You are on page 1of 6

Colloids and Surfaces B: Biointerfaces 224 (2023) 113205

Contents lists available at ScienceDirect

Colloids and Surfaces B: Biointerfaces


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

Encapsulation of resveratrol within size-controlled nanoliposomes: Impact


on solubility, stability, cellular permeability, and oral bioavailability
Youjin Baek , Eun Woo Jeong , Hyeon Gyu Lee *
Department of Food and Nutrition, Hanyang University, 17 Haengdang-dong, Seongdong-gu, Seoul 133-791, South Korea

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

Keywords: This study examined the influence of the nanoliposomes (LPs) particle size on the solubility, antioxidant stability,
Particle size in vitro release profile, Caco-2 cellular transport activity, cellular antioxidant activity, and in vivo oral
Nanoliposomes bioavailability of resveratrol (RSV). LPs with sizes of 300, 150, and 75 nm were prepared using the thin-lipid film
Resveratrol
hydration method, followed by ultrasonication for 0, 2, and 10 min, respectively. Formulating small LPs (< 100
Oral administration
Poorly soluble compound
nm) was effective to enhance the solubility, in vitro release profile, cellular permeability, and cellular antioxidant
Stability activity of RSV. A similar pattern was observed for in vivo oral bioavailability. However, the size reduction of
RSV-loaded LPs did not promote the antioxidant stability of RSV, owing to their large surface area used to
interact with harsh environments. This study provides the better understanding of the appropriate particle size
range of LPs to improve their in vitro and in vivo performances of RSV as an effective carrier for oral
administration.

1. Introduction of a hydrophilic core surrounded by a lipid bilayer, are considered po­


tential lipid-based nanocarriers owing to their capacity to improve the
Resveratrol (3,5,6′ -trihydroxy-stilbene, RSV), one of the most high­ solubility, stability, and bioavailability of poorly water-soluble com­
lighted naturally occurring antioxidants, is a polyphenol phytoalexin pounds [11]. In addition, LPs have received significant attention for
generally abundant in plants such as soybean, peanuts, and grapes [1]. applications in the food industry because of their high biocompatibility
RSV exists as two geometric isomers, trans and cis-RSV; generally, and non-toxicity [12].
trans-RSV demonstrates higher therapeutic characteristics than cis-RSV Particle size of LPs is one of the main factors influencing in vivo
[2]. Previous studies have reported that RSV has the ability to prevent performances of encapsulated compounds after injection [13]. In the
obesity, inhibit cancer, and reduce the likelihood of cardiovascular pharmaceutical industry, the in vivo performance of LPs with different
disease, owing to its high antioxidant activity [3,4]. particle size ranges has been investigated to determine how the particle
However, the application of RSV is strictly restricted in the food in­ size of LPs influences their biodistribution after lateral tail vein, intra­
dustry owing to its degradation by high temperature, alkaline pH, or muscular, and penile vein injection [14–16]. On the other hand, rela­
ultraviolet light, resulting in the conversion of the trans to cis- form [5]. tively little attention in food industry has been paid to the effects of LPs
When administered orally, RSV would demonstrate low intestinal ab­ with different particle sizes on the properties of phytochemicals found in
sorption and low bioavailability due to its poor aqueous solubility food after oral administration. In general, LPs would be transported
(20–30 μg/mL) [1]. Moreover, the low bioavailability of RSV may be across the intestinal mucosa by several mechanisms after oral ingestion;
explained by its conjugation with glucuronic acid and sulfates in the endocytosis, paracellular uptake, and uptake by microfold cell (M-cell)
liver and intestinal epithelial cells after administration, resulting in [17]. Endocytosis is one of passage mechanisms through the epithelial
rapid elimination [6,7]. Several attempts have been made to overcome cell membrane via transcellular route, which would embed the LPs and
the limitations of RSV associated with solubility, stability, and form the membrane vesicles, followed by the release of compounds
bioavailability by incorporating RSV into various lipid-based nano­ within LPs into the cytoplasm [18–20]. Meanwhile, paracellular uptake
carriers, such as nanoliposomes, nanoemulsions, and nanostructured is the mechanism by which LPs pass through the intercellular space
lipid carriers [8–10]. Among them, nanoliposomes (LPs), which consist between small intestinal epithelial cells controlled by the tight junctions

* Corresponding author.
E-mail address: hyeonlee@hanyang.ac.kr (H.G. Lee).

https://doi.org/10.1016/j.colsurfb.2023.113205
Received 7 December 2022; Received in revised form 2 February 2023; Accepted 10 February 2023
Available online 12 February 2023
0927-7765/© 2023 Elsevier B.V. All rights reserved.
Y. Baek et al. Colloids and Surfaces B: Biointerfaces 224 (2023) 113205

[19,21]. Finally, M-cells, mainly located in Peyer’s patch, could uptake The ZP values of RSV-loaded size-controlled LPs ranged from − 47.7
LPs via Peyer’s patch and transport LPs directly to the lymphatic systems ± 0.7 to − 41.1 ± 5.2 mV. The ZP values in this study exhibited no
[22]. Several previous studies have reported the enhanced cellular up­ significant differences regardless of the ultrasonication time, indicating
take of bioactive compounds via those three mechanisms by nano­ that the influence of various ultrasonication times on the surface charge
encapsulation within LPs with small particle sizes because LPs with of RSV-loaded LPs was minimal. In general, a higher ZP value is related
particle size less than 100 nm have been reported to demonstrate high to better LPs stability because the ZP value illustrates the repulsion
cellular transport efficiency [23–26]. However, oral absorption could be between charged LPs, and absolute ZP values greater than 30 mV are
restricted by various physiological obstacles which could be managed by regarded as stable [11]. These results demonstrated that ultrasonication
the particular particle size range of LPs. Therefore, it is crucial to time did not influence the surface charge of RSV-loaded size-controlled
investigate the overall in vitro and in vivo characteristics of LPs with wide LPs, leading to electrostatic stability.
size ranges, including solubility, antioxidant stability, cellular proper­
ties, and in vivo oral bioavailability, for manufacturing effective orally 2.1.2. EE
administered LPs. The EE of the RSV-loaded LPs 300, 150, and 75 were 80.88 ± 0.05,
To the best of our knowledge, the impacts of the particle size of LPs 80.24 ± 0.01, and 79.82 ± 0.02%, respectively, without significant
on the in vitro and in vivo performances of RSV after orally administra­ differences, as shown in Table 1. This result indicated that the amount of
tion, which are of great importance, are not fully investigated. Consid­ RSV entrapped within the LPs was not affected by the ultrasonication
ering that the performance of RSV may differ depending on the particle time. The high EE of RSV-loaded LPs, regardless of the ultrasonication
size of the LPs, this present study aimed to investigate how RSV-loaded time, could be due to the affinity of the core materials for lipid bilayer.
LPs with different particle sizes influence RSV solubility, antioxidant Generally, hydrophobic components, entrapped within the lipid bilayers
stability, and in vivo oral bioavailability. RSV-loaded LPs consisting of L- of LPs, have a high affinity for lipid bilayers and low affinity for water,
α-phosphatidylcholine, β-sitosterol, and resveratrol were formulated resulting in high EE and low drug leakage [30]. In a previous study, the
using the thin-lipid film hydration method, and their sizes were authors formulated tacrolimus-loaded LPs with ultrasonication for 0–7
controlled by ultrasonication. The physicochemical features of the RSV- min to investigate the effects of ultrasonication time on the EE of
loaded LPs, such as particle size, polydispersity index (PDI), zeta po­ tacrolimus. The authors affirmed that the EE of tacrolimus was stable at
tential (ZP), and entrapment efficiency (EE), were investigated. In all ultrasonication times owing to its high affinity for lipid bilayers [31].
addition, the impact of size-controlled LPs on the characteristics of RSV, Hence, RSV-loaded LPs subjected to 0, 2, and 10 min of ultrasonication
such as solubility, antioxidant stability, in vitro release properties using exhibited three distinct particle sizes with high EE, indicating that these
simulated gastric fluid (SGF) and simulated intestinal fluid (SIF) models, size-controlled LPs successfully encapsulated RSV and could be consid­
cellular transport assay, cellular antioxidant assay using Caco-2 mono­ ered adequate for further analysis.
layers, and in vivo pharmacokinetic assay after oral administration to
Sprague-Dawley (SD) rats were investigated. 2.2. Solubility of RSV-loaded size-controlled LPs

2. Results and discussions The effects of LPs size on RSV solubility were demonstrated in Fig. 1.
LPs 75, 150, and 300 increased the solubility of RSV by 13.49, 12.96,
2.1. Physical properties of RSV loaded size-controlled LPs and 12.04 times more, respectively, than RSV dissolved in water, sug­
gesting that encapsulation within LPs can significantly enhance the
2.1.1. Particle size, PDI, and ZP solubility of hydrophobic compounds (p < 0.05) [11]. In addition, LPs
The physical properties such as particle size, PDI, and ZP of the RSV- 75 and 300 showed a significant difference in the aqueous solubility of
loaded size-controlled LPs were illustrated in Table 1. RSV-loaded LPs RSV, indicating that the particle size of the LPs can remarkably influence
with ultrasonication times of 0, 2, and 10 min exhibited the particle size
of 318.6 ± 15.9 (LPs 300), 152.6 ± 7.4 (LPs 150), and 73.3 ± 4.9 nm
(LPs 75), respectively. Ultrasonication time longer than 10 min was
considered to have no significant effect on reducing the particle size of
RSV-loaded LPs as predetermined in our preliminary experiments. The
PDI of RSV-loaded LPs 300, 150, and 75 were 0.636 ± 0.070, 0.353 ±
0.075, and 0.244 ± 0.008, respectively. The PDI of LPs less than 0.5
indicates highly narrow particle size distribution, and the PDI higher
than 0.5 illustrates the extensive particle size distribution [27]. Ultra­
sonication times of 2 and 10 min significantly narrowed the PDI of the
LPs compared to LPs without ultrasonication, indicating that the uni­
formity of the LPs can be enhanced by the increasing the ultrasonication
time (p < 0.05) [28]. Thus, the longer the ultrasonication time, the more
sonication energy is delivered to the liposome suspension, leading to a
smaller particle size and a homogenous size distribution [29]. These
results indicated that ultrasonication time (0, 2, and 10 min) can highly
control the particle size and PDI of RSV-loaded LPs, without modifying Fig. 1. Solubility of free RSV, LPs 75, LPs 150, and LPs 300. a-c Columns with
the lipid composition. different letters are significantly different from each other (p < 0.05).

Table 1
Physical characteristics of resveratrol loaded size-controlled nanoliposomes (LPs).
Sample Ultrasonication time (min) Target size(nm) Measured size (nm) Polydispersity index Zeta potential (mV) Entrapment efficiency (%)

LPs 300 0 300.0 318.6 ± 15.9a 0.636 ± 0.070a − 47.7 ± 0.7a 80.88 ± 0.05a
LPs 150 2 150.0 152.6 ± 7.4b 0.353 ± 0.075b − 41.1 ± 5.2a 80.24 ± 0.01a
LPs 75 10 75.0 73.3 ± 4.9c 0.244 ± 0.008c − 45.2 ± 7.1a 79.82 ± 0.02a

a-c Different letters in the same column indicate significant differences (p < 0.05)

2
Y. Baek et al. Colloids and Surfaces B: Biointerfaces 224 (2023) 113205

the solubility of RSV (p < 0.05). This result could be explained by the core materials which are susceptible to degradation during heat
fact that the smaller particle size of the LPs can generate a larger surface processing.
area of encapsulated compounds, which can affect surface reactivity,
leading to enhanced aqueous solubility [32]. A previous study reported 2.4. In vitro release profile of RSV-loaded size-controlled LPs
that RSV-loaded nanoparticles (NPs) with a particle size of 72 nm
increased RSV aqueous solubility compared to NPs with a particle size of The in vitro release profile of RSV-loaded size-controlled LPs in the
269 nm owing to their larger contact surface area between RSV and SGF environment followed by that in the SIF environment was demon­
aqueous media [1]. Thus, this study confirmed that LPs 75, which strated to investigate the characteristics of RSV being released in gastro-
exhibited smaller particle sizes, appeared to have a higher aqueous intestinal tract (Fig. 3). Free RSV was rapidly released within the first
solubility of RSV than LPs 300. 2 h, and the cumulative amount of RSV reached 81.40% after 12 h.
Meanwhile, the cumulative amounts of RSV released from LPs 75, 150,
2.3. Antioxidant stability of RSV-loaded size-controlled LPs and 300 were 52.90%, 45.44%, and 41.32%, respectively, after 12 h,
and LPs 75 released a significantly higher amount of RSV than LPs 150
The antioxidant activity of RSV before and after thermal treatment at and 300 (p < 0.05). Within the first 30 min, RSV that was free or
90 ℃ for 4 h in its free and encapsulated forms with different particle attached to the surface of the bilayer membranes immediately diffused
sizes was compared in Fig. 2. At 0 h, the DPPH radical scavenging effect in the SGF environment, leading to the burst release of RSV [35]. After
of free RSV was 91.19% and that of RSV-loaded size-controlled LPs was 30 min, all LPs samples exhibited delayed release of RSV within 2 h,
ranged from 84.71% to 88.62%, indicating that antioxidant activity of which could be explained by two different mechanisms. One was pepsin,
free RSV was higher than that of RSV encapsulated within size- which cannot diffuse through bilayer membranes [36]. Another possi­
controlled LPs. Similar findings with the present study have been re­ bility was the aggregation or precipitation of LPs structures due to
ported that RSV-loaded LPs and proniosomes exhibited lower DPPH electrostatic interactions at low pH, which prevents RSV from being
radical scavenging effect than free RSV [4,33]. These results could be released [37]. Similar results were obtained by Liu et al., who observed
explained by the location of hydrophobic compounds in LPs as afore­ aggregation of LPs structures as the ZP value of LPs changed to zero at
mentioned [30]. As RSV is not prone to move across the liposomal acidic pH [38].
membrane due to its high affinity for lipid bilayer, it would be released On the other hand, the release profile of RSV in the SIF environment
from LPs slowly; therefore, it requires more time to interact with the demonstrated an entirely different trend owing to the presence of
reaction medium, resulting in lower antioxidant activity than its free pancreatin, which degrades the LPs structures primarily composed of
form. phospholipids. Likewise, previous literature has affirmed that LPs
On the other hand, the antioxidant activity of RSV at 4 h showed showed a higher release of curcumin in SIF than in SGF environments
entirely different trend. The antioxidant activities of all samples due to the lipid-degrading effect of lipolytic enzymes in pancreatin [39].
exhibited a decreasing trend with the increase of thermal treatment Moreover, LPs 75 in this study released a significantly higher amount of
time, indicating that 90 ℃ was high enough to induce the thermal RSV than LPs 150 and 300, indicating that the larger surface area of LPs
degradation of RSV and LPs. The antioxidant activity of free RSV was 75 was more susceptible to the reaction with pancreatin to disrupt the
only 45.94%, while that of RSV-loaded size-controlled LPs was greater phospholipid bilayer. Therefore, the release profile of RSV within LPs in
than 56.63% after 4 h of thermal treatment, indicating that LPs could the SGF environment suggests that encapsulating RSV within LPs may
enhance the thermal stability of bioactive compounds [34]. The differ­ protect RSV from acidic or enzymatic degradation along the way to the
ence in the antioxidant activities depending on the particle size of LPs small intestine. In particular, LPs with smaller particle sizes allow for the
was clearly observed; the DPPH radical scavenging effect of LPs 75 was release of higher amount of RSV into the SIF environment, which is
significantly lower than that of LPs 300 (p < 0.05). Similarly, a previous sequentially associated with bioavailability.
study illustrated that NPs with a particle size of 87 nm showed signifi­
cantly lower stability of RSV than those with a particle size of 258 nm 2.5. Cellular transport activity of RSV-loaded size-controlled LPs
[1]. LPs with smaller particle sizes were highly exposed to the heat
owing to their larger surface area, which provides more sites for RSV The Papp of free RSV, LPs 75, 150, and 300 across the Caco-2
interaction, leading to higher degradation than LPs with larger particle monolayers was demonstrated in Fig. 4. Compared to free RSV, the
sizes. Thus, it is tempting to presume that the preparation of LPs with Papp value of RSV increased by 1.53, 1.31, and 1.21 times through
large particle sizes may efficiently enhance the antioxidant stability of nanoencapsulation within LPs 75, 150, and 300, respectively. The

Fig. 2. DPPH radical scavenging activity of free RSV, LPs 75, LPs 150, LPs 300 Fig. 3. In vitro release profile of free RSV, LPs 75, LPs 150, and LPs 300. a-c
at 90 ℃ for 4 h. a-c Means with different letters indicate significant differ­ Means with different letters are significantly different from each
ences (p < 0.05). other (p < 0.05).

3
Y. Baek et al. Colloids and Surfaces B: Biointerfaces 224 (2023) 113205

Fig. 4. Apparent permeability coefficient (Papp) of free RSV, LPs 75, LPs 150, Fig. 5. Cellular antioxidant activity of free RSV, LPs 75, LPs 150, and LPs 300
and LPs 300 transported across Caco-2 cell monolayers. a-c Means with different in Caco-2 monolayer. a-c Columns with different letters indicates significant
letters are significantly different (p < 0.05). differences (p < 0.05).

results indicated that cellular transport of RSV significantly increased by monolayers than those with particle sizes over 100 nm [21,44]. Be­
nanoencapsulation within the LPs (p < 0.05). In addition, the Papp value sides, reduced particle size leads to an increased surface area between
of LPs was dramatically higher than that of LPs 150 and 300 (p < 0.05). entrapped RSV and 2′ ,7′ -dichlorodihydrofluorescein (DCFH), resulting
In general, free RSV in an aqueous solution with poor solubility is in decreased 2′ ,7′ -dichlorofluorescein (DCF) formation, followed by
limited to reaching the cell surfaces, while RSV-loaded LPs improved the increased antioxidant activity in the Caco-2 monolayer [32]. Therefore,
cellular permeability of RSV owing to their improved solubility [40]. LPs 75, which has a smaller size and higher surface area than other LPs
RSV-loaded LPs 75, which had the smallest particle size and the highest formulations, could have more chances to permeate the Caco-2 mono­
solubility of RSV, demonstrated the highest cellular permeability among layer and interact with DCFH, leading to lower degradation of DCFH to
other LPs formulations. DCF and prevention of free radical formation.
Additionally, nanocarriers can be transported into epithelial cell
layers via two routes. One is the paracellular route, which is controlled
2.7. In vivo bioavailability of RSV loaded size-controlled LPs
by tight junctions between the enterocytes. The transport of nano­
carriers via the paracellular route is limited because the width between
The mean plasma RSV concentrations after the oral administration of
enterocyte cells only increases to 20 nm, even though the tight junction
free RSV and RSV-loaded size-controlled LPs are shown in Fig. 6. The
opens entirely [21,41]. Nanocarriers with a positive charge, such as
plasma concentration of RSV was 0.404 ± 0.249 μg/mL at 1 h,
chitosan nanoparticles or chitosan-coated LPs, have positive effects on
decreased rapidly from 1 to 8 h, and then became undetectable after 8 h.
the opening of tight junction, which has a negative charge due to the
In contrast, the concentration of RSV in plasma was highly elevated by
presence of negatively charged sialic acid groups on the mucus mem­
incorporating within size-controlled LPs. The pharmacokinetic param­
brane [42,43]. In other words, RSV-loaded size-controlled LPs prepared
eters of free RSV, LPs 75, 150, and 300 after oral administration are
in our study may not affect the opening of intercellular tight junctions
demonstrated in Table 2. The Cmax and Tmax values of free RSV were
because of RSV-loaded size-controlled LPs’ negative charge. Another
0.404 ± 0.249 μg/mL and 1.08 h, respectively. In contrast, the Cmax of
mechanism is endocytosis via a transcellular route [19]. Generally,
LPs 75, 150, and 300 were 1.486 ± 0.427, 0.910 ± 0.174, and 0.684
nanocarriers can be easily transported into enterocytes when their
± 0.346 μg/mL, and Tmax values of LPs 75, 150, and 300 were at 2.67,
particle size is 50–100 nm, whereas nanocarriers with a particle size of
2.17, and 1.83 h, respectively. Moreover, the oral bioavailability of RSV
20–500 nm can be transported into M-cells, which can be expressed by a
increased by 12.531, 1.733 and 2.801 times by encapsulation within LPs
co-culture with Raji B cells [44]. This result indicated that modifying LPs
75, compared to that by encapsulation within free RSV, LPs 150, and
to a smaller size could potentially enhance the permeability of bioactive
300, respectively, based on the AUC0–12 values.
materials across the Caco-2 monolayer, leading to an increased chances
In general, free RSV exhibited a peak plasma concentration at
of reaching the bloodstream.

2.6. Cellular antioxidant activity of RSV-loaded size-controlled LPs

As shown in Fig. 5, the cellular antioxidant activity (CAA) unit of free


RSV, LPs 75, 150, and 300 in Caco-2 cell was 20.54 ± 2.63, 41.35
± 2.01, 35.03 ± 1.56, and 30.54 ± 1.69%, respectively, indicating that
the CAA of RSV can be increased significantly by encapsulation within
size-controlled LPs compared to that of free RSV (p < 0.05). Moreover,
among the size-controlled LPs, the CAA value of LPs 75 was significantly
higher than that of LPs 300 (p < 0.05). This result indicated that the
enhanced CAA value of RSV by incorporation into LPs with smaller
particle sizes may be attributed to the improved solubility and perme­
ability of RSV through the Caco-2 monolayer. The increase in CAA with
a decrease in the particle size of LPs may be attributed to two sequential
reasons: the particle size-dependent endocytosis mechanism and
increased surface area of LPs. As aforementioned, nanocarriers with Fig. 6. Mean plasma resveratrol concentration versus time profiles after oral
particle size of 50–100 nm are more prone to endocytosis in Caco-2 administration of free RSV, LPs 75, LPs 150, and LPs 300.

4
Y. Baek et al. Colloids and Surfaces B: Biointerfaces 224 (2023) 113205

Table 2
Pharmacokinetic parameters, including Cmax, Tmax, AUC0–12, and MRT after oral administration of free RSV, LPs 75, LPs 150, and LPs 300 at 20 mg/kg body weight.
Formulation Free RSV LPs 75 LPs 150 LPs 300

Dose mg /kg body weight 20 20 20 20


Pharmacokinetic parameters
***
Peak plasma concentration (Cmax) [μg/mL] 0.404 ± 0.249 1.486 ± 0.427 0.910 ± 0.174* 0.684 ± 0.346
Time to reach peak plasma concentration (Tmax) [h] 1.08 ± 0.49 2.67 ± 1.03** 2.17 ± 0.98* 1.83 ± 0.41
Area under curve0–12 (AUC0–12) [μg/mL×h] 0.825 ± 0.306 10.340 ± 0.866*** 5.968 ± 1.025*** 3.692 ± 1.918***
Mean residence time (MRT) [h] 1.875 ± 0.378 7.363 ± 2.111*** 6.090 ± 1.530*** 5.779 ± 1.347***

Data presented as mean ± standard deviation (n = 6). * Values are significantly different from free resveratrol at the level of p < 0.05. ** Values are significantly
different from free resveratrol at the level of p < 0.01. *** Values are significantly different from free resveratrol at the level of p < 0.001.

approximately 0.5–1 h, and the oral bioavailability of RSV has been release profile, cellular properties, and in vivo oral availability, while
reported to be less than 1% [45]. Moreover, as mentioned above, RSV large LPs (≥300 nm) could increase the thermal stability of RSV. Our
demonstrates low bioavailability due to highly rapid glucuronic acid findings may have significant implications for understanding that
and sulfate conjugation in the intestine and liver [6,7]. Our study appropriate particle size range of LPs for improving the solubility,
affirmed that encapsulation within LPs could greatly enhance the oral antioxidant stability, cellular properties, and in vivo oral bioavailability
bioavailability of RSV, mainly owing to the increment of RSV solubility, of RSV to establish them as an effective delivery system.
stability in the gastrointestinal tract, and absorption properties [11,46].
Among the size-controlled LPs, LPs 75 showed the highest Cmax, Tmax CRediT authorship contribution statement
AUC0–12, and MRT, indicating that controlling the particle size is a
crucial factor in improving the oral bioavailability of RSV. The differ­ Youjin Baek: Conceptualization, Investigation, Methodology, Data
ences in RSV oral bioavailability between size-controlled LPs may be curation, Visualization, Writing – original draft, Writing – review &
explained by two main mechanisms. One is the increased surface area of editing. Eun Woo Jeong: Conceptualization, Data curation. Hyeon Gyu
LPs 75, which had more chances of increasing the solubility and Lee: Conceptualization, Supervision, Project administration, Funding
permeation of RSV to the intestinal environment, resulting in increased acquisition.
oral bioavailability [11,46]. The other is the particle size of the LPs that
can be taken up by lymphatic vessels [22,47]. LPs with particle size Declaration of Competing Interest
smaller than 10 nm are more prone to escape from the lymphatic vessel,
while nanocarriers with particle sizes ranging from 10 to 100 nm are The authors declare that they have no known competing financial
appropriate for delivering bioactive compounds to the lymphatic vessel, interests or personal relationships that could have appeared to influence
thus avoiding first-pass metabolism in the liver [48]. Consequently, the work reported in this paper.
decreased first-pass metabolism of bioactive compounds may lead to
increased bioavailability [49]. A similar result was observed in a pre­ Data Availability
vious study which showed that erythropoietin-loaded LPs with a particle
size of 100 nm were more effectively absorbed in rats after oral Data will be made available on request.
administration than those with a particle size of 200 nm [49]. Therefore,
LPs 75, which exhibited particle sizes smaller than 100 nm, may be a Acknowledgement
promising lipid-based nanocarrier to overcome the low oral bioavail­
ability of RSV. This work was supported by a National Research Foundation of
Korea (NRF) grant funded by the Korean government (MSIT)
2.8. Conclusion (No.2021R1A2C2013460).

In this study, encapsulation of RSV in the LPs system was an effective Appendix A. Supporting information
strategy to overcome the limitations of RSV, including its low solubility,
stability, and bioavailability. Moreover, RSV-loaded LPs were success­ Supplementary data associated with this article can be found in the
fully formulated in different particle sizes (approximately 75, 150, and online version at doi:10.1016/j.colsurfb.2023.113205.
300 nm) with the same concentration of lipid compositions, enabling
the comparison of solubility, antioxidant stability after thermal treat­ References
ment, cellular permeability, cellular antioxidant activity, and in vivo oral
bioavailability. All in vitro and in vivo performances of RSV showed a [1] J.H. Chung, J.-S. Lee, H.G. Lee, Resveratrol-loaded chitosan–γ-poly (glutamic acid)
nanoparticles: optimization, solubility, UV stability, and cellular antioxidant
size-dependent manner. The results obviously demonstrated that LPs activity, Colloids Surf. B Biointerfaces 186 (2020), 110702, https://doi.org/
with small particle size (75 nm) were positively related to enhancing the 10.1016/j.colsurfb.2019.110702.
solubility and in vitro release profile due to their large surface area to [2] I. Choi, N. Li, Q. Zhong, Enhancing bioaccessibility of resveratrol by loading in
natural porous starch microparticles, Int. J. Biol. Macromol. 194 (2022) 982–992,
interact with the reaction environment compared with LPs with large https://doi.org/10.1016/j.ijbiomac.2021.11.157.
particle size (300 nm). Besides, LPs with a particle size of 75 nm [3] E.C. Alexandre, F.B. Calmasini, M.G. de Oliveira, F.H. Silva, C.P.V. da Silva, D.
enhanced the cellular permeability and cellular antioxidant activity of M. Andre’, F.C. Leonardo, M.A. Delbin, E. Antunes, Chronic treatment with
resveratrol improves overactive bladder in obese mice via antioxidant activity, Eur.
RSV compared with those with 300 nm, indicating that their primary J. Pharmacol. 788 (2016) 29–36, https://doi.org/10.1016/j.ejphar.2016.06.017.
permeation route may be mediated by endocytosis. This tendency was [4] H. Jeong, K.J. Samdani, D.H. Yoo, D.W. Lee, N.H. Kim, I.S. Yoo, J.H. Lee,
consistent with the result of the pharmacokinetic study after oral Resveratrol cross-linked chitosan loaded with phospholipid for controlled release
and antioxidant activity, Int. J. Biol. Macromol. 93 (2016) 757–766, https://doi.
administration, indicating that LPs with small particle size are more
org/10.1016/j.ijbiomac.2016.09.018.
effective in improving the bioavailability of RSV. In contrast, when the [5] Š. Zupančič, Z. Lavrič, J. Kristl, Stability and solubility of trans-resveratrol are
particle size of LPs was increased to 300 nm, the antioxidant stability strongly influenced by pH and temperature, Eur. J. Pharm. Biopharm. 93 (2015)
was improved, which affirmed that a larger particle size is more 196–204, https://doi.org/10.1016/j.ejpb.2015.04.002.
[6] A.Y. Berman, R.A. Motechin, M.Y. Wiesenfeld, M.K. Holz, The therapeutic potential
acceptable to enhance the stability of RSV than a smaller particle size. of resveratrol: a review of clinical trials, Npj Precis. Oncol. 1 (1) (2017) 1–9,
Therefore, small LPs (<100 nm) could enhance the solubility, in vitro https://doi.org/10.1038/s41698-017-0038-6.

5
Y. Baek et al. Colloids and Surfaces B: Biointerfaces 224 (2023) 113205

[7] A.R. Neves, M. Lucio, J.L. Lima, S. Resis, Resveratrol in medicinal chemistry: a [28] J.H. Nam, S.-Y. Kim, H. Seong, Investigation on physicochemical characteristics of
critical review of its pharmacokinetics, drug-delivery, and membrane interactions, a nanoliposome-based system for dual drug delivery, Nanoscale Res. Lett. 13 (1)
Curr. Med. Chem. 19 (11) (2012) 1663–1681, https://doi.org/10.2174/ (2018) 1–11, https://doi.org/10.1186/s11671-018-2519-0.
092986712799945085. [29] A.K. Sahu, V. Jain, Screening of process variables using Plackett–Burman design in
[8] M. Huang, C. Liang, C. Tan, S. Huang, R. Ying, Y. Wang, Z. Wang, Y. Zhang, the fabrication of gedunin-loaded liposomes, Artif. Cells Nanomed. Biotechnol. 45
Liposome co-encapsulation as a strategy for the delivery of curcumin and (5) (2017) 1011–1022, https://doi.org/10.1080/21691401.2016.1200057.
resveratrol, Food Funct. 10 (10) (2019) 6447–6458, https://doi.org/10.1039/ [30] E.A.A. Azim, S.A. Elkheshen, R.M. Hathout, M.A. Fouly, N.M.E. Hoffy, Augmented
c9fo01338e. in vitro and in vivo profiles of brimonidine tartrate using gelatinized-core liposomes,
[9] H. Cheng, H. Zhang, D. Li, H. Duan, L. Liang, Impact of oil type on the location, Int. J. Nanomed. 17 (2022) 2753, https://doi.org/10.2147/ijn.s370192.
partition and chemical stability of resveratrol in oil-in-water emulsions stabilized [31] Y. Dai, R. Zhou, L. Liu, Y. Lu, J. Qi, W. Wu, Liposomes containing bile salts as novel
by whey protein isolate plus gum Arabic, Food Hydrocoll. 109 (2020), 106119, ocular delivery systems for tacrolimus (FK506): in vitro characterization and
https://doi.org/10.1016/j.foodhyd.2020.106119. improved corneal permeation, Int. J. Nanomed. 8 (2013) 1921, https://doi.org/
[10] C. Houacine, D. Adams, K.K. Singh, Impact of liquid lipid on development and 10.2147/ijn.s44487.
stability of trimyristin nanostructured lipid carriers for oral delivery of resveratrol, [32] J.B. Min, E.S. Kim, J.-S. Lee, H.G. Lee, Preparation, characterization, and cellular
J. Mol. Liq. 316 (2020), 113734, https://doi.org/10.1016/j.molliq.2020.113734. uptake of resveratrol-loaded trimethyl chitosan nanoparticles, Food Sci.
[11] Y. Zhu, M. Wang, J. Zhang, W. Peng, C.K. Firempong, W. Deng, Q. Wang, S. Wang, Biotechnol. 27 (2) (2018) 441–450, https://doi.org/10.1007/s10068-017-0272-2.
F. Shi, J. Yu, X. Xu, W. Zhang, Improved oral bioavailability of capsaicin via [33] P.A. Shruthi, H.A. Pushpadass, M.E.E. Franklin, S.N. Battula, N.L. Naik,
liposomal nanoformulation: preparation, in vitro drug release and Resveratrol-loaded proniosomes: Formulation, characterization and fortification,
pharmacokinetics in rats, Arch. Pharm. Res. 38 (2015) 512–521, https://doi.org/ LWT 134 (2020), 110127, https://doi.org/10.1016/j.lwt.2020.110127.
10.1007/s12272-014-0481-7. [34] Y. Liu, D. Liu, L. Zhu, Q. Gan, X. Le, Temperature-dependent structure stability and
[12] B.S. Esposto, P. Jauregi, D.R. Tapia-Blácido, M. Martelli-Tosi, Liposomes vs. in vitro release of chitosan-coated curcumin liposome, Food Res. Int. 74 (2015)
chitosomes: encapsulating food bioactives, Trends Food Sci. Technol. 108 (2021) 97–105, https://doi.org/10.1016/j.foodres.2015.04.024.
40–48, https://doi.org/10.1016/j.tifs.2020.12.003. [35] H.J. Je, E.S. Kim, J.-S. Lee, H.G. Lee, Release properties and cellular uptake in
[13] M. Leitgeb, Ž. Knez, M. Primožič, Sustainable technologies for liposome Caco-2 cells of size-controlled chitosan nanoparticles, J. Agric. Food Chem. 65 (50)
preparation, J. Supercrit. Fluids 165 (2020), 104984, https://doi.org/10.1016/j. (2017) 10899–10906, https://doi.org/10.1021/acs.jafc.7b03627.
supflu.2020.104984. [36] S. Hu, M. Niu, F. Hu, Y. Lu, J. Qi, Z. Yin, W. Wu, Integrity and stability of oral
[14] Y.-D. Dong, E. Tchung, C. Nowell, S. Kaga, N. Leong, D. Mehta, L.M. Kaminskas, B. liposomes containing bile salts studied in simulated and ex vivo gastrointestinal
J. Boyd, Microfluidic preparation of drug-loaded PEGylated liposomes, and the media, Int. J. Pharm. 441 (1–2) (2013) 693–700, https://doi.org/10.1016/j.
impact of liposome size on tumour retention and penetration, J. Liposome Res. 29 ijpharm.2012.10.025.
(1) (2019) 1–9, https://doi.org/10.1080/08982104.2017.1391285. [37] W. Liu, J. Liu, L.J. Salt, M.J. Ridout, J. Han, P.J. Wilde, Structural stability of
[15] G. Lou, G. Anderluzzi, S. Woods, C.W. Roberts, Y. Perrie, A novel microfluidic- liposome-stabilized oil-in-water pickering emulsions and their fate during in vitro
based approach to formulate size-tuneable large unilamellar cationic liposomes: digestion, Food Funct. 10 (11) (2019) 7262–7274, https://doi.org/10.1039/
formulation, cellular uptake and biodistribution investigations, Eur. J. Pharm. c9fo00967a.
Biopharm. 143 (2019) 51–60, https://doi.org/10.1016/j.ejpb.2019.08.013. [38] W. Liu, J. Liu, W. Liu, T. Li, C. Liu, Improved physical and in vitro digestion stability
[16] G. Phan, A. Herbet, S. Cholet, H. Benech, J.R. Deverre, E. Fattal, Pharmacokinetics of a polyelectrolyte delivery system based on layer-by-layer self-assembly
of DTPA entrapped in conventional and long-circulating liposomes of different size alginate–chitosan-coated nanoliposomes, J. Agric. Food Chem. 61 (17) (2013)
for plutonium decorporation, J. Control. Release 110 (1) (2005) 177–188, https:// 4133–4144, https://doi.org/10.1021/jf305329n.
doi.org/10.1016/j.jconrel.2005.09.029. [39] M. Hasan, K. Elkhoury, C.J. Kahn, E. Arab-Tehrany, M. Linder, Preparation,
[17] S. Haddadzadegan, F. Dorkoosh, A. Bernkop-Schnurch, Oral delivery of therapeutic characterization, and release kinetics of chitosan-coated nanoliposomes
peptides and proteins: Technology landscape of lipid-based nanocarriers, Adv. encapsulating curcumin in simulated environments, Molecules 24 (10) (2019)
Drug Deliv. Rev. 182 (2022), 114097, https://doi.org/10.1016/j. 2023, https://doi.org/10.3390/molecules24102023.
addr.2021.114097. [40] J.-J. Ma, Y.-G. Yu, S.-W. Yin, G.-H. Tang, X.-Q. Yang, Cellular uptake and
[18] M.S. de Almeida, E. Susnik, B. Drasler, P. Taladriz-Blanco, A. Petri-Fink, B. Rothen- intracellular antioxidant activity of zein/chitosan nanoparticles incorporated with
Rutishauser, Understanding nanoparticle endocytosis to improve targeting quercetin, J. Agric. Food Chem. 66 (48) (2018) 12783–12793, https://doi.org/
strategies in nanomedicine, Chem. Soc. Rev. 50 (9) (2021) 5397–5434, https://doi. 10.1021/acs.jafc.8b04571.
org/10.1039/d0cs01127d. [41] J. Sheng, L. Han, J. Qin, G. Ru, R. Li, L. Wu, D. Cui, P. Yang, Y. He, J. Wang, N-
[19] D.J. McClements, Edible lipid nanoparticles: digestion, absorption, and potential trimethyl chitosan chloride-coated PLGA nanoparticles overcoming multiple
toxicity, Prog. Lipid Res. 52 (4) (2013) 409–423, https://doi.org/10.1016/j. barriers to oral insulin absorption, ACS Appl. Mater. Interfaces 7 (28) (2015)
plipres.2013.04.008. 15430–15441, https://doi.org/10.1021/acsami.5b03555.
[20] M.R.I. Shishir, N. Karim, V. Gowd, X. Zheng, W. Chen, Liposomal delivery of [42] H.M. Ezzat, Y.S. Elnaggar, O.Y. Abdallah, Improved oral bioavailability of the
natural product: a promising approach in health research, Trends Food Sci. anticancer drug catechin using chitosomes: design, in-vitro appraisal and in-vivo
Technol. 85 (2019) 177–200, https://doi.org/10.1016/j.tifs.2019.01.013. studies, Int. J. Pharm. 565 (2019) 488–498, https://doi.org/10.1016/j.
[21] E. Roger, F. Lagarce, E. Garcion, J.P. Benoit, Lipid nanocarriers improve paclitaxel ijpharm.2019.05.034.
transport throughout human intestinal epithelial cells by using vesicle-mediated [43] I. Fiebrig, S.E. Harding, S.S. Davis, Sedimentation analysis of potential interactions
transcytosis, J. Control. Release 140 (2) (2021) 174–181, https://doi.org/10.1016/ between mucins and a putative bioadhesive polymer, Prog. Coll. Polym. Sci. 94
j.jconrel.2009.08.010. (1994) 66–73, https://doi.org/10.1007/bfb0115603.
[22] P. Pandya, P. Giram, R.P. Bhole, H.I. Chang, S.Y. Raut, Nanocarriers based oral [44] C. Dima, E. Assadpour, S. Dima, S.M. Jafari (Eds.), In Vitro Assays for Evaluating
lymphatic drug targeting: Strategic bioavailability enhancement approaches, the Release of Nanoencapsulated Food Ingredients, Vol. 5, Elsevier, Amsterdam,
J. Drug Deliv. Sci. Technol. 64 (2021), 102585, https://doi.org/10.1016/j. Netherlands, 2020, https://doi.org/10.1016/b978-0-12-815665-0.00004-7.
jddst.2021.102585. [45] A. Amri, J.C. Chaumeil, S. Sfar, C. Charrueau, Administration of resveratrol: what
[23] Y. Sun, J. Chi, X. Ye, S. Wang, J. Liang, P. Yue, H. Xiao, X. Gao, Nanoliposomes as formulation solutions to bioavailability limitations, J. Control. Release 158 (2)
delivery system for anthocyanins: physicochemical characterization, cellular (2012) 182–193, https://doi.org/10.1016/j.jconrel.2011.09.083.
uptake, and antioxidant properties, LWT 139 (2021), 110554, https://doi.org/ [46] M. Takahashi, S. Uechi, K. Takara, Y. Asiin, K. Wada, Evaluation of an oral carrier
10.1016/j.lwt.2020.110554. system in rats: bioavailability and antioxidant properties of liposome-encapsulated
[24] M. Yang, H. Zhong, R. Guan, X. Liu, W. Zhang, X. Lu, J. Xu, Cellular uptake, curcumin, J. Agric. Food Chem. 57 (19) (2009) 9141–9146, https://doi.org/
transport mechanism and anti-inflammatory effect of cyanidin-3-glucoside 10.1021/jf9013923.
nanoliposomes in Caco-2/RAW264.7 co-culture model, Front. Nutr. 9 (2022), [47] S.G.M. Ong, L.C. Ming, K.S. Lee, K.H. Yuen, Influence of the encapsulation
995391, https://doi.org/10.3389/fnut.2022.995391. efficiency and size of liposome on the oral bioavailability of griseofulvin-loaded
[25] J. Xian, X. Zhong, H. Gu, X. Wang, J. Li, J. Li, Y. Wu, C. Zhang, J. Zhang, Colonic liposomes, Pharmaceutics 8 (3) (2016) 25, https://doi.org/10.3390/
delivery of celastrol-loaded layer-by-layer liposomes with pectin/trimethylated pharmaceutics8030025.
chitosan coating to enhance its anti-ulcerative colitis effects, Pharmaceutics 13 [48] R.S. Managuli, S.Y. Raut, M.S. Reddy, S. Mutalik, Targeting the intestinal
(12) (2021) 2005, https://doi.org/10.3390/pharmaceutics13122005. lymphatic system: a versatile path for enhanced oral bioavailability of drugs,
[26] R.S. Managuli, J.T.W. Wang, F.M. Faruqu, A. Pandey, S. Jain, K.T. Al-Jamal, Expert Opin. Drug Deliv. 15 (8) (2018) 787–804, https://doi.org/10.1080/
S. Mutalik, Surface engineered nanoliposomal platform for selective lymphatic 17425247.2018.1503249.
uptake of asenapine maleate: in vitro and in vivo studies, Mater. Sci Eng. C: Mater. [49] Y. Maitani, M. Hazama, Y. Tojo, N. Shimoda, T. Nagai, Oral administration of
Biol. Appl. 109 (2020), 110620, https://doi.org/10.1016/j.msec.2019.110620. recombinant human erythropoietin in liposomes in rats: influence of lipid
[27] M.M. Seyedabadi, H. Rostami, S.M. Jafari, M. Fathi, Development and composition and size of liposomes on bioavailability, J. Pharm. Sci. 85 (4) (1996)
characterization of chitosan-coated nanoliposomes for encapsulation of caffeine, 440–445, https://doi.org/10.1021/js950477m.
Food Biosci. 40 (2021), 100857, https://doi.org/10.1016/j.fbio.2020.100857.

You might also like