In Engineering in Translational Medicine; Springer: London, UK, 2013; pp. 667–689. [Google Scholar] [CrossRef]
  109. Di Mascolo, D.; Basnett, P.; Palange, A.L.; Francardi, M.; Roy, I.; Decuzzi, P. Tuning core hydrophobicity of spherical polymeric nanoconstructs for docetaxel delivery. Polym. Int. 2016, 65, 741–746. [Google Scholar] [CrossRef]
  110. Hoshyar, N.; Gray, S.; Han, H.; Bao, G. The effect of nanoparticle size on in vivo pharmacokinetics and cellular interaction. Nanomedicine 2016, 11, 673–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Yu, X.; Trase, I.; Ren, M.; Duval, K.; Guo, X.; Chen, Z. Design of Nanoparticle-Based Carriers for Targeted Drug Delivery. Available online: https://www.hindawi.com/journals/jnm/2016/1087250/ (accessed on 14 February 2021).
  112. Ageitos, J.M.; Chuah, J.-A.; Numata, K. Chapter Design Considerations for Properties of Nanocarriers on Disposition and Efficiency of Drug and Gene Delivery. In Nanomedicines: Design, Delivery and Detection; Royal Society of Chemistry: London, UK, 2016; pp. 1–22. [Google Scholar] [CrossRef]
  113. Chidambaram, M.; Krishnasamy, K. Application of Plackett-Burman Factorial Design in The Development of Curcumin Loaded Eudragit E 100 Nanoparticles. Nano Biomed. Eng. 2013, 5, 28–33. [Google Scholar] [CrossRef] [Green Version]
  114. Hickey, J.; Santos, J.L.; Williford, J.-M.; Mao, H.-Q. Control of polymeric nanoparticle size to improve therapeutic delivery. J. Control. Release 2015, 219, 536–547. [Google Scholar] [CrossRef] [Green Version]
  115. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, Applications and Toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  116. Ding, S.; Serra, C.A.; Vandamme, T.F.; Yu, W.; Anton, N. Double Emulsions Prepared by Two–Step Emulsification: History, State-Of-The-Art and Perspective. J. Control. Release 2019, 295, 31–49. [Google Scholar] [CrossRef]
  117. Goodarzi, F.; Zendehboudi, S. A Comprehensive Review on Emulsions and Emulsion Stability in Chemical and Energy Industries. Can. J. Chem. Eng. 2018, 97, 281–309. [Google Scholar] [CrossRef] [Green Version]
  118. Aubry, J.; Ganachaud, F.; Addad, J.-P.C.; Cabane, B. Nanoprecipitation of Polymethylmethacrylate by Solvent Shifting:Boundaries. Langmuir 2009, 25, 1970–1979. [Google Scholar] [CrossRef]
  119. Mora-Huertas, C.E.; Garrigues, O.; Fessi, H.; Elaissari, A. Nanocapsules prepared via nanoprecipitation and emulsification–diffusion methods: Comparative study. Eur. J. Pharm. Biopharm. 2012, 80, 235–239. [Google Scholar] [CrossRef]
  120. Lepeltier, E.; Bourgaux, C.; Couvreur, P. Nanoprecipitation and the “Ouzo effect”: Application to Drug Delivery Devices. Adv. Drug Deliv. Rev. 2014, 71, 86–97. [Google Scholar] [CrossRef]
  121. Amoyav, B.; Benny, O. Controlled and tunable polymer particles’ production using a single microfluidic device. Appl. Nanosci. 2018, 8, 905–914. [Google Scholar] [CrossRef] [Green Version]
  122. Vauthier, C.; Dubernet, C.; Fattal, E.; Pinto-Alphandary, H.; Couvreur, P. Poly(alkylcyanoacrylates) as biodegradable materials for biomedical applications. Adv. Drug Deliv. Rev. 2003, 55, 519–548. [Google Scholar] [CrossRef]
  123. Mishra, B.; Patel, B.B.; Tiwari, S. Colloidal nanocarriers: A review on formulation technology, types and applications toward targeted drug delivery. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 9–24. [Google Scholar] [CrossRef] [PubMed]
  124. Moshfeghi, A.A.; Peyman, G.A. Micro-and Nanoparticulates. Adv. Drug Deliv. Rev. 2005, 57, 2047–2052. [Google Scholar] [CrossRef]
  125. Álvarez-Román, R.; Cavazos-Rodríguez, M.; Chávez-Montes, A.; Castro-Ríos, R.; Waksman de Torres, N.; Salazar-Cavazos, M.; Galindo Rodríguez, S. Formulación y caracterización de nanocápsulas con un antioxidante natural para su aplicación cutánea (Formulation and Characterization of Nanocapsules as a Natural Antioxidant for Cutaneous Application). Química Hoy (Chem. Sci.) 2011, 1, 29–35. Available online: http://eprints.uanl.mx/13446/1/Art6.pdf (accessed on 15 November 2021).
  126. Shi, W.; Zhang, Z.-J.; Yuan, Y.; Xing, E.-M.; Qin, Y.; Peng, Z.-J.; Zhang, Z.-P.; Yang, K.-Y. Optimization of parameters for preparation of docetaxel-loaded PLGA nanoparticles by nanoprecipitation method. J. Huazhong Univ. Sci. Technol. 2013, 33, 754–758. [Google Scholar] [CrossRef]
  127. Teng, Z.; Luo, Y.; Wang, T.; Zhang, B.; Wang, Q. Development and Application of Nanoparticles Synthesized with Folic Acid Conjugated Soy Protein. J. Agric. Food Chem. 2013, 61, 2556–2564. [Google Scholar] [CrossRef] [PubMed]
  128. Kim, S.; Kim, Y.S. Production of gliadin-poly(ethyl cyanoacrylate) nanoparticles for hydrophilic coating. J. Nanopart. Res. 2014, 16, 2277. [Google Scholar] [CrossRef]
  129. Lebouille, J.G.J.L.; Stepanyan, R.; Slot, J.J.M.; Stuart Cohen, M.A.; Tuinier, R. Nanoprecipitation of Polymers in a Bad Solvent. Colloids Surf. A Physicochem. Eng. Asp. 2014, 460, 225–235. [Google Scholar] [CrossRef]
  130. Moorthi, C.; Kathiresan, K. Fabrication of highly stable sonication assisted curcumin nanocrystals by nanoprecipitation method. Drug Inven. Today 2013, 5, 66–69. [Google Scholar] [CrossRef]
  131. He, Y.; Huang, Y.; Cheng, Y. Structure Evolution of Curcumin Nanoprecipitation from a Micromixer. Cryst. Growth Des. 2010, 10, 1021–1024. [Google Scholar] [CrossRef]
  132. Peltonen, L.; Aitta, J.; Hyvönen, S.; Karjalainen, M.; Hirvonen, J. Improved Entrapment Efficiency of Hydrophilic Drug Substance During Nanoprecipitation of Poly(I)Lactide Nanoparticles. AAPS PharmSciTech 2004, 5, 115. [Google Scholar] [CrossRef] [Green Version]
  133. He, W.; Lu, Y.; Qi, J.; Chen, L.; Hu, F.; Wu, W. Food proteins as novel nanosuspension stabilizers for poorly water-soluble drugs. Int. J. Pharm. 2013, 441, 269–278. [Google Scholar] [CrossRef]
  134. Yan, X.; Delgado, M.; Fu, A.; Alcouffe, P.; Gouin, S.G.; Fleury, E.; Katz, J.L.; Ganachaud, F.; Bernard, J. Simple but Precise Engineering of Functional Nanocapsules through Nanoprecipitation. Angew. Chem. Int. Ed. 2014, 53, 6910–6913. [Google Scholar] [CrossRef]
  135. Sah, E.; Sah, H. Recent Trends in Preparation of Poly(lactide- co -glycolide) Nanoparticles by Mixing Polymeric Organic Solution with Antisolvent. J. Nanomater. 2015, 2015, 1–22. [Google Scholar] [CrossRef] [Green Version]
  136. Barreras-Urbina, C.G.; Ramírez-Wong, B.; López-Ahumada, G.A.; Ibarra, S.E.B.; Martínez-Cruz, O.; Tapia-Hernández, J.A.; Félix, F.R. Nano- and Micro-Particles by Nanoprecipitation: Possible Application in the Food and Agricultural Industries. Int. J. Food Prop. 2016, 19, 1912–1923. [Google Scholar] [CrossRef]
  137. Ansari, M.J.; Alshahrani, S. Nano-encapsulation and characterization of baricitinib using poly-lactic-glycolic acid co-polymer. Saudi Pharm. J. 2019, 27, 491–501. [Google Scholar] [CrossRef]
  138. Sharma, N.; Madan, P.; Lin, S. Effect of process and formulation variables on the preparation of parenteral paclitaxel-loaded biodegradable polymeric nanoparticles: A co-surfactant study. Asian J. Pharm. Sci. 2016, 11, 404–416. [Google Scholar] [CrossRef] [Green Version]
  139. Calzoni, E.; Cesaretti, A.; Polchi, A.; Di Michele, A.; Tancini, B.; Emiliani, C. Biocompatible Polymer Nanoparticles for Drug Delivery Applications in Cancer and Neurodegenerative Disorder Therapies. J. Funct. Biomater. 2019, 10, 4. [Google Scholar] [CrossRef] [Green Version]
  140. Chang, J.; Jallouli, Y.; Kroubi, M.; Yuan, X.-B.; Feng, W.; Kang, C.-S.; Pu, P.-Y.; Betbeder, D. Characterization of endocytosis of transferrin-coated PLGA nanoparticles by the blood–brain barrier. Int. J. Pharm. 2009, 379, 285–292. [Google Scholar] [CrossRef]
  141. Nassar, T.; Rom, A.; Nyska, A.; Benita, S. Novel double coated nanocapsules for intestinal delivery and enhanced oral bioavailability of tacrolimus, a P-gp substrate drug. J. Control. Release 2009, 133, 77–84. [Google Scholar] [CrossRef]
  142. de Assis, D.N.; Mosqueira, V.C.F.; Vilela, J.M.C.; Andrade, M.S.; Cardoso, V.N. Release Profiles and Morphological Characterization by Atomic Force Microscopy and Photon Correlation Spectroscopy of 99mTechnetium-Fluconazole Nanocapsules. Int. J. Pharm. 2008, 349, 152–160. [Google Scholar] [CrossRef]
  143. Mugheirbi, N.A.; Paluch, K.J.; Tajber, L. Heat induced evaporative antisolvent nanoprecipitation (HIEAN) of itraconazole. Int. J. Pharm. 2014, 471, 400–411. [Google Scholar] [CrossRef] [Green Version]
  144. Salatin, S.; Barar, J.; Barzegar-Jalali, M.; Adibkia, K.; Kiafar, F.; Jelvehgari, M. Development of a nanoprecipitation method for the entrapment of a very water soluble drug into Eudragit RL nanoparticles. Res. Pharm. Sci. 2017, 12, 1–14. [Google Scholar] [CrossRef] [Green Version]
  145. Huang, W.; Zhang, C. Tuning the Size of Poly(lactic-co-glycolic Acid) (PLGA) Nanoparticles Fabricated by Nanoprecipitation. Biotechnol. J. 2017, 13, 1700203. [Google Scholar] [CrossRef]
  146. Madani, F.; Esnaashari, S.S.; Mujokoro, B.; Dorkoosh, F.; Khosravani, M.; Adabi, M. Investigation of Effective Parameters on Size of Paclitaxel Loaded PLGA Nanoparticles. Adv. Pharm. Bull. 2018, 8, 77–84. [Google Scholar] [CrossRef] [Green Version]
  147. Pandey, S.K.; Patel, D.K.; Thakur, R.; Mishra, D.P.; Maiti, P.; Haldar, C. Anti-cancer evaluation of quercetin embedded PLA nanoparticles synthesized by emulsified nanoprecipitation. Int. J. Biol. Macromol. 2015, 75, 521–529. [Google Scholar] [CrossRef]
  148. Vuddanda, P.R.; Mishra, A.; Singh, S.K. Development of polymeric nanoparticles with highly entrapped herbal hydrophilic drug using nanoprecipitation technique: An approach of quality by design. Pharm. Dev. Technol. 2014, 20, 579–587. [Google Scholar] [CrossRef]
  149. Kara, A.; Ozturk, N.; Sarisozen, C.; Vural, I. Investigation of Formulation Parameters of PLGA Nanoparticles Prepared by Nanoprecipitation Technique. In Proceedings of the 5th International Conference on Nanotechnology: Fundamentals and Application, Prague, Czech Republic, 11–13 August 2014; p. 94. [Google Scholar]
  150. Lucas, P.; Vaysse, M.; Aubry, J.; Mariot, D.; Sonnier, R.; Ganachaud, F. Finest nanocomposite films from carbon nanotube-loaded poly(methyl methacrylate) nanoparticles obtained by the Ouzo effect. Soft Matter 2011, 7, 5528–5531. [Google Scholar] [CrossRef]
  151. Yadav, K.S.; Sawant, K.K. Modified Nanoprecipitation Method for Preparation of Cytarabine-Loaded PLGA Nanoparticles. AAPS PharmSciTech 2010, 11, 1456–1465. [Google Scholar] [CrossRef] [Green Version]
  152. Khan, S.A.; Schneider, M. Improvement of Nanoprecipitation Technique for Preparation of Gelatin Nanoparticles and Potential Macromolecular Drug Loading. Macromol. Biosci. 2013, 13, 455–463. [Google Scholar] [CrossRef]
  153. Zhang, J.-Y.; Shen, Z.-G.; Zhong, J.; Hu, T.-T.; Chen, J.-F.; Ma, Z.-Q.; Yun, J. Preparation of amorphous cefuroxime axetil nanoparticles by controlled nanoprecipitation method without surfactants. Int. J. Pharm. 2006, 323, 153–160. [Google Scholar] [CrossRef]
  154. Ansari, M. Factors Affecting Preparation and Properties of Nanoparticles by Nanoprecipitation Method. Indo Am. J. Pharm. Sci. 2017, 4, 4854–4858. [Google Scholar]
  155. Zhuang, J.; Fang, R.H.; Zhang, L. Preparation of Particulate Polymeric Therapeutics for Medical Applications. Small Methods 2017, 1, 1700147. [Google Scholar] [CrossRef]
  156. Bilati, U.; Allémann, E.; Doelker, E. Development of a nanoprecipitation method intended for the entrapment of hydrophilic drugs into nanoparticles. Eur. J. Pharm. Sci. 2005, 24, 67–75. [Google Scholar] [CrossRef]
  157. Lince, F.; Marchisio, D.; Barresi, A. Strategies to control the particle size distribution of poly-ε-caprolactone nanoparticles for pharmaceutical applications. J. Colloid Interface Sci. 2008, 322, 505–515. [Google Scholar] [CrossRef]
  158. Liu, Y.; Lu, Y.C.; Luo, G.S. Modified nanoprecipitation method for polysulfone nanoparticles preparation. Soft Matter 2014, 10, 3414–3420. [Google Scholar] [CrossRef]
  159. de Oliveira, A.M.; Jäger, E.; Jäger, A.; Stepánek, P.; Giacomelli, F.C. Physicochemical aspects behind the size of biodegradable polymeric nanoparticles: A step forward. Colloids Surf. A Physicochem. Eng. Asp. 2013, 436, 1092–1102. [Google Scholar] [CrossRef]
  160. Olenius, T.; Yli-Juuti, T.; Elm, J.; Kontkanen, J.; Riipinen, I. New Particle Formation and Growth: Creating a New Atmospheric Phase Interface. In Physical Chemistry of Gas-Liquid Interface; Faust, J.A., House, J.E., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; Chapter 11; pp. 315–352. [Google Scholar] [CrossRef]
  161. Cheng, J.C.; Vigil, R.; Fox, R. A competitive aggregation model for Flash NanoPrecipitation. J. Colloid Interface Sci. 2010, 351, 330–342. [Google Scholar] [CrossRef]
  162. Rao, J.P.; Geckeler, K.E. Polymer nanoparticles: Preparation techniques and size-control parameters. Prog. Polym. Sci. 2011, 36, 887–913. [Google Scholar] [CrossRef]
  163. Chorny, M.; Fishbein, I.; Danenberg, H.D.; Golomb, G. Lipophilic drug loaded nanospheres prepared by nanoprecipitation: Effect of formulation variables on size, drug recovery and release kinetics. J. Control. Release 2002, 83, 389–400. [Google Scholar] [CrossRef]
  164. Asadi, H.; Rostamizadeh, K.; Salari, D.; Hamidi, M. Preparation of biodegradable nanoparticles of tri-block PLA–PEG–PLA copolymer and determination of factors controlling the particle size using artificial neural network. J. Microencapsul. 2011, 28, 406–416. [Google Scholar] [CrossRef]
  165. Alshamsan, A. Nanoprecipitation is more efficient than emulsion solvent evaporation method to encapsulate cucurbitacin I in PLGA nanoparticles. Saudi Pharm. J. 2013, 22, 219–222. [Google Scholar] [CrossRef]
  166. Patel, K.C.; Pramanik, S. Formulation and Characterization of Mefenamic Acid Loaded Polymeric Nanoparticles. World J. Pharm. Pharm. Sci. 2014, 3, 1391–1405. [Google Scholar]
  167. Lancheros, R.; Guerrero, C.A.; Godoy-Silva, R.D. Improvement of N-Acetylcysteine Loaded in PLGA Nanoparticles by Nanoprecipitation Method. J. Nanotechnol. 2018, 2018, 1–11. [Google Scholar] [CrossRef] [Green Version]
  168. Chidambaram, M.; Krishnasamy, K. Modifications to the Conventional Nanoprecipitation Technique: An Approach to Fabricate Narrow Sized Polymeric Nanoparticles. Adv. Pharm. Bull. 2013, 4, 205–208. [Google Scholar] [CrossRef]
  169. Bilati, U.; Allémann, E.; Doelker, E. Nanoprecipitation versus emulsion-based techniques for the encapsulation of proteins into biodegradable nanoparticles and process-related stability issues. AAPS PharmSciTech 2005, 6, E594–E604. [Google Scholar] [CrossRef] [Green Version]
  170. Luo, C.; Okubo, T.; Nangrejo, M.; Edirisinghe, M. Preparation of polymeric nanoparticles by novel electrospray nanoprecipitation. Polym. Int. 2014, 64, 183–187. [Google Scholar] [CrossRef]
  171. Zhao, C.-X.; He, L.; Qiao, S.Z.; Middelberg, A.P. Nanoparticle synthesis in microreactors. Chem. Eng. Sci. 2011, 66, 1463–1479. [Google Scholar] [CrossRef]
  172. Heshmatnezhad, F.; Nazar, A.R.S. On-chip controlled synthesis of polycaprolactone nanoparticles using continuous-flow microfluidic devices. J. Flow Chem. 2020, 10, 533–543. [Google Scholar] [CrossRef]
  173. Zhang, L.; Chen, Q.; Ma, Y.; Sun, J. Microfluidic Methods for Fabrication and Engineering of Nanoparticle Drug Delivery Systems. ACS Appl. Bio. Mater. 2019, 3, 107–120. [Google Scholar] [CrossRef]
  174. Senthilkumar, P.; Dawn, S.S.; Saipriya, C.; Samrot, A.V. Synthesis of polyhydroxybutyrate nanoparticles using surfactant (SPAN20) for hydrophobic drug delivery. Rasayan J. Chem. 2018, 11, 1686–1695. [Google Scholar] [CrossRef]
  175. Shakeri, F.; Shakeri, S.; Hojjatoleslami, M. Preparation and Characterization of Carvacrol Loaded Polyhydroxybutyrate Nanoparticles by Nanoprecipitation and Dialysis Methods. J. Food Sci. 2014, 79, N697–N705. [Google Scholar] [CrossRef]
  176. Anton, N.; Benoit, J.-P.; Saulnier, P. Design and production of nanoparticles formulated from nano-emulsion templates—A review. J. Control. Release 2008, 128, 185–199. [Google Scholar] [CrossRef]
  177. Pal, S.L.; Jana, U.; Manna, P.K.; Mohanta, G.P.; Manavalan, R. Nanoparticle: An Overview of Preparation and Characterization. J. Appl. Pharm. Sci. 2011, 1, 228–234. [Google Scholar]
  178. Håkansson, A.; Rayner, M. General Principles of Nanoemulsion Formation by High-Energy Mechanical Methods. Nanoemulsions 2018, 103–139. [Google Scholar] [CrossRef]
  179. Goyal, A.K.; Garg, T.; Bhandari, S.; Rath, G. Advancement in pulmonary drug delivery systems for treatment of tuberculosis. In Nanostructures for Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2017; pp. 669–695. [Google Scholar] [CrossRef]
  180. Wang, J.; Shi, A.; Agyei, D.; Wang, Q. Formulation of water-in-oil-in-water (W/O/W) emulsions containing trans-resveratrol. RSC Adv. 2017, 7, 35917–35927. [Google Scholar] [CrossRef] [Green Version]
  181. Stauffer, F.; Peter, B.; Alem, H.; Funfschilling, D.; Dumas, N.; Serra, C.; Roques-Carmes, T. Polyelectrolytes layer-by-layer surface modification of PDMS microchips for the production of simple O/W and double W/O/W emulsions: From global to localized treatment. Chem. Eng. Process. Process Intensif. 2019, 146, 107685. [Google Scholar] [CrossRef] [Green Version]
  182. Khadem, B.; Sheibat-Othman, N. Modeling droplets swelling and escape in double emulsions using population balance equations. Chem. Eng. J. 2019, 382, 122824. [Google Scholar] [CrossRef]
  183. Khadem, B.; Khellaf, M.; Sheibat-Othman, N. Investigating Swelling-Breakdown in Double Emulsions. Colloids Surf. A Physicochem. Eng. Asp. 2020, 585, 124181. [Google Scholar] [CrossRef]
  184. Mendoza-Muñoz, N.; Alcalá-Alcalá, S.; Quintanar-Guerrero, D. Preparation of Polymer Nanoparticles by the Emulsification-Solvent Evaporation Method: From Vanderhoff’s Pioneer Approach to Recent Adaptations. In Polymer Nanoparticles for Nanomedicines; Springer: Cham, Switzerland, 2016; pp. 87–121. [Google Scholar] [CrossRef]
  185. Wang, Y.; Li, P.; Peng, Z.; She, F.H.; Kong, L.X. Microencapsulation of Nanoparticles with Enhanced Drug Loading for pH-Sensitive Oral Drug Delivery for the Treatment of Colon Cancer. J. Appl. Polym. Sci. 2013, 129, 714–720. [Google Scholar] [CrossRef]
  186. Rosca, I.D.; Watari, F.; Uo, M. Microparticle formation and its mechanism in single and double emulsion solvent evaporation. J. Control. Release 2004, 99, 271–280. [Google Scholar] [CrossRef]
  187. Soppimath, K.S.; Aminabhavi, T.M.; Kulkarni, A.R.; Rudzinski, W.E. Biodegradable polymeric nanoparticles as drug delivery devices. J. Control. Release 2001, 70, 1–20. [Google Scholar] [CrossRef]
  188. Pisani, E.; Fattal, E.; Paris, J.; Ringard, C.; Rosilio, V.; Tsapis, N. Surfactant dependent morphology of polymeric capsules of perfluorooctyl bromide: Influence of polymer adsorption at the dichloromethane–water interface. J. Colloid Interface Sci. 2008, 326, 66–71. [Google Scholar] [CrossRef]
  189. Jeong, Y.-I.; Cho, C.-S.; Kim, S.-H.; Ko, K.-S.; Kim, S.-I.; Shim, Y.-H.; Nah, J.-W. Preparation of poly(DL-lactide-co-glycolide) nanoparticles without surfactant. J. Appl. Polym. Sci. 2001, 80, 2228–2236. [Google Scholar] [CrossRef]
  190. Staff, R.H.; Landfester, K.; Crespy, D. Recent Advances in the Emulsion Solvent Evaporation Technique for the Preparation of Nanoparticles and Nanocapsules. In Hierarchical Macromolecular Structures: 60 Years after the Staudinger Nobel Prize II; Springer: Berlin/Heidelberg, Germany, 2013; pp. 329–344. [Google Scholar] [CrossRef]
  191. Deshmukh, R.; Wagh, P.; Naik, J. Solvent evaporation and spray drying technique for micro- and nanospheres/particles preparation: A review. Dry. Technol. 2016, 34, 1758–1772. [Google Scholar] [CrossRef]
  192. Musyanovych, A.; Schmitz-Wienke, J.; Mailänder, V.; Walther, P.; Landfester, K. Preparation of Biodegradable Polymer Nanoparticles by Miniemulsion Technique and Their Cell Interactions. Macromol. Biosci. 2008, 8, 127–139. [Google Scholar] [CrossRef]
  193. Budhian, A.; Siegel, S.J.; Winey, K.I. Haloperidol-loaded PLGA nanoparticles: Systematic study of particle size and drug content. Int. J. Pharm. 2007, 336, 367–375. [Google Scholar] [CrossRef]
  194. Bilati, U.; Allémann, E.; Doelker, E. Sonication Parameters for the Preparation of Biodegradable Nanocapsules of Controlled Size by the Double Emulsion Method. Pharm. Dev. Technol. 2003, 8, 1–9. [Google Scholar] [CrossRef]
  195. Mainardes, R.M.; Evangelista, R.C. Praziquantel-loaded PLGA nanoparticles: Preparation and characterization. J. Microencapsul. 2005, 22, 13–24. [Google Scholar] [CrossRef]
  196. Potineni, A.; Lynn, D.M.; Langer, R.; Amiji, M.M. Poly (Ethylene Oxide)-Modified Poly (β-amino ester) Nanoparticles as A pH-Sensitive Biodegradable System for Paclitaxel Delivery. J. Control. Release 2003, 86, 223–234. [Google Scholar] [CrossRef]
  197. Kılıçay, E.; Demirbilek, M.; Türk, M.; Güven, E.; Hazer, B.; Denkbas, E.B. Preparation and characterization of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHX) based nanoparticles for targeted cancer therapy. Eur. J. Pharm. Sci. 2011, 44, 310–320. [Google Scholar] [CrossRef]
  198. Iqbal, M.; Valour, J.-P.; Fessi, H.; Elaissari, A. Preparation of biodegradable PCL particles via double emulsion evaporation method using ultrasound technique. Colloid Polym. Sci. 2014, 293, 861–873. [Google Scholar] [CrossRef]
  199. Zhang, C.; Zhang, Z.; Zhao, L. Folate-decorated Poly(3-Hydroxybutyrate-co-3-Hydroxyoctanoate) Nanoparticles for Targeting Delivery: Optimization and In Vivo Antitumor Activity. Drug Deliv. 2016, 23, 1830–1837. [Google Scholar] [CrossRef] [Green Version]
  200. Lu, X.-Y.; Li, M.; Zhu, X.-L.; Fan, F.; Wang, L.-L.; Ma, J.-G. Microbial synthesized biodegradable PHBHHxPEG hybrid copolymer as an efficient intracellular delivery nanocarrier for kinase inhibitor. BMC Biotechnol. 2014, 14, 4. [Google Scholar] [CrossRef] [Green Version]
  201. Shah, M.; Ullah, N.; Choi, M.H.; Kim, M.O.; Yoon, S.C. Amorphous Amphiphilic P(3HV-co-4HB)-b-mPEG Block Copolymer Synthesized from Bacterial Copolyester Via Melt Transesterification: Nanoparticle Preparation, Cisplatin-Loading for Cancer Therapy and In Vitro Evaluation. Eur. J. Pharm. Biopharm. 2012, 80, 518–527. [Google Scholar] [CrossRef] [PubMed]
  202. Kilicay, E.; Karahaliloglu, Z.; Hazer, B.; Tekin, I.; Denkbas, E.B. Concanavaline A conjugated bacterial polyester-based PHBHHx nanoparticles loaded with curcumin for breast cancer therapy. J. Microencapsul. 2016, 33, 274–285. [Google Scholar] [CrossRef]
  203. Masood, F.; Chen, P.; Yasin, T.; Fatima, N.; Hasan, F.; Hameed, A. Encapsulation of Ellipticine in poly-(3-hydroxybutyrate-co-3-hydroxyvalerate) based nanoparticles and its in vitro application. Mater. Sci. Eng. C 2013, 33, 1054–1060. [Google Scholar] [CrossRef] [PubMed]
  204. Shah, M.; Ullah, N.; Choi, M.H.; Yoon, S.C. Nanoscale Poly(4-Hydroxybutyrate)-mPEG Carriers for Anticancer Drugs Delivery. J. Nanosci. Nanotechnol. 2014, 14, 8416–8421. [Google Scholar] [CrossRef]
  205. Ben David-Naim, M.; Grad, E.; Aizik, G.; Nordling-David, M.M.; Moshel, O.; Granot, Z.; Golomb, G. Polymeric nanoparticles of siRNA prepared by a double-emulsion solvent-diffusion technique: Physicochemical properties, toxicity, biodistribution and efficacy in a mammary carcinoma mice model. Biomaterials 2017, 145, 154–167. [Google Scholar] [CrossRef] [PubMed]
  206. Haque, S.; Boyd, B.J.; McIntosh, M.P.; Pouton, C.W.; Kaminskas, L.M.; Whittaker, M. Suggested Procedures for The Reproducible Synthesis of Poly(d,l-Lactide-co-Glycolide) Nanoparticles using The Emulsification Solvent Diffusion Platform. Curr. Nanosci. 2018, 14, 448–453. [Google Scholar] [CrossRef] [PubMed]
  207. Baena-Aristizábal, C.M.; Fessi, H.; Elaissari, A.; Mora-Huertas, C.E. Biodegradable microparticles preparation by double emulsification—Solvent extraction method: A Systematic study. Colloids Surf. A Physicochem. Eng. Asp. 2016, 492, 213–229. [Google Scholar] [CrossRef]
  208. Miyazaki, Y.; Sugihara, H.; Nishiura, A.; Kadota, K.; Tozuka, Y.; Takeuchi, H. Appropriate Selection of An Aggregation Inhibitor of Fine Particles Used for Inhalation Prepared by Emulsion Solvent Diffusion. Drug Dev. Ind. Pharm. 2017, 43, 30–41. [Google Scholar] [CrossRef] [PubMed]
  209. Quintanar-Guerrero, D.; Allémann, E.; Doelker, E.; Fessi, H. A mechanistic study of the formation of polymer nanoparticles by the emulsification-diffusion technique. Colloid Polym. Sci. 1997, 275, 640–647. [Google Scholar] [CrossRef] [Green Version]
  210. Mohanraj, V.; Chen, Y. Nanoparticles-A Review. Trop. J. Pharm. Res. 2006, 5, 561–573. [Google Scholar] [CrossRef] [Green Version]
  211. Nagavarma, B.; Yadav, H.K.; Ayaz, A.; Vasudha, L.; Shivakumar, H. Different Techniques for Preparation of Polymeric Nanoparticles-A Review. Asian J. Pharm. Clin. Res. 2012, 5, 16–23. [Google Scholar]
  212. Perez, C.; Sanchez, A.; Putnam, D.; Ting, D.; Langer, R.; Alonso, M. Poly(lactic acid)-poly(ethylene glycol) nanoparticles as new carriers for the delivery of plasmid DNA. J. Control. Release 2001, 75, 211–224. [Google Scholar] [CrossRef]
  213. Ma, J.; Feng, P.; Ye, C.; Wang, Y.; Fan, Y. An improved interfacial coacervation technique to fabricate biodegradable nanocapsules of an aqueous peptide solution from polylactide and its block copolymers with poly(ethylene glycol). Colloid Polym. Sci. 2001, 279, 387–392. [Google Scholar] [CrossRef]
  214. Hassou, M. Modélisation et Simulation de la Formation des Nanocapsules Polymériques Par la Méthode D’émulsion-Diffusion. Ph.D. Thesis, Université Claude Bernard-Lyon I, Villeurbanne, France, 2007. Available online: https://tel.archives-ouvertes.fr/tel-00264755v2/document (accessed on 25 August 2021).
  215. Moinard-Chécot, D.; Chevalier, Y.; Briancon, S.; Beney, L.; Fessi, H. Mechanism of nanocapsules formation by the emulsion–diffusion process. J. Colloid Interface Sci. 2008, 317, 458–468. [Google Scholar] [CrossRef]
  216. Pramual, S.; Assavanig, A.; Bergkvist, M.; Batt, C.A.; Sunintaboon, P.; Lirdprapamongkol, K.; Svasti, J.; Niamsiri, N. Development and characterization of bio-derived polyhydroxyalkanoate nanoparticles as a delivery system for hydrophobic photodynamic therapy agents. J. Mater. Sci. Mater. Med. 2015, 27, 1–11. [Google Scholar] [CrossRef]
  217. Lu, X.-Y.; Ciraolo, E.; Stefenia, R.; Chen, G.-Q.; Zhang, Y.; Hirsch, E. Sustained release of PI3K inhibitor from PHA nanoparticles and in vitro growth inhibition of cancer cell lines. Appl. Microbiol. Biotechnol. 2011, 89, 1423–1433. [Google Scholar] [CrossRef]
  218. Chen, C.; Yang, W.; Wang, D.-T.; Chen, C.-L.; Zhuang, Q.-Y.; Kong, X.-D. A modified spontaneous emulsification solvent diffusion method for the preparation of curcumin-loaded PLGA nanoparticles with enhanced in vitro anti-tumor activity. Front. Mater. Sci. 2014, 8, 332–342. [Google Scholar] [CrossRef]
  219. Guinebretière, S.; Briancon, S.; Lieto, J.; Mayer, C.; Fessi, H. Study of the emulsion-diffusion of solvent: Preparation and characterization of nanocapsules. Drug Dev. Res. 2002, 57, 18–33. [Google Scholar] [CrossRef]
  220. Trimaille, T.; Pichot, C.; Fessi, H.; Delair, T. Poly(d,l-lactic acid) nanoparticle preparation and colloidal characterization. Colloid Polym. Sci. 2003, 281, 1184–1190. [Google Scholar] [CrossRef]
  221. Surassmo, S.; Min, S.-G.; Bejrapha, P.; Choi, M.-J. Effects of surfactants on the physical properties of capsicum oleoresin-loaded nanocapsules formulated through the emulsion–diffusion method. Food Res. Int. 2010, 43, 8–17. [Google Scholar] [CrossRef]
  222. Sahana, D.; Mittal, G.; Bhardwaj, V.; Kumar, M. PLGA Nanoparticles for Oral Delivery of Hydrophobic Drugs: Influence of Organic Solvent on Nanoparticle Formation and Release Behavior In Vitro and In Vivo Using Estradiol as a Model Drug. J. Pharm. Sci. 2008, 97, 1530–1542. [Google Scholar] [CrossRef]
  223. Jain, A.K.; Swarnakar, N.K.; Godugu, C.; Singh, R.; Jain, S. The effect of the oral administration of polymeric nanoparticles on the efficacy and toxicity of tamoxifen. Biomaterials 2011, 32, 503–515. [Google Scholar] [CrossRef]
  224. Fonseca, A.C.; Ferreira, P.; Cordeiro, R.A.; Mendonça, P.V.; Góis, J.R.; Gil, M.H.; Coelho, J.F.J. Drug Delivery Systems for Predictive Medicine: Polymers as Tools for Advanced Applications. In New Strategies to Advance Pre/Diabetes Care: Integrative Approach by PPPM; Springer: Dordrecht, The Netherlands, 2013; pp. 399–455. [Google Scholar] [CrossRef]
  225. Zhang, J.; Xiong, X. Salting-Out Assisted Liquid–Liquid Extraction (SALLE) in LC-MS Bioanalysis. In Sample Preparation in LC-MS Bioanalysis; Li, W., Jian, W., Fu, Y., Eds.; Wiley: New York, NY, USA, 2019; Chapter 5; pp. 68–75. [Google Scholar] [CrossRef]
  226. De, A.; Bose, R.; Kumar, A.; Mozumdar, S. Targeted Delivery of Pesticides Using Biodegradable Polymeric Nanoparticles; Springer: Cham, Switzerland, 2014. [Google Scholar]
  227. Li, T.; Ma, L.; Sun, D.; Liu, L.; Qayum, A.; Jiang, Z.; Hou, J. Purification of lactoperoxidase from bovine milk by integrating the technique of salting-out extraction with cation exchange chromatographic separation. J. Food Meas. Charact. 2019, 13, 1400–1410. [Google Scholar] [CrossRef]
  228. Eley, J.G.; Pujari, V.D.; McLane, J. Poly (Lactide-co-Glycolide) Nanoparticles Containing Coumarin-6 for Suppository Delivery: In Vitro Release Profile and In Vivo Tissue Distribution. Drug Deliv. 2004, 11, 255–261. [Google Scholar] [CrossRef]
  229. Grumezescu, A.M. Drug Targeting and Stimuli Sensitive Drug Delivery Systems; William Andrew: London, UK, 2018. [Google Scholar]
  230. Niknafs, M.; Kaviani, R.; Gharekhani, A.; Jouyban, A.; Shayanfar, A. Salting-out liquid–liquid microextraction to the determination of mycophenolic acid in plasma samples. Chem. Pap. 2019, 74, 1663–1668. [Google Scholar] [CrossRef]
  231. Shaik, M.; Shivanna, D.K.; Kamate, M.; AB, V.; Tp, K.-V. Single Lysis-Salting Out Method of Genomic DNA Extraction From Dried Blood Spots. J. Clin. Lab. Anal. 2016, 30, 1009–1012. [Google Scholar] [CrossRef] [Green Version]
  232. Fujii, T.; Kawasaki, S.-I. Salting-out effects on vanillin extraction by supercritical carbon dioxide from aqueous vanillin solution containing salts. J. Supercrit. Fluids 2019, 152, 104550. [Google Scholar] [CrossRef]
  233. Fan, Y.F.; Wang, Y.N.; Ma, J.B. Preparation of insulin nanoparticles and their encapsulation with biodegradable polyelectrolytes via the layer-by-layer adsorption. Int. J. Pharm. 2006, 324, 158–167. [Google Scholar] [CrossRef] [PubMed]
  234. Jaeghere, F.; Allémann, E.; Feijen, J.; Kissel, T.; Doelker, E.; Gurny, R. Cellular Uptake of PEO Surface-Modified Nanoparticles: Evaluation of Nanoparticles Made of PLA: PEO Diblock and Triblock Copolymers. J. Drug Target. 2000, 8, 143–153. [Google Scholar] [CrossRef] [PubMed]
  235. Lim, K.; Hamid, Z.A. Polymer Nanoparticle Carriers in Drug Delivery Systems: Research Trend. In Applications of Nanocomposite Materials in Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2018; Chapter 10; pp. 217–237. [Google Scholar] [CrossRef]
  236. Dubey, S.; Mody, N.; Sharma, R.; Agrawal, U.; Vyas, S.P. Nanobiomaterials: Novel Nanoplatforms for Protein and Peptide Delivery. In Nanobiomaterials in Drug Delivery; Elsevier: Amsterdam, The Netherlands, 2016; Chapter 4; pp. 111–146. [Google Scholar] [CrossRef]
  237. Zhang, Z.; Grijpma, D.W.; Feijen, J. Poly(trimethylene carbonate) and monomethoxy poly(ethylene glycol)-block-poly(trimethylene carbonate) nanoparticles for the controlled release of dexamethasone. J. Control. Release 2006, 111, 263–270. [Google Scholar] [CrossRef] [PubMed]
  238. Galindo-Rodriguez, S.; Allémann, E.; Fessi, H.; Doelker, E. Physicochemical Parameters Associated with Nanoparticle Formation in the Salting-Out, Emulsification-Diffusion, and Nanoprecipitation Methods. Pharm. Res. 2004, 21, 1428–1439. [Google Scholar] [CrossRef]
  239. Krishnasailaja, A.; Sarithareddy, A.; Abbaraju, K.; Aenugu, S. Preparation and Characterisation of Sulfasalazine Loaded Polymeric Nanoparticles by Salting Out Technique. J. Bionanosci. 2017, 11, 17–23. [Google Scholar] [CrossRef]
  240. Song, X.; Zhao, Y.; Wu, W.; Bi, Y.; Cai, Z.; Chen, Q.; Li, Y.; Hou, S. PLGA nanoparticles simultaneously loaded with vincristine sulfate and verapamil hydrochloride: Systematic study of particle size and drug entrapment efficiency. Int. J. Pharm. 2008, 350, 320–329. [Google Scholar] [CrossRef]
  241. Konan, Y.N.; Gurny, R.; Allémann, E. Preparation and characterization of sterile and freeze-dried sub-200 nm nanoparticles. Int. J. Pharm. 2001, 233, 239–252. [Google Scholar] [CrossRef]
  242. Gazi, A.S.; Sailaja, A.K. Preparation and Characterization of Paracetamol Loaded Eudragit S100 Nanoparticles by Salting Out Technique. J. Dev. Drugs 2018, 7, 1–4. [Google Scholar] [CrossRef]
  243. Zweers, M.L.; Engbers, G.H.; Grijpma, D.W.; Feijen, J. Release of anti-restenosis drugs from poly(ethylene oxide)-poly(dl-lactic-co-glycolic acid) nanoparticles. J. Control. Release 2006, 114, 317–324. [Google Scholar] [CrossRef]
  244. Zweers, M.L.; Engbers, G.H.; Grijpma, D.W.; Feijen, J. In vitro degradation of nanoparticles prepared from polymers based on dl-lactide, glycolide and poly(ethylene oxide). J. Control. Release 2004, 100, 347–356. [Google Scholar] [CrossRef]
  245. Nguyen, C.A.; Allémann, E.; Schwach, G.; Doelker, E.; Gurny, R. Synthesis of A Novel Fluorescent Poly (D, L-Lactide) End-Capped with 1-Pyrenebutanol Used for The Preparation of Nanoparticles. Eur. J. Pharm. Sci. 2003, 20, 217–222. [Google Scholar] [CrossRef]
  246. Chern, C. Emulsion polymerization mechanisms and kinetics. Prog. Polym. Sci. 2006, 31, 443–486. [Google Scholar] [CrossRef]
  247. Verma, G.; Rajagopalan, M.D.; Valluru, R.; Sridhar, K.A. Nanoparticles: A Novel Approach to Target Tumors. In Nano-and Microscale Drug Delivery Systems; Elsevier: Amsterdam, The Netherlands, 2017; Chapter 7; pp. 113–129. [Google Scholar]
  248. Yang, Y.; Fang, Z.; Chen, X.; Zhang, W.; Xie, Y.; Chen, Y.; Liu, Z.; Yuan, W. An Overview of Pickering Emulsions: Solid-Particle Materials, Classification, Morphology, and Applications. Front. Pharmacol. 2017, 8, 287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Guerra, F.E.B. Emulsion Polymerization of Superhydrophobic Monomers. Ph.D. Thesis, University of Basque Country, Donostia-San Sebastian, Spain, 2017. Available online: https://addi.ehu.es/bitstream/handle/10810/24229/TESIS_BOSCAN_GUERRA_FREDDY%20ENRIQUE.pdf?sequence=1 (accessed on 20 November 2021).
  250. Khan, M.U.; Reddy, K.R.; Snguanwongchai, T.; Haque, E.; Gomes, V.G. Polymer brush synthesis on surface modified carbon nanotubes via in situ emulsion polymerization. Colloid Polym. Sci. 2016, 294, 1599–1610. [Google Scholar] [CrossRef]
  251. Gharieh, A.; Khoee, S.; Mahdavian, A.R. Emulsion and miniemulsion techniques in preparation of polymer nanoparticles with versatile characteristics. Adv. Colloid Interface Sci. 2019, 269, 152–186. [Google Scholar] [CrossRef] [PubMed]
  252. Muñoz-Bonilla, A.; van Herk, A.M.; Heuts, J.P. Preparation of Hairy Particles and Antifouling Films using Brush-Type Amphiphilic Block Copolymer Surfactants in Emulsion Polymerization. Macromolecules 2010, 43, 2721–2731. [Google Scholar] [CrossRef]
  253. Garay-Jimenez, J.C.; Gergeres, D.; Young, A.; Lim, D.V.; Turos, E. Physical properties and biological activity of poly(butyl acrylate–styrene) nanoparticle emulsions prepared with conventional and polymerizable surfactants. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 443–451. [Google Scholar] [CrossRef] [Green Version]
  254. Lu, S.; Qu, R.; Forcada, J. Preparation of magnetic polymeric composite nanoparticles by seeded emulsion polymerization. Mater. Lett. 2009, 63, 770–772. [Google Scholar] [CrossRef]
  255. Gao, J.; Wu, C. Modified Structural Model for Predicting Particle Size in the Microemulsion and Emulsion Polymerization of Styrene under Microwave Irradiation. Langmuir 2004, 21, 782–785. [Google Scholar] [CrossRef]
  256. Costa, C.; Santos, A.; Fortuny, M.; Araújo, P.; Sayer, C. Kinetic advantages of using microwaves in the emulsion polymerization of MMA. Mater. Sci. Eng. C 2009, 29, 415–419. [Google Scholar] [CrossRef]
  257. Zhang, J.; Cao, Y.; He, Y. Ultrasonically irradiated emulsion polymerization of styrene in the presence of a polymeric surfactant. J. Appl. Polym. Sci. 2004, 94, 763–768. [Google Scholar] [CrossRef]
  258. Yoon, S.-J.; Chun, H.; Lee, M.-S.; Kim, N. Preparation of poly(N-vinylcarbazole) (PVK) nanoparticles by emulsion polymerization and PVK hollow particles. Synth. Met. 2009, 159, 518–522. [Google Scholar] [CrossRef]
  259. Bourgeat-Lami, E.; França, A.J.P.G.; Chaparro, T.C.; Silva, R.D.; Dugas, P.-Y.; Alves, G.M.; Santos, A.M. Synthesis of Polymer/Silica Hybrid Latexes by Surfactant-Free RAFT-Mediated Emulsion Polymerization. Macromolecules 2016, 49, 4431–4440. [Google Scholar] [CrossRef]
  260. Errezma, M.; Ben Mabrouk, A.; Magnin, A.; Dufresne, A.; Boufi, S. Surfactant-free emulsion Pickering polymerization stabilized by aldehyde-functionalized cellulose nanocrystals. Carbohydr. Polym. 2018, 202, 621–630. [Google Scholar] [CrossRef] [PubMed]
  261. Farias-Cepeda, L.; Herrera-Ordonez, J.; Estevez, M.; Luna-Barcenas, G.; Rosales-Marines, L. New Insights on surfactant-free styrene emulsion polymerization in The presence of sodium styrene sulfonate. Colloid Polym. Sci. 2016, 294, 1571–1576. [Google Scholar] [CrossRef]
  262. Sahiner, N.; Sengel, S.B. Surfactant Free Synthesis and Characterization of Poly(Vinyl Carbazole) Microgel and its Chemical Modifications. Colloids Surf. A Physicochem. Eng. Asp. 2017, 514, 243–250. [Google Scholar] [CrossRef]
  263. Heo, H.J.; Park, I.J.; Lee, S.G.; Ha, J.-W.; Lee, S.-B.; Sohn, E.-H. Surfactant-free preparation of poly(vinylidene fluoride) nanoparticle dispersions and their use as surface coating agents. Green Chem. 2017, 20, 502–505. [Google Scholar] [CrossRef]
  264. Kassim, S.; Zahari, S.B.; Tahrin, R.A.A.; Harun, N.A. Co-polymerization of methyl methacrylate and styrene via surfactant-free emulsion polymerization, as a potential material for photonic crystal application. AIP Conf. Proc. 2017, 1885, 20018. [Google Scholar] [CrossRef]
  265. Chen, Z.; Zhao, Y.; Zhao, Y.; Thomas, H.; Zhu, X.; Möller, M. Inclusion of Phase-Change Materials in Submicron Silica Capsules Using a Surfactant-Free Emulsion Approach. Langmuir 2018, 34, 10397–10406. [Google Scholar] [CrossRef]
  266. An, Z.; Tang, W.; Hawker, C.J.; Stucky, G.D. One-Step Microwave Preparation of Well-Defined and Functionalized Polymeric Nanoparticles. J. Am. Chem. Soc. 2006, 128, 15054–15055. [Google Scholar] [CrossRef] [PubMed]
  267. Chiu, T.; Don, T. Synthesis and characterization of poly(methyl methacrylate) nanoparticles by emulsifier-free emulsion polymerization with a redox-initiated system. J. Appl. Polym. Sci. 2008, 109, 3622–3630. [Google Scholar] [CrossRef]
  268. Fang, F.F.; Kim, J.H.; Choi, H.J.; Kim, C.A. Synthesis and electrorheological response of nano-sized laponite stabilized poly(methyl methacrylate) spheres. Colloid Polym. Sci. 2009, 287, 745–749. [Google Scholar] [CrossRef]
  269. Camli, S.T.; Buyukserin, F.; Balci, O.; Budak, G.G. Size controlled synthesis of sub-100nm monodisperse poly(methylmethacrylate) nanoparticles using surfactant-free emulsion polymerization. J. Colloid Interface Sci. 2010, 344, 528–532. [Google Scholar] [CrossRef]
  270. Cui, X.; Zhong, S.; Wang, H. Emulsifier-free core–shell polyacrylate latex nanoparticles containing fluorine and silicon in shell. Polymer 2007, 48, 7241–7248. [Google Scholar] [CrossRef]
  271. Faridi-Majidi, R.; Sharifi-Sanjani, N. Emulsifier-free miniemulsion polymerization of styrene and the investigation of encapsulation of nanoparticles with polystyrene via this procedure using an anionic initiator. J. Appl. Polym. Sci. 2007, 105, 1244–1250. [Google Scholar] [CrossRef]
  272. Öztürk, N.; Bereli, N.; Akgöl, S.; Denizli, A. High capacity binding of antibodies by poly(hydroxyethyl methacrylate) nanoparticles. Colloids Surf. B Biointerfaces 2008, 67, 14–19. [Google Scholar] [CrossRef]
  273. Lee, J.M.; Lee, S.J.; Jung, Y.J.; Kim, J.H. Fabrication of nano-structured polythiophene nanoparticles in aqueous dispersion. Curr. Appl. Phys. 2008, 8, 659–663. [Google Scholar] [CrossRef]
  274. Frank, H.; Ziener, U.; Landfester, K. Formation of Polyimide Nanoparticles in Heterophase with an Ionic Liquid as Continuous Phase. Macromolecules 2009, 42, 7846–7853. [Google Scholar] [CrossRef]
  275. Kim, S.W.; Cho, H.G.; Park, C.R. Fabrication of Unagglomerated Polypyrrole Nanospheres with Controlled Sizes From a Surfactant-Free Emulsion System. Langmuir 2009, 25, 9030–9036. [Google Scholar] [CrossRef]
  276. Engström, J.; Brett, C.J.; Körstgens, V.; Müller-Buschbaum, P.; Ohm, W.; Malmström, E.; Roth, S.V. Core–Shell Nanoparticle Interface and Wetting Properties. Adv. Funct. Mater. 2020, 30, 1907720. [Google Scholar] [CrossRef]
  277. Kedzior, S.A.; Marway, H.S.; Cranston, E.D. Tailoring Cellulose Nanocrystal and Surfactant Behavior in Miniemulsion Polymerization. Macromolecules 2017, 50, 2645–2655. [Google Scholar] [CrossRef]
  278. Nauman, N.; Zaquen, N.; Junkers, T.; Boyer, C.; Zetterlund, P.B. Particle Size Control in Miniemulsion Polymerization via Membrane Emulsification. Macromolecules 2019, 52, 4492–4499. [Google Scholar] [CrossRef]
  279. Li, W.S.J.; Negrell, C.; Ladmiral, V.; Lai-Kee-Him, J.; Bron, P.; Lacroix-Desmazes, P.; Joly-Duhamel, C.; Caillol, S. Cardanol-based polymer latex by radical aqueous miniemulsion polymerization. Polym. Chem. 2018, 9, 2468–2477. [Google Scholar] [CrossRef]
  280. Bao, Z.; Smith, K.W. Miniemulsion Polymerization to Prepare Drag Reducers. US20110184121A1, 24 April 2018. [Google Scholar]
  281. Landfester, K.; Antonietti, M. The polymerization of acrylonitrile in miniemulsions: “Crumpled latex particles” or polymer nanocrystals. Macromol. Rapid Commun. 2000, 21, 820–824. [Google Scholar] [CrossRef]
  282. Landfester, K.; Willert, M.; Antonietti, M. Preparation of Polymer Particles in Nonaqueous Direct and Inverse Miniemulsions. Macromolecules 2000, 33, 2370–2376. [Google Scholar] [CrossRef]
  283. Chern, C.-S.; Sheu, J.-C. Effects of carboxylic monomers on the styrene miniemulsion polymerizations stabilized by SDS/alkyl methacrylates. Polymer 2001, 42, 2349–2357. [Google Scholar] [CrossRef]
  284. Arias, J.L.; Gallardo, V.; Gómez-Lopera, S.A.; Plaza, R.; Delgado, A. Synthesis and characterization of poly(ethyl-2-cyanoacrylate) nanoparticles with a magnetic core. J. Control. Release 2001, 77, 309–321. [Google Scholar] [CrossRef]
  285. Ham, H.T.; Choi, Y.S.; Chee, M.G.; Chung, I.J. Singlewall carbon nanotubes covered with polystyrene nanoparticles byin-situ miniemulsion polymerization. J. Polym. Sci. Part A Polym. Chem. 2005, 44, 573–584. [Google Scholar] [CrossRef]
  286. Ziegler, A.; Landfester, K.; Musyanovych, A. Synthesis of phosphonate-functionalized polystyrene and poly(methyl methacrylate) particles and their kinetic behavior in miniemulsion polymerization. Colloid Polym. Sci. 2009, 287, 1261–1271. [Google Scholar] [CrossRef] [Green Version]
  287. van Herk, A.; Forcada, J.; Pastorin, G. Synthetic Strategies for Synthesis of Polymer Nanoparticles. In Controlled Release Systems; Jenny Stanford Publishing: Boca Raton, FL, USA, 2016; Chapter 5; pp. 61–78. [Google Scholar] [CrossRef]
  288. Ponzio, R.A.; Marcato, Y.L.; Gomez, M.L.; Waiman, C.V.; Chesta, C.A.; Palacios, R.E. Crosslinked polymer nanoparticles containing single conjugated polymer chains. Methods Appl. Fluoresc. 2017, 5, 024001. [Google Scholar] [CrossRef]
  289. Ghayempour, S.; Montazer, M. A modified microemulsion method for fabrication of hydrogel Tragacanth nanofibers. Int. J. Biol. Macromol. 2018, 115, 317–323. [Google Scholar] [CrossRef]
  290. Cankaya, N. Recent Research in Polymerization; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef]
  291. Puig, J.E.; Rabelero, M. Semicontinuous Microemulsion Polymerization. Curr. Opin. Colloid Interface Sci. 2016, 25, 83–88. [Google Scholar] [CrossRef]
  292. Candau, F.; Anquetil, J.-Y. New Developments in Polymerization in Bicontinuous Microemulsions. In Micelles, Microemulsions, and Monolayers; Routledge: Oxfordshire, UK, 2018; pp. 193–213. [Google Scholar] [CrossRef]
  293. Sarov, Y.; Capek, I. Kinetic events of (micro)emulsion polymerization of styrene. Polym. Bull. 2019, 77, 4851–4865. [Google Scholar] [CrossRef]
  294. Guerrero-Ramírez, L.G.; Nuño-Donlucas, S.M.; Cesteros, L.C.; Katime, I. Novel Functionalized Nanohydrogels, Synthesis and Some Applications. J. Phys. Conf. Ser. IOP Publ. 2008, 127, 012010. [Google Scholar] [CrossRef]
  295. Sosa, N.; Zaragoza, E.A.; Lopez, R.G.; Peralta, R.D.; Katime, I.; Becerra, F.; Mendizabal, E.; Puig, J.E. Unusual Free Radical Polymerization of Vinyl Acetate in Anionic Microemulsion Media. Langmuir 2000, 16, 3612–3619. [Google Scholar] [CrossRef]
  296. Barrère, M.; da Silva, S.C.; Balic, R.; Ganachaud, F. Synthesis of Monodisperse Poly(dimethylsiloxane) Micro- and Macroemulsions. Langmuir 2001, 18, 941–944. [Google Scholar] [CrossRef]
  297. Hermanson, K.D.; Kaler, E.W. Kinetics and Mechanism of the Multiple Addition Microemulsion Polymerization of Hexyl Methacrylate. Macromolecules 2003, 36, 1836–1842. [Google Scholar] [CrossRef]
  298. Xu, X.-J.; Chen, F. Semi-Continuous Emulsion Copolymerization of Butyl Methacrylate with Polymerizable Anionic Surfactants. Polymer 2004, 45, 4801–4810. [Google Scholar] [CrossRef]
  299. Jang, J.; Oh, J.H.; Stucky, G.D. Fabrication of Ultrafine Conducting Polymer and Graphite Nanoparticles. Angew. Chem. Int. Ed. 2002, 41, 4016–4019. [Google Scholar] [CrossRef]
  300. Ramírez, A.G.; López, R.G.; Tauer, K. Studies on Semibatch Microemulsion Polymerization of Butyl Acrylate: Influence of the Potassium Peroxodisulfate Concentration. Macromolecules 2004, 37, 2738–2747. [Google Scholar] [CrossRef]
  301. Song, R.; Murphy, M.; Li, C.; Ting, K.; Soo, C.; Zheng, Z. Current development of biodegradable polymeric materials for biomedical applications. Drug Des. Dev. Ther. 2018, 12, 3117–3145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Nguyen, P.Q.; Courchesne, N.M.D.; Duraj-Thatte, A.; Praveschotinunt, P.; Joshi, N.S. Engineered Living Materials: Prospects and Challenges for Using Biological Systems to Direct the Assembly of Smart Materials. Adv. Mater. 2018, 30, 1704847. [Google Scholar] [CrossRef] [PubMed]
  303. Stanley, S. Biological nanoparticles and their influence on organisms. Curr. Opin. Biotechnol. 2014, 28, 69–74. [Google Scholar] [CrossRef]
  304. Pertici, G. Introduction to Bioresorbable Polymers for Biomedical Applications. In Bioresorbable polymers for Biomedical Applications; Elsevier: Amsterdam, The Netherlands, 2017; Chapter 1; pp. 3–29. [Google Scholar] [CrossRef]
  305. Hiroe, A.; Chek, M.F.; Hakoshima, T.; Sudesh, K.; Taguchi, S. Synthesis of Polyesters III: Acyltransferase as Catalyst. In Enzymatic Polymerization towards Green Polymer Chemistry; Springer: Singapore, 2019; pp. 199–231. [Google Scholar] [CrossRef]
  306. Zain, N.-A.A.; Ng, L.-M.; Foong, C.P.; Tai, Y.T.; Nanthini, J.; Sudesh, K. Complete Genome Sequence of a Novel Polyhydroxyalkanoate (PHA) Producer, Jeongeupia sp. USM3 (JCM 19920) and Characterization of Its PHA Synthases. Curr. Microbiol. 2020, 77, 500–508. [Google Scholar] [CrossRef] [PubMed]
  307. Kourmentza, C.; Plácido, J.; Venetsaneas, N.; Burniol-Figols, A.; Varrone, C.; Gavala, H.N.; Reis, M.A.M. Recent Advances and Challenges towards Sustainable Polyhydroxyalkanoate (PHA) Production. Bioengineering 2017, 4, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Mukheem, A.; Hossain, M.; Shahabuddin, S.; Muthoosamy, K.; Manickam, S.; Sudesh, K.; Saidur, R.; Sridewi, N.; Campus, N.M. Bioplastic Polyhydroxyalkanoate (PHA): Recent Advances in Modification and Medical Applications. arXiv 2018, arXiv:10.20944/preprints201808.0271.v1. Available online: https://europepmc.org/article/ppr/ppr48118 (accessed on 30 August 2021).
  309. Chek, M.F.; Hiroe, A.; Hakoshima, T.; Sudesh, K.; Taguchi, S. PHA synthase (PhaC): Interpreting the functions of bioplastic-producing enzyme from a structural perspective. Appl. Microbiol. Biotechnol. 2018, 103, 1131–1141. [Google Scholar] [CrossRef]
  310. Maestro, B.; Sanz, J.M. Polyhydroxyalkanoate-Associated Phasins as Phylogenetically Heterogeneous, Multipurpose Proteins. Microb. Biotechnol. 2017, 10, 1323–1337. [Google Scholar] [CrossRef]
  311. Hokamura, A.; Fujino, K.; Isoda, Y.; Arizono, K.; Shiratsuchi, H.; Matsusaki, H. Characterization and identification of the proteins bound to two types of polyhydroxyalkanoate granules in Pseudomonas sp. 61-3. Biosci. Biotechnol. Biochem. 2015, 79, 1369–1377. [Google Scholar] [CrossRef] [Green Version]
  312. Dinjaski, N.; Prieto, M.A. Smart polyhydroxyalkanoate nanobeads by protein based functionalization. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 885–899. [Google Scholar] [CrossRef] [PubMed]
  313. Rodríguez-Carmona, E.; Villaverde, A. Nanostructured bacterial materials for innovative medicines. Trends Microbiol. 2010, 18, 423–430. [Google Scholar] [CrossRef]
  314. Koller, M.J.M. Biodegradable and Biocompatible Polyhydroxy-alkanoates (PHA): Auspicious Microbial Macromolecules for Pharmaceutical and Therapeutic Applications. Molecules 2018, 23, 362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Prakash, P.; Lee, W.-H.; Loo, C.-Y.; Wong, H.S.J.; Parumasivam, T. Advances in Polyhydroxyalkanoate Nanocarriers for Effective Drug Delivery: An Overview and Challenges. Nanomaterials 2022, 12, 175. [Google Scholar] [CrossRef] [PubMed]
  316. Williams, S.F.; Martin, D.P. Applications of Polyhydroxyalkanoates (PHA) in Medicine and Pharmacy. Biopolym 2002. [Google Scholar] [CrossRef]
  317. Yao, Y.-C.; Zhan, X.-Y.; Zhang, J.; Zou, X.-H.; Wang, Z.-H.; Xiong, Y.-C.; Chen, J.; Chen, G.-Q. A specific drug targeting system based on polyhydroxyalkanoate granule binding protein PhaP fused with targeted cell ligands. Biomaterials 2008, 29, 4823–4830. [Google Scholar] [CrossRef] [PubMed]
  318. Errico, C.; Bartoli, C.; Chiellini, F.; Chiellini, E. Poly(hydroxyalkanoates)-Based Polymeric Nanoparticles for Drug Delivery. J. Biomed. Biotechnol. 2009, 2009, 1–10. [Google Scholar] [CrossRef]
  319. Shishatskaya, E.I.; Goreva, A.V.; Voinova, O.N.; Inzhevatkin, E.V.; Khlebopros, R.G.; Volova, T.G. Evaluation of antitumor activity of rubomycin deposited in absorbable polymeric microparticles. Bull. Exp. Biol. Med. 2008, 145, 358–361. [Google Scholar] [CrossRef]
  320. Zhang, C.; Zhao, L.; Dong, Y.; Zhang, X.; Lin, J.; Chen, Z. Folate-mediated poly(3-hydroxybutyrate-co-3-hydroxyoctanoate) nanoparticles for targeting drug delivery. Eur. J. Pharm. Biopharm. 2010, 76, 10–16. [Google Scholar] [CrossRef]
  321. Gursel, I.; Yagmurlu, F.; Korkusuz, F.; Hasirci, V. In vitro antibiotic release from poly(3-hydroxybutyrate-co-3-hydroxyvalerate) rods. J. Microencapsul. 2002, 19, 153–164. [Google Scholar] [CrossRef]
  322. Rossi, S.; Azghani, A.O.; Omri, A. Antimicrobial efficacy of a new antibiotic-loaded poly(hydroxybutyric-co-hydroxyvaleric acid) controlled release system. J. Antimicrob. Chemother. 2004, 54, 1013–1018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Wang, Z.; Itoh, Y.; Hosaka, Y.; Kobayashi, I.; Nakano, Y.; Maeda, I.; Umeda, F.; Yamakawa, J.; Nishimine, M.; Suenobu, T. Mechanism of Enhancement Effect of Dendrimer on Transdermal Drug Permeation Through Polyhydroxyalkanoate Matrix. J. Biosci. Bioeng. 2003, 96, 537–540. [Google Scholar] [CrossRef]
Figure 1. The nanoprecipitation process illustrated in a diagram. The enlarged image (inset) illustrates the process of nanoparticle (yellow spheres) formation owing to the surface tension difference between the aqueous phase (high surface tension) and organic phase (low surface tension). Adapted from Wang et al. (2016) [78].
Figure 1. The nanoprecipitation process illustrated in a diagram. The enlarged image (inset) illustrates the process of nanoparticle (yellow spheres) formation owing to the surface tension difference between the aqueous phase (high surface tension) and organic phase (low surface tension). Adapted from Wang et al. (2016) [78].
Nanomaterials 12 00576 g001
Figure 2. Scanning electron micrographs of synthesized P(3HB) NPs. (ad) NPs were prepared using chloroform and (a) DMSO, (b) DMSO (loaded), (c) water (d) water (loaded), (eh) ethyl acetate, and (e) DMSO, (f) DMSO (loaded) (g) water, (h) water (loaded). Adapted from Senthilkumar et al. (2018) [174].
Figure 2. Scanning electron micrographs of synthesized P(3HB) NPs. (ad) NPs were prepared using chloroform and (a) DMSO, (b) DMSO (loaded), (c) water (d) water (loaded), (eh) ethyl acetate, and (e) DMSO, (f) DMSO (loaded) (g) water, (h) water (loaded). Adapted from Senthilkumar et al. (2018) [174].
Nanomaterials 12 00576 g002
Figure 3. The emulsification-solvent evaporation technique is depicted schematically. Adapted from Wang et al. (2016) [78].
Figure 3. The emulsification-solvent evaporation technique is depicted schematically. Adapted from Wang et al. (2016) [78].
Nanomaterials 12 00576 g003
Figure 4. Diagrammatic representation of the emulsification solvent diffusion method.
Figure 4. Diagrammatic representation of the emulsification solvent diffusion method.
Nanomaterials 12 00576 g004
Figure 5. Diagrammatic interpretation of the salting-out technique.
Figure 5. Diagrammatic interpretation of the salting-out technique.
Nanomaterials 12 00576 g005
Figure 6. Schematic representation of the emulsion polymerization method.
Figure 6. Schematic representation of the emulsion polymerization method.
Nanomaterials 12 00576 g006
Figure 7. Diagram depiction of the mini-emulsion polymerization method.
Figure 7. Diagram depiction of the mini-emulsion polymerization method.
Nanomaterials 12 00576 g007
Figure 8. TEM image of nanosized PHA granules inside bacterial cells.
Figure 8. TEM image of nanosized PHA granules inside bacterial cells.
Nanomaterials 12 00576 g008
Table 1. Advantages and limitations of two types of polymer synthesis methods; dispersion of preformed polymer and direct polymerization.
Table 1. Advantages and limitations of two types of polymer synthesis methods; dispersion of preformed polymer and direct polymerization.
MethodAdvantagesLimitations References
Dispersion of preformed polymers
nanoprecipitationRequires low energy
Reproducible
Single step
Scalability
Size of NPs can be affected by stirring rate
Low efficiency of drug encapsulation
[78,79]
emulsification solvent evaporationScalability
Single step emulsion for hydrophobic agents
Double or multiple step emulsion for hydrophilic agents
Requires heating or vacuum for evaporation
Residual solvent or stabilizer
Not stable
[80,81]
emulsification solvent diffusionDoes not require homogenizer
High reproducibility
Easy to scale up
Uses high volumes of water
Probable leakage of water-soluble drugs into external phase
Lower efficiency in lipophilic drug encapsulation
[82,83]
salting outDoes not require heating
Avoids chlorinated solvents
Suitable for DNA, RNA, and proteins
Requires high speed homogenization
Exclusive for the encapsulation of lipophilic drugs
Time-consuming
Limited scalability
[84,85]
Direct polymerization
emulsionProduce polymers with high molar mass
Uses water as dispersion medium
Excellent heat dissipation
Requires removal of surfactant
Time consuming
High cost
[86,87]
surfactant-free emulsionDoes not require surfactant
Simple and green process
Uses water-soluble initiators
Requires the preparation of monodisperse and uniformly distributed particle sizes[88,89]
mini emulsionUses a low molecular mass co-stabilizer
Small particle size
Low volume of surfactant
Uses a high-shear device
Surfactant is retained in the polymer
[90,91]
micro-emulsionUses water-soluble initiators
Thermodynamically stable
Formation of empty micelles
Destabilized microdroplets
Increased particle size
Requires a high ratio of surfactant
[92,93]
microbialNon-toxic
Eco-friendly
Biocompatible
High production cost[94,95]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pulingam, T.; Foroozandeh, P.; Chuah, J.-A.; Sudesh, K. Exploring Various Techniques for the Chemical and Biological Synthesis of Polymeric Nanoparticles. Nanomaterials 2022, 12, 576. https://doi.org/10.3390/nano12030576

AMA Style

Pulingam T, Foroozandeh P, Chuah J-A, Sudesh K. Exploring Various Techniques for the Chemical and Biological Synthesis of Polymeric Nanoparticles. Nanomaterials. 2022; 12(3):576. https://doi.org/10.3390/nano12030576

Chicago/Turabian Style

Pulingam, Thiruchelvi, Parisa Foroozandeh, Jo-Ann Chuah, and Kumar Sudesh. 2022. "Exploring Various Techniques for the Chemical and Biological Synthesis of Polymeric Nanoparticles" Nanomaterials 12, no. 3: 576. https://doi.org/10.3390/nano12030576

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