Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Molecular Liquids 394 (2024) 123720

Contents lists available at ScienceDirect

Journal of Molecular Liquids


journal homepage: www.elsevier.com/locate/molliq

Stable nanoemulsions for poorly soluble curcumin: From production to


digestion response in vitro
Qianyu Ye , Sophie Kwon , Zi Gu , Cordelia Selomulya *
School of Chemical Engineering, UNSW Sydney, NSW 2052, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Curcumin, a polyphenol, can induce anticancer activity depending on dose. However, oral curcumin adminis­
Curcumin tration is limited by its low bioavailability due to aqueous insolubility and instability against physiological
Nanoemulsion conditions. This study aims at formulating nanoemulsions by phase inversion temperature to enhance curcumin
Stability
loading, stability, antioxidant performance, bioaccessibility, and in vitro absorption. The selection mechanisms
Antioxidant activity
In vitro digestion
for oil phase (coconut oil), surfactant (polyoxyl 40 hydrogenated castor oil), co-surfactant (soy phospholipid),
and aqueous phase (2 % wt citrate buffer at pH 4.5) are established. The nanoemulsions show tunable mean
droplet size (26–129 nm), high curcumin loading (9.53 ± 0.49 mg/mL), polydispersity < 0.3, and ζ-potential <
-26 mV. Nanoencapsulation increases curcumin photostability by 36–42 % and the antioxidant activity after 24-
hour UV radiation is unchanged (P > 0.05). The curcumin nanoemulsions show ~ 11 %, 24 %, and 57 % higher
retention and ~ 10 %, 12 %, and 17 % higher antioxidant activity than raw curcumin after 3-hour simulated
gastric, intestinal, and physiological incubations, respectively. During in vitro digestion and absorption, the
encapsulated curcumin shows higher bioaccessibility and absorption than free curcumin (P < 0.05). The samples
are stable during 4-week storage at 4˚C and room temperature without preservatives. These findings suggest the
potential to develop a nanoencapsulation strategy, particularly for an oral delivery system of oil-soluble drugs.

1. Introduction liposomes [11], complex powders [6], nanoemulsions [12], and others.
Nanoemulsions are a nano-sized colloidal particulate system acting
Despite advances in cancer treatment, breast cancer remains one of as carriers for improving the delivery of active drug molecules. In oil-in-
the leading causes of death for women, with over 2 million women water (o/w) nanoemulsions, oil droplets (d < 100 nm) are encapsulated
diagnosed with breast cancer and about 700,000 deaths globally in 2020 within a protective coating of surfactant and co-surfactant molecules
[1]. It has been demonstrated that several dietary components like and are dispersed in an aqueous continuous phase [13]. The composition
curcumin from turmeric spice enable to induce anticancer activity and preparation conditions affect the physicochemical and functional
against breast cancer cell lines by increasing reactive oxygen species properties of nanoemulsions such as droplet size distribution, droplet
production and promoting apoptosis in cancer cells but not in normal concentration, interfacial tension, storage stability, and transparency
cells [2,3]. Curcumin has been classified as a Generally Regarded as Safe [12]. When the dimension of oil droplets is significantly smaller than the
(GRAS) compound by the US Food and Drug Administration and has wavelength of light, the resulting weak scattering will make the nano­
been established an adequate daily intake value of 3 mg/kg body weight emulsions optically transparent [14]. Compared with conventional
per day by Joint FAO/WHO Expert Committee on Food Additives and emulsions with a droplet size of 1–20 µm, nanoemulsions display some
European Food Safety Authority [4]. However, the anticancer activity of advantages e.g., transparent or translucent appearance, longer shelf-life,
curcumin is dose-dependent [5] and restrained by its poor aqueous low viscosity and interfacial tension, higher surface area and cellular
solubility (0.11 ± 0.01 µg/ml in Milli-Q water) [6], low photostability uptake, and potential to enhance bioavailability and absorption [14,15].
[7], instability under neutral and alkaline conditions [8], low bio­ Therefore, nanoemulsions have been considered as potential carriers for
accessibility, and extensive first-pass metabolism [9,10]. To solve these entrapping, stabilising, and incorporating water-insoluble drugs like
issues, various encapsulation strategies have been developed, including curcumin into transparent aqueous-based medical [16], food [17],

* Corresponding author.
E-mail address: cordelia.selomulya@unsw.edu.au (C. Selomulya).

https://doi.org/10.1016/j.molliq.2023.123720
Received 11 October 2023; Received in revised form 15 November 2023; Accepted 30 November 2023
Available online 2 December 2023
0167-7322/© 2023 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

beverage [18], and cosmetic products [19]. To produce nanoemulsions liquid coconut oil, octanoic acid, oleic acid, and linoleic acid) and
with smaller oil droplets, high-energy emulsification methods are nor­ nonionic surfactants (Tween 40, Tween 80, PEG 200, PEG 400, PEG 600,
mally applied e.g., high-pressure homogenisation [13] and ultra­ and RH 40), was determined following a modified method from Mandal,
sonication [20]. The high energy consumption during processing may Mandal [16]. An excess of curcumin (~200 mg) was added into a glass
incur costs and degradation of active ingredients, making these methods vial containing 2 g of each vehicle. After sealing, the mixtures were
unfavourable for scaling up production. By contrast, the phase inversion heated at 60 ◦ C for 1 h in a water bath to accelerate the solubilisation of
temperature (PIT) method is a low-energy emulsification method that curcumin and then were shaken at 200 rpm and 40 ◦ C for 24 h to reach
utilises the chemical energy of the surfactant i.e., the temperature- equilibrium. After centrifugation at 18,000g for 10 min using a Micro­
dependent solubility and molecular geometry of nonionic surfactants fuge® 22R centrifuge (Beckman Coulter, USA), the supernatants were
[21]. At the PIT, nonionic surfactants show the minimum of interfacial diluted with absolute ethanol for quantification using a microplate
tension and the same affinity for the aqueous and oil phases, caused by reader (Infinite M Nano+, TECAN, Switzerland) at 425 nm based on
the surfactant spontaneous curvature passing zero [22], which favour­ calibrations made with known amounts of curcumin dissolved in
ably forms kinetically stable actives-loaded nanoemulsions. ethanol.
Although there are several reports on the preparation of curcumin-
fortified nanoemulsions by the PIT method [12,23,24], to the best of 2.2.2. Cloud point of nonionic surfactant solutions
our knowledge, the selection criteria for oil phase, surfactant, co- Citrate buffer was selected as the aqueous phase of the o/w nano­
surfactant and aqueous phase, and the mechanisms underlying the emulsions to dissolve surfactants, whose selection criteria and mecha­
process have never been systematically studied. This study also aims to nisms will be discussed in detail in section 3.4. The cloud point (CP) of
formulate curcumin-fortified nanoemulsions by the PIT method to surfactant solutions was measured based on the procedure adopted by
enhance the loading capacity, storage, pH and UV stability, antioxidant Hasan, Leanpolchareanchai [26] with minor modifications. Nonionic
performance, bioaccessibility, and in vitro intestinal absorption of surfactants were dissolved in 0–2 % wt citrate buffer at surfactant con­
curcumin. tents of 10 %, 20 %, 30 %, 40 %, and 50 % wt, and then were gradually
heated to 100 ◦ C under magnetic stirring, followed by three temperature
2. Materials and methods cycles (100–60-100–60-100–60 ◦ C). When a clear surfactant solution
reversed into a cloudy system, the temperature range was recorded as
2.1. Materials the CP temperature of the surfactant solutions i.e., phase inversion
temperature (PIT). The three temperature cycles were used to check the
Curcumin from Curcuma longa (Turmeric) (≥65 %), soy phospho­ reproducibility of PIT.
lipids, polyoxyl 40 hydrogenated castor oil (RH 40), octanoic acid, oleic
acid, linoleic acid, poly(ethylene glycol) 200 (PEG 200), PEG 400, PEG 2.2.3. Nanoemulsion preparation and pseudo-ternary phase diagram
600, propylene glycol, pepsin from porcine gastric mucosa powder construction
(≥3,200 units/mg protein), and bile salts were obtained from Sigma- Blank and curcumin-loaded nanoemulsions were prepared by the PIT
Aldrich, St Louis, Mo, USA. Liquid coconut oil was purchased from method using a modified method from Jintapattanakit, Hasan [12]. The
Natural Raw C Pty Ltd, Australia. Sunflower oil was from Crisco Pre­ selected components were mixed at room temperature (20–22 ◦ C) and
mium Oil, Australia. Canola oil was from Farmers Harvest, Australia. gradually heated to 90 ◦ C under magnetic stirring at 200 rpm, followed
Pancreatin from porcine pancreas (8 x USP specifications) was pur­ by three temperature cycles (90–60-90–60-90–75 ◦ C) to achieve the
chased from MP Biomedicals. Other chemicals used were from Chem phase inversion process. Then, twice the amount of cold milli-Q water
Supply, Australia. Milli-Q water from a Millipore system (18.2 MΩ cm− 1 was added into the system for fast cooling down from 75 ◦ C, followed by
at 25 ◦ C). All chemicals were used without further purification. 20-min magnetic stirring at 200 rpm at room temperature. Formulations
Simulated Gastric Fluid (SGF) and Simulated Intestinal Fluid (SIF) containing different ratios of the oil phase, surfactant, co-surfactant, and
were prepared following the standard INFOGEST protocol [25]. Elec­ aqueous phase were examined accordingly to find out the nanoemulsion
trolyte stock solutions containing 0.5 mol/L potassium chloride, 0.5 existing zone in the pseudo-ternary phase diagrams.
mol/L potassium dihydrogen phosphate, 1 mol/L sodium bicarbonate, 2
mol/L sodium chloride, 0.15 mol/L magnesium chloride, and 0.5 mol/L
ammonium carbonate were prepared. 100 mL SGF (adjusted to pH 1.6 2.3. Nanoemulsion characterisation
using 6 M hydrochloric acid) was composed of 80 mL of the electrolyte
stock solutions, 0.05 mL 0.3 M calcium chloride solution, 19.95 mL 2.3.1. Loading efficiency of curcumin nanoemulsions
Milli-Q water, and porcine pepsin with an enzyme activity of 2000 The loading efficiency of freshly prepared curcumin nanoemulsions
units/mL in the final digestion mixtures. 200 mL SIF (adjusted to pH 7.0 was measured to determine the appropriate curcumin incorporation,
with 6 M hydrochloric acid) was prepared with 160 mL of the electrolyte using Eq. (1):

Amount of curcumin encapsulated


Loading efficiency(LE)(%) = × 100 (1)
Amount of curcumin added in preparing nanoemulsion

stock solutions, 0.4 mL 0.3 M calcium chloride solution, 39.6 mL Milli-Q The amount of curcumin encapsulated was determined by centrifuging
water, 10 mM bile salts, and pancreatin with a pancreatic lipase activity the nanoemulsions at 18,000g for 20 min at 20 ◦ C, collecting an aliquot
of 16 or 400 units/mL in the final digestion samples. of the supernatants, diluting the supernatants with absolute ethanol, and
measuring the curcumin absorbance at 425 nm with the microplate
2.2. Nanoemulsion preparation reader. No clear phase separation and curcumin precipitation were
observed after centrifugation due to the small droplet size of nano­
2.2.1. Curcumin solubility in various vehicles emulsions (d < 100 nm). The supernatants collected were akin to the
The solubility of curcumin in oil phases (canola oil, sunflower oil, original nanoemulsions.

2
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

2.3.2. Physical and chemical stability of nanoemulsions during storage 2.3.7. In vitro gastrointestinal digestion and intestinal absorption
The nanoemulsions were stored in the darkness at room temperature For standardised in vitro gastric digestion [25], 5 mL of the curcumin-
and 4 ◦ C for 4 weeks without additional preservatives. The physical loaded nanoemulsions were mixed with 5 mL of the SGF, followed by
stability of nanoemulsions was examined by the change in droplet size rapid pH adjustment to 3.0 with 1 M sodium hydroxide solution. A
and zeta potential of nanoemulsions using a Zetasizer Nano ZS (Malvern certain amount of raw curcumin was dispersed into milli-Q water to
Instruments, UK), during this storage. The chemical stability was achieve the same curcumin addition of nanoemulsions containing 2 mg/
determined by the variation in the curcumin concentration of the mL curcumin, as a control. The gastric digestion samples were incubated
nanoemulsions during the same storage. The curcumin concentration in an orbital shaker incubator (Bioline, Edwards Instrument, Australia)
was evaluated using the same method in section 2.3.1. at 200 rpm and 37 ◦ C for 2 h. 200 µL of gastric digestion samples were
collected every 30 min and centrifuged at 18,000g and 4 ◦ C for 10 min,
2.3.3. IR spectra collection while no curcumin precipitation and phase separation were found. 100
IR spectra of nanoemulsions and various components were recorded µL of the supernatants were diluted with ethanol and the absorbance of
and analysed using an IR Alpha (Bruker) instrument with platinum ATR curcumin was tested at 425 nm using the microplate reader. After each
(Bruker Optics GmbH, Germany) and OPUS software version 7.5 (Bruker sample collection, 200 µL of the SGF were added to the system to make
Optics GmbH, Germany). sure the final volume was unchanged.
After 2-hour in vitro gastric incubation, the 10 mL digestion samples
2.3.4. Photostability of curcumin in nanoemulsions were mixed with 10 mL of the SIF, adjusted to pH 7.0 using 6 M hy­
2 g of nanoemulsions were loaded into an open glass vial, separately, drochloric acid and 1 M sodium hydroxide solution, with a pancreatic
covered by a single layer of cling film to avoid moisture evaporation and lipase activity of 400 units/mL in the final mixtures [25]. The intestinal
allow light to pass through [27]. As a control, a certain amount of raw digestion samples were incubated at 37 ◦ C and 200 rpm for 2 h. 200 µL of
curcumin was dissolved in 2 g of absolute ethanol at the same concen­ the chyme were collected every 30 min and centrifuged at 18,000g and
tration as that of nanoemulsions. Soft ultraviolet (UV) radiation (18 W, 4 ◦ C for 10 min. The curcumin concentration was determined using the
350–400 nm) was performed using a fluorescent blacklight blue lamp same method above. 200 µL of the SIF were added to the system after
(F18T8 BLB 1709, Crompton Lighting, Australia) with an 8-cm distance each collection.
at room temperature for 24 h to accelerate the photodegradation of
curcumin. The curcumin content as a function of radiation time was 2.3.8. In vitro intestinal absorption using a dialysis model
tested using the method in section 2.3.1. The in vitro intestinal absorption of lipophilic curcumin was simu­
lated with a dialysis model (MWCO 12,000–14,000 Da) following a
2.3.5. Curcumin stability under different physiological conditions modified method from Hu, Wu [28]. The gastrointestinal (GI) digestion
As curcumin is unstable at neutral and alkaline pH, the stability of samples were dialysed against PBS-T buffer (PBS buffer containing 0.05
nanoemulsions was evaluated under simulated gastric, intestinal, and % wt Tween 80 outside of the dialysis bag) at 1:5 v/v, incubated at 37 ◦ C
physiological conditions [6]. Curcumin ethanolic solution with the same and 130 rpm for 3 h. Tween 80 was added to the PBS buffer as an
curcumin concentration as that of nanoemulsions was used as a control. emulsifier to dissolve the water-insoluble curcumin [6]. As curcumin in
The samples were dispersed and incubated in the SGF at pH 3.0, the SIF this study was encapsulated in lipid nanodroplets, this would facilitate
with 16 units/mL pancreatic lipase at pH 7.0, and phosphate buffered drug-loaded droplets to pass through the intestinal lymphatic transport
saline (PBS) at pH 7.4, separately, at 37 ◦ C for 3 h. At incubation time 0, [29]. The intestinal lymphatic fluid contains large amounts of albumin
the curcumin content in all final mixtures was 16 µg/mL. The curcumin up to 0.012–0.036 g/mL [30], which enables to solubilise curcumin via
concentration retained with time was determined with the microplate binding [31]. Therefore, the addition of 0.05 % wt Tween 80 to PBS
reader at 425 nm. buffer was regarded as an alternative way to simulate the solubilisation
and transport of curcumin in the lymphatic fluid.
2.3.6. DPPH scavenging activity of curcumin nanoemulsions The first approximation of the rate of a drug’s ability to cross the
Before and after photodegradation: DPPH (2,2-diphenyl-1-picrylhy­ intestinal barrier is expressed as the apparent permeability [32]. In this
drazyl) radical scavenging activity was measured to evaluate the anti­ study, the apparent permeability coefficient (Papp) of perfusate transport
oxidant activity of curcumin in nanoemulsions following a modified across the dialysis model was obtained from the steady-state level of the
method [6]. The samples before and after 24-hour UV radiation in sec­ perfusate diffusing out of the dialysis membrane using Eq (3) as follows
tion 2.3.4 were diluted with milli-Q water at a fixed ratio, separately. [33]:
Afterwards, 0.5 mL of the diluted samples was mixed with 0.5 mL of 0.2 ( ) dQnormalised dQ
mM DPPH ethanolic solution, followed by vigorously mixing and 30-min Papp cm min− 1 = = (3)
dt A × (Ci − C0 )dt
incubation in darkness. Then, the final mixtures were centrifuged at
18,000g for 5 min and the absorbance of the supernatants was measured where Q represents the amount of drug tested in the container outside of
at 527 nm using the microplate reader. The mixtures of 0.5 mL of the the intestinal barrier (mimicked by the dialysis tube), mg; A means the
diluted samples and 0.5 mL of ethanol were considered a blank. The effective surface area of the dialysis membrane available for diffusion,
mixture of 0.5 mL of milli-Q water and 0.5 mL of DPPH ethanolic so­ cm2; Ci and C0 are the drug content inside and outside the dialysis tube at
lution was used as the control. The DPPH scavenging activity was t = 0 min, mg/mL; and Qnormalised is the normalised Q given the differ­
calculated using Eq. (2): ences in the surface area of the dialysis membrane and the initial drug
(
OD of sample − OD of blank
) content inside and outside the dialysis tube, cm.
Radical scavenging activity (%) = 1 − The overall ability of a drug to cross the intestinal barrier can be
OD of control
expressed as the absorption ratio, using Eq (4) as follows [28]:
× 100
(2) Q0 − Qt
Absorption ratio (%) = × 100% (4)
Q0
Before and after physiological incubation: In section 2.3.5, the samples
freshly mixed with the SGF, SIF, and PBS and the samples mixed and where Q0 represents the dose of the drug added to the dialysis tube at t =
incubated for 3 h were added to 0.2 mM DPPH ethanolic solution at 1: 1 0 min, mg; and Qt means the amount of the drug still in the dialysis tube
v/v, separately. Other steps for the DPPH scavenging activity mea­ at t minutes, mg.
surement were the same as those above.

3
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

2.4. Statistical analysis nontoxic and nonionic surfactants with HLB values ranging from 11 to
18 suitable for o/w emulsions. To increase the loading capacity of cur­
All sample collections and measurements were conducted in tripli­ cumin in the nanoemulsions, RH 40 with the highest curcumin solubility
cate. The average and standard deviation were calculated using Micro­ was selected as the surfactant (Table 1). The difference in curcumin
soft Excel. The statistical differences between the groups were evaluated solubility in six surfactants was seemingly irrelevant to their HLB values,
using one-way analysis of variance (ANOVA) in IBM SPSS Statistics 26.0 as RH 40, Tween 40, and Tween 80 have similar HLB values ranging
(IBM Inc., USA) at a probability level of 0.05. Origin 2019 (OriginLab from 14 to 16 in Table 1 [36]. The higher curcumin solubility in RH 40
Corporation, USA) was applied to plot some figures. was presumably caused by its three hydrophobic tails in one molecule
which provided more lipophilic structures to interact with curcumin,
3. Results and discussion while other surfactants only had one hydrophobic tail in each molecule.
A similar phenomenon was observed in the lower curcumin solubility in
3.1. Selection criteria for the oil phase octanoic acid with one hydrophobic tail (C8) and the higher solubility in
coconut oil with three hydrophobic tails (mainly C8) in section 3.1.
As shown in Table 1, sunflower oil, canola oil, and coconut oil were Therefore, surfactants with HLB values ranging from 14 to 16 and more
tested as commonly used vegetable oils and exhibited statistically lipophilic structures can effectively stabilise nanoemulsions and in­
different solubilities for curcumin, probably due to their different fatty crease curcumin loading.
acid profiles. According to the nutrition information provided by the
manufacturers, canola oil, sunflower oil, and coconut oil contained 93.3 3.3. Selection criteria for co-surfactant
% wt of unsaturated fat, 88.2 % wt of unsaturated fat, and 98.6 % wt
saturated fat, respectively. The main fatty acids in canola oil were re­ Soy phospholipid, glycerol, and propylene glycol are three
ported to be 62.4 % wt of oleic acid and 20.1 % wt of linoleic acid, while commonly used co-surfactants with a short to medium hydrophobic
the principal fatty acids in sunflower oil were 15.3 % wt of oleic acid and chain, weak amphiphilicity, and terminal small hydrophilic group [37].
71.2 % wt of linoleic acid [34]. By contrast, the major fatty acids in These features allow them to disturb the packing and long-range order of
coconut oil were composed of 53.1 % wt of octanoic acid and 30.8 % wt surfactant molecules and improve droplet curvature and interfacial
of capric acid based on the manufacturer. Therefore, the solubility of fluidity in the formation of nanoemulsions. Additionally, combining
curcumin in three major fatty acids was measured as the controls: hydrophilic surfactants and hydrophobic co-surfactants can create
octanoic acid ≫ linoleic acid > oleic acid, which matches the curcumin smaller and more uniform oil droplets in nanoemulsions with improved
solubility in these oils: coconut oil (mainly containing octanoic acid) ≫ stability [38]. Thus, molecular structure, the ability to lower interfacial
sunflower oil (linoleic acid) > canola oil (oleic acid). This difference in tension, and HLB value are three key factors in the screening of co-
curcumin solubility in three fatty acids was probably associated with surfactants [38,39]. Soy phospholipid (HLB: 4–7 [40]) can lower the
their hydrophile-lipophile balance (HLB) values (Table 1). The HLB of a interfacial tension of its aqueous solutions more effectively [41] than
surfactant measures its balance of hydrophilic and lipophilic moieties to other co-surfactants such as glycerol [42] and propylene glycol [43].
indicate the type of emulsions (o/w or w/o) that the surfactant is suit­ Consequently, soy phospholipid was used as the co-surfactant in the
able for [22]. Compared to the other two fatty acids, the shorter carbon system.
chain in octanoic acid reduced its hydrophobicity and brought about a
higher HLB value of 5.8 to dissolve curcumin. The higher affinity of C– –C 3.4. Cloud point determination and selection criteria for the aqueous
bond for water than C–C bond may cause a higher HLB value and cur­ phase
cumin solubility in linoleic acid than in oleic acid [35], despite similar
estimated HLB values of these two fatty acids. This indicates that The dissolution of nonionic surfactants in water (e.g., RH 40 in this
selecting an oil with a suitable fatty acid profile and molecular structure study) is achieved through the hydrogen bonding between water mol­
can effectively increase the solubility of curcumin or other lipophilic ecules and the repeating oxyethylene units in RH 40 [44]. The hydrogen
drugs. Hence, coconut oil was selected as the oil phase in this study. bonding would be damaged with increasing temperature, causing RH 40
dehydration, the change in solubility behaviour of RH 40 from hydro­
philic to lipophilic, and then phase separation at a certain temperature,
3.2. Selection criteria for surfactant referred to as cloud point (CP). As seen in Fig. 1(A), the CP temperature
of 20 % wt RH 40 aqueous solution was higher than 100 ◦ C and thus salts
Tween 40, Tween 80, PEG 200, PEG 400, PEG 600, and RH 40 are should be introduced into the systems to lower the CP.
Screening of salts depends on their Hofmeister series or lyotropic
Table 1 sequence [45]. The kosmotropic anions (anti-solvents and salting out
Curcumin solubility in different oil phases and surfactants with HLB values. (a-f effect) in a sequence of Hofmeister series (citrate3- > tartrate2- > sul­
represent that under the same letter, there is statistical insignificances at P > phate2- > hydrogen phosphate2- > chromate- > carbonate2- > chloride-)
0.05, n = 3.).
can hydrate better with water molecules, which in turn causes the
Vehicle Component Curcumin solubility HLB oxyethylene units in RH 40 to be more hydrophobic to separate from the
type (mg/mL)
aqueous phase at lower temperatures (i.e., a lower CP). While the cha­
Oil Sunflower 1.340 ± 0.039c / otropic anions (co-solvents and salting in effect) in a sequence of hex­
oil
d
afluorophosphate- > thiocyanate- > perchlorate- > iodide- > nitrate- >
Canola oil 1.209 ± 0.049 /
Coconut oil 2.891 ± 0.032a /
bromide- show weak hydration ability and make the nonionic surfac­
Octanoic 1.506 ± 0.006b 5.8 (estimated by the tants more hydrophilic, which promotes the solubilisation of RH 40 even
acid method of Davies [65]) at higher temperatures (a higher CP). Therefore, citrate3- was selected as
Oleic acid 0.501 ± 0.046f 1.0 (estimated) the anion in the salt.
Linoleic acid 0.829 ± 0.041e 1.0 (estimated)
The anti-solvent cations in a sequence of Hofmeister series (tet­
Surfactant Tween 40 47.369 ± 2.245c 15.6
Tween 80 51.823 ± 1.924b 15.0 ramethylammonium+ > sodium ion+ > potassium ion+ > lithium ion+)
PEG 200 47.501 ± 1.482c 18.1 [66] salt out of nonionic surfactants, lowering the CP [45]. The co-solvent
PEG 400 43.285 ± 2.590d 11.3 [67] cations (guanidinium+ > barium ion2+ > calcium ion2+ > magnesium
PEG 600 42.658 ± 0.867d 12 [68] ion2+) function in the opposite way. Considering food safety and cost,
RH 40 81.142 ± 0.846a 14–16 [36]
sodium ion was chosen as the cation and thus sodium citrate was

4
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

Fig. 1. Change in CP (A) with increasing concentration of citrate buffer at pH 4.5 and (B) with the addition of RH 40. (a-c mean that under the same letter, there is
statistical insignificances at P > 0.05, n = 3.).

selected as the salt added into the aqueous. On the other hand, curcumin > 100 nm and PDI > 0.2 were referred to as non-nanoemulsions. As seen
is sensitive to pH and shows better stability at acidic pH [6]. A pH of less in Table S1 and Fig. 2(D, E, F), the oil droplet size distribution of
than 4.0 is potentially damaging to the teeth [46]. Therefore, instead of samples at Smix 0.3 was in the range of 18 to 47 nm. With increasing
sodium citrate, citrate buffer at pH 4.5 (containing citric acid and so­ Smix, nanoemulsions with adjustable oil droplet size from ~ 17 to 103
dium citrate) would be selected as the aqueous phase candidate. nm can be produced. This should not be simply explained as the addition
In Fig. 1(A), the CP of RH 40 decreased from over 100 ◦ C to 75–88 ◦ C of co-surfactants would cause an increase in droplet size. Instead, when
with increasing citrate buffer concentration from 0 % to 2 % wt (0.084 observing the samples with different Smix values but the same compo­
M). A similar reduction in CP was achieved by using 6 % wt NaCl so­ nent mass ratio (e.g., 0.3–15, 0.5–15, and 0.7–15), the droplet size
lution as an aqueous phase [26], demonstrating that citrate can lower decreased from 19.56 ± 0.26 nm, 18.83 ± 0.23 nm, to 18.45 ± 0.22 nm,
the CP of nonionic surfactants more efficiently than chloride ions due to respectively. The droplet size of formulation number 21 also reduced
the salt kosmotrope and chaotrope explained above. With increasing RH from 23.64 ± 0.06 nm at Smix 0.3, 21.76 ± 0.36 nm at Smix 0.5, to
40 concentration, the CP reduced slightly and the time duration for 20.88 ± 0.21 nm at Smix at 0.7. For most formulations, the reduction in
cloudy state completion increased, as shown in Fig. 1(B). This is pre­ droplet size with increasing Smix was because adding co-surfactants can
sumably due to less available non-associated water molecules to hydrate increase the fluidity of interfacial film, the rate of intermicellar ex­
the oxyethylene units in RH 40 at higher RH 40 contents [44]. Curcumin change, and the curvature of oil droplets, resulting in the formation of
is sensitive to heat and starts partial degradation at around 50 ◦ C in smaller oil droplets [49]. For some formulations such as formulation
methanolic solutions and 150 ◦ C in oily solutions [47]. To reduce cur­ numbers 13 and 14, the excess addition of co-surfactants may lead to the
cumin thermal degradation and costs, the CP of RH 40 should be lower formation of micelles and vesicles, causing nanoemulsion destabilisation
than the boiling point of water but higher than the storage/shipping and the formation of transition-state emulsions and non-nanoemulsions.
temperature of nanoemulsions. Otherwise, the nanoemulsions with a CP For samples with the same Smix, oil droplet size increased with the oil
close to or lower than the storage/shipping temperature would undergo phase ratio (Table S1). The oil droplet size of samples was controlled by
phase separation during storage and shipping, causing nanoemulsions the component mass ratio and Smix in a complex manner.
destabilisation [48]. In this study, the CP of RH 40 was adjusted to The effect of Smix on PDI was not as evident as that on droplet size,
75–88 ◦ C and met the criteria above. Consequently, 2 % wt citrate buffer as shown in Fig. 2(G, H, I) and Table S1. However, for some formula­
at pH 4.5 would be used as the aqueous phase. tions like 13 and 14, an excess of co-surfactants also increased the PDI of
samples, presumably due to the nanoemulsion destabilisation
mentioned above. Fig. 2(J, K, L) exhibit the negative zeta potential of
3.5. Pseudo-ternary phase diagrams and storage stability nonpolar oil droplets covered by soy phospholipids and nonionic RH 40,
which is presumably caused by the adsorption of hydroxyl ions at the o/
As shown in Fig. 2(A), (B), and (C), blank samples without curcumin w interface via hydrogen bonding to the repeating oxyethylene units in
were prepared first to understand the effect of component mass ratios on RH 40 [50]. The addition of soy phospholipids elevated the absolute
samples after the PIT method and rapid cooling-dilution process. Each value of zeta potential, making the systems more stable. This is because
number in Fig. 2(A-C) represents a component mass ratio and is used as a zeta potential measurement was performed on samples by diluting with
formulation number. Smix represents the mixture of co-surfactants and milli-Q water to reach an attenuator setting of 10. The pH values of the
surfactants. Smix = 0.5 means that the molecular ratio of co-surfactant original and diluted samples were 4.7–5.0 and around 5.0, respectively,
(soy phospholipid) to surfactant (RH 40) was fixed at 0.5. The number which were higher than the isoelectric point of phospholipids ~ 4.5
and component mass ratio of nanoemulsions (red dots) varied with [51]. Hence, oil droplets became more negatively charged with
Smix. A larger nanoemulsion existing zone implies a more stable system increasing Smix.
to support the formation of nanoemulsions [12], and thus 0.3 and 0.5 The nanoemulsions were then kept in darkness for 4 weeks at 4 ◦ C
were considered the candidate Smix values. and room temperature without additional pasteurisation to evaluate
The detailed information and statistical analyses of nanoemulsions their storage stability based on the oil droplet size variation (Fig. 3).
and transition-state emulsions are summarised in Table S1 in the Sup­ Droplet size hardly changed with time, whereas PDI was more sensitive
plementary Materials, in which Sample name 0.3–9 represents the to long-term storage. The storage stability of nanoemulsions was
sample prepared with Smix 0.3 and formulation number 9. The samples possibly due to the inhibition effect of citric acid in citrate buffer on the
with droplet size < 100 nm and polydispersity (PDI) < 0.2 were regar­ growth of bacteria due to its Ca2+ chelating activity [52]. Considering
ded as nanoemulsions. The samples with droplet size < 100 nm, 0.2 < the nanoemulsion existing zone and the effects of Smix on droplet size,
PDI < 0.3, or 100 nm < droplet size < 200 nm, PDI < 0.2 were PDI, electrostatic stability (zeta potential), and storage stability, 0.5 was
considered as transition-state emulsions. The samples with droplet size

5
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

Fig. 2. Pseudo-ternary phase diagrams with Smix at 0.3 (A), Smix 0.5 (B), and Smix 0.7 (C) for developing nanoemulsions. The Z-average of the nanoemulsions and
emulsions in the transition state with Smix 0.3 (D), Smix 0.5 (E), and Smix 0.7 (F). The PDI of the samples above with Smix 0.3 (G), Smix 0.5 (H), and Smix 0.7 (I). The zeta
potential of the samples above with Smix 0.3 (J), Smix 0.5 (K), and Smix at 0.7 (L). Non-nanoemulsions are excluded in (D-L).

selected as the optimal Smix. weak scattering explained in the Introduction. Considering the PDI
Fig. 3(C) and (G) reflect that the appearance of blank nanoemulsions stability and appearance in Fig. 3(D) and (G), Samples 0.5–12, 0.5–13,
was strongly affected by the oil droplet size. With increasing droplet size and 0.5–14 would be used in the following curcumin loading
from 18.83 ± 0.23 nm to 84.49 ± 1.00 nm, the blank nanoemulsions experiments.
changed from transparent to hazy and milky appearance, due to the

6
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

Fig. 3. Storage stability of blank nanoemulsions prepared at Smix 0.3 (A, B), 0.5 (C, D) and 0.7 (E, F) and kept at 4 ◦ C and room temperature in darkness for 4 weeks
without additional pasteurisation. (G) The appearance of freshly prepared blank nanoemulsions at Smix 0.5. (Room temperature is abbreviated as r.t.).

3.6. Curcumin loading capacity and location oil: Smix ratio of 2: 1, loading curcumin caused a statistically significant
increase in droplet size (P < 0.05), implying curcumin molecules
In Table S2 in the Supplementary Materials, all curcumin-loaded migrated into the oil droplets rather than micelles. Sample 0.5–13 was
samples showed loading efficiency > 95 %, electrostatic stability (zeta somewhere in between. The impact of curcumin content on PDI followed
potential < -25 mV), monodispersed oil droplets (PDI < 0.3), adjustable the similar trend above, in Fig. 4(B). It also implies curcumin loading
droplet size distribution from about 25 to 129 nm, and acceptable pH in tended to elevate the polydispersity of nanoemulsions, no matter which
the range of 4.7–5.2. With increasing oil phase ratio from 6.7 % to 13.3 vehicle dominated the systems, micelles or oil droplets. In Fig. 4(C) and
% wt, the maximum curcumin loading capacity of Sample 0.5–14, (D), Sample name like 0.5–12-2C represents the sample with Smix 0.5,
0.5–13, and 0.5–12 reached 4.76, 6.20, and 9.53 mg/mL, respectively, Formulation number 12, and curcumin content 2 mg/mL. The appear­
indicating the critical role of oil phase content in tuning curcumin ance of curcumin-loaded nanoemulsions became more orange with
loading. higher curcumin loading. When curcumin loading was fixed at 2 mg/mL,
Fig. 4(A) shows the different trends of droplet size as a function of different formulations showed various transparency, due to the droplet
curcumin loading for three samples. In the presence of an excess Smix size-dependent scattering mentioned above.
(Sample 0.5–14 with an oil: Smix ratio of 1: 2), non-adsorbed Smix (RH To investigate the effect of curcumin content on the nanoemulsions,
40 and soy phospholipids) would exist as micelles in the aqueous phase. the samples containing 2 and 4 mg/mL of curcumin were selected in the
Micelles are thermodynamically stable and enable to incorporate hy­ following functional property measurements.
drophobic curcumin within their structures [12], and thus the droplet
size hardly increased with curcumin loading. For Sample 0.5–12 with an

7
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

suffered curcumin degradation during room-temperature storage, from


4.2 mg/mL to ~ 3.6 mg/mL in the 1st week and 3.4 mg/mL in the 4th
week, a drop of ~ 19 % in total from initial to week 4. During storage at
4 ◦ C, the curcumin retention of the above samples reduced to ~ 3.7 mg/
mL in the 1st week and 3.6 mg/mL in the 4th week (around 14 %
decrease). For the oil droplet-dominated formulation (Sample 0.5-12-
4C), its curcumin retention decreased from ~ 4.2 mg/mL to 3.9 mg/
mL in the 1st week and 3.8 mg/mL in the 4th week (9.5 % drop),
regardless of storage temperature. The micelle-oil droplet balanced
samples (0.5-13-4C) were somewhere in between. This indicates that the
low-dose samples and the oil droplet-dominant high-dose ones can adapt
to a broader storage temperature range (i.e., better storage stability),
while the micelle-dominant high-dose samples preferred a lower storage
temperature. Compared to surfactant micelles, oil droplets stabilised by
surfactants showed better protection of curcumin from degradation,
probably due to the double isolation (oil phase plus surfactant layer).
The IR spectra of components and nanoemulsions are shown in
Fig. S1(A) and (B) in the Supplementary Materials, respectively. After
encapsulation by the PIT method and cooling-dilution process, the
nanoemulsions still exhibited the characteristic bands of curcumin,
including the band at 1634 cm− 1 attributed to the overlapping stretch­
ing vibrations of alkenes (C––C) and carbonyl (C– –O) groups, band at
1280 cm− 1 ascribed to the bending vibration of the ν(C-O) phenolic
group, and band at 1515 cm− 1 assigned to the in-plane bending vibra­
tions around aliphatic δ(CC-C) and δ(CC– –O), in-plane bending vibra­
tions around aromatic δ(CC-H) of keto and enol configurations, and
stretching vibrations around aromatic ν(C–C) bonds of keto and enolic
form of curcumin [53]. In Fig. S1(B), the bands of coconut oil, RH 40,
and soy phospholipids consisted of 1462 cm− 1 originating from the
–C–H deformation vibrations in CH2 and CH3 groups (bending vibra­
tions), 1745 cm− 1 ascribed to the stretching vibrations of the C– –O
groups in esters, and 2958, 2925, and 2854 cm− 1 attributed to the
–C–H stretching vibrations in –CH3 and CH2 groups in triglycerides
[54]. This indicates that each ingredient, especially curcumin, was well
retained in the nanoemulsions during storage.

3.8. Photostability of curcumin nanoemulsions

Curcumin degrades under light exposure and UV has a more signif­


icant effect on curcumin photodegradation than daylight [7]. Therefore,
the photostability of free curcumin and curcumin-loaded nanoemulsions
under UV radiation is shown in Fig. 6(A). 24-hour UV radiation reduced
the retention of curcumin to 66–72 % in nanoemulsions and ~ 30 % in
raw curcumin, suggesting that encapsulation may partially shield UV
radiation, causing lower degradation in the encapsulated samples. The
similar degradation percentages of micelle-dominant (Sample 0.5-14-
2C) and oil droplet-dominant formulations (0.5-12-2C) imply that the
Fig. 4. Change in (A) droplet size and (B) PDI as a function of curcumin delivery vehicle hardly affected the photostability of curcumin. Major
loading. Changes in nanoemulsion appearance with (C) curcumin loading for curcumin decomposition seemingly occurred in the first five hours in
Sample 0.5–13, and (D) component mass ratio at the same curcumin loading, 2 Fig. 6(A). However, some curcumin photodegradation products present
mg/mL. (a-h represent statistical significances at P < 0.05, n = 3. The numbers a yellow colour and remain detected at 425 nm [27], which can be
on top of the Saturated column mean the saturated curcumin contents loaded in considered as the limitation of absorbance spectra for curcumin
the samples.). measurement.
Additionally, the DPPH inhibition activity of samples before and
3.7. Storage stability and IR spectra of curcumin nanoemulsions after 24-hour radiation was tested to evaluate the effect of UV radiation
on curcumin antioxidant activity. In Fig. 6(B), the antioxidant activities
Storage stability of nanoemulsions includes the variations of droplet of curcumin-loaded nanoemulsions after 24-hour UV radiation were as
size, PDI (physical stability), and curcumin retention (chemical stabil­ high as the fresh ones, while raw curcumin displayed much lower
ity) with storage condition and time. As seen in Fig. 5(A) and (B), both antioxidant activity after radiation by half. When the DPPHCurcumin
formulations containing 2 and 4 mg/mL curcumin showed physical sample-DPPHBlank sample, the antioxidant activities of curcumin nano­
stability during 4-week storage at 4 ◦ C and room temperature in dark­ emulsions were statistically similar. This suggests that after 24-hour
ness. In Fig. 5(C), the samples with low-dose curcumin (2 mg/mL) photodecomposition, the degradation products of curcumin still had
retained most of the curcumin during storage at 4 ◦ C and room tem­ good radical scavenging activity. This is because the antioxidant
perature. Interestingly, storage temperature and delivery vehicle mechanisms of curcumin include phenoxide radicals formed via electron
affected the curcumin retention of samples with high-dose curcumin transfer and carbon-centred radicals via direct hydrogen abstraction
simultaneously. The micelle-dominated formulation (Sample 0.5-14-4C) [55]. The photodegradation products of curcumin probably lost the

8
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

Fig. 5. Storage stability of curcumin-loaded nanoemulsions kept at 4 ◦ C and room temperature in darkness for 4 weeks without additional pasteurisation. (Room
temperature is abbreviated as r.t.).

9
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

Fig. 6. Photostability (A) and DPPH radical scavenging activity (B) of raw curcumin and curcumin-loaded nanoemulsions against 24-hour UV radiation. (Curcumin is
abbreviated as CUR. a-e represent that under the same letter, there is statistical insignificances at P > 0.05, n = 3.).

Fig. 7. Chemical stability of curcumin against gastric (A), intestinal (B) and PBS incubation (C). DPPH radical scavenging activity of curcumin before (D) and after 3-
hour gastric (E), intestinal (F), and PBS incubation (G). (Curcumin is abbreviated as CUR and normalised samples mean DPPHCurcumin sample-DPPHBlank sample. a-h
indicate that under the same letter, there is statistical insignificances at P > 0.05, n = 3.).

β-diketone moiety but retained the active hydroxyl methoxyphenyl 3.9. Curcumin retention against various physiological conditions
group after UV radiation, which is also crucial for antioxidant activity
[7,56]. This well-retained antioxidant activity of curcumin after radia­ Curcumin decomposes in neutral and alkaline environments i.e., in­
tion indicates a great potential for nanoemulsions as an oral delivery testinal and physiological conditions. [6]. As seen in Fig. 7(A), (B), and
system with a longer shelf life. (C), raw curcumin and curcumin nanoemulsions were dispersed in SGF

10
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

at pH 3.0, SIF at pH 7.0, and PBS at pH 7.4 for 3 h at 37 ◦ C. To fully 3.10. In vitro GI digestion and intestinal absorption
expose curcumin to the surrounding environments, all the samples
(including raw curcumin) were diluted 125 times with the media. The in vitro GI digestion and intestinal absorption of curcumin are
Compared to raw curcumin, the encapsulated curcumin showed higher shown in Fig. 8(A) and (B). The release (bioaccessibility) curves of
retention up to > 93 % and > 87 % during 3-hour gastric and PBS in­ nanoencapsulated curcumin showed an ultra-fast release in the initial
cubations, respectively. Note that the pancreatic lipase activity was gastric phase followed by a relatively slow gastric release. This specific
fixed at 16 units/mL in the intestinal digestion samples to slow down the “Γ” shape-like release profile is poorly fitted by zero-order, Higuchi or
lipid hydrolysis and observe curcumin decomposition. Fig. 7(B) implies first-order models, and is considered as a rapid and substantial release
that lipid hydrolysis by lipase probably damaged the integrity of oil drug delivery system. The slight reduction in the release curve in the
droplets and RH 40 micelles, causing a continuous curcumin release and intestinal phase is attributed to the degradation of curcumin under
exposure to neutral pH, and a gradual decrease in curcumin retention neutral pH.
during simulated intestinal digestion. By comparing the raw curcumin In the gastric phase, raw and nanoencapsulated curcumin exhibited
retentions in the SIF at pH 7.0 to those in PBS at pH 7.4, we found in the completely different bioaccessibility of ~ 1.6 % and ~ 99 %, respec­
intestinal phase, bile salts might emulsify raw curcumin and form mi­ tively, due to the aqueous insolubility of raw curcumin [8]. Note that the
celles [57], which alleviated curcumin degradation in the first two pancreatic lipase activity was set at 400 units/mL in the final intestinal
hours. Whereas a neutral pH led to the destabilisation of curcumin after digestion samples, to simulate human lipolysis following a modified
3 h. INFOGEST protocol [25]. During 2-hour intestinal digestion, most
The DPPH inhibition activity of samples before and after 3-hour in­ nanoemulsions showed good curcumin bioaccessibility up to 81 %,
cubation was measured to understand the impact of various physiolog­ except for Sample 0.5–14-2C which was a low-dose micelle-dominant
ical conditions on curcumin antioxidant activity, in Fig. 7(D-G). The formulation with bioaccessibility of ~ 67 %. This further indicates the
antioxidant activity of blank nanoemulsions presumably resulted from advantage of oil droplet dominant formulations in curcumin stabilisa­
the phenolic compounds of coconut oil [58]. Normalised samples mean tion. The bioaccessibility of raw curcumin increased slightly to ~ 19 %,
DPPHCurcumin sample-DPPHBlank sample. In water, the normalised DPPH presumably due to the emulsification and dissolution effect of bile salts
scavenging activities of encapsulated samples were slightly lower than on lipophilic compounds like curcumin [57]. At the endpoint of intes­
that of raw CUR, presumably because one phenoxy group in curcumin tinal digestion, coconut oil droplets should be hydrolysed into a free
with free radical scavenging activity formed molecularly hydrogen fatty acid-monoglyceride mixture (lipid digesta), while RH 40 was less
bonding with the head group of soy phospholipids [59]. DPPH assay is effectively hydrolysed [62]. Thus, curcumin was possibly loaded in the
strongly pH-dependent and presents lower results under acidic pH [60], lipid digesta droplets stabilised by partially hydrolysed RH 40
which caused the lower scavenging activities of samples in SGF (pH 3.0) molecules.
than those in water. Under the unfavourable SIF and PBS conditions for The in vitro intestinal absorption was simulated with a dialysis model
curcumin, the DPPH scavenging activity of 0.5-14-2C was lower than [28], implying that the size of delivery vehicles would play a key role in
those of 0.5-12-2C and 0.5-13-2C, as the curcumin in 0.5-14-2C tended the amount of curcumin transported. During 3-hour intestinal absorp­
to stay in the aqueous surfactant micelles and thus became more tion, the low-dose oil droplet-dominant formulations (Samples 0.5–12-
accessible to the physiological pH than other two samples. Compared to 2C and 0.5–13-2C) exhibited less curcumin transport, probably owing to
PBS, lipolysis in SIF caused higher release, exposure, and degradation of their relatively large droplet size and low curcumin concentration gra­
curcumin. However, the normalised DPPH scavenging activities were dients during dialysis. Raw curcumin had significantly faster transport
similar in SIF and PBS (33–37 %). This is possibly because nano­ because it was emulsified in small-size bile salt micelles. However, this
encapsulation alleviated curcumin degradation during in vitro intestinal does not mean bile salt micelle was an ideal vehicle. The apparent
incubation, and the degradation products of curcumin at physiological permeability coefficient (Papp) and absorption ratio of curcumin in
pH (bicyclopentadione, vanillin, ferulic acid, etc.) still had the active different samples are summarised in Fig. 8(B) and Table S3 in the
hydroxyl methoxyphenyl group for antioxidant activity [61]. Supplementary Materials. All the nanoemulsions showed continuous
curcumin transport, indicating that the encapsulated curcumin
remained stable and accumulated after 4-hour digestion and 20-hour

Fig. 8. The bioaccessibility of curcumin during in vitro GI digestion (green and yellow parts) and the normalised content of curcumin in PBS-T buffer diffusing from
the GI digestion samples of nanoemulsions or raw curcumin in the dialysis model (blue part) for 3 h (A) and for 20 h (B). The data were linearly fitted where the slope
of the straight line represents the apparent permeability coefficient (Papp). (Curcumin is abbreviated as CUR.). (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

11
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

intestinal transport. The raw curcumin decomposed in bile salt micelles Declaration of competing interest
under unfavourable physiological pH. Sample 0.5–13-4C transported
more curcumin than others, owing to its high curcumin dosage and oil The authors declare that they have no known competing financial
droplet-dominant feature. Sample 0.5–14-2C had the second highest interests or personal relationships that could have appeared to influence
curcumin in vitro absorption ratio, indicating that the small-size RH 40 the work reported in this paper.
micelles had a higher cross-membrane transport ability than oil droplets
and could stabilise curcumin better than bile salt micelles. Sample Data availability
0.5–13-6C had the saturated curcumin loading and oil droplet-dominant
formulation, whose less transport suggests that the saturated loading Data will be made available on request.
might cause curcumin to distribute in the outside layer of oil droplets or
micelles, causing curcumin destabilisation. The absorption ratio of free Acknowledgements
fatty acid was reported to be 26.62 ± 7.26 % using an Ex vivo rat small
intestine model [28]. In this study, the highest absorption ratio of oil- This project was supported by ARC Discovery grant (DP200100642).
soluble curcumin was found to be 3.36 ± 0.24 % in Sample 0.5–13- Sophie Kwon was a Taste of Research scholarship recipient. The authors
4C. The difference in absorption ratios may be due to the complex would like to acknowledge Dr Nicholas Bedford for access to Malvern
structure of intestinal epithelial cell layers which can generate Zetasizer Nano.
membrane-associated fatty acid binding proteins to promote the uptake
of fatty acids and oil-soluble nutrients [63]. Therefore, a Caco-2 cell Appendix A. Supplementary material
monolayer model will be applied in future to simulate the in vitro in­
testinal absorption of curcumin-loaded nanoemulsions [64]. Supplementary data to this article can be found online at https://doi.
org/10.1016/j.molliq.2023.123720.
4. Conclusion
References
The selection criteria for oil phase (coconut oil), surfactant (RH 40),
co-surfactant (soy phospholipid), aqueous phase (2 % wt citrate buffer at [1] H. Sung, et al., Global cancer statistics 2020: GLOBOCAN estimates of incidence
and mortality worldwide for 36 cancers in 185 countries, CA Cancer J. Clin. 71 (3)
pH 4.5), and Smix (0.5) were proposed via systematic studies to prepare (2021) 209–249.
nanoemulsions by the PIT method. After curcumin was loaded, the [2] G. Wang, et al., Curcumin sensitizes carboplatin treatment in triple negative breast
nanoemulsions had 4-week storage stability without the need for pre­ cancer through reactive oxygen species induced DNA repair pathway, Mol. Biol.
Rep. 49 (4) (2022) 3259–3270.
servatives and pasteurisation. The nanoemulsions exhibited curcumin [3] S. Eghbaliferiz, M. Iranshahi, Prooxidant Activity of Polyphenols, Flavonoids,
loading capacity of up to 9.53 mg/mL, adjustable transparency with Anthocyanins and Carotenoids: Updated Review of Mechanisms and Catalyzing
mean droplet size from 26 to 129 nm and low polydispersity < 0.29, and Metals, Phytother. Res. 30 (9) (2016) 1379–1391.
[4] B. Kocaadam, N. Şanlier, Curcumin, an active component of turmeric (Curcuma
electrostatic stability (zeta potential < -26 mV). Decreasing the mass longa), and its effects on health, Crit. Rev. Food Sci. Nutr. 57 (13) (2017)
ratio of Smix to coconut oil shifted the nanoemulsions from micelle- 2889–2895.
dominant formulation to oil droplet-dominant formulation and caused [5] H. Zong, et al., Curcumin inhibits metastatic progression of breast cancer cell
through suppression of urokinase-type plasminogen activator by NF-kappa B
the migration of curcumin from micelles to oil droplets. The use of
signaling pathways, Mol. Biol. Rep. 39 (2012) 4803–4808.
nanoemulsions also increased the photostability of curcumin by 36–42 [6] Q. Ye, et al., On improving bioaccessibility and targeted release of curcumin-whey
% and kept the antioxidant activity unchanged after 24-hour UV radi­ protein complex microparticles in food, Food Chem. 346 (2021), 128900.
[7] M.L.R. del Castillo, et al., Stabilization of curcumin against photodegradation by
ation, owing to the retained active phenolic groups in the encapsulated
encapsulation in gamma-cyclodextrin: A study based on chromatographic and
samples. Compared to free curcumin, the nanoencapsulated curcumin spectroscopic (Raman and UV–visible) data, Vib. Spectrosc 81 (2015) 106–111.
showed ~ 11 %, 24 %, and 57 % higher retention and 10 %, 12 %, and [8] Q. Ye, et al., Digestion of curcumin-fortified yogurt in short/long gastric residence
17 % higher antioxidant activity after 3-hour simulated gastric, intes­ times using a near-real dynamic in vitro human stomach, Food Chem. 372 (2022),
131327.
tinal, and physiological incubations, respectively. Nanoemulsions [9] B. Zheng, D.J. McClements, Formulation of more efficacious curcumin delivery
exhibited significantly higher bioaccessibility than free curcumin during systems using colloid science: enhanced solubility, stability, and bioavailability,
in vitro GI digestion, while free curcumin displayed a faster apparent Molecules 25 (12) (2020) 2791.
[10] B. Wahlang, Y.B. Pawar, A.K. Bansal, Identification of permeability-related hurdles
permeability coefficient in the first three hours, due to size-dependent in oral delivery of curcumin using the Caco-2 cell model, Eur. J. Pharm. Biopharm.
cross-membrane transport in the dialysis model. However, the high- 77 (2) (2011) 275–282.
dose oil droplet-dominant nanoemulsions had a higher curcumin ab­ [11] T. Feng, et al., Liposomal curcumin and its application in cancer, Int. J. Nanomed.
(2017) 6027–6044.
sorption ratio than the other samples, due to the higher curcumin con­ [12] A. Jintapattanakit, H.M. Hasan, V.B. Junyaprasert, Vegetable oil-based
centration gradient and protective effect of encapsulation. The proposed nanoemulsions containing curcuminoids: Formation optimization by phase
formulation also had higher curcumin retention and antioxidant activity inversion temperature method, J. Drug Deliv. Sci. Technol. 44 (2018) 289–297.
[13] C. Qian, D.J. McClements, Formation of nanoemulsions stabilized by model food-
during storage and various in vitro physiological incubations. The actual
grade emulsifiers using high-pressure homogenization: Factors affecting particle
intestinal absorption and lymphatic transport of the curcumin-fortified size, Food Hydrocoll. 25 (5) (2011) 1000–1008.
nanoemulsions remained unknown under the current experimental [14] S.N. Kale, S.L. Deore, Emulsion micro emulsion and nano emulsion: a review,
System. Rev. Pharmacy 8 (1) (2017) 39.
conditions and will be further studied with a Caco-2 cell monolayer
[15] A. Karthikeyan, N. Senthil, T. Min, Nanocurcumin: A promising candidate for
model in our next study. A three-dimensional tumour spheroid model therapeutic applications, Front. Pharmacol. 11 (2020) 487.
incorporating a typical breast cancer cell line (4 T1 cells) will be [16] S.D. Mandal, S. Mandal, J. Patel, Brain targeting efficiency of Curcumin loaded
leveraged to evaluate the anti-cancer activity and cytotoxicity of the mucoadhesive microemulsion through intranasal route, J. Pharm. Investig. 46
(2016) 179–188.
nanoemulsion digesta permeating from the Caco-2 cell model. [17] J.B. Aswathanarayan, R.R. Vittal, Nanoemulsions and their potential applications
in food industry, Front. Sustain. Food Syst. 3 (2019) 95.
CRediT authorship contribution statement [18] A. Molet-Rodríguez, et al., Incorporation of antimicrobial nanoemulsions into
complex foods: A case study in an apple juice-based beverage, LWT 141 (2021),
110926.
Qianyu Ye: Conceptualization, Methodology, Investigation, Visual­ [19] U.M. Musazzi, et al., Emulsion versus nanoemulsion: how much is the formulative
ization, Formal analysis, Writing – original draft. Sophie Kwon: shift critical for a cosmetic product? Drug Deliv. Transl. Res. 8 (2018) 414–421.
[20] N. Ahmad, et al., Preparation of a novel curcumin nanoemulsion by ultrasonication
Investigation, Methodology. Zi Gu: Supervision, Writing – review & and its comparative effects in wound healing and the treatment of inflammation,
editing. Cordelia Selomulya: Supervision, Writing – review & editing, RSC Adv. 9 (35) (2019) 20192–20206.
Project administration, Funding acquisition.

12
Q. Ye et al. Journal of Molecular Liquids 394 (2024) 123720

[21] G. Ren, et al., Nanoemulsion formation by the phase inversion temperature method [45] B. Kang, et al., Hofmeister Series: Insights of Ion Specificity from Amphiphilic
using polyoxypropylene surfactants, J. Colloid Interface Sci. 540 (2019) 177–184. Assembly and Interface Property, ACS Omega 5 (12) (2020) 6229–6239.
[22] R. Strey, Phase behavior and interfacial curvature in water-oil-surfactant systems, [46] A. Reddy, et al., The pH of beverages in the United States, J. Am. Dent. Assoc. 147
Curr. Opin. Colloid Interface Sci. 1 (3) (1996) 402–410. (4) (2016) 255–263.
[23] E. Rivera-Pérez, et al., Encapsulation of spray-dried curcumin nanoemulsions to [47] N.A. Siddiqui, Evaluation of thermo sensitivity of curcumin and quantification of
develop a supplement with ingredients for the control of osteoarthritis, J. Drug ferulic acid and vanillin as degradation products by a validated HPTLC method,
Deliv. Sci. Technol. 82 (2023), 104299. Pak. J. Pharm. Sci. 28 (1) (2015) 299–305.
[24] G.R. Vaz, et al., Development of nasal lipid nanocarriers containing curcumin for [48] G.C. Kalur, S.R. Raghavan, Anionic wormlike micellar fluids that display cloud
brain targeting, J. Alzheimers Dis. 59 (3) (2017) 961–974. points: rheology and phase behavior, J. Phys. Chem. B 109 (18) (2005) 8599–8604.
[25] M. Minekus, et al., A standardised static in vitro digestion method suitable for [49] J. Eastoe, M.J. Hollamby, L. Hudson, Recent advances in nanoparticle synthesis
food–an international consensus, Food Funct. 5 (6) (2014) 1113–1124. with reversed micelles, Adv. Colloid Interface Sci. 128–130 (2006) 5–15.
[26] H.M. Hasan, J. Leanpolchareanchai, A. Jintapattanakit, Preparation of virgin [50] W. Liu, et al., Formation and stability of paraffin oil-in-water nano-emulsions
coconut oil nanoemulsions by phase inversion temperature method, Adv. Mat. Res. prepared by the emulsion inversion point method, J. Colloid Interface Sci. 303 (2)
1060 (2015) 99–102. (2006) 557–563.
[27] B.H. Lee, et al., Changes in chemical stability and bioactivities of curcumin by [51] R. Pichot, R.L. Watson, I.T. Norton, Phospholipids at the interface: current trends
ultraviolet radiation, Food Sci. Biotechnol. 22 (2013) 279–282. and challenges, Int. J. Mol. Sci. 14 (6) (2013) 11767–11794.
[28] Z. Hu, et al., Intestinal absorption of DHA microcapsules with different [52] S. Brul, P. Coote, Preservative agents in foods: Mode of action and microbial
formulations based on ex vivo rat intestine and in vitro dialysis models, Food resistance mechanisms, Int. J. Food Microbiol. 50 (1) (1999) 1–17.
Funct. 14 (4) (2023) 2008–2021. [53] T.M. Kolev, et al., DFT and experimental studies of the structure and vibrational
[29] J.-S. Baek, C.-W. Cho, Surface modification of solid lipid nanoparticles for oral spectra of curcumin, Int. J. Quantum Chem 102 (6) (2005) 1069–1079.
delivery of curcumin: Improvement of bioavailability through enhanced cellular [54] T. Oniszczuk, et al., Impact of storage temperature and time on Moldavian
uptake, and lymphatic uptake, Eur. J. Pharm. Biopharm. 117 (2017) 132–140. dragonhead oil–spectroscopic and chemometric analysis, Open Chem. 17 (1)
[30] J.E. Moore Jr., C.D. Bertram, Lymphatic System Flows, Annu. Rev. Fluid Mech. 50 (2019) 609–620.
(1) (2018) 459–482. [55] M.L. Del Prado-Audelo, et al., Formulations of curcumin nanoparticles for brain
[31] T.H. Kim, et al., Preparation and characterization of water-soluble albumin-bound diseases, Biomolecules 9 (2) (2019) 56.
curcumin nanoparticles with improved antitumor activity, Int. J. Pharm. 403 (1–2) [56] H.H. Tc̵ nnesen, J.V. Greenhill, Studies on curcumin and curcuminoids. XXII:
(2011) 285–291. Curcumin as a reducing agent and as a radical scavenger, Int. J. Pharm. 87 (1–3)
[32] P. Palumbo, et al., A general approach to the apparent permeability index, (1992) 79–87.
J. Pharmacokinet Pharmacodyn. 35 (2008) 235–248. [57] H. Zhang, et al., Conversion of bile salts from inferior emulsifier to efficient smart
[33] A.-Y. Abaut, F. Chevanne, P. Le Corre, Oral bioavailability and intestinal secretion emulsifier assisted by negatively charged nanoparticles at low concentrations,
of amitriptyline: Role of P-glycoprotein? Int. J. Pharm. 330 (1) (2007) 121–128. Chem. Sci. 12 (35) (2021) 11845–11850.
[34] R.C. Zambiazi, et al., Fatty acid composition of vegetable oils and fats. Boletim do [58] A. Marina, et al., Antioxidant capacity and phenolic acids of virgin coconut oil, Int.
Centro de Pesquisa de Processamento de Alimentos, 25 (1) (2007). J. Food Sci. Nutr. 60 (sup2) (2009) 114–123.
[35] C. Black, G.G. Joris, H.S. Taylor, The solubility of water in hydrocarbons, J. Chem. [59] Y. Niu, et al., Temperature-dependent stability and DPPH scavenging activity of
Phys. 16 (5) (1948) 537–543. liposomal curcumin at pH 7.0, Food Chem. 135 (3) (2012) 1377–1382.
[36] O.M. Feeney, et al., ‘Stealth’lipid-based formulations: Poly (ethylene glycol)- [60] A. Pękal, K. Pyrzynska, Effect of pH and metal ions on DPPH radical scavenging
mediated digestion inhibition improves oral bioavailability of a model poorly activity of tea, Int. J. Food Sci. Nutr. 66 (1) (2015) 58–62.
water soluble drug, J. Control. Release 192 (2014) 219–227. [61] O.N. Gordon, C. Schneider, Vanillin and ferulic acid: not the major degradation
[37] H. Choudhury, et al., Improvement of cellular uptake, in vitro antitumor activity products of curcumin, Trends Mol. Med. 18 (7) (2012) 361–363.
and sustained release profile with increased bioavailability from a nanoemulsion [62] J.F. Cuine, et al., Evaluation of the impact of surfactant digestion on the
platform, Int. J. Pharm. 460 (1–2) (2014) 131–143. bioavailability of danazol after oral administration of lipidic self-emulsifying
[38] I.K. Hong, S.I. Kim, S.B. Lee, Effects of HLB value on oil-in-water emulsions: formulations to dogs, J. Pharm. Sci. 97 (2) (2008) 995–1012.
Droplet size, rheological behavior, zeta-potential, and creaming index, J. Ind. Eng. [63] R.W. Schwenk, et al., Fatty acid transport across the cell membrane: regulation by
Chem. 67 (2018) 123–131. fatty acid transporters, Prostaglandins, Leukotrienes and Essential Fatty Acids
[39] D. Andrews, T. Nann, R.H. Lipson, Comprehensive nanoscience and (PLEFA) 82 (4–6) (2010) 149–154.
nanotechnology. 2019: Academic press. [64] M. Yao, et al., Design of nanoemulsion-based delivery systems to enhance intestinal
[40] F. Bot, D. Cossuta, J.A. O’Mahony, Inter-relationships between composition, lymphatic transport of lipophilic food bioactives: Influence of oil type, Food Chem.
physicochemical properties and functionality of lecithin ingredients, Trends Food 317 (2020), 126229.
Sci. Technol. 111 (2021) 261–270. [65] J.T. Davies, Interfacial phenomena. 2012: Elsevier.
[41] C. Arancibia, et al., Comparing the effectiveness of natural and synthetic [66] A.S. Suliman, R.J. Anderson, A.A. Elkordy, Norfloxacin as a model hydrophobic
emulsifiers on oxidative and physical stability of avocado oil-based nanoemulsions, drug with unique release from liquisolid formulations prepared with PEG200 and
Innov. Food Sci. Emerg. Technol. 44 (2017) 159–166. Synperonic PE/L-61 non-volatile liquid vehicles, Powder Technol. 257 (2014)
[42] B. Saleh, A. Abouel-Kasem, S. Ahmed, Quantitative analysis of wear particles 156–167.
generated by cavitation erosion in glycerol-water mixtures, Indus. Lubric. Tribol. [67] Y.-T. Zhang, et al., A novel microemulsion-based isotonic perfusate modulated by
69 (5) (2017) 627–637. Ringer’s solution for improved microdialysis recovery of liposoluble substances,
[43] I.S. Khattab, et al., Density, viscosity, surface tension, and molar volume of J. Nanobiotechnol. 16 (1) (2018) 91.
propylene glycol+water mixtures from 293 to 323K and correlations by the [68] R. Nazari-Vanani, et al., A novel self-nanoemulsifying formulation for sunitinib:
Jouyban-Acree model, Arab. J. Chem. 10 (2017) S71–S75. Evaluation of anticancer efficacy, Colloids Surf. B Biointerfaces 160 (2017) 65–72.
[44] H. Akbaş, Ç. Batıgöç, Spectrometric studies on the cloud points of Triton X-405,
Fluid Phase Equilib. 279 (2) (2009) 115–119.

13

You might also like