Next Article in Journal
Oral Bioavailability of Kinsenoside in Beagle Dogs Measured by LC-MS/MS: Improvement of Ex Vivo Stability of a Lactone-Containing Compound
Next Article in Special Issue
Ginsenoside Drug Nanocomposites Prepared by the Aerosol Solvent Extraction System for Enhancing Drug Solubility and Stability
Previous Article in Journal
A Novel Desloratadine-Benzoic Acid Co-Amorphous Solid: Preparation, Characterization, and Stability Evaluation
Previous Article in Special Issue
Solubility Improvement of Benexate through Salt Formation Using Artificial Sweetener
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bioavailability Enhancement of Poorly Water-Soluble Drugs via Nanocomposites: Formulation–Processing Aspects and Challenges

Otto H. York Department of Chemical and Materials Engineering, New Jersey Institute of Technology, Newark, NJ 07102, USA
*
Author to whom correspondence should be addressed.
Submission received: 19 June 2018 / Revised: 1 July 2018 / Accepted: 1 July 2018 / Published: 8 July 2018
(This article belongs to the Special Issue Dissolution Enhancement of Poorly Soluble Drugs)

Abstract

:
Drug nanoparticles embedded in a dispersant matrix as a secondary phase, i.e., drug-laden nanocomposites, offer a versatile delivery platform for enhancing the dissolution rate and bioavailability of poorly water-soluble drugs. Drug nanoparticles are prepared by top-down, bottom-up, or combinative approaches in the form of nanosuspensions, which are subsequently dried to prepare drug-laden nanocomposites. In this comprehensive review paper, the term “nanocomposites” is used in a broad context to cover drug nanoparticle-laden intermediate products in the form of powders, cakes, and extrudates, which can be incorporated into final oral solid dosages via standard pharmaceutical unit operations, as well as drug nanoparticle-laden strip films. The objective of this paper is to review studies from 2012–2017 in the field of drug-laden nanocomposites. After a brief overview of the various approaches used for preparing drug nanoparticles, the review covers drying processes and dispersant formulations used for the production of drug-laden nanocomposites, as well as various characterization methods including quiescent and agitated redispersion tests. Traditional dispersants such as soluble polymers, surfactants, other water-soluble dispersants, and water-insoluble dispersants, as well as novel dispersants such as wet-milled superdisintegrants, are covered. They exhibit various functionalities such as drug nanoparticle stabilization, mitigation of aggregation, formation of nanocomposite matrix–film, wettability enhancement, and matrix erosion/disintegration. Major challenges such as nanoparticle aggregation and poor redispersibility that cause inferior dissolution performance of the drug-laden nanocomposites are highlighted. Literature data are analyzed in terms of usage frequency of various drying processes and dispersant classes. We provide some engineering considerations in comparing drying processes, which could account for some of the diverging trends in academia vs. industrial practice. Overall, this review provides rationale and guidance for drying process selection and robust nanocomposite formulation development, with insights into the roles of various classes of dispersants.

Graphical Abstract

1. Introduction

The number of newly developed drug molecules with greater lipophilicity, higher molecular weight, and poor water solubility has increased over the last few decades due to the emerging trends in combinatorial chemistry and drug design [1,2,3]. About 40% of drugs with market approval and nearly 90% of molecules in the discovery pipeline are poorly water-soluble [4]. The majority of failures in new drug development have been attributed to poor water solubility of the drug. It is well-known that poor solubility and slow dissolution can lead to low bioavailability, resulting in suboptimal drug delivery [5,6]. Commonly used approaches for enhancing the dissolution rate of these molecules include nanoparticle-based formulations [7,8], lipid-based drug delivery systems [9,10], pro-drugs [11,12], amorphous solid dispersions [13,14], salt formation [15,16], co-crystals [17,18], and cyclodextrin complexes [19,20].
Among the several approaches mentioned above, preparation of drug nanoparticles has been shown to be successful for improving the dissolution rate of a multitude of drugs, and 16 drugs have been marketed using drug nanoparticles (see Table 1). Drug nanoparticles have larger specific surface area and higher overall solute transfer coefficient than their micron-sized counterparts [21,22,23]. Moreover, ultrafine particles, especially those with sizes less than ~100 nm, tend to show higher saturation solubility, which can be explained via the Ostwald–Freundlich equation [24]. Overall, all these features exhibited by nanoparticles improve the dissolution rates according to the Noyes–Whiney equation [25]; this in turn enhances bioavailability [26,27]. Besides enhanced solubility and dissolution rates leading to improved bioavailability, other advantages of drug nanoparticles include the elimination of food effects, safe dose escalation, and enhanced efficacy and tolerability profiles [28,29,30].
Drug nanoparticles can be prepared in the form of suspensions, referred to as nanosuspensions, by top-down, bottom-up, or combinative methods. Top-down methods such as high-pressure homogenization (HPH) [32], stirred media milling [21,33,34], and ball milling [35] involve high shear–impact forces to achieve size reduction of coarse, as-received drug crystals down to micro or nanometer scale. Bottom-up methods involve building up particles by precipitation of dissolved molecules via liquid antisolvent precipitation (LASP) [36] and precipitation by supercritical fluids [37,38]. Melt emulsification is another example of bottom-up technique which can be used for drugs with low melting points. In this method, the drug is dispersed in an aqueous stabilizer solution and heated to melt crystals, followed by flash cooling to produce drug nanosuspensions [39,40]. Combinative methods [41] include a combination of bottom-up and top-down approaches.
Drug nanosuspensions must be physically stable during processing and storage for proper downstream processing or adequate shelf-life, depending on the intended final dosage form [23,28]. Important benefits resulting from high surface area can be lost if nanoparticles grow and/or form large clusters (aggregates). Moreover, suspensions that exhibit severe aggregation can pose significant downstream processing challenges due to high zero-shear viscosity and/or yield stress. As compared with microparticles, nanoparticles in a suspension show a strong tendency to aggregate because they have a high number concentration (given solid loading), small interparticle distances, enhanced Brownian motion, and relatively high surface energy [33,42,43,44]. During the preparation of drug nanosuspensions, turbulent mixing and high shear can cause faster aggregation if the nanosuspension is not properly stabilized upon use of various stabilizers [45,46]. The Brownian motion of nanoparticles may also contribute to the high collision rates during processing [44,47]. During the storage of the nanosuspensions, the Brownian motion is the major driving force for nanoparticle collisions besides gravity. Once the nanoparticles collide, they can aggregate due to van der Waals or hydrophobic forces, depending on their surface charge, which is quantified by zeta potential [48]. According to Müller [49], a zeta potential value of at least ±30 mV is required for an electrostatically stabilized suspension. About ±20 mV provides only a short-term stability, and values in the range of −5 mV to +5 mV indicate fast aggregation [50]. In the case of combined electrostatic and steric (electrosteric) stabilization, a minimum zeta potential of ±20 mV is desirable [51].
Stabilization of drug nanosuspensions can be achieved by electrostatic, steric, and electrosteric interactions of nanoparticle surfaces with adsorbing polymers and surfactants [48,52], also known as stabilizers. Most poorly water-soluble drugs exhibit hydrophobic behavior and cannot be well-dispersed in aqueous media without the addition of stabilizers, which also serve as wetting agents. On the other hand, many BCS Class II drugs exhibit finite solubility in the dispersion medium, which can be enhanced by the addition of stabilizers, especially surfactants. With increasing solubility, the particles may grow due to Ostwald ripening [53,54] especially if the suspensions are stored for a long time before down-stream processing such as filtration and drying.
While drug nanosuspensions can be used as oral suspensions and injectables, most marketed products are developed as oral solid dosage forms (see Table 1), because the latter are preferred by patients and doctors for their relative ease of administration, accurate dosing, and stability [28,55]. Moreover, despite the use of stabilizers, it is challenging to ensure the long-term physical stability of drug nanosuspensions. In fact, drying is generally perceived as a stabilization step for nanocrystals to avoid typical deterioration occurring in a liquid nanosuspension, such as Ostwald ripening, particle aggregation, sedimentation, and creaming [56,57]. For all the aforementioned reasons, drug nanosuspensions have been dried, as illustrated in Figure 1, via spray drying [33,34,58,59,60,61,62], fluid bed coating/granulation/drying [8,60,63,64,65,66], spray-freeze drying [7,67], freeze drying [68,69,70,71,72], vacuum drying [73,74], nanoextrusion [75,76,77,78], and wet casting–drying [40,79,80,81,82,83]. Drying processes convert drug nanosuspensions into nanocomposites that encapsulate or carry drug nanoparticles and their clusters dispersed as a secondary phase in the matrix of dispersants (stabilizers used in nanosuspensions and other excipients). Depending on the drying method, nanocomposites can be in the form of powders, cakes, or extrudates, which can be integrated into tablets, capsules, and sachets via standard pharmaceutical unit operations. Alternatively, they are in the form of polymeric strip films.
A major formulation challenge in dissolution enhancement upon use of drug-laden nanocomposites is that drug nanoparticles in nanocomposites may be released too slowly and/or in the form of large clusters (a.k.a. aggregates) during in vivo or vitro dissolution [66,84,85,86,87]. Besides the aggregation that may take place during the preparation/storage of drug nanosuspensions, drug nanoparticles can also aggregate into larger sub-micron clusters or even micron-sized clusters during the removal of water or solvents in the drying process, depending on the type/concentration of the dispersants [74,85]. Consequently, the advantages of drug nanoparticles with inherently large surface areas could be lost upon drying. The aggregates may be broadly classified as irreversible and reversible, as shown in Figure 2 [66,74,86], based on the redispersion behavior of dried nanosuspensions (nanocomposites) in liquids. Nanocomposite particles may contain aggregates of drug nanoparticles that have formed during the nanosuspension preparation step and/or drying step. Hard aggregates can be formed by the fusion of drug nanoparticles due to the removal of water/solvents during drying, especially when the dispersant concentration in the nanosuspension is too low (see e.g., [86]). They are most likely held together by solid bridges formed upon re-crystallization of some dissolved drug during drying. Agglomerates, another type of irreversible aggregates, could also form during drying. Although the exact mechanism leading to nanoparticle agglomeration is unknown [34], the capillary pressure theory is one theory that explains agglomeration due to the capillary forces encountered during the drying process [88]; others attributed agglomeration to polymer chain entanglement and/or potential micro-phase separation of polymeric stabilizer–other dispersants from particles upon increase in particle concentration with reduced water content [74,89,90]. Since irreversible aggregates do not redisperse back to primary nanoparticles, significant loss of drug surface area occurs, leading to inferior dissolution rate enhancement [66,85,91,92]. Unless otherwise indicated, aggregates in nanocomposites refer to irreversible aggregates in this paper.
A relevant concern is that even the reversible aggregates and primary drug nanoparticles in nanocomposites may be released too slowly from the dispersant matrix in aqueous media, which will lead to inferior dissolution rate and bioavailability. Not only does the dispersants’ type/concentration in the nanocomposite matrix affect the aggregation of drug nanoparticles in the nanosuspensions and dried nanocomposites, it also modulates the release of drug nanoparticles and their clusters, thus controlling the overall drug release rate [56,84]. Therefore, an understanding of nanoparticle recovery from nanocomposites after suspending nanocomposites in water (redispersion phenomenon) and its impact on drug dissolution rate is critically important [55,66,84,87,92]. The study of redispersion also sheds light on the functionalities of different classes of dispersants used in nanocomposites.
About 600 publications are available in the Scopus database on drug nanoparticles and drug nanocomposites, with growing interest over the past 7 years (Figure 3). Several excellent review papers are available on this topic. Chin et al. [55] reviewed formulations and processes used for converting drug nanosuspensions to final drug products in publications up to 2012, as well as providing a review of patents. Brough and Williams [93] provided a comparative analysis of amorphous solid dispersions and nanocrystal technologies for poorly water-soluble drugs for oral delivery. Junghanns and Muller [29] summarized the approaches used to formulate the currently marketed products containing poorly water-soluble drugs. Kesisoglou et al. [28] described the principles of nanosizing, production, and characterization of nanoformulations, and in vivo impact of these formulations. Chogale et al. [94] mainly described characterization techniques for various characteristics of nanocrystals (particle size, saturation solubility, dissolution velocity), which have an impact on the improved performance of nanocrystals. Peltonen and Hirvonen [95] presented the most important properties of nanocrystalline drug compounds, with multiple examples of the development and characterization of nanocrystalline drug formulations and a focus on the role of higher saturation solubility. They explained the impact of polymers and surfactants on the stabilization of nanocrystals, with a few examples from the literature and marketed products, but did not do an in-depth analysis on the roles of various dispersants. Malamatari et al. [31] outlines the advantages, stabilization, and production of drug nanocrystals, with an emphasis on wet milling, while highlighting their pharmaceutical applications. Although there is some overlap among the aforementioned review papers, each one has a unique focus and different duration of literature covered. In general, most reviews neither highlighted the critical role of redispersion on dissolution rate improvement of poorly water-soluble drugs, nor did they discuss recently developed redispersion methods in detail. Unlike some of the previous reviews, we provide here a systematic critical analysis of different classes of dispersants in terms of their functionalities and impact on the aggregation–redispersion. As the most comprehensive review paper on nanosuspension drying, this paper provides a statistical analysis of the usage of various preparation methods for drug nanosuspensions and nanocomposites from 92 studies from 2012–2017. This review also covers novel drying methods such as nanoextrusion and wet film-casting–drying, as well as novel dispersants, which are largely missing from the previous reviews. Finally, it provides significant guidance and insight into the rational selection of a drying process and dispersant, in order to develop robust, redispersible, fast-dissolving nanocomposite formulations.
The organization of this review paper closely follows the sequence of preparation steps in Figure 1. Section 2 presents a short review of various approaches used for the preparation of drug nanosuspensions; the characterization of nanosuspensions is not discussed at length, as several review papers cover this topic in detail; readers are referred to these review papers and the references cited therein [23,55,94]. Section 3 presents a comprehensive review of various drying methods, methods for characterizing the drug-laden nanocomposites including the newly developed redispersion test methods, and formulation aspects such as functionalities of dispersants and their impact on redispersion/drug dissolution. It will also present a statistical analysis of the usage frequency of various preparation methods for drug nanoparticles–nanocomposites and various dispersant classes. Section 4 will present important engineering considerations that must be taken into account for the selection of a drying process, and explain some of the diverging trends between academic studies and industrial practice. Section 5 will provide various insights gained from the analysis of the data presented in Section 3, as well as practical guidance for the rational selection of dispersants for robust, streamlined formulation development of redispersible, fast-dissolving drug nanocomposites. Finally, specific applications of drug nanosuspensions and drug-laden nanocomposites in drug delivery, as well as patent landscape, have been extensively covered in previous reviews (e.g., [31,55]); hence, they are outside the scope of this review paper.

2. Preparation of Drug Nanosuspensions and Their Stabilization

2.1. Preparation Methods

Various methods used for the preparation of drug nanosuspensions in 2012–2017 studies are summarized in Table 2. Top-down approaches aim to break micron-sized drug crystals down to smaller micro or nanoparticles via shear–impact. Wet media milling (WMM) and high-pressure homogenization (HPH) are the most commonly used top-down approaches for particle size reduction (Figure 4). WMM is an organic solvent-free process that has several distinct advantages, such as production of suspensions with high drug loading, ability to run continuously, and good scalability. Moreover, it can be universally applied to most drug candidates with poor water-solubility [21,23,96]. In WMM, drug suspensions are prepared by dispersing drug particles in a stabilizer solution followed by particle size reduction in a media mill, where coarse drug particles are broken down into smaller particles by bead–bead collisions. Particle size during milling generally depends on process–equipment parameters, mechanical and physicochemical properties of drug particles, and the physical stability of the milled suspension, i.e., extent of aggregation and/or Ostwald ripening in the presence of various stabilizers [23,97,98]. Li et al. [23] provided a holistic view of various formulation–processing aspects of WMM, and concluded that preparation of a drug nanosuspension with desired particle size and adequate storage stability entails selecting a proper stabilizer formulation and effective process–equipment parameters for the WMM process. While either a polymer or a surfactant alone can be used for stabilization, a combination of a cellulosic polymer and an anionic surfactant has been shown to be effective in stabilizing multiple drug nanosuspensions [99,100,101]. The impact of bead size–loading, rotor speed, and drug loading on breakage kinetics, drug particle size, and operational efficiency was studied extensively via experimentation and microhydrodynamic models [102,103,104].
High-pressure homogenization (HPH) is another popular top-down method to produce drug nanoparticles [105]. It uses jet-stream homogenization by pumping drug, dispersion medium, and stabilizers through a micro fluidizing nozzle. The particle size reduction is caused by cavitation forces, shear forces, and collisions through multiple homogenization cycles. The process parameters that control the particle size are homogenization pressure, number of passes, drug loading, and stabilizer type–loading. Shen et al. [106] and Sun et al. [107] used HPC to produce stable drug nanosuspensions with a combination of steric and electrostatic stabilizers. Compared to WMM, HPH has the advantage of reduced product contamination due to wear because it does not use milling media.
Liquid antisolvent precipitation (LASP) [108,109], supercritical fluid precipitation [37,38], acid–base precipitation [110,111], and melt emulsification [39,40] are some of the common bottom-up methods. Here, we briefly describe LASP, as it is the most widely used bottom-up method (Figure 4), and melt emulsification, as it is a facile, cheap, and solvent-free bottom-up method. Particle formation by LASP involves the mixing of solution–antisolvent streams to generate supersaturation and fast precipitation of particles [36,69]. Uniform mixing conditions ensure rapid and uniform supersaturation, making it a precipitation-controlled process which results in the precipitation of ultra-fine particles with narrow particle-size distribution. However, the residual solvent should be quickly removed from the resulting suspension [112]; otherwise, it could lead to severe aggregation and particle growth [62,109]. Stabilizer screening involves selecting a favorable solvent and antisolvent system and stabilizers that can adsorb on the crystal surface as they form, thereby inhibiting crystal growth. Despite the simplicity of its design, low energy consumption, and absence of product contamination without any moving parts like in wet media milling, the LASP process has many challenges, such as residual solvents in suspensions, inadequate physical stability, and low drug loading in the suspensions [23,62,109].
Melt emulsification (ME) is a facile bottom-up method for preparing drug nanosuspensions [39,40]. In this process, an aqueous suspension of drug particles is heated to temperatures above the melting point of the drug to form an oil-in-water emulsion owing to the immiscibility of the molten drug–water. The hot emulsion is broken into smaller droplets by applying mechanical agitation via ultrasonication, homogenization, magnetic stirring, etc. Subsequent cooling of the emulsion leads to solidification–recrystallization of the drug droplets into nanoparticles. Obviously, ME is only applicable to drugs with melting points below the boiling point of water. The process parameters that affect particle size include sonication energy, cooling rate, drug loading, and suitable stabilizer selection. Knieke et al. [39] screened stabilizers using the hydrophilic–lipophilic balance concept to produce 30 wt.% fenofibrate nanosuspensions. They were able to produce nanosuspensions with smallest size of 150 nm using poloxamer 188 as the stabilizer by optimizing sonication energy, speed, drug loading, and stabilizer loading; however, the suspensions were physically stable at that size only for few min, thus requiring immediate solidification of the nanosuspension via drying.
Combinative methods have been claimed to increase efficiency of particle size reduction [41]. In general, they can be described as a combination of a bottom-up method followed by a top-down method such as LASP–HPH and ME–HPH. For example, Fu et al. [113] combined LASP and HPH to prepare nanosuspensions of Nimodipine. Nimodipine was dissolved in dimethyl sulfoxide (DMSO) and instantaneously precipitated in aqueous phase containing poloxamer 127, HPMC E5, and sodium deoxycholate. DMSO was removed by lyophilization to improve the stability of the nanosuspensions. If DMSO was not removed quickly enough, the drug particles would have grown. Apparently, combinative methods have not yet attracted as much attention as either WMM or LASP alone (Figure 4).
The suspensions prepared by bottom-up approaches typically have low drug loading and poor physical stability. The solvent used in LASP must be removed quickly, either by filtration–drying [109] or continuous drying [62,112], which entails more processing steps or elaborate process design–integration, respectively. In contrast, top-down methods, especially wet stirred media milling, offer significant advantages: they are considered more universal, i.e., applicability to a large class of BCS Class II drugs because of their capability of achieving high drug loading, organic solvent-free processing, continuous operation capability, and ease of scale up [22,23,31,98]. While wet stirred media milling is energy intensive and more costly and may cause unacceptable media wear/product contamination, recent investigations [102,103,104] and wide industrial practice exemplified in the manufacture of a multitude of marketed drug products (refer to Table 1) suggest that these issues can be easily mitigated by the judicious choice of process–equipment parameters. Hence, it is not surprising to see from Figure 4 that WMM was the most popular method of drug nanosuspension production in 2012–2017 studies. Similarly, it is the most widely used method for drug nanoparticle production in the marketed products (refer to Table 1). This convergence of published academic/industrial research and actual industry practice indeed suggests that WMM is the preferred method for pharmaceutical drug nanoparticle production.

2.2. More on the Stabilization of Drug Nanosuspensions

Since there are several review papers on the stabilization of drug nanosuspensions (e.g., [23,28,114]), this section will provide a brief review of the topic only. Electrostatic forces, steric forces, entropic forces, and van der Waals forces among nanoparticles determine the overall physical stability of a drug nanosuspension [57]. Stabilization of nanoparticles can be achieved by soluble polymers and/or surfactants as a class of dispersants known as stabilizers [56,115,116]. In fact, most of the soluble polymers and surfactants presented in Table 2 serve as stabilizers in the preparation of drug nanosuspensions. As mentioned in the Introduction, zeta potential of drug nanosuspensions is important to their stability. According to Müller [49], a zeta potential value of at least ±30 mV is required for an electrostatically stabilized suspension. About ±20 mV provides only a short-term stability, and values in the range −5 mV to +5 mV indicate fast aggregation [50]. In the case of a combined electrostatic and steric stabilization, a.k.a. electrosteric stabilization, a minimum zeta potential of ±20 mV is desirable [51]. However, drug nanosuspensions with zeta potentials below 20 mV (absolute) were physically stable in some earlier work [101,117,118], which could be explained by the adsorption of nonionic polymer or nonionic surfactant and ensuing steric effect alone. Hence, the use of zeta potential alone, especially for predicting the stability of drug nanosuspensions stabilized with combinations of polymers–surfactants, should be considered with caution [119,120,121,122].
The selection of an optimal stabilizer formulation is a laborious experimental task, yet an important one to produce a stable drug nanosuspension. A poorly formulated drug nanosuspension may undergo aggregation, Ostwald ripening, fast sedimentation of particles, and cake formation during milling/storage, which will lead to various issues in downstream processing of the respective suspensions, and poor product performance from the final oral solid dosages such as slow drug release [23,33,65,123]. As a general principle, if used at insufficiently low concentrations, stabilizers such as polymers and surfactants in drug nanosuspensions may not prevent aggregation, while their excessive use, especially for surfactants, can promote Ostwald ripening [53,54,100] or raise the viscosity so much that downstream processing may be negatively affected, e.g., inability to spray a drug nanosuspension in spray-drying and fluidized bed coating.
The first systematic investigations of the stabilizing capability of adsorbed polymers were carried out by Lee et al. [124,125]. A connection between the hydrophobicity of the polymer and the ability to stabilize drug nanocrystals was indicated [124]. In addition, differences in the surface energy between the particle and the polymer were found to play a role in the stabilization process [125]. Choi et al. [126] concluded that not only the surface energy, but also the specific interaction between the stabilizer and the drug appears to play important role. George and Ghosh [127] investigated the correlation between drug–stabilizer properties and critical quality attributes (CQAs) of drug nanosuspension formulations. Their study suggested that logP and fusion enthalpy of the drugs had a direct impact on the feasibility of a stable nanosuspension, and that the most likely candidate for WMM was a drug with high enthalpy and hydrophobicity. In contrast, in a more comprehensive study, Eerdenbrugh et al. [119] used 13 stabilizers at three different concentrations in wet-milled suspensions of nine drug compounds, and concluded that no correlation between physicochemical drug properties (molecular weight, melting point, logP, solubility, and density) and stable nanosuspension formation exists.
Recent modeling and experimental investigations [63,66,85,99,100,128,129] have suggested that the combined use of non-ionic cellulosic polymers such as hydroxypropyl cellulose (HPC), hydroxypropyl methyl cellulose (HPMC), etc. and surfactants, especially anionic surfactants such as sodium dodecyl sulfate (SDS), dioctyl sodium sulfosuccinate (DOSS), etc., can have synergistic stabilization effects on drug nanosuspensions. Bilgili et al. [101] demonstrated HPC–SDS combinations for adequate stabilization of five BCS Class II drugs, attributing the synergistic stabilization to an electrosteric mechanism, similar to that described in previous studies [65,99]. When SDS was used below the critical micelle concentration (CMC) to stabilize a griseofulvin nanosuspension along with HPC, the significant synergistic stabilizing action of HPC–SDS was attributed to enhanced drug wettability (lower surface tension and higher wetting effectiveness factor) and ensuing higher deaggregation effectiveness afforded by the presence of SDS, in addition to the steric stabilization afforded by HPC [122].

3. Preparation, Characterization, and Formulation of Drug-Laden Nanocomposites

3.1. Drying Processes for the Production of Nanocomposites

Drug nanosuspensions are dried to prepare drug-laden nanocomposites. Based on 92 studies published from 2012–2017, Table 2 lists the method of nanoparticle formation, drying method, and dispersant formulations used. It also presents the results from redispersion tests, including the characteristic particle sizes of the drug particles in the drug nanosuspension (before redispersion) and those sizes after redispersion of the nanocomposites in the redispersion medium. The data in Table 2 was analyzed and the usage frequency of various drying methods is presented in Figure 5.
Freeze drying is the most popular method for drying drug nanosuspensions into nanocomposites (Figure 5). During freeze drying, a.k.a. lyophilization, water is removed from a frozen sample by sublimation and desorption under vacuum. The freeze drying cycle can be divided into three steps: freezing (solidification), primary drying (ice sublimation) and secondary drying (desorption of unfrozen water). The general purpose of pharmaceutical freeze-drying is to achieve long-term stability of heat-labile active compounds in a formulation [195,196]. This process generates various stresses during the freezing and drying steps. So, protectants are usually added to the nanosuspension formulation to protect the nanoparticles from freezing and desiccation stresses, as well as to prevent aggregation during drying. Most commonly used dispersants include cryoprotectants such as trehalose, glucose, sucrose, and mannitol [197]. The immobilization of nanoparticles within a glassy matrix of cryoprotectant can prevent their aggregation and protect them against the mechanical stress of ice crystals [90,197]. Type–concentration of stabilizers (polymer–surfactant) and cryoprotectants, drug loading, and the drug:cryoprotectant ratio affect drug nanoparticle aggregation and redispersibility [74,197]. Usually, a freeze-thawing study should be realized before freeze drying to select the cryoprotectant which is best able to conserve the properties of the nanoparticles. The critical process conditions during freeze drying are the velocity of freezing with or without annealing, the pressure and shelf temperature, and the duration of each stage of the process i.e., freezing time, primary drying time, and secondary drying time [196,197].
Spray drying is the second most popular drying technique for converting nanosuspensions into dry powders (Figure 5). In spray drying, solutions or suspensions are atomized through a bi-fluid or pressure nozzle and sprayed into a drying chamber, and hot air passing co-currently dries the atomized droplets [198]. The particles formed upon evaporation of the liquid are further dried in the cyclone and separated into a collecting chamber. The critical process parameters during spray drying are droplet size, process air temperature, atomization air flow rate, suspension/solution flow rate, viscosity, and solids loading in the feed [89,199]. Vehring [89] provides an extensive review of pharmaceutical spray drying processes in general. Several investigations have demonstrated the formation of high drug-loaded, fast-dissolving drug-laden nanocomposite microparticles upon drying of drug nanosuspensions [27,85,121,185]. Spray-dried powders usually have a low bulk density and particle size, i.e. in the range of ~5­–50 µm, depending on the formulation [89,200].
In fluid bed coating/drying, drug nanosuspension is atomized via a bi-fluid nozzle and sprayed onto fluidized carrier/substrate particles, also known as beads, and coating–drying of the nanoparticles on the beads occurs (bead layering) [64]. Usually, lactose or microcrystalline cellulose particles with sizes from 50–1000 µm are used as carriers. The suspension flow rate, atomization air pressure, fluidization air flow rate, and air temperature are controlled to prevent agglomeration of the coated particles. In an alternative processing route, i.e., fluid bed granulation/drying, drug nanosuspensions containing polymers, which serve as granulation binder fluid, can be sprayed onto a fluidized bed of excipients to form granules or agglomerates [130,192]. There are 7 papers on fluidized bed coating vs. two on fluid bed granulation (see Table 2). Hence, we focus on fluid bed coating here. Bhakay et al. [66,84] and Azad et al. [91,201] fluid bed coated nanosuspensions of griseofulvin and fenofibrate onto various grades of lactose carrier particles and investigated the impact of formulation and carrier size on the formation of redispersible nanocomposite particles. Knieke et al. [152] successfully coated sub-50 µm lactose carrier particles with fenofibrate nanosuspension without appreciable agglomeration, and prepared free-flowing powders for proper downstream processing. In spite of having a median particle size less than 100 µm, the final composite powders were free flowing, had a high bulk density, and exhibited fast fenofibrate release during the dissolution. Azad et al. [91] examined the impact of carrier particle size on redispersion and dissolution of itraconazole (ITZ) and fenofibrate (FNB) nanocomposites. The drug nanosuspensions were prepared by WMM with HPMC–SDS as stabilizers, followed by their coating onto the GranuLac 200 (d50: 27.7 µm) and PrismaLac 40 (d50: 321.1 µm) particles. As seen in Figure 6, the nanocomposite particles with GranuLac 200 exhibited significantly faster drug dissolution than those with PrismaLac 40. Finer carrier particles, owing to their higher surface area per unit mass and corresponding thinner coating layer (drug–polymer shell), provide faster drug release than coarser carrier particles.
Recently, a nanoextrusion process has been developed to disperse drug nanoparticles in a polymeric matrix (nanocomposites) in the form of extrudates using a variant of the traditional hot melt extrusion (HME) process [75,76,77,78]. Unlike a traditional HME process, the nanoextrusion process uses a wet-milled suspension of the drug as feed, along with an extrusion polymer, and disperses drug nanoparticles in the polymeric matrix while evaporating the water, thus yielding dry extrudates in the form of nanocomposites. Simply put, nanoextrusion acts as a continuous drier unlike the traditional HME process. This technique was first presented by Khinast et al. [75] as a one-step process for converting a stabilized nanosuspension into nanocomposite, where crystalline nano-titanium oxide was used as a model substance. In a follow-up study by Baumgartner et al. [76], crystalline phenytoin nanoparticles were used to demonstrate applicability to pharmaceutical products. Ye et al. [77] combined the use of high-pressure homogenization and extrusion for the production of nanocomposite with low drug concentration, i.e., 1–2%. Li et al. [78] prepared extrudates with nanocrystalline griseofulvin (GF) particles dispersed in the HPC matrix as a secondary phase (nanocomposites) and extrudates with amorphous GF molecularly dispersed within the Soluplus® matrix (amorphous solid dispersion), demonstrating the versatility of the nanoextrusion process. Li et al. [87] milled the GF–HPC extrudates into various sieve cuts and examined the impact of matrix (particle) size on drug release; they observed that finer extrudate particles led to faster drug release, but such an impact quickly attains a plateau below ~200 μm.
Besides the interest in the aforementioned drying methods that yielded powders/cakes or extrudates, there has also been an interest in incorporating drug nanoparticles into polymeric strip films (Figure 5). Thin strip films offer significant advantages for oral delivery of drugs, such as rapid disintegration and dissolution in the oral cavity, especially for geriatric–pediatric populations [202]. The wet film casting–drying commonly used with soluble drugs has been applied to dry drug nanosuspensions in polymeric strip films, which encapsulate drug nanoparticles in a polymeric matrix uniformly (nanocomposite) [79,203]. In this method, drug nanosuspensions are mixed with aqueous solutions of HPMC E15LV, HPMC K4M, or pullulan, along with a plasticizer for desired viscosity; the resulting slurry is cast into film using a casting knife to control the film thickness, followed by drying. The resulting film is cut into different sizes according to the required dose. Drug nanoparticles prepared by both top-down and bottom-up approaches have been incorporated into thin polymeric strip-films that exhibit significantly enhanced drug release and acceptable content uniformity, even for low drug doses [40,79,109]. Sievens et al. [79] wet-cast and dried a slurry formed by mixing griseofulvin, fenofibrate, and naproxen nanosuspensions prepared by WMM with aqueous HPMC–glycerin solutions. The films exhibited good redispersibility for all drugs. The dissolution testing was only performed for fenofibrate; films exhibited faster drug release compared with the as-received drug, physical mixture, and the compact of the drug. Susarla et al. [131] investigated the impact of convection drying parameters such as air velocity and temperature, as well as film precursor viscosity on film properties and redispersion of griseofulvin nanoparticles. The griseofulvin nanoparticles were fully recovered upon redispersion in water from all convection-dried films, suggesting that the film formation process, including faster drying, did not lead to irreversible drug nanoparticle aggregation. Krull et al. extensively studied the impact of different plasticizers [81], polymer molecular weight [82], and the impact of drug loading [83] in a series of papers on redispersion of BCS class II drugs from polymeric strip films and drug release during in vitro dissolution. In a separate study, Krull et al. [162] prepared griseofulvin nanoparticle laden pullulan–xanthan gum films. Thinner films, films with lower xanthan gum loading, and smaller drug nanoparticles led to faster drug release from the films, while drug loading had no discernible effect within the range studied.

3.2. Characterization of the Nanocomposites

Nanosuspension characterization methods have been discussed in several review papers [23,28,94]. In this section, we briefly review various characterization methods for the characterization of the nanocomposites. Several methods are routinely used, except redispersion methods and drug wettability testing, which will be discussed in more detail.

3.2.1. Particle Sizing

Particle size of nanocomposite particles is usually measured by laser diffraction using e.g., Sympatec Rodos/Helos particle sizer, Beckman Coulter, or Malvern Mastersizer, in the dry particle sizing mode. Approximately 1 g of nanocomposite powder was placed on the chute of a Rodos dispersing system and particle size was measured at an optimum dispersing pressure to avoid attrition of nanocomposite particles [27,86]. The particle size of nanocomposite particles after redispersion in a liquid medium and drug particle size in nanosuspensions can be measured using the liquid module of Beckman Coulter or Malvern Mastersizer by laser diffraction (e.g., [8,66,85,152]). Malvern zetasizer and Delsa nano are also used to measure size of nanosuspensions and redispersed nanocomposite particles via dynamic light scattering (DLS) (see e.g., [103,150,151,182]). Since DLS can measure particle sizes accurately up to few microns, it cannot measure coarse aggregates or clusters present in the nanosuspensions or redispersed nanocomposite particles. On the other hand, laser diffraction instruments cannot accurately determine particle sizes below 100 nm. Unless median drug particle size is smaller than 100 nm, laser diffraction is the first method of choice, as it can detect the primary particles as well as large aggregates/clusters in a given sample. Otherwise, DLS alone or in combination with an orthogonal method like scanning/transmission electron microscopy coupled with image analysis could give valuable information (e.g., [103]).
Atomization of nanosuspensions via a bi-fluid or pressure nozzle is an integral part of spray drying and fluid bed coating/drying processes. Malvern Spraytec/Insitec RTSizer or similar laser-diffraction based system can be used to measure size of droplets emanating from a nozzle (e.g., [204,205]), which allows formulators to examine the impact of atomization air pressure, nanosuspension flow rate, and viscosity on droplet size. Once the droplet size is known at a small scale, it should be maintained across different scales to ensure proper scale-up of the drying process.

3.2.2. Scanning Electron Microscope (SEM) Imaging

The morphology and structure of nanocomposites can be examined under a scanning electron microscope (SEM). A few nanocomposite particles can be placed on a carbon tape which is stuck onto an SEM stub, and images are taken at different magnifications under SEM (e.g., [27,206]). The cross-section of nanocomposite particles can be seen by embedding the nanocomposite particles in epoxy, slicing the epoxy using a microtome, and placing a thin cut of epoxy under SEM (e.g., [207]). Another method for the visualization of cross-section of nanocomposite particles is to spread the particles onto a glass slide, cut them with a sharp knife, and focus on the broken particles under SEM (e.g., [66,207]).
Elemental analysis of nanocomposite powders can be done by energy dispersive X-ray analyzer to confirm uniform dispersion of API in the nanocomposite particles. An electron microscope with a detector for characteristic X-rays, i.e., energy-dispersive X-ray spectroscopy (EDS), is a powerful analytical method for quantitative and qualitative surface analysis. Generated X-rays inside the SEM are characteristic for each atom in the periodic table (excluding He and H), and their detection by EDS allows for measuring the elemental composition of the sample. The amount of emitted X-rays from each element is directly proportional to its concentration (mass or atomic fraction) in the sample. EDS analysis will only work if an element present in the drug molecule gives a distinct X-ray signal [207]. For example, Azad et al. [27] employed EDS to show the presence of milled superdisintegrant particles along with fenofibrate nanoparticles in nanocomposite particles.
The suspension obtained after redispersion of nanocomposites particles can be observed under SEM by placing a drop of the suspension on a silicon chip [8,86]. The silicon chip is placed onto a carbon tape stuck onto a SEM stub, and dried in a vacuum oven before imaging using SEM. The objective of SEM imaging after redispersion of nanocomposite particles in water is to look for nanoparticles and nanoparticle aggregates [66,85].

3.2.3. Drug Crystallinity

The crystallinity of the drug in nanocomposite particles is usually examined via X-Ray diffraction (XRD) (e.g., [183,208]) and differential scanning calorimetry (DSC) (e.g., [133,209]). The crystallinity of the drug in nanocomposites is affected by the nanosuspension preparation and drying method, as well as the formulation, especially drug–polymeric dispersant interactions. For most wet media milled drugs dried into nanocomposite particles with polymeric dispersants, the diffractograms show that the peak positions are preserved with some broadening and intensity reduction due to the formation of fine nanocrystals, defect formation, and even some partial amorphization [210,211,212]. For similar reasons, DSC thermograms clearly show a distinct melting endotherm, albeit with a melting point depression and reduced fusion enthalpy, as compared with as-received drug and physical mixtures [22,27,213]. Freeze-drying of drug nanosuspensions prepared by top-down approaches could either preserve drug crystallinity or result in some reduction in crystallinity, depending on the concentration/type of dispersant used [179,182,190]. Freeze drying of drug nanosuspensions prepared via LASP tends to result in a pronounced loss of crystallinity, and even complete amorphization [194,214,215].

3.2.4. Redispersion Methods

Drug nanoparticles in the dispersant matrix of nanocomposites must be released into the dissolution medium in vitro or in vivo, preferably in the form of discrete primary nanoparticles, so that large surface area of the nanoparticles could allow for fast, immediate drug release. Hence, developing an understanding of the redispersion phenomenon and nanoparticle release mechanisms could significantly help robust nanocomposite formulation development [66,84,92]. Redispersion testing entails dispersing nanocomposite particles in water or a physiologically relevant fluid upon application of agitation/shear, so that drug nanoparticles are recovered in the fluid [66,216]. Table 2 shows that common agitation methods include sonication, magnetic stirring, pipette stirring, paddle stirring, manual shaking, as well as in-situ shearing/agitation in the sample cell of laser diffraction equipment. Unlike dissolution testing, redispersion testing judiciously uses a dispersion fluid volume–nanocomposite mass so that all dispersants present in the nanocomposite dissolves, while drug nanoparticles simply disperse in the fluid without significant extent of dissolution; hence, their particle sizes can be measured.
In redispersion studies, it is customary to compare the particle size of the milled suspension with that of the suspension obtained from the redispersion of the nanocomposites in water. This comparison has been used to assess the recovery and redispersion behavior of drug nanoparticles [66,67,92,147,217]. If the redispersed suspension contains water-insoluble dispersant particles, these particles can be removed by centrifugation or filtration to avoid interference during particle size measurements. If possible, either filtration should be avoided, because drug particle aggregates and agglomerates may also be removed in this process, or filter opening size should be carefully chosen to separate only water-insoluble dispersants. The supernatant obtained after centrifugation can be used to measure the particle size of the redispersed suspension. The centrifugation time and speed should be optimized such that drug aggregates are present in the supernatant. Another approach is to analyze the supernatant by dissolving the drug in a solvent, measuring the concentration via UV or HPLC, and calculating the fractional recovery by comparing it with the original amount of drug concentration present in the nanocomposite particles [92].
A cursory look at the literature prior to 2012 reveals that many studies on the drying of pharmaceutical nanosuspensions did not consider the redispersion and recovery of nanoparticles at all [34,59,60,64,69,70,218,219,220]. Redispersion testing was performed, but little to no detail was provided in other studies [33,65,197,221,222]. Yet another group of studies investigated the recovery of polymeric nanoparticles, silica, and nanocapsules without drugs [58,61,217,223]. In a smaller group of studies, the recovery of drug nanoparticles during redispersion was investigated for spray-dried [62,63], spray-freeze dried [67], freeze-dried [74], and fluid-bed dried [63] nanocomposites. These studies described a single redispersion method without investigating the potential impact of redispersion method itself, due to differences in agitation/shearing conditions. Table 2 shows that 73% of the nanocomposite formulations in 2012–2017 studies were subjected to redispersion testing. The data in a 2014 review paper by Chin et al. [55] suggest this ratio to be 40% (33% for studies before 2011). This comparison shows that there is significantly more interest in developing an understanding of the redispersion behavior of the nanocomposites in recent studies.
To the best of authors’ knowledge, there are only two studies [66,92] which solely investigated the redispersion of drug-laden nanocomposites using multiple redispersion test methods. Bhakay et al. [66] studied the impact of different redispersion methods on particle size of drug-laden, core (lactose)–shell nanocomposite microparticles. About a gram of the nanocomposite particles were weighed and dispersed in 30 mL water for 2 min. The maximum amount of the drug that can dissolve in water during the redispersion test in their method was very small (e.g., about 0.2% of drug particles). Dispersants such as HPC, SDS, lactose, and mannitol in nanocomposite particles dissolve in water and the particle sizes obtained from particle size instruments were mainly the sizes of undissolved drug particles and their clusters. Fast redispersion and recovery of nanoparticles (e.g., within 2 min) under low agitation conditions could be a desirable attribute of robust BCS Class II drug formulations in immediate release, solid dosage forms. Gentle pipette stirring (manual), magnetic stirring at 100 rpm, stirring with an overhead laboratory stirrer at 200 and 1200 rpm, and sonication in an ultrasonic bath were used. After dispersing the nanocomposite particles in water or buffer solutions, an aliquot of the sample was taken while the sample was being agitated, and the particle size was measured by laser diffraction. This study established that the nanosuspension formulation (dispersant type/concentration) has more impact on the redispersed particle size than the agitation/shearing method. In principle, any of the above redispersion methods may be used to rank different dispersant formulations according to their capabilities of releasing nanoparticles. Different redispersion methods had only a slight impact on the recovery of nanoparticles when SDS was present either in the formulation or in the redispersion medium (water). Sonication appears to be more effective than stirring-based methods in deaggregating the clusters and aggregates formed during the redispersion in water.
Bhakay et al. [92] have assessed the use of quiescent redispersion methods in comparison to agitated redispersion methods. Approximately 100 mg nanocomposite particles were dispersed in 4 mL water in a cuvette, as shown in Figure 7 (left panel), with the dimensions of 1.2 × 1.2 × 4.6 cm (L × W × H). As the nanocomposite particles descend and sediment, they release drug particles that can be observed visually and quantified by determination of turbidity, particle size, and drug assay. Several pictures were captured during the descent of nanocomposite particles every few seconds. During the settling, nanocomposite particles with no stabilizers, or with HPC alone, did not release significant amount of nanoparticles, as signified by low turbidity and low drug nanoparticle concentration in the supernatant. In a separate experiment, nanocomposite particles were dispersed in water kept in a cuvette without external agitation, and particle size was measured via dynamic light scattering (DLS) with Delsa nano. In a separate method, nanocomposite particles were redispersed in 4 mL water for 5 min and the resulting suspensions were centrifuged for 5 min to remove particles that did not redisperse. Centrifugation removed the large particles and clusters and mitigated high scattering during dynamic light scattering size measurements which would otherwise cause imprecise results. Another method is to study redispersion dynamics via DLS by dispersing the nanocomposite particles in water and measuring the size to develop an understanding of how larger particles redisperse over a period of time. Redispersion phenomena can also be visualized under optical microscope by adding a drop of water or buffer solution to a nanocomposite particle and capturing images at different time intervals, as shown in Figure 7 (right panel). The appearance of a “cloud” around the nanocomposite particle was a signature of drug nanoparticle recovery, and those formulations that exhibit such clouds also exhibited high turbidity generation during the sedimentation (see Figure 7, left panel). Overall, Bhakay et al. [92] concluded that both quiescent redispersion and agitated redispersion methods yielded similar rank-ordering of the dispersants used in the nanocomposites; hence, both types of tests could be used for screening/optimizing dispersants for fast-dissolving drug nanocomposite formulations.

3.2.5. Dissolution Testing

Nanocomposite samples having drug amounts equivalent to saturation solubility, or lower than saturation solubility of BCS Class II drugs in the dissolution medium have been used preferentially because such non-sink conditions have been shown to better discriminate between various nanocomposite formulations [59,169,224,225]. Water, aqueous surfactant solutions, buffer solutions, or other simulated biological fluids can be chosen as dissolution media. USP Type I apparatus (basket method) [150,159], USP Type II apparatus with a rotating paddle [66,135,146], and USP IV flow-through cell apparatus [79,131,153] were used to measure drug release from nanocomposites. Aliquots can be taken manually or in an automated fashion, at different time intervals to analyze percent drug dissolved. The amount of drug dissolved can be determined by measuring absorbance using UV spectroscopy or HPLC.

3.2.6. Drug Wettability

Wettability of drug particles and wettability enhancement upon use of various dispersants [78,122,226] can be examined via liquid penetration into a drug powder bed, also known as the modified Washburn method [227,228]. A typical powder tensiometer, e.g., Attension Sigma 700 set-up (Biolin Scientific, Linthicum, MD, USA), consists of a sample holder in the form of a cylindrical metallic tube with small holes at the bottom, as well as a hook at the top of the cover equipped with screw threads. About ~1 g of drug powder was packed uniformly into the tube before each measurement (e.g., [78]). A filter paper was placed at the perforated end of the sample holder to support the drug powder sample. A petri dish containing liquid, typically water or a dispersant solution, was placed below the perforated end of the holder on the mechanical platform. The powder tensiometer measures the mass of liquid penetrated into the drug powder bed as a function of time. The apparent shear viscosity and surface tension of the liquids must be known or measured. The contact angle for the deionized water/dispersant solution and drug can be determined using the modified Washburn equation, which provides a relationship between liquid penetration rate and contact angle θas follows:
T = ( η C ρ 2 γ cos θ ) M 2
where T, M, η, ρ, and γ are time after contact, mass of the liquid penetrated into the drug powder bed, viscosity of the liquid, density of the liquid, and surface tension of the liquid, respectively. C is a characteristic parameter of the drug powder sample. It could be determined independently using a completely wetting liquid such as hexane, heptane, etc. In examining the impact of various dispersants on wettability, the same drug powder is used as the powder sample; hence, C remains relatively invariant for different dispersant solutions and deionized water under similar packing conditions. One can therefore calculate the ratio of cosθd/cosθw, i.e., wetting effectiveness factor, as a crude measure of wettability enhancement by the dispersants. Here, θd is the contact angle between drug and the dispersant solution, and θw is the contact angle between drug and deionized water. The wettability enhancement with different dispersants on the wetting of drug particles may be assessed by comparing this ratio, taking the wettability by water as a basis of comparison [78]. The slope of the modified Washburn equation is obtained by fitting the linear region of a liquid penetration curve. Using the slope for different stabilizer solutions and water, cosθd/cosθw is calculated. Dispersants, especially surface-active adsorbing polymers and surfactants, improve drug wettability, and thus, help deaggregation during WMM [122] and enable faster redispersion and drug release from drug nanocomposites [78,87].

3.3. Nanocomposite Formulations and Functionalities of Dispersants

Table 3 classifies various dispersants and their functionalities–action mechanisms in drug nanosuspensions–nanocomposites, while Figure 8 presents their usage frequency in nanocomposite formulations. In the statistical analysis of Table 2 data with various dispersants used in a given study/formulation, a redispersible formulation was regarded as either the one that led to the finest drug particle size upon redispersion of the nanocomposites (see Table 2), or that led to the fastest dissolution if redispersion testing was not performed in that particular study. In the latter case, similar to Eerdenbrugh et al. [34,59], redispersibility was not directly studied, but inferred from the dissolution testing. Different classes of dispersants, dispersant loading, drug type, drug loading, ratio of drug:dispersant concentration, drying method, and particle formation methods have an impact on the redispersion of nanoparticles from nanocomposites and drug release. Some of the salient features and impacts of various classes of dispersants are presented below.

3.3.1. Impact of Soluble Polymers on Redispersion–Drug Release

Figure 8a shows that about ~83% of the nanocomposite formulations used soluble polymers either alone or in combination with other dispersants. Soluble polymers include cellulosic polymers, polyvinyl pyrrolidone, polyethylene glycol, etc. [33,86]; different grades of HPC, HPMC and PVP are the most popular (refer to Table 2). The typical use levels of these polymers are 1–7.5% w/v in nanosuspension formulations. They usually play a dual rule: firstly, being soluble dispersants that adsorb on drug nanoparticles, they act as steric stabilizers during the preparation of drug nanosuspensions [23,31,100]. Secondly, they serve as primary matrix/film former in the nanocomposites encapsulating/covering drug nanoparticles upon drying of drug nanosuspensions and mitigate nanoparticle aggregation during the drying [28,55]. Moreover, higher viscosity of the drug nanosuspension upon use of soluble polymers could help to retard aggregation due to slower Brownian motion during storage, and even mitigate phase separation-induced aggregation during drying [100], depending on the specific drug–polymer and their mass ratio.
Lee [33] prepared 16 wt.% drug (unspecified) nanosuspensions by WMM with HPC as the stabilizer, and spray-dried nanosuspensions to obtain nanocomposite particles. Increasing HPC concentration from 0.38–3.08 wt.% in suspension resulted in greater HPC adsorption onto the drug surface, and produced nanosuspensions with smaller particle size during milling. At a high concentration of HPC (3.08%), polymer adsorption on drug surface seemed to be sufficient to cover the drug surface and prevent aggregation of nanoparticles by steric stabilization during milling. It took about 25 h to completely redisperse the nanocomposite particles back into drug nanoparticles in water. Lee [33] attributed this slow nanoparticle recovery solely to aggregation during drying, perhaps due to the low (~1:5) mass ratio of HPC:drug. Other possible explanations include the formation of drug nanoparticle aggregates during storage prior to spray-drying (no stability study provided in Lee [33]), aggregation due to increased capillary forces during drying [34,88], and slow wetting and penetration of water into the spray-dried particles, besides the relatively slow dissolution of HPC. While Lee [33] did not study drug release, Li et al. [86] coated wet media milled griseofulvin (GF) suspensions stabilized with HPC onto Pharmatose carrier particles via fluidized bed coating/drying, and examined the impact of HPC concentration on GF release. The findings were somewhat similar to those of Lee [33], except that most of the drug aggregates already formed during the milling–storage, which led to poor redispersion and slow dissolution; immediate drug release was not achieved. Li et al. [86] demonstrated that higher HPC concentration led to smaller aggregates during milling and after redispersion, which in turn enabled faster drug release from the nanocomposites; however, since aggregation could not be suppressed completely, immediate drug release was not achieved.
In other studies such as He et al. [160], Ho and Lee [120], and Kumar et al. [139], drug nanoparticles were recovered from nanocomposites with soluble polymers that serve as stabilizers/dispersants. He et al. [160] formed nanoparticles of teniposide in the presence of PVP K30 by antisolvent precipitation followed by freeze drying; Ho and Lee [120] milled naproxen down to the nanosize domain with HPC and formed nanocomposite particles by electrospray drying; and Kumar et al. [139] wet-milled naproxen particles with HPMC E15 followed by spray drying to form nanocomposite particles. While the above-mentioned studies used different drugs and different particle formation–drying methods, the nanocomposites completely redispersed into nanoparticles during manual redispersion or sonication in less than 5 min. Ho and Lee [120] did not study dissolution of spray-dried nanocomposites. He et al. [160] and Kumar et al. [139] corroborated the redispersion results with dissolution; teniposide and naproxen nanocomposite particles formulated with PVP K30 and HPMC E15 showed significantly improved dissolution.
The aforementioned studies demonstrate that (i) preventing aggregation in drug nanosuspensions is a necessary condition for fast redispersion and drug release from the nanocomposites; (ii) soluble polymers may not be able to stabilize drug nanoparticles, regardless of the concentration used, leading to the formation of aggregates in the nanosuspensions; (iii) depending on the drug:polymer ratio, aggregates may also form during drying, which will cause inferior dissolution enhancement; and (iv) higher polymer concentration may alleviate the aforementioned issues at the expense of reduced drug loading in the nanocomposites. The upshot of these results is that other classes of dispersants, in addition to soluble polymers, may help to resolve the aforementioned issues; hence, it is no surprise to see only ~12% of the redispersible formulations used soluble polymers alone (Figure 8b).

3.3.2. Impact of Surfactants on Redispersion–Drug Release

~73% of the formulations in Table 2 used surfactants alone or in combination with other dispersants (Figure 8a). Being surface-active agents, surfactants reduce the surface tension of suspension liquids, help to disperse drug particles in suspension liquid by enhancing drug wettability, and help to stabilize drug nanosuspensions either alone or in combination with polymers [23,28,55,163]. They also impart wettability enhancement to the nanocomposites which help to speed up redispersion and drug nanoparticle recovery [78,86]. Charged surfactants such as SDS and DOSS stabilize drug nanosuspensions by electrostatic stabilization, while polymeric surfactants like poloxamers, TPGS, tween 80 etc. stabilize nanoparticles by steric repulsion. The typical use levels of surfactants are between 0.05–5% w/v in drug nanosuspensions. Surfactant usage below critical micellar concentration (CMC) is preferred, because at concentrations greater than CMC, drugs tend to dissolve in the micelles, which may lead to particle size growth via Ostwald ripening during storage of the nanosuspensions [53,54].
Eerdenbrugh et al. [34] wet media milled 9 model BCS Class II drugs with Vitamin E TPGS at 25 wt.% (with respect to drug) and dried the suspensions into nanocomposite particles by freeze drying and spray drying. Drugs with more hydrophobic surfaces formed aggregates during drying, and compromised the dissolution rate as compared to corresponding milled nanosuspensions. Other drugs formed redispersible nanocomposite particles with a polymeric surfactant alone. Bhakay et al. [66] reported poor redispersion from one-month aged nanocomposites prepared by fluid bed drying/coating of GF–SDS nanosuspension, while redispersibility was achieved by GF–HPC–SDS nanocomposites. Li et al. [86], for the same type of nanocomposites, found that when 0.5 wt.% SDS w.r.t. GF was used, significant aggregation occurred during the drying, which led to very slow GF release. The above findings, along with the observation that only ~8% of the redispersible formulations in Table 2 have surfactants alone (Figure 8b), suggest that other classes of dispersants, in addition to surfactants, must be considered to develop a more general approach for development of redispersible, fast-dissolving nanocomposite formulations.

3.3.3. Impact of Soluble Polymer–Surfactant on Redispersion–Drug Release

Figure 8b shows that ~29% of the redispersible formulations had soluble polymer–surfactant, which is the most widely used strategy for formulating drug-laden nanocomposites based on the data in Table 2. Combinations of HPC–SDS, HPMC–SDS and PVP–SDS are some of the most common choices for forming redispersible nanocomposite particles. Besides the beneficial effects and mechanisms of action for the soluble polymers and surfactants mentioned in Section 3.3.1 and Section 3.3.2, their combined use could lead to synergistic effects in drug nanosuspension stabilization, mitigation of nanoparticle aggregation, and fast redispersion–drug release from the nanocomposites. For example, when anionic surfactants along with soluble polymers were used, electrosteric stabilization of the drug nanosuspensions, as well as faster deaggregation during the milling due to enhanced wettability of the hydrophobic drug particles, occurred (refer to Section 2.2.). Bhakay et al. [66], Niwa et al. [7], and Basa et al. [65] reported the positive impact of polymer–surfactant combination in the formation of redispersible nanocomposite particles for different drugs and drying methods. In general, the stabilization of drug nanosuspensions and fast redispersion/drug release require much less soluble polymer–surfactant compared to the case when either soluble polymers or surfactants are used alone. The use of less dispersants in the nanocomposite has several benefits, such as reduced extent of Ostwald ripening (at reduced surfactant loading) and higher drug loading in the nanocomposite.
A systematic and comprehensive study was carried out by Li et al. [86] to examine the impact of HPC molecular weight and loading, SDS concentration, and combination of HPC–SDS on redispersion and dissolution performance of fluidized bed coated Pharmatose with wet-milled griseofulvin (GF) suspensions. In the absence of SDS, even at 7.5% HPC (any grade), GF nanoparticles severely aggregated in the wet-milled suspension, and dissolution from the nanocomposites did not achieve fast, immediate release, signifying the criticality of the surfactant. Figure 9 shows the impact of HPC molecular weight, in combination with 0.05 wt.% SDS, on dissolution of griseofulvin from nanocomposite particles. Immediate drug release was achieved when low molecular weight HPC SSL was used ≥1% w/w concentration (w.r.t. water) along with 0.05% w/w SDS. The highest molecular weight grade of HPC (HPC L) entailed the use of high concentrations of polymer (7.5% w/w) to immediately release drug nanoparticles despite of addition of 0.05% SDS. Besides showing the positive impact of polymer–surfactant combination, their study exemplifies some general trends: (i) higher dispersant concentration enabled faster drug release, albeit at a reduced drug loading; (ii) use of SDS at a small concentration (0.05%) was necessary to form almost aggregate-free nanosuspensions and mitigate aggregate formation during drying; (iii) proper stabilization of drug suspensions is a necessary condition for fast drug nanoparticle recovery and fast drug dissolution from the nanocomposites, but not sufficient because aggregation may also occur during drying, and nanoparticle recovery from the matrix of nanocomposites is controlled by wettability and dissolution of the dispersants; and (iv) immediate drug release was achieved when a lower molecular weight (SSL grade) HPC ≥1% concentration was used along with 0.05% SDS.
While polymer–surfactant combinations appear to be the most widely used formulation strategy (Figure 8b), additional dispersant classes may be needed for a variety of reasons. Surfactants may pose challenges such as particle growth via Ostwald ripening during milling and/or storage [53,54,100,229], and for anionic surfactants, incompatibilities with other ionic molecules, sensitivity to pH, salt or temperature changes, and GIT irritation [230,231]. Another potential issue with surfactants is that they can be toxic if used in excess [230], especially in inhalation and parenteral applications [232,233,234]. Surfactants can be chosen from the handbook of pharmaceutical excipients which provides the typical use levels for different administration routes [235]. Additionally, a list of inactive ingredients (excipients including the surfactants) used in approved drug products is available on the U.S. Food and Drug Administration (FDA) website [236]. This list provides information on dosage form, route of administration, and maximum potency of the excipient per unit dose in already approved products. For new surfactants or different usage (dosage form/route of administration) of existing surfactants, FDA and The International Council for Harmonisation (ICH, Geneva, Switzerland) guidance documents lay out toxicity studies to be performed in rodent species to define use levels (readers are referred to [237] and the ICH guidance documents cited therein).
Some of the aforementioned issues can be mitigated by reducing the surfactant loading using streamlined and material-sparing formulation approaches (e.g., [100,238,239]), and by the combined use of polymers–surfactants which allow for lower surfactant loading. In fact, Ostwald ripening issue can be practically resolved by simply adding/mixing most or all of the surfactant with the nanosuspension containing drug–polymer right before the drying step. Another formulation strategy is to minimize the use of surfactants or eliminate their use completely by making use of other classes of dispersants. In combination with soluble polymer–surfactants, the use of other dispersant classes can reduce the surfactant loading required otherwise, while still ensuring, or even improving the performance of the nanocomposites. Alternatively, surfactants can be completely replaced by additional dispersant classes. The use of water-soluble dispersants (WSD) [197], water-insoluble dispersants (WID) [59], and crosslinked polymers (swellable, water-insoluble dispersants) [8,27] will be covered below.

3.3.4. Impact of Other Classes of Dispersants

Water-soluble dispersants (WSD) refer to sugars, sugar alcohols, cyclodextrin, and some proteins present in the nanocomposites other than soluble polymers and surfactants. They dissolve on contact with water and aid in breakage/erosion of the nanocomposite matrix, as well as drug particle aggregates during redispersion/dissolution [28,55,138,152]. WSD are the most widely dispersants after polymers and surfactants (Figure 8a), and usage frequency of polymer–surfactant–WSD is second to that of polymer–surfactant in redispersible formulations (Figure 8b) because, having low molecular weight, WSD dissolve faster than soluble polymers, thus creating pores in nanocomposite matrix and facilitating polymer dissolution and drug nanoparticle release [152]. During freeze drying, sugars and sugar alcohols act as cryoprotectants by reducing ice formation during freezing, and thereby prevent formation of nanoparticle aggregates [197,240]. As can be seen from Figure 8b, they are rarely used alone because they cannot stabilize drug nanosuspensions—unlike polymers/surfactants—and they are relatively poor film formers. WSD:drug mass ratio must be set relatively high (for some drugs up to 10:1 in freeze drying [240]) so that they can physically separate the drug nanoparticles and prevent their aggregation. They are almost exclusively used along with polymer/surfactants and polymer–surfactants. For example, even with 1:1 mannitol:griseofulvin (GF) mass ratio, GF–HPC–mannitol nanocomposites prepared by either fluid bed coating [8] or spray drying [27] could not achieve fast, immediate release in the absence of a surfactant although addition of Mannitol significantly improved the drug release rate. Hence, WSD concentration must be optimized, or alternative classes of dispersants besides polymers/surfactants should be considered.
Water-insoluble dispersants (WID) such as microcrystalline cellulose, anhydrous dicalcium phosphate, colloidal fumed silica, and clay/montmorillonite particles prevent direct contact of drug nanoparticles with each other, thus preventing the formation of aggregates during drying and aid in redispersion [59,192]. Eerdenbrugh et al. [59] used WID such as Avicel PH101 (microcrystalline cellulose), Fujicalin (anhydrous dicalcium phosphate), and Aerosil 200 (colloidal fumed silica), as well as a polymeric surfactant (Inutec SP1). Nanosuspensions of phenylbutazone, itraconazole, and cinnarizine were prepared in an aqueous solution of 25 wt.% Vitamin E TPGS (surfactant) and spray-dried after the addition of the aforementioned dispersants. Among all three WID examined, only Aerosil 200 improved dissolution performance of all spray-dried powders, which was attributed to its large surface area and good matrix forming capability. However, the best dissolution performance resulted from the use of Inutec, i.e., the polymeric surfactant. Unfortunately, there is no comparative assessment of Aerosil 200 as a WID with commonly-used WSDs such as sugars and sugar alcohols. Hence, the use of such insoluble dispersants warrants further investigation.
Crosslinked polymers (CLP), especially wet-milled superdisintegrants [8,27,241], form a novel class of dispersants which erode/disintegrate the nanocomposite matrix by swelling on contact with water and release nanoparticles during redispersion–dissolution. Azad et al. [27] and Bhakay et al. [8] have used common superdisintegrants such as SSG, CCS, and CP as a novel class of dispersants with the goal of replacing surfactants in nanocomposite formulations. Superdisintegrants were wet-milled along with griseofulvin particles in the presence of HPC; the resultant suspension was either fluid bed coated onto Pharmatose carrier particles [8] or spray dried [27]. Depending on the duration of milling, superdisintegrant particles can exist as a binary mixture of colloidal and micron-sized particles [8,27]. The stabilizing action of colloidal superdisintegrant particles was discussed extensively in Azad et al. [212], and will not be covered here. The aforementioned studies [8,27] and recent work [87,242] have demonstrated that (i) wet-milled CCS/SSG at 10% w.r.t. drug allows for replacement of an anionic surfactant (SDS) without significant deterioration in drug release rate and drug loading; (ii) even at 10% w.r.t. drug, they are superior to Mannitol at 100%, a commonly used WSD; (iii) the positive impact of the superdisintegrants correlated positively with their swelling capacity, thus signifying a swelling-induced erosion/disintegration as their action mechanism; and (iv) high drug-loaded (>65%), redispersible, fast-dissolving surfactant-free nanocomposites can be prepared with the use of soluble polymer–superdisintegrant combinations.

3.4. Impact of Drying Method

Eerdenbrugh et al. [34] compared freeze drying and spray drying of 9 model compounds and concluded that dissolution performance was dependent on the drug and its formulation, and that the drying method did not have a significant impact. Bhakay et al. [85] compared redispersion and dissolution performance of griseofulvin and azodicarbonamide nanocomposite particles formed by spray drying versus fluid bed coating/drying. They concluded that both drying methods yielded nanocomposite particles that can be rapidly redispersed into nanoparticles when polymer–surfactant (HPC–SDS) combination was used as stabilizers/dispersants (Figure 10). Sievens et al. [79] wet-cast and dried griseofulvin nanosuspensions to prepare redispersible strip films. The formulation of griseofulvin nanosuspension used by Sievens et al. [79] was very similar that used by Bhakay et al. [8], which allows a comparison of fluid bed coating vs. wet film casting–drying. All these findings signify that the impact of different drying methods on redispersibility–drug dissolution appear to be less than that of the drug–dispersant formulation. To put it differently, redispersible, fast-dissolving nanocomposites can be prepared using any drying method by judicious selection of dispersants’ type/concentration. However, the selection of a drying process is still important to downstream processability of the dried intermediate and scalability, cost, manufacturability, etc., which will be discussed in Section 4. Typical drug loadings in nanocomposite particles formed by fluid bed coating are 10–50% w/w, whereas higher drug loadings, i.e., as high as ~90% w/w, can be obtained via freeze and spray drying. Knieke et al. [152] and Bhakay et al. [85] have produced nanocomposite particles with 20% and 48% drug loadings, respectively, by fluid bed coating. In both studies, at higher drug loadings, agglomeration of the coated carrier particles occurred, but the nanocomposites still redispersed back into drug nanoparticles. Even though the fluid bed coating/drying process could form some granules, the combination of polymer–surfactant as dispersants prevented the formation of drug nanoparticle aggregation in the granules, and allowed complete redispersion, as seen in Figure 10. Spray-dried powders, even with 77% drug loading, also released drug nanoparticles without aggregation in the redispersion test (Figure 10).

3.5. Incorporating Nanocomposite Intermediate into Final Oral Solid Dosage Forms

Compared to the voluminous information available on the production of drug nanosuspensions and their drying into nanocomposites, relatively scant information is available on the downstream processing of the nanocomposites into final solid dosages such as tablets, capsules, sachets, etc. [55]. In this section, we review some of the representative publications. Basa et al. [65] studied the development of a tablet formulation containing ketoconazole nanoparticles. Ketoconazole nanosuspension stabilized with HPMC–SDS was produced using a wet media mill and fluid bed coated onto lactose monohydrate carrier particles. The layered particles were compressed into tablets, keeping the formulation’s physical properties and dimensional characteristics similar to a marketed tablet formulation. The results showed that the nanocomposites were redispersible, and their tablet exhibited faster dissolution than a marketed tablet. However, drug release from tablets was significantly slower than that from nanosuspensions, which signifies the importance of tabletting formulation–process parameters. Kumar et al. [192] wet media milled lurasidone hydrochloride with HPMC E3–polysorbate and Poloxamer F188 separately, and sprayed the resulting nanosuspensions on mannitol–microcrystalline cellulose into a fluid bed to prepare granules. The granules were compressed, along with tabletting excipients, into an orally disintegrating tablet with <30 s disintegration time; this was shown to significantly enhance drug release compared to a tablet of the as-received drug of identical composition.
Nekkanti et al. [209] prepared a nanosuspension by WMM of candesartan with HPMC–SDS as stabilizers and spray dried the nanosuspension following the addition of mannitol to prepare nanocomposite particles. The spray-dried powder was then directly compressed into tablets, followed by blending with various tabletting excipients. The results showed that the spray-dried nanocomposites were redispersible, and their tablets showed significantly faster dissolution than tablets with a micronized drug of identical formulation and a commercial tablet formulation. Dolenc et al. [243] prepared a celecoxib nanosuspension using PVP–SDS as stabilizer via the emulsion-diffusion method, spray-dried the nanosuspension, blended the dried powder with microcrystalline cellulose, and compressed it into tablets. Tablets with celecoxib nanoparticles dissolved significantly faster than those with micron-sized celecoxib particles.
Tuomela et al. [123] performed a formulation optimization study on tableting of freeze-dried nanosuspensions of itraconazole and indomethacin that were prepared by WMM and stabilized with Poloxamer F127 and Poloxamer F68, respectively. Freeze-dried powders (nanocomposite particles) of both drugs were compressed into tablets via direct compression and wet granulation–compression using various excipients for granulation and tabletting. All the nanocomposite powders dissolved immediately, while the dissolution was slower from tablets; tablet dissolution rates decreased with an increase in the loading of nanocomposite particles in tablets. As the nanocomposite particle loading in tablets increased, more nanoparticle contact was established, which led to a decrease in the porosity of the tablets. Due to decreased porosity, the dissolution medium cannot enter the dense structure, causing longer disintegration and slower dissolution. Hence, an optimum loading of nanocomposite particles in tablets must be used to obtain tablets with suitable strength and dissolution properties, while the advantages of large surface areas of nanoparticles are retained. For both drugs, an optimum concentration of 40% w/w nanocomposite particles in tablets gave favorable dissolution results. Mauludin et al. [244] prepared a rutin nanosuspenion with SDS as a stabilizer via HPH and freeze-dried it. The powder was blended with standard tabletting excipients and then compressed into tablets. The results showed that the drug release rate from the rutin-nanocrystal loaded tablet was faster than those of a rutin-microcrystal loaded tablet and a marketed tablet.
Besides tablets, drug nanoparticle-laden strip films have been developed recently. Sievens et al. [79], Susarla et al. [131], and Krull et al. [81,82,83], have performed extensive formulation and process optimization studies to manufacture redispersible strip films. The drug release rate is slower in strip films than in the precursor drug nanosuspensions, as the polymeric film provides a diffusion barrier while it swells–erodes–dissolves. Krull et al. [83] investigated the impact of drug loading on film properties and the redispersion/dissolution of griseofulvin nanoparticles. Approximately 40% w/w griseofulvin (GF) loading was found to be the upper limit that yields films with acceptable mechanical and dissolution properties. Drug loadings of 50 wt.% and 73 wt.% were achieved in HPMC E15 and E4M films, respectively; however, films with drug loadings above 40–50 wt.% were unacceptably brittle. At drug loadings of 10–30 wt.%, redispersion and dissolution properties of the GF films were good. Poor redispersion and significantly slower GF release were observed at drug loadings above 50 wt.%. These results suggest that the greatest barriers to producing pharmaceutical films with high loadings of poorly water-soluble drug nanoparticles are overcoming poor film mechanical properties, and ensuring the recovery of the embedded drug nanoparticles, both of which can be conceivably overcome by further formulation development.

4. Some Engineering Considerations for Rational Selection of a Drying Process

For early development/formulation screening and small/bench scale production of nanocomposites, any of the drying processes in Section 3 can be used. The studies reported in Table 2 used small/lab-scale drying equipment. Figure 5 shows that 45% of the studies used freeze drying and 25% used spray drying, which can be explained by their material-sparing nature and cheaper/easier access to small-scale equipment. The popularity of freeze drying and spray drying was also indicated in a previous review paper (Chin et al. [55]). While fluid bed drying (coating or granulation) and nanoextrusion can also be used, they will likely use more material/drug for formulation development. We see a diverging trend between the published, small-scale work and industrial practice in large-scale development and manufacturing; there is no marketed product with freeze-drying. Triglide® was produced using spray drying, and Emend®, Tricor® and Rapamune® were produced using fluid bed coating/drying, wet granulation, and tablet coating, respectively. Chin et al. [55] argues that one major reason for the disagreement between academic work and industrial practice, especially at large/commercial scales, is the poor flowability and low bulk density of the spray-dried and freeze-dried products, which demand further unit operations to obtain suitable material characteristics for compression. However, we assert that there are many factors that can influence the selection of a drying process, especially for late-stage development and manufacturing. These factors include operational/engineering considerations such as cost, scalability, cycle time, existing equipment, and available capacity, as well as the prevalent culture in companies, which received much less attention in a unified perspective in pharmaceutical literature.
Freeze-drying is a promising technique for temperature sensitive or thermally labile drugs, and is ideally suited for vaccines and biologics. When used for drying drug nanosuspensions, the resulting powders are porous, which allows for fast disintegration and redispersion [41]. On the other hand, the product is usually in the form of a cake or lumpy powder that needs to be milled/sieved to obtain a powder. Another disadvantage of such powders is the poor flowability and low bulk density, which requires further unit operations such as granulation and addition of a considerable amount of excipients to achieve suitable material characteristics for compression [55,245]. Most importantly, long processing/cycle times, i.e., on the order of day(s), as well as high cost and energy consumption [196,246], render freeze drying unfavorable for commercial-scale production for drying nanosuspensions of small molecule drugs [55,208,246]. Based on all these considerations, we maintain that freeze drying will not be of choice for large-scale development and manufacturing, unless the drug is thermally labile or extremely temperature/shear sensitive.
Spray drying is preferred over freeze-drying, especially for large-scale development and manufacturing, because it is an inherently continuous one-step process that consumes less time and energy [243,247]. In general, spray drying yields powders with rather low densities and poor flowability [55], which requires additional processing steps, such as roller compaction to obtain tablettable intermediate [248]. These challenges can be managed with appropriate downstream processing steps and the use of tabletting excipients; however, some operational and commercial aspects must be considered before process selection. The feed liquid to a spray drier can be an aqueous drug nanosuspension (top down methods) or nanoparticles in water–solvent mixture (e.g., antisolvent precipitation). Spray drying of aqueous nanosuspensions is less challenging, as air can be used as drying medium, the residual solvent is harmless, and water vapor can be exhausted into the atmosphere directly, whereas spray drying of suspensions with organic solvents requires direct integration of the anti-solvent precipitation equipment with the spray drier (see e.g., [62,141]) and the use of nitrogen or other inert gases, along with specialized solvent recovery systems. It must be noted that large-scale spray driers require specialized facilities, and there are few available pharmaceutical spray drying facilities in the market for product and process development purposes [249]. Hence, the selection of spray drying entails consideration of available plants (in-house vs. CMOs) with appropriate equipment, available capacity, and potential new investment in expertise, installation, and operation of capital-intensive large-scale spray driers and facilities.
Fluid bed coating of a drug nanosuspension onto inert beads such as sugar, cellulose, or other inert excipients, i.e., bead-layering, appears to have several advantages, such good flowability and high bulk density of the coated powders compared with freeze-dried and spray-dried powders. As a result, downstream processing of these powders, such as direct filling into capsules/sachets or compressing them into tablets with additional excipients, is simpler. In addition, fluid bed coating is a versatile process and can also be used to apply a second functional coating onto the beads, which could, for example, afford enteric or mucoadhesive properties [64,248]. Although fluid bed coating has certain disadvantages such as limited drug load achievable and the rather long coating time [250], it has already been used for several oral dosage forms on the market, including Emend. It is also possible to granulate excipients using the drug nanosuspension as “binder solution”, with additional binder polymer if needed, and to produce a nanocomposite in the form of granulated powder, which also has good flowability and high bulk density [192]. Fluid bed granulation is much faster than fluid bed coating because high spray rates can be used to reduce cycle time, which favors particle agglomeration as a desirable transformation unlike in coating. From an operational/engineering perspective, fluid bed processing is the best available option to most pharmaceutical companies, as fluid bed drying is a standard pharmaceutical unit operation as part of wet granulation process train, and even traditional fluid bed driers can easily be retrofitted to perform coating/granulation, i.e., with the addition of spray guns, without major capital investment.
Spray drying and fluid bed coating/granulation/drying processes entail atomization of the feed drug nanosuspensions into droplets. Generating small droplets during atomization puts a limit on the viscosity; highly viscous suspensions that have high solids loading (drug nanoparticles and dissolved species) and/or high MW polymers or swellable dispersants cannot be processed without dilution. Practically, certain polymer types and MWs are not conducive to these spraying processes, which restrict the choice of polymers/polymer grades. Being a continuous process and having the capability to handle viscous polymers, a recently-developed nanoextrusion process could be advantageous for drying drug nanosuspensions [75,76,78]. The product, in the form of extrudates, is dry milled into a powder, and the powder can be blended with excipients for tabletting. For the nanoextrusion process to be thermally efficient and the extrudates to have sufficiently low moisture content, the solids loading in the feed must be maximized. Drug nanosuspensions prepared with media milling can have up to ~50% drug loading, whereas bottom-up methods tend to produce dilute nanosuspensions (<10% drug loading) [23]. Hence, drug nanosuspensions may need to be filtered prior to nanoextrusion, and this aspect, as well as downstream processing of the extrudates into tablets, warrants further investigation.
Wet film casting–drying of a precursor drug–polymer suspension, which is prepared by mixing a drug nanosuspension with a film-forming polymer–plasticizer solution, yields drug nanoparticle-laden polymer strip films. This relatively new continuous drying method has shown itself to be promising [81,82,83], and, like the nanoextrusion process, it can handle suspensions that may not be effectively atomized without dilution. Unlike all the previously mentioned drying methods that eventually produce a nanocomposite powder/extrudate, this method yields a completely different dosage format, i.e., ~50–200 μm strip films that disintegrate and release the drug rapidly in the oral cavity. Such films can be taken with little to no water, and thus, have significant advantages over tablets regarding patient compliance for the geriatric–pediatric–dysphagic population.

5. Additional Insights into Development of Redispersible, Fast-Dissolving Nanocomposites

Unlike for BCS Classes I and III drugs (high solubility), the ideal drug release for poorly water-soluble drugs (BCS Class II drugs) from solid oral dosages containing nanocomposites is generally the one that corresponds to the fastest drug release in vitro, because the low drug solubility and ensuing slow dissolution of such drugs is the rate limiting step in absorption through the gastrointestinal tract. It is well-known that solid oral dosages containing drug-laden nanocomposites of poorly water-soluble drugs exhibit significantly higher drug bioavailability and exposure in in vivo testing on animals and/or clinical testing on humans compared with dosages containing the as-received micron-sized drug particles [113,209,251,252,253]. Hence, the major formulation criterion for developing nanocomposites incorporating poorly water-soluble drug nanoparticles and final solid dosages containing drug-laden nanocomposites has been to attain significantly faster in vitro drug dissolution than that from as-received micron-sized drug particles, their physical mixtures with the excipients of the nanocomposites, and finally, the same oral solid dosages containing as-received micron-sized drug particles (e.g., [66,85,86,87]). In fact, the use of this criterion has already been illustrated with several examples throughout this review paper, especially in Section 3.5, for final oral dosages prepared with various drying methods. Ideally, the fastest dissolution profile with drug nanoparticles is attained upon dosing a nanosuspension, which is not a desirable dosage form from a patient preference/compliance perspective. Unlike drug release from nanosuspensions, drug release from nanocomposites or final solid dosage forms involves the rate-limiting step of nanoparticle recovery during in vitro or in vivo redispersion–dissolution; therefore, the drug release is slower from nanocomposites than nanosuspension. A comparative assessment of in vitro drug release from nanocomposites (dried nanosuspension)/final solid dosages, with respect to that from a nanosuspension, may be used to gauge the effectiveness of the dissolution rate enhancement upon the use of nanocomposites. Considering that some of the biggest challenges in dissolution rate enhancement originate from poor and slow redispersion of the nanocomposites leading to slow recovery of drug nanoparticles [66,92], redispersible drug nanocomposite formulations offer a promising solution, as discussed below.
Section 3 summarized findings of various studies which examine the impact of various classes of dispersants on redispersion–drug release from the nanocomposites. It should be noted that there is no “universal formulation approach” that allows for proper stabilization of all drug nanosuspensions and redispersible, fast drug release from the nanocomposites. Chemical structure, solubility, and hydrophobicity/logP of drugs are different, which need different wetting agents and stabilizers to stabilize the drug nanoparticles [23]. Moreover, aggregation of drug nanoparticles during drying is a complex thermodynamic–kinetic phenomenon that cannot be solely predicted based on the properties of the drugs or drug–dispersant pairs. For example, it is well-known that transport processes such as drying rate and diffusion rate of dispersants during the evaporation [89], as well changing viscosity of the suspensions during drying, could have an impact on the nanoparticle aggregation and redispersibility [74].
Eerdenbrugh et al. [34] attempted to correlate the dissolution of 9 wet media milled poorly water-soluble drugs namely cinnarizine, griseofulvin, indomethacin, itraconazole, loviride, mebendazole, naproxen, phenylbutazone and phenytoin, with their logP alone. TPGS was used as the only stabilizer/dispersant. All nanosuspensions were spray-dried and compared for dissolution performance, taking the dissolution of non-nanosized drug product as basis of comparison. The SLS concentration in the aqueous dissolution medium, corresponding to identical drug compound solubility of 0.375 mg/mL, was determined from the linear fit of the concentration of the solubilized compound as a function of SLS concentration, taking into account the 0.5, 1 and 2% (w/v) SLS data points. Keeping the same drug solubiliy in the dissolution medium, the impact of drug logP was investigated. The results show that drugs with high logP values such as phenylbutazone (4.2), itraconazole (8.5), and cinnarizine (6.1) exhibited slower dissolution. For indomethacin (3.3), loviride (3.7), and phenytoin (2.3), dissolution was not compromised upon drying of the drug nanosuspensions. The more hydrophobic drugs were presumed to form agglomerates (irreversible aggregates) during drying, which reduced the dissolution rate. Hence, Eerdenbrugh et al. [34] suggest that for such drugs, additional matrix formers (dispersants) are required to ensure rapid dissolution. While this study appears to be unique in the literature, providing excellent “general guidance” about drugs with high logP, several caveats must be noted. First, for griseofulvin (2.2), mebendazole (3.3), and naproxen (3.0), the dissolution results were inconclusive due to poor discrimination between nanosized and non-nanosized products. Despite having the lowest logP in all drugs studied in ref. [34], griseofulvin nanosuspensions require a soluble polymer–surfactant combination to prevent aggregate formation during drying [85,86]. Second, to keep identical solubility in the dissolution medium, Eerdenbrugh et al. [34] varied the SLS concentration in the dissolution medium, which could have affected the drug wettability and drug release during dissolution. Such an effect could confound results vis a vis the role of logP alone on the formation of agglomerates and its impact on the dissolution. In fact, it is speculated that just the mere presence of SLS and ensuing improved drug wettability of the nanocomposites could have positively impacted the dissolution rate. Such an impact of SLS in the redispersion–dissolution medium on the redispersion of drug-laden nanocomposites and drug release has been demonstrated elsewhere [66,86]. Finally, actual redispersion of the nanocomposites was not studied in ref. [34], which could have shed more light on the dissolution results. Even Eerdenbrugh et al. claimed that logP can be used as a quick, but “rough” prediction tool. It is clear that further research is needed to understand the impact of drug properties before general conclusions can be drawn as to whether just a surfactant, like TPGS, or a polymer alone would suffice, or whether additional dispersants are required to prevent the aggregation of a given drug.
Our analysis of Table 2 reveals that 73% (88/120) of the formulations were subjected to redispersion testing, and 63/120 formulations were subjected to both redispersion and dissolution testing. Out of 63 formulations, 60 formulations that exhibited the best redispersibility also exhibited the best dissolution performance. This finding is not surprising, as redispersion and drug nanoparticle recovery is the preliminary step in drug release from drug-laden nanocomposites; redispersion testing could be used to make inferences on the formation of irreversible aggregates during drying. Also, as indicated in Section 3.2.4, during 2012–2017, there is a resurgence of interest in redispersion phenomena: 73% vs. 40% [55] of the formulations reported were subjected to redispersion. Hence, both quiescent and agitated redispersion tests can be used to “predict” fast drug dissolution and rank-order dispersant formulations [8,66,92]. In fact, the relationship between redispersion and dissolution can be quantitative, as demonstrated in Figure 11 [8]. A good correlation between griseofulvin (GF) released in 2 min from drug nanocomposites with various dispersant formulations during a dissolution test and percentage of nanoparticles released during an agitated redispersion test was obtained. Similar correlations were obtained between dissolution test results and measured turbidity in a quiescent redispersion test [92].
Analysis of the redispersible formulations in Table 2 allows us to assess which dispersant classes are widely used in such formulations (refer to Figure 8). Soluble polymers and surfactants are the most commonly used dispersants, whereas water-insoluble dispersants and crosslinked polymers were used in fewer formulations. Most redispersible formulations have a combination of different dispersant classes; polymers alone (12%) or surfactants alone (8%) are not as widely used. Polymer–surfactant are the most used dispersants in redispersible formulations, followed by polymer–surfactant–WSD (mostly sugar/sugar alcohols). This finding is not surprising, because the use of multiple classes of dispersants could have a synergistic or additive impact on prevention of aggregation in the nanosuspensions during preparation/storage and during drying, as discussed in Section 3. As a novel class of dispersants in drug-laden nanocomposites, crosslinked polymers such as wet-milled superdisintegrants have recently been shown to be superior to WSDs like mannitol and sucrose in the preparation of surfactant-free nanocomposites. Considering that higher dispersant concentrations usually favor redispersion and drug release at the expense of drug loading, a formulation optimization must be performed following initial screening of polymer–surfactant or polymer–surfactant–WSD. In the screening, various quiescent or agitated redispersion methods can be used, because those formulations that do not redisperse in water or other biorelevant fluids will most likely lead to inferior dissolution enhancement.

6. Concluding Remarks and Future Research Directions

This comprehensive review paper provides a holistic view of drug nanosuspension and nanocomposite formation methods, selection of dispersants based on their functionalities in view of the statistical analysis of 92 publications from 2012–2017, characterization methods, and novel approaches in formulation and process development. The statistical analysis shows that wet media milling is the most commonly-used method for the preparation of drug nanosuspensions in both academic studies and industrial practice/marketed products. On the other hand, we note a divergence of the choice of a drying method in academia vs. industry for the preparation of drug-laden nanocomposites, which can be explained by the development scale, operational/engineering considerations such as cost, scalability, cycle time, existing equipment and available capacity as well as the prevalent culture in companies, which has received much less attention in the pharmaceutical literature. Hence, the selection of a drying method should take these factors into account, as well as the intended final dosage form and delivery route, physico-chemical properties, and degradation characteristics of the drug, and its target patient population.
The following general guidance and considerations for robust formulation development of fast-dissolving nanocomposites emerge from this review: (i) preparation of stabilized, aggregate-free nanosuspensions is necessary, but insufficient for significant dissolution enhancement; (ii) the type–concentration of dispersants must be selected to ensure the stability of the nanosuspensions prior to drying, as well as to minimize aggregation during the drying and fast recovery of the primary drug nanoparticles during dissolution; (iii) in general, higher dispersant concentration favors both drug nanosuspension stabilization, except for some surfactants, and faster redispersion–drug release at the expense of reduced drug payload in the nanocomposites; (iv) thus, usage of dispersants must be optimized to ensure sufficiently high drug payload; (v) the combined use of different types of dispersants with various functionalities, especially soluble polymer–surfactant, soluble polymer–surfactant–WSD or soluble polymer–superdisintegrant (for surfactant-free formulations), may help to enhance nanosuspension stabilization and dissolution, and could even bring synergistic improvements; and (vi) analysis of drug logP and drug wettability enhancement by dispersants, as well as redispersion tests, could help in selecting robust redispersible formulations. This review paper has also demonstrated the criticality of the redispersion phenomenon and the use of various agitated–quiescent redispersion tests, in addition to standardized dissolution tests, for robust and streamlined development of redispersible, fast-dissolving drug-laden nanocomposites. The use of redispersion testing has increased recently, and it is expected that such tests will be used as commonly as dissolution tests in nanoformulation development.
Future research should be directed to a more fundamental understanding of aggregate formation during nanosuspension drying, including capillary pressure mechanism and nanoparticle phase separation phenomena, impact of drug properties, as well as drug–dispersant interactions for both nanosuspension stabilization and drying, and development of “standardized” redispersion tests for the screening of various nanocomposite formulations. This review also points to a significant need for more research on the incorporation of the nanocomposites in standard oral solid dosage forms, such as tablets and capsules, as well as on relatively new oral solid dosages, such as drug nanoparticle-laden polymeric strip films and extrudate-based products.

Author Contributions

M.R. conducted the literature search and prepared the tables. A.B. performed the statistical analysis of the data from the literature, prepared the figures, and wrote the first draft of the manuscript along with E.B. R.N.D. helped revised the manuscript, provided feedback, and valuable suggestions. E.B. prepared the final manuscript and is responsible for the overall quality and content of the manuscript as the lead and corresponding author. Finally, except M.R., all authors have contributed materially and intellectually through several joint papers on the preparation of drug nanoparticles and drug nanocomposites, which helped to shape the intellectual basis of the current review paper.

Funding

This research was funded by Engineering Research Center for Structured Organic Particulate Systems of the U.S. National Science Foundation grant number EEC-0540855.

Acknowledgments

We would like to thank previous members of our research group Afolawemi Afolabi, Lu Zhang, Lucas Sievens-Figueroa, Meng Li, Mohammad Azad, and Scott Krull, who contributed to the knowledge base behind this review paper over the last decade.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lipinski, C.A. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods 2000, 44, 235–249. [Google Scholar] [CrossRef]
  2. Lipinski, C. Poor aqueous solubility—An industry wide problem in drug discovery. Am. Pharm. Rev. 2002, 5, 82–85. [Google Scholar]
  3. Keserü, G.M.; Makara, G.M. The influence of lead discovery strategies on the properties of drug candidates. Nat. Rev. Drug Discov. 2009, 8, 203–212. [Google Scholar] [CrossRef] [PubMed]
  4. Kalepu, S.; Nekkanti, V. Insoluble drug delivery strategies: Review of recent advances and business prospects. Acta Pharm. Sin. B 2015, 5, 442–453. [Google Scholar] [CrossRef] [PubMed]
  5. Fasano, A. Innovative strategies for the oral delivery of drugs and peptides. Trends Biotechnol. 1998, 16, 152–157. [Google Scholar] [CrossRef]
  6. Müllertz, A.; Ogbonna, A.; Ren, S.; Rades, T. New perspectives on lipid and surfactant based drug delivery systems for oral delivery of poorly soluble drugs. J. Pharm. Pharmacol. 2010, 62, 1622–1636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Niwa, T.; Danjo, K. Design of self-dispersible dry nanosuspension through wet milling and spray freeze-drying for poorly water-soluble drugs. Eur. J. Pharm. Sci. 2013, 50, 272–281. [Google Scholar] [CrossRef] [PubMed]
  8. Bhakay, A.; Azad, M.; Vizzotti, E.; Dave, R.N.; Bilgili, E. Enhanced recovery and dissolution of griseofulvin nanoparticles from surfactant-free nanocomposite microparticles incorporating wet-milled swellable dispersants. Drug Dev. Ind. Pharm. 2014, 40, 1509–1522. [Google Scholar] [CrossRef] [PubMed]
  9. Humberstone, A.J.; Charman, W.N. Lipid-based vehicles for the oral delivery of poorly water soluble drugs. Adv. Drug Deliv. Rev. 1997, 25, 103–128. [Google Scholar] [CrossRef]
  10. Hauss, D.J.; Fogal, S.E.; Ficorilli, J.V.; Price, C.A.; Roy, T.; Jayaraj, A.A.; Keirns, J.J. Lipid-based delivery systems for improving the bioavailability and lymphatic transport of a poorly water-soluble LTB4 inhibitor. J. Pharm. Sci. 1998, 87, 164–169. [Google Scholar] [CrossRef] [PubMed]
  11. Rautio, J.; Kumpulainen, H.; Heimbach, T.; Oliyai, R.; Oh, D.; Järvinen, T.; Savolainen, J. Prodrugs: Design and clinical applications. Nat. Rev. Drug Discov. 2008, 7, 255–270. [Google Scholar] [CrossRef] [PubMed]
  12. Rumondor, A.C.; Dhareshwar, S.S.; Kesisoglou, F. Amorphous solid dispersions or prodrugs: Complementary strategies to increase drug absorption. J. Pharm. Sci. 2016, 105, 2498–2508. [Google Scholar] [CrossRef] [PubMed]
  13. Nakagami, H. Solid dispersions of indomethacin and griseofulvin in non-porous fumed silicon dioxide, prepared by melting. Chem. Pharm. Bull. 1991, 39, 2417–2421. [Google Scholar] [CrossRef]
  14. Serajuddin, A. Solid dispersion of poorly water-soluble drugs: Early promises, subsequent problems, and recent breakthroughs. J. Pharm. Sci. 1999, 88, 1058–1066. [Google Scholar] [CrossRef] [PubMed]
  15. Rahman, Z.; Zidan, A.S.; Samy, R.; Sayeed, V.A.; Khan, M.A. Improvement of physicochemical properties of an antiepileptic drug by salt engineering. AAPS PharmSciTech 2012, 13, 793–801. [Google Scholar] [CrossRef] [PubMed]
  16. Elder, D.P.; Holm, R.; de Diego, H.L. Use of pharmaceutical salts and cocrystals to address the issue of poor solubility. Int. J. Pharm. 2013, 453, 88–100. [Google Scholar] [CrossRef] [PubMed]
  17. Schultheiss, N.; Newman, A. Pharmaceutical cocrystals and their physicochemical properties. Cryst. Growth Des. 2009, 9, 2950–2967. [Google Scholar] [CrossRef] [PubMed]
  18. Yadav, A.; Shete, A.; Dabke, A.; Kulkarni, P.; Sakhare, S. Co-crystals: A novel approach to modify physicochemical properties of active pharmaceutical ingredients. Indian J. Pharm. Sci. 2009, 71, 359–370. [Google Scholar] [CrossRef] [PubMed]
  19. Aleem, O.; Kuchekar, B.; Pore, Y.; Late, S. Effect of β-cyclodextrin and hydroxypropyl β-cyclodextrin complexation on physicochemical properties and antimicrobial activity of cefdinir. J. Pharm. Biomed. Anal. 2008, 47, 535–540. [Google Scholar] [CrossRef] [PubMed]
  20. Srivalli, K.M.R.; Mishra, B. Improved aqueous solubility and antihypercholesterolemic activity of ezetimibe on formulating with hydroxypropyl-β-cyclodextrin and hydrophilic auxiliary substances. AAPS PharmSciTech 2016, 17, 272–283. [Google Scholar] [CrossRef] [PubMed]
  21. Bhakay, A.; Merwade, M.; Bilgili, E.; Dave, R.N. Novel aspects of wet milling for the production of microsuspensions and nanosuspensions of poorly water-soluble drugs. Drug Dev. Ind. Pharm. 2011, 37, 963–976. [Google Scholar] [CrossRef] [PubMed]
  22. Monteiro, A.; Afolabi, A.; Bilgili, E. Continuous production of drug nanoparticle suspensions via wet stirred media milling: A fresh look at the Rehbinder effect. Drug Dev. Ind. Pharm. 2013, 39, 266–283. [Google Scholar] [CrossRef] [PubMed]
  23. Li, M.; Azad, M.; Davé, R.; Bilgili, E. Nanomilling of drugs for bioavailability enhancement: A holistic formulation-process perspective. Pharmaceutics 2016, 8, 17. [Google Scholar] [CrossRef] [PubMed]
  24. Müller, R.H.; Benita, S.; Böhm, B.H. Nanosuspensions. In Emulsions and Nanosuspensions for the Formulation of Poorly Soluble Drugs; Benita, S., Böhm, B.H., Eds.; Medpharm Scientific: Stuttgart, Germany, 1998; pp. 149–173. [Google Scholar]
  25. Noyes, A.A.; Whitney, W.R. The rate of solution of solid substances in their own solutions. J. Am. Chem. Soc. 1897, 19, 930–934. [Google Scholar] [CrossRef]
  26. Zhang, X.; Guan, J.; Ni, R.; Li, L.C.; Mao, S. Preparation and solidification of redispersible nanosuspensions. J. Pharm. Sci. 2014, 103, 2166–2176. [Google Scholar] [CrossRef] [PubMed]
  27. Azad, M.; Arteaga, C.; Abdelmalek, B.; Davé, R.; Bilgili, E. Spray drying of drug-swellable dispersant suspensions for preparation of fast-dissolving, high drug-loaded, surfactant-free nanocomposites. Drug Dev. Ind. Pharm. 2015, 41, 1617–1631. [Google Scholar] [CrossRef] [PubMed]
  28. Kesisoglou, F.; Panmai, S.; Wu, Y. Nanosizing—Oral formulation development and biopharmaceutical evaluation. Adv. Drug Deliv. Rev. 2007, 59, 631–644. [Google Scholar] [CrossRef] [PubMed]
  29. Junghanns, J.-U.A.; Müller, R.H. Nanocrystal technology, drug delivery and clinical applications. Int. J. Nanomed. 2008, 3, 295–309. [Google Scholar] [Green Version]
  30. Srivalli, K.M.R.; Mishra, B. Drug nanocrystals: A way toward scale-up. Saudi Pharm. J. 2016, 24, 386–404. [Google Scholar] [CrossRef] [PubMed]
  31. Malamatari, M.; Taylor, K.M.; Malamataris, S.; Douroumis, D.; Kachrimanis, K. Pharmaceutical nanocrystals: Prodcution by wet media milling and applications. Drug Discov. Today 2018, 23, 534–547. [Google Scholar] [CrossRef] [PubMed]
  32. Muller, R.; Bohm, B.; Grau, J. Nanosuspensions: A formulation approach for poorly soluble and poorly bioavailable drugs. Handb. Pharm. Control. Release Technol. 2000, 17, 345–357. [Google Scholar]
  33. Lee, J. Drug nano-and microparticles processed into solid dosage forms: Physical properties. J. Pharm. Sci. 2003, 92, 2057–2068. [Google Scholar] [CrossRef] [PubMed]
  34. Van Eerdenbrugh, B.; Froyen, L.; Van Humbeeck, J.; Martens, J.A.; Augustijns, P.; Van den Mooter, G. Drying of crystalline drug nanosuspensions—The importance of surface hydrophobicity on dissolution behavior upon redispersion. Eur. J. Pharm. Sci. 2008, 35, 127–135. [Google Scholar] [CrossRef] [PubMed]
  35. Liversidge, G.G.; Cundy, K.C.; Bishop, J.F.; Czekai, D.A. Surface Modified Drug Nanoparticles. U.S. Patent 5,145,684, 8 September 1992. [Google Scholar]
  36. Beck, C.; Dalvi, S.V.; Dave, R.N. Controlled liquid antisolvent precipitation using a rapid mixing device. Chem. Eng. Sci. 2010, 65, 5669–5675. [Google Scholar] [CrossRef]
  37. De Zordi, N.; Moneghini, M.; Kikic, I.; Grassi, M.; Castillo, A.E.D.R.; Solinas, D.; Bolger, M.B. Applications of supercritical fluids to enhance the dissolution behaviors of Furosemide by generation of microparticles and solid dispersions. Eur. J. Pharm. Biopharm. 2012, 81, 131–141. [Google Scholar] [CrossRef] [PubMed]
  38. Imperiale, J.C.; Bevilacqua, G.; Rosa, P.d.T.V.e.; Sosnik, A. Production of pure indinavir free base nanoparticles by a supercritical anti-solvent (SAS) method. Drug Dev. Ind. Pharm. 2014, 40, 1607–1615. [Google Scholar] [CrossRef] [PubMed]
  39. Knieke, C.; Rawtani, A.; Davé, R.N. Concentrated fenofibrate nanoparticle suspensions from melt emulsification for enhanced drug dissolution. Chem. Eng. Technol. 2014, 37, 157–167. [Google Scholar] [CrossRef]
  40. Bhakay, A.; Vizzotti, E.; Li, M.; Davé, R.; Bilgili, E. Incorporation of fenofibrate nanoparticles prepared by melt emulsification into polymeric films. J. Pharm. Innov. 2016, 11, 53–63. [Google Scholar] [CrossRef]
  41. Salazar, J.; Ghanem, A.; Müller, R.H.; Möschwitzer, J.P. Nanocrystals: Comparison of the size reduction effectiveness of a novel combinative method with conventional top-down approaches. Eur. J. Pharm. Biopharm. 2012, 81, 82–90. [Google Scholar] [CrossRef] [PubMed]
  42. Bruno, J.A.; Doty, B.D.; Gustow, E.; Illig, K.J.; Rajagopalan, N.; Sarpotdar, P. Method of Grinding Pharmaceutical Substances. U.S. Patent No. 5,518,187, 21 May 1996. [Google Scholar]
  43. Sommer, M.; Stenger, F.; Peukert, W.; Wagner, N. Agglomeration and breakage of nanoparticles in stirred media mills—A comparison of different methods and models. Chem. Eng. Sci. 2006, 61, 135–148. [Google Scholar] [CrossRef]
  44. Knieke, C.; Sommer, M.; Peukert, W. Identifying the apparent and true grinding limit. Powder Technol. 2009, 195, 25–30. [Google Scholar] [CrossRef]
  45. Su, J.-C.; Liang, S.Y.; Liu, W.L.; Jan, T.C. Ceramic micro/nanoparticle size evolution in wet grinding in stirred ball mill. J. Manuf. Sci. Eng. 2004, 126, 779–786. [Google Scholar] [CrossRef]
  46. Peukert, W.; Schwarzer, H.-C.; Stenger, F. Control of aggregation in production and handling of nanoparticles. Chem. Eng. Process. Process Intensif. 2005, 44, 245–252. [Google Scholar] [CrossRef]
  47. Cerdeira, A.M.; Mazzotti, M.; Gander, B. Miconazole nanosuspensions: Influence of formulation variables on particle size reduction and physical stability. Int. J. Pharm. 2010, 396, 210–218. [Google Scholar] [CrossRef] [PubMed]
  48. Napper, D.H. Colloid stability. Ind. Eng. Chem. Prod. Res. Dev. 1970, 9, 467–477. [Google Scholar] [CrossRef]
  49. Muller, R. Zetapotential und Partikelladung—Kurze Theorie, Praktische Meûdurchfu Ehrung, Daten Interpretation; Wissenschaftliche Verlagsgesellschaft: Stuttgart, Germany, 1996. [Google Scholar]
  50. Riddick, T.M. Control of Colloid Stability Through Zeta Potential; Zeta-Meter Inc. via Livingston Publishing Company: Lynnewood, PA, USA, 1968. [Google Scholar]
  51. Lakshmi, P.; Kumar, G.A. Nanosuspension technology: A review. Int. J. Pharm. Sci. 2010, 2, 35–40. [Google Scholar]
  52. Kim, C.-J. Advanced Pharmaceutics: Physicochemical Principles; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  53. Ghosh, I.; Bose, S.; Vippagunta, R.; Harmon, F. Nanosuspension for improving the bioavailability of a poorly soluble drug and screening of stabilizing agents to inhibit crystal growth. Int. J. Pharm. 2011, 409, 260–268. [Google Scholar] [CrossRef] [PubMed]
  54. Verma, S.; Kumar, S.; Gokhale, R.; Burgess, D.J. Physical stability of nanosuspensions: Investigation of the role of stabilizers on ostwald ripening. Int. J. Pharm. 2011, 406, 145–152. [Google Scholar] [CrossRef] [PubMed]
  55. Chin, W.W.L.; Parmentier, J.; Widzinski, M.; Tan, E.H.; Gokhale, R. A brief literature and patent review of nanosuspensions to a final drug product. J. Pharm. Sci. 2014, 103, 2980–2999. [Google Scholar] [CrossRef] [PubMed]
  56. Peltonen, L.; Hirvonen, J. Pharmaceutical nanocrystals by nanomilling: Critical process parameters, particle fracturing and stabilization methods. J. Pharm. Pharmacol. 2010, 62, 1569–1579. [Google Scholar] [CrossRef] [PubMed]
  57. Wu, L.; Zhang, J.; Watanabe, W. Physical and chemical stability of drug nanoparticles. Adv. Drug Deliv. Rev. 2011, 63, 456–469. [Google Scholar] [CrossRef] [PubMed]
  58. Tewa-Tagne, P.; Briançon, S.; Fessi, H. Preparation of redispersible dry nanocapsules by means of spray-drying: Development and characterisation. Eur. J. Pharm. Sci. 2007, 30, 124–135. [Google Scholar] [CrossRef] [PubMed]
  59. Van Eerdenbrugh, B.; Froyen, L.; Van Humbeeck, J.; Martens, J.A.; Augustijns, P.; Van Den Mooter, G. Alternative matrix formers for nanosuspension solidification: Dissolution performance and X-ray microanalysis as an evaluation tool for powder dispersion. Eur. J. Pharm. Sci. 2008, 35, 344–353. [Google Scholar] [CrossRef] [PubMed]
  60. Vogt, M.; Vertzoni, M.; Kunath, K.; Reppas, C.; Dressman, J.B. Cogrinding enhances the oral bioavailability of EMD 57033, a poorly water soluble drug, in dogs. Eur. J. Pharm. Biopharm. 2008, 68, 338–345. [Google Scholar] [CrossRef] [PubMed]
  61. Kho, K.; Cheow, W.S.; Lie, R.H.; Hadinoto, K. Aqueous re-dispersibility of spray-dried antibiotic-loaded polycaprolactone nanoparticle aggregates for inhaled anti-biofilm therapy. Powder Technol. 2010, 203, 432–439. [Google Scholar] [CrossRef]
  62. Hu, J.; Ng, W.K.; Dong, Y.; Shen, S.; Tan, R.B. Continuous and scalable process for water-redispersible nanoformulation of poorly aqueous soluble APIs by antisolvent precipitation and spray-drying. Int. J. Pharm. 2011, 404, 198–204. [Google Scholar] [CrossRef] [PubMed]
  63. Ryde, N.P.; Ruddy, S.B. Solid dose Nanoparticulate Compositions Comprising a Synergistic Combination of a Polymeric Surface Stabilizer and Dioctyl Sodium Sulfosuccinate. U.S. Patent 6,375,986, 23 April 2002. [Google Scholar]
  64. Möschwitzer, J.; Müller, R.H. Spray coated pellets as carrier system for mucoadhesive drug nanocrystals. Eur. J. Pharm. Biopharm. 2006, 62, 282–287. [Google Scholar] [CrossRef] [PubMed]
  65. Basa, S.; Muniyappan, T.; Karatgi, P.; Prabhu, R.; Pillai, R. Production and in vitro characterization of solid dosage form incorporating drug nanoparticles. Drug Dev. Ind. Pharm. 2008, 34, 1209–1218. [Google Scholar] [CrossRef] [PubMed]
  66. Bhakay, A.; Davé, R.; Bilgili, E. Recovery of BCS Class II drugs during aqueous redispersion of core-shell type nanocomposite particles produced via fluidized bed coating. Powder Technol. 2013, 236, 221–234. [Google Scholar] [CrossRef]
  67. Cheow, W.S.; Ng, M.L.L.; Kho, K.; Hadinoto, K. Spray freeze-drying production of thermally sensitive polymeric nanoparticle aggregates for inhaled drug delivery: Effect of freeze-drying adjuvants. Int. J. Pharm. 2011, 404, 289–300. [Google Scholar] [CrossRef] [PubMed]
  68. De Waard, H.; Hinrichs, W.; Frijlink, H. A novel bottom-up process to produce drug nanocrystals: Controlled crystallization during freeze-drying. J. Control. Release 2008, 128, 179–183. [Google Scholar] [CrossRef] [PubMed]
  69. Dong, Y.; Ng, W.K.; Shen, S.; Kim, S.; Tan, R.B. Controlled antisolvent precipitation of spironolactone nanoparticles by impingement mixing. Int. J. Pharm. 2011, 410, 175–179. [Google Scholar] [CrossRef] [PubMed]
  70. Li, W.; Yang, Y.; Tian, Y.; Xu, X.; Chen, Y.; Mu, L.; Zhang, Y.; Fang, L. Preparation and in vitro/in vivo evaluation of revaprazan hydrochloride nanosuspension. Int. J. Pharm. 2011, 408, 157–162. [Google Scholar] [CrossRef] [PubMed]
  71. Beirowski, J.; Inghelbrecht, S.; Arien, A.; Gieseler, H. Freeze-drying of nanosuspensions, part 3: Investigation of factors compromising storage stability of highly concentrated drug nanosuspensions. J. Pharm. Sci. 2012, 101, 354–362. [Google Scholar] [CrossRef] [PubMed]
  72. Chung, N.-O.; Lee, M.K.; Lee, J. Mechanism of freeze-drying drug nanosuspensions. Int. J. Pharm. 2012, 437, 42–50. [Google Scholar] [CrossRef] [PubMed]
  73. Choi, J.-Y.; Park, C.H.; Lee, J. Effect of polymer molecular weight on nanocomminution of poorly soluble drug. Drug Deliv. 2008, 15, 347–353. [Google Scholar] [CrossRef] [PubMed]
  74. Kim, S.; Lee, J. Effective polymeric dispersants for vacuum, convection and freeze drying of drug nanosuspensions. Int. J. Pharm. 2010, 397, 218–224. [Google Scholar] [CrossRef] [PubMed]
  75. Khinast, J.; Baumgartner, R.; Roblegg, E. Nano-extrusion: A one-step process for manufacturing of solid nanoparticle formulations directly from the liquid phase. AAPS PharmSciTech 2013, 14, 601–604. [Google Scholar] [CrossRef] [PubMed]
  76. Baumgartner, R.; Eitzlmayr, A.; Matsko, N.; Tetyczka, C.; Khinast, J.; Roblegg, E. Nano-extrusion: A promising tool for continuous manufacturing of solid nano-formulations. Int. J. Pharm. 2014, 477, 1–11. [Google Scholar] [CrossRef] [PubMed]
  77. Ye, X.; Patil, H.; Feng, X.; Tiwari, R.V.; Lu, J.; Gryczke, A.; Kolter, K.; Langley, N.; Majumdar, S.; Neupane, D. Conjugation of hot-melt extrusion with high-pressure homogenization: A novel method of continuously preparing nanocrystal solid dispersions. AAPS PharmSciTech 2016, 17, 78–88. [Google Scholar] [CrossRef] [PubMed]
  78. Li, M.; Ioannidis, N.; Gogos, C.; Bilgili, E. A comparative assessment of nanocomposites vs. amorphous solid dispersions prepared via nanoextrusion for drug dissolution enhancement. Eur. J. Pharm. Biopharm. 2017, 119, 68–80. [Google Scholar] [CrossRef] [PubMed]
  79. Sievens-Figueroa, L.; Bhakay, A.; Jerez-Rozo, J.I.; Pandya, N.; Romañach, R.J.; Michniak-Kohn, B.; Iqbal, Z.; Bilgili, E.; Davé, R.N. Preparation and characterization of hydroxypropyl methyl cellulose films containing stable BCS Class II drug nanoparticles for pharmaceutical applications. Int. J. Pharm. 2012, 423, 496–508. [Google Scholar] [CrossRef] [PubMed]
  80. Susarla, R.; Afolabi, A.; Patel, D.; Bilgili, E.; Davé, R.N. Novel use of superdisintegrants as viscosity enhancing agents in biocompatible polymer films containing griseofulvin nanoparticles. Powder Technol. 2015, 285, 25–33. [Google Scholar] [CrossRef] [Green Version]
  81. Krull, S.M.; Patel, H.V.; Li, M.; Bilgili, E.; Davé, R.N. Critical material attributes (CMAs) of strip films loaded with poorly water-soluble drug nanoparticles: I. Impact of plasticizer on film properties and dissolution. Eur. J. Pharm. Sci. 2016, 92, 146–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Krull, S.M.; Ammirata, J.; Bawa, S.; Li, M.; Bilgili, E.; Davé, R.N. Critical material attributes of strip films loaded with poorly water-soluble drug nanoparticles: II. Impact of polymer molecular weight. J. Pharm. Sci. 2017, 106, 619–628. [Google Scholar] [CrossRef] [PubMed]
  83. Krull, S.M.; Moreno, J.; Li, M.; Bilgili, E.; Davé, R.N. Critical material attributes (CMAs) of strip films loaded with poorly water-soluble drug nanoparticles: III. Impact of drug nanoparticle loading. Int. J. Pharm. 2017, 523, 33–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bhakay, A. Improved Recovery and Dissolution from Dried Stabilized Aqueous Nanomilled Suspensions of Poorly Water-Soluble Drugs. Ph.D. Thesis, New Jersey Institute of Technology, Newark, NJ, USA, 2013. [Google Scholar]
  85. Bhakay, A.; Azad, M.; Bilgili, E.; Dave, R. Redispersible fast dissolving nanocomposite microparticles of poorly water-soluble drugs. Int. J. Pharm. 2014, 461, 367–379. [Google Scholar] [CrossRef] [PubMed]
  86. Li, M.; Lopez, N.; Bilgili, E. A study of the impact of polymer–surfactant in drug nanoparticle coated pharmatose composites on dissolution performance. Adv. Powder Technol. 2016, 27, 1625–1636. [Google Scholar] [CrossRef]
  87. Li, M. Assessment of Nanocomposites Versus Amorphous Solid Dispersions for Dissolution Enhancement of BCS Class II Drugs. Ph.D. Thesis, New Jersey Institute of Technology, Newark, NJ, USA, 2017. [Google Scholar]
  88. Wang, B.; Zhang, W.; Zhang, W.; Mujumdar, A.S.; Huang, L. Progress in drying technology for nanomaterials. Dry. Technol. 2005, 23, 7–32. [Google Scholar] [CrossRef]
  89. Vehring, R. Pharmaceutical particle engineering via spray drying. Pharm. Res. 2008, 25, 999–1022. [Google Scholar] [CrossRef] [PubMed]
  90. Lee, M.K.; Kim, M.Y.; Kim, S.; Lee, J. Cryoprotectants for freeze drying of drug nano-suspensions: Effect of freezing rate. J. Pharm. Sci. 2009, 98, 4808–4817. [Google Scholar] [CrossRef] [PubMed]
  91. Azad, M.; Moreno, J.; Bilgili, E.; Davé, R. Fast dissolution of poorly water soluble drugs from fluidized bed coated nanocomposites: Impact of carrier size. Int. J. Pharm. 2016, 513, 319–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Bhakay, A.; Davé, R.; Bilgili, E. Quiescent and agitated redispersion as a tool for evaluating dispersant effectiveness in dissolution enhancement of drug-laden nanocomposites. AAPS PharmSciTech 2018, 19, 436–447. [Google Scholar] [CrossRef] [PubMed]
  93. Brough, C.; Williams, R.O. Amorphous solid dispersions and nano-crystal technologies for poorly water-soluble drug delivery. Int. J. Pharm. 2013, 453, 157–166. [Google Scholar] [CrossRef] [PubMed]
  94. Chogale, M.M.; Ghodake, V.N.; Patravale, V.B. Performance parameters and characterizations of nanocrystals: A brief review. Pharmaceutics 2016, 8, 26. [Google Scholar] [CrossRef] [PubMed]
  95. Peltonen, L.; Hirvonen, J. Drug nanocrystals—Versatile option for formulation of poorly soluble materials. Int. J. Pharm. 2018, 537, 73–83. [Google Scholar] [CrossRef] [PubMed]
  96. Müller, R.; Jacobs, C.; Kayser, O. Nanosuspensions as particulate drug formulations in therapy: Rationale for development and what we can expect for the future. Adv. Drug Deliv. Rev. 2001, 47, 3–19. [Google Scholar] [CrossRef]
  97. Kwade, A. Wet comminution in stirred media mills—Research and its practical application. Powder Technol. 1999, 105, 14–20. [Google Scholar] [CrossRef]
  98. Merisko-Liversidge, E.; Liversidge, G.G.; Cooper, E.R. Nanosizing: A formulation approach for poorly-water-soluble compounds. Eur. J. Pharm. Sci. 2003, 18, 113–120. [Google Scholar] [CrossRef]
  99. Bilgili, E.; Afolabi, A. A combined microhydrodynamics-polymer adsorption analysis for elucidation of the roles of stabilizers in wet stirred media milling. Int. J. Pharm. 2012, 439, 193–206. [Google Scholar] [CrossRef] [PubMed]
  100. Knieke, C.; Azad, M.; Davé, R.; Bilgili, E. A study of the physical stability of wet media-milled fenofibrate suspensions using dynamic equilibrium curves. Chem. Eng. Res. Des. 2013, 91, 1245–1258. [Google Scholar] [CrossRef]
  101. Bilgili, E.; Li, M.; Afolabi, A. Is the combination of cellulosic polymers and anionic surfactants a good strategy for ensuring physical stability of BCS Class II drug nanosuspensions? Pharm. Dev. Technol. 2016, 21, 499–510. [Google Scholar] [CrossRef] [PubMed]
  102. Afolabi, A.; Akinlabi, O.; Bilgili, E. Impact of process parameters on the breakage kinetics of poorly water-soluble drugs during wet stirred media milling: A microhydrodynamic view. Eur. J. Pharm. Sci. 2014, 51, 75–86. [Google Scholar] [CrossRef] [PubMed]
  103. Li, M.; Yaragudi, N.; Afolabi, A.; Dave, R.; Bilgili, E. Sub-100 nm drug particle suspensions prepared via wet milling with low bead contamination through novel process intensification. Chem. Eng. Sci. 2015, 130, 207–220. [Google Scholar] [CrossRef]
  104. Li, M.; Alvarez, P.; Bilgili, E. A microhydrodynamic rationale for selection of bead size in preparation of drug nanosuspensions via wet stirred media milling. Int. J. Pharm. 2017, 524, 178–192. [Google Scholar] [CrossRef] [PubMed]
  105. Möschwitzer, J.P. Drug nanocrystals in the commercial pharmaceutical development process. Int. J. Pharm. 2013, 453, 142–156. [Google Scholar] [CrossRef] [PubMed]
  106. Shen, B.D.; Shen, C.Y.; Yuan, X.D.; Bai, J.X.; Lv, Q.Y.; Xu, H.; Dai, L.; Yu, C.; Han, J.; Yuan, H.L. Development and characterization of an orodispersible film containing drug nanoparticles. Eur. J. Pharm. Biopharm. 2013, 85, 1348–1356. [Google Scholar] [CrossRef] [PubMed]
  107. Sun, W.; Ni, R.; Zhang, X.; Li, L.C.; Mao, S. Spray drying of a poorly water-soluble drug nanosuspension for tablet preparation: Formulation and process optimization with bioavailability evaluation. Drug Dev. Ind. Pharm. 2015, 41, 927–933. [Google Scholar] [CrossRef] [PubMed]
  108. Liu, D.; Xu, H.; Tian, B.; Yuan, K.; Pan, H.; Ma, S.; Yang, X.; Pan, W. Fabrication of carvedilol nanosuspensions through the anti-solvent precipitation–ultrasonication method for the improvement of dissolution rate and oral bioavailability. AAPS PharmSciTech 2012, 13, 295–304. [Google Scholar] [CrossRef] [PubMed]
  109. Beck, C.; Sievens-Figueroa, L.; Gärtner, K.; Jerez-Rozo, J.I.; Romañach, R.J.; Bilgili, E.; Davé, R.N. Effects of stabilizers on particle redispersion and dissolution from polymer strip films containing liquid antisolvent precipitated griseofulvin particles. Powder Technol. 2013, 236, 37–51. [Google Scholar] [CrossRef]
  110. Freag, M.S.; Elnaggar, Y.S.; Abdallah, O.Y. Development of novel polymer-stabilized diosmin nanosuspensions: In vitro appraisal and ex vivo permeation. Int. J. Pharm. 2013, 454, 462–471. [Google Scholar] [CrossRef] [PubMed]
  111. Elham, G.; Mahsa, P.; Vahid, R. Spray drying of nanoparticles to form fast dissolving glipizide. Asian J. Pharm. 2015, 213–218. [Google Scholar]
  112. Kumar, V.; Prud’homme, R.K. Nanoparticle stability: Processing pathways for solvent removal. Chem. Eng. Sci. 2009, 64, 1358–1361. [Google Scholar] [CrossRef]
  113. Fu, Q.; Ma, M.; Li, M.; Wang, G.; Guo, M.; Li, J.; Hou, Y.; Fang, M. Improvement of oral bioavailability for nisoldipine using nanocrystals. Powder Technol. 2017, 305, 757–763. [Google Scholar] [CrossRef]
  114. Tuomela, A.; Hirvonen, J.; Peltonen, L. Stabilizing agents for drug nanocrystals: Effect on bioavailability. Pharmaceutics 2016, 8, 16. [Google Scholar] [CrossRef] [PubMed]
  115. Van Eerdenbrugh, B.; Van den Mooter, G.; Augustijns, P. Top-down production of drug nanocrystals: Nanosuspension stabilization, miniaturization and transformation into solid products. Int. J. Pharm. 2008, 364, 64–75. [Google Scholar] [CrossRef] [PubMed]
  116. Merisko-Liversidge, E.; Liversidge, G.G. Nanosizing for oral and parenteral drug delivery: A perspective on formulating poorly-water soluble compounds using wet media milling technology. Adv. Drug Deliv. Rev. 2011, 63, 427–440. [Google Scholar] [CrossRef] [PubMed]
  117. Müller, R.; Jacobs, C. Buparvaquone mucoadhesive nanosuspension: Preparation, optimisation and long-term stability. Int. J. Pharm. 2002, 237, 151–161. [Google Scholar] [CrossRef]
  118. Mishra, P.R.; Al Shaal, L.; Müller, R.H.; Keck, C.M. Production and characterization of hesperetin nanosuspensions for dermal delivery. Int. J. Pharm. 2009, 371, 182–189. [Google Scholar] [CrossRef] [PubMed]
  119. Van Eerdenbrugh, B.; Vermant, J.; Martens, J.A.; Froyen, L.; Van Humbeeck, J.; Augustijns, P.; Van den Mooter, G. A screening study of surface stabilization during the production of drug nanocrystals. J. Pharm. Sci. 2009, 98, 2091–2103. [Google Scholar] [CrossRef] [PubMed]
  120. Ho, H.; Lee, J. Redispersible drug nanoparticles prepared without dispersant by electro-spray drying. Drug Dev. Ind. Pharm. 2012, 38, 744–751. [Google Scholar] [CrossRef] [PubMed]
  121. Cerdeira, A.M.; Mazzotti, M.; Gander, B. Formulation and drying of miconazole and itraconazole nanosuspensions. Int. J. Pharm. 2013, 443, 209–220. [Google Scholar] [CrossRef] [PubMed]
  122. Li, M.; Alvarez, P.; Orbe, P.; Bilgili, E. Multi-faceted characterization of wet-milled griseofulvin nanosuspensions for elucidation of aggregation state and stabilization mechanisms. AAPS PharmSciTech 2018, 19, 1789–1801. [Google Scholar] [CrossRef] [PubMed]
  123. Tuomela, A.; Laaksonen, T.; Laru, J.; Antikainen, O.; Kiesvaara, J.; Ilkka, J.; Oksala, O.; Rönkkö, S.; Järvinen, K.; Hirvonen, J. Solid formulations by a nanocrystal approach: Critical process parameters regarding scale-ability of nanocrystals for tableting applications. Int. J. Pharm. 2015, 485, 77–86. [Google Scholar] [CrossRef] [PubMed]
  124. Lee, J.; Lee, S.-J.; Choi, J.-Y.; Yoo, J.Y.; Ahn, C.-H. Amphiphilic amino acid copolymers as stabilizers for the preparation of nanocrystal dispersion. Eur. J. Pharm. Sci. 2005, 24, 441–449. [Google Scholar] [CrossRef] [PubMed]
  125. Lee, J.; Choi, J.-Y.; Park, C. Characteristics of polymers enabling nano-comminution of water-insoluble drugs. Int. J. Pharm. 2008, 355, 328–336. [Google Scholar] [CrossRef] [PubMed]
  126. Choi, J.-Y.; Yoo, J.Y.; Kwak, H.-S.; Nam, B.U.; Lee, J. Role of polymeric stabilizers for drug nanocrystal dispersions. Curr. Appl. Phys. 2005, 5, 472–474. [Google Scholar] [CrossRef]
  127. George, M.; Ghosh, I. Identifying the correlation between drug/stabilizer properties and critical quality attributes (CQAs) of nanosuspension formulation prepared by wet media milling technology. Eur. J. Pharm. Sci. 2013, 48, 142–152. [Google Scholar] [CrossRef] [PubMed]
  128. Gupta, R.B.; Kompella, U.B. Nanoparticle Technology for Drug Delivery; Taylor & Francis: New York, NY, USA, 2006. [Google Scholar]
  129. Zhu, W.; Romanski, F.S.; Dalvi, S.V.; Dave, R.N.; Tomassone, M.S. Atomistic simulations of aqueous griseofulvin crystals in the presence of individual and multiple additives. Chem. Eng. Sci. 2012, 73, 218–230. [Google Scholar] [CrossRef]
  130. Bose, S.; Schenck, D.; Ghosh, I.; Hollywood, A.; Maulit, E.; Ruegger, C. Application of spray granulation for conversion of a nanosuspension into a dry powder form. Eur. J. Pharm. Sci. 2012, 47, 35–43. [Google Scholar] [CrossRef] [PubMed]
  131. Susarla, R.; Sievens-Figueroa, L.; Bhakay, A.; Shen, Y.; Jerez-Rozo, J.I.; Engen, W.; Khusid, B.; Bilgili, E.; Romanach, R.J.; Morris, K.R. Fast drying of biocompatible polymer films loaded with poorly water-soluble drug nano-particles via low temperature forced convection. Int. J. Pharm. 2013, 455, 93–103. [Google Scholar] [CrossRef] [PubMed]
  132. Cheng, K.K.; Yeung, C.F.; Ho, S.W.; Chow, S.F.; Chow, A.H.; Baum, L. Highly stabilized curcumin nanoparticles tested in an in vitro blood–brain barrier model and in alzheimer’s disease Tg2576 mice. AAPS J. 2013, 15, 324–336. [Google Scholar] [CrossRef] [PubMed]
  133. He, W.; Lu, Y.; Qi, J.; Chen, L.; Yin, L.; Wu, W. Formulating food protein-stabilized indomethacin nanosuspensions into pellets by fluid-bed coating technology: Physical characterization, redispersibility, and dissolution. Int. J. Nanomed. 2013, 8, 3119–3128. [Google Scholar]
  134. Fu, Q.; Sun, J.; Zhang, D.; Li, M.; Wang, Y.; Ling, G.; Liu, X.; Sun, Y.; Sui, X.; Luo, C. Nimodipine nanocrystals for oral bioavailability improvement: Preparation, characterization and pharmacokinetic studies. Colloids Surf. B Biointerfaces 2013, 109, 161–166. [Google Scholar] [CrossRef] [PubMed]
  135. Ige, P.P.; Baria, R.K.; Gattani, S.G. Fabrication of fenofibrate nanocrystals by probe sonication method for enhancement of dissolution rate and oral bioavailability. Colloids Surf. B Biointerfaces 2013, 108, 366–373. [Google Scholar] [CrossRef] [PubMed]
  136. Zuo, B.; Sun, Y.; Li, H.; Liu, X.; Zhai, Y.; Sun, J.; He, Z. Preparation and in vitro/in vivo evaluation of fenofibrate nanocrystals. Int. J. Pharm. 2013, 455, 267–275. [Google Scholar] [CrossRef] [PubMed]
  137. Wei, X.-L.; Han, Y.-R.; Quan, L.-H.; Liu, C.-Y.; Liao, Y.-H. Oily nanosuspension for long-acting intramuscular delivery of curcumin didecanoate prodrug: Preparation, characterization and in vivo evaluation. Eur. J. Pharm. Sci. 2013, 49, 286–293. [Google Scholar] [CrossRef] [PubMed]
  138. Kumar, S.; Gokhale, R.; Burgess, D.J. Sugars as bulking agents to prevent nano-crystal aggregation during spray or freeze-drying. Int. J. Pharm. 2014, 471, 303–311. [Google Scholar] [CrossRef] [PubMed]
  139. Kumar, S.; Xu, X.; Gokhale, R.; Burgess, D.J. Formulation parameters of crystalline nanosuspensions on spray drying processing: A DoE approach. Int. J. Pharm. 2014, 464, 34–45. [Google Scholar] [CrossRef] [PubMed]
  140. Amsa, P.; Tamizharasi, S.; Jagadeeswaran, M.; Sivakumar, T. Preparation and solid state characterization of simvastatin nanosuspensions for enhanced solubility and dissolution. Int. J. Pharm. Pharm. Sci. 2014, 6, 265–269. [Google Scholar]
  141. Dong, Y.; Ng, W.K.; Hu, J.; Shen, S.; Tan, R.B. Continuous production of redispersible and rapidly-dissolved fenofibrate nanoformulation by combination of microfluidics and spray drying. Powder Technol. 2014, 268, 424–428. [Google Scholar] [CrossRef]
  142. Patel, G.V.; Patel, V.B.; Pathak, A.; Rajput, S.J. Nanosuspension of efavirenz for improved oral bioavailability: Formulation optimization, in vitro, in situ and in vivo evaluation. Drug Dev. Ind. Pharm. 2014, 40, 80–91. [Google Scholar] [CrossRef] [PubMed]
  143. Obeidat, W.M.; Sallam, A.-S.A. Evaluation of tadalafil nanosuspensions and their PEG solid dispersion matrices for enhancing its dissolution properties. AAPS PharmSciTech 2014, 15, 364–374. [Google Scholar] [CrossRef] [PubMed]
  144. Papdiwal, A.P.; Pande, V.V.; Aher, S. Investigation of effect of different stabilizers on formulation of zaltoprofen nanosuspension. Int. J. Pharm. Sci. Rev. Res. 2014, 27, 244–249. [Google Scholar]
  145. Homayouni, A.; Sadeghi, F.; Varshosaz, J.; Garekani, H.A.; Nokhodchi, A. Promising dissolution enhancement effect of soluplus on crystallized celecoxib obtained through antisolvent precipitation and high pressure homogenization techniques. Colloids Surf. B Biointerfaces 2014, 122, 591–600. [Google Scholar] [CrossRef] [PubMed]
  146. Elsayed, I.; Abdelbary, A.A.; Elshafeey, A.H. Nanosizing of a poorly soluble drug: Technique optimization, factorial analysis, and pharmacokinetic study in healthy human volunteers. Int. J. Nanomed. 2014, 9, 2943–2953. [Google Scholar]
  147. Yao, Q.; Tao, X.; Tian, B.; Tang, Y.; Shao, Y.; Kou, L.; Gou, J.; Li, X.; Yin, T.; Tang, X. Improved oral bioavailability of core–shell structured beads by redispersion of the shell-forming nanoparticles: Preparation, characterization and in vivo studies. Colloids Surf. B Biointerfaces 2014, 113, 92–100. [Google Scholar] [CrossRef] [PubMed]
  148. Patel, J.; Dhingani, A.; Garala, K.; Raval, M.; Sheth, N. Design and development of solid nanoparticulate dosage forms of telmisartan for bioavailability enhancement by integration of experimental design and principal component analysis. Powder Technol. 2014, 258, 331–343. [Google Scholar] [CrossRef]
  149. Homayouni, A.; Sadeghi, F.; Varshosaz, J.; Garekani, H.A.; Nokhodchi, A. Comparing various techniques to produce micro/nanoparticles for enhancing the dissolution of celecoxib containing PVP. Eur. J. Pharm. Biopharm. 2014, 88, 261–274. [Google Scholar] [CrossRef] [PubMed]
  150. Bhalekar, M.R.; Upadhaya, P.G.; Reddy, S.; Kshirsagar, S.J.; Madgulkar, A.R. Formulation and evaluation of acyclovir nanosuspension for enhancement of oral bioavailability. Asian J. Pharm. 2014, 8, 110–118. [Google Scholar] [CrossRef]
  151. Kumar, S.; Jog, R.; Shen, J.; Zolnik, B.; Sadrieh, N.; Burgess, D.J. In vitro and in vivo performance of different sized spray-dried crystalline itraconazole. J. Pharm. Sci. 2015, 104, 3018–3028. [Google Scholar] [CrossRef] [PubMed]
  152. Knieke, C.; Azad, M.A.; To, D.; Bilgili, E.; Davé, R.N. Sub-100 micron fast dissolving nanocomposite drug powders. Powder Technol. 2015, 271, 49–60. [Google Scholar] [CrossRef]
  153. Krull, S.M.; Susarla, R.; Afolabi, A.; Li, M.; Ying, Y.; Iqbal, Z.; Bilgili, E.; Davé, R.N. Polymer strip films as a robust, surfactant-free platform for delivery of BCS Class II drug nanoparticles. Int. J. Pharm. 2015, 489, 45–57. [Google Scholar] [CrossRef] [PubMed]
  154. Kumar, S.; Shen, J.; Zolnik, B.; Sadrieh, N.; Burgess, D.J. Optimization and dissolution performance of spray-dried naproxen nano-crystals. Int. J. Pharm. 2015, 486, 159–166. [Google Scholar] [CrossRef] [PubMed]
  155. Yue, P.; Wang, C.; Dan, J.; Liu, W.; Wu, Z.; Yang, M. The importance of solidification stress on the redispersibility of solid nanocrystals loaded with harmine. Int. J. Pharm. 2015, 480, 107–115. [Google Scholar] [CrossRef] [PubMed]
  156. Ge, Z.-Q.; Du, X.-Y.; Huang, X.-N.; Qiao, B. Enhanced oral bioavailability of ursolic acid nanoparticles via antisolvent precipitation with TPGS1000 as a stabilizer. J. Drug Deliv. Sci. Technol. 2015, 29, 210–217. [Google Scholar] [CrossRef]
  157. Malamatari, M.; Somavarapu, S.; Bloxham, M.; Buckton, G. Nanoparticle agglomerates of indomethacin: The role of poloxamers and matrix former on their dissolution and aerosolisation efficiency. Int. J. Pharm. 2015, 495, 516–526. [Google Scholar] [CrossRef] [PubMed]
  158. Mishra, B.; Sahoo, J.; Dixit, P.K. Formulation and process optimization of naproxen nanosuspensions stabilized by hydroxypropyl methylcellulose. Carbohydr. Polym. 2015, 127, 300–308. [Google Scholar] [CrossRef] [PubMed]
  159. Lai, F.; Franceschini, I.; Corrias, F.; Sala, M.C.; Cilurzo, F.; Sinico, C.; Pini, E. Maltodextrin fast dissolving films for quercetin nanocrystal delivery. A feasibility study. Carbohydr. Polym. 2015, 121, 217–223. [Google Scholar] [CrossRef] [PubMed]
  160. He, S.; Yang, H.; Zhang, R.; Li, Y.; Duan, L. Preparation and in vitro–in vivo evaluation of teniposide nanosuspensions. Int. J. Pharm. 2015, 478, 131–137. [Google Scholar] [CrossRef] [PubMed]
  161. Pan, Z.; Cui, B.; Zeng, Z.; Feng, L.; Liu, G.; Cui, H.; Pan, H. Lambda-cyhalothrin nanosuspension prepared by the melt emulsification-high pressure homogenization method. J. Nanomater. 2015, 16, 263–270. [Google Scholar] [CrossRef]
  162. Krull, S.M.; Ma, Z.; Li, M.; Davé, R.N.; Bilgili, E. Preparation and characterization of fast dissolving pullulan films containing BCS class II drug nanoparticles for bioavailability enhancement. Drug Dev. Ind. Pharm. 2016, 42, 1073–1085. [Google Scholar] [CrossRef] [PubMed]
  163. Bonda, A.F.; Rinaldi, M.; Segale, L.; Palugan, L.; Cerea, M.; Vecchio, C.; Pattarino, F. Nanonized itraconazole powders for extemporary oral suspensions: Role of formulation components studied by a mixture design. Eur. J. Pharm. Sci. 2016, 83, 175–183. [Google Scholar] [CrossRef] [PubMed]
  164. Paredes, A.J.; Llabot, J.M.; Sánchez Bruni, S.; Allemandi, D.; Palma, S.D. Self-dispersible nanocrystals of albendazole produced by high-pressure homogenization and spray-drying. Drug Dev. Ind. Pharm. 2016, 42, 1564–1570. [Google Scholar] [CrossRef] [PubMed]
  165. Ma, Y.-Q.; Zhang, Z.-Z.; Li, G.; Zhang, J.; Xiao, H.-Y.; Li, X.-F. Solidification drug nanosuspensions into nanocrystals by freeze-drying: A case study with ursodeoxycholic acid. Pharm. Dev. Technol. 2016, 21, 180–188. [Google Scholar] [CrossRef] [PubMed]
  166. Yuan, H.; Li, X.; Zhang, C.; Pan, W.; Liang, Y.; Chen, Y.; Chen, W.; Liu, L.; Wang, X. Nanosuspensions as delivery system for gambogenic acid: Characterization and in vitro/in vivo evaluation. Drug Deliv. 2016, 23, 2772–2779. [Google Scholar] [CrossRef] [PubMed]
  167. Dan, J.; Ma, Y.; Yue, P.; Xie, Y.; Zheng, Q.; Hu, P.; Zhu, W.; Yang, M. Microcrystalline cellulose-carboxymethyl cellulose sodium as an effective dispersant for drug nanocrystals: A case study. Carbohydr. Polym. 2016, 136, 499–506. [Google Scholar] [CrossRef] [PubMed]
  168. Han, M.; Ma, L.; Yu, X.; Li, Z.; Guo, Y.; Wang, X. A nanoparticulate drug-delivery system for 20 (S)-protopanaxadiol: Formulation, characterization, increased oral bioavailability and anti-tumor efficacy. Drug Deliv. 2016, 23, 2410–2418. [Google Scholar] [CrossRef] [PubMed]
  169. Shah, D.A.; Patel, M.; Murdande, S.B.; Dave, R.H. Influence of spray drying and dispersing agent on surface and dissolution properties of griseofulvin micro and nanocrystals. Drug Dev. Ind. Pharm. 2016, 42, 1842–1850. [Google Scholar] [CrossRef] [PubMed]
  170. Alaei, S.; Ghasemian, E.; Vatanara, A. Spray drying of cefixime nanosuspension to form stabilized and fast dissolving powder. Powder Technol. 2016, 288, 241–248. [Google Scholar] [CrossRef]
  171. Steiner, D.; Finke, J.H.; Kwade, A. Efficient production of nanoparticle-loaded orodispersible films by process integration in a stirred media mill. Int. J. Pharm. 2016, 511, 804–813. [Google Scholar] [CrossRef] [PubMed]
  172. Taneja, S.; Shilpi, S.; Khatri, K. Formulation and optimization of efavirenz nanosuspensions using the precipitation-ultrasonication technique for solubility enhancement. Artif. Cells Nanomed. Biotechnol. 2016, 44, 978–984. [Google Scholar] [CrossRef] [PubMed]
  173. Shen, G.; Cheng, L.; Wang, L.Q.; Zhang, L.H.; Shen, B.D.; Liao, W.B.; Li, J.J.; Zheng, J.; Xu, R.; Yuan, H.L. Formulation of dried lignans nanosuspension with high redispersibility to enhance stability, dissolution, and oral bioavailability. Chin. J. Nat. Med. 2016, 14, 757–768. [Google Scholar] [CrossRef]
  174. Xie, Y.; Ma, Y.; Xu, J.; Dan, J.; Yue, P.; Wu, Z.; Yang, M.; Zheng, Q. Roles of cryo/thermal strength for redispersibility of drug nanocrystals: A representative study with andrographolide. Arch. Pharm. Res. 2016, 39, 1404–1417. [Google Scholar] [CrossRef] [PubMed]
  175. Mishra, B.; Sahoo, J.; Dixit, P.K. Enhanced bioavailability of cinnarizine nanosuspensions by particle size engineering: Optimization and physicochemical investigations. Mater. Sci. Eng. C 2016, 63, 62–69. [Google Scholar] [CrossRef] [PubMed]
  176. Karakucuk, A.; Celebi, N.; Teksin, Z.S. Preparation of ritonavir nanosuspensions by microfluidization using polymeric stabilizers: I. A design of experiment approach. Eur. J. Pharm. Sci. 2016, 95, 111–121. [Google Scholar] [CrossRef] [PubMed]
  177. Frank, K.J.; Boeck, G. Development of a nanosuspension for iv administration: From miniscale screening to a freeze dried formulation. Eur. J. Pharm. Sci. 2016, 87, 112–117. [Google Scholar] [CrossRef] [PubMed]
  178. Chonkar, A.D.; Rao, J.V.; Managuli, R.S.; Mutalik, S.; Dengale, S.; Jain, P.; Udupa, N. Development of fast dissolving oral films containing lercanidipine HCl nanoparticles in semicrystalline polymeric matrix for enhanced dissolution and ex vivo permeation. Eur. J. Pharm. Biopharm. 2016, 103, 179–191. [Google Scholar] [CrossRef] [PubMed]
  179. Mu, S.; Li, M.; Guo, M.; Yang, W.; Wang, Y.; Li, J.; Fu, Q.; He, Z. Spironolactone nanocrystals for oral administration: Different pharmacokinetic performances induced by stabilizers. Colloids Surf. B Biointerfaces 2016, 147, 73–80. [Google Scholar] [CrossRef] [PubMed]
  180. Toziopoulou, F.; Malamatari, M.; Nikolakakis, I.; Kachrimanis, K. Production of aprepitant nanocrystals by wet media milling and subsequent solidification. Int. J. Pharm. 2017, 533, 324–334. [Google Scholar] [CrossRef] [PubMed]
  181. Zhang, H.; Chen, W.; Zhao, Z.; Dong, Q.; Yin, L.; Zhou, J.; Ding, Y. Lyophilized nanosuspensions for oral bioavailability improvement of insoluble drugs: Preparation, characterization, and pharmacokinetic studies. J. Pharm. Innov. 2017, 12, 271–280. [Google Scholar] [CrossRef]
  182. Colombo, M.; Orthmann, S.; Bellini, M.; Staufenbiel, S.; Bodmeier, R. Influence of drug brittleness, nanomilling time, and freeze-drying on the crystallinity of poorly water-soluble drugs and its implications for solubility enhancement. AAPS PharmSciTech 2017, 18, 2437–2445. [Google Scholar] [CrossRef] [PubMed]
  183. Nguyen, D.N.; Clasen, C.; Van den Mooter, G. Encapsulating darunavir nanocrystals within Eudragit L100 using coaxial electrospraying. Eur. J. Pharm. Biopharm. 2017, 113, 50–59. [Google Scholar] [CrossRef] [PubMed]
  184. Sartori, G.J.; Prado, L.D.; Rocha, H.V.A. Efavirenz dissolution enhancement IV—Antisolvent nanocrystallization by sonication, physical stability, and dissolution. AAPS PharmSciTech 2017, 18, 3011–3020. [Google Scholar] [CrossRef] [PubMed]
  185. Hou, Y.; Shao, J.; Fu, Q.; Li, J.; Sun, J.; He, Z. Spray-dried nanocrystals for a highly hydrophobic drug: Increased drug loading, enhanced redispersity, and improved oral bioavailability. Int. J. Pharm. 2017, 516, 372–379. [Google Scholar] [CrossRef] [PubMed]
  186. Chen, Y.; Liu, Y.; Xu, J.; Xie, Y.; Zheng, Q.; Yue, P.; Yang, M. A natural triterpenoid saponin as multifunctional stabilizer for drug nanosuspension powder. AAPS PharmSciTech 2017, 18, 2744–2753. [Google Scholar] [CrossRef] [PubMed]
  187. Steiner, D.; Finke, J.H.; Kwade, A. Redispersion of nanoparticle-loaded orodispersible films: Preservation of particle fineness. Chem. Ing. Tech. 2017, 89, 1034–1040. [Google Scholar] [CrossRef]
  188. Xu, J.; Ma, Y.; Xie, Y.; Chen, Y.; Liu, Y.; Yue, P.; Yang, M. Design and evaluation of novel solid self-nanodispersion delivery system for andrographolide. AAPS PharmSciTech 2017, 18, 1572–1584. [Google Scholar] [CrossRef] [PubMed]
  189. Yang, L.; Hong, J.; Di, J.; Guo, Y.; Han, M.; Liu, M.; Wang, X. 10-Hydroxycamptothecin (HCPT) nanosuspensions stabilized by mPEG1000-HCPT conjugate: High stabilizing efficiency and improved antitumor efficacy. Int. J. Nanomed. 2017, 12, 3681–3695. [Google Scholar] [CrossRef] [PubMed]
  190. Iurian, S.; Bogdan, C.; Tomuță, I.; Szabó-Révész, P.; Chvatal, A.; Leucuța, S.E.; Moldovan, M.; Ambrus, R. Development of oral lyophilisates containing meloxicam nanocrystals using QbD approach. Eur. J. Pharm. Sci. 2017, 104, 356–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Liu, C.; Chang, D.; Zhang, X.; Sui, H.; Kong, Y.; Zhu, R.; Wang, W. Oral fast-dissolving films containing lutein nanocrystals for improved bioavailability: Formulation development, in vitro and in vivo evaluation. AAPS PharmSciTech 2017, 18, 2957–2964. [Google Scholar] [CrossRef] [PubMed]
  192. Kumar, R.S. Development and characterization of oral disintegrating tablet containing nanosuspension of lurasidone hydrochloride antipsychotic drug. Asian J. Pharm. 2017, 11, 102–111. [Google Scholar]
  193. Konnerth, C.; Braig, V.; Ito, A.; Schmidt, J.; Lee, G.; Peukert, W. Formation of mefenamic acid nanocrystals with improved dissolution characteristics. Chem. Ing. Tech. 2017, 89, 1060–1071. [Google Scholar] [CrossRef]
  194. Geng, T.; Banerjee, P.; Lu, Z.; Zoghbi, A.; Li, T.; Wang, B. Comparative study on stabilizing ability of food protein, non-ionic surfactant and anionic surfactant on BCS type II drug carvedilol loaded nanosuspension: Physicochemical and pharmacokinetic investigation. Eur. J. Pharm. Sci. 2017, 109, 200–208. [Google Scholar] [CrossRef] [PubMed]
  195. Wang, W. Lyophilization and development of solid protein pharmaceuticals. Int. J. Pharm. 2000, 203, 1–60. [Google Scholar] [CrossRef]
  196. Tang, X.C.; Pikal, M.J. Design of freeze-drying processes for pharmaceuticals: Practical advice. Pharm. Res. 2004, 21, 191–200. [Google Scholar] [CrossRef] [PubMed]
  197. Abdelwahed, W.; Degobert, G.; Fessi, H. Investigation of nanocapsules stabilization by amorphous excipients during freeze-drying and storage. Eur. J. Pharm. Biopharm. 2006, 63, 87–94. [Google Scholar] [CrossRef] [PubMed]
  198. Masters, K. Spray Drying Handbook; Halsted Press: New York, NY, USA, 1985. [Google Scholar]
  199. Davis, M. Recent strategies in spray drying for the enhanced bioavailability of poorly water-soluble drugs. J. Control. Release 2018, 269, 110–127. [Google Scholar] [CrossRef] [PubMed]
  200. Patel, B.B.; Patel, J.K.; Chakraborty, S.; Shukla, D. Revealing facts behind spray dried solid dispersion technology used for solubility enhancement. Saudi Pharm. J. 2015, 23, 352–365. [Google Scholar] [CrossRef] [PubMed]
  201. Azad, M. Dissolution Enhanced Nanocomposites from Stabilized Suspensions of Poorly Water-Soluble Drugs. Ph.D. Thesis, New Jersey Institute of Technology, Newark, NJ, USA, 2013. [Google Scholar]
  202. Dixit, R.; Puthli, S. Oral strip technology: Overview and future potential. J. Control. Release 2009, 139, 94–107. [Google Scholar] [CrossRef] [PubMed]
  203. Krull, S.M. Formulation and Dissolution of Polymer Strip Films for the Delivery of Poorly Water-Soluble Drug Nanoparticles. Ph.D. Thesis, New Jersey Institute of Technology, Newark, NJ, USA, 2017. [Google Scholar]
  204. Maury, M.; Murphy, K.; Kumar, S.; Shi, L.; Lee, G. Effects of process variables on the powder yield of spray-dried trehalose on a laboratory spray-dryer. Eur. J. Pharm. Biopharm. 2005, 59, 565–573. [Google Scholar] [CrossRef] [PubMed]
  205. Bilgili, E.; Rosen, L.A.; Ko, J.S.; Chen, A.; Smith, E.J.; Fliszar, K.; Wong, G. Experimental study of fluidized bed co-granulation of two active pharmaceutical ingredients: An industrial scale-up perspective. Part. Sci. Technol. 2011, 29, 285–309. [Google Scholar] [CrossRef]
  206. El-Meliegy, E.; Mabrouk, M.; Kamal, G.M.; Awad, S.M.; El-Tohamy, A.M.; El Gohary, M.I. Anticancer drug carriers using dicalcium phosphate/dextran/CMCnanocomposite scaffolds. J. Drug Deliv. Sci. Technol. 2018, 45, 315–322. [Google Scholar] [CrossRef]
  207. Grigorov, P.I.; Glasser, B.J.; Muzzio, F.J. Cross-sectional analysis of impregnated excipient particles by energy dispersive X-ray spectroscopy. Powder Technol. 2018, 332, 197–209. [Google Scholar] [CrossRef]
  208. Parmentier, J.; Tan, E.H.; Low, A.; Möschwitzer, J.P. Downstream drug product processing of itraconazole nanosuspension: Factors influencing drug particle size and dissolution from nanosuspension-layered beads. Int. J. Pharm. 2017, 524, 443–453. [Google Scholar] [CrossRef] [PubMed]
  209. Nekkanti, V.; Pillai, R.; Venkateshwarlu, V.; Harisudhan, T. Development and characterization of solid oral dosage form incorporating candesartan nanoparticles. Pharm. Dev. Technol. 2009, 14, 290–298. [Google Scholar] [CrossRef] [PubMed]
  210. Feng, T.; Pinal, R.; Carvajal, M.T. Process induced disorder in crystalline materials: Differentiating defective crystals from the amorphous form of griseofulvin. J. Pharm. Sci. 2008, 97, 3207–3221. [Google Scholar] [CrossRef] [PubMed]
  211. Deng, Z.; Xu, S.; Li, S. Understanding a relaxation behavior in a nanoparticle suspension for drug delivery. Int. J. Pharm. 2008, 351, 236–243. [Google Scholar] [CrossRef] [PubMed]
  212. Azad, M.; Afolabi, A.; Bhakay, A.; Leonardi, J.; Davé, R.; Bilgili, E. Enhanced physical stabilization of fenofibrate nanosuspensions via wet co-milling with a superdisintegrant and an adsorbing polymer. Eur. J. Pharm. Biopharm. 2015, 94, 372–385. [Google Scholar] [CrossRef] [PubMed]
  213. Van Eerdenbrugh, B.; Sofie, V.; Martens, J.A.; Jan, V.; Ludo, F.; Jan, V.H.; Van Den Mooter, G.; Patrick, A. Microcrystalline cellulose, a useful alternative for sucrose as a matrix former during freeze-drying of drug nanosuspensions—A case study with itraconazole. Eur. J. Pharm. Biopharm. 2008, 70, 590–596. [Google Scholar] [PubMed]
  214. Pu, X.; Sun, J.; Wang, Y.; Wang, Y.; Liu, X.; Zhang, P.; Tang, X.; Pan, W.; Han, J.; He, Z. Development of a chemically stable 10-hydroxycamptothecin nanosuspensions. Int. J. Pharm. 2009, 379, 167–173. [Google Scholar] [CrossRef] [PubMed]
  215. Han, M.; Liu, X.; Guo, Y.; Wang, Y.; Wang, X. Preparation, characterization, biodistribution and antitumor efficacy of hydroxycamptothecin nanosuspensions. Int. J. Pharm. 2013, 455, 85–92. [Google Scholar] [CrossRef] [PubMed]
  216. Schwarz, C.; Mehnert, W. Freeze-drying of drug-free and drug-loaded solid lipid nanoparticles (SLN). Int. J. Pharm. 1997, 157, 171–179. [Google Scholar] [CrossRef]
  217. Konan, Y.N.; Gurny, R.; Allémann, E. Preparation and characterization of sterile and freeze-dried sub-200 nm nanoparticles. Int. J. Pharm. 2002, 233, 239–252. [Google Scholar] [CrossRef]
  218. Stamm, A.; Seth, P. Fenofibrate Pharmaceutical Composition Having High Bioavailability and Method for Preparing It. U.S. Patent 6,074,670, 13 June 2000. [Google Scholar]
  219. Biradar, S.V.; Patil, A.R.; Sudarsan, G.V.; Pokharkar, V.B. A comparative study of approaches used to improve solubility of roxithromycin. Powder Technol. 2006, 169, 22–32. [Google Scholar] [CrossRef]
  220. Li, C.; Li, C.; Le, Y.; Chen, J.-F. Formation of bicalutamide nanodispersion for dissolution rate enhancement. Int. J. Pharm. 2011, 404, 257–263. [Google Scholar] [CrossRef] [PubMed]
  221. Van Eerdenbrugh, B.; Froyen, L.; Martens, J.; Blaton, N.; Augustijns, P.; Brewster, M.; Van den Mooter, G. Characterization of physico-chemical properties and pharmaceutical performance of sucrose co-freeze–dried solid nanoparticulate powders of the anti-HIV agent loviride prepared by media milling. Int. J. Pharm. 2007, 338, 198–206. [Google Scholar] [CrossRef] [PubMed]
  222. Chaubal, M.V.; Popescu, C. Conversion of nanosuspensions into dry powders by spray drying: A case study. Pharm. Res. 2008, 25, 2302–2308. [Google Scholar] [CrossRef] [PubMed]
  223. Layre, A.-M.; Couvreur, P.; Richard, J.; Requier, D.; Eddine Ghermani, N.; Gref, R. Freeze-drying of composite core-shell nanoparticles. Drug Dev. Ind. Pharm. 2006, 32, 839–846. [Google Scholar] [CrossRef] [PubMed]
  224. Liu, P.; De Wulf, O.; Laru, J.; Heikkilä, T.; van Veen, B.; Kiesvaara, J.; Hirvonen, J.; Peltonen, L.; Laaksonen, T. Dissolution studies of poorly soluble drug nanosuspensions in non-sink conditions. AAPS PharmSciTech 2013, 14, 748–756. [Google Scholar] [CrossRef] [PubMed]
  225. Deng, J.; Staufenbiel, S.; Hao, S.; Wang, B.; Dashevskiy, A.; Bodmeier, R. Development of a discriminative biphasic in vitro dissolution test and correlation with in vivo pharmacokinetic studies for differently formulated racecadotril granules. J. Control. Release 2017, 255, 202–209. [Google Scholar] [CrossRef] [PubMed]
  226. Talukder, R.; Reed, C.; Dürig, T.; Hussain, M. Dissolution and solid-state characterization of poorly water-soluble drugs in the presence of a hydrophilic carrier. AAPS PharmSciTech 2011, 12, 1227–1233. [Google Scholar] [CrossRef] [PubMed]
  227. Washburn, E.W. The dynamics of capillary flow. Phys. Rev. 1921, 17, 273–283. [Google Scholar] [CrossRef]
  228. Hołownia, D.; Kwiatkowska, I.; Hupka, J. An investigation on wetting of porous materials. Physicochem. Prob. Miner. Process. 2008, 42, 251–262. [Google Scholar]
  229. Yalkowsky, S.H.; Roseman, T.J. Techniques of Solubilization of Drugs; M. Dekker: New York, NY, USA, 1981. [Google Scholar]
  230. Liversidge, G.G.; Cundy, K.C. Particle size reduction for improvement of oral bioavailability of hydrophobic drugs: I. Absolute oral bioavailability of nanocrystalline danazol in beagle dogs. Int. J. Pharm. 1995, 125, 91–97. [Google Scholar] [CrossRef]
  231. Gupta, S.; Kesarla, R.; Omri, A. Formulation strategies to improve the bioavailability of poorly absorbed drugs with special emphasis on self-emulsifying systems. ISRN Pharm. 2013, 2013, 848043. [Google Scholar] [CrossRef] [PubMed]
  232. Suzuki, M.; Machida, M.; Adachi, K.; Otabe, K.; Sugimoto, T.; Hayashi, M.; Awazu, S. Histopathological study of the effects of a single intratracheal instillation of surface active agents on lung in rats. J. Toxicol. Sci. 2000, 25, 49–55. [Google Scholar] [CrossRef] [PubMed]
  233. Wong, J.; Brugger, A.; Khare, A.; Chaubal, M.; Papadopoulos, P.; Rabinow, B.; Kipp, J.; Ning, J. Suspensions for intravenous (IV) injection: A review of development, preclinical and clinical aspects. Adv. Drug Deliv. Rev. 2008, 60, 939–954. [Google Scholar] [CrossRef] [PubMed]
  234. Lebhardt, T.; Roesler, S.; Uusitalo, H.P.; Kissel, T. Surfactant-free redispersible nanoparticles in fast-dissolving composite microcarriers for dry-powder inhalation. Eur. J. Pharm. Biopharm. 2011, 78, 90–96. [Google Scholar] [CrossRef] [PubMed]
  235. Rowe, R.C.; Sheskey, P.J.; Quinn, M.E. Handbook of Pharmaceutical Excipients, 6th ed.; Royal Pharmaceutical Society: London, UK, 2009. [Google Scholar]
  236. U.S. Food and Drug Administration (FDA). Database for Inactive Ingredient Search for Approv-ed Drug Products. 2018. Available online: https://www.accessdata.fda.gov/scripts/cder/iig/index.cfm?event=BasicSearch.page (accessed on 28 June 2018).
  237. U.S. Food and Drug Administration (FDA). Guidance for Industry: Nonclinical Studies for the Safety Evaluation of Pharmaceutical Excipients; Center for Drug Evaluation and Research (CDER) and the Center for Biologics Evaluation and Research (CBER): Rockville, MD, USA, 2005. Available online: https://www.fda.gov/downloads/Drugs/GuidanceComplianceRegulatoryInformation/Guidances/UCM079250.pdf (accessed on 28 June 2018).
  238. Leung, D.H.; Lamberto, D.J.; Liu, L.; Kwong, E.; Nelson, T.; Rhodes, T.; Bak, A. A new and improved method for the preparation of drug nanosuspension formulations using acoustic mixing technology. Int. J. Pharm. 2014, 473, 10–19. [Google Scholar] [CrossRef] [PubMed]
  239. Li, M.; Zhang, L.; Davé, R.N.; Bilgili, E. An intensified vibratory milling process for enhancing the breakage kinetics during the preparation of drug nanosuspensions. AAPS PharmSciTech 2016, 17, 389–399. [Google Scholar] [CrossRef] [PubMed]
  240. Abdelwahed, W.; Degobert, G.; Stainmesse, S.; Fessi, H. Freeze-drying of nanoparticles: Formulation, process and storage considerations. Adv. Drug Deliv. Rev. 2006, 58, 1688–1713. [Google Scholar] [CrossRef] [PubMed]
  241. Bilgili, E.; Dave, R.; Bhakay, A.; Azad, M. Systems and Methods for Superdisintegrant-Based Composite Particles for Dispersion and Dissolution of Agents. U.S. Patent 9,452,107B2, 27 September 2016. [Google Scholar]
  242. Li, M.; Dave, R.; Bilgili, E. High drug-loaded surfactant-free nanocomposite microparticles for enhanced dissolution of poorly soluble drugs. In Proceedings of the 8th World Congress on Particle Technology, Orlando, FL, USA, 22–26 April 2018. [Google Scholar]
  243. Dolenc, A.; Kristl, J.; Baumgartner, S.; Planinšek, O. Advantages of celecoxib nanosuspension formulation and transformation into tablets. Int. J. Pharm. 2009, 376, 204–212. [Google Scholar] [CrossRef] [PubMed]
  244. Mauludin, R.; Müller, R.H.; Keck, C.M. Development of an oral rutin nanocrystal formulation. Int. J. Pharm. 2009, 370, 202–209. [Google Scholar] [CrossRef] [PubMed]
  245. Cal, K.; Sollohub, K. Spray drying technique. I: Hardware and process parameters. J. Pharm. Sci. 2010, 99, 575–586. [Google Scholar] [CrossRef] [PubMed]
  246. Walters, R.H.; Bhatnagar, B.; Tchessalov, S.; Izutsu, K.-I.; Tsumoto, K.; Ohtake, S. Next generation drying technologies for pharmaceutical applications. J. Pharm. Sci. 2014, 103, 2673–2695. [Google Scholar] [CrossRef] [PubMed]
  247. Mujumdar, A.; Alterman, D. Drying in the pharmaceutical and biotechnology fields. In Handbook of Downstream Processing; Springer: Dordrecht, The Netherlands, 1997; pp. 235–260. [Google Scholar]
  248. Möschwitzer, J.P.; Müller, R.H. Factors influencing the release kinetics of drug nanocrystal-loaded pellet formulations. Drug Dev. Ind. Pharm. 2013, 39, 762–769. [Google Scholar] [CrossRef] [PubMed]
  249. Schwartzbach, H. The Possibilities and Challenges of Spray Drying; Pharmaceutical Technology Europe: London, UK, 2010. [Google Scholar]
  250. Teunou, E.; Poncelet, D. Batch and continuous fluid bed coating–review and state of the art. J. Food Eng. 2002, 53, 325–340. [Google Scholar] [CrossRef]
  251. Kesisoglou, F.; Mitra, A. Crystalline nanosuspensions as potential toxicology and clinical oral formulations for BCS II/IV compounds. AAPS J. 2012, 14, 677–687. [Google Scholar] [CrossRef] [PubMed]
  252. Shah, S.M.H.; Ullah, F.; Khan, S.; Shah, S.M.M.; de Matas, M.; Hussain, Z.; Minhas, M.U.; AbdEl-Salam, N.M.; Assi, K.H.; Isreb, M. Smart nanocrystals of artemether: Fabrication, characterization, and comparative in vitro and in vivo antimalarial evaluation. Drug Des. Dev. Ther. 2016, 10, 3837–3850. [Google Scholar] [CrossRef] [PubMed]
  253. Rahim, H.; Sadiq, A.; Khan, S.; Khan, M.A.; Shah, S.M.H.; Hussain, Z.; Ullah, R.; Shahat, A.A.; Ibrahim, K. Aceclofenac nanocrystals with enhanced in vitro, in vivo performance: Formulation optimization, characterization, analgesic and acute toxicity studies. Drug Des. Dev. Ther. 2017, 11, 2443–2452. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustrating the steps involved in the preparation of drug-laden nanocomposites including their characterization. Nanocomposites in the form of powders, cakes, or extrudates are intermediate products that are incorporated into final solid oral dosage forms such as tablets, capsules, and sachets via standard pharmaceutical unit operations upon use of additional excipients. Polymeric strip films prepared by wet film casting–drying are the final product.
Figure 1. Schematic illustrating the steps involved in the preparation of drug-laden nanocomposites including their characterization. Nanocomposites in the form of powders, cakes, or extrudates are intermediate products that are incorporated into final solid oral dosage forms such as tablets, capsules, and sachets via standard pharmaceutical unit operations upon use of additional excipients. Polymeric strip films prepared by wet film casting–drying are the final product.
Pharmaceutics 10 00086 g001
Figure 2. Classification of various types of aggregates that may be present in nanocomposite particles based on their redispersion behavior.
Figure 2. Classification of various types of aggregates that may be present in nanocomposite particles based on their redispersion behavior.
Pharmaceutics 10 00086 g002
Figure 3. The number of published journal articles from 2010–2017 which reported the preparation of drug nanoparticles and drug nanocomposites. Source: Scopus database, key words used: “drug nanoparticles” or “drug nanocomposites” or “drug + drying + nanocrystals” or “drug + drying + nanosuspensions”.
Figure 3. The number of published journal articles from 2010–2017 which reported the preparation of drug nanoparticles and drug nanocomposites. Source: Scopus database, key words used: “drug nanoparticles” or “drug nanocomposites” or “drug + drying + nanocrystals” or “drug + drying + nanosuspensions”.
Pharmaceutics 10 00086 g003
Figure 4. The usage frequency of various drug nanosuspension preparation methods in the studies reported in Table 2. The sample size for the analysis here is 94, even though the number of publications in Table 2 is 92, because two studies compared two different preparation methods.
Figure 4. The usage frequency of various drug nanosuspension preparation methods in the studies reported in Table 2. The sample size for the analysis here is 94, even though the number of publications in Table 2 is 92, because two studies compared two different preparation methods.
Pharmaceutics 10 00086 g004
Figure 5. The usage frequency of various drying methods in the studies reported in Table 2. The sample size for the analysis here is 103, even though the number of publications in Table 2 is 92, because some studies compared two drying methods.
Figure 5. The usage frequency of various drying methods in the studies reported in Table 2. The sample size for the analysis here is 103, even though the number of publications in Table 2 is 92, because some studies compared two drying methods.
Pharmaceutics 10 00086 g005
Figure 6. Comparison of drug release from the nanocomposite particles: (a) ITZ coated on PrismaLac 40 vs. GranuLac 200, (b) FNB coated on PrismaLac 40 vs. GranuLac 200, during the USP II dissolution test. Dissolution was performed using 7.2 mg/mL SDS solution for (a) and 2.88 mg/mL SDS solution for (b). (Figure adapted from Azad et al. [91] with permission from Elsevier, www.elsevier.com).
Figure 6. Comparison of drug release from the nanocomposite particles: (a) ITZ coated on PrismaLac 40 vs. GranuLac 200, (b) FNB coated on PrismaLac 40 vs. GranuLac 200, during the USP II dissolution test. Dissolution was performed using 7.2 mg/mL SDS solution for (a) and 2.88 mg/mL SDS solution for (b). (Figure adapted from Azad et al. [91] with permission from Elsevier, www.elsevier.com).
Pharmaceutics 10 00086 g006
Figure 7. Images of redispersion dynamics of griseofulvin (GF) nanocomposite particles, which were prepared by fluidized bed coating of GF nanosuspension on Pharmatose carrier particles with various dispersants, in quiescent water (no external agitation/shear) (left panel), and after addition of a drop of water on a single nanocomposite particle, visualized under optical microscope (right panel). Nanocomposite formulations in the figure from top to bottom contain GF nanoparticles without stabilizers, GF nanoparticles with HPC, GF nanoparticles with HPC–SDS, and GF with HPC–croscarmellose sodium (CCS) milled for 60 min. (Adapted from Bhakay et al. [92] with permission from Springer Nature, www.springernature.com).
Figure 7. Images of redispersion dynamics of griseofulvin (GF) nanocomposite particles, which were prepared by fluidized bed coating of GF nanosuspension on Pharmatose carrier particles with various dispersants, in quiescent water (no external agitation/shear) (left panel), and after addition of a drop of water on a single nanocomposite particle, visualized under optical microscope (right panel). Nanocomposite formulations in the figure from top to bottom contain GF nanoparticles without stabilizers, GF nanoparticles with HPC, GF nanoparticles with HPC–SDS, and GF with HPC–croscarmellose sodium (CCS) milled for 60 min. (Adapted from Bhakay et al. [92] with permission from Springer Nature, www.springernature.com).
Pharmaceutics 10 00086 g007
Figure 8. The usage frequency of (a) various classes of dispersants and (b) classes of dispersants in the formulations that led to formation of redispersible nanocomposites in the studies reported in Table 2. The sample size for the analysis here is 120, even though the number of publications in Table 2 is 92, because some studies investigated more than one drug. WSD: other water-soluble dispersant (besides soluble polymer–surfactant), WID: water-insoluble dispersant, CLP: crosslinked polymers.
Figure 8. The usage frequency of (a) various classes of dispersants and (b) classes of dispersants in the formulations that led to formation of redispersible nanocomposites in the studies reported in Table 2. The sample size for the analysis here is 120, even though the number of publications in Table 2 is 92, because some studies investigated more than one drug. WSD: other water-soluble dispersant (besides soluble polymer–surfactant), WID: water-insoluble dispersant, CLP: crosslinked polymers.
Pharmaceutics 10 00086 g008
Figure 9. Drug (GF) release profiles of the nanocomposites prepared using wet-milled GF suspensions containing 0.05% SDS and HPC with different molecular weights of HPC: (a) HPC SSL (40 kDa); (b) HPC SL (100 kDa); and (c) HPC L (140 kDa) (Adapted from Li et al. [86] with permission from Elsevier, www.elsevier.com).
Figure 9. Drug (GF) release profiles of the nanocomposites prepared using wet-milled GF suspensions containing 0.05% SDS and HPC with different molecular weights of HPC: (a) HPC SSL (40 kDa); (b) HPC SL (100 kDa); and (c) HPC L (140 kDa) (Adapted from Li et al. [86] with permission from Elsevier, www.elsevier.com).
Pharmaceutics 10 00086 g009
Figure 10. Impact of drug loading in the nanocomposite and drying method on redispersion of griseofulvin (GF)–HPC–SDS nanocomposite microparticles (NCMP) in water after 2 min paddle stirring at 200 rpm. FBD refers to fluid bed coating/drying of the GF nanosuspension onto Pharmatose; SD stands for spray drying. “Before redisp” particle sizes refer to the particle sizes in the wet media milled drug nanosuspension with HPC–SDS. (Adapted from Bhakay et al. [85] with permission from Elsevier, www.elsevier.com).
Figure 10. Impact of drug loading in the nanocomposite and drying method on redispersion of griseofulvin (GF)–HPC–SDS nanocomposite microparticles (NCMP) in water after 2 min paddle stirring at 200 rpm. FBD refers to fluid bed coating/drying of the GF nanosuspension onto Pharmatose; SD stands for spray drying. “Before redisp” particle sizes refer to the particle sizes in the wet media milled drug nanosuspension with HPC–SDS. (Adapted from Bhakay et al. [85] with permission from Elsevier, www.elsevier.com).
Pharmaceutics 10 00086 g010
Figure 11. A correlation between percentage of GF dissolved in 2 min in the USP II dissolution test and percentage of GF nanoparticles recovered in 2 min during the agitated redispersion test with a paddle stirrer. Markers correspond to various dispersant formulations in the GF nanocomposites prepared by fluidized bed coating of wet media milled GF suspensions on Pharmatose. The redispersion–dissolution medium is water. (Adapted from Bhakay et al. [8] with permission from Taylor & Francis Ltd., www.tandfonline.com.
Figure 11. A correlation between percentage of GF dissolved in 2 min in the USP II dissolution test and percentage of GF nanoparticles recovered in 2 min during the agitated redispersion test with a paddle stirrer. Markers correspond to various dispersant formulations in the GF nanocomposites prepared by fluidized bed coating of wet media milled GF suspensions on Pharmatose. The redispersion–dissolution medium is water. (Adapted from Bhakay et al. [8] with permission from Taylor & Francis Ltd., www.tandfonline.com.
Pharmaceutics 10 00086 g011
Table 1. Drug nanoparticle-based marketed products approved by FDA (Adapted from Malamatari et al. [31] with permission from Elsevier, www.elsevier.com).
Table 1. Drug nanoparticle-based marketed products approved by FDA (Adapted from Malamatari et al. [31] with permission from Elsevier, www.elsevier.com).
Product Name/CompanyDrugNanoparticle Preparation Method aFinal DosageYear Approved
Avinza®/King PharmaMorphine sulfateWMMCapsule2002
Azopt®/AlconBrinzolmidWMMSuspension1998
Cesamet®/LillyNabilonPrecipitationCapsule2005
Emend®/MerckAprepitantWMMCapsule2003
Focalin XR®/NovartisDexmethylphenidate HClWMMCapsule2001
Gris-Peg®/NovartisGriseofulvinPrecipitationTablet1982
Herbesser®/MitsubishiDiltiazemWMMTablet2002
Invega Sustenna®/Johnson & JohnsonPaliperidone palmitateWMMSuspension2009
Megace ES®/Par PharmaceuticalMegestrol acetateWMMSuspension2005
Neprelan®/WyethNaproxen sodiumWMMTablet2006
Rapamune®/WyethSirolimus (rapamycin)WMMSuspension, Tablet2000
Ritalin LA®/NovartisMethylphenidate HClWMMCapsule2002
Theodur®/Mitsubishi Tanabe PharmaTheophyllineWMMTablet, Capsule2008
Tricor®/AbbottFenofibrateWMMTablet2004
Triglide®/SkyePharmaFenofibrateHPHTablet2005
Verelan PM®/Schwarz PharmaVerapamil HClWMMCapsule1998
Zanaflex®/AcordaTizanidine HClWMMCapsule2002
a HPH: High-pressure homogenization; WMM: Wet media milling.
Table 2. Review of recent literature (2012–2017) regarding the preparation methods used for drug nanosuspension preparation, drying methods used for converting nanosuspensions into nanocomposites, dispersants used in the formulations, and redispersion methods used for nanoparticle recovery.
Table 2. Review of recent literature (2012–2017) regarding the preparation methods used for drug nanosuspension preparation, drying methods used for converting nanosuspensions into nanocomposites, dispersants used in the formulations, and redispersion methods used for nanoparticle recovery.
Method for Drug Nanosuspension PreparationDrying MethodDrug and Its Assay in Nanocomposites (% w/w)Dispersants in Nanocomposites a,bRedispersion Method (If Used)d50c, dvm d, Cumulant e Size Before and After Redispersion (µm)Reference
BeforeAfter
LASP–UltrasonicationFreeze dryingCarvedilol (–)(Alpha tocopherol succinate, SDS, Maltose)_0.212_fLiu et al., (2012) [108]
WMMFreeze dryingNaproxen (–)(HPC, PEG Carrageenan), SucroseDried powders were dispersed in 150 mL water and sonicated for 1 min0.148 d0.150 d,gChung et al., (2012) [72]
WMMFreeze dryingModel drug (–)(Poloxamer 338, PVP K15), Cremophor EL Sucrose, TrehaloseAn aqueous solution of 5 mg/mL poloxamer 338 was used as a medium_0.165 gBeirowski et al., (2012) [71]
WMMWet film casting–dryingGriseofulvin
(3.8 mg/cm2)
Naproxen
(3.3 mg/cm2)
Fenofibrate
(4.8 mg/cm2)
(HPMC E15LV, SDS, Glycerin)0.71 cm2 circular films were put in 15 mL water and stirred for 10 min via magnetic stirrer0.163
0.144
0.207
0.175 g
0.145 g
0.256 f
Sievens et al., (2012) [79]
WMMElectrospray dryingNaproxen (–)(HPC)Dried powders were placed in 150 mL water and sonicated for 4 min~0.110 d0.100 d,gHo and Lee (2012) [120]
WMMFluid bed granulation/dryingCompound A (9.19%)(Vitamin E TPGS, HPMC 3, Mannitol DC), Lactose monohydrate_~0.220_ fBose et al., (2012) [130]
WMMFluid bed coating/dryingGriseofulvin (12.4%)(HPC SL, SDS), Mannitol, Pharmatose (core)1 g dried sample was dispersed in 30 mL water for 2 min using paddle stirring (200 rpm), pipette stirring, magnetic stirring (100 rpm) and sonication0.1450.150 fBhakay et al., (2013) [66]
LASP–Ultrasonication Wet film casting–dryingGriseofulvin (3.95%)(HPMC E15LV, HPMC E4M, Pluronic F127, Glycerin)Dried films were dispersed in water0.580~2.000 fBeck et al., (2013) [109]
WMMSpray dryingMiconazole (45%)
Itraconazole (44%)
(HPC, SDS, Mannitol), MCCDried samples were dispersed in water and shaken manually0.157 d
0.144 d
~0.200 d,f
~0.150 d,f
Cerdeira et al., (2013) [121]
Freeze dryingMiconazole (47%)
Itraconazole (42%)
(HPC, SDS, Mannitol), MCCDried samples were dispersed in water and shaken manually0.182 d
0.192 d
~0.198 d,f
~0.200 d
WMMWet film casting–dryingGriseofulvin (1.87 mg/cm2)(SDS, HPMC E15, Glycerin)0.715 cm2 circular films were put in 15 mL water and stirred for 10 min using magnetic stirrer 0.1630.164 fSusarla et al., (2013) [131]
WMMSpray-freeze dryingPhenytoin (–)(PVP, SLS)Powders equivalent to 2 mg phenytoin were dispersed in 10 mL of dissolution media (pH 1.2 and 6.8) and stirred up to 30 min via magnetic stirrer 0.170–0.180~0.400 fNiwa et al., (2013) [7]
LASPFreeze dryingCurcumin (37.6%)(PEG-PLA, PVP BP, HPBCD)Dried samples were dispersed in DI-water0.055 e0.076 e,gCheng et al., (2013) [132]
LASP–UltrasonicationFluid bed coating/dryingIndomethacin (–)(β-lactoglobulin, PVP K30, Trehalose), Nonpareil (core), Soybean Protein Isolate, Whey protein isolate 100 mg dried product was dispersed in 10 mL DI-water via manual shaking for 1 min0.243 e0.289 e,fHe et al., (2013) [133]
Acid-base neutralizationSpray dryingDiosmin (–)(HPMC, Mannitol), MCDried powders were dispersed in distilled water 0.336 e0.316 e,fFreag et al., (2013) [110]
LASP–HPHFreeze dryingNimodipine (–)(Poloxamer 407, Sodium deoxycholate, HPMC E5, Mannitol, Maltose)Beckmann Coulter LS 2300.159 d0.148 d,fFu et al., (2013) [134]
UltrasonicationFreeze dryingFenofibrate (–)(Poloxamer 188, Mannitol), PVP K25, Poloxamer 407, SDS, Tween 80_0.460 e_ fIge et al., (2013) [135]
WMMSpray dryingFenofibrate (–)(HPMC E5, SDS, Mannitol), Sucrose, Glucose, Maltose, Lactose20 mg dry powder was added in 5 mL of DI water and shaken manually0.4520.499 fZuo et al., (2013) [136]
UltrasonicationNanoextrusionTiO2 (–)(Citric acid monohydrate, SDS, Soluplus), Tween60, Cremophor EL, Cremophor RH 40_~0.182 e_ gKhinast et al., (2013) [75]
HPHWet film casting–dryingHerpetrione (10 mg/4 cm2)(PVP K30, SDS, L-HPC, HPMC E50, MCC, PEG 400, Mannitol)2 × 2 cm2 film was placed into distilled water and manually shaken for 30 s 0.260 e0.280 e,fShen et al., (2013) [106]
WMMFreeze dryingCurcumin didecanoate (–)(Poloxamer 188)2 mg powder was suspended in peanut oil and sonicated for 1 min~0.500 e0.517 e,gWei et al., (2013) [137]
WMMSpray dryingNaproxen (–)
Indomethacin (–)
(HPMC E15),
Dowfax 2A
1(Dowfax 2A1),
HPMC E15
Powders were suspended in saturated and filtered solution of the drug in 30% glycerin solution0.309 e
0.223 e
0.400 e,f
0.351 e,f
Kumar et al., (2014a) [138]
WMMSpray dryingIndomethacin (–)(Dowfax 2A1, Maltose), Trehalose, Lactose, Mannitol, Ficoll PM70, MaltodextrinPowders were suspended in saturated and filtered solution of indomethacin in 30% glycerin solution0.200–0.300 e0.179 e,gKumar et al., (2014b) [139]
Freeze dryingIndomethacin (–)(Dowfax 2A1, Sucrose), Trehalose, Lactose, Mannitol, Ficoll PM70, Maltodextrin 0.197 e0.208 e,g
HPHFreeze dryingSimvastatin (–)(Soya Lecithin, Mannitol)_0.316 e_ fAsma et al., (2014) [140]
WMMNanoextrusionPhenytoin (–)(Tween 80, Soluplus), Tween 20, Kolliphor P188, Kollicoat IR_0.335 e_ fBaumgartner et al., (2014) [76]
LASPSpray dryingFenofibrate (–)(PVP 10, MMT), Lactose_<1.00 e_ fDong et al., (2014) [141]
WMMFreeze dryingEfavirenz (–)(PVP K30, SLS, Trehalose), Mannitol, Poloxamer 188 and 407_0.320 e_ fPatel et al., (2014) [142]
HPHSpray dryingLovastatin (–)(PVP K17, SDS), HPMC 2910, Polyvinyl alcohol, PVP K30 and K12, Poloxamer 188 and 407Dried powders were dispersed in distilled water and shaken manually for 10 s
RDI = mean redispersion size/nanosuspesion size ×100
0.380 e110% fZhang et al., (2014) [26]
LASP–UltrasonicationFreeze dryingTadalafil (–)(Tween 80, Span80), SLS PEG 4000, PVP K30, Pluronic F-127, Methocel E50 and E5, Span 20 and 60_0.193 e_ fObeidat et al., (2014) [143]
LASP–UltrasonicationFreeze dryingZaltoprofen (–)(Poloxamer 188, SLS), Poloxamer 407Dried samples were diluted with 20 mL distilled water to observe redispersibility0.179 e_ fPapdiwal et al., (2014) [144]
Freeze dryingQuercetin dehydrate (–)(Tween 80)_~0.430 e_ f
LASPFreeze dryingCelecoxib (–)(Soluplus)_0.293 e_ fHomayouni et al., (2014) [145]
LASP–HPHFreeze dryingCelecoxib (–)(Soluplus)_0.577 e_ f
LASP–Ultrasonication–HPHFreeze dryingDiacerein (–)(SDS, Mannitol), PVA, Sodium deoxycholate, SucroseMalvern Zetasizer® Nano ZS900.374 e0.374 e,fElsayed et al., (2014) [146]
WMMFluid bed coating/dryingBifendate (–)(HPC SL, SLS, Mannitol), MCC (core), HPMC E5, PVP K30, Poloxamer 407 and 1885 g dried products were dispersed in 100 mL purified water and paddle stirred at 100 rpm for 5 min 0.139 d0.360 d,fYao et al., (2014) [147]
WMMFreeze dryingTelmisartan (–)(Poloxamer 188, Trehalose), Poloxamer 407, PVA, PVP K30, SLS, Tween 80, HPMC E5, HPC, Glucose, Lactose, Mannitol, Sucrose50 mg dried product was dispersed in 5 mL distilled water and sonicated for 15–30 s~0.335 e~0.337 e,fPatel et al., (2014) [148]
LASPSpray dryingCelecoxib (–)(PVP K30)0.321 e_ fHomayouni et al., (2014) [149]
Freeze dryingCelecoxib (–)(PVP K30)_0.321 e_ f
LASP–HPHFreeze dryingCelecoxib (–)(PVP K30)_0.450 e_ f
LASP–UltrasonicationFreeze dryingAcyclovir (–)(Poloxamer 188, PVP K30, Sucrose), TPGS, Tween 80, Mannitol, MCCMalvern Mastersizer0.274 e0.353 e,fBhalekar et al., (2014) [150]
WMMFluid bed coating/dryingGriseofulvin (48.8%)
Azodicarbonamide (40.4%)
(HPC SL, SDS), Pharmatose (core)Dried product equivalent to 5 wt.% drug (w.r.t. suspension) dispersed in 20 mL water stirred for 2 min using impeller at 200 rpm0.160
0.250
~0.160 f
~0.250 f
Bhakay et. al., (2014) [85]
Spray dryingGriseofulvin (76.9%)
Azodicarbonamide (74%)
(HPC SL, SDS), Mannitol 0.160
0.250
~0.160 f
~0.250 f
WMMSpray dryingItraconazole (–)(PVP40, SLS), Methocel E3, Methylcellulose A15, HPMC E5, HPMC E15, HPC, PVA, Kollidon 30, PVP40, Poloxamer 188 and 407Powders were suspended in saturated and filtered solution of indomethacin in 30% glycerin solution0.283 e0.310 e,fKumar et al., (2015) [151]
WMMFluid bed coating/dryingFenofibrate (14.1%)(HPMC E3, SDS, Mannitol), Pharmatose (core), GranuLac (core), potato starch (core)100 mg dried powders were dispersed in 8 mL DI water and shaken manually for 30 s 0.1600.265 fKnieke et al., (2015) [152]
WMMSpray dryingGriseofulvin (75.2%)(HPC SL, SDS), CCS, SSG, MannitolQualitative assessment: 10 mg dried powder was dispersed in 4 mL DI water and observed the cloudiness0.161Cloudy supernatant fAzad et al., (2015) [27]
WMMWet film casting–dryingFenofibrate (17%)
Griseofulvin (14.5%)
Naproxen (14.2%)
Phenylbutazone (15.1%)
Azodicarbonamide (17.7%)
(SDS, HPMC E15 LV, Glycerin)~0.71 cm2 circular films were dispersed in 3 mL DI water and vortex mixed for 3–5 min0.178
0.16
10.136
0.176
0.278
0.283 f
0.164 f
0.134 f
0.184 f
0.352 f
Krull et al., (2015) [153]
WMMSpray dryingNaproxen (–)(HPMC E15, Trehalose), Tween 80, LactosePowders were suspended in saturated and filtered solution of naproxen in 30% glycerin solution0.243 e0.282 e,fKumar et al., (2015) [154]
HPHFreeze dryingHarmine (–)(HPMC E15, Sorbitol), TPGS, Tween 80, CMS-Na, RH40, Sucrose, Glucose, Trehalose, MannitolMalvern Mastersizer
RDI = mean redispersion size/nanosuspension size ×100
0.500–0.700~100% gYue et al., (2015) [155]
Spray dryingHarmine (–)(CMS-Na), HPMC E15, TPGS, Tween 80, CMS-Na, RH40 0.500–0.700~100% g
WMMWet film casting–dryingGriseofulvin (15.8%)(HPMC E15, SDS, Glycerin, CCS/HPMC E4), SSG, CP, GG, XG, Pectin~0.71 cm2 circular films were dispersed in 10 mL DI water and vortex mixed for 1 min 0.1600.160 fSusarla et al., (2015) [80]
LASPFreeze dryingUrsolic acid (–)(TPGS 1000, Trehalose), Maltose, Glucose, Sucrose, PEG2000Dried powders were dispersed in water and shaken manually0.127 e0.239 e,fGe et al., (2015) [156]
WMMSpray dryingIndomethacin (36.3%)(Poloxamers 188, Mannitol, L-leucine), Poloxamers 407 and184, TPGS 1000Malvern Nano ZS0.263 e0.417 e,fMalamatari et al., (2015) [157]
HPHSpray dryingItraconazole (–)(Poloxamers 407, SLS, Mannitol)Malvern Zetasizer~0.316 e~0.320 e,fSun et al., (2015) [107]
LASP–UltrasonicationFreeze dryingNaproxen (9.42%)(HPMC)_0.530 e_ fMishra et al., (2015) [158]
HPHWet film casting–dryingQuercetin
(10 mg/6 cm2)
(Maltodextrins, Glycerin, Tween 80, Span 80)Malvern Zetasizer® Nano ZS0.753 e0.781 e,fLai et al., (2015) [159]
Freeze dryingQuercetin (–)(Tween 80) 0.753 e~0.921 e,f
LASP–UltrasonicationFreeze dryingTeniposide (–)(PVP K30), Poloxamer 188, HPMCPowders were suspended with 5% glucose and vortexed for 5 s0.151 e0.151 e,fHe et al., (2015) [160]
Acid-base neutralization–UltrasonicationSpray dryingGlipizide (–)(SLS, MCC), Poloxamer, PVP, HPMC, Tween 80, Mannitol, Sorbitol_0.262_ fElham et al., (2015) [111]
ME–HPHFreeze dryingLambda-cyhalothrin (–)(MRES, SDS), PVP K30, PVP K90, SL, Tween 80, PEGNPE, HPMC, Poloxamer 188, Span 80, Mannitol_0.016 e_ gPan et al., (2015) [161]
WMMWet film casting–dryingGriseofulvin (8.75%)(Pullulan, SDS, XG, Glycerin)10 mg of the dry film was dispersed in 20 mL DI water and vortexed for 2 min at 1500 rpm0.1680.379Krull et al., (2016) [162]
WMMWet film casting–dryingGriseofulvin (15.4%)(HPMC E15LV, SDS, PEG 400), Triacetin, Glycerin~0.7 cm2 circular dried film was dispersed in 3 mL DI water and vortexed at 1500 rpm for 3–5 min 0.1590.169 fKrull et al., (2016) [81]
MEWet film casting–dryingFenofibrate (5.03%)(Pluronic F68, HPMC E15 LV, Glycerin)2.54 cm diameter film was dispersed in 10 mL DI water and stirred for 5 min at 475 rpm using a magnetic stirrer~0.500~0.800 fBhakay et al., (2016) [40]
WMMFluid bed coating/dryingGriseofulvin (12.0%)(HPC L, SDS), Pharmatose (core), HPC SSL, HPC SL1 g of the dried powders was dispersed in 30 mL DI water and paddle stirred for 2 min at 200 rpm ~0.250~0.250 fLi et al., (2016) [86]
HPHSpray dryingItraconazole (–)(Tween 20, Methocel E5)Powder corresponding to 0.2 g drug was dispersed in 25 mL water and shaken manually for 30 s0.313 e0.425 eBonda et al., (2016) [163]
HPHSpray dryingAlbendazole (–)(Poloxamer 188), SDS, Poloxamer 40720 mg powders were dispersed in 5 mL DI water and manually shaken for 1 min0.450 e0.550 e,fParades et al., (2016) [164]
HPHNanoextrusionEfavirenz (1.02%)(SLS, Kollidon® 30, Soluplus)Extrudates were dispersed in water and vortexed for 30 s0.320 e~0.030 fYe et al., (2016) [77]
LASP–HPHFreeze dryingUrsodeoxycholic (–)(PVP K30, Sucrose/Glucose), TPGS, Poloxamer 188, Lactose Poloxamer 407, RH40, HPMC, SDS, Tween 80, PEG4000, Trehalose, CMS-Na, MCCS, Mannitol, SorbitolMalvern Mastersizer
RDI = mean redispersion size/nanosuspension
size ×100
0.600–0.900~100% gMa et al., (2016) [165]
LASPFreeze dryingGambogenic acid (29.7%)(PVP K30, PEG2000, Trehalose, Mannitol)Malvern Zetasizer~0.184 e~0.198 e,fYuan et al., (2016) [166]
HPHSpray dryingModel drug (–)(HPC L, MCCS), HPMC, MCC, CCS, Trehalose, LactoseMalvern Mastersizer0.4610.478 fDan et al., (2016) [167]
LASP–UltrasonicationFreeze drying20(S)-protopanaxadiol (50%)(Bovine serum)_0.222 e_ fHan et al., (2016) [168]
WMMSpray dryingGriseofulvin (–)(HPC SL, Docusate sodium, Mannitol)Dried powders were dispersed in water and simulated gastric fluid (pH 1.2) using USP II 0.205 e0.200–0.300 e,fShah et al., (2016) [169]
LASP–UltrasonicationSpray dryingCefixime trihydrate (–)(PVP K30, Lactose monohydrate), Mannitol, Sorbitol Excess amount of dried powder was dispersed in 2 mL distilled water and sonicated for short time0.266 e~0.300 e,fAlaei et al., (2016) [170]
WMMWet film casting–dryingNaproxen (–)
Anthraquinone (–)
(HPMC, Glycerol, PVP VA64, SDS)4 cm2 film was dispersed in 0.9 mL liquid (DI water, tap water and saliva substitute) for 1–15 min and further diluted with DI water 0.270
0.273
0.280–0.315 g
0.340–0.420 g
Steiner et al., (2016) [171]
WMMFluid bed coating/dryingItraconazole (14.8%)
Fenofibrate (13.6%)
(HPMC, SDS), Hydrophilic silica), GranuLac® 200 (core), PrismaLac® 40 (core)1 g of dried products were dispersed in 30 mL of 7.2 and 2.88 mg/mL SDS and stirred at 110 rpm for 2 min0.172
0.171
0.490 f
0.290 f
Azad et al., (2016) [91]
LASP–UltrasonicationFreeze dryingEfavirenz (–)(Poloxamer 407, Soya lecithin, Mannitol)_~0.184 e_ fTaneja et al.,
(2016) [172]
Acid-base neutralization–HPHFreeze dryingHerpetospermum caudigerum lignans (–)(SDS, PVP K30, Mannitol), Poloxamer 188, Tween 80, HPMC, PVA, LecithinDried products were dispersed in DI water and shaken for 1 min0.243 e0.286 e,fGang et al., (2016) [173]
HPHSpray dryingAndrographolide (–)(HPMC E15/MCCS, Lactose), RH 40, TPGS, PVP K30, Tween 80, Sucrose, Trehalose, Mannitol, SorbitolMalvern Mastersizer
RDI = mean redispersion size/nanosuspension
size ×100
0.500–0.900~100% gXie et al., (2016) [174]
Freeze dryingAndrographolide (–)(HPMC E15, Tween 80, Sucrose/Trehalose/Sorbitol), RH 40, TPGS, MCCS, PVP K30, Mannitol, Lactose 0.500–0.900~100% g
LASP–UltrasonicationFreeze dryingCinnarizine (9.53%)(PVA)_0.621 e_ fMishra et al., (2016) [175]
HPHFreeze dryingRitonavir (–)(HPMC 3, SDS, Mannitol), PVP K30 _0.562 e_ fKarakucuk et al., (2016) [176]
WMMFreeze dryingBI XX (–)(Mannitol, Arginine), Polysorbate 80, PEG 400, Proline, Benzalkonium chloride2 mL DI water was added to the dried product and shaken manually 0.192~0.250 gFrank and Boeck (2016) [177]
LASP–UltrasonicationWet film casting–dryingLercanidipine HCl hemihydrate (3.06 mg)(TPGS 1000, Hypromellose E15), Hypromellose E5, PVA, PEG 400, Sodium alginate, MC, HPC, HPMC E54 cm2 films were placed in 10 mL DI water and stirred for 10 min using magnetic stirrer~0.277 e~0.240 e,fChonkar et al., (2016) [178]
WMMFreeze dryingSpironolactone (–)(HPMC E5, Sorbitol), Sodium deoxycholate, Poloxamer 407, Poloxamer 188, MannitolDried powders were dispersed in water by gentle shaking 0.374 e0.399 e,fMu et al., (2016) [179]
WMMNanoextrusionGriseofulvin (24.1%)(HPC SL, SDS), Soluplus_0.154_ fLi et al., (2017) [78]
WMMSpray dryingAprepitant (–)(Pharmacoat 603, SDS, sucrose), HPMC E15, HPC SSL, PVP K30, TPGS 1000, Poloxamer P188, MannitolDried powders were dispersed in few milliliters of distilled water and shaken manually for 30 s0.395 e0.420 e,gToziopoulou et al., (2017) [180]
Freeze dryingAprepitant (–)(Pharmacoat 603, SDS, Mannitol) 0.395 e0.395 e,g
LASP–HPHFreeze dryingP2X7 receptor antagonist (–)(Poloxamer 188, Mannitol), HPMC, SDS_0.245 e_ fZhang et al., (2017) [181]
WMMFreeze dryingNisoldipine (–)(PVP K30, SDS, Trehalose), HPMC E5_0.240 e_ fFu et al., (2017) [113]
WMMFreeze dryingDexamethasone (–)
Tacrolimus (–)
(Poloxamer 407)
(Poloxamer 407)
Malvern Zetasizer® Nano ZS900.403 e0.511 e0.300–0.700 e,g
0.300–0.800 e,g
Colombo et al., (2017) [182]
WMMElectrospraying–Freeze dryingDarunavir (–)(Tween 20, SLS, Eudragit L100, Mannitol), Poloxamer 338, 188, HPMC, Vitamin E TPGS 400 and 1000, Tween 80_0.295 e_ fNguyen et al., (2017) [183]
LASP–UltrasonicationFreeze dryingEfavirenz (–)(HPMC E5, SLS), HPC ELF, Poloxamer 188 and 407, PVP K30_0.252 e_ fSartori et al., (2017) [184]
WMMSpray dryingAllisartan Isoproxil (61.7%)(PVP K30, Mannitol, SDS), Pluronic F127 and F68, Tween 80, HPMC E5, HPMC E50, HPC, PEG 4000 and 6000, Glucose, Sucrose, Maltose, SorbitolDried powders were dispersed in water at a drug concentration of 2.5% (w/v) and diluted to 900-fold with water~0.300 e~0.304 e,fHou et al., (2017) [185]
HPHFreeze dryingAndrographolide (–)(Glycyrrhizin, Trehalose), Poloxamer 188, Tween 80, TPGS, Sucrose, LactoseMalvern Mastersizer0.4870.491 fChen et al., (2017) [186]
WMMWet film casting–dryingGriseofulvin (50%)(HPMC E15LV, SDS, Glycerin), HPMC 4M~0.7 cm2 circular films were dispersed in DI water and vortexed at 1500 rpm for 3–5 min 0.150<0.200 fKrull et al., (2017) [83]
WMMWet film casting–dryingAnthraquinone (–)(HPMC, Glycerol), HPC, Kollidon VA64, SDS4 cm2 film dispersed in 0.9 mL distilled water for 5 min0.3020.337 gSteiner et al., (2017) [187]
HPHSpray dryingAndrographolide (–)(TPGS, PVP K30, MCC, Lactose), PVPPMalvern Mastersizer
RDI = mean redispersion size/nanosuspension size ×100
~0.514101.5% fXu et al., (2017) [188]
LASP–HPHFreeze drying10-hydroxycamptothecin (94.9%) Camptothecin (91.2%)
7-ethyl-10-hydroxycamptothecin (90.1%)
(mPEG1000-HCPT)Dried products were reconstituted with water~0.093 e
~0.121 e
0.133 e
~0.094 e,f
0.137 e,g
0.133 e,g
Yang et al., (2017) [189]
Ultrasonication–HPHFreeze dryingMeloxicam (–)(Poloxamer 188, Mannitol), PVP K25, PEG 4000Dried products corresponding 0.5 mL suspension were diluted to 15 mL with purified water and sonicated for 60 s~0.463 e~0.501 e,fIurian et al., (2017) [190]
LASP–UltrasonicationWet film casting–dryingLutein (0.23 mg/cm2)(SPC, SDS, HPMC, PEG 400, Cremophor EL), TPGS, PVP K30, Poloxamer 188, PVAA piece of film was dispersed in 5 mL water and sonicated for 3 min0.220 e~0.378 e,fLiu et al., (2017) [191]
WMMFluid bed granulation/dryingLurasidone hydrochloride (–)(HPMC E3, Polysorbate 80, Mannitol, MCC), HPMC E5, Poloxamer 188, SLS, Span 20, Labrasol_0.245 e_ fKumar et al., (2017) [192]
WMMSpray dryingMefenamic acid I (–)
Mefenamic acid II (–)
(HPC SSL, Lactose)_0.188
0.189
_ f_ fKonnerth et al., (2017) [193]
LASP–UltrasonicationFreeze dryingCarvedilol (–)(SDS), Whey protein isolate, Poloxamer 1881 mg sample was dispersed in 3 mL of water and shaken manually for 1 min ~0.225 e0.227 e,fGeng et al., (2017) [194]
a BCN-Baicalin; CCS: Carboxymethyl cellulose sodium; CMS-Na: sodium carboxymethyl starch; CP: Crospovidone; Cremophor® RH 40: Polyoxyethylene hydrogenated castor oil; Dowfax 2A1: alkyldiphenyloxide disulfonate; GG: Guar gum; HPBCD: hydroxypropyl β-cyclodextrin; HPMC: Hydroxypropylmethylcellulose; MC: Methylcellulose; MCC: Microcrystalline cellulose; MCCS: Microcrystalline cellulose and carboxymethyl cellulose sodium; MRES: Maleic rosin-polyoxypropylene-polyoxyethylene ether sulfonate; MMT: Montmorillonite (clay); NGN-Naringenin; OCA-Oleanolic acid; PEG-PLA: Polyethyleneglycol-polylactide; PEG4000: Polyethylene glycol; PVP: Polyvinylpyrrolidone; mPEG1000-HCPT: Polyethylene glycol (PEG)ylated 10-hydroxycamptothecin; RCN-Rubescensin; RPN-Rutacarpine; RVL-Resveratrol; SDS: Sodium dodecyl sulfate; SLS: Sodium lauryl sulfate; SPC: Soy phosphatidylcholine; SSG: Sodium starch glycolate; TPGS 1000: D-α-tocopheryl polyethylene glycol 1000 succinate; XG: Xanthan gum; RDI: Redispersibility index. b The dispersants in the parenthesis refer to the particular nanocomposite formulation that led to the finest drug particle sizes upon redispersion, or that led to the fastest dissolution if redispersion was not performed. c Unless otherwise indicated, all particle sizes refer to the median size (d50) of the nanosuspension (before redispersion) and the redispersed suspension (after redispersion) for the particular nanocomposite formulation that led to the finest drug particle sizes. If redispersion was not performed, “before redispersion” particle size refers to the particle size of the nanosuspension with the finest particle size. If multiple redispersion methods were used, the specific redispersion method used for the reported particle sizes is underlined. d Volume-mean diameter. e Cumulant size was measured by dynamic light scattering. f Fastest dissolving formulation. g Dissolution testing was not performed.
Table 3. Classification of dispersants used in drug-laden nanocomposites.
Table 3. Classification of dispersants used in drug-laden nanocomposites.
ClassExamplesPossible Mechanisms of ActionReferences
Soluble polymersHPC, HPMC, PVP, Soluplus, PEG, PVA, Sodium CMC, methyl cellulose, PEG-PLA, Sodium alginateSteric stabilizer in nanosuspension, enhanced drug wettability in nanocomposites, primary matrix/film former that prevents aggregation during drying and facilitates erosion/disintegration of nanocomposites via dissolution.Lee et al., 2003 [33], Li et al., 2016 [23], Chin et al. [55]
SurfactantsAlpha tocopherol succinate, SDS, TPGS, Poloxamer 338, Pluronic F127, Pluronic F68, Span 20/60, Tween 20/60, Dowfax 2A1, Soy lecithin, Docussate sodium, Cremophor, Sodium deoxycholateSteric or electrostatic stabilizer depending on its charge, wetting agent, enhanced drug wettability in nanocompositesBonda et al., 2016 [163], Li et al., 2016 [23], Chin et al., [55]
Other water-soluble dispersants (WSD) Sucrose, Lactose, Trehalose, Glucose, Maltose, Maltodextrin, Ficoll PM70, HPβ-cyclodextrin, L-Leucine, Mannitol, Arginine, sodium carboxymethyl starch (CMS-Na)Secondary matrix former that prevents aggregation during drying, facilitates erosion/disintegration of nanocomposite particles via dissolution, and act as cryoprotectant in freeze drying Abdelwahed et al., 2006 [197], Kesisoglou et al. [28], Chin et al. [55]
Water-insoluble dispersants (WID)Microcrystalline cellulose, anhydrous dicalcium phosphate, colloidal fumed silica, montmorilloniteSecondary matrix former that prevents aggregation during drying and facilitates erosion/disintegration of nanocompositesEerdenbrugh et al., 2008 [59]
Crosslinked polymers (CLP)SSG, CCS, CPSwellable dispersant that facilitates erosion/disintegration of nanocomposite particles via swelling-induced breakage/erosion of nanocomposite matrix Bhakay et al., 2014 [8], Azad et al., 2015 [27]

Share and Cite

MDPI and ACS Style

Bhakay, A.; Rahman, M.; Dave, R.N.; Bilgili, E. Bioavailability Enhancement of Poorly Water-Soluble Drugs via Nanocomposites: Formulation–Processing Aspects and Challenges. Pharmaceutics 2018, 10, 86. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics10030086

AMA Style

Bhakay A, Rahman M, Dave RN, Bilgili E. Bioavailability Enhancement of Poorly Water-Soluble Drugs via Nanocomposites: Formulation–Processing Aspects and Challenges. Pharmaceutics. 2018; 10(3):86. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics10030086

Chicago/Turabian Style

Bhakay, Anagha, Mahbubur Rahman, Rajesh N. Dave, and Ecevit Bilgili. 2018. "Bioavailability Enhancement of Poorly Water-Soluble Drugs via Nanocomposites: Formulation–Processing Aspects and Challenges" Pharmaceutics 10, no. 3: 86. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics10030086

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