Open access peer-reviewed chapter - ONLINE FIRST

Nanoemulsions: A Recent Drug Delivery Tool

Written By

Vaibhav Changediya

Submitted: 04 March 2024 Reviewed: 04 March 2024 Published: 03 June 2024

DOI: 10.5772/intechopen.1005266

Nanoemulsions - Design and Applications IntechOpen
Nanoemulsions - Design and Applications Edited by Juan Mejuto

From the Edited Volume

Nanoemulsions - Design and Applications [Working Title]

Prof. Juan C. Mejuto and Dr. Mihalj Poša

Chapter metrics overview

10 Chapter Downloads

View Full Metrics

Abstract

The use of nano/sub-micron particles in food, cosmetic, and pharmaceutical technology is becoming more and more popular. In particular, this interest has been growing in tandem with improved stabilization and emulsification methods. High-energy and low-energy spontaneous emulsification techniques are the two primary categories of nanoemulsion preparation techniques. Stability ranging from a few hours to years is influenced by important characteristics related to preparation procedures and components used. Nanoemulsions do not worry about issues like creaming, coalescence sedimentation, and flocculation because of their small droplet size. Ostwald ripening for them is the primary destabilizing process, though. This chapter provides a thorough overview of nanoemulsions including an explanation of their preparation techniques and assessments.

Keywords

  • nanoemulsion
  • spontaneous emulsification
  • Ostwald ripening
  • creaming
  • flocculation

1. Introduction

Recently, there has been a lot of interest in lipid-based formulations as a way to increase the permeability and bioavailability of poorly soluble drugs. As a result, many innovative drug delivery strategies have been used, and in each case, the nanoemulsion is essential to the active pharmaceutical ingredient’s transport to the target organ or location. Among many different technologies, nanoemulsions have proven to advance drug delivery systems more than others.

These are considered the best alternatives for raising the oral bioavailability of drugs listed in the Biopharmaceutical Drug Classification System (BCS) under classes II and IV. A nanoemulsion is a clear solution of two nonsoluble liquids, such as water and oil, that is thermodynamically stable and stabilized by an interfacial layer of surfactant molecules. Nanoemulsions are a novel approach to medicine administration that uses emulsified water and oil systems with mean droplet sizes ranging from 50 to 1000 nm.

The main distinctions between emulsions and nanoemulsions are the size and morphology of the particles dispersed in continuous phases. The particle size in typical emulsions is 1/20 μm, but in nanoemulsions, it is 10/200 nm. A nanoemulsion is a kinetically stable liquid that consists of an oil phase and a water phase that possesses the appropriate surfactant. The dispersed phase has a low oil/water interfacial tension and is primarily composed of tiny particles with sizes ranging from 5 to 200 nm. They are regarded as the greatest alternative colloidal dispersions with exact ratios of an oil phase, an aqueous phase, a surfactant, and a co-surfactant are called nanoemulsions. To create the nanoemulsions, both high energy and low-energy emulsification procedures were applied. While low-energy emulsification methods take advantage of the system’s physicochemical properties, which exploit phase transitions to produce nanoemulsion, high-energy emulsification methods utilize high shear mixing, high/pressure homogenization, or ultrasonification. Because oil, surfactant, and co-surfactant nanoemulsions are nontoxic and nonirritating, the Food and Drug Administration (FDA) has approved them for consumption by humans as “generally recognised as safe” [1, 2].

Advertisement

2. Composition of nanoemulsion

The following are the primary elements used in the creation of nanoemulsions:

  1. Oil (for the solubilization of drugs or lipophilic molecules)

  2. Surfactant

  3. Water

  4. Co-surfactant (amplify the surfactant’s effect).

The stability of the system is greatly dependent upon and impacted by the selection of the appropriate components for nanoemulsions. Thus, choosing them carefully is essential. Surfactants and/or co-surfactants/co-solvents, along with an oily phase and an aqueous phase, make up most nanoemulsions.

2.1 Oil

Thus, oil is the most crucial ingredient in nanoemulsions. This facilitates the completion of the intended tasks. As such, it may operate as a carrier for the lipophilic active pharmaceutical ingredients or act as an active agent in and of itself (the essential oil in our study, for example). The selected oil’s characteristics, including its polarity (low polarity oils are quickly disrupted by applied external force), viscosity (low viscosity oils are disrupted by applied external energy more quickly, resulting in the creation of tiny droplets more quickly), and interfacial tension (low interfacial tension in the oil facilitates size reduction and reduces the energy required to shrink the droplet size).

2.2 Surfactants

Without surfactants, the nanoemulsion system that stabilizes the thermodynamically unstable mixing of two immiscible liquids by lowering interfacial tension and altering dispersion entropy would be incomplete. The primary criteria for the surfactants used in the production of nanoemulsions are stability, high drug loading capacity, efficacious emulsification, and safety. Rapid adsorption of an appropriate surfactant used in the nanoemulsion formulation onto the interface between the two immiscible phases is necessary to prevent the coalescence of the nanodroplets and to dramatically lower interfacial tension. The four subgroups of surfactants are cationic, nonionic, zwitter-ionic, and anionic surfactants. Out of these four types, nonionic surfactants are used the most since they are known to have high biological absorption, are less sensitive to pH variations, and have a lower toxicity and irritating profile than ionic ones. The lipids in the stratum corneum can also be fluidized and solubilized by nonionic surfactants, which enhances the skin’s capacity to absorb the applied solution.

The scale used to classify nonionic surfactants is called the hydrophilic–lipophilic balance (HLB) which is a specific empirical expression of nonionic surfactants that characterizes the relationship between the hydrophilic and hydrophobic portions of the surfactant. In 1954, William C. Gryphon developed it. The values on the HLB scale range from 0 to 20. In order to form water/oil nanoemulsions, surfactants with HLB values between 3 and 6 are lipophilic (like SPANS) sorbitan monolaurate, whereas those with HLB values between 8 and 18 are hydrophilic (like TWEENS) polyoxyethylene sorbitan monolaurate. Co-surfactants are surfactants with an HLB value greater than 20. It is not a random addition of surfactants; rather, the surface becomes saturated with surfactants, and micelles begin to form when the critical micelle concentration is reached. It is claimed that a mixture of surfactants makes the nanoemulsion system more stable. The four subgroups of surfactants are cationic, nonionic, zwitter-ionic, and anionic surfactants. Nonionic surfactants are the most often used of these four types because they are less sensitive to pH changes and have a lower toxicity and irritating profile than ionic ones. They are also well known for having a high level of biological acceptability. The stratum corneum’s lipids can also be fluidized and dissolved by nonionic surfactants.

They thereby enhance the skin’s capacity to absorb and permeate the applied solution. William C. Gryphon developed a scale in 1954 that is used to group nonionic surfactants.

2.3 Co-surfactant

Co-surfactant helps the surfactant in the nanoemulsion system emulsify the oil in an aqueous phase. By combining with surfactants and penetrating the surfactant layer, co-surfactants in these systems dissolve the interfacial film, provide the necessary fluidity, reduce interfacial tension, and aid in the emulsification process. The interfacial film is made more flexible by the addition of a co-surfactant because surfactant alone frequently fails to produce temporary negative interfacial tension and fluid interfacial film.

Co-surfactants can also aid in the solubilization of the oil by altering the oil-water interface’s curvature. Because the interaction between the surfactant and co-surfactant impacts how therapeutic compounds or lipophilic drugs partition into the aqueous and oil phases, choosing the right co-surfactant is crucial. 1,2-Propylene glycol, Transcutol® HP (diethylene glycol monoethyl ether), polyethylene glycol-400, and absolute carbohol. Alcohols may increase the miscibility of the two phases because of the way they partition between the aqueous and oily phases. Thus, ethanol, isopropyl alcohol, 1-butanol, and propylene glycol were selected as co-surfactants. Additionally, because carbosol and polyethylene glycol 400 are generally well-tolerated and show increased permeability when added to formulations, they were selected. The selection of surfactants and co-surfactants in the preparation of nanoemulsions is based on their % transmittance. Transcutol P (diethylene glycol monoethyl ether) had a higher % transmittance than propylene glycol. It was also found that transcutol P (diethylene glycol monoethyl ether) worked well as a potent permeation enhancer. It has been found that the surfactant’s performance during the emulsification process is influenced by the co-surfactant concentration. Selecting a co-surfactant involves a number of factors, one of which is assessing the phase diagram’s nanoemulsion area. The capacity of the ratio of surfactant to co-surfactant to form nanoemulsions is significantly influenced by the size of the nanoemulsion area in the phase diagram [3, 4, 5].

Advertisement

3. Methods of preparation of nanoemulsion

The stratum corneum (SC), the top layer of the epidermis, acts as a powerful barrier to the applied components, as was previously discussed. This issue makes it difficult for the formulations to penetrate the skin. Consequently, numerous solutions including micro and nanosystems were explored and developed in order to get around this problem. Adding the active substance or ingredients to a nanoemulsion carrier was one of the better ways. An emulsion with droplets ranging in size from 20 to 200 nm is called a nanoemulsion. It is a colloidal system with two phases that are incompatible with one another: an aqueous phase and an oily phase. The system’s dispersion of the oily component into the aqueous phase is referred to as an oil-in-water (oil/water) nanoemulsion.

To maintain physical stability over the long term without any flocculation or coalescence in the system. A nanoemulsion is a transparent or translucent liquid-in-liquid dispersion system that is stable both thermodynamically and kinetically. Nanoemulsions offer an advantage over regular emulsions in that they are able to better encapsulate medications in the body. The ability of the nanometer-sized droplets to impede phase separation, coalescence, and flocculation improves the dispersibility of the system. Better long-term stability follows as a consequence [6, 7].

3.1 Phase inversion method

Chemical energy from phase changes that happen during the emulsification process produces fine dispersion. The required phase transitions are provided by changes in the polymer chain or composition at constant temperature. Predicated on the notion that temperature affects the solubility of polyoxyethylene/type surfactants. The surfactant monolayer exhibits a notable positive temperature change with constant composition at low temperatures. When temperatures rise, this surfactant becomes lipophilic and produces an oil-swollen micellar solution phase due to spontaneous curvature brought on by dehydration, as shown in Figure 1.

Figure 1.

Method of preparation of nanoemulsion.

3.2 Phase inversion temperature (PIT)

This method changes the composition without changing the temperature. It is important to note that nonionic surfactants, including polyethoxylated surfactants, have temperature-dependent solubility. Emulsification is the process of changing the affinities of surfactants for water and oil that are dependent on temperature. Heat causes polyethoxylated surfactants to become lipophilic due to the dehydration of polyoxyethylene groups. As a result, this circumstance validates the applicability of the PIT method for producing nanoemulsions. To make nanoemulsions with this method, the sample temperature must achieve the PIT (hydrophile–lipophile balance) or PIT level.

The PIT method produces the smallest droplet sizes and interfacial tensions that are feasible. This method improves emulsification by utilizing extraordinarily low interfacial tensions at the HLB temperature. It has been observed, therefore, that despite spontaneous emulsification at the HLB temperature, emulsions are incredibly delicate because of the incredibly fast coalescence rate. It has been reported that stable, fine emulsion droplets can be formed by rapidly cooling the emulsion to a temperature near PIT [8, 9, 10].

3.3 Phase inversion composition (PIC)

This method keeps the temperature constant while changing the composition. Nanoemulsions can be made by progressively adding water or oil to the water/surfactant or oil/surfactant mixture. The PIC technique is more appropriate for large-scale production than the PIT method since adding a single component to an emulsion is easier than producing a sudden change in temperature. When more water is supplied to the system, the water volume increases, and the system reaches a transition composition. As stated otherwise, the surfactant’s spontaneous curvature goes from negative to zero as its polyoxyethylene chains get more hydrated. The transition composition achieves a balance between the lipophilic and hydrophilic properties of the surfactant, much like the HLB temperature, as shown in Figure 1.

3.4 Sonication method

Sonication is the most efficient way to make nanoemulsions. This method reduced the droplet size of conventional emulsions or microemulsions by using a sonication process. This method can only be used to generate small batches of nanoemulsions; large batches cannot be prepared this way. Even though ultrasound can create emulsion directly, it is best to create coarse emulsion before applying acoustic power since breaking an interface requires a lot of energy. Due to its low product throughput, the ultrasonic emulsification process is mostly utilized in laboratories to create emulsion droplets as thin as 0.2 micrometers, as shown in Figure 1 [11].

3.5 Ultrasonic system

“Sonotrodes” (sonicator probes) are composed of piezoelectric quartz crystals that may expand and contract in response to alternating electrical voltage and provide the energy input for ultrasonic emulsification. Cavitations, which occur when the sonicator probe’s tip mechanically vibrates the liquid, are the main process that produces ultrasonically generated effects. Cavitation is the formation and collapse of vapor cavities in a moving liquid. This type of vapor cavity arises when fluctuations in local velocity cause the local pressure to decrease to the temperature of the flowing liquid. When these cavities collapse, strong shock waves are produced that break up the scattered droplets by radiating across the solution along the radiating face of the tip, as shown in Figure 1.

3.6 Microfluidizer

When making emulsion, much higher pressures can be reached—up to about 700 Mpa. This is accomplished at the nozzle of the microfludizer, in the interaction chamber, where two jets of crude emulsion from different channels collide with one another. The process stream is delivered by a pneumatically powered pump that can pressurize the on-site compressed air (150/650 Mpa) to a maximum of 150 Mpa. A high-pressure flow stream driven via microchannels and into an impingement spot results in an impressive shearing motion. This has the potential to create a very fine emulsion.

3.7 High/energy emulsification method

Systems known as nanoemulsions are incapable of self-formation or equilibrium. This is why mechanical or chemical energy must be added to them in order for them to form. Among the mechanical energy input tools utilized in the creation of nanoemulsions are high-pressure homogenizers, high-shear stirring, and ultrasonic generators. These mechanical devices generate strong forces that separate the oil and water phases to form nanoemulsions. In high-energy technologies, the input energy density is approximately 108/1010 W kg/1. In the shortest length of time, the system obtains the energy required to generate uniform, small-sized particles. High/pressure homogenizers are the most widely used instruments for producing nanoemulsions because of their ability to achieve this.

3.8 High-pressure homogenizer

It is the method for producing nanoemulsions that is most frequently utilized. Nanoemulsions with particle sizes as small as 1 nm can be produced using this technology by employing a piston homogenizer or a high-pressure homogenizer. During the process, a small aperture is used to force the macroemulsion through at a pressure of 500–5000 psi. When cavitation, severe turbulence, and hydraulic shear come together, the process creates incredibly small droplet-sized nanoemulsions.

This process can be continued until the final product reaches the desired polydispersity index (PDI) and droplet size. The uniformity of droplet size in nanoemulsions is specified by PDI. Higher PDI in nanoemulsions is connected with lower droplet size uniformity. A PDI of less than 0.08 indicates a monodisperse sample; a PDI of more than 0.3 indicates a narrow size distribution; and a PDI between 0.08 and 0.3 indicates a broad size distribution. However, the production of small droplets at the submicron level requires a large amount of energy. This energy combined with the increasing temperatures during the high-pressure homogenization process, could cause the components to degrade. Enzymes, proteins, and nucleic acids are a few examples of molecules that heat can damage.

3.9 High/shear stirring

This method uses high-energy mixers and rotor/stator systems to prepare nanoemulsions. The droplet sizes of the internal phase can be significantly decreased by increasing the mixing intensity of these devices. However, creating emulsions with an average droplet size of less than 200–300 nm could be difficult [12, 13, 14, 15, 16].

3.10 Low/energy emulsification method

Nanoemulsification can also be accomplished with low-energy methods, resulting in more homogeneous and smaller droplets. These methods such as phase inversion temperature and phase inversion component, produce smaller and more homogeneous droplets by making use of the physicochemical properties of the system. The types of oils and emulsifiers that low-energy techniques can employ are restricted, such as proteins and polysaccharides, even though they are frequently more effective at creating minute droplets than high-energy approaches. To solve this problem, large amounts of artificial surfactants are used in low-energy methods to produce nanoemulsions; nonetheless, this limits their range of uses, especially for many food processing as shown in Figure 1.

3.11 Spontaneous nanoemulsification

It benefits from the chemical energy replacement during the emulsification process, which is based on dilution with the continuous phase and usually proceeds at constant temperature without any phase transitions in the system. This method can produce nanoemulsions at room temperature, hence no specialized equipment is required. Essentially, the following variables influenced it: bulk and interfacial viscosity, phase transition region, surfactant concentration, surfactant structure, and interfacial tension. In the pharmaceutical industry, systems developed with this method are usually called self-emulsifying drug/delivery systems (SEDDS) or self-nano/emulsifying drug/delivery systems (SNEDDS). When water is mixed with an oil phase that has a water-soluble component, oil droplets naturally form. The process is driven by the movement of a chemical that dissolves in water from the oil phase to the water phase. This leads to interfacial turbulence, which helps oil droplets form naturally, as shown in Figure 1 [17, 18, 19].

Advertisement

4. Patents related to nanoemulsion

As the most effective type of intellectual property protection, patents are crucial to a nanotechnology company’s expansion. Patents will be crucial to the success of the global nanotechnology revolution, just as they were to the growth of the biotechnology and information technology industries. In fact, patents are already influencing the young, quickly developing field of nanoscience and small technologies. A company’s long-term survival will depend on its ability to get legitimate and defendable patent protection as it develops nanotechnology-related products and processes and starts looking for commercial uses for its ideas, as stated in Table 1 [20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49].

Patent application titlePatent App. No.Date
Nanoparticles and nanoemulsions14/893,1232017/12/05
Nanoemulsion therapeutic compositions and methods of using the same12/5675712015/02/24
Antimicrobial nanoemulsion compositions and methods82363352012/08/07
Antimicrobial nanoemulsion compositions and methods82323202012/07/31
Topical compositions and methods of detection and treatment201200398142012/02/16
Cancer vaccine compositions and methods of using the same201102809112011/11/17
Methods of using nanoemulsion compositions having anti/inflammatory activity201102006572011/08/18
Stable nanoemulsions for ultrasound/mediated of drug delivery and imaging201101770052011/07/21
Method for the preparation of nanoparticles from nanoemulsion201101357342011/06/09
Antimicrobial nanoemulsion compositions and methods201100703062011/03/24
Nanoemulsion formulations for direct delivery201100450502011/02/24
Lyophilized nanoemulsion201100152662011/01/20
Nanoemulsion vaccines201003166732010/12/16
Nanoemulsion of resveratrol/phospholipid complex and method for preparing the same and applications thereof201002971992010/11/25
Per fluorocarbonnanoemulsion containing quantum dot nanoparticles and method for preparing the same201002330942010/09/16
Compositions for treatment and prevention of acne, methods for making the compositions, and methods of use thereof201002269832010/09/09
Nanoemulsion therapeutic compositions and methods of using the same201000925262010/04/15
Stable mixed emulsions201000695112010/03/18
Antimicrobial nanoemulsion compositions and methods76552522010/02/02
Antimicrobial nanoemulsion compositions and methods201000033302010/01/07
Nanoemulsion influenza vaccine200903047992009/12/10
Nanoemulsion adjuvants200902910952009/11/26
Oil/in/water nanoemulsion, a cosmetic composition, and a cosmetic product comprising it, a process for preparing said nanoemulsion200902085412009/08/20
Nanoemulsion therapeutic compositions and methods of using the same200803177992008/12/25
PCT/AU2008/0017142008/11/18
Nanoemulsion vaccines200802540662008/10/16
Nanoemulsion vaccines200801819052008/07/31
Nanoemulsion vaccines73146242008/01/01
Nanoemulsion compositions having anti/inflammatory activity200700368312007/02/15
Nanoemulsion formulations200201550842002/10/24
Nanomulsion based on nonionic and cationic amphiphilic lipids and uses thereof60399362000/03/21
Solid fat nanoemulsions as vaccine delivery vehicles57166371998/02/10
Solid fat nanoemulsions as drug delivery vehicles55760161996/11/19

Table 1.

Patents on nanoemulsion.

Advertisement

5. Evaluation of nanoemulsion

5.1 Determination of encapsulation efficiency

To determine the amount of drug entrapped in the formulation, a weighed quantity of the formulation is ultrasonically dispersed in an organic solvent to release the drug, which is then extracted into a suitable buffer. Once the right dilutions are made and compared to an adequate blank, the extraction is subjected to spectrophotometric analysis at the drug’s λmax to determine the drug content. The drug’s loading efficiency (LE) and entrapment efficiency (EE) can be calculated using these formulas. Drug LE represents the amount of drug in the achieved product (mg)/total product weight (mg) × 100, while drug EE represents the amount of drug in the obtained product (mg)/total quantity of drug added (mg) × 100. To assess drug concentration, high-performance liquid chromatography (HPLC) in reverse phase can also be employed [50, 51, 52, 53, 54, 55, 56, 57].

5.2 Determination of particle size and polydispersity index (PDI)

Photon correlation spectroscopy (PCS) is used to measure the PDI and particle size of nanoemulsions using a Malvern Zetasizer. PCS calculates the variation in light scattering over time brought on by particle Brownian motion.

The fundamental tenet of PCS is that smaller particles accelerate larger ones. The laser beam is distorted by the submicron particles present in the solution. The rate at which particle diffusion causes the laser scattering intensity to vary around a mean value at a fixed angle depends on the particle size. The calculated photoelectron time correlation function yields a line width distribution histogram that is associated with particle size. A weighed quantity of formulation is combined with double or distilled water to generate a homogenous dispersion, which is then used to measure the particle size. The PDI and particle size must be measured using this mixture immediately. A 0 (zero) PDI represents a monodisperse system while a 1 PDI represents a polydisperse particle dispersion [20].

5.3 Determination of zeta potential

The zeta potential is a method for figuring out a particle’s surface charge in a liquid. Zeta potential is a helpful tool for predicting dispersion stability; the presence and adsorption of electrolytes, as well as the physicochemical properties of the drug, polymer, and medium, all affect its value. The Malvern Zetasizer apparatus is employed for its measurement. Zeta potential is determined by diluting the nanoemulsion and calculating its value using the electrophoretic mobility of the oil droplets. It is believed that a zeta potential of ±30 mV suffices to ensure the nanoemulsion’s physical stability.

5.4 Morphological study of nanoemulsion

The morphology of nanoemulsions is analyzed using transmission electron microscopy (TEM). In a transmission electron microscope (TEM), an electron beam is directed at a thin foil specimen. Upon interaction with the material, these incident electrons become unscattered, elastically scattered, or elastically scattered electrons. The distances between the objective lens and the specimen and between the objective lens and its image plane regulate the magnification.

The electromagnetic lenses focused the scattered or unscattered electrons and projected them onto a screen to produce a contrast/amplitude image, a phase/contrast image, an electron diffraction image, or a phantom image with distinctly darkened parts, depending on the density of unscattered electrons. At higher magnifications, diffraction modes can be used in conjunction with bright field imaging to reveal the size and structure of nanoemulsion droplets. To immobilize the sample for TEM investigation, a suspension of lyophilized nanoparticles or a few drops of nanoemulsion are prepared in double-distilled water and placed onto a holey film grid. The excess solution must be squeezed off the grid and dyed after immobilization [21, 22].

5.5 Atomic force microscope (AFM)

These days, the surface morphology of nanoemulsion formulations is studied using a relatively new technique called AFM. Nanoemulsions are diluted with water and then applied to a glass slide in order to conduct AFM. After that, the coated drops are dried in an oven and scanned at 100 mV/s.

5.6 In vitro drug release study

Performance of drug formulation drug release studies conducted in vitro can help estimate in vivo. The in vitro release rate of a medication is usually studied using a United state pharmacopoeia (USP) dissolving apparatus. Dried nanoparticles or nanoemulsion containing 10 mg of medication were applied to dialysis membrane pouches and placed in a flask containing buffer after being dissolved in buffer. This experiment is run at 37 ± 0.5° with a stirring speed of 50 rpm. Samples are removed on a regular basis and are always replaced with the same volume of fresh dissolving media.

Spectrophotometry is used to detect the absorbance of materials at a certain wavelength after they have been suitably diluted. The absorbance of the sample is used to compute the percentage of drug release at different time intervals using a calibration curve.

5.7 In vitro skin permeation studies

With the Keshary–Chien/diffusion cell, permeation investigations can be studied both in vitro and in vivo. For permeation studies, the abdominal skin of adult male rats weighing 250 ± 10 g is frequently utilized. The donor and recipient chambers of the diffusion cell are separated by the rat skin. Twenty percent ethanol-infused fresh water is added to receiver chambers that are kept at 37°C and rotated at 300 rpm all the time. The formulas are kept in the donor room.

At predetermined intervals, such as 2, 4, 6, and 8 hours, a predetermined volume (0.5 ml) of the receiver chamber’s solution was removed for gas chromatographic analysis. Each time, the sample was quickly replaced with an equivalent volume of brand-new solution. Three runs are made for each sample. The total amount of medication absorbed via rat skins at each time interval is plotted against a function of time following cumulative adjustments. In steady-state conditions, the plot slope is utilized to calculate medicine penetration rates [23].

5.8 Stability studies

The purpose of stability studies is to assess how stable a medicinal ingredient is in the presence of various environmental factors, including temperature, humidity, and light. The formulation is stored for 24 months in a dispersed and freeze/dried state before the stability studies of the nanoemulsion are carried out in accordance with the guidelines set forth by the International Conference on Harmonization. There are three storage conditions that are utilized: ambient (25 ± 2°/60 ± 5% RH), refrigerated (5 ± 3°), and freeze (−20 ± 5°). The necessary quantity of nanoemulsion is stored in glass vials with hermetically sealed closures. Samples are removed and analyzed for characteristics such as loading, particle size, EE, and the in vitro drug release profile at predetermined intervals [24].

5.9 Shelf life determination

Expedited stability studies are performed to determine the shelf life of a nanoemulsion. The formulations are stored at three distinct temperatures and relative humidity levels (30°, 40°, and 50 ± 0.5°) for almost 3 months. After a predetermined period of time (0, 30, 60, and 90 days), samples are taken out and analyzed using HPLC at λmax to find out how much medication remains. The samples that are removed at zero time are known as control samples. This determines the order of the reaction. Next, using the following equation, the reaction rate constant (K) for the deterioration is calculated at each elevated temperature based on the slope of the lines: An Arrhenius plot with slope = −K/2.303 is produced by plotting the logarithms of K at different elevated temperatures against the reciprocal of absolute temperature. This data yields the plot value of K at 25°, which is then used to calculate shelf life by plugging the value into the calculation t0.9 = 0.1052/K25. Here, t0.9—also referred to as shelf life—is the amount of time required for a 10% degradation of the drug [25].

5.10 Thermodynamic stability studies

The majority of thermodynamic stability studies are carried out in three steps. Initially, a cycle of heating and cooling is conducted to determine whether adjusting the temperature affects the stability of the nanoemulsion in any way. The nanoemulsion is subjected to six cycles of temperature between 4° (the refrigeration temperature) and 40° while the formulation is stored at each temperature for a minimum of 48 hours.

The formulations that show stability at these temperatures are used in centrifugation experiments. The second technique involves centrifuging the manufactured nanoemulsions for 30 minutes at 5000 rpm in order to check for phase separation, creaming, or cracking. Those who showed no signs of instability could experience a freeze-thaw cycle. The third technique is the freeze/thaw cycle, which entails freezing and thawing nanoemulsion formulations three times at temperatures between −21° and + 25°.

If a formulation shows no signs of instability, this test finds that it has good stability. These mixtures are then subjected to dispersibility tests to determine how well they self- or emulsify. This test determines that a formulation has good stability if it exhibits no indicators of instability. The next step is to put these mixtures through dispersibility tests to see how effectively they self- or emulsify.

5.11 Dispersibility studies

Dispersibility tests are performed utilizing conventional USP XXII dissolving equipment to assess the effectiveness of self-emulsification or nanoemulsion formation. Each formulation is added to 500 ml of distilled water that is kept at 37 ± 0.5°; 2.1 ml of each formulation is used. A typical dissolution paddle made of stainless steel spins at 50 revolutions per minute to gently agitate the mixture. The nanoemulsion formulations’ In vitro performance is assessed visually by the application of a grading system that is explained below. Level A nanoemulsions seem clear or bluish and form quickly—within 1 minute. Although they are somewhat less transparent, Grade B nanoemulsions develop quickly and have a bluish-white appearance. Within 2 minutes, grade C nanoemulsions, which are fine, milky emulsions form. Grade D emulsions are slower and more dull, with a grayish-white look and a hint of oiliness [26].

5.12 Determination of viscosity

Viscosity measurement is a crucial component of a nanoemulsion’s physicochemical assessment. Viscosity is measured using a number of tools, such as the Ostwald viscometer, Hoeppler falling ball viscometer, Stormer viscometer, Brookfield viscometer, and Ferranti/Shirley viscometer. Of all these devices, the Brookfield viscometer is the one that is suggested for figuring out the viscosity of a nanoemulsion.

The assessment of viscosity verifies whether the system is an O/W or W/O emulsion. Systems that exhibit low viscosity are classified as O/W types, whereas those that exhibit high viscosity are classified as water-in-oil types. Nonetheless, the survismeter has emerged as the most often utilized piece of apparatus since it assesses the particle size, hydrodynamic volumes, contact angle, dipole moment, interfacial tension, surface tension, and viscosity of the nanoemulsions.

5.13 Refractive index

The refractive index provides information on the transparency of the nanoemulsion and the transmission of light through the medium. The ratio of the wave’s phase speed (vp) in the medium to its speed (c) in the reference medium is the formula for determining a medium’s refractive index (n), and it is expressed as n = c/vp. The refractive index of the nanoemulsion can be determined by placing a drop on a slide and comparing it to the refractive index of water (1.333) using an Abbes-type refractometer set at 25 ± 0.5. In order for a nanoemulsion to be considered transparent, its refractive index must equal that of water.

5.14 pH and osmolarity measurements

A pH meter is used to determine a nanoemulsion’s pH, whereas a microosmometer uses the freezing point method to determine an emulsion’s osmolarity. Once 100 μl of nanoemulsion has been transferred into a microtube, measurements are taken.

5.15 Dye solubilization

A water soluble dye is dispersible in an O/W globule as opposed to soluble in the aqueous phase of the W/O globule. Similarly, an oil-soluble dye is dispersible in the W/O globule but dissolves in the oily phase of the O/W globule. An O/W nanoemulsion will absorb color evenly when water-soluble dye is applied; however, if the emulsion is W/O, the dye will only remain in the dispersed phase, and the color will not absorb evenly [27].

5.16 Dilutability test

The stability of a nanoemulsion can be preserved when a continuous phase is added in larger quantities which justifies the dilution test. Consequently, unlike O/W nanoemulsions, W/O nanoemulsions do not dilute with water and instead experience phase inversion into O/W nanoemulsions. W/O nanoemulsion can only be diluted with oil [28].

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Abbreviations

BCS

biopharmaceutical drug classification system

HLB

hydrophilic-lipophilic balance

SC

stratum corneum

PIT

phase inversion temperature

PIC

phase inversion composition

PDI

polydispersity index

SEDDS

self-emulsifying drug/delivery systems

SNEDDS

self-nano/emulsifying drug/delivery systems

LE

loading efficiency

EE

entrapment efficiency

HPLC

high-performance liquid chromatography

PCS

Photon correlation spectroscopy

TEM

transmission electron microscopy

AFM

atomic force microscope

References

  1. 1. Jaiswal M, Dudhe R. Nanoemulsion: An advanced mode of drug deliverysystem. Biotech. 2015;5:123-127. DOI: 2010.1007/s13205-014-0214-0
  2. 2. Dorkoosh FA, Brussee J, Verhoef JC, Borchard G, Rafiee Tehrani M, Junginger HE. Preparation and NMR characterization of superporous hydrogels and Superporous hydrogel composites. Polymer. 2000;41(23):8213-8220. DOI: 2010.1016/S0032-3861(00)00200-7
  3. 3. Okyar A, Özsoy Y, GüngÖr S. Novel formulation approaches for dermal and transdermal delivery of non-steroidal anti-inflammatory drugs. In: Lemmey A, editor. Rheumatoid Arthritis–Treatment. Croatia: Intech; 2012. pp. 25-48. DOI: 10.5772/28461
  4. 4. Sunil KY, Manoj KM, Anumapaa T, Ashutosh S. Emulgel a new approach for enhanced for topical drug delivery. International Journal of Current Pharmaceutical Research. 2017;9(1):15-19. DOI: 10.22159/ijcpr.16628
  5. 5. Hardenia A, Jayronia S, Jain S. Emulgel: An emergent tool in topical drug delivery. International Journal of Pharmaceutical Sciences Review and Research. 2014;5(5):1653-1660. DOI: 2010.13040/IJPSR.0975-8232.5(5).1653-60
  6. 6. Jana S, Ali SA, Nayak AK, Sen KK, Basu SK. Development of topical gel containing aceclofenac-crospovidone solid dispersion by “quality by design (QbD)” approach. Chemical Engineering Research and Design. 2014;9(2):2095-2105. DOI: 10.1016/j.cherd.2014.01.025
  7. 7. Mohamed MI. Optimization of chlorphenesin emulgel formulation. The AAPS Journal. 2004;6(3):81-87. DOI: 2010.1208/aapsj060326
  8. 8. Jones DS, Woolfson AD, Brown AF. Textural, viscoelastic and mucoadhesive properties of pharmaceutical gels composed of cellulose polymers. International Journal of Pharmaceutics. 1997;151(2):223-233. DOI: 10.10 16/S0378-5173(97)04904-1
  9. 9. Andrews GP, Gorman SP, Jones DS. Rheological characterization of primary and binary interactive bioadhesive gels composed of cellulose derivatives designed as ophthalmic viscosurgical devices. Biomaterials. 2015;26(5):571-580. DOI: 2010.1016/j.biomaterials. 2004.02.062
  10. 10. Rai VK, Mishra N, Yadav KS, Yadav NP. Nanoemulsion as pharmaceutical carrier for dermal and transdermal drug delivery: Formulation development, stability issues, basic considerations and applications. Journal of Controlled Release. 2018;270:203-225. DOI: 2010.1016/ j.jconrel.2017.11.049
  11. 11. John J, Bhattacharya M, Raynor PC. Emulsions containing vegetable oils for cutting fluid application. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2004;237(1):141-150. DOI: 2010.1016/ j.colsurfa.2003.12.029
  12. 12. Mehmood T, Ahmad A, Ahmed A, Ahmed Z. Optimization of olive oil based O/W nanoemulsions prepared through ultrasonic homogenization: A response surface methodology approach. Food Chemistry. 2017;229:790-796. DOI: 2010.1016/j.foodchem.2017.03.023
  13. 13. Lu W-C, Chiang B-H, Huang D-W, Li P-H. Skin permeation of d-limonene-based nanoemulsions as a transdermal carrier prepared by ultrasonic emulsification. Ultrasonics Sonochemistry. 2014;21(2):826-832. DOI: 2010.1016/j.ultsonch.2013.10.013
  14. 14. Jafari SM, He Y, Bhandari B. Nano-emulsion production by sonication and microfluidization-a comparison. International Journal of Food Properties. 2006;9(3):475-485. DOI: 2010.1080/10942910600596464
  15. 15. Saberi AH, Fang Y, McClements DJ. Effect of glycerol on formation, stability, and properties of vitamin-E enriched nanoemulsions produced using spontaneous emulsification. Journal of Colloid and Interface Science. 2013;411:105-113. DOI: 2010.1016/j.jcis.2013.08.041
  16. 16. Sriamornsak P, Limmatvapirat S, Piriyaprasarth S, Mansukmanee P, Huang Z. A new self-emulsifying formulation of mefenamic acid with enhanced drug dissolution. Asian Journal of Pharmaceutical Sciences. 2015;10:121-127. DOI: 10.1016/j.ajps.2014.10.003
  17. 17. Patel KC, Pramanik S. Formulation and characterization of mefenamic acid loaded polymeric nanoparticles. World Journal of Pharmacy and Pharmaceutical Sciences. 2014;3(6):1391-1405. 2097348896
  18. 18. Oswal T, Naik S. Formulation and evaluation of mefenamic acid emulgel. IJPRD. 2014;5(12):91-100. DOI: 10.1016/j.jsps.2011.08.001
  19. 19. Kamboj S, Saini V, Bala S, Sharma G. Formulation and characterization of drug loaded niosomal gel for anti-inflammatory activity. International Journal of Medical, Health, Biomedical, Bioengineering and Pharmaceutical Engineering. 2013;7(2):541-545. 207231875
  20. 20. Nanoparticles and nanoemulsions 14/893,123. 2017
  21. 21. Nanoemulsion therapeutic compositions and methods of using the same 12/567571. 2015
  22. 22. Antimicrobial nanoemulsion compositions and methods 8236335. 2012
  23. 23. Antimicrobial nanoemulsion compositions and methods 8232320. 2012
  24. 24. Topical compositions and methods of detection and treatment 20120039814. 2012
  25. 25. Cancer vaccine compositions and methods of using the same 20110280911. 2011
  26. 26. Methods of using nanoemulsion compositions having anti/inflammatory activity 20110200657. 2011
  27. 27. Stable nanoemulsions for ultrasound/mediated of drug delivery and imaging 20110177005. 2011
  28. 28. Method for the preparation of nanoparticles from nanoemulsion 20110135734. 2011
  29. 29. Antimicrobial nanoemulsion compositions and methods 20110070306. 2011
  30. 30. Nanoemulsion formulations for direct delivery 20110045050. 2011
  31. 31. Lyophilized nanoemulsion 20110015266. 2011
  32. 32. Nanoemulsion vaccines 20100316673. 2010
  33. 33. Nanoemulsion of resveratrol/phospholipid complex and method for preparing the same and applications thereof 20100297199. 2010
  34. 34. Per fluorocarbon nanoemulsion containing quantum dot nanoparticles and method for preparing the same 20100233094. 2010
  35. 35. Compositions for treatment and prevention of acne, methods for making the compositions, and methods of use thereof 20100226983. 2010
  36. 36. Nanoemulsion therapeutic compositions and methods of using the same 20100092526. 2010
  37. 37. Stable mixed emulsions 20100069511. 2010
  38. 38. Antimicrobial nanoemulsion compositions and methods 7655252. 2010
  39. 39. Antimicrobial nanoemulsion compositions and methods 20100003330. 2010
  40. 40. Nanoemulsion influenza vaccine 20090304799. 2009
  41. 41. Nanoemulsion adjuvants 20090291095. 2009
  42. 42. Oil/in/water nanoemulsion, a cosmetic composition and a cosmetic product comprising it, a process for preparing said nanoemulsion 20090208541. 2009
  43. 43. Nanoemulsion therapeutic compositions and methods of using the same 20080317799. 2008
  44. 44. Nanoemulsion vaccines 20080254066. 2008
  45. 45. Nanoemulsion vaccines 20080181905. 2008
  46. 46. Nanoemulsion vaccines 7314624. 2008
  47. 47. Nanoemulsion compositions having anti/inflammatory activity 20070036831. 2000
  48. 48. Solid fat nanoemulsions as vaccine delivery vehicles 5716637. 1998
  49. 49. Solid fat nanoemulsions as drug delivery vehicles 5576016. 1996
  50. 50. Faiyaz S, Shafiq S, Sushma T, Farhan J, Khar R, Ali M. Development and bioavailability assessment of ramipril nanoemulsion formulation. European Journal of Pharmaceutics and Biopharmaceutics. 2007;66:227-243
  51. 51. NagibA. Nanoemulsion and Nanoemulgel as a TopicalFormulation. IOSR Journal of Pharmacy. 2015;5(10):43-47. DOI: 283046065
  52. 52. Funk JL, Frye JB, Oyarzo JN, Chen J, Zhang H, Timmermann BN. Anti-inflammatory effects of the essential oils of ginger (Zingiber officinale roscoe) in experimental rheumatoid arthritis. Pharma Nutrition. 2016;4(3):123-131. DOI: 10.1016/j.phanu.2016.02.004
  53. 53. Pavoni L, Perinelli DR, Bonacucina G, Cespi M, Palmieri GF. An overview of micro- and nanoemulsions as vehicles for essential oils: Formulation, preparation, and stability. Nanomaterials. 2020;10(1):135. DOI: 10.3390/nano10010135
  54. 54. Salim N, Basri M, Rahman MA, Abdullah DK, Basri H, Salleh AB. Phase behaviour, formation and characterization of palm-based esters nanoemulsion formulation containing ibuprofen. Journal of Nanomedicine and Nanotechnology. 2011;2(4):1-5. DOI: 10.4172/2157-7439.1000113
  55. 55. Chuacharoen T, Prasongsuk S, Sabliov CM. Effects of surfactant concentrations on physicochemical properties and functionality curcumin nanoemulsions under conditions relevant to commercial utilization. Molecules. 2019;24:2744. DOI: 10.3390/molecules 24152744
  56. 56. Sakulku U, Nuchuchua O, Uawongyart N, Puttipipatkhachorn S, Soottitantawat A, Ruktanonchai U. Characterization and mosquito repellent activity of citronella oil nanoemulsion. International Journal of Pharmaceutics. 2009;372(1):105-111. DOI: 1016/j.ijpharm.2008.12.029
  57. 57. Thakkar HP, Khunt A, Dhande RD, Patel AA. Formulation and evaluation of itraconazole nanoemulsion for enhanced oral bioavailability. Journal of Microencapsulation. 2015;32(6):559-569. DOI: 10.3109/02652048.2015.1065917

Written By

Vaibhav Changediya

Submitted: 04 March 2024 Reviewed: 04 March 2024 Published: 03 June 2024