ArticlePDF Available

Preparation and Characterization of Lidocaine-Loaded, Microemulsion-Based Topical Gels

Authors:

Abstract and Figures

: Microemulsion-based gels (MBGs) were prepared for transdermal delivery of lidocaine and evaluated for their potential for local anesthesia. Lidocaine solubility was measured in various oils, and phase diagrams were constructed to map the concentration range of oil, surfactant, cosurfactant, and water for oil-in-water (o/w) microemulsion (ME) domains, employing the water titration method at different surfactant/cosurfactant weight ratios. Refractive index, electrical conductivity, droplet size, zeta potential, pH, viscosity, and stability of fluid o/w MEs were evaluated. Carbomer® 940 was incorporated into the fluid drug-loaded MEs as a gelling agent. Microemulsion-based gels were characterized for spreadability, pH, viscosity, and in-vitro drug release measurements, and based on the results obtained, the best MBGs were selected and subsequently subjected to ex-vivo rat skin permeation anesthetic effect and irritation studies. Data indicated the formation of nano-sized droplets of MEs ranging from 20 - 52 nm with a polydispersity of less than 0.5. In-vitro release and ex-vivo permeation studies on MBGs showed significantly higher drug release and permeation in comparison to the marketed topical gel. Developed MBG formulations demonstrated greater potential for transdermal delivery of lidocaine and advantage over the commercially available gel product, and therefore, they may be considered as potential vehicles for the topical delivery of lidocaine.
Content may be subject to copyright.
Iranian Journal of Pharmaceutical Research 21 (2022) e1: 1-21
DOI: 10.22037/ijpr.2021.115214.15256
Received: March 2021
Accepted: June 2021
Original Article
Preparation and Characterization of Lidocaine-loaded,
Microemulsion-Based Topical Gels
Mahshid Daryaba, Mehrdad Faizib, Arash Mahboubia, c* and Reza Aboofazelia, d*
aDepartment of Pharmaceutics, School of Pharmacy, Shahid Beheshti University of Medical
Sciences, Tehran, Iran. bDepartment of Pharmacology and Toxicology, School of Pharmacy,
Shahid Beheshti University of Medical Sciences, Tehran, Iran. cFood Safety Research Center,
Shahid Beheshti University of Medical Sciences, Tehran, Iran. dProtein Technology Research
Center, Shahid Beheshti University of Medical Sciences, Tehran, Iran.
* Corresponding author:
E-mail: a.mahboubi@sbmu.ac.ir; raboofazeli@sbmu.ac.irE-mail: a.mahboubi@sbmu.ac.ir; raboofazeli@sbmu.ac.ir
Abstract
Microemulsion-based gels (MBGs) were prepared for transdermal delivery of lidocaine and
evaluated for their potential for local anesthesia. Lidocaine solubility was measured in various
oils, and phase diagrams were constructed to map the concentration range of oil, surfactant,
cosurfactant, and water for oil-in-water (o/w) microemulsion (ME) domains, employing the
water titration method at dierent surfactant/cosurfactant weight ratios. Refractive index,
electrical conductivity, droplet size, zeta potential, pH, viscosity, and stability of uid o/w MEs
were evaluated. Carbomer 940 was incorporated into the uid drug-loaded MEs as a gelling
agent. Microemulsion-based gels were characterized for spreadability, pH, viscosity, and in-vitro
drug release measurements, and based on the results obtained, the best MBGs were selected and
subsequently subjected to ex-vivo rat skin permeation anesthetic eect and irritation studies. Data
indicated the formation of nano-sized droplets of MEs ranging from 20-52 nm with a polydispersity
of less than 0.5. In-vitro release and ex-vivo permeation studies on MBGs showed signicantly
higher drug release and permeation in comparison to the marketed topical gel. Developed MBG
formulations demonstrated greater potential for transdermal delivery of lidocaine and advantage
over the commercially available gel product, and therefore, they may be considered as potential
vehicles for the topical delivery of lidocaine.
Keywords: Lidocaine; Microemulsion; Microemulsion-based gel; Phase diagrams; Skin
permeation; Local anesthesia.
Introduction
Topical anesthetics, as valuable tools in
the eld of dermatology, are widely used
to control cutaneous pain associated with
medical procedures, prevent or treat chronic
conditions such as post-herpetic neuralgia,
complex regional pain syndrome, and cancer-
related pains. These compounds are expected
to cause painless, cutaneous analgesia with a
quick onset of action and sucient duration
(1-3). Based on the chemical structure of
their intermediate chain, these weak bases are
classied into aminoesters (e.g., benzocaine,
procaine, tetracaine, etc.) and aminoamides
(e.g., bupivacaine, lidocaine, prilocaine)
classes (1, 4).
Local anesthetic (LA) drugs are
commercially available in various
pharmaceutical dosage forms such as gels,
2
Daryab M et al. / IJPR 21 (2022) e1: 1-21
creams, ointments, solutions, and patches (5,
6), and many of them are available over-the-
counter without the need for a prescription.
The purpose of such transdermal formulations
is to increase skin permeability, reduce the
eective concentration, provide painless,
cutaneous analgesia and numbness with a
quick onset of action and sucient duration of
action and minimize side eects (6, 7).
Lidocaine is a low soluble and good
penetrating drug (Biopharmaceutical
Classication System, class II) with a
molecular weight of 234.3 g/mol and log p
equal to 2.84, which makes it a good candidate
for skin delivery (Scheme 1). This amide
derivative is the most commonly used as well
as an eective and reliable LA drug formulated
as topical products due to its desirable
properties such as the low risk of allergic
reactions, intermediate duration of action, and
low systemic toxicity (8-10). However, the
major problems encountered with the use of
commercial lidocaine products are late-onset
and insuciency of local anesthetic eect. It
has been reported that commercial creams or
gels are not capable of eectively delivering
lidocaine base (or its HCl salt) through the
intact skin. To overcome these problems, an
increase in the skin permeability to dermally
applied lidocaine is required (8).
Several strategies have been adopted to
enhance the skin permeability of LAs and
improve their onset and duration of action,
as well as prevent systemic absorption
and reduce side eects. Among the most
common physical techniques, one can
mention iontophoresis, sonophoresis,
magnetophoresis, electroporation,
microporation, and microneedle technologies
(9, 11 and 12). However, using these methods
is restricted because of the high cost, need for
special devices, and qualied sta (13). Other
delivery strategies include the incorporation
of LAs into innovative colloidal carrier
delivery systems such as liposomes, niosomes,
ethosomes, nanospheres, nanoparticles, and
microemulsions (5, 13).
Microemulsions (ME), rst introduced
by Hoar and Schulman (14), are transparent,
spontaneously formed, dispersed systems in
which the interfacial layer is stabilized by
a layer of surfactant molecules (usually in
combination with a co-surfactant) (15). These
transparent, low viscose, thermodynamically
stable (no tendency for occulation or
coalescence) colloidal dispersions with
droplets less than 120 nm in diameter oer
several advantages for ecient transdermal
delivery of drugs. MEs can be formulated
as water-in-oil (w/o), oil-in-water (o/w),
and bicontinuous systems (16-18). The ease
of preparation, relatively high solubilizing
capacity for a variety of hydrophilic and
lipophilic molecules because of the existence
of two microdomains in a single-phase
solution, long-term thermodynamic stability,
and good production feasibility have made
them promising drug delivery systems (19-
21). The greater amount of drug incorporated
in MEs, compared to conventional topical
formulations, could increase the ux of drug
through the skin. Moreover, enhancement of
drug solubility can increase the concentration
gradient and thermodynamic activity of
the drug, which could favor its partitioning
into the skin. The possibility of employing
ingredients for ME formulation with skin
penetration enhancing eect can also aect
the barrier function of stratum corneum (SC),
promoting permeation of drug (22-25).
Since MEs are less viscose in nature, their
low skin adherence has restricted their topical
application (26). To overcome this challenge
and retain the applied dose on the skin for
a sucient time, ME-based gel (MBG)
formulations have been, therefore, developed,
utilizing a suitable thickening agent to modify
the rheological behavior. MBGs, also known
Scheme 1. Structure of lidocaine
Scheme 1. Structure of lidocaine
3
Lidocaine-loaded, Microemulsion-based Topical Gels
as hydrogel-thickened MEs, are nanocarriers
derived from o/w MEs composed of dispersed
oil phase within a continuous aqueous phase,
which is thickened with a suitable hydrophilic
gelling agent (27, 28). By the addition of
gelling components, the application of
MEs to the skin becomes easier compared
to runny uid MEs. Various gelling agents
such as Carbopol, xanthan gum, chitosan,
poloxamer, hydroxypropyl methylcellulose,
and carrageenan have been utilized for the
preparation of MBGs (27, 28).
Based on the type of polymer used, dierent
procedures for the preparation of MBGs have
been employed. A mixture of oil, surfactant,
and cosurfactant with the dissolved drug is
added to the previously prepared hydrogel
matrix in a two-stage procedure. Alternatively,
o/w ME is prepared and then gelled by directly
dispersing a suitable thickening agent (28).
These gels have the advantages of both MEs
and hydrogels, including ease of preparation,
enhanced drug solubility and permeability,
optical clarity, longer shelf-life, water
solubility, and spreadability (29, 30). In recent
years, numerous studies have demonstrated
that MBGs are potential transdermal
delivery systems for a wide variety of drugs
commonly used in dierent skin disorders or
even systemic diseases (31, 32). Negi et al.
showed that phospholipid MBGs containing
lidocaine and prilocaine have enhanced skin
permeation and improved the analgesic eect
signicantly, compared to the commercial
cream (13), while a remarkable analgesic
activity has been observed with ropivacaine-
loaded MBGs, formulated by Transcutol
HP and Capryol 90 (33). In 2017, Okur et
al. investigated the permeation of Carbopol
940-based, benzocaine-loaded MBGs and
conrmed high permeability through the skin
with less systemic side eects with no sign of
inammation and irritation (34).
Since the dermal delivery of lidocaine is
still a concern, the suitability of MBGs for
its transdermal delivery was examined in this
investigation. Thus, this study was planned
to develop and characterize lidocaine-loaded
MBGs, formulated with pharmaceutically
acceptable components. It was hypothesized
that a rapid onset and longer duration of
anesthetic eect of lidocaine might be
produced when incorporated in MBGs.
Experimental
Materials
Lidocaine base and Cremophor RH40
(PEG-40 hydrogenated castor oil) were
supplied by Daroupakhsh Pharma Chem.
Co. (Tehran, Iran) and Osvah Pharmaceutical
Co. (Tehran, Iran), respectively. Kolliphor®
ELP (Cremophor EL, PEG-35 hydrogenated
castor oil) was purchased from BASF
(Germany). Labrasol (caprylocaproyl
macrogol-8-glycerides) and Transcutol P
(diethylene glycol monoethyl ether) were
gifted by Gattefossé (France). Olive & castor
oils were provided from Sigma Aldrich (USA)
and GPR Rectapur (France), respectively.
Triacetin (glycerol triacetate), isopropyl
myristate (IPM), polysorbate 80 (Tween 80),
polyethylene glycol 400 (PEG 400), propylene
glycol (PG), ethanol 96% (v/v), glacial acetic
acid (HPLC grade), acetonitrile (HPLC grade),
sodium hydroxide and potassium dihydrogen
phosphate were all obtained from Merck
Chemical Co. (Germany). Carbomer 934
and 940 were purchased from BF Goodrich
(USA). Metolose 90SH (hydroxypropyl
methylcellulose) was provided from Shin-
Etsu Chemical Co., Ltd. (Japan). Puried
water was prepared by a Millipore Milli-Q
water purication system (USA).
High-performance liquid chromatography
(HPLC) method
A quantitative assay of lidocaine base
was carried out by an HPLC method outlined
in the United States Pharmacopoeia (USP
41-NF 36). Chromatographic studies were
implemented on the Knauer HPLC system
(Germany), equipped with a UV detector
(Smartline 2500), pump (Smartline 1000), and
software (Chromgate V3.1.7). The separation
process was carried out on a reversed-phase
C18 column (5 μ, 250 × 4.6 mm), using a
freshly prepared and degassed mobile phase
consisting of acetonitrile (A) and water/glacial
acetic acid (930:50; pH 3.4) in the ratio of 1:4.
The ow rate was xed at 1.2 mL/min, and
all measurements were performed at room
temperature. The injection of samples was
performed on a Reodyne injector equipped
4
Daryab M et al. / IJPR 21 (2022) e1: 1-21
with a 20 μL loop. The UV detector was set at
254 nm. The calibration curve was found to be
linear in the concentration range of 10-100 μg/
mL (r2 = 0.9985).
Determination of lidocaine oil solubility
The solubility of lidocaine was evaluated in
triacetin, IPM, castor oil, and olive oil by the
shake-ask method (35). An excess amount of
lidocaine was added to 1 mL of the oil. The
mixture was then continuously stirred, using
a magnetic stirrer, at room temperature (25
℃) for 72 h in order to achieve equilibrium.
After removing the undissolved drug, samples
were centrifuged (Sigma 1-14, Osterode am
Harz, Germany) at 5000 rpm for 10 min;
the supernatant was separated and ltered
through a 0.22 μm membrane lter and then
diluted with a suitable solvent (chloroform or
ethanol 96% v/v). Finally, the amount of the
drug dissolved in each oil was assayed by a
UV spectrophotometer (UV-2601, Rayleigh,
China) at the wavelength of 263 nm, using oil
samples with known drug concentrations.
Construction of phase diagrams
Castor oil and triacetin (selected based
on the oil solubility studies), four non-ionic
surfactants (Tween 80, Labrasol, Cremophor®
EL and Cremophor RH40) and three co-
surfactants (PEG 400, Transcutol P and PG)
were chosen to construct the phase diagrams
and determine the o/w microemulsion domains.
Surfactant/co-surfactant weight ratios (Rsm)
were kept constant at dierent values of
1:1, 1:2, 2:1. Clear oil-surfactant mixtures
with various weight ratios of 1:9 to 9:1 were
prepared by weighing appropriate amounts
of each component into screw-capped vials
and mixing thoroughly at room temperature.
Samples were then titrated with small aliquots
of triple distilled water while stirring for
a sucient time to attain equilibrium. The
course of each titration was inspected visually
and through cross-polaroids for determining
the clarity and the possible formation of a
birefringent liquid crystalline phase. The
triangle diagrams were mapped with the top
apex representing a xed Rsm (1:1, 1:2, or
2:1), the right and left apices representing
the oil and water, respectively. All mixtures
which produced optically transparent, non-
birefringent solutions at relatively water-rich
parts of the phase diagrams were designated
as o/w MEs.
Preparation of lidocaine-loaded MEs
Following the determination of o/w ME
regions on the phase diagrams, those oil/
surfactant/oil systems which demonstrated
a relatively extended o/w ME area on the
phase diagrams, composed of a minimum
of 5% (w/w) oil and not more than 25%
(w/w) surfactant mixture, were selected for
drug loading. Lidocaine-loaded MEs were
prepared by the spontaneous emulsication
method. A given amount of lidocaine base was
dissolved gradually in the oil phase to which
the surfactant mixtures were then added, and
the required amount of distilled water was
nally added dropwise while stirring the
mixture gently until a transparent solution
was obtained. The formulations were stored
at room temperature and were evaluated
for clarity, drug precipitation, and phase
separation within 72 h.
Characterization of uid MEs
Refractive index (RI), pH, and conductivity
The Refractive index of drug-loaded MEs
was measured by Abbe Refractometer (2WAJ,
Bluewave Industry Co., Ltd., Shanghai,
China). The pH and electrical conductivity
of MEs were determined using a calibrated
pH-meter (744, Metrohm AG, Switzerland)
and a conductivity meter (712, Metrohm AG,
Switzerland), respectively.
Determination of particle size and zeta
potential
Mean droplet size (Z-ave), polydispersity
index (PDI), and zeta potential of ME
formulations were measured at 25 °C, using
a Malvern Zetasizer (Nano-ZS, Malvern
Instruments, Worcestershire, UK), equipped
with a Nano ZS Software for data acquisition
and analysis. Each sample was analyzed in
triplicate, and the results were reported as
mean ± SEM.
Determination of viscosity and rheological
behavior
A Brookeld DV2T cone and plate
viscometer (LV, Brookeld Engineering
5
Lidocaine-loaded, Microemulsion-based Topical Gels
Laboratories, Middlesboro, USA), equipped
with a CP-42 spindle, was used to measure the
viscosity and examine the rheological behavior
of ME formulations. To evaluate thixotropic
behavior, measurements were carried out at a
rotation speed ranging from 2 to 70 rpm for
both up curves and down curves, at 25 ± 1 °C.
Results within the 10-100% range of torque
were considered acceptable and recorded. The
shear stress (Pa) was plotted vs. shear rate
(1/s), and the viscosity was calculated based
on the slope of the linear portion of the plots.
Preparation of lidocaine-loaded MBGs
MBGs containing 5 wt% lidocaine were
fabricated by dispersing dierent amounts
of various polymers, namely Carbomer 934
(1-3 wt%), Carbomer 940 (0.5-1.5 wt%)
or Metolose 90SH (3-6 wt%), in the drug-
loaded uid MEs under magnetic stirring.
Gel characterization
Spreadability
To measure the spreadability of MBGs, a
circle with 1 cm in diameter was marked on
a glass plate. Half a gram of the test gel was
placed on the circle, and a second glass plate
was placed on the gel. A 5 g weight was put
on the upper glass plate, and after 5 min, the
weight was removed, and the diameter of the
spread gel was measured and reported (36).
pH measurement
One gram of MBGs was mixed with 99 g
distilled water and stirred thoroughly until a
uniform mixture was obtained. The pH was
measured in triplicate, using a calibrated pH
meter (744, Metrohm AG, Switzerland).
Determination of viscosity and rheological
behavior
Viscosity and rheological properties of the
MBGs were determined at 25 ± 1 °C, using
a Brookeld DV-III Ultra Programmable
Rheometer (Brookeld Engineering
Laboratories, Middlesboro, USA), attached
with spindle no. 51. Measurements were
performed at a rotation speed ranging from 0.5
to 250 rpm for both upward and downward ow
curves. Flow curves (rheograms) were plotted,
and the viscosities were then calculated based
on the slope of the linear portion of the plots.
Stability tests
Fluid MEs were stored in sealed glass vials
at 25 °C for 15 months and observed for any
macroscopic changes, including turbidity,
phase separation, drug precipitation, and color
change (37). The stability of the selected
MBGs was also evaluated for 6 months at
ambient temperature, 2-8 °C and 40 ± 2 °C
(Relative humidity: 75 ± 5%), and checked for
their appearance and viscosity. In addition, the
optimum gel formulations were centrifuged
(5702, Eppendorf AG, Hamburg, Germany) at
5000 rpm for 30 min, subsequently subjected
to seven heating/cooling cycles (24 h at 4 °C
followed by 24 h at 40 °C) and three 24-hour
freeze-thaw (FT) cycles (-5 and 25 °C) (38-
40).
In-vitro drug release
Cellulose acetate membrane
In-vitro permeation study of lidocaine-
loaded MBGs was carried out using vertical
Franz diusion cell with 1.767 cm2 eective
diusion surface area. Synthetic cellulose
acetate membrane (MW cut-o 12,000 Da),
previously soaked in phosphate buer pH
7.4 for 24 h at 2-8 °C, was placed between
the donor and receptor compartments of the
diusion cell. The receptor chamber (25 mL)
was lled with phosphate buer solution
(0.1 M, pH 7.4) and thermostated at 37 ± 0.5
°C, while continuously stirring (400 rpm). A
quantity of 200 mg of the gel was applied to
the membrane, and the donor chamber was
covered with Paralm. At predetermined time
intervals (5, 7, 10, 15, 20, 30, 45, 60, 75, 90,
105, and 120 min), an aliquot of 2 mL sample
was taken from the release medium, and the
same volume of the fresh buer was added
to the receptor chamber to maintain the sink
condition. During the test, the diusion cells
were checked for the presence of a bubble on
both sides of the membrane. The cumulative
amount of drug released from MBGs at
each time was measured. As a control, a
commercially available 5 wt% lidocaine gel
was used.
Ex-vivo permeation study
The ex-vivo permeability study protocol
was approved by the local Animal Ethics
Committee of Shahid Beheshti University of
6
Daryab M et al. / IJPR 21 (2022) e1: 1-21
Medical Sciences, Tehran, Iran (approval No:
1399.002). Male Wistar albino rats (200-250
g) were sacriced by ether inhalation. The
hair of test animals was carefully trimmed
with electrical clippers, and the full-thickness
skin was excised carefully from the abdominal
region and wiped with acetone to remove
adhering fat and connecting tissues. The
prepared skins were wrapped in aluminum foil
and stored at -20 °C for further use. Prior to the
test, the skins were kept for 30 min at ambient
temperature, then mounted between the donor
and receiver compartments of a static Franz
diusion cell, while the dermis side was in
contact with the release medium for 12 h
(41). After skin hydration and replacement
of fresh phosphate buer (pH 7.4), the skin
permeation study was carried out with the
same procedure described for the in-vitro drug
release experiments, except that the samples
were taken from the receptor chamber after 7,
10, 15, 20, 30, 45, 60, 75, 90, 105, 120, 150,
180, 210, 240, 300, 360, 480 and 600 minutes
and ltered through a 0.45 μm membrane
lter. The cumulative percentage of lidocaine
in withdrawn samples was calculated, and the
results were plotted as a function of time (in
a minute) and compared with those obtained
from the commercial gel. The ex-vivo release
prole was tted into various mathematical
models, i.e., zero-order, rst-order, Higuchi
and Korsmeyer-Peppas, in order to elucidate
the kinetic release model. All the experiments
were performed in triplicate, and the results
were reported as mean ± SEM. Data were
statistically analyzed by one-way analysis
of variance (ANOVA), followed by Tukey’s
post hoc using GraphPad Prism version 8.0.1
(GraphPad Software, Inc., USA). A 0.05 level
of probability was considered as the level of
signicant dierence (*p < 0.05: signicant,
**p < 0.01: very signicant, and ***p < 0.001:
extremely signicant).
Evaluation of the local anesthetic eect
Male Wistar albino rats (200-250 g) and
New Zealand white male albino rabbits (2.0-
2.5 kg) were obtained from Pasteur Institute
(Tehran, Iran) and used for local anesthetic
studies and skin irritation tests, respectively.
The animals were housed in suitable cages
at a controlled temperature (20-24 °C), on a
12:12 h, day/night cycle with free access to
a pellet diet and water ad libitum. All animal
experiments were performed in accordance
with the National Institute of Health (NIH)
Guide for the Care and Use of Laboratory
Animals (8th edition), approved by the
Institutional Animal Care and Use Committee
and local Animal Ethics Committee of Shahid
Beheshti University of Medical Sciences (No.
IR.SBMU.PHARMACY.REC.1399.002). The
local anesthetic eect of formulations was
assessed by performing a manual von Frey
test. All experiments were carried out between
9:00 and 16:00. Before the experiments, the
adult male rats were placed in an individual
clear acrylic box with an elevated plastic wire
mesh oor which allowed acclimating for 30
min in the testing environment. Animals were
divided randomly into the following groups (n
= 8):
Placebo control group (control A, B, and C)
Treated group with selected lidocaine-
loaded MBGs (formulation A, B, and C)
Treated group with the commercial
lidocaine gel
In this behavioral study, 0.5 g of each gel
was topically applied to the rat hind paw.
Then, a series of 10 von Frey laments with
logarithmic incremental stiness (4, 6, 8, 10,
15, 26, 60, 100, 180, and 300 g) in ascending
order was used to determine the mechanical
allodynia threshold of the animals, 10-210
min following the application of the gel with
10-min intervals. Each nylon lament was
applied ve times through the mesh oor
on the plantar surface of the rat paw till it
bends (buckles). Brisk paw withdrawal,
licking, or shaking of the stimulated paw
was considered as a positive response (42).
The strongest lament inducing up to two
responses out of ve stimuli was recorded
as the mechanical threshold at each time
point. Results were reported as mean ± SEM.
Statistical dierences were evaluated using
two-way ANOVA with Bonferroni’s post-
test to compare the mechanical threshold at
each time, and one-way ANOVA followed by
Tukey’s post-test to evaluate the created area
under the time-course curve (AUC10-210 min) of
the mechanical threshold by each formulation
during the test. As stated earlier, p < 0.05 was
considered statistically signicant (*p < 0.05,
7
Lidocaine-loaded, Microemulsion-based Topical Gels
**p < 0.01 and ***p < 0.001).
Skin irritation test
The acute dermal irritation potential of the
nal formulation was evaluated in accordance
with the OECD guideline (43). The animals
were acclimatized for one week before the
beginning of the study and had access to a
standard diet and food. The hairs on the back of
rabbits were trimmed by an electrical clipper
24 h prior to administration of the formulation.
The animals were divided into three groups (n
= 3) as follows:
No application (control)
Blank MG4
Drug-loaded MG4
Half a gram of gel formulations was applied
uniformly to the test area (approximately 6
cm2). At the end of the 4-h exposing duration,
the residual gel was wiped o with water. All
rabbits were observed for any visible change
such as erythema or edema after 1, 24, 48, and
72 h of the gel application. If skin damage
cannot be recognized as irritation or corrosion
after 72 h, the observation should continue
until day 14 in an attempt to determine the
reversibility of the eects. Erythema and
edema were graded according to the following
criteria: 0, no visible reaction; 1, very slight
reaction; 2, well-dened erythema; 3, moderate
to severe reaction; 4, severe reaction. Finally,
the irritation scores of the test area were
calculated using the following equation and
interpreted according to Table 1.
Equation 1.
Results and Discussion
Drug solubility in the oil phase
The ability of the oil phase for drug
solubilization is considered as the most
important criterion for o/w microemulsion
formulations (44). Lidocaine solubility results
for dierent oils are given in Figure 1. As can
be seen, the highest solubility was obtained in
castor oil and triacetin (538.460 ± 7.457 mg/
mL and 530.727 ± 6.029 mg/mL, respectively).
This represents the potential of these oils to
Table 1. Interpretation of PII scores.*
PII score Irritation grade
PII = 0 no irritation
0 < PII ≤ 2 mild
2 < PII ≤ 5 moderate
5 <PII 8 severe
*Appraisal of the safety of chemicals in foods, drugs, and cosmetics: Association of Food
and Drug Officials of the United States, 1959.
Figure 1. Solubility study of lidocaine in various oils. Data expressed in mean SEM (n = 3).
Figure 1. Solubility study of lidocaine in various oils. Data expressed in mean ± SEM (n = 3).
Table 1. Interpretation of PII scores.*
󰇛󰇜



8
Daryab M et al. / IJPR 21 (2022) e1: 1-21
solubilize lidocaine, and therefore, they were
selected for ME and MBG preparations.
Phase diagrams and o/w ME domains
Phase diagrams of four-component systems
were constructed to determine the appropriate
concentration ranges of the components to form
MEs. Unlike the systems containing castor oil,
the o/w ME region was observed on triacetin-
based phase diagrams. Figure 2 indicates that in
nearly all phase diagrams (except for triacetin/
Tween 80/PEG 400/water systems at Rsm of
1:2 and triacetin/Labrasol/PEG 400/water,
regardless of Rsm), a transparent, isotropic
o/w ME region was formed in the oil-poor
part of the phase diagrams. It should be noted
that because of the diculties in accurately
determining the boundaries between the ME
domains and surfactant-rich area on the top
of the phase diagrams, samples with up to 50
wt% surfactant mixture were considered as
MEs, above which the area was considered as
surfactant-rich area.
The following generalizations could be
Figure 2. Phase diagrams of systems consisting of triacetin as the oil phase (right apex), distilled
water (left apex), (A) Transcutol P, (B) PG, and (C) PEG 400 as co-surfactant and various
surfactants (top apex) namely Tween 80 (green), Labrasol (red), Cremophor EL (yellow), and
Cremophor RH40 (purple) at various Rsm of a) 1:1, b) 1:2, and c) 2:1. The colored area in the oil-
poor part of the phase diagram represents the o/w ME domain.
Figure 2. Phase diagrams of systems consisting of triacetin as the oil phase (right apex), distilled water (left apex), (A)
Transcutol P, (B) PG, and (C) PEG 400 as co-surfactant and various surfactants (top apex) namely Tween 80 (green),
Labrasol (red), Cremophor EL (yellow), and Cremophor RH40 (purple) at various Rsm of a) 1:1, b) 1:2, and c) 2:1.
The colored area in the oil-poor part of the phase diagram represents the o/w ME domain.
9
Lidocaine-loaded, Microemulsion-based Topical Gels
made about the investigated systems:
Tween-based systems showed higher water
solubilization capacity in comparison to the
other surfactants.
Irrespective of the type of surfactant and
Rsm, the largest and smallest ME areas were
seen in the presence of Transcutol P and PEG
400, respectively.
Regardless of the type of surfactant and
co-surfactant, Rsm did not have a signicant
inuence on the extent of the o/w ME region.
Various ME formulations were selected
from the relatively extended o/w ME area on
the phase diagrams, considering the minimum
possible concentration of surfactants. Those
with no drug precipitation and phase separation
at the time of preparation and after 72 h storage
were chosen for further characterization tests
(Table 2).
ME characteristics
The prepared formulations (Table 2) were
found to be macroscopically identical, i.e.,
homogeneous, single-phase, and transparent
by visual inspection. The colloidal nature
of these systems was also conrmed by
observing the Tyndall eect. Table 3 lists
the data of Z-average, PDI, zeta potential,
pH, conductivity, and viscosity of the drug-
loaded ME formulations. The isotropic nature
of the formulations was also conrmed as a
completely dark eld was observed under the
cross-polarized light microscope.
The refractive indices of all formulations
were ranged between 1.3750 and 1.3890 and
close to that of water as the external phase
(1.334). Electrical conductivity measurement
is a useful tool to dierentiate w/o droplets
from o/w-type droplets and bicontinuous
structures. Generally, low conductivity
exhibits the formation of w/o droplet MEs
(because water makes the internal phase),
while systems showing high conductivity are
dened as bicontinuous or o/w-type MEs as
the presence of water in the continuous phase
allows the measurement of conductivity. In
this study, data obtained from both RI and
conductivity measurements (162.592-198.340
Table 2. Composition of the selected microemulsion systems for drug solubilization.
Microemulsion Triacetin (%) (oil phase) Surfactant (%) Co-surfactant (%) Water
(aqueous phase)
X1 X
2 X
3 X
4 X
5 X
6
ME1 5.04 16.59 - - 8.29 - - 70.08
ME2 5.04 12.49 - - - 12.49 - 69.98
ME3 5.04 16.59 - - - 8.29 - 70.08
ME4 5.04 16.59 - - - - 8.29 70.08
ME5 5.04 - 12.49 - 12.49 - - 69.98
ME6 5.04 - 16.59 - 8.29 - - 70.08
ME7 5.04 - 9.97 - - 9.97 - 75.02
ME8 5.04 - 8.29 - - 16.59 - 70.08
ME9 5.04 - 13.23 - - 6.74 - 74.99
ME10 5.04 - 9.97 - - - 9.97 75.02
ME11 5.04 - 16.59 - - - 8.29 70.08
ME12 5.04 - - 16.59 8.29 - - 70.08
ME13 5.04 - - 12.49 - 12.49 - 69.98
ME14 5.04 - - 16.59 - 8.29 - 70.08
X1: Tween 80; X2: Cremopho
r
EL; X3: Cremopho
r
RH40; X4: PEG 400, X5: TranscutolP; X6: PG.
Table 2. Composition of the selected microemulsion systems for drug solubilization.
Table 3. Characterization of fluid MEs (mean SEM).
ME Z-average (nm) PDI Zeta potential (mV) pH Conductivity (
S/cm) Refractive index Viscosity (mPa.s)
ME1 47.073 0.524 0.408 0.005 0.635 0.029 8.04 0.030 194.005 0.116 1.3850 0.0002 27.10 0.058
ME2 36.710 0.541 0.460 0.009 -0.711 0.075 8.28 0.015 167.327 0.122 1.3815 0.0002 11.90 0.033
ME3 45.290 0.593 0.389 0.004 0.0886 0.035 7.82 0.012 193.940 0.035 1.3830 0.0003 18.50 0.044
ME4 44.140 1.400 0.380 0.009 0.0255 0.040 7.76 0.010 194.721 0.124 1.3830 0.0002 19.70 0.015
ME5 52.196 0.248 0.404 0.002 0.523 0.046 7.80 0.010 190.037 0.043 1.3840 0.0002 20.90 0.010
ME6 30.950 0.163 0.363 0.003 -0.265 0.058 7.90 0.012 169.172 0.588 1.3865 0.0003 26.8 00.007
ME7 38.850 0.117 0.357 0.007 0.366 0.035 8.11 0.010 198.340 0.116 1.3770 0.0005 8.59 0.006
ME8 59.506 0.532 0.452 0.008 0.285 0.046 8.09 0.015 194.288 0.130 1.3805 0.0002 7.85 0.006
ME9 32.380 0.121 0.371 0.001 0.134 0.023 8.04 0.025 188.365 0.289 1.3765 0.0003 11.20 0.007
ME10 28.016 0.112 0.314 0.001 0.832 0.080 8.10 0.009 176.834 0.324 1.3750 0.0005 8.41 0.010
ME11 20.626 0.225 0.350 0.001 -0.447 0.052 7.98 0.015 162.591 0.176 1.3820 0.0002 19.80 0.030
ME12 45.400 5.506 0.428 0.050 0.0318 0.064 8.16 0.010 179.647 0.472 1.3890 0.0005 18.80 0.046
ME13 57.060 14.475 0.178 0.006 -0.310 0.035 8.20 0.015 173.398 1.041 1.3810 0.0005 9.10 0.006
ME14 26.910 0.736 0.232 0.006 0.201 0.098 8.22 0.012 173.012 1.882 1.3820 0.0003 14.00 0.055
Table 3. Characterization of uid MEs (mean ± SEM).
10
Daryab M et al. / IJPR 21 (2022) e1: 1-21
µS/cm) approved the o/w structure of the MEs
studied (45). The conductivity results also
depicted that the addition of lidocaine to the
internal phase of MEs did not aect the system
stability.
The average particle size and size
distribution of MEs were evaluated by the
dynamic light scattering technique. PDI was
also determined to provide information about
the deviation from the mean size. Table 3
represents the results of size and PDI analysis.
As can be seen, in all systems, the average
size of the ME droplets was less than 60 nm
(ranged from 20.626-59.506 nm) which lies
in the proposed range for ME systems (<120
nm). All formulations exhibited unimodal
droplet size distribution patterns (diagrams not
shown). In most cases, as a measure of droplet
size uniformity, PDI values were found to be
less than 0.4, suggesting that droplets in nearly
all MEs were relatively uniform-sized.
To analyze the charge of the droplets, zeta
potential is determined. Zeta potential values
indicated that the interface had a low surface
charge (-0.711 to +0.832 mV). The charge in
the interfacial area, in general, may originate
from many factors the composition of oil, the
presence of electrolytes in the water phase,
and the nature of surfactants. In this study,
very low zeta potential (nearly zero potential)
values obtained in this study could be ascribed
to the presence of non-ionic surfactants.
Zeta potential is not usually considered as an
important measure for the stability prediction
of MEs prepared with non-ionic surfactants
(46).
The pH values of all MEs were found to vary
between 7.76 and 8.28. A slight increase in pH
could be attributed to the presence of lidocaine
in the formulations. MEs possessed very low
viscosity (7.85 to 27.1 mPa.s), independent
of shear rate. For all formulations, a linear
section was observed on the ow curves,
constructed with shear stress vs. shear rate (r2
0.99). Figure 3 illustrates the rheogram of
the ME5 formulation.
MBGs
Three dierent gelling agents were used to
increase the viscosity of MEs. In the presence
of Carbomer 934, all gels were found
opaque. HPMC was also unable to yield clear,
homogenous MBGs with desired viscosity.
Turbidity and lack of homogeneity were
resolved by substituting HPMC and Carbomer
934 with Carbomer 940 (47). As described by
Chen et al., the reason might be associated with
the dissociation of Carbomer 934 and HPMC
matrices from the hydrated state by surfactant
and co-surfactant in the microemulsion (48).
The gels were also evaluated in terms of
stickiness, ease of spreading, and coarseness
Figure 3. Rheogram of ME5 formulation.
11
Lidocaine-loaded, Microemulsion-based Topical Gels
by rubbing a sucient amount of gels between
index and thumb ngers. In general, results
showed that all Carbomer 940-based gels
were homogeneous, transparent, and smooth
without any particulate matter, grittiness,
or lumps and, therefore, MBGs with 1 wt%
of Carbomer 940 were nally prepared for
further investigations.
MBGs properties
The data obtained from the characterization
of MBGs in terms of spreadability, pH, and
viscosity are given in Table 4. The spreadability
of gel formulations, that is, the ability of gels
to spread uniformly on the skin surface, is a
property upon which the therapeutic eciency
of a gel depends and helps in the uniform
gel application. Values in Table 4 refer to
the extent to which the formulations readily
spread on the glass plates by applying a small
amount of shear.
Results indicated that the highest
spreading diameter (4.2 cm) was obtained
for the formulation MG8, which possessed
the lowest viscosity, whereas the lowest
spreading diameter (3.1 cm) was found for the
system MG11 with the highest viscosity. The
appropriate spreadability of MBGs may be
related to the loose gel matrix nature of MBGs
due to the presence of oil globules (49).
As depicted in Table 4, it was observed
that pH values of MBGs were within the
physiological range varying from 6.87 to 7.42.
This pH range suggests that the gels could
result in less irritation to the skin. A decrease
in pH of MBGs in comparison with MEs
may be attributed to the acidic properties of
Carbomer 940 (17).
The use of MEs on the skin is very
dicult because of their uidity. For a dermal
pharmaceutical or cosmetic product, an
appropriate viscosity with sucient retention
time on the skin is required. Hence, MBGs
were developed by using Carbomer 940 in
an attempt to modify rheological behavior.
Viscosity values for lidocaine-loaded gels are
also shown in Table 4. As expected, following
the incorporation of the gelling agent into
MEs, the viscosity of systems increased
signicantly (from 224.83 to 871.62 mPa.s),
and pseudoplastic behavior was observed.
The latter could facilitate and improve the
spreading features of the formulation. The
ow indices (n) were found to be less than
1 (0.2788-0.4479), indicating that all MBGs
were shear-thinning in nature according to
the power law equation (13). Rheograms also
revealed the absence of thixotropy in the gels
investigated (Figure 4).
Stability studies of uid MEs and MBGs
The stability of MEs was evaluated after 15
months of storage at room temperature. MGBs
were also kept at dierent storage conditions
(5 ± 3 °C, 25 ± 2 °C and 40 ± 2 °C) for 9
months, and their transparency and consistency
were monitored. As shown in Figure 5, all
ME formulations (except ME13) were clear
without any turbidity or sedimentation. The
gels also remained clear with homogenous
structures and displayed no macroscopic
physical changes following storage at ambient
temperature and in a refrigerator (Figure 6).
However, loss of their viscosity was observed
Table 4. Characterization of the MBGs (mean ± SEM).
Formulations Spread diameter (cm) pH Viscosity (mPa.s) Flow index Nature/type of flow
MG1 3.3 0.058 7.23 0.012 696.65 0.870 0.3397 Shear-thinning / pseudoplastic
MG2 3.6 0.033 7.42 0.015 428.85 0.731 0.3438 Shear-thinning / pseudoplastic
MG3 3.7 0.058 6.98 0.010 705.75 0.463 0.3247 Shear-thinning / pseudoplastic
MG4 3.8 0.033 6.88 0.009 796.08 0.511 0.3245 Shear-thinning / pseudoplastic
MG5 3.6 0.058 6.97 0.012 712.89 0.932 0.3156 Shear-thinning / pseudoplastic
MG6 3.7 0.067 7.10 0.030 773.89 1.261 0.3404 Shear-thinning / pseudoplastic
MG7 4.1 0.058 7.21 0.010 235.51 0.334 0.3467 Shear-thinning / pseudoplastic
MG8 4.2 0.033 7.16 0.015 224.83 0.327 0.4479 Shear-thinning / pseudoplastic
MG9 3.9 0.100 7.18 0.009 379.50 0.464 0.3140 Shear-thinning / pseudoplastic
MG10 3.2 0.058 7.22 0.030 660.12 1.359 0.2788 Shear-thinning / pseudoplastic
MG11 3.1 0.067 7.14 0.003 871.62 1.458 0.2988 Shear-thinning / pseudoplastic
MG12 3.3 0.033 7.28 0.010 677.93 1.034 0.3265 Shear-thinning / pseudoplastic
MG13 3.8 0.100 7.31 0.009 397.85 1.172 0.3489 Shear-thinning / pseudoplastic
MG14 3.5 0.153 7.33 0.015 516.76 1.023 0.3502 Shear-thinning / pseudoplastic
Marketed gel 3.9 0.100 6.87 0.025 328.18 0.572 0.4776 Shear-thinning / pseudoplastic
Table 4. Characterization of the MBGs (mean ± SEM).
12
Daryab M et al. / IJPR 21 (2022) e1: 1-21
after 60 days of storage at 40 °C.
The stability of MBGs was evaluated
under stressed conditions by visual inspection.
When subjected to centrifugation at 5000
rpm for 30 min, it was found that this stress-
induced no damage, and the formulations
remained homogeneous and exhibited no sign
of phase separation or breakdown. The eect
Figure 4. Rheogram of MG6 formulation.
Figure 4. Rheogram of MG6 formulation.
Figure 5. Stability of MEs after 15 months of storage at room temperature. As seen, all
formulations except ME13 were clear without any turbidity or sedimentation.
Figure 5. Stability of MEs after 15 months of storage at room temperature. As seen, all formulations except ME13 were
clear without any turbidity or sedimentation.
13
Lidocaine-loaded, Microemulsion-based Topical Gels
of heating-cooling cycles on the stability of
MBGs was also veried. In each heating-
cooling cycle, the sample was rst heated to
40 °C for 24 h and subsequently cooled to 4 °C
for 24 h. Seven heating-cooling cycles were
run to record the gel responses to temperature
uctuations. Finally, the inuence of the
repeatedly freeze-thawed treatment (-5 and 25
°C for 24 h) on the stability of the gels was
investigated. The results obtained from these
stability tests foresee MBGs to have good
physical stability since no phase separation
was observed and the textural properties were
not inuenced by temperature variation.
Permeation study
Drug release and permeation studies through
cellulose acetate membrane and rat skin,
respectively, from MBGs, were carried out
using vertical Franz diusion cells. Although
human skin is considered the gold standard
in permeation study of topical formulations,
however, limited availability, variability, and
ethical reasons have led to employ the animal
skin models (50). Some structural similarities
between rat skin and human skin (e.g., thickness,
lipid content, and water uptake) can propose
rat skin as a surrogate for permeation studies
(51, 52). For preliminary drug permeation
screening, the lipophilic articial membrane
was employed, and subsequently, an ex-vivo
permeation study on rat skin was conducted
for the formulations with the highest ux value
through an articial membrane.
In-vitro drug release through an articial
membrane
This part of the investigation was aimed
to select the best MBG formulations for ex-
vivo skin permeation and animal tests. The
cumulative percentage of released lidocaine
was plotted as a function of time (Figure 7).
In general, it was observed that in all MBGs,
the drug release percentage at all sampling
points was signicantly greater than that of
the commercial gel, suggesting that MBGs
could improve the release pattern of the drug
in comparison with the marketed product. In-
vitro drug release proles also revealed that
the formulations MG3 (triacetin/Tween 80/
Transcutol P at Rsm of 2:1), MG5 (triacetin/
Cremophor EL/PEG 400 at Rsm of 1:1), and
MG4 (triacetin/Tween 80/PG at Rsm of 2:1)
released the maximum amount of lidocaine
(61.65 ± 1.62%, 61.24 ± 0.70% and 61.04 ±
0.76%, respectively) after 2 hours (p < 0.01)
(Figure 7), while the system MG11 (triacetin/
Cremophor EL/PG at Rsm of 2:1) displayed
the lowest amount of released drug (50.42
± 0.76%) with no statistically signicant
dierence compared to the commercial gel
(p > 0.05). Therefore, MG3, MG4, and MG5
were considered as the optimum gel systems
and chosen for ex-vivo drug permeation
investigations.
Ex-vivo drug permeation through the skin
The ex-vivo drug permeation through
the skin was carried out in an attempt to
formulate a vehicle with suitable skin uptake
Figure 6. Stability of MBGs after 9 months of storage at room temperature. No textural change or
breakdown was observed in the formulations investigated.
Figure 6. Stability of MBGs after 9 months of storage at room temperature. No textural change or breakdown was
observed in the formulations investigated.
14
Daryab M et al. / IJPR 21 (2022) e1: 1-21
and penetration. Results are depicted in Figure
8. As can be seen, the drug permeation from
formulations MG3, MG4, and MG5 started
immediately without any lag phase, followed
by a continuous increase over time. By
comparing the skin permeation proles, it is
observed that MG4 (triacetin/Tween 80/PG at
Rsm of 2:1) exhibited the highest cumulative
Figure 7. In-vitro drug release profiles of formulation MG3, MG4 and MG5 through an artificial
membrane. Data are shown as means ± SEM (n = 3). One-way ANOVA followed by Tukey’s post-
test multiple comparisons were conducted (*p < 0.05: significant, **p < 0.01: very significant, and
***p < 0.001: extremely significant, in comparison with the marketed product).
Figure 8. Drug permeation from MG3, MG4, and MG5 systems through abdominal rat skin. Data
are shown as means ± SEM (n = 3). The difference between the release percentage of the
formulations was statistically analyzed by one-way ANOVA followed by Tukey's post-test (*p <
0.05: significant, **p < 0.01: very significant, and ***p < 0.001: extremely significant, in
comparison with the marketed product).
Figure 7. In-vitro drug release proles of formulation MG3, MG4 and MG5 through an articial membrane. Data are
shown as means ± SEM (n = 3). One-way ANOVA followed by Tukey’s post-test multiple comparisons were conduct-
ed (*p < 0.05: signicant, **p < 0.01: very signicant, and ***p < 0.001: extremely signicant, in comparison with the
marketed product).
Figure 8. Drug permeation from MG3, MG4, and MG5 systems through abdominal rat skin. Data are shown as means ±
SEM (n = 3). The dierence between the release percentage of the formulations was statistically analyzed by one-way
ANOVA followed by Tukey’s post-test (*p < 0.05: signicant, **p < 0.01: very signicant, and ***p < 0.001: extremely
signicant, in comparison with the marketed product).
15
Lidocaine-loaded, Microemulsion-based Topical Gels
amount of lidocaine permeation versus time
after 10 h (5300.705 mg/cm2) and signicantly
enhanced lidocaine permeation compared
to MG5 and the control gel. No statistically
signicant dierence in permeation was
found between MG3 and MG4 until 15 min,
suggesting that their onset of action could be
almost the same. However, higher drug release
was observed from MG4, which supports a
longer duration of action.
Rapid onset of action for LAs is very
important, and therefore, a high initial
permeation is immediately required.
Cumulative drug release per unit area of
skin surface from all formulations after 10,
20, and 60 min demonstrated a signicant
enhancement of ux in comparison with the
commercial gel (p < 0.001), so that for MG4
as the optimized system, 4.09, 3.54, and
1.91-fold increase in the ux were observed,
respectively. It is crucial to determine the
minimum amount of drug permeation that
induces local anesthesia (i.e., the anesthetic
threshold). Lidocaine anesthetic threshold was
calculated to be 500 μg/cm2, based on the data
obtained from in-vitro drug release and in-vivo
anesthetic examination (tail-ick test) (8).
Considering this value, it could be expected
that formulation MG4 causes local anesthesia
faster than the other MGBs within 7 minutes
after applying the gel. Formulations MG3 and
MG5 also induced their eects after 10-15
min; however, the amount of drug needed to
initiate the anesthetic eect of the commercial
gel was released in 30 to 45 min. In general,
it is concluded that the faster local anesthesia
was achieved by the use of MBGs, compared
to the commercially available gel.
An increase in the permeation rate
within the rst two hours could be explained
as follows. Due to the presence of both
hydrophilic and lipophilic components and
the resulting combined eects, MEs possess
a favorable solubilizing behavior. This
increases the thermodynamic activity of the
drug, which is a driving force for drug release
and its penetration (49). Besides, it has been
previously reported that topically applied MEs
are expected to penetrate the skin and exist
intact in the stratum corneum (SC). Kweon
et al. have suggested that MEs, once entered
into the SC, could alter both the polar and lipid
pathways, and the subsequent interaction of the
lipid portion of the MEs with the SC makes the
dissolved drug partition into the existing lipids.
On the other hand, the bilayer structure of the
SC could be destabilized by the intercalation of
ME droplets between its lipid chains (53). The
hydration eect on the drug uptake of the SC
by the hydrophilic domain of MEs should also
be considered. It is thought that the aqueous
phase of MEs would increase the interlamellar
volume and disrupt the lipid bilayers due to the
swelling of the intercellular proteins, causing
a more easily penetration of the drug through
the lipid pathway of the SC (53). In conclusion,
the greater penetration enhancing the activity of
MEs may be attributed to the combined eects
of both the lipophilic and hydrophilic domains
of microemulsions.
As can be seen in Figure 8, the release of
lidocaine molecules from the investigated
MBGs (MG3-MG5) was sustained for 10
h. This phenomenon may be explained by
considering the release of the loaded lidocaine
from the internal phase, which might act as
a drug reservoir, to the external phase and
then from the continuous phase to the skin
through passive diusion. Lidocaine can also
be partially solubilized in the external phase
and interfacial lm of ME that can supply fast
release at the initial time of study leading to
the fast onset of action without any lag time.
It has been suggested that the gel formation in
ME limits the diusion of the drug dissolved
in the droplets and therefore slows down its
release. Thus, one can conclude that MBGs
are potentially able to sustain the release of
drugs as compared with their uid systems.
Therefore, the high permeation rate of MG4
could be related to its ability to create a high-
saturated vehicle which can result in high
thermodynamic activity (54).
The particle size of ME droplets plays
an important role in percutaneous drug
absorption. It has been reported in the
literature that by decreasing the droplet size,
the number of particles that can interact with
the skin surface is probably increased (49, 53
and 55). In this investigation, the particle size
of all ME formulations was in the range of
20-52 nm. This suggests that a large surface
area for the transfer of lidocaine to the skin is
available.
16
Daryab M et al. / IJPR 21 (2022) e1: 1-21
The higher lidocaine ux from formulated
MBGs compared to the commercial gel
originates from the penetration-enhancing
eect of applied components. Cao et al.
prepared celecoxib-loaded MBG using Tween
80 and Transcutol P and evaluated the ex-
vivo permeation of the drug into the mouse
skin. The results revealed that the interested
formulation could have a 4-fold greater
permeability than the conventional gel (56).
These ndings have also been previously
obtained by Shakeel et al., using penetration
enhancers such as Labral, triacetin, Tween
80, and Transcutol P for aceclofenac (57).
Similarly, other researchers have developed
MBGs containing a mixture of Cremophor
EL and PEG 400 as the surfactant phase to
co-delivery of evodiamine and rutaecarpine.
By application of this nano-based gel
formulation, it was shown possible to achieve
approximately 2.6-fold higher transdermal
ux compared with control hydrogel (58).
In general, numerous studies on ME gels
prepared with Tween 80 and PG have shown
that this surfactant mixture has an important
impact on increasing the skin permeability as
well as the stability of systems, and it has also
been stated that PG can exhibit an additive
eect on drug permeation in combination with
other penetration enhancers (59).
Drug release kinetics
To determine the kinetics of permeation
from these vehicles, the data obtained from ex-
vivo permeation experiments were kinetically
analyzed according to zero order, rst order,
Higuchi and Korsmeyer-Peppas models, and
the results of data tting into these models
were evaluated by the highest correlation
coecient (R2). Based on the best goodness
of t (see Table 5), it was found that MG3,
MG4, and the marketed product were followed
Higuchi kinetic model (MG3: R2 = 0.9942,
MG4: R2 = 0.9862, marketed gel: R2 = 0.9832).
Higuchi model-based permeation, previously
reported for indomethacin (chitosan-based),
terbinane (chitosan-based), itraconazole
(Lutrol F127-based) and ibuprofen
(Carbopol 940-based) MBGs and for topical
ketoprofen and pentoxifylline MEs (24, 60-
64), suggests that the release process could be
mainly controlled by the Fickian diusion of
dissolved lidocaine through the gel network of
Carbomer 940. However, the analysis of the
release plot for MG5 revealed that lidocaine
followed the rst-order model for controlled
permeation, suggesting that the release rate is
concentration-dependent (23, 65).
If diusion is the main drug release
mechanism regarding the Higuchi equation,
then a plot of the drug amount released versus
the square root of time should result in a straight
line. However, a deviation from the Fickian
equation may be observed, and the mechanism
of diusion from polymeric dosage forms may
follow a non-Fickian behavior. Korsmeyer-
Peppas equation (Equation 2) is a more
general relationship that describes a mixed
mechanism of drug release (polymer swelling
and/or diusion) from a polymeric system:
Equation 2.
where k is a constant incorporating the
geometrics and structural characteristics
of dosage form, n is the release exponent
indicative of the release mechanism, and
Mt/M is the fractional release of the drug.
This equation relates the drug release to the
elapsed time (t). In this study, to elucidate
the drug release mechanism, the rst 60%
of drug release data was used to calculate
Table 5. Models used to assess the release kinetics from the best MBGs and the corresponding Korsmeyer-Peppas parameters.
Formulation R2 values Korsmeyer-Peppas parameters
Zero ordera First orderb Higuchic Korsmeyer-Peppasd n k
MG3 0.9409 0.9913 0.9942 0.9919 0.5062 2.9798
MG4 0.9613 0.9628 0.9862 0.9890 0.4840 3.7818
MG5 0.9528 0.9964 0.9936 0.9915 0.5609 2.2289
Commercial gel 0.9775 0.9585 0.9832 0.9940 0.8543 0.4476
aCumulative amount of drug permeated (g) versus time.
bLog of the amount of remaining drug (g) versus time.
cCumulative amount of drug permeated (g) versus square root of time.
dSee equation 2 in the text.
Table 5. Models used to assess the release kinetics from the best MBGs and the corresponding Korsmeyer-Peppas
parameters.

17
Lidocaine-loaded, Microemulsion-based Topical Gels
values of n, k, and correlation coecient
(R2) (Table 5). Values of the release exponent
for MG3, MG4, and MG5 formulations and
the marketed product were calculated to be
between 0.484 and 0.854. Therefore, it was
concluded that the mechanism of transport for
all these formulations followed an anomalous
(non-Fickian) behavior, as described in Table
6, possibly including both diusion and/or
polymer erosion phenomena. These results
are in accordance with those reported for
zaltoprofen and griseofulvin MBGs (47, 66
and 67) and contraceptive vagino-adhesive
propranolol HCl gel (68).
Anesthetic eect
Paw withdrawal threshold (PWT) values
of the lidocaine-treated rats were found to
be signicantly higher than their respective
controls, conrming the induction of the
anesthetic eect of lidocaine (p < 0.001).
Repeated-measure, two-way ANOVA
(followed by Bonferroni’s post-test) revealed
that MG4 formulation showed a markedly
greater anesthetic eect in comparison with
the marketed gel. This nding supports the
results of the ex-vivo permeation test (Figure
9). Also, it was observed that MG3 showed
no statistically signicant dierence in PWT
value approximately during the rst two hours
of the study, and for MG5, the induction of local
anesthesia was similar to the marketed gel. In
order to compare the average pain threshold
during the complete period of observation
following the application of the formulations,
the area under the time-course curve was
calculated. As can be clearly seen in Figure
10, MG4 and MG3 induced a statistically
signicant high pain threshold in comparison
to the commercial product (p < 0.001, one-
way ANOVA followed by Tukey’s post-test),
although the dierence was not signicant
between that of MG5 and the marketed gel.
Skin irritation test
The irritation potential of any transdermal
formulation is a critical factor that could limit
its use and patient acceptability. In the present
Table 6. Interpretation of diffusion release mechanism.
Release exponent (n) Release mechanism
n ≤ 0.5 Fickian diffusion
0.45 < n < 0.89 non-Fickian (anomalous) transpor
t
n = 0.89 case II transpor
t
n > 0.89 super case II transpor
t
Table 6. Interpretation of diusion release mechanism.
Figure 9. Paw withdrawal threshold of the selected formulations (MG3, MG4, & MG5) to
mechanical stimulation (von Frey filaments). Data are shown as means ± SEM, n = 8 rats per group
(n = 8). Two-way ANOVA followed by Bonferroni post-test.
Figure 9. Paw withdrawal threshold of the selected formulations (MG3, MG4, and MG5) to mechanical stimulation
(von Frey laments). Data are shown as means ± SEM, n = 8 rats per group (n = 8). Two-way ANOVA followed by
Bonferroni post-test.
18
Daryab M et al. / IJPR 21 (2022) e1: 1-21
study, special consideration was given to the
selection of components used in the formulations
on the basis of solubility and the minimal skin
irritation tendency. Draize primary skin irritation
test was performed on the albino rabbit skin to
study the irritability of the optimum formulation.
The results obtained from skin irritation studies
after 1, 24, 48, and 72 h of the gel application
are listed in Table 7. The prepared gels were not
found to be skin irritants.
Conclusion
In the present study, various formulations
of lidocaine-loaded MBGs were prepared and
characterized. It was concluded that MBGs
could be considered as a more promising
approach for the transdermal delivery of
lidocaine due to their appropriate viscosity and
rheological behavior, spreadability, pH, high
penetration ability, and skin tolerability with
no irritation, high stability, and improvement
in PWT and anesthetic eect. However,
further research and clinical investigations
need to be conducted to elucidate the possible
mechanism(s) of lidocaine delivery to the skin
and conrm the therapeutic ecacy.
Acknowledgments
This research was nancially supported
by the Vice-Chancellor of Research, Shahid
Beheshti University of Medical Sciences. The
authors are grateful to Dr. Mona Khorram
Joy and Dr. Bahareh Alizadeh for their great
eorts and assistance in conducting the animal
studies and HPLC analysis, respectively.
Conict of interests
The authors declared that there is no
conict of interest.
Figure 10. The area under the curve (AUC10-210 min) of withdrawal threshold time-course of the
selected formulations (MG3, MG4, & MG5) to mechanical stimulation (von Frey filaments). Data
are shown as means ± SEM, n = 8 rats per group (n = 8). One-way ANOVA followed by Tukey's
post-test multiple comparisons were conducted.
Figure 10. The area under the curve (AUC10-210 min) of withdrawal threshold time-course of the selected formulations
(MG3, MG4, and MG5) to mechanical stimulation (von Frey laments). Data are shown as means ± SEM, n = 8 rats per
group (n = 8). One-way ANOVA followed by Tukey’s post-test multiple comparisons were conducted.
Table 7. Average response scores of skin irritation.
Group Primary Irritation Index (Mean
SEM; n = 3)
1 h 24 h 48 h 72 h
No application (control) 0 0 0 0 0 0 0 0
Blank MG4 0 0 0 0 0 0 0 0
lidocaine-loaded MG4 0.44 0.11 0.22 0.11 0 0 0 0
Table 7. Average response scores of skin irritation.
19
Lidocaine-loaded, Microemulsion-based Topical Gels
Author contributions
All authors have made substantial
contributions to the experimental design of
this project. Experimental procedures and data
collection were carried out by M. Daryab.
Analysis and interpretation of data were
performed by all authors. This manuscript
was initially drafted by M. Daryab and
revised by A. Mahboubi and R. Aboofazeli
before submission. All authors approved the
nal manuscript for submission. This study
was the subject of the Pharm.D. thesis of M.
Daryab, proposed and approved by the School
of Pharmacy, Shahid Beheshti University of
Medical Sciences, Iran.
References
(1) Shah J, Votta-Velis EG and Borgeat A. New local
anesthetics. Best Pract. Res. Clin. Anaesthesiol.
(2018) 32: 179-85.
(2) Heavner JE. Local anesthetics. Curr. Opin.
Anaesthesiol. (2007) 20: 336-42.
(3) Shrestha M and Chen A. Modalities in managing
postherpetic neuralgia. Korean J Pain. (2018) 31:
235-43.
(4) Alster T. Review of lidocaine/tetracaine cream
as a topical anesthetic for dermatologic laser
procedures. Pain Ther. (2013) 2: 11-9.
(5) de Araújo DR, da Silva DC, Barbosa RM, Franz-
Montan M, Cereda CM, Padula C, Santi P and de
Paula E. Strategies for delivering local anesthetics
to the skin: focus on liposomes, solid lipid
nanoparticles, hydrogels and patches. Expert Opin.
Drug Deliv. (2013) 10: 1551-63.
(6) Shipton EA. New formulations of local
anaesthetics-part I. Anesthesiol Res. Pract. (2012)
2012: 546409.
(7) Shainhouse T and Cunningham BB. Topical
anesthetics. Am. J. Drug Deliv. (2004) 2: 89-99.
(8) Dogrul A, Arslan SA and Tirnaksiz F. Water/oil
type microemulsion systems containing lidocaine
hydrochloride: in-vitro and in-vivo evaluation. J.
Microencapsul. (2014) 31: 448-60.
(9) Wang Y, Su W, Li Q, Li C, Wang H, Li Y, Cao Y,
Chang J and Zhang L. Preparation and evaluation
of lidocaine hydrochloride-loaded TAT-conjugated
polymeric liposomes for transdermal delivery. Int.
J. Pharm. (2013) 441: 748-56.
(10) Yuan JS, Ansari M, Samaan M and Acosta
EJ. Linker-based lecithin microemulsions for
transdermal delivery of lidocaine. Int. J. Pharm.
(2008) 349: 130-43.
(11) Baek SH, Shin JH and Kim YC. Drug-coated
microneedles for rapid and painless local
anesthesia. Biomed. Microdevices. (2017) 19: 2.
(12) Shipton EA. New delivery systems for local
anaesthetics-part 2. Anesthesiol Res Pract. (2012)
2012: 289373.
(13) Negi P, Singh B, Sharma G, Beg S, Raza K and
Katare OP. Phospholipid microemulsion-based
hydrogel for enhanced topical delivery of lidocaine
and prilocaine: QbD-based development and
evaluation. Drug Deliv. (2016) 23: 951-67.
(14) Hoar TP and Schulman JH. Transparent water-
in-oil dispersions: the Oleopathic Hydro-Micelle.
Nature (1943) 152: 102-3.
(15) Tashtoush BM, Bennamani AN and Al-Taani BM.
Preparation and characterization of microemulsion
formulations of nicotinic acid and its prodrugs for
transdermal delivery. Pharm. Dev. Technol. (2013)
18: 834-43.
(16) Baroli B, López-Quintela MA, Delgado-Charro MB,
Fadda AM and Blanco-Méndez J. Microemulsions
for topical delivery of 8-methoxsalen. J. Control.
Release. (2000) 69: 209-18.
(17) Coneac G, Vlaia V, Olariu I, Muţ AM, Anghel DF, Ilie
C, Popoiu C, Lupuleasa D and Vlaia L. Development
and evaluation of new microemulsion-based hydrogel
formulations for topical delivery of uconazole.
AAPS PharmSciTech. (2015) 16: 889-904.
(18) Formariz TP, Sarmento VH, Silva-Junior AA,
Scarpa MV, Santilli CV and Oliveira AG.
Doxorubicin biocompatible O/W microemulsion
stabilized by mixed surfactant containing
soya phosphatidylcholine. Colloids Surf. B.
Biointerfaces (2006) 51: 54-61.
(19) Chhibber T, Wadhwa S, Chadha P, Sharma G and
Katare OP. Phospholipid structured microemulsion
as eective carrier system with potential in
methicillin sensitive Staphylococcus aureus
(MSSA) involved burn wound infection. J. Drug
Target. (2015) 23: 943-52.
(20) Reis MY, Santos SM, Silva DR, Silva MV, Correia
MT, Navarro D, Ferraz MA, Santos GK, Hallwass
F, Bianchi O and Silva AG. Anti-inammatory
activity of babassu oil and development of a
microemulsion system for topical delivery. Evid.
Based Complementary Altern. Med. (2017) 2017:
3647801.
(21) Tessema EN, Gebre-Mariam T, Paulos G, Wohlrab
J and Neubert RH. Delivery of oat-derived
phytoceramides into the stratum corneum of the skin
using nanocarriers: Formulation, characterization
and in-vitro and ex-vivo penetration studies. Eur. J.
Pharm. Biopharm. (2018) 127: 260-9.
20
Daryab M et al. / IJPR 21 (2022) e1: 1-21
(22) Delgado-Charro MB, Iglesias-Vilas G, Blanco-
Méndez J, López-Quintela MA, Marty JP and Guy
RH. Delivery of a hydrophilic solute through the
skin from novel microemulsion systems. Eur. J.
Pharm. Biopharm. (1997) 43: 37-42.
(23) Sah AK, Jain SK and Pandey RS. Microemulsion
based hydrogel formulation of methoxsalen for the
eective treatment of psoriasis. Asian J. Pharm.
Clin. Res. (2011) 4: 140-5.
(24) Cavalcanti AL, Reis MY, Silva GC, Ramalho
ÍM, Guimarães GP, Silva JA, Saraiva KL and
Damasceno BP. Microemulsion for topical
application of pentoxifylline: In-vitro release and in-
vivo evaluation. Int. J. Pharm. (2016) 506: 351-60.
(25) Savić V, Todosijević M, Ilić T, Lukić M, Mitsou E,
Papadimitriou V, Avramiotis S, Marković B, Cekić
N and Savić S. Tacrolimus loaded biocompatible
lecithin-based microemulsions with improved skin
penetration: structure characterization and in-vitro/
in vivo performances. Int. J. Pharm. (2017) 529:
491-505.
(26) Patel HK, Barot BS, Parejiya PB, Shelat PK and
Shukla A. Topical delivery of clobetasol propionate
loaded microemulsion based gel for eective
treatment of vitiligo: Ex-vivo permeation and skin
irritation studies. Colloids Surf. B. Biointerfaces.
(2013) 102: 86-94.
(27) Yang C, Shen Y, Wang J, Ouahab A, Zhang T and
Tu J. Cationic polymer-based micro-emulgel with
self-preserving ability for transdermal delivery of
diclofenac sodium. Drug Deliv. (2015) 22: 814-22.
(28) Djekic L, Martinovic M and Primorac M.
Microemulsion Hydrogels – Properties and
current applications in drug delivery. In: Torres
T. (ed.) Microemulsions: systems, properties and
applications. 1st ed. Nova Science Publishers, New
York, USA (2016) 1-36.
(29) Patel RR, Patel ZK, Patel KR and Patel MR. Micro
emulsion based gel: recent expansions for topical
drug delivery system. J. Med. Pharm. Allied Sci.
(2014) 1: 1-15.
(30) Djekic L, Martinovic M, Stepanović-Petrović R,
Micov A, Tomić M and Primorac M. Formulation
of hydrogel-thickened nonionic microemulsions
with enhanced percutaneous delivery of ibuprofen
assessed in-vivo in rats. Eur. J. Pharm. Sci. (2016)
92: 255-65.
(31) Vlaia L, Olariu I, Coneac G, Muţ AM, Popoiu
C, Corina S, Anghel DF, Maxim ME, Kalas S
and Vlaia VI. Development of microemulsion-
loaded hydrogel formulations for topical
delivery of metoprolol tartrate: Physicochemical
characterization and ex-vivo evaluation.
FARMACIA (2016) 64: 901-13.
(32) Wani RR, Patil MP, Dhurjad P, Chaudhari CA and
Kshirsagar SJ. Microemulsion based gel: A novel
approach in delivery of hydrophobic drugs. Int. J.
Pharm. Res. Scholars. (2015) 4: 398-410.
(33) Zhao L, Wang Y, Zhai Y, Wang Z, Liu J and
Zhai G. Ropivacaine loaded microemulsion and
microemulsion-based gel for transdermal delivery:
preparation, optimization, and evaluation. Int. J.
Pharm. (2014) 477: 47-56.
(34) Okur NÜ, Çağlar EŞ, Arpa MD and Karasulu HY.
Preparation and evaluation of novel microemulsion-
based hydrogels for dermal delivery of benzocaine.
Pharm. Dev. Technol. (2017) 22: 500-10.
(35) Ghai D and Sinha VR. Nanoemulsions as self-
emulsied drug delivery carriers for enhanced
permeability of the poorly water-soluble
selective β1-adrenoreceptor blocker Talinolol.
Nanomedicine (2012) 8: 618-26.
(36) Chhatrani BM and Shah DP. A review on
microemulsion based gel: A novel approach for
enhancing topical delivery of hydrophobic drug.
Int. J. Pharm. Pharm. Res. (2017) 8: 19-35.
(37) Kaur A, Sharma G, Gupta V, Ratho RK, Shishu
and Katare OP. Enhanced acyclovir delivery using
w/o type microemulsion: preclinical assessment of
antiviral activity using murine model of zosteriform
cutaneous HSV-1 infection. Artif. Cells Nanomed.
Biotechnol. (2018) 46: 346-54.
(38) Baboota S, Alam MS, Sharma S, Sahni JK,
Kumar A and Ali J. Nanocarrier-based hydrogel
of betamethasone dipropionate and salicylic acid
for treatment of psoriasis. Int. J. Pharm. Investig.
(2011) 1: 139-47.
(39) Almeida IF and Bahia MF. Evaluation of the
physical stability of two oleogels. Int. J. Pharm.
(2006) 327: 73-7.
(40) Farghaly DA, Aboelwafa AA, Hamza MY and
Mohamed MI. Microemulsion for topical delivery
of fenoprofen calcium: in-vitro and in-vivo
evaluation. J. Liposome Res. (2018) 28: 126-36.
(41) Zeb A, Qureshi OS, Kim HS, Cha JH, Kim HS and
Kim JK. Improved skin permeation of methotrexate
via nanosized ultradeformable liposomes. Int. J.
Nanomedicine. (2016) 11: 3813-24.
(42) Deuis JR, Dvorakova LS and Vetter I. Methods
used to evaluate pain behaviors in rodents. Front.
Mol. Neurosci. (2017) 10: 284.
(43) OECD. Test No. 404: Acute Dermal Irritation/
Corrosion. OECD Guidelines for the Testing
of Chemicals, Section 4, OECD Publishing,
Paris (2015) Avaiable from URL: https://doi.
org/10.1787/9789264242678-en.
(44) Shaq-un-Nabi S, Shakeel F, Talegaonkar S,
Ali J, Baboota S, Ahuja A, Khar RK and Ali M.
21
Lidocaine-loaded, Microemulsion-based Topical Gels
Formulation development and optimization using
nanoemulsion technique: a technical note. AAPS
PharmSciTech. (2007) 8: Article 28.
(45) Mishra R, Prabhavalkar KS and Bhatt LK.
Preparation, optimization, and evaluation
of zaltoprofen-loaded microemulsion and
microemulsion-based gel for transdermal delivery.
J. Liposome Res. (2016) 26: 297-306.
(46) Tadros T. Principles of emulsion stabilization
with special reference to polymeric surfactants. J.
Cosmet .Sci. (2006) 57: 153-69.
(47) Olariu I, Coneac G, Vlaia L, Vlaia V, Anghel DF,
Ilie C, Popoiu C and Lupuleasa D. Development
and evaluation of microemulsion-based hydrogel
formulations for topical delivery of propranolol
hydrochloride. Dig. J. Nanomater. Biostruct.
(2014) 9: 395-412.
(48) Chen L, Zhao X, Cai J, Guan Y, Wang S, Liu H, Zhu
W and Li J. Triptolide-loaded microemulsion-based
hydrogels: physical properties and percutaneous
permeability. Acta Pharm. Sin. B (2013) 3: 185–92.
(49) Patel MR, Patel RB, Parikh JR and Patel BG. Novel
microemulsion-based gel formulation of tazarotene
for therapy of acne. Pharm. Dev. Technol. (2016)
21: 921-32.
(50) Abd E, Yousef SA, Pastore MN, Telaprolu K,
Mohammed YH, Namjoshi S, Grice JE and Roberts
MS. Skin models for the testing of transdermal
drugs. Clin. Pharmacol. (2016) 8: 163-76.
(51) Morimoto Y, Hatanaka T, Sugibayashi K and
Omiya H. Prediction of skin permeability of drugs:
comparison of human and hairless rat skin. J.
Pharm. Pharmacol. (1992) 44: 634-9.
(52) Godin B and Touitou E. Transdermal skin delivery:
Predictions for humans from in vivo, ex-vivo and
animal models. Adv. Drug Deliv. Rev. (2007) 59:
1152-61.
(53) Kweon JH, Chi SC and Park ES. Transdermal
delivery of diclofenac using microemulsions. Arch.
Pharmacal Res. (2004) 27: 351-6.
(54) Higuchi T. Physical chemical analysis of
percutaneous absorption process from creams and
ointments. J. Soc. Cosmet. Chem. (1960) 11: 85-97.
(55) Radwan SA, ElMeshad AN and Shoukri RA.
Microemulsion loaded hydrogel as a promising
vehicle for dermal delivery of the antifungal
sertaconazole: Design, optimization and ex-vivo
evaluation. Drug Dev. Ind. Pharm. (2017) 43:
1351-65.
(56) Cao M, Ren L and Chen G. Formulation,
optimization and ex-vivo and in vivo evaluation of
celecoxib microemulsion-based gel for transdermal
delivery. AAPS PharmSciTech. (2017) 18: 1960-71.
(57) Shakeel F, Baboota S, Ahuja A, Ali J, Aqil M
and Shaq S. Nanoemulsions as vehicles for
transdermal delivery of aceclofenac. AAPS
PharmSciTech. (2007) 8: Article 104.
(58) Zhang YT, Li Z, Zhang K, Zhang HY, He ZH,
Xia Q, Zhao JH and Feng NP. Co-delivery of
evodiamine and rutaecarpine in a microemulsion-
based hyaluronic acid hydrogel for enhanced
analgesic eects on mouse pain models. Int. J.
Pharm. (2017) 528: 100-6.
(59) Williams AC and Barry BW. Penetration enhancers.
Adv. Drug Deliv. Rev. (2012) 64: 128-37.
(60) Starýchová L, Žabka M, Spaglová M, Čuchorová
M, Vitková M, Čierna M, Bartoníková K and
Gardavská K. In-vitro liberation of indomethacin
from chitosan gels containing microemulsion
in dierent dissolution mediums. J. Pharm. Sci.
(2014) 103: 3977-84.
(61) Çelebi N, Ermiş S and Özkan S. Development of
topical hydrogels of terbinane hydrochloride and
evaluation of their antifungal activity. Drug Dev.
Ind. Pharm. (2015) 41: 631-9.
(62) Chudasama A, Patel V, Nivsarkar M, Vasu K and
Shishoo C. Investigation of microemulsion system
for transdermal delivery of itraconazole. J. Adv.
Pharm. Technol. Res. (2011) 2: 30-8.
(63) Aliberti AL, de Queiroz AC, Praça FS, Eloy
JO, Bentley MV and Medina WS. Ketoprofen
microemulsion for improved skin delivery and in-
vivo anti-inammatory eect. AAPS PharmSciTech.
(2017) 18: 2783-91.
(64) Gohel MC and Nagori SA. Fabrication and
evaluation of hydrogel thickened microemulsion
of ibuprofen for topical delivery. Indian J. Pharm.
Edu. Res. (2010) 44: 189-96.
(65) Bachhav YG and Patravale VB. Microemulsion-
based vaginal gel of clotrimazole: formulation,
in-vitro evaluation, and stability studies. AAPS
PharmSciTech. (2009) 10: 476-81.
(66) Mishra B, Sahoo SK and Sahoo S. Liranaftate
loaded xanthan gum based hydrogel for topical
delivery: Physical properties and ex-vivo
permeability. Int. J. Biol. Macromol. (2018) 107:
1717-23.
(67) Aggarwal N, Goindi S and Khurana R. Formulation,
characterization and evaluation of an optimized
microemulsion formulation of griseofulvin for
topical application. Colloids Surf. B Biointerfaces.
(2013) 105: 158-66.
(68) Tasdighi E, Jafari Azar Z and Mortazavi SA.
Development and in-vitro evaluation of a
contraceptive vagino-adhesive propranolol
hydrochloride gel. Iran. J. Pharm. Res. (2012) 11:
13-26.
This article is available online at http://www.ijpr.ir
... It contains between 52% and 68% carboxylic acid groups, calculated on a dry basis (Bialik et al., 2021;Draganoiu et al., 2012) and expresses the dissociation constant as pKa of 6.0 ± 0.5 (Jaworski et al., 2021). Therefore, it is considered a weak acidic gelling agent (Daryab, Faizi, Mahboubi, & Aboofazeli, 2022;Jaworski et al., 2021) that conveyed acidity to the CBP gel bases. ...
... In addition to the gel concentration, gel viscosity can be manipulated by combining gelling agents at various compositions. Since there is an inverse relationship between the viscosity and the spreadability and/or extrudability of gels (Sanjana, Ahmed, & Bh, 2021;Budiman, Praditasari, Rahayu, & Aulifa, 2019;Daryab et al., 2022;Lee et al., 2009;Lucero, García, Vigo, & León, 1995;Maraie, & Kadhium, 2019;Pavithra, Jeganath, & Iqbal, 2018), which affects the ease of application (Jaber et al., 2020), gels need to be optimized to possess a proper viscositythat is, to simultaneously hold the formulations and to stay at site of application with sufficient retention time (Daryab et al., 2022;Jaber et al., 2020;Kotwal, Bhise, & Thube, 2007). ...
... In addition to the gel concentration, gel viscosity can be manipulated by combining gelling agents at various compositions. Since there is an inverse relationship between the viscosity and the spreadability and/or extrudability of gels (Sanjana, Ahmed, & Bh, 2021;Budiman, Praditasari, Rahayu, & Aulifa, 2019;Daryab et al., 2022;Lee et al., 2009;Lucero, García, Vigo, & León, 1995;Maraie, & Kadhium, 2019;Pavithra, Jeganath, & Iqbal, 2018), which affects the ease of application (Jaber et al., 2020), gels need to be optimized to possess a proper viscositythat is, to simultaneously hold the formulations and to stay at site of application with sufficient retention time (Daryab et al., 2022;Jaber et al., 2020;Kotwal, Bhise, & Thube, 2007). ...
Article
Full-text available
For prevention of dental caries in xerostomic patients, 2% w/w sodium fluoride oral gel is frequently prescribed; however, the producthasnotbeen officially approved for clinical use in Thailand.Therefore, this study aimedto develop 2% w/w sodium fluoride oral gelsdispensed specifically to patients in Rajavithi Hospital.Bothsingle andcombined gel bases were prepared from different gelling agents, including gelatin, xanthan gum, hydroxypropyl methylcellulose, sodium carboxymethylcellulose, and Carbopol 940P.They were subsequently characterized according totheir physicochemical properties, i.e.,clarity, spreadability, pH values, and apparent viscosity.All the gel bases were clear viscous liquids of good spreadability, except the xanthan gum gel bases, which were slightly cloudy.The pH values indicatethatthe gelatin, xanthan gum, and Carbopol 940P gel bases were acidic,while the cellulose gel bases were neutral. The apparent viscosity ranged from3 to11.5x 105cPs, depending mainly on type and concentration of the gelling agents.After preparation, the gel bases were then incorporated with 1 M pH 7.4 phosphate buffer solution to the final concentration of 5% w/w.Macroscopic characteristics of the buffered gel bases were generally unchanged, exceptfortheapparentviscosity,which decreased slightly.Sodium fluoride was subsequently added to the selected bufferedgel bases.The resulting sodium fluoride gels were neutral and transparent viscous liquids of good spreadability and exhibited good stability against the heating-cooling cycle and accelerated testing.Thus, it is conceivablethat 2% w/w sodium fluoride oral gels with acceptable physicochemical characteristics and excellent stability were successfully developed.
... Unlike hydroxypropyl methylcellulose, which must be produced in hot water, carbopol gelling agents have the advantage of being able to be developed in room temperature water. Furthermore, the wide viscosity range of 40,000-60,000 cps of carbopol 940 led to its selection for hydrogel preparations (Daryab et al. 2022). The concentration of carbopol 940 gelling agent directly influences the viscosity of the preparation, which also influences its physical characteristics (Daood et al. 2019). ...
Article
Full-text available
Octadecyl 3-(3,5-di-tert-butyl-4-hydroxyphenyl) propanoate (ODHP) was extracted in a previous study from the culture broth of soil isolate Alcaligenes faecalis MT332429 and showed a promising antimycotic activity. This study was aimed to formulate ODHP loaded β-cyclodextrins (CD) nanosponge (NS) hydrogel (HG) to control skin fungal ailments since nanosponges augment the retention of tested agents in the skin. Box-Behnken design was used to produce the optimized NS formulation, where entrapment efficiency percent (EE%), polydispersity index (PDI), and particle size (PS) were assigned as dependent parameters, while the independent process parameters were polyvinyl alcohol % (w/v %), polymer-linker ratio, homogenization time, and speed. The carbopol 940 hydrogel was then created by incorporating the nanosponges. The hydrogel fit Higuchi’s kinetic release model the best, according to in vitro drug release. Stability and photodegradation studies revealed that the NS-HG remained stable under tested conditions. The formulation also showed higher in vitro antifungal activity against Candida albicans compared to the control fluconazole. In vivo study showed that ODHP-NS-HG increased survival rates, wound contraction, and healing of wound gap and inhibited the inflammation process compared to the other control groups. The histopathological examinations and Masson’s trichrome staining showed improved healing and higher records of collagen deposition. Moreover, the permeability of ODHP-NS-HG was higher through rats’ skin by 1.5-folds compared to the control isoconazole 1%. Therefore, based on these results, NS-HG formulation is a potential carrier for enhanced and improved topical delivery of ODHP. Our study is a pioneering research on the development of a formulation for ODHP produced naturally from soil bacteria. Key points • Octadecyl 3-(3,5-di-tert-butyl-4-hydroxyphenyl) propanoate was successfully formulated as a nanosponge hydrogel and statistically optimized. • The new formula exhibited in vitro good stability, drug release, and higher antifungal activity against C. albicans as compared to the fluconazole. • Ex vivo showed enhanced skin permeability, and in vivo analysis showed high antifungal activity as evidenced by measurement of various biochemical parameters and histopathological examination.
... Microemulsions (Mes) consist of an oil phase, water phase, surfactant, and cosurfactant, and can be divided into water-inoil (w/o) type, water-in-oil (o/w) type, and two-phase continuous type according to the structure (Karasulu, 2008;Gradzielski et al., 2021). Mes systems exhibit several advantages, including a strong solubilizing power, which can dissolve various drugs with different polarities and improve the solubility and bioavailability of drugs; thermodynamic stability, which enhances the stability of drugs; optical clarity, which is easy to prepare; strong permeability, high levels of diffusion, and a high absorption rate, and the ability to easily penetration into the keratin layer (vadlamudi et al., 2014;Daryab et al., 2022). therefore, Mes are a good vehicle for drug delivery. ...
Article
Full-text available
Pain management remains among the most common and largely unmet clinical problems today. Local anesthetics play an indispensable role in pain management. The main limitation of traditional local anesthetics is the limited duration of a single injection. To address this problem, catheters are often placed or combined with other drugs in clinical practice to increase the time that local anesthetics act. However, this method does not meet the needs of clinical analgesics. Therefore, many researchers have worked to develop local anesthetic extended-release types that can be administered in a single dose. In recent years, drug extended-release systems have emerged dramatically due to their long duration and efficacy, providing more possibilities for the application of local anesthetics. This paper summarizes the types of local anesthetic drug delivery systems and their clinical applications, discusses them in the context of relevant studies on local anesthetics, and provides a summary and outlook on the development of local anesthetic extended-release agents.
... At 24 h after application, the cumulative permeability of Cur from the Cur solution, Cur-ME and CME-KCS gel was 1.08 ± 0.19, 4.83 ± 0.45 and 3.78 ± 0.26 µg/cm 2 , respectively. Both the Cur-ME and CME-KCS gel showed better enhancement in Cur penetration than the Cur solution, which might be due to the fact that ingredients such as PEG400 and GTCC in ME act as permeation enhancers and could significantly change the structure of the epidermis [46]. In addition, the nanosized ME with very small droplets and a huge surface area could increase the concentration of the drug in the epidermal layer, thus causing the transfer of the drug into the skin, forming a higher concentration gradient and providing the driving force for drug penetration through the stratum corneum [47]. ...
Article
Full-text available
Curcumin (Cur) is a kind of polyphenol with a variety of topical pharmacological properties including antioxidant, analgesic and anti-inflammatory activities. However, its low water solubility and poor skin bioavailability limit its effectiveness. In the current study, we aimed to develop microemulsion-based keratin–chitosan gel for the improvement of the topical activity of Cur. The curcumin-loaded microemulsion (CME) was formulated and then loaded into the keratin–chitosan (KCS) gel to form the CME-KCS gel. The formulated CME-KCS gel was evaluated for its characterization, in vitro release, in vitro skin permeation and in vivo activity. The results showed that the developed CME-KCS gel had an orange-yellow and gel-like appearance. The particle size and zeta potential of the CME-KCS gel were 186.45 ± 0.75 nm and 9.42 ± 0.86 mV, respectively. The CME-KCS gel showed desirable viscoelasticity, spreadability, bioadhesion and controlled drug release, which was suitable for topical application. The in vitro skin permeation and retention study showed that the CME-KCS gel had better in vitro skin penetration than the Cur solution and achieved maximum skin drug retention (3.75 ± 0.24 μg/cm2). In vivo experimental results confirmed that the CME-KCS gel was more effective than curcumin-loaded microemulsion (Cur-ME) in analgesic and anti-inflammatory activities. In addition, the CME-KCS gel did not cause any erythema or edema based on a mice skin irritation test. These findings indicated that the developed CME-KCS gel could improve the skin penetration and retention of Cur and could become a promising formulation for topical delivery to treat local diseases.
... They offer several advantages, including improved drug solubility, increased bioavailability of poorly water-soluble drugs, and enhanced chemical stability [16,[21][22][23]. The intensive utilization of MEs spans various administration routes, including oral [24][25][26][27], topical [28][29][30][31], and parenteral [32][33][34][35], establishing them as promising carrier systems for delivering a broad spectrum of active ingredients, including drugs that are either water-soluble or poorly water-soluble [11,36]. ...
Article
Full-text available
Microemulsions (MEs) have gained prominence as effective drug delivery systems owing to their optical transparency, low viscosity, and thermodynamic stability. MEs, when stabilized with surfactants and/or co-surfactants, exhibit enhanced drug solubilization, prolonged shelf life, and simple preparation methods. This review examines the various types of MEs, explores different preparation techniques, and investigates characterization approaches. Plant extracts and bioactive compounds are well established for their utilization as active ingredients in the pharmaceutical and cosmetic industries. Being derived from natural sources, they serve as preferable alternatives to synthetic chemicals. Furthermore, they have demonstrated a wide range of therapeutic effects, including anti-inflammatory, antimicrobial, and antioxidant activities. However, the topical application of plant extracts and bioactive compounds has certain limitations, such as low skin absorption and stability. To overcome these challenges, the utilization of MEs enables enhanced skin absorption, thereby making them a valuable mode of administration. However, considering the significant surfactant content in MEs, this review evaluates the potential skin irritation caused by MEs containing herbal substances. Additionally, the review explores the topical application of MEs specifically for herbal substances, with an emphasis on their anti-inflammatory properties.
... Nanoscale drug delivery systems (DDS) have been designed as inert systems that can transport drugs at the target site with high selectivity and controllable kinetic release profile [6,7]. Various DDS, such as niosomes, liposomes, transferosomes, and solid lipid nanoparticles, have already been employed to improve LID efficacy due to the role of these systems as skin permeation enhancers able to increase drug permeability, and consequently, ensure a higher drug accumulation at the target site and thus, an improved analgesic effect [8][9][10][11]. Their better permeation ability allows for lower drug doses and could reduce cardiovascular side effects [12,13]. ...
Article
Full-text available
Lidocaine is a local anaesthetic drug with an amphiphilic structure able to self-associate, under certain conditions, in molecular aggregates playing the role of both carrier and drug. The aim of this study was to determine the optimal conditions for obtaining vesicular carriers, called lidosomes. The new formulations were obtained using both lidocaine and lidocaine hydrochloride and different hydration medias (distilled water, acid, and basic aqueous solution). Lidosomes formulations were characterized in terms of size, ζ-potential, drug retained, stability formulation, and ex vivo permeation profile. Moreover, lidosomes were incorporated in two different gel structures: one based on carboxymethylcellulose and one based on pluronic F-127 to achieve suitable properties for a topical application. Results obtained showed that lidocaine showed a better performance to aggregate in vesicular carriers in respect to hydrochloride form. Consequently, only formulations comprised of lidocaine were studied in terms of skin permeation performance and as carriers of another model drug, capsaicin, for a potential combined therapy. Lidocaine, when in form of vesicular aggregates, acted as percutaneous permeation enhancer showing better permeation profiles with respect to drug solutions. Moreover, lidosomes created a significant drug depot into the skin from which the drug was available for a prolonged time, a suitable feature for a successful local therapy.
... After the abdominal hair of the rats was shaved with an electric shaver, the animals were sacrificed by ether inhalation anesthesia, the abdominal skin was excised, adipose tissue and adhesions were removed, and then the skin was hydrated with physiological saline and stored at 4°C for use (Daryab et al., 2022). ...
Article
Full-text available
The present study focused on the development of Cur-loaded SOHA nanogels (Cur-SHNGs) to enhance the topical administration of Cur. The physiochemical properties of Cur-SHNGs were characterized. Results showed that the morphology of the Cur-SHNGs was spherical, the average size was 171.37 nm with a zeta potential of −13.23 mV. Skin permeation experiments were carried out using the diffusion cell systems. It was found that the skin retention of Cur-SHNGs was significantly improved since it showed the best retention value (0.66 ± 0.17 μg/cm²). In addition, the hematoxylin and eosin staining showed that the Cur-SHNGs improved transdermal drug delivery by altering the skin microstructure. Fluorescence imaging indicated that Cur-SHNGs could effectively deliver the drug to the deeper layers of the skin. Additionally, Cur-SHNGs showed significant analgesic and anti-inflammatory activity with no skin irritation. Taken together, Cur-SHNGs could be effectively used for the topical delivery of therapeutic drugs.
Article
Full-text available
We previously demonstrated that baicalin had efficacy against gouty arthritis (GA) by oral administration. In this paper, a novel baicalin-loaded microemulsion-based gel (B-MEG) was prepared and assessed for the transdermal delivery of baicalin against GA. The preparation method and transdermal capability of B-MEG was screened and optimized using the central composite design, Franz diffusion cell experiments, and the split-split plot design. Skin irritation tests were performed in guinea pigs. The anti-gout effects were evaluated using mice. The optimized B-MEG comprised of 50 % pH 7.4 phosphate buffered saline, 4.48 % ethyl oleate, 31.64 % tween 80, 13.88 % glycerin, 2 % borneol, 0.5 % clove oil and 0.5 % xanthan gum, with a baicalin content of (10.42 ± 0.08) mg/g and particle size of (15.71 ± 0.41) nm. After 12 h, the cumulative amount of baicalin permeated from B-MEG was (672.14 ± 44.11) μg·cm⁻². No significant skin irritation was observed following B-MEG application. Compared to the model group, B-MEG groups significantly decreased the rate of auricular swelling (P < 0.01) and number of twists observed in mice (P < 0.01); and also reduced the rate of paw swelling (P < 0.01) and inflammatory cell infiltration in a mouse model of GA. In conclusion, B-MEG represents a promising transdermal carrier for baicalin delivery and can be used as a potential therapy for GA.
Article
Pharmaceutical industries and drug regulatory agencies are inclining towards continuous manufacturing due to better control over the processing conditions and in view to improve product quality. In the present work, continuous manufacturing of O/W emulgel by melt extrusion process was explored using lidocaine as an active pharmaceutical ingredient. Emulgel was characterized for pH, water activity, globule size distribution, and in vitro release rate. Additionally, effect of temperature (25°C and 60°C) and screw speed (100, 300, and 600 rpm) on the globule size and in vitro release rate was studied. Results indicated that at a given temperature, emulgel prepared under screw speed of 300 rpm resulted in products with smaller globules and faster drug release.Graphical Abstract
Article
Full-text available
Postherpetic neuralgia (PHN) is the most troublesome side effect of Herpes Zoster (HZ), which mainly affects the elderly and immunocompromised populations. Despite the current advancement of treatments, PHN persists in many individuals influencing their daily activities and reducing their quality of life. Anticonvulsants, antidepressants, topical therapies including lidocaine and capsaicin, and opioids, are the most widely used therapies for the treatment of PHN. These medications come with their adverse effects, so they should be used carefully with the elderly or with patients with significant comorbidities. Other measures like botulinum toxin, nerve blocks, spinal cord stimulation, and radiofrequency have also contributed significantly to the management of PHN. However, the efficacy, safety, and tolerability of these invasive methods need to be carefully monitored when administering them. Early diagnosis and early initiation of treatment can reduce the burden associated with PHN. The zoster vaccine has effectively reduced the incidence of HZ and PHN. In this article, we discuss the treatment options available for the management of PHN, mainly focusing on the efficacy and safety of different therapeutic modalities.
Article
Full-text available
Babassu oil extraction is the main income source in nut breakers communities in northeast of Brazil. Among these communities, babassu oil is used for cooking but also medically to treat skin wounds and inflammation, and vulvovaginitis. This study aimed to evaluate the anti-inflammatory activity of babassu oil and develop a microemulsion system with babassu oil for topical delivery. Topical anti-inflammatory activity was evaluated in mice ear edema using PMA, arachidonic acid, ethyl phenylpropiolate, phenol, and capsaicin as phlogistic agents. A microemulsion system was successfully developed using a Span® 80/Kolliphor® EL ratio of 6 : 4 as the surfactant system (S), propylene glycol and water (3 : 1) as the aqueous phase (A), and babassu oil as the oil phase (O), and analyzed through conductivity, SAXS, DSC, TEM, and rheological assays. Babassu oil and lauric acid showed anti-inflammatory activity in mice ear edema, through inhibition of eicosanoid pathway and bioactive amines. The developed formulation (39% A, 12.2% O, and 48.8% S) was classified as a bicontinuous to o/w transition microemulsion that showed a Newtonian profile. The topical anti-inflammatory activity of microemulsified babassu oil was markedly increased. A new delivery system of babassu microemulsion droplet clusters was designed to enhance the therapeutic efficacy of vegetable oil.
Article
Full-text available
Rodents are commonly used to study the pathophysiological mechanisms of pain as studies in humans may be difficult to perform and ethically limited. As pain cannot be directly measured in rodents, many methods that quantify “pain-like” behaviors or nociception have been developed. These behavioral methods can be divided into stimulus-evoked or non-stimulus evoked (spontaneous) nociception, based on whether or not application of an external stimulus is used to elicit a withdrawal response. Stimulus-evoked methods, which include manual and electronic von Frey, Randall-Selitto and the Hargreaves test, were the first to be developed and continue to be in widespread use. However, concerns over the clinical translatability of stimulus-evoked nociception in recent years has led to the development and increasing implementation of non-stimulus evoked methods, such as grimace scales, burrowing, weight bearing and gait analysis. This review article provides an overview, as well as discussion of the advantages and disadvantages of the most commonly used behavioral methods of stimulus-evoked and non-stimulus-evoked nociception used in rodents.
Article
Local anesthetics are used for performing various regional anesthesia techniques to provide intraoperative anesthesia and analgesia, as well as for the treatment of acute and chronic pain. Older medications such as lidocaine and bupivacaine as well as newer ones such as mepivacaine and ropivacaine are being used successfully for decades. Routes of administration include neuraxial, perineural, intravenous, various infiltrative approaches, topical, and transdermal. There are new innovations with the use of older local anesthetics in a novel manner, in addition to the development and use of new formulations. This chapter seeks to summarize the pharmacokinetics of local anesthetics and address the role of newer local anesthetics, as well as clinical implications, safety profiles, and the future of local anesthetic research. Finally, some clinical pearls are highlighted.
Article
Deficiency or altered composition of stratum corneum (SC) lipids such as ceramides (CERs), causing skin barrier dysfunction and skin dryness, have been associated with skin diseases such as atopic dermatitis and psoriasis, and ageing. Replenishing the depleted native CERs with exogenous CERs has also been shown to have beneficial effects in restoring the skin barrier. Phyto-derived CERs such as oat CERs were shown to be potential for skin barrier reinforcement. To effect this, however, the oat CERs should overcome the SC barrier and delivered deep into the lipid matrix using the various novel formulations. In an attempt to demonstrate the potential use of oat CERs, lecithin-based microemulsions (MEs) and starch-based nanoparticles (NPs) were formulated and characterized. Besides, ME gel and NP gel were also prepared using Carbopol®980 as a gelling agent. The in vitro release and penetration (using artificial four-layer membrane system) and ex vivo permeation (using excised human skin) of oat CERs from the various formulations were investigated. The results revealed ME enhanced the in vitro release and penetration oat CERs compared to the other formulations. On the other hand, the NPs retarded the release of oat CERs and small quantities of oat CERs incorporated into NP gel penetrated into the deeper layers of the multilayer membranes. The penetration-enhancing effect of ME was also observed in the ex vivo permeation studies where significant quantities of oat CERs were found in the acceptor compartment. Compared to the ME, the ME gel exhibited reduced depth and extent of oat CERs permeation. As compared to NP gel, ME gel enhanced the degree of permeation of oat CERs into the deeper layer of the skin. Generally the gel formulations were effective in concentrating oat CERs in the SC where they are needed to be.
Article
A topical microemulsion (ME)-based hydrogel was developed to enhance permeation of an antifungal drug, liranaftate (LRFE) for effective eradication of cutaneous fungal infection. Pseudo-ternary phase diagrams were used to determine the existence of MEs region. ME formulations were prepared with Di-isopropyl adaptate, Cremophore-EL, Ethanol and distilled water. Xanthan Gum (1.5% w/w) was used for preparation of hydrogel of LRFE microemulsion (HLM) and characterized. The in vitro and in vivo evaluation of prepared HLM and saturated drug solution were compared. The viscosity, average droplet size and pH of HLM were 142.30±0.42 to 165.15±0.21Pa.s, 52.53-93.40nm and 6.6-7.1, respectively. Permeation rate of LRFE from optimized formulation (HLM-3), composed with Di-isopropyl adaptate (4.5% w/w), Cremophor-EL (30% w/w), Ethanol (10% w/w) and water (52% w/w) was observed higher in compare with other HLMs and saturated drug solution. HLM-3 was stable, six times higher drug deposition capacity in skin than saturated drug solution and did not caused any erythema based on skin sensitivity study on rat. The average zone of inhibition of HLM-3 (25.52±0.26mm) was higher in compare with saturated drug solution (13.44±0.40mm) against Candida albicans.
Article
The aim of this study was to improve the analgesic effect of evodiamine and rutaecarpine, using a microemulsion-based hydrogel (ME-Gel) as the transdermal co-delivery vehicle, and to assess hyaluronic acid as a hydrogel matrix for microemulsion entrapment. A microemulsion was formulated with ethyl oleate as the oil core to improve the solubility of the alkaloids and was loaded into a hyaluronic acid-structured hydrogel. Permeation-enhancing effects of the microemulsion enabled evodiamine and rutaecarpine in ME-Gel to achieve 2.60- and 2.59-fold higher transdermal fluxes compared with hydrogel control (p < 0.01). The hyaluronic acid hydrogel-containing microemulsion exhibited good skin biocompatibility, whereas effective ME-Gel co-delivery of evodiamine and rutaecarpine through the skin enhanced the analgesic effect in mouse pain models compared with hydrogel. Notably, evodiamine and rutaecarpine administered using ME-Gel effectively down-regulated serum levels of prostaglandin E2, interleukin 6, and tumor necrosis factor α in formaldehyde-induced mouse pain models, possibly reflecting the improved transdermal permeability of ME-Gel co-delivered evodiamine and rutaecarpine, particularly with hyaluronic acid as the hydrogel matrix.
Article
The purpose of the present study was to develop and optimize sertaconazole microemulsion-loaded hydrogel (STZ ME G) to enhance the dermal delivery and skin retention of the drug. Following screening of various oils for maximum drug solubility, twelve pseudoternary phase diagrams were constructed using oils (Peceol®, Capryol® 90), surfactants (Tween® 80, Cremophor® EL), a cosurfactant (Transcutol® P) and water. A 2¹x 3¹x 2¹x 3¹ full factorial design was employed to optimize a ME of desirable characteristics. The MEs were formulated by varying the oil type, oil concentration, surfactant type, and surfactant: cosurfactant ratio. Optimized ME formulae F22 [5% Peceol®, 55% Tween® 80: Transcutol® (1:2), 40% water] and F31 [5% Peceol®, 55% Cremophor® EL: Transcutol® (1:2), 40% water] acquired mean droplet size of 75.21 and 8.68 nm, and zeta potential of 34.65 and 24.05 mV, respectively. Since F22 showed higher STZ skin retention during ex vivo studies (686.47 µg/cm²) than F31 (338.11 µg/cm²); hence it was incorporated in 0.5% Carbopol 934 gel to augment STZ skin retention capability. STZ ME G exhibited higher STZ skin retention (1086.1 µg/cm²) than the marketed product ‘Dermofix® cream’ (270.3 µg/cm²). The antimycotic activity against C.albicans revealed increased zones of inhibition for F22 and STZ ME G (35.75 and 30.5 mm, respectively) compared to Dermofix® cream (26 mm). No histopathological changes were observed following topical application of STZ ME G on rats’ skin (n = 9). Overall, the obtained results confirmed that the fabricated formulation could be a promising vehicle for the dermal delivery of STZ.
Chapter
The current research in cutaneous and percutaneous drug delivery is focused on design of advanced carriers with enhanced applicability, stability, drug-loading capacity, and drug penetration/permeation ability. Microemulsion hydrogels (MHs) are carriers which attract the attention of a growing number of researchers who aimed to enhace dermal or transdermal delivery of drugs which are common therapeutics for different skin disorders or systemic deseases such as infections, androgenic alopecia, rheumatoid arthritis, osteoarthritis, and spondylitis, respectively. Microemulsion hydrogels is a heterogeneous group of the drug delivery systems which usually represent thermodynamically stable systems comprising dispersed oil phase within a continuous aqueous phase which is thickened with a suitable hydrophilic polymer. This chapter summarises a novel observations regarding physicochemical properties, physical and chemical stability, and drug delivery potential of this type of drug delivery systems, based on comprehensive review of the research results from the relevant scientific publications. Particularly, the chapter describes the rheological behaviour of microemulsion hydrogels and elucidates the role of the most commonly used synthetic and natural hydrophilic polymers (such as carbomers, hydroxypropyl cellulose (HPC), hydroxypropyl methylcellulose (HPMC), xanthan gum, poloxamers) as gelling agents. Furthermore, the recently established models of the complex structure of different microemulsion hydrogels, the in vitro drug release profiles, the ex vivo permeation profiles, and the in vivo drug delivery availability of different drug molecules from the investigated systems, are described. The skin compatibility and skin irritation potential aspect of the selected hydrogel-thickened microemulsion systems is commented. The most important findings of the selected studies on microemulsion hydrogels were presented in order to illustrate their potential for achievement of topical, regional or transdermal drug delivery, including sustained drug delivery, as well as to compare to that of conventional hydrogels, microemulsions, and creams.