Skip to the main content

Review article

https://doi.org/10.2478/acph-2024-0015

Taste-masking methods in multiparticulate dosage forms with a focus on poorly soluble drugs

TILEN SIMŠIČ ; Faculty of Pharmacy, University of Ljubljana, 1000 Ljubljana, Slovenia; Alterno Labs d.o.o., 1231 Ljubljana-Črnuče, Slovenia
ODON PLANINŠEK ; Faculty of Pharmacy, University of Ljubljana, 1000 Ljubljana, Slovenia
ANA BAUMGARTNER ; Faculty of Pharmacy, University of Ljubljana, 1000 Ljubljana, Slovenia


Full text: english pdf 655 Kb

versions

page 177-199

downloads: 0

cite

Download JATS file


Abstract

In the past, the administration of medicines for children mainly involved changes to adult dosage forms, such as crushing tablets or opening capsules. However, these methods often led to inconsistent dosing, resulting in under- or overdosing. To address this problem and promote adherence, numerous initiatives, and regulatory frameworks have been developed to develop more child-friendly dosage forms. In recent years, multiparticulate dosage forms such as mini-tablets, pellets, and granules have gained popularity. However, a major challenge that persists is effectively masking the bitter taste of drugs in such formulations. This review therefore provides a brief overview of the current state of the art in taste masking techniques, with a particular focus on taste masking by film coating. Methods for evaluating the effectiveness of taste masking are also discussed and commented on. Another important issue that arises frequently in this area is achieving sufficient dissolution of poorly water-soluble drugs. Since the simultaneous combination of sufficient dissolution and taste masking is particularly challenging, the second objective of this review is to provide a critical summary of studies dealing with multiparticulate formulations that are tackling both of these issues.

Keywords

multiparticulatedosage forms; taste masking; paediatric; film coating; solubility enhancement; evaluation methods

Hrčak ID:

313324

URI

https://hrcak.srce.hr/313324

Publication date:

30.6.2024.

Visits: 0 *




INTRODUCTION

Historically, the majority of oral dosage forms administered to children were off-label or unlicensed. To ensure accurate dosing for neonates and children, healthcare professionals often resorted to modifying adult dosage forms, such as crushing tablets and mixing them with liquid, breaking tablets into smaller portions, or opening capsules and dividing the enclosed powder for administration. (1). Since the beginning of this millennium, several tools for personalised medicine have been introduced in paediatrics, such as mini-tablets or 3D printing of drugs (1, 2), but pharmaceutical compounding (dose adaptation done by pharmacists) and manipulation (dose adjustment at the point of administration by nurses or parents) are still the most commonly used alternatives. Several studies have highlighted the difficulty of obtaining an accurate dose from manipulated tablets, with large dose variations and discrepancies between manipulations due to galenics, prescribed dose, or manipulator experience (3–5). Therefore, these practises cannot be considered equivalent to commercial forms manufactured according to good manufacturing practises (6, 7). The consequences of using compounding/manipulation of conventional drugs approved for adults can translate into suboptimal adherence due to bad taste, underdosing with therapeutic failure, or overdosing with adverse effects (8–10).

In the last two decades, the field of paediatric pharmacy has been advanced by the development of improved paediatric formulations. The main focus has been on the development of age-appropriate and child-friendly medicines, with the introduction of a regulatory framework as an incentive. Since 2007, the European pharmaceutical industry has been prohibited from developing new medicines intended only for use in adults. Paediatric Regulation requires that children be included in a form of paediatric development called Paediatric Investigation Plan (PIP), which requires the agreement of the European Medicines Agency (EMA) and its Paediatric Committee (11). In 2007, the World Health Organisation (WHO) also launched its “Make Medicines Child Size” initiative (12). Since 2009, all applications with a new aspect (indication, route of administration, dosage form) must include information and study results that comply with PIP (2). In 2014, the EMA introduced guidelines for the pharmaceutical development of medicinal products for paediatric use (for children younger than 18 years of age) (13). In the United States, the Food and Drug Administration (FDA) enacted the Best Pharmaceuticals for Children Act in 2002 to provide financial incentives for voluntary paediatric studies conducted by companies. In 2003, the Paediatric Research Equity Act required companies to evaluate drug safety and effectiveness of medicines in paediatric patients (14). These two acts were reauthorized in 2007 as part of the FDA Amendment Act. In the United States, the document corresponding to PIP is called the Paediatric Study Plan. The outcomes arising from of all the mentioned paediatric regulation were examined by the EMA, more specifically the Paediatric Committee, as documented in their 2017 10-year report submitted to the European Commission (E.C., 2017). Three main objectives were defined in this report: to promote research and development of pharmaceutical formulations for children; to ensure that paediatric pharmaceutical formulations are designed and approved for their use in the near future; to increase the availability of information on this topic by encouraging stakeholders cooperate with each other and share new information (15). There are already some excellent reviews that address the progress and challenges of paediatric formulation development (7, 16–18).

The term “paediatric” encloses several age groups, including neonates (0–27 days), infants (28 days–23 months), children (2–11 years), and adolescents (12–17 years), leading to differences based on their physiologic development that may be critical to the development of dosage forms (19).

Liquid formulations have traditionally been considered the most optimal medications for children because they are easy to swallow and their doses can be easily adjusted. However, some of their major drawbacks include poor stability, low potential for taste masking and controlled release, and a greater number of excipients. In 2008, WHO and the EMA, in collaboration with experts in the field, stated that flexible solid oral dosage forms are probably best suited for children. These are defined as those that can be administered in more than one way, such as dispersed in water or breast milk or taken orally as a whole (7). Multiparticulate formulations, or in short multiparticulates, definitely meet this requirement well.

Multiparticulates are small solid multiple-unit dosage forms including microparticles, nanoparticles, granules, pellets, beads, and mini-tablets, usually less than 4 mm in diameter, where a dose consists of multiple particles. Depending on formulation characteristics, multiparticulates can be mixed with beverages or soft foods, or they can even be placed directly in the oral cavity. They are not intended to be dissolved in liquid for administration, because this may affect palatability as well as influence pharmacokinetic profile (7). A study assessing the acceptability of placebo multiparticulates of different sizes in healthy children and adults concluded that multiparticulates could serve as a viable means for administering drugs to children aged four years and older (20). It was also found that multiparticulates were the most common solid oral formulation among paediatric oral formulations during 2007–2018, confirming their suitability for children (21). Because dose adjustment is straightforward with such formulations, especially with the novel dispensing systems (26–31), the main focus is on taste masking of poor-tasting active pharmaceutical ingredients (APIs), which can be challenging when dealing with small particles (22–24). The unpleasant, particularly bitter taste of an API has been cited as a key obstacle in paediatric drug formulation development, as children are known to be more sensitive to bitter and unfamiliar-tasting substances than adults (16, 25). Therefore, it is crucial to develop approaches to mask the taste of APIs in order to comply with the children's needs.

The aim of this review is firstly to provide a general overview of taste masking methods of multiparticulates and a critical review of taste masking evaluation methods. Second, the focus will be switched specifically on multiparticulates containing poorly water-soluble APIs where taste masking is implemented. Their formulation, excipients used, effectiveness of taste masking, and potential improvements of dissolution profile will be discussed in detail. It is the authors' belief that this is a particularly interesting research niche because the aim of most taste-masking approaches is to slow the dissolution of the drug, which is exactly the opposite of what we typically want to achieve with poorly water-soluble APIs, especially when rapid release is desired. Therefore, achieving taste masking and sufficient, or even enhanced, dissolution of a poorly water-soluble API in one formulation is particularly challenging.

TASTE MASKING METHODS

Many methods have been developed to mask the taste of bitter-tasting APIs in multiparticulates. The topic has been discussed in several excellent reviews in recent years (32–35). Briefly, three categories of taste masking approaches can be distinguished, which aim to mask the taste at three different levels: the formulation level, the particle level, and the chemical level (35, 36).

At the formulation level, the use of sweeteners, flavors, bitter blockers, and taste modifiers is most common. Both sweeteners and flavors are very intuitive and simple methods that can mask any taste, not only bitter ones, but their success is limited for highly bitter or very water-soluble APIs as well as with APIs with severe and persistent aftertastes (32, 37). Bitter blockers and taste suppressants act by interfering with taste receptors or taste transduction mechanisms (38). Because there are approximately 25 known bitter receptors, bitter blockers often have the disadvantage of acting on only a specific receptor or a few receptors, depending on their structure. On the other hand, if they act on one of the later stages of taste transduction, they may also block other tastes, making such compounds less commercially desirable (32, 38).

At the particle level, the main objective is usually to reduce or prevent the contact of a bitter-tasting API with taste receptors in the mouth by a coating process, as shown in Fig. 1.

image1.png

Fig. 1. Coating process for taste masking; adapted from Worku et al. (46)

The most common approaches are microencapsulation, polymer film coating, and lipid barrier systems (35). Polymer coating has been reported as the most viable option for taste masking (33, 37). Different types of polymers can be used, either water-soluble, water-insoluble, or with pH-dependent solubility, as shown in Table I (39). The latter may be the most promising because they do not dissolve at salivary pH, whereas they are highly soluble at gastric pH and release the drug rapidly. Eudragit E PO is a cationic copolymer commonly used for taste masking and seems to provide the most promising results (22, 40–42). Lipid barriers could be an interesting alternative because they are easy to use (melting) and allow for an anhydrous process and formulation, which is suitable for moisture-sensitive APIs. In addition, taste masking with lipids can work in different ways, i.e., the API can be dispersed into the lipid structures, bind directly to the lipid molecules, or adsorb on the insoluble surface of the lipid particles (32, 43). However, impurities can affect the melting interval and thus the whole process, which is a drawback of this method. Scale-up can also be very challenging due to difficulties in temperature regulation (44). Microencapsulation can be considered a modern extension of film coating, intended for small particles and therefore very suitable for multi-unit dosage forms A good example of this is the Actimask technology developed by SPI Pharma. Particles of ibuprofen and acetaminophen are coated with gelatin by coacervation from water, which should provide a smooth mouthfeel in addition to taste masking in the first 2 minutes (45).

Table I. Examples of polymers used for taste masking (34, 47)

Acid-soluble polymers Basic-soluble polymers
PolymerTrade namePolymerTrade name
Aminodimethylmethacrylate copolymerEUDRAGIT® E 100 EUDRAGIT® E PO EUDRAGIT® E 12.5 Copolymer of methacrylic acid, type A

EUDRAGIT® L 30 D-55/

L 100-55

Kollicoat® MAE 30 DP/100 P

Copolymer of methacrylic acid, type BEUDRAGIT® L 12,5 EUDRAGIT® L 100
Copolymer of methacrylic acid, type CEUDRAGIT® S 12,5 EUDRAGIT® S 100
Copolymer of methacrylic acidEUDRAGIT® FS 30 D
Sodium alginateKeltone LV CR
Aminodiethylmethacrylate copolymerKollicoat® Smartseal 30 D ShellacSSB 55 Pharma
CarboxymethylcelluloseAkucell
Cellulose acteate phtalate

Aquacoat® CPD

Eastman® C-A-P NF

Cellulose acetate butyrateCAB Eastman
Water-soluble polymers Water-nonsoluble polymers
PolymerTrade namePolymerTrade name
HydroxyethylcelluloseNatrosolTM, Oxycellulose Ethylcellulose

EthocelTM

Auqacoat® ECD

Surelease®

Hydroxypropyl methylcelluloseMethocel® Pharmacoat® Walocel® SpectracelTM Cellulose acetateEastman CA
Sodium carboxymethylcelluloseWalocel® CRT A Poly(ethyl acrylate-co-methyl methacrylate) 2:1EUDRAGIT® NM 30 D
PolyvinylpyrrolidoneKollidon® Ammonium methacrylate Type A: poly(ethyl acrylate-co-methyl methacrylate-co-trimethylammonioethyl methacrylate chloride) 1:2:0.2EUDRAGIT® RL 30 D EUDRAGIT® RL 100/ RL PO EUDRAGIT® RL 12.5
Polyvinylalcohol-polyethylene glycol copolymerKollicoat® IR AquaPolish® Ammonium methacrylate Type B: poly(ethyl acrylate-co-methyl methacrylate-co-trimethylammonioethyl methacrylate chloride) 1:2:0.1EUDRAGIT® RS 30 D EUDRAGIT® RS 100/ RS PO EUDRAGIT® RS 12.5
Polyvinylalcohol-polyethylene glycol copolymer; Polyvinyl alcoholKollicoat® IR Protect Poly(ethyl acrylate-co-methyl methacrylate) 2:1EUDRAGIT® NM 30 D
Polyvinyl alcoholOpadry® AMB Polyvinyl acetateKollicoat® SR 30 D

The chemical level includes a variety of methods, most of which work by reducing the solubility of API in saliva, reducing the number of API particles exposed to taste buds, or reducing its affinity for receptors. This is achieved by chemical modification of API into a prodrug or into a salt, by conversion into another (less soluble) solid form (polymorph, cocrystal), or by conversion into an inclusion complex with ion-exchange resins or cyclodextrins (32, 33). Complexation with cyclodextrins is considered the most challenging platform for taste masking and also requires large amounts of cyclodextrins, so it is not considered very promising. In addition, there is limited data on the potential toxicity of cyclodextrins in children younger than 2 years of age (36, 48, 49).

Recently, there have been some publications on the use of nanocarrier systems and procedures like nanoencapsulation for taste masking. Various additional advantages have been attributed to these, depending on the type of nanocarrier system (improved stability for submicron lipid emulsions, specific targeting for nanogels, high loading capacity for nanostructured lipid carriers). However, the procedures to prepare such systems are usually lengthy, costly, and complicated, so research is still limited (50). In addition, it has been reported that very small particles may remain trapped on the tongue instead of being swallowed, which is not desirable (51, 52).

TASTE MASKING OF MULTIPARTICULATES BY FILM COATING

As mentioned above, polymer film coating has been reported as the most appropriate approach for taste masking, and several studies using this concept have been published to date (Table II).

Polymers

Polymers used for taste masking are based on polysaccharide, polypeptide, or vinyl polymer chemistry. They can be water-soluble, cationic, anionic, or neutral insoluble polymers (34). They must be non-toxic and compatible with the API and other formulation ingredients. It is important to precisely know their solubility in water and organic solvents, pH-dependent solubility, and their glass transition temperature, as these properties affect the film coating process, the necessity to add other ingredients to the coating solution, the effectiveness of film coating and its stability (33, 34).

Kollicoat® Smartseal 30 D and 100 P are both copolymers consisting of methyl methacrylate (MMA) and diethylaminoethyl methacrylate (DEAEMA) in a 7:3 ratio. The difference between them is that Kollicoat Smartseal 30 D is an aqueous dispersion, while 100 P is a spray-dried powder. Both also contain macrogol cetostearyl ether (0.6 and 2 % in 30 D and 100 P, respectively) and sodium lauryl sulfate (0.8 and 2.5 % in 30 D and 100 P, respectively) for stabilization (79, 80). Both are soluble below a pH of 5.5 and insoluble at neutral pH, making them enterosoluble polymers.

Eudragit® E 100 and E PO are cationic copolymers based on dimethylaminoethyl methacrylate, butyl methacrylate, and methyl methacrylate in a ratio of 2:1:1. The difference between them is that Eudragit E 100 is a granulate, while E PO is a powder. Both are soluble below pH 5.0 (81). T The manufacturer also states in its guidelines that E 100 should be used with organic solvents as coating liquid, while E PO should be used with water (49).

Eudragit L 30 D-55 is another polymer with pH-dependent solubility, but unlike the previously described polymers, it is soluble above pH 5.5. However, when combined with other polymers such as Eudragit RL 30 D in the right ratio, a certain dissolution pH can be achieved. For example, Chen et al. (74) coated azithromycin-loaded pellets with a mixture of Eudragit RL 30 D and L 30 D-55 in a ratio of 1:4 (w/w). The coated pellets showed rapid release in the gastric environment (pH up to 3), whereas the release of the drug in the oral environment (pH 5–7) was avoided. Eudragit RL 30 D, on the other hand, has pH-independent solubility. It is provided as an aqueous dispersion of a copolymer of ethyl acrylate, methyl methacrylate, and a small amount of methacrylic acid ester with quaternary ammonium groups. Eudragit L 30 D-55 is also provided as an aqueous dispersion, but the components of this copolymer are methacrylic acid and ethyl acrylate copolymer in the ratio 1:1 (82–84).

Other ingredients used in film coating

Usually, the taste-masking polymer is not the only component of the coating. Plasticisers are a common component of a film coating to ensure better distribution of the coating solution and thus the formation of a coherent film. They act by lowering the glass transition temperature of the polymers, which improves the mechanical properties of the film coating and reduces its brittleness (85). The type and concentration of a plasticizer can significantly affect film properties (86). The most commonly used plasticizers are triethyl citrate and polyethylene glycol, but this list is not exhaustive. It should be mentioned that the use of a plasticizer in the coating also has disadvantages. For example, Stange et al. (71) reported that some polymers, such as Eudragit E 100, are sticky, which can lead to aggregation and agglomeration, and this can be exacerbated by the addition of a plasticizer. Instead of adding an additional component, they suspended the polymer in a mixture of water and ethanol before film coating. Since water itself acts as a plasticizer, 12 % ( m/m) water in the coating suspension was sufficient for a coherent film.

Anti-tacking agents are also frequently used in film coatings to reduce agglomeration and aggregation. Talc or magnesium stearate are commonly used for this purpose. The amount of talc in the coating is often quite high, i.e., even up to 50 % ( m/m) (71, 73), while magnesium stearate is usually used in small amounts, i.e., up to 2 % ( m/m) (60, 63), although there are exceptions with as much as 35 % ( m/m) magnesium stearate in the coating (68).

Another important ingredient in a coating that is commonly used is a surfactant, such as sodium lauryl sulfate. This is because the addition of a surfactant can improve the penetration and spreading properties of the coating liquid, which can otherwise be hindered by the hydrophobic components of the coating or core dosage form. Other ingredients that may also be used include colorants and antioxidants (86). Pigments can also be used in film coatings because they are insoluble in water and thus tend not to migrate into and between coatings, as can be the case when using colorants (87).

Table II. Studies dealing with taste masking of multiparticulates by film coating

API; bitterness threshold value (µg mL–1) Dosage formCoating excipient (best one)(Minimal) amount of coating to achieve taste maskingOther excipients in coatingCoating methodRef.
Chlorpheniramine maleate; 100pelletsKollicoat Smartseal 100P9 % ( m/m) fluid bed53, 54
Praziquantel; 30ground extrudateEudragit E PO56.5 % ( m/m) glyceryl monostearatehot melt extrusion, spray drying55
Sodium benzoate; NApelletsKollicoat SR 30D4 mg cm–2 propylene glycol, talcbottom spray fluid bed56
Paracetamol; 1080pelletsHPMC40 % ( m/m) Na2CO3, cetanol salting out in fluid bed57, 58
Cetirizine; NAcoated drug particles; ODTsEudragit RL-30 DNot giventalc, triethyl citratefluid bed59
Clindamycin; 12.5coated beads; ODTsEudragit E PO35 % ( m/m) stearic acid, Mg stearate, SLSfluid bed60, 61
Prednisone; NAmini tabletsEudragit E PONot givendip coating62
Diclofenac sodium; 20microcapsulesEthyl cellulose50 % ( m/m) diethyl phthalate or PEG 600, Mg stearateorganic phase separation63, 64
Mirtazapine; 100granules; ODTsEudragit E 1008 % ( m/m) coacervation65, 66
Crizotinib; NAmicrospheresKollicoat Smartseal 30 DNot givenbutylated hydroxytoluene, triethyl citrate, talcfluid bed67
Quinine sulphate; 0.007pelletsEudragit E PO20 % ( m/m) SLS, Mg stearate, stearic acid/dibutyl sebacatefluid bed68, 69
Chlorphenamine maleate; 100pelletsKollicoat Smartseal 100P9 % ( m/m) fluid bed54, 70
Naproxen; 1124granulesEudragit E 10061 % ( m/m) talcfluid bed71
Berberine hydrochloride; NAmicrocapsules; ODTsEudragit E 10044 % ( m/m) fluid bed72
Diclofenac sodium; 20pellets; ODTsEudragit E 1005 % ( m/m) talc, PEG 6000fluid bed64, 73
Azithromycin; 27.9pelletsEudragit RL-30 D and L 30 D-5536.3 % ( m/m) triethyl citratefluid bed74, 75
Granisetron hydrochloride; NApellets; ODTsKollicoat Smartseal 30DNot giventalc, colorant, plasticizerfluid bed76
Paracetamol; 1080drug particles; ODTsEudragit E PO28.8 % ( m/m) Stearic acid, talc, SLSfluid bed42
Ibuprofen; 1000drug paticlesHPMC/EC10 % ( m/m) Talcfluid bed77, 78

NA – not available, ODTs – orodispersible tablets, HPMC – hydroxypropyl methylcellulose, PEG – polyethylene glycol, SLS – sodium lauryl sulphate, EC – ethyl cellulose

Coating (film) amount

As shown in Table II, the amount of coating required to achieve complete taste masking is case-specific and ranges from 5 to 62 % ( m/m). Interestingly, both of these values were achieved with the same polymer, Eudragit E 100, applied to the multiparticulates by the fluid bed method. In the formulation with 5 % coating, diclofenac sodium was taste-masked, while in the formulation with 62 % coating, the used API was naproxen sodium. Considering this, the bitter taste of naproxen seems to be detected faster than that of diclofenac sodium. However, looking at the bitterness threshold values in Table II, it appears that diclofenac sodium is far more bitter than naproxen, which is somewhat surprising. Nevertheless, Stange et al. (71) acknowledged in their study that their formulation could contain a lower percentage of coating and still be an effective taste masking agent. Looking more closely at the relationship between the reported bitterness threshold value and the percentage of coating, there is no clear correlation, meaning that a low bitterness threshold value does not necessarily mean that the percentage of coating was high. This suggests that other factors besides the amount of coating are important for taste masking, such as the coating method, the type of polymer, and coating coherency. In addition, particle properties are also of utmost importance. For example, smaller particles have a higher specific surface area, so larger coating amounts are required. The morphology of the particles and their friability can also play an important role. For example, irregularly shaped and brittle particles are more difficult to coat, and thus taste masking is less efficient (24).

EVALUATION OF TASTE MASKING

To assess the effectiveness of taste masking, different evaluation methods can be used, which can be briefly divided into in vivo and in vitro methods.

In vivo methods are usually performed with human volunteers and, despite their high cost and ethical and toxicological problems, are still a very common approach for evaluating taste masking (88, 89). Moreover, their use in the paediatric population is even more problematic due to restrictions and safety concerns. Therefore, it is more appropriate to use the adult population, even though their taste perception may differ from that of children The common procedure of these methods is as follows: A group of subjects is asked to hold the sample in the oral cavity for a specified time, after which they must rate the taste using a numerical or descriptive scale. In addition, subjects are sometimes asked to rate other sensory properties of the samples, such as palatability, mouthfeel, roughness, aftertaste, etc. (63, 71–73). The fact that such properties can only be evaluated by humans could be considered an important advantage of in vivo methods, as these properties can significantly influence the acceptability of the product by patients (71). Because there are no standard protocols for taste masking evaluation (88), the in vivo methods described in research are quite versatile in terms of the number of subjects, time points to assess the taste, and the rating system. The group of volunteers is usually quite small, i.e., about 10 subjects, and it is usually emphasised that they were healthy at the time of testing, which is important because some diseases (influenza, covid-19, etc.) can affect taste perception. Usually, the volunteers are asked not to swallow the investigated sample and should rinse their mouths with water afterwards. The time required to evaluate the taste of the sample can range from 5 seconds to several minutes, which limits to some extent the comparability of the results of different studies. Because multiparticulates are generally easy to swallow, time intervals longer than a minute might seem long, but on the other hand, it is possible that small particles get stuck between the teeth and are swallowed later (71, 90). Therefore, it is useful to determine the amount of fluid to be taken with the medication to ensure that an adequate amount of particles is flushed from the oral cavity. Scoring systems typically include up to five numeric or descriptive options or levels to assess bitterness or other taste qualities. The literature also mentions that facial hedonic scales are appropriate for younger children but are less commonly used because they provide less information than other scoring systems (91).

In vitro methods can generally be divided into electronic tongue measurements and drug release testing (36, 92). Electronic tongues or e-tongues are sensory array systems that can detect individual substances or mixtures by sensor membranes and electrochemical techniques (93). These sensors are designed to mimic taste buds by eliciting changes in electrochemical potentials that should be compared with the actual physiological action potentials that are induced during taste perception (94). After measuring the samples, which must be solutions, the results are usually analyzed by principal component analysis (PCA) or similar statistical methods. To assess whether taste masking has been achieved, the distances between samples and placebos or pleasant and bitter taste references are evaluated in two-dimensional PCA maps (93–96). Although e-tongues have become an important and useful tool in taste masking research, one must be cautious in interpreting the results. For example, sensors respond differently to charged and uncharged molecules, and furthermore, interactions between formulation components can lead to misleading results (93).

Drug release assays for taste masking evaluation are a widely used method, most likely due to their simplicity, objectivity, low cost, and time efficiency. However, there is a lack of a standardized procedure for conducting the analysis. Therefore, there are a variety of in vitro drug release methods that differ in terms of the volume of dissolution medium, its composition, pH, duration of the test, and the device in which the test is performed. The most commonly used dissolution medium is phosphate buffer with a pH of 6.8 and a temperature of 37 °C (58, 60, 71, 97, 98). There are also some cases where distilled water (sometimes with a surfactant) was used instead (68, 93, 95). Sometimes simulated salivary fluid (SSF) is used to closely mimic conditions in the oral cavity, but the composition is not the same in every study (55, 65, 73, 99). Regarding the volume of the dissolution medium, it would probably be most appropriate to use the approximate volume of residual saliva in the oral cavity, which is 1–2 mL (88). However, the volume of the dissolution medium has been found to range from 2 to 900 mL, which is very different from the principles of physiological conditions. There is also considerable variation in the sampling time required to estimate the efficiency of taste masking, which goes even up to 15 min. Interestingly, there are large differences between in vivo and in vitro time points for estimating taste masking, and it would probably be more appropriate to use similar and, more importantly, physiologically relevant time intervals in both types of experiments. Indeed, drug release after 5 minutes can hardly serve as a metric for successful taste masking, since the residence time of oral dosage forms in the mouth is probably much shorter.

Although the USP I/ II apparatus or the shake flask method are commonly used to perform drug release assays, some attempts have been made to better simulate the oral cavity. For example, Keeley et al. developed a flow-through dissolution column (Fig. 2) and compared it to the USP I dissolution test. While both systems were able to distinguish between different types and levels of coating to mask chlorphenamine taste, only the novel dissolution test could give absolute quantification of taste masking based on previously determined taste thresholds. Although the authors also mention the limited size of the dosage forms as a disadvantage of this method, this is irrelevant for multiparticulates, as they are small enough to be introduced into the device (70).

image2.png

Fig. 2. Flow-through dissolution column designed by Keeley et al. Reprinted from reference (70) with permission from the American Chemical Society (ACS).

It is also important to note that the results of the drug release experiment should be compared to a specific reference value in order to claim the effectiveness of taste masking with greater confidence. The most commonly used, and probably most advisable, reference value to which to compare results is the bitterness threshold value, which is the concentration of the substance above which a bitter taste can be detected. This value is substance-specific and its determination is defined in the European Pharmacopeia in Test 2.8.15, "Bitterness Value" (100). A minimum of six volunteers are recommended, who should rinse their mouths before testing. To correct for differences in taste perception among panel members, a correction factor is first determined across different concentrations of quinine hydrochloride solution. To determine at which concentration the bitter taste of the API is detectable, several dilutions of the drug are prepared according to a specific protocol. If the concentration in the dissolution medium after a certain time point is higher than the specified value, taste masking has not been fully achieved. However, it should be noted that the value reported in the literature may also vary. In addition, if sweeteners or aromas are added to the formulation, they may compensate for the initial release of the bitter API, and taste masking may still be achieved. Another possible reference for comparing release results is the FIP /AAPS (Federation International Pharmaceutique/American Association of Pharmaceutical Sciences) guideline, which states that less than 10 % ( m/V) of the API should be released within 5 minutes to ensure taste masking (101). However, since this is an arbitrary value, it may not be applicable to every single substance and also depends on the dissolution method used (95). Although there are studies that evaluate taste masking based on the dissolution profile of their formulation in the first few minutes compared with the dissolution of the pure API, this comparison may not be appropriate and could lead to misinterpretation of the results. In this case, it is advisable to use a complementary method to evaluate taste masking, such as e-tongue or in vivo tests.

TASTE MASKING OF POORLY WATER-SOLUBLE DRUGS IN MULTIPARTICULATES

As mentioned earlier, the principle of taste masking is usually to prevent contact between a bitter-tasting API and the taste buds by slowing the dissolution of the drug in the oral cavity. However, it is possible that this may also reduce the dissolution of the API further down the gastrointestinal tract. For highly soluble APIs, this may not be a problem if their concentration remains high enough to reach the therapeutic range after absorption. For poorly soluble APIs, on the other hand, this could result in concentrations that are below therapeutic levels. For this reason, taste masking of poorly soluble APIs presents a particular technological challenge. It is therefore not surprising that there are only a limited number of studies in which taste-masked multiparticulate drugs have been formulated using a (model) poorly water-soluble API (Table III).

Table III. Overview of publications dealing with taste-masked multiparticulates including a poorly water-soluble API

APIBCS classTaste masking approachTaste masking excipienta Ref.
PraziquantelIIEncapsulationEudragit E PO, glyceryl monostearate55
IbuprofenIILipid matrixCocoa butter99
NaproxenIIFilm coatingEudragit E71
KetoprofenIISolid dispersionEudragit E PO94
Diclofenac sodiumIIFilm coatingEudragit E73
MirtazepineIICoacervationEudragit E65
Ondansetron HClIILipid matrixGeleol97
Diclofenac sodiumIIMicroencapsulationEthyl cellulose63
AzithromycinIIFilm coatingEudragit R, L74
QuercetinIVMicroencapsulationCarnauba wax102

a If more than one excipient was tested, only the one giving the best results is given. See the text below for further explanation.

As shown in Table III, most researchers used either Eudragit E or Eudragit E PO, which are both enterosoluble polymers, to mask the taste of bitter-tasting, poorly water-soluble substances. In this way, the polymer prevents the taste buds from coming into contact with the API, but immediately after entering the stomach, the polymer dissolves and no longer interferes with the dissolution process of the API.

Münster et al. (55) attempted to develop a formulation that combined immediate release and taste masking of poorly water-soluble praziquantel. They tested several polymers (Eudragit E PO, RL PO, L100, S100, Kollidon SR) to determine which exhibited the best miscibility with praziquantel. Since dynamic scanning calorimetry (DSC) measurements showed no difference between samples, a small-scale drug release study was conducted in SSF, simulated gastric fluid sine pepsin (spSGF), and fasted state simulated intestinal fluid (FaSSIF) with films of amorphous solid dispersions in a 96-well plate. It was found that Eudragit E PO reduced drug release by 6-fold in SSF compared to pure API, but allowed the highest degree of supersaturation in the other two media. Hot melt extrusion was used to prepare a multiparticulate formulation, and glyceryl monostearate was used as an additional excipient for taste masking, since it is insoluble in water, and as a lubricant to ensure better flowability of the mixture. In fact, it enhanced the taste masking effect by up to 11-fold compared to the pure polymer, as determined by dissolution in SSF, and concentrations of praziquantel remained below the reported bitterness threshold value for the first 5 min. However, as the authors pointed out, these results would need to be confirmed in vivo. Although praziquantel was present in the extrudates in a crystalline state, dissolution in spSGF and FaSSIF was nevertheless improved compared with the pure API due to the improved wettability provided by Eudragit E PO and glyceryl monostearate, as well as the solubilizing effect of Eudragit E PO. The fact that the API was not amorphous also contributed to the stability of the formulations after eight weeks of storage under various conditions, as evidenced by only minor differences in drug release.

Alshetaili et al. (94) used ketoprofen as a poorly soluble, bitter-tasting model drug and mixed it with Eudragit E PO by hot melt extrusion. In this case, no other excipient was needed because ketoprofen itself acted as a plasticizer. The multiparticulate dosage form they formed the extrudate into was mini tablets. DSC measurements confirmed that ketoprofen was completely amorphous in the mini-tablets. When dissolution in an acidic medium was tested, formulations containing 20, 30 and 40 % ketoprofen showed 100 % release within 20 minutes, which was attributed to the solubilizing effect of Eudragit E PO and salt formation between the acidic groups in ketoprofen and the tertiary ammonium groups in the polymer. This was also confirmed by Fourier transform infrared spectroscopy (FTIR). Pure ketoprofen was released only about 20 % in 120 min, showing a significant improvement. In SSF with a pH of 6.8, drug release was less than 0.5 % in the first 2 minutes. Although nothing was said about the threshold value for ketoprofen, it could be concluded that taste masking was indeed achieved, since the formulations were also tested with e-tongue and it was found that the "taste" of the samples was very similar to that of the buffer. Interestingly, after three months of storage, the samples remained amorphous and showed similar release profiles to the freshly prepared samples, which was also attributed to strong interactions between the drug and the polymer.

Stange et al. (71) developed a multiparticulate formulation with naproxen coated with Eudragit E as a taste masking agent. Unlike previous studies in which Eudragit E was used as a film coating polymer, their coating did not contain plasticizers because, as mentioned earlier, these can increase tacking tendency. Instead, they used only talc and a mixture of water and ethanol for the coating, which was done in a fluidized bed coater. As the amount of water increased, the coating efficiency increased and provided better taste masking. Drug release was tested in a two-stage experiment, first in acidic media and after 30 minutes in pH 6.8. After 30 minutes in acidic media, which is supposed to be the shortest residence time of the pellets in the stomach, more than 80 % of naproxen was released from the selected formulation. At a pH of 6.8, only the results for the first 5 minutes are reported, so it is not possible to determine how much API was actually released at the end of the experiment. In addition, no comparison with the dissolution of the pure API is given, so it is difficult to say whether the dissolution rate has changed due to the coating.

Alotaibi et al. (73) developed taste-masked pellets that contained diclofenac sodium, which were subsequently formulated into orodispersible tablets. The sugar pellets were triple-layered; the first layer consisted of diclofenac sodium and HPMC (binder), the second layer was enteric coating consisting of Eudragit L100, talc (anti-adherent), and triethyl citrate (plasticizer), and the third layer was a taste-masking coating made of Eudragit E100, talc (anti-adherent) and PEG 6000 (plasticizer). Dissolution experiments with pellets prepared in a two-step manner (first 2 h at pH 1.2, then 3 h at pH 6.8) showed that the taste-masking coating did not affect drug release in SIF, as Eudragit E100 dissolves in an acidic medium. In SSF, the enteric coating did not significantly delay drug release, but even with 5 % Eudragit E100, drug release was only 3.0 % in 3 minutes. Orodispersible tablets (200 mg) were prepared with the pellets that exhibited the best properties, i.e., the fastest dissolution at pH 6.8 and the slowest dissolution in SSF (12 % Eudragit L100, 5 % Eudragit E100). Different formulations were prepared with varying amounts of superdisintegrants (croscarmellose sodium or crospovidone), microcrystalline cellulose, sodium stearyl fumarate, colloidal silicon dioxide, and mannitol. Mannitol is a commonly used excipient in orodispersible formulations due to its cool, sweet, and mild taste and pleasant mouthfeel due to its negative heat of solution. The formulation that showed the fastest drug release at pH 6.8 and the lowest release at acidic pH was compared with the commercially available Voltaren 25 mg. Drug release in the acidic medium was limited in both cases. However, at pH 6.8, the release rate of the selected formulation was significantly higher than that of Voltaren, suggesting a shorter onset of action and thus faster pain relief. In addition, in vivo studies have shown that the developed tablets effectively mask the bitter taste and aftertaste as well as provide good mouthfeel. The fact that the tablets are small and contain the correct dose of the drug also speaks to their superiority over the marketed formulation.

Yıldız et al. (65) also used Eudragit E (6 or 8 % m/m) to prepare taste-masked mirtazapine granules by coacervation, which is considered an economical approach. They also produced orodispersible tablets from the multiparticulates using some additional excipients to ensure the possibility of direct compression, which is also an economical method. The mass of the tablets produced was somewhat large (> 500 mg), which is not desirable for orodispersible formulations. However, the disintegration times were less than 30 seconds, so a large mass was not an issue. The dissolution profiles at three different pH values (1.2, 4.5, and 6.8) were compared with the marketed formulation Remeron SolTab and showed almost no differences. However, taste masking studies in SSF showed that the marketed formulation had lower dissolution and thus more efficient taste masking than all developed formulations, regardless of the coating ratio. The authors concluded that a higher coating ratio would be required to fully mask the taste of mirtazapine.

Only a few studies used other types of polymers to cover the taste of poorly water-soluble APIs. Al-Omran et al. (63) attempted to mask the taste of diclofenac sodium without reducing its release rate by incorporating it into microcapsules. Their core consisted of the API, lactose, and microcrystalline cellulose, while their wall was based on ethylcellulose. Diethyl phthalate or PEG 600 in various ratios were used as plasticisers in the wall, and magnesium stearate was used to prevent aggregation of the particles. Organic phase separation with toluene as solvent and petroleum ether as non-solvent was used to formulate the microcapsules. It should be noted that toluene in particular is toxic and its use for pharmaceutical products is therefore questionable. Dissolution experiments in phosphate buffer with a pH of 7.4 showed that the release rate depends on the type and concentration of plasticizer, the shape of the core material, the core size, and the core-to-wall ratio. They also compared their formulation to a marketed product Voltaren 25 mg and found that the release rate, as well as the amount of drug released, was higher for a formulation with a core-to-wall ratio of 1:1 and 20 % ( m/m) PEG 600 as a plasticiser, and in formulation with a core-to-wall ratio of 2:1 and 20 % ( m/m) diethyl phthalate as a plasticiser. However, in vivo tests to evaluate the palatability and taste of the formulations showed the superiority of the formulations with a higher wall ratio, making the former the better formulation. Since the article does not provide numerical values or comparisons to evaluate taste and palatability, it is difficult to draw further conclusions.

Chen et al. (74) dealt with taste masking and enhanced absorption of azithromycin, which is poorly soluble in the alkaline but well soluble in the acidic pH. The pellets loaded with the API were therefore coated with a mixture of two Eudragit-type polymers, RL30D and L30D-55. As mentioned earlier, the first has pH-independent swelling properties and is insoluble in water, while the second dissolves only in media with a pH above 5.5. Triethyl citrate was added to the coating as a plasticizer. The coating was applied in a fluidized bed processor. The dissolution test at a pH of 1.2 showed that the entire API was released within 120 minutes. However, after 40 minutes, the drug release was less than 20 %. Since it is known that the residence time of the pellets in the stomach may be shorter, this result might not be very promising. In addition, the dissolution study was conducted with 100 mg of API, while the recommended dose is either 500 mg or 250 mg, which also limits the relevance of these experiments.

In contrast to the previously described studies, Kharb et al. (97) did not use a polymeric substance to mask the taste of a bitter-tasting substance. Instead, Geleol (a mixture of glycerol monostearate and glyceryl stearate) was used to prepare lipid-matrix pellets with ondansetron hydrochloride dihydrate as the API. Several formulations with different ratios of Geleol were tested, but only one had drug release in phosphate buffer at pH 6.8 below the bitterness threshold concentration, indicating successful taste masking. Dissolution tests at pH 1.2 showed that drug release from this formulation was significantly slower than from pure ondansetron (< 40 % vs. 100 % after 120 minutes). To improve the slow drug release, fumed silica (Cab-o-sil) was added to the formulation to ensure faster drug release due to its hydrophilicity and gelling properties. After 120 minutes, almost 70 % of the API from the batch with 3 % Cab-o-sil had been released, so an improvement in the release of the API was evident, although it did not reach the release rate of the pure API.

Obaidat et al. (99) also prepared a taste-masked ibuprofen solid dispersion by incorporating it into a lipid substance, specifically cocoa butter microparticles, using supercritical fluid technology, a technique rarely used in taste masking research. The prepared ratios of ibuprofen were 10, 20 and 30 % ( m/m). For all samples, PXRD measurements showed a decrease in the crystallinity of the drug, while DSC showed complete miscibility in the molten carrier. However, the authors suggest that some of the drug was precipitated in crystalline form due to the weak interactions between the carrier and the drug. Dissolution experiments in phosphate buffer with a pH of 7.2 interestingly showed a significantly lower drug release in the first 20 minutes compared to the drug release of pure ibuprofen, despite the partially amorphous state of the samples. The authors suggest that this may be due to the hydrophobic nature of cocoa butter. However, after 45 minutes, the 10 and 20 % samples showed no statistically significant difference in drug release compared to pure ibuprofen or physical mixtures. However, it is possible that drug release from such formulations is faster in vivo due to enzymatic activity and surface-active agents that contribute to the degradation of lipid components. Since the amorphous state is prone to physical instability, the effect of storage conditions on the physical state and dissolution profile was also tested. Refrigerator storage appeared to provide satisfactory stability of the samples, while room temperature storage resulted in an increase in the crystallinity of the samples. Since the storage time was only 6 weeks, the prepared formulations did not seem to exhibit high physical stability.

Khor et al. (102) attempted to mask the taste of plant-derived flavonoid quercetin with three different water-insoluble excipients: shellac (natural resin secreted by the insect Kerria lacca composed of polyhydroxy polycarboxylic esters and single esters, carnauba wax (Brazilian palm tree wax composed mainly of esters), and zein (water-insoluble prolamine isolated from corn gluten). They use hot-melt extrusion to formulate microcapsules with varying ratios of excipient to API and 20 % ( m/m) water as plasticisers. Taste masking experiments performed in vitro in SSF with pH 6.8 and with an electronic tongue system showed that both carnauba wax and zein decreased the dissolution of quercetin, while shellac actually increased it. This was explained by the high solubility of shellac at pH 6.8 and/or the lower degree of quercetin crystallinity. This may even suggest that shellac was not the optimal choice for taste masking in the first place. Electronic tongue tests also confirmed that the other two excipients performed better in taste masking, but unlike the dissolution tests, shellac performed better than pure quercetin. This shows that the taste masking methods are not absolute and that it is advisable to perform more tests to obtain more meaningful results. The in vitro dissolution test, conducted at a pH of 1.0 (the first 120 minutes) and a pH of 6.8 (120–240 minutes), showed that the formulation containing 60 % shellac had the highest percentage of released quercetin after 240 minutes. However, since it was found to be ineffective in taste masking, the formulation with 70 % carnauba wax might be the better one, as the percentage of released API was similar to that of pure quercetin and its dissolution rate at pH 1.0 was also the highest among the tested formulations.

From the research reviewed in this section, it can be seen that the simultaneous implementation of techniques to improve the dissolution of poorly water-soluble drugs and to conceal their taste is rare. In fact, only 3 studies (55, 94, 99) used techniques such as preparation of solid dispersions and/or amorphization to improve dissolution and also showed that it is improved compared to pure API. There is some research (63, 65, 73) that compared their formulation to the marketed product and showed either equivalent or improved drug release properties, which is also a relevant comparison since it at least ensures that drug release is sufficient to achieve the therapeutic effect. On the other hand, in some research (74, 97), drug release from the prepared formulation was worsened compared to pure API due to taste masking approaches, while Stange et al. (71) do not show results that would deal with drug release in detail. Moreover, a significant lack of in vivo results was noticed; not only regarding taste masking, but also regarding drug absorption. In fact, these results would probably give the best indication as to whether taste masking of poorly water-soluble drugs is truly feasible.

CONCLUSIONS

The increasing use of multiparticulate formulations definitely represents a step forward in the development of child-friendly medications, as they ensure precise dosing and are well tolerated by children of all ages. Although much research has been done to ensure taste masking of bitter-tasting substances, this remains a challenging task as there is no optimal approach and no universal optimal method. Taste masking by film coating appears to be the preferred approach for multiparticulate as well as for other drug delivery systems, especially with polymers that are insoluble in saliva but dissolve in gastric fluid. In fact, this approach also appears to be the best when the drug is poorly water-soluble, as it is least likely to interfere with API dissolution after it leaves the oral cavity. However, because masking the taste of poorly water-soluble APIs is a major challenge, there are not many studies addressing this issue. In addition, it is often not clear how much the dissolution profile of such APIs is compromised by taste-masking excipients. In vivo approaches to assess taste masking are still considered the gold standard, but this is ethically questionable, particularly for the paediatric population. A combination of in vitro methods, i.e. dissolution in SSF and e-tongue, might thus be more feasible for paediatric medicine. Future research and development should therefore deal with evaluating their formulations regarding taste masking and drug release with methods that would give more relevant and meaningful results. Furthermore, it should put a greater emphasis on utilizing techniques to improve water solubility when dealing with poorly water-soluble drugs so as to avoid compromising sufficient drug release with taste masking. It is expected that taste-masked multiparticulate formulations will continue to be the subject of research and that more such formulations will be marketed in the future.

Availability of data and materials. – The datasets generated during and/or analysed during the current study are available from the corresponding author upon reasonable request.

Funding. – This research was co-funded by the Slovenian Research Agency (research core funding P1-0189). The Slovenian Research Agency had no role in the design and conduct of the study; analysis and interpretation of data; preparation, approval and submission of the manuscript for publication.

Conflict of interest. – The authors declare that they have no competing interests.

Authors contributions. – All authors contributed to the study conception, design and contributed to the data collection. A. Baumgartner interpreted the data and drafted the manuscript. All authors have read and approved the entire manuscript.

References

1 

R. H. Richey, U. U. Shah, M. Peak, J. V. Craig, J. L. Ford, C. E. Barker, A. J. Nunn and M. A. Turner, Manipulation of drugs to achieve the required dose is intrinsic to paediatric practice but is not supported by guidelines or evidence,. BMC Pediatr. 13: 2013 Article ID 81 (8 pages);. https://doi.org/10.1186/1471-2431-13-81

2 

M. Preis, Orally disintegrating films and mini-tablets-innovative dosage forms of choice for pediatric use,. AAPS PharmSciTech. 16: 2015 234–241. https://doi.org/10.1208/s12249-015-0313-1

3 

V. M. Vaz and L. Kumar, 3D printing as a promising tool in personalized medicine,. AAPS PharmSciTech. 22: 2021 Article ID 49 (20 pages);. https://doi.org/10.1208/s12249-020-01905-8

4 

C. Watson, E. A. Webb, S. Kerr, J. H. Davies, H. Stirling and H. Batchelor, How close is the dose? Manipulation of 10 mg hydrocortisone tablets to provide appropriate doses to children,. Int. J. Pharm. 545: 2018 57–63. https://doi.org/10.1016/j.ijpharm.2018.04.054

5 

C. Verrue, E. Mehuys, K. Boussery, J.-P. Remon and M. Petrovic, Tablet-splitting: a common yet not so innocent practice,. J. Adv. Nurs. 67: 2011 26–32. https://doi.org/10.1111/j.1365-2648.2010.05477.x

6 

T. B. Ernest, J. Craig, A. Nunn, S. Salunke, C. Tuleu, J. Breitkreutz, R. Alex and J. Hempenstall, Preparation of medicines for children – a hierarchy of classification,. Int. J. Pharm. 435: 2012 124–130. https://doi.org/10.1016/j.ijpharm.2012.05.070

7 

C. Litalien, S. Bérubé, C. Tuleu, A. Gilpin, É. K. Landry, M. Valentin, R. Strickley and M. A. Turner, From paediatric formulations development to access: Advances made and remaining challenges,. Br. J. Clin. Pharmacol. 88: 2022 4349–4383. https://doi.org/10.1111/bcp.15293

8 

V. Bhatt-Mehta, R. B. MacArthur, R. Löbenberg, J. J. Cies, I. Cernak and R. H. P. Ii, An algorithm to identify compounded non-sterile products that can be formulated on a commercial scale or imported to promote safer medication use in children,. Pharm. Basel Switz. 3: 2015 284–294. https://doi.org/10.3390/pharmacy3040284

9 

E. Rawlence, A. Lowey, S. Tomlin and V. Auyeung, Is the provision of paediatric oral liquid unlicensed medicines safe?, Arch.Dis. Child. - Educ. Amp Pract. Ed. , editor. 103: 2018 Article ID 310 (5 pages);. https://doi.org/10.1136/archdischild-2016-312132

10 

J. M. Rood, M. J. Engels, S. L. Ciarkowski, L. D. Wagenknecht, C. J. Dickinson and J. G. Stevenson, Variability in compounding of oral liquids for pediatric patients: a patient safety concern,. J. Am. Pharm. Assoc. JAPhA. 54: 2014 383–389. https://doi.org/10.1331/JAPhA.2014.13074

11 

D. A. van Riet-Nales, E. G. A. W. Römkens, A. Saint-Raymond, P. Kozarewicz, A. F. A. M. Schobben, T. C. G. Egberts and C. M. A. Rademaker, Oral medicines for children in the European paediatric investigation plans,. PloS One. 9: 2014 983488:https://doi.org/10.1371/journal.pone.0098348

12 

E. Finney, Make medicine child size, Children’s medicines: A situational analysis,. 2011 https://www.who.int/childmedicines/progress/CM_analysis.pdf;last access. 2023

13 

European Medicines Agency, Guideline on pharmaceutical development of medicines for paediatric use,. 2013 https://www.ema.europa.eu/en/documents/scientific-guideline/guideline-pharmaceutical-development-medicines-paediatric-use_en.pdf;last access. 2023

14 

J. Zisowsky, A. Krause and J. Dingemanse, Drug development for pediatric populations: Regulatory aspects,. Pharmaceutics. 2: 2010 364–388. https://doi.org/10.3390/pharmaceutics2040364

15 

European Commission, State of Paediatric Medicines in the EU: 10 years of the EU Paediatric Regulation,. 2017 https://health.ec.europa.eu/system/files/2017-11/2017_childrensmedicines_report_en_0.pdf;last access. 2023

16 

M. Moreira and M. Sarraguça, How can oral paediatric formulations be improved? A challenge for the XXI century,. Int. J. Pharm. 590: 2020 Article ID 119905 (9 pages);. https://doi.org/10.1016/j.ijpharm.2020.119905

17 

K. A. Mfoafo, M. Omidian, C. D. Bertol, Y. Omidi and H. Omidian, Neonatal and pediatric oral drug delivery: Hopes and hurdles,. Int. J. Pharm. 597: 2021 120296https://doi.org/10.1016/j.ijpharm.2021.120296

18 

R. G. Strickley, Q. Iwata, S. Wu and T. C. Dahl, Pediatric drugs – a review of commercially available oral formulations,. J. Pharm. Sci. 97: 2008 1731–1774. https://doi.org/10.1002/jps.21101

19 

Batchelor H, Paediatric Development: Anatomy. Age, Weight, Body Surface and Stature, Organ Development,. Springer,; New York: 2014 p. 3–7

20 

F. Lopez, P. Mistry, H. Batchelor, J. Bennett, A. Coupe, T. Ernest, M. Orlu and C. Tuleu, Acceptability of placebo multiparticulate formulations in children and adults, Sci. Rep. 8: 2018 Article ID 9210 (10 pages);. https://doi.org/10.1038/s41598-018-27446-6

21 

R. G. Strickley, Pediatric oral formulations: An updated review of commercially available pediatric oral formulations since 2007,. J. Pharm. Sci. 108: 2019 1335–1365. https://doi.org/10.1016/j.xphs.2018.11.013

22 

T. Simšič, B. Nolimal, J. Minova, A. Baumgartner and O. Planinšek, A straw for paediatrics: How to administer highly dosed, bitter tasting paracetamol granules,. Int. J. Pharm. 602: 2021 Article ID 120615 (11 pages);. https://doi.org/10.1016/j.ijpharm.2021.120615

23 

A. Hejduk and J. Lulek, Dispensing of minitablets – Has the problem been resolved?,. Int. J. Pharm. 619: 2022 Article ID 121666;. https://doi.org/10.1016/j.ijpharm.2022.121666

24 

Overview of taste masking excipients & technologies for solid & liquid oral pharmaceuticals offered by leading contract manufacturers (CDMOs, CMOs);. https://www.pharmacompass.com/pharma-blog/overview-of-taste-masking-excipients-technologies-for-solid-liquid-oral-pharmaceuticals-offered-by-leading-contract-manufacturers-cdmos-cmos;last access. 2023

25 

H. K. Batchelor and J. F. Marriott, Formulations for children: problems and solutions,. Br. J. Clin. Pharmacol. 79: 2015 405–418. https://doi.org/10.1111/bcp.12268

26 

Sympfiny^(®): Delivery device for multiparticulate drugs;. https://www.cphi-online.com/sympfiny-delivery-device-for-multiparticulate-prod500849.html;last access. 2023

27 

Unique minitablet dispenser meets flexible oral dosing need; https://www.ondrugdelivery.com/unique-minitablet-dispenser-meets-flexible-oral-dosing-need/; last access. 2023

28 

IQ-Dose for micro-tablets dosage wins Award;. http://www.industry24h.com/iq-dose-for-micro-tablets-dosage-wins-award/;last access. 2023

30 

K. Wening, E. J. Laukamp, M. Thommes and J. Breitkreutz, Individual oral therapy with immediate release and effervescent formulations delivered by the solid dosage pen,. J. Pers. Med. 2: 2012 217–231. https://doi.org/10.3390/jpm2040217

31 

J. Schomakers, A. Grummel, Dosing Stick Containing Rod-shaped Tablets, U.S. Pat. 01782181: 2004

32 

J. Walsh, A. Cram, K. Woertz, J. Breitkreutz, G. Winzenburg, R. Turner and C. Tuleu, Playing hide and seek with poorly tasting paediatric medicines: do not forget the excipients, Adv. Drug Deliv. Rev. 73: 2014 14–33. https://doi.org/10.1016/j.addr.2014.02.012

33 

W. Faisal, F. Farag, A. A. H. Abdellatif and A. Abbas, Taste masking approaches for medicines, Curr. Drug Deliv. 15: 2018 167–185. https://doi.org/10.2174/1567201814666171013145958

34 

S. Joshi and H.-U. Petereit, Film coatings for taste masking and moisture protection,. Int. J. Pharm. 457: 2013 395–406. https://doi.org/10.1016/j.ijpharm.2013.10.021

35 

X. Zheng, F. Wu, Y. Hong, L. Shen, X. Lin and Y. Feng, Developments in taste-masking techniques for traditional chinese medicines,. Pharmaceutics. 10: 2018 Article. 157:https://doi.org/10.3390/pharmaceutics10030157

36 

K. Al-Japairai, S. Hamed Almurisi, A. Almonem Doolaanea, S. Mahmood, F. Alheibshy, A. Alobaida, N. Abdul-Halim and B. Chatterjee, A review on taste masked multiparticulate dosage forms for paediatric,. Int. J. Pharm. 632: 2022 Article ID 122571;. https://doi.org/10.1016/j.ijpharm.2022.122571

37 

H. Sohi, Y. Sultana and R. K. Khar, Taste masking technologies in oral pharmaceuticals: Recent developments and approaches, Drug Dev. Ind. Pharm. 30: 2004 429–448. https://doi.org/10.1081/DDC-120037477

38 

D. Andrews, S. Salunke, A. Cram, J. Bennett, R. S. Ives, A. W. Basit and C. Tuleu, Bitter-blockers as a taste masking strategy: A systematic review towards their utility in pharmaceuticals,. Eur. J. Pharm. Biopharm. 158: 2021 35–51. https://doi.org/10.1016/j.ejpb.2020.10.017

39 

L. A. Felton, Use of polymers for taste-masking pediatric drug products, Drug Dev. Ind. Pharm. 44: 2018 1049–1055. https://doi.org/10.1080/03639045.2018.1430822

40 

M. Guhmann, M. Preis, F. Gerber, N. Pöllinger, J. Breitkreutz and W. Weitschies, Design, development and in-vitro evaluation of diclofenac taste-masked orodispersible tablet formulations, Drug Dev. Ind. Pharm. 41: 2015 540–551. https://doi.org/10.3109/03639045.2014.884122

41 

Y. Nakano, A. Maeda, S. Uchida and N. Namiki, Preparation and evaluation of unpleasant taste-masked pioglitazone orally disintegrating tablets,. Int. J. Pharm. 446: 2013 160–165. https://doi.org/10.1016/j.ijpharm.2013.02.019

42 

M. Drašković, D. Medarević, I. Aleksić and J. Parojčić, In vitro and in vivo investigation of taste-masking effectiveness of Eudragit E PO as drug particle coating agent in orally disintegrating tablets, Drug Dev. Ind. Pharm. 43: 2017 723–731. https://doi.org/10.1080/03639045.2016.1220572

43 

D. Douroumis, Practical approaches of taste masking technologies in oral solid forms, Expert Opin. Drug Deliv. 4: 2007 417–426. https://doi.org/10.1517/17425247.4.4.417

44 

V. Jannin and Y. Cuppok, Hot-melt coating with lipid excipients, Prog. Film Coat. 457: 2013 480–487. https://doi.org/10.1016/j.ijpharm.2012.10.026

45 

SPI Pharma, Actimask^(®): Taste Masked Actives,. 2017 https://www.spipharma.com/media/2860/actimask-psb-oct-2017.pdf;last access. 2023

46 

Z. Worku, V. Puri, L. Kumar and A. Bansal, Trends in pharmaceutical taste masking technologies: A patent review, Recent Pat. Drug Deliv. Formul. 3: 2009 26–39. https://doi.org/10.2174/187221109787158364

47 

Evonik Industries, EUDRAGIT: Application Guidelines,; 13th ed,. Evonik Nutrition & Care GmbH,; Germany: 2019

48 

European Medicines Agency, Cyclodextrins used as excipients,. 2017 https://www.ema.europa.eu/en/documents/scientific-guideline/questions-answers-cyclodextrins-used-excipients-medicinal-products-human-use_en.pdf;last access date February 3, 2023.

49 

M. S. S. Cunha-Filho, B. Dacunha-Marinho, J. J. Torres-Labandeira, R. Martínez-Pacheco and M. Landin, Characterization of β-lapachone and methylated β-cyclodextrin solid-state systems,. AAPS PharmSciTech. 8(3) 2007 Article ID 60 (10 pages);. https://doi.org/10.1208/pt0803060

50 

N. E. Nasr, A. N. ElMeshad and A. R. Fares, Nanocarrier systems in taste masking, Sci. Pharm. 90: 2022 Artile ID 20 (14 pages);. https://doi.org/10.3390/scipharm90010020

51 

B. Al-kasmi, MHD. B. Alsirawan, M. Bashimam and H. El-zein, Mechanical microencapsulation: The best technique in taste masking for the manufacturing scale – Effect of polymer encapsulation on drug targeting,. J. Controll. Release. 260: 2017 134–141. https://doi.org/10.1016/j.jconrel.2017.06.002

52 

D. Bar-Shalom, Necessity of rethinking oral pediatric formulations, Clin. Ther. 36: 2014 180–183. https://doi.org/10.1016/j.clinthera.2014.01.010

53 

M. Ghimire, Z. N. Mahmoudi, J. Teckoe and A. Rajabi-Siahboomi, Application of a pH dependent taste-mask film coating for pediatric multiparticulate formulations,. Int. J. Pharm. 536: 2018 490–522. https://doi.org/10.1016/j.ijpharm.2017.08.021

54 

M. Jelvehgari, L. Barghi and F. Barghi, Preparation of chlorpheniramine maleate-loaded alginate/chitosan particulate systems by the ionic gelation method for taste masking,. Jundishapur J. Nat. Pharm. Prod. 9: 2014 39–48. https://doi.org/10.17795/jjnpp-12530

55 

M. Münster, C. Schoch, C. Schmidt and J. Breitkreutz, Multiparticulate system combining taste masking and immediate release properties for the aversive compound praziquantel,. Eur. J. Pharm. Sci. 109: 2017 446–454. https://doi.org/10.1016/j.ejps.2017.08.034

56 

E. Combescot, G. Morat, P. de Lonlay and V. Boudy, Pediatric drug formulation of sodium benzoate extended-release granules,. Pharm. Dev. Technol. 21: 2016 261–267. https://doi.org/10.3109/10837450.2014.996898

57 

T. Yoshida, H. Tasaki, A. Maeda, M. Katsuma, K. Sako and T. Uchida, Salting-out taste-masking system generates lag time with subsequent immediate release,. Int. J. Pharm. 365: 2009 81–88. https://doi.org/10.1016/j.ijpharm.2008.08.026

58 

B. Albertini, C. Cavallari, N. Passerini, D. Voinovich, M. L. González-Rodrıguez, L. Magarotto and L. Rodriguez, Characterization and taste-masking evaluation of acetaminophen granules: comparison between different preparation methods in a high-shear mixer,. Eur. J. Pharm. Sci. 21: 2004 295–303. https://doi.org/10.1016/j.ejps.2003.10.017

59 

D. D. Douroumis, A. Gryczke and S. Schminke, Development and evaluation of cetirizine HCl taste-masked oral disintegrating tablets,. AAPS PharmSciTech. 12: 2011 141–151. https://doi.org/10.1208/s12249-010-9569-7

60 

S. L. Cantor, M. A. Khan and A. Gupta, Development and optimization of taste-masked orally disintegrating tablets (ODTs) of clindamycin hydrochloride, Drug Dev. Ind. Pharm. 41: 2015 1156–1164. https://doi.org/10.3109/03639045.2014.935392

61 

A. Alayoubi, B. Daihom, H. Adhikari, S. Mishra, R. Helms and H. Almoazen, Development of a taste-masked oral suspension of clindamycin HCl using ion exchange resin Amberlite IRP 69 for use in pediatrics, Drug Dev. Ind. Pharm. 42: 2016 1579–1589. https://doi.org/10.3109/03639045.2016.1160102

62 

H. Lou, M. Liu, L. Wang, S. R. Mishra, W. Qu, J. Johnson, E. Brunson and H. Almoazen, Development of a mini-tablet of co-grinded prednisone-Neusilin complex for pediatric use,. AAPS PharmSciTech. 14: 2013 950–958. https://doi.org/10.1208/s12249-013-9981-x

63 

M. F. Al-Omran, S. A. Al-Suwayeh, A. M. El-Helw and S. I. Saleh, Taste masking of diclofenac sodium using microencapsulation,. J. Microencapsul. 19: 2002 45–52. https://doi.org/10.1080/02652040110055612

64 

P. S. Sona and C. Muthulingam, Formulation and evaluation of taste masked orally disitegrating tablets of diclofenac sodium,. Int. J. PharmTech Res. 3: 2011 819–826. https://doi.org/10.1081/DDC-120037477

65 

S. Yıldız, E. Aytekin, B. Yavuz, S. Bozdağ Pehlivan, İ. Vural and N. Ünlü, Development and evaluation of orally disintegrating tablets comprising taste-masked mirtazapine granules,. Pharm. Dev. Technol. 23: 2018 488–495. https://doi.org/10.1080/10837450.2017.1315670

66 

P. Shrestha, R. Shrestha and S. Shrestha, Formulation and evaluation of orally disintegrating tablet containing taste masked mirtazapine,. Pharm. Res. 10(2) 2021 71–80. https://doi.org/10.15406/japlr.2021.10.00368

67 

B. S. Silva, M. Santangelo, M.-J. Colbert, C. Fauteux-Lefebvre, J. A. Bartlett, P.-P. Lapointe-Garant and R. Gosselin, Building process understanding of fluid bed taste mask coating of microspheres,. AAPS PharmSciTech. 20: 2019 Article ID 173 (14 pages);. https://doi.org/10.1208/s12249-019-1384-1

68 

P. C. Kayumba, N. Huyghebaert, C. Cordella, J. D. Ntawukuliryayo, C. Vervaet and J. P. Remon, Quinine sulphate pellets for flexible pediatric drug dosing: Formulation development and evaluation of taste-masking efficiency using the electronic tongue,. Eur. J. Pharm. Biopharm. 66: 2007 460–465. https://doi.org/10.1016/j.ejpb.2006.11.018

69 

V. Patravale and N. B. Prabhu, Taste masking of quinine sulphate,. Ind. J. Pharm. Sci. 67: 2005 233–235

70 

A. Keeley, M. Teo, Z. Ali, J. Frost, M. Ghimire, A. Rajabi-Siahboomi, M. Orlu and C. Tuleu, In vitro dissolution model can predict the in vivo taste masking performance of coated multiparticulates, Mol. Pharm. 16: 2019 2095–2105. https://doi.org/10.1021/acs.molpharmaceut.9b00060

71 

U. Stange, C. Führling and H. Gieseler, Taste masking of naproxen sodium granules by fluid-bed coating,. Pharm. Dev. Technol. 19: 2014 137–147. https://doi.org/10.3109/10837450.2012.757784

72 

X. Hu, Y. Li, E. Zhang, X. Wang, M. Xing, Q. Wang, J. Lei and H. Huang, Preparation and evaluation of orally disintegrating tablets containing taste-masked microcapsules of berberine hydrochloride,. AAPS PharmSciTech. 14: 2013 29–37. https://doi.org/10.1208/s12249-012-9880-6

73 

H. F. Alotaibi, S. Elsamaligy, G. M. Mahrous, M. A. Bayomi and H. A. Mahmoud, Design of taste masked enteric orodispersible tablets of diclofenac sodium by applying fluid bed coating technology,. Saudi Pharm. J. 27: 2019 354–362. https://doi.org/10.1016/j.jsps.2018.12.003

74 

Y. Chen, Y. Liu, C. Wu, X. Pan and T. Peng, Dry suspension containing coated pellets with pH-dependent drug release behavior for the taste-masking of azithromycin,. AAPS PharmSciTech. 24: 2022 Article ID 21;. https://doi.org/10.1208/s12249-022-02484-6

75 

R. Huang, Y. Zhang, T. Wang, L. Shen, Z. Zhang, Y. Wang and D. Quan, Creation of an assessment system for measuring the bitterness of azithromycin-containing reverse micelles,. Asian J. Pharm. Sci. 13: 2018 343–352. https://doi.org/10.1016/j.ajps.2018.02.001

76 

N. Choudhary and J. Avari, Formulation and evaluation of taste mask pellets of granisetron hydrochloride as oro dispersible tablet,. Braz. J. Pharm. Sci. 51: 2015 569–578. https://doi.org/10.1590/S1984-82502015000300009

77 

T. Hamashita, M. Matsuzaki, T. Ono, M. Ono, Y. Tsunenari, T. Aketo and S. Watano, Granulation of core particles suitable for film coating by agitation fluidized bed ii. A proposal of a rapid dissolution test for evaluation of bitter taste of ibuprofen,. Chem. Pharm. Bull. (Tokyo). 56: 2008 883–887. https://doi.org/10.1248/cpb.56.883

78 

T. Hamashita, Y. Nakagawa, T. Aketo and S. Watano, Granulation of core particles suitable for film coating by agitation fluidized bed I. Optimum formulation for core particles and development of a novel friability test method,. Chem. Pharm. Bull. (Tokyo). 55: 2007 1169–1174. https://doi.org/10.1248/cpb.55.1169

79 

BASF, Kollicoat^(®) Smartseal 100 P: Technical Information,. 2019 https://www.pharmaexcipients.com/product/kollicoat-smartseal-100-p/;last access. 2023

80 

BASF, Kollicoat^(®) Smartseal 30 D:. Technical Information. 2019 https://www.pharmaexcipients.com/product/kollicoat-smartseal-30-d/;last access. 2023

81 

Eudragit^(®) E 100, Eudragit^(®). E PO and Eudragit^(®) E. 125:https://www.pharmaexcipients.com/news/eudragit-e-100-12-5-po/;last access. 2023

82 

84 

Evonik EUDRAGIT^(®) RL 30 D Copolymer;. https://www.matweb.com/search/datasheettext.aspx?matguid=97322e67e21c46f4950bd8b64f71c82d;last access. 2023

85 

H. Lim and S. W. Hoag, Plasticizer effects on physical-mechanical properties of solvent cast Soluplus^(®) films,. AAPS PharmSciTech. 14: 2013 903–910. https://doi.org/10.1208/s12249-013-9971-z

86 

A. Roy, A. Ghosh, S. Datta, S. Das, P. Mohanraj, J. Deb and M. E. Bhanoji Rao, Effects of plasticizers and surfactants on the film forming properties of hydroxypropyl methylcellulose for the coating of diclofenac sodium tablets,. Saudi Pharm. J. SPJ Off. Publ. Saudi Pharm. Soc. 17: 2009 233–241. https://doi.org/10.1016/j.jsps.2009.08.004

87 

L. A. Felton and J. W. McGinity, Influence of insoluble excipients on film coating systems, Drug Dev. Ind. Pharm. 28: 2002 225–243. https://doi.org/10.1081/DDC-120002839

88 

M. Pein, M. Preis, C. Eckert and F. E. Kiene, Taste-masking assessment of solid oral dosage forms – a critical review,. Int. J. Pharm. 465: 2014 239–254. https://doi.org/10.1016/j.ijpharm.2014.01.036

89 

S. Gittings, N. Turnbull, C. J. Roberts and P. Gershkovich, Dissolution methodology for taste masked oral dosage forms,. J. Control. Release. 173: 2014 32–42. https://doi.org/10.1016/j.jconrel.2013.10.030

90 

Y. Thabet, M. Slavkova and J. Breitkreutz, 10 years EU regulation of pediatric medicines – impact on cardiovascular drug formulations,. Expert Opin. Drug Deliv. 15: 2018 261–270. https://doi.org/10.1080/17425247.2018.1420054

91 

E. H. Davies and C. Tuleu, Medicines for children: A matter of taste,. J. Pediatr. 153: 2008 599–6042. https://doi.org/10.1016/j.jpeds.2008.06.030

92 

B. Al-Kasmi, O. Al Rahal, H. El-Zein and A.-H. Nattouf, Structural and in vitro in vivo evaluation for taste masking,. Expert Opin. Drug Deliv. 15: 2018 1105–1116. https://doi.org/10.1080/17425247.2018.1535590

93 

G. F. Petrovick, J. Breitkreutz and M. Pein-Hackelbusch, Taste-masking properties of solid lipid based micropellets obtained by cold extrusion-spheronization,. Int. J. Pharm. 506: 2016 361–370. https://doi.org/10.1016/j.ijpharm.2016.04.058

94 

A. S. Alshetaili, B. K. Almutairy, R. V. Tiwari, J. T. Morott, S. M. Alshehri, X. Feng, B. B. Alsulays, J.-B. Park, F. Zhang and M. A. Repka, Preparation and evaluation of hot-melt extruded patient-centric ketoprofen mini-tablets, Curr. Drug Deliv. 13: 2016 730–741. https://doi.org/10.2174/1567201812666151012113806

95 

G. F. Petrovick, P. Kleinebudde and J. Breitkreutz, Orodispersible tablets containing taste-masked solid lipid pellets with metformin hydrochloride: Influence of process parameters on tablet properties,. Eur. J. Pharm. Biopharm. 122: 2018 137–145. https://doi.org/10.1016/j.ejpb.2017.10.018

96 

C. Eckert, M. Pein and J. Breitkreutz, Lean production of taste improved lipidic sodium benzoate formulations,. Eur. J. Pharm. Biopharm. 88: 2014 455–461. https://doi.org/10.1016/j.ejpb.2014.05.013

97 

V. Kharb, V. A. Saharan, V. Kharb, H. Jadhav and S. Purohit, Formulation and evaluation of lipid based taste masked granules of ondansetron HCl,. Eur. J. Pharm. Sci. 62: 2014 180–188. https://doi.org/10.1016/j.ejps.2014.05.012

98 

D. G. Lopes, V. Garsuch, K. Becker, A. Paudel, M. Stehr, A. Zimmer and S. Salar-Behzadi, Improving the granule strength of roller-compacted ibuprofen sodium for hot-melt coating processing,. Int. J. Pharm. 510: 2016 285–295. https://doi.org/10.1016/j.ijpharm.2016.06.049

99 

R. Obaidat, H. Aleih, H. Mashaqbeh, B. Altaani, M. M. Alsmadi and M. Alnaief, Development and evaluation of cocoa butter taste masked ibuprofen using supercritical carbon dioxide,. AAPS PharmSciTech. 22: 2021 Article ID 106;. https://doi.org/10.1208/s12249-021-01962-7

100 

Council of Europe, European Pharmacopoeia. 10th Edition. Strasbourg: 2019

101 

M. Siewert, J. Dressman, C. K. Brown and V. P. Shah, FIP/AAPS guidelines to dissolution/in vitro release testing of novel/special dosage forms,. AAPS PharmSciTech. 4: 2003 Article ID E7;. https://doi.org/10.1208/pt040107

102 

C. M. Khor, W. K. Ng, P. Kanaujia, K. P. Chan and Y. Dong, Hot-melt extrusion microencapsulation of quercetin for taste-masking,. J. Microencapsul. 34: 2017 29–37. https://doi.org/10.1080/02652048.2017.1280095


This display is generated from NISO JATS XML with jats-html.xsl. The XSLT engine is libxslt.