European Journal of Pharmaceutical Sciences: Meng Li, Yunjing Li, Weiwei Liu, Rongli Li, Cuiying Qin, Nan Liu, Jing Han

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

European Journal of Pharmaceutical Sciences 93 (2016) 224–232

Contents lists available at ScienceDirect

European Journal of Pharmaceutical Sciences

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

The preparation of Cistanche phenylethanoid glycosides liquid


proliposomes: Optimized formulation, characterization and proliposome
dripping pills in vitro and in vivo evaluation
Meng Li, Yunjing Li, Weiwei Liu, Rongli Li, Cuiying Qin, Nan Liu, Jing Han ⁎
School of Pharmaceutical Engineering, Shenyang Pharmaceutical University, China

a r t i c l e i n f o a b s t r a c t

Article history: Water-soluble Cistanche phenylethanoid glycosides (CPhGs) have poor permeability and low bioavailability.
Received 26 January 2016 However, liposomes can improve the permeability of such drugs and their poor stability, and proliposomes
Received in revised form 13 July 2016 have been used to overcome these problems. Based on this, Cistanche phenylethanoid glycoside liquid
Accepted 30 July 2016
proliposomes (CPhGsP) and dripping(?) pills were prepared and optimized using response surface methodology.
Available online 1 August 2016
The properties of CPhGsP were evaluated in terms of their encapsulation efficiency, particle size, zeta potential,
Keywords:
and morphology. The results obtained showed that the optimal formulation was drug/soybean phospholipid/
Liquid proliposomes poloxamer-188/sodium deoxycholate/propylene glycol 1:22.38:3.52:0.84:80 (w/w/w/w/v). This resulted in an
Cistanche phenylethanoid glycosides encapsulation efficiency, particle size, and zeta potential of hydrated proliposomes with phosphate buffer solu-
Stability tion (pH 7.4) of 51.97%, 671.7 nm, and −25.49 mV, respectively. Stability testing of CPhGsP and CPhGs ordinary
Dripping pills liposomes was carried out for 3 months at 4 ± 2 °C, 25 ± 2 °C, 40 ± 2 °C, 75 ± 5% RH. The results obtained
In vitro release showed that the stability of the proliposomes was better than that of ordinary liposomes at the same tempera-
ture, while a lower temperature of 4 °C is ideal for storage. Cistanche phenylethanoid glycoside liquid
proliposomes dripping pills (CPhGsPD) are efficiently released in gastrointestinal solution as shown by in vitro
release experiments and the structure of the liposomes does not destroy the proliposome dripping pills by hydra-
tion. In vivo experiments showed that the areas under the plasma level-time curves and peak concentrations of
CPhGsPD and hydrated proliposomes were higher than those of CPhGs. Moreover, with CPhGsPD, the pharmaco-
kinetic parameters were similar to those with hydrated proliposomes. These results showed that CPhGsPD offer a
good way to improve the oral delivery of CPhGs.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction hepatoprptective effects (Morikawa et al., 2010), anti-inflammatory ac-


tivity (Nan et al., 2013; Lin et al., 2002) and the ability to increase anti-
The stem of Cistanche deserticola Y.C. Ma is a common traditional body production (Maruyama et al., 2008). However, PhGs have poor
Chinese medicine which is widely used in the treatment of kidney defi- permeability and low in vivo bioavailability (Li et al., 2015; Jia et al.,
ciency, body weakness and constipation. It is known as “desert ginseng”, 2006; Cui et al., 2016; Jia et al., 2009).
because it can grow in arid deserts. The main bioactive ingredient is Liposomes are colloidal particles that are almost spherical micro-
phenylethanoid glycosides (PhGs), which is a group of water-soluble scopic vesicles and are composed of one or more lipid bilayers arranged
natural products. Currently, over 34 kinds of PhGs have been found in in a concentric fashion enclosing a number of aqueous compartments
Cistanche (Jiang and Tu, 2009), and echinacoside (Fig. 1a) and acteoside (Bangham et al., 1965). These phospholipid vesicles formed by hydro-
(Fig. 1b) are the representative and major active ingredients exhibiting philic and hydrophobic molecules have the potential to encapsulate li-
a variety of bioactivities, especially echinacoside. Modern pharmacolog- pophilic and hydrophilic substances. The structure of liposome
ical studies have shown that PhGs from Cistanche have many medicinal vesicles is similar to that of biofilms, which can improve drug perme-
activities such as neuroprotective effects (Geng et al., 2004; Chen et al., ability and promote drug absorption. The location of a drug in the vesi-
2007; Zhang et al., 2015), antifatigue (Cai et al., 2010), and cles depends on its physicochemical characteristics and the types of
lipids used in preparation of the liposomes (Gregoriadis, 1976). Lipo-
somes are composed of phospholipids which are biodegradable, non-
⁎ Corresponding author at: Shenyang Pharmaceutical University, Wenhua Road No.
toxic and free from any antigenic, pyrogenic or allergic reactions.
103, Shenyang 110016, China. Accordingly, liposomes have been extensively investigated for drug de-
E-mail address: huagonglou314@163.com (J. Han). livery, drug targeting, controlled release and to increase solubility

http://dx.doi.org/10.1016/j.ejps.2016.07.020
0928-0987/© 2016 Elsevier B.V. All rights reserved.
M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232 225

OGlc
a O
O
O
HO
O
OH

RhaO OH
HO
OH

OH
b O
O O
HO
O
OH
RhaO OH
HO
OH
Glc=glucose Rha=rhamnose

Fig. 1. Chemical structures of echinacoside (a) and acteoside (b).

(Yadav et al., 2011). In recent years, liposomes have attracted much at- values and the interaction of the variables (Vicente et al., 1998;
tention as drug carriers to improve the effect of treatment, reduce side Xiong et al., 2009). In addition, RSM saves time and effort com-
effects, and improve stability by protecting drugs from degradation or pared with other approaches, because it reduces the number of
transformation (Mi and Burke, 1994; Gabizon et al., 1982). However, li- experimental trials needed (Liyana-Pathirana and Shahidi, 2005;
posomes have a number of physicochemical stability issues such as Lee et al., 2006). Therefore, RSM is widely used to optimize process
phospholipid degradation by hydrolysis or oxidation and aggregation, parameters.
sedimentation and leakage of encapsulated drug in aqueous dispersions The aim of this study was to develop liquid proliposomes for the
which limits their use for oral drug delivery (Ariën et al., 1993; Grit et al., water-soluble drug Cistanche phenylethanoid glycosides (CPhGs) and
1989; Hunt and Tsang, 1981; Kumar et al., 2001). further improve its bioavailability and increase the liposomal stability.
In order to solve these problems and help in the development of a In addition, RSM was used to optimize the preparation of CPhGs liquid
new drug delivery system, proliposomes can be used. The concept of proliposomes. Dripping pills were selected as containers for Cistanche
proliposomes was first proposed by Payne et.al in 1986, who defined phenylethanoid glycoside proliposomes (CPhGsP). During dissolution
proliposomes as dry, free flowing powders (Payne et al., 1986). of the dripping pills, proliposomes were released and hydrated to
Proliposomes are capable of forming liposomal suspensions on form Cistanche phenylethanoid glycoside liposomes (CPhGsL). The
being added to the water phase or under in vivo conditions (Payne proliposomes were characterized by their particle size, drug encapsula-
et al., 1986). In recent years, liquid proliposomes have become a tion efficiency (EE %), zeta potential and morphology after hydration.
“hot topic” for research. According to a number of reports, The drug dissolution rate of Cistanche phenylethanoid glycoside liquid
proliposomes are a kind of transparent solution which form drug- proliposomes dripping pills (CPhGsPD) were evaluated in phosphate
loaded liposomes when blank proliposomes are mixed with a physi- buffer solution (pH 7.4) and hydrochloric acid buffer solution
ological saline solution containing the drug (Junping et al., 2000). (pH 1.2). Finally, the accelerated stability testing of CPhGsP and CPhGs
Compared with solid proliposomes, the preparation process for ordinary liposomes were also studied for 3 months at 4 ± 2 °C, 25 ±
liquid proliposomes is simpler and requires no special equipment. 2 °C, 40 ± 2 °C, 75 ± 5% RH.
In addition, the liquid proliposomes can automatically and rapidly
produce a liposome suspension and dissolve completely in the 2. Materials and methods
water phase. The mean particle size of liposomes hydrated from
liquid proliposomes is smaller and uniformly distributed. Currently, 2.1. Materials
less-relevant studies about liquid proliposomes for oral administra-
tion are being reported. CPhGs (purity N 80%) were obtained from the Hetian Emperor Chen
Response surface methodology (RSM) is a technique used to Medicine Biological Technology Co., Ltd. (Xinjiang, China). Echinacoside
optimize multifactor experiments and empirically derive a func- was obtained from the Chengdu Yi Te Mu Biological Technology Co., Ltd.
tional relationship between one or more experimental responses (Chengdu, China). Chlorogenic acid (CA, purity ≥ 96.2%) was obtained
and a set of input variables (Hamsaveni et al., 2001; Chiang et al., from the China Food and Drug Inspection Institute (Beijing, China).
2003; Zhang et al., 2007). RSM can be easily applied in response Tween 80 and Span 20 were obtained from Sinopharm Chemical Re-
to the value of each factor level through the process of regression, agent Co., Ltd. (Shenyang, China) and sodium deoxycholate was from
response surface and contour lines. On the basis of the response Shanghai Suo Lai Bao Biological Technology Co., Ltd. (Shanghai, China).
value of each factor on the level, it can predict the optimum re- Poloxamer-188 (P-188) was obtained from BASF Co., Ltd.
sponse values and the corresponding experimental conditions, (Ludwigshafen, Germany) and soybean phospholipid (SP) was a gener-
and describe the relationship between the response and the inde- ous gift from Jiangsu Man Shi Biological Technology Co., Ltd. (Jiangsu,
pendent variables, and the impression between the response China). All reagents used were of analytical grade or better.
226 M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232

Fig. 2. Proliposomes, hydrated liposomes, and dropping pills preparation process.

2.2. Preparation of CPhGsP, CPhGsPD and hydrated liposomes 2.4. Determination of Cistanche phenylethanoid glycosides

The preparation of proliposomes, hydrated liposomes, and dripping The content of CPhGs in proliposomes was determined by ultraviolet
pills are shown in Fig. 2. A mixture of CPhGs, soybean phospholipids, (UV)-visible detection, and 333 nm was selected as detection wave-
poloxamer-188 and sodium deoxycholate were dissolved in propylene length. Echinacoside was used as a reference. Eight milligrams of
glycol by sonication(ultrasound 3 s, interval 2 s, 160 W, 15 min) echinacoside was placed in a 50 mL volumetric flask and made up to
(JY92-2D, Ningbo, China), and then filled into ampoules and sealed the mark with methanol. Volumes of 0.5, 1.0, 1.5, 2.0, and 2.5 mL were
after the removal of oxygen using sterile nitrogen gas. It was a red pipetted into a 10 mL volumetric flask and made up to the mark with
brown transparent liquid. The dripping pills were made in our laborato- methanol and the absorbances were measured at 333 nm, with metha-
ry and consisted of proliposome/P-1881:4 (v/w). The coolant was poly- nol as a blank control. The standard curve was constructed by plotting
dimethylsiloxane, the cooling temperature was 2 ~ 10 °C, the melting the concentration of echinacoside versus the absorbance: A =
temperature was 80 °C, and a drip rate of 45 drops per minute was 0.0233C-0.0167 (r = 0.9997). Thelinearity was good over the range
used when preparing the dropping pills. These dropping pills were 8 ~ 40 μg·mL−1. All the recoveries were 97 ~ 102%, with relative stan-
dried for 12 h at room temperature. dard deviations b 2%. The intra- and inter-day relative standard devia-
Hydrated liposomes were formed automatically by dropping phos- tions were b2%.
phate buffer solution (pH 7.4) on to CPhGsP and shaking the mixture
manually for 2 min. 2.5. Characterization of CPhGsP

2.5.1. Measurement of size and zeta potential


2.3. Box-Behnken design of the preparation conditions The particle size, zeta potential and distribution of the hydrated lipo-
somes were measured using a laser particle size analyzer (PSS
Base on the single factor experiments, we found that the ratio of Nicomp380ZLS, USA). Proliposomes were hydrated in phosphate buffer
soybean phospholipids to drug (A), the ratio of poloxamer-188 to (pH 7.4) for size measurements. The measurements were carried out in
drug (B) and the ratio of sodium deoxycholate to drug (C) were triplicate and the results were expressed as mean ± standard deviation.
the main factors influencing the entrapment rate of CPhGsP and
particle size. Therefore, the three factors mentioned above at three Table 1
levels were used and seventeen experiments were carried out Levels of factors used in BBD.
according to the Box-Behnken design (BBD). The range and values
Range and levels
of the experimental variable investigated in this study are presented Factors Code
−1 0 1
in Table 1. The response surfaces that described the interaction
between each factor and index were drawn according to the fitting Ratio of soybean phospholipids to drug (w/w) A 15:1 20:1 25:1
equation. Experimental data were analyzed by Design-Expert trial Ratio of poloxamer-188 to drug (w/w) B 3:1 4:1 5:1
Ratio of sodium deoxycholate to drug (w/w) C 0.5:1 1:1 1.5:1
software version 8.0.6.
M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232 227

2.5.2. Transmission electron microscopy were above 80%, which was within the acceptable limits to meet the
The morphology of the hydrated CPhGsL was examined by Trans- guidelines for bioanalytical methods.
mission Electron Microscopy (TEM) with negative staining (Dayan
and Touitou, 2000). Samples were dyed using 2% phosphotungstic 2.8.2. Pharmacokinetic design
acid, and deposited on a carbon-coated copper grid to form a carbon Eighteen male New Zealand rabbits with an average weight of 2.1 ±
film. The samples were the examined by TEM after drying at room 0.42 kg were used in this study. The rabbits were obtained from the Lab-
temperature. oratory Animal Center of Shenyang Pharmaceutical University (Shen-
yang, China). They were kept in an environmentally controlled
2.5.3. Evaluation of entrapment efficiency breeding room before starting the experiments and fed with standard
Proliposomes were hydrated to form liposomal suspensions with 10 laboratory food and water ad libitum and fasted overnight before the
volumes of phosphate buffer (pH 7.4). Then 0.5 mL of the liposomes was test. The rabbits were divided into three groups and each animal re-
diluted with 10 mL methanol followed by ultraviolet (UV)-visible detec- ceived a dose in one of the following dosage forms: (1) the CPhGs
tion at 333 nm, the absorbance was `, and the calculated concentration group was given CPhGs at a dose of 25 mg·kg− 1; (2) the hydrated
was C1. One milliliter samples of liposomes were placed in 1.5 mL cen- proliposome group was given hydrated proliposomes at a dose of
trifuge Tubes (Eppendorf Tubes, EP Tubes) and placed in a TGL-16G cen- 25 mg·kg−1; (3) the CPhGsPD group was given CPhGsPD at a dose of
trifuge (Shanghai, China) and centrifuged at 16,000 rpm for 35 min. A 25 mg·kg−1.
0.5 ml sample of the supernatant was diluted with methanol to 10 mL, After oral administration of CPhGs, hydrated proliposomes, and
and the absorbance A2 was measured, and the concentration of C2 was CPhGsPD, 2 mL blood samples were collected from the marginal ear
calculated. The encapsulation efficiency (EE %) was calculated using vein at 5, 15, 30, 45, 60, 90, 120, 150, 180, 210, 240 and 360 min from
the following formula: the three groups. Plasma was separated by centrifugation at 4000 rpm
for 15 min and stored at −20 °C until analysis.
EE% ¼ ðC1−C2 Þ=C1  100%
2.8.3. Plasma sample preparation
Chlorogenic acid was used as the internal standard. When the plas-
Where C1 is the amount of drug in the liposome suspension and C2 is ma had thawed, 100 μL plasma, 20 μL chlorogenic acid (CA,
the amount of drug in the eluate; their units are μg·mL−1. 0.2 mg·mL− 1) and 20 μL menthol were added to 1.5 mL centrifuge
tubes, and then vortex-mixed for 60 s. After adding 80 μL 10% trichloro-
2.6. Storage stability of CPhGsP and CPhGs ordinary liposomes acetic acid, each mixture was shaken for 120 s and then centrifuged
(12,000 rpm, 10 min) at 4 °C. The upper layer (160 μL) was transferred
Liposomal suspensions are thermodynamically unstable and so they to a clean centrifuge tube and dried under nitrogen at 50 °C. The residue
can flocculate and leak after a long storage period. Hence, temperature was reconstituted in methanol (100 μL) and aliquots (20 μL) of the su-
plays a key role in liposome stability (Laridi et al., 2003; Nguyen et al., pernatant were passed through a 0.22 μm membrane filter and then
2013; Park et al., 2013). Therefore, the stability of CPhGsP and CPhGs or- injected into the HPLC system for analysis.
dinary liposomes were studied at 4 ± 2 °C, 25 ± 2 °C, 40 ± 2 °C. The en-
capsulation efficiency was measured after 0, 7, 15, 30, 45, 60, 75 and 3. Results and discussion
90 days and determinations were carried out in triplicate.
3.1. Effect of surfactants on the encapsulation efficiency and particle size
2.7. In vitro CPhGsPD release study
The ratio of drug to propylene glycol was fixed a 1:80 for the test.
The in vitro release studies of CPhGsPD were carried out at 37 ± 0.5 ° The results of the effects of surfactants on the encapsulation efficiency
C using phosphate buffer solution (pH 7.4) and hydrochloric acid buffer and particle size are shown in Table 2. It was found that the hydrated li-
solution (pH 1.2). The total amount of drug was determined at 0, 5, 10, posome encapsulation efficiency was inversely proportional to the par-
20, 30, 45, 60, 90 and 120 min. Samples were withdrawn at specific ticle size compared with liposomes without surfactant. The
times, and all samples were passed through a 0.45 μm membrane filter. encapsulation efficiency of samples with a single component of
Also, the same amount of corresponding fresh dissolution medium was poloxamer-188 was higher than that of samples with added sodium
added to the stripping cup. The CPhGs content was determined from the deoxycholate, but in terms of particle size, the former was higher than
standard curve of echinacoside. At the same time, the particle size, zeta the latter. For the samples with added poloxamer-188 and sodium
potential and entrapment efficiency were measured after hydration of deoxycholate, the encapsulation efficiency and the particle size were
the dropping pills. All the experiments were carried out in triplicate. somewhere in the middle. It has been reported that the smaller the par-
ticle size, the faster the particles will infiltrate plasma (Stuart and Allen,
2.8. Pharmacokinetic studies in rabbits 2000). Therefore, the method of mixing surfactants was selected during
the experiments. In addition, with an increase in surfactant contents,
2.8.1. High performance liquid chromatography (HPLC) the particle size of the liposomes was reduced and the encapsulation
The content of echinacoside was measured using an HPLC system
(SHIMADZU, Japan) equipped with a binary pump, and an ultraviolet
Table 2
(UV) detector at 330 nm. A C18 analytical column (150 mm × 4.6 mm Surfactants of the characteristics of CPhGsP in terms of their particle size, and encapsula-
I.D. 5 μm) was used and the column was maintained at 30 °C. The mobile tion efficiency CPhGsP (n = 3).
phase consisted of acetonitrile-methanol-0.1% formic acid (10:15:75, v/
Encapsulation Particle
v) and was used throughout the analysis at a flow rate of 1.0 mL·min−1
Formulation Proportion(w/w) efficiency (%) size(nm)
and the sample injection volume was 20 μL. The solutions were all
D + SP 1:20:1 35.43 ± 0.67 900.7 ± 15
passed through a 0.22 μm membrane before injection. Echinacoside
D + SP + Tween 80 1:20:1 37.54 ± 0.86 881.3 ± 23
was the reference. A standard curve was constructed by plotting the D + SP + Span 20 1:20:1 43.57 ± 0.95 806.3 ± 20
concentration of echinacoside versus the absorbance: Y = D + SP + SD 1:20:1 36.97 ± 1.04 643.5 ± 26
0.88364X + 0.00173 (r = 0.998). Where Y is the echinacoside and CA D + SP + P-188 1:20:1 48.73 ± 0.79 739.0 ± 22
concentration, and A is the echinacoside and CA peak area. The standard D + SP + SD + P-188 1:20:0.5:0.5 46.81 ± 1.13 693.1 ± 19

curve ranged from 22.4–4480 ng·mL−1 and was linear. All recoveries D, drug; SD, sodium deoxycholate; SP, soybean phospholipid; P-188, Poloxamer-188.
228 M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232

Table 3 Table 5
Results of response surface experiments. The results of fitting second-order equations.

Levels of independent factors Response EE (%) Std. Dev. 1.03 R-Squared 0.9938
Run Mean 44.97 Adj R-Squared 0.9859
A B C
C.V. % 2.28 Pred R-Squared 0.9596
1 15 4 1.5 34.54 PRESS 48.21 Adeq Precision 32.914
2 20 4 1 51.12
3 20 3 1.5 47.34
4 15 4 0.5 33.33
5 20 4 1 52.73 negative correlation in the above regression equation. The analysis of
6 20 5 0.5 44.58 the fitting result is presented in Table 5. The adjusted R2 and predicted
7 20 4 1 53.63
8 25 3 1 48.33
R2 for the predictive model was 0.9859 and 0.9596, respectively, and
9 15 5 1 26.58 the statistical test results of equation parameters, showed that the
10 20 4 1 52.45
11 20 3 0.5 48.33
12 20 5 1.5 42.57
13 25 4 1.5 50.29
14 20 4 1 53.90
15 15 3 1 30.54
16 25 5 1 43.73
17 25 4 0.5 50.48

efficiency can be reduced if it is excessive. Hence, in response to the sur-


face experiments, the content of poloxamer-188 and sodium
deoxycholate were investigated as factors.

3.2. Analysis of response surface

There were seventeen experimental runs for optimizing three indi-


vidual parameters in the Box-Behnken design (BBD) using Design-Ex-
pert 8.0.6 software, and the experimental conditions and relevant
results are shown in Table 3. The results obtained showed that the en-
capsulation efficiency ranged from 26.58 to 53.90%.
The statistical significance of the regression model was checked by
the F-test and p-value, and the analysis of variance for the experimental
results is presented in Table 4. The small p-value for the model
(b0.0001) shows that the model is significant, and a p-value b 0.05
shows that the model term is significant. The p-value for the ‘lack of
fit’ test was 0.5996, indicating the model term was not significant and
the quadratic model was adequate.
Using statistical processing and fitting, a multiple second-order
equation was obtained as follows:
Final equation in terms of coded factors:
EE = + 52.77 + 8.48A − 2.14B − 0.25C − 0.16AB − 0.35AC −
0.26BC − 9.51A2 − 5.96B2 − 1.10C2The above regression equation de-
scribes the effects of three independent variables on the index and
their correlation. The positive and negative coefficients relate to the var-
iable and response representative of the correlation, in which a positive
coefficient represents a positive correlation, otherwise there was a

Table 4
Statistical analysis of variance for the encapsulation efficiency of experiment results.

Sum of Mean F
Source dƒ p value Prob N F
squares squares value

Model 1186.35 9 131.82 124.97 b0.0001 Significant


A 575.28 1 575.28 545.41 b0.0001
B 36.47 1 36.47 34.57 0.0006
C 0.49 1 0.49 0.46 0.5174
AB 0.1 1 0.1 0.097 0.7644
AC 0.49 1 0.49 0.46 0.5174
BC 0.26 1 0.26 0.25 0.6347
Fig. 3. Response surface of soybean phosphatide to the CPhGs ratio, the poloxamer-188 to
A2 380.64 1 380.64 360.88 b0.0001
CPhGs ratio, and the sodium deoxycholate to CPhGs ratio on the encapsulation efficiency
B2 149.72 1 149.72 141.94 b0.0001
A: the 3-D response surface plot at different soybean phosphatide to CPhGs ratios,
C2 5.08 1 5.08 4.81 0.0643
poloxamer-188 to CPhGs ratios at fixed sodium deoxycholate to CPhGs ratios; B: the 3-D
Residual 7.38 7 1.05
response surface plot at different soybean phosphatide to CPhGs ratios and sodium
Lack of fit 2.54 3 0.85 0.7 0.5996 Not significant
deoxycholate to CPhGs ratios at a fixed poloxamer-188 to CPhGs ratio; C: the 3-D
Pure error 4.84 4 1.21
response surface plot at different poloxamer-188 to CPhGs ratios, and sodium
Cor total 1193.73 16
deoxycholate to CPhGs ratios at a fixed soybean phosphatide to CPhGs ratio.
M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232 229

experimental results adequately fit the equation and the selected variable interaction graphs are shown in Fig. 3 and Fig. 4, respectively.
model. So, it can be used to predict a response and obtain a random for- With an increase in the ratio of soybean phospholipids to drug the en-
mula within the range of the designed factor and level by regression capsulation efficiency increases then finally changes little as shown in
equations. Fig. 3A, and the factor was the most significant (p b 0.0001). With an in-
In order to better comprehend the predictive models of the results crease in the ratio of poloxamer-188 to drug, the encapsulation efficien-
three-dimensional and variable interaction graphs of the models, the re- cy initially increases then falls as shown in Fig. 3B, and the factor was
sponse surface diagram of the encapsulation efficiency of CPhGs and the significant (p b 0.0006). In addition, as shown in Fig. 3C, with an increase
in the ratio of sodium deoxycholate to drug, the encapsulation efficiency
does not change significantly, and the factor was not significant
(p b 0.5174). As shown in Fig. 4, the interaction of the ratio of soybean
phospholipids to drug (A) and the ratio of poloxamer-188 to drug (B)
is the weakest, while the ratio of soybean phospholipids to drug (A)
and the ratio of sodium deoxycholate to drug (C) is the strongest.
The feasibility of the model equation for predicting the optimum re-
sponse values was tested under optimum conditions. The set of opti-
mum conditions, determined by using the RSM optimization
approach, was tested experimentally and parallel tests were carried
out in triplicate according to the optimized formulation. The mean

Fig. 4. Factor interaction graph showing the effect of the ratio of soybean phospholipids to
CPhGs, the ratio of poloxamer-188 to CPhGs, and the ratio of sodium deoxycholate to
CPhGs A: Interaction between the ratio of soybean phospholipids to CPhGs and the ratio
of poloxamer-188 to CPhGs; B: Interaction between the ratio of soybean phospholipids Fig. 5. Intensity weighting particle size distribution a.CPhGsP hydrated liposomes
to CPhGs and the ratio of sodium deoxycholate to CPhGs; C: Interaction between the b.CPhGsPD hydrated liposomes with phosphate buffer solution (pH 7.4) c.CPhGsPD
ratio of poloxamer-188 to CPhGs and the ratio of sodium deoxycholate to CPhGs. hydrated liposomes with hydrochloric acid buffer solution (pH 1.2).
230 M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232

Fig. 8. Release curve of CPhGsPD in phosphate buffer (pH 7.4) and hydrochloric acid buffer
(pH 1.2).

Fig. 6. Transmission electron micrographs of CPhGsP hydrated liposomes with phosphate


buffer solution (pH 7.4).
3.3.2. Transmission electron microscopy images of proliposomes
TEM images of the proliposomes after hydration are shown in Fig. 6,
which shows that the liposome vesicles were adjacent to the spherical
particles with distinct boundaries, which confirms the formation of lipo-
experimental encapsulation efficiency was 51.97% ± 1.07%, which was
somes from proliposomes. The liposome vesicles as shown under TEM
close to the predicted value of 54.30% (the ratio of soybean phospho-
have multilamellar vesicle walls, which confirms the formation of
lipids to drug of 22.38, the ratio of poloxamer-188 to drug of 3.52, and
multilamellar vesicles (MLVs) from proliposomes.
the ratio of sodium deoxycholate to drug of 0.84).

3.4. Storage stability results of CPhGsP and CPhGs ordinary liposomes


3.3. Characterization of CPhGsP
The stability curves of CPhGs ordinary liposomes and CPhGsP at dif-
3.3.1. Measurement of particle size and zeta potential ferent temperatures are shown in Fig. 7. Compared with CPhGs ordinary
The particle size distribution of hydrated CPhGsP is presented in Fig. liposomes, the CPhGsP encapsulation efficiency was much higher. The
5a. The average particle size and the average zeta potential were encapsulation efficiency of proliposomes did not change markedly,
671.7 nm and − 25.49 mV, respectively. The zeta potential showed while the encapsulation efficiency of CPhGs ordinary liposomes and
that the liposomes obtained have sufficient charge to inhibit the aggre- CPhGsP at 4 °C and 25 °C, especially at 4 °C, were basically unchanged.
gation of vesicles (Xiong et al., 2009). The zeta potential of CPhGsL was However, with an increase in time, the encapsulation efficiency of
negative, because the SP gave some ingredients a negative charge, such CPhGs ordinary liposomes was significantly reduced, and the encapsu-
as phosphatidylinositol and phosphatidylserine. lation efficiency of CPhGsP groups was slightly reduced. These results
show that CPhGs ordinary liposomes and CPhGsP should be stored at
low temperatures, and the stability of proliposomes is better than that
of ordinary liposomes.

Fig. 9. Mean plasma concentration-time profile of echinacoside following oral


Fig. 7. Stability curve of CPhGs ordinary liposomes and CPhGsP at different temperatures. administration of CPhGs, hydrated proliposomes, and CPhGsPD in rabbits (n = 6).
M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232 231

Table 6
The main pharmacokinetic parameters of CPhGs, hydrated proliposomes, and CPhGsPD in rabbits (n = 6).

Parameter CPhGs Hydrated proliposomes CPhGsPD

Tmax (min) 15.00 ± 0.00 30 ± 0.00 30 ± 0.00


Cmax (ng·mL−1) 193.17 ± 9.30 324.23 ± 11.38 315.17 ± 12.7
AUC0 − ∞ (ng·mL−1·min) 30,275.39 ± 4765.95 41,687.38 ± 8081.50 38,331.51 ± 4881.45
AUC0 − t (ng·mL−1·min) 15,494.70 ± 1183.66 26,322.50 ± 2133.74 24,552.14 ± 1400.94
MRT0 − ∞ (min) 301.289 ± 38.833 130.514 ± 40.062 149.553 ± 30.774
MRT0 − t (min) 80.790 ± 3.400 74.466 ± 2.579 72.039 ± 3.338
CL (L·h−1·kg−1) 0.001 ± 0.000 0.001 ± 0.000 0.001 ± 0.000
Vd (L·kg−1) 0.284 ± 0.028 0.159 ± 0.041 0.126 ± 0.013

Data are expressed as mean ± SD; Tmax time of peak concentration, Cmax peak concentration, AUC area under the curve, CL clearance rate, Vd volume of distribution, MRT mean residence
time.

3.5. Analysis of CPhGsPD in vitro release long residence time. Moreover, with CPhGsPD, the pharmacokinetic pa-
rameters were similar to those of hydrated proliposomes, showing that
The release curves of CPhGsPD in phosphate buffer (pH 7.4) and hy- the CPhGsPD absorption properties of liposomes confirmed they had
drochloric acid buffer (pH 1.2) are shown in Fig. 8. In phosphate buffer dissolved in vivo.
(pH 7.4) and hydrochloric acid buffer (pH 1.2), dripping pills can be al-
most completely released in 30 min. However, the release is faster and
higher in phosphate buffer (pH 7.4). These results indicated that 4. Conclusions
CPhGsPD is efficiently released in gastrointestinal solution in theory.
In addition, the CPhGsPD hydrated particle size is shown in Fig. 5b and In this study, a form of liquid CPhGsP for oral administration was
Fig. 5c hydrated by phosphate buffer (pH 7.4) and hydrochloric acid prepared using a more efficient and simpler method. The preparation
buffer (pH 1.2). The particle size was 671.7 nm, 645.7 nm and of proliposomes was found to be well suited to obtain an optimal
532.8 nm for hydrated CPhGsP, and CPhGsPD in phosphate buffer proliposomal formulation of CPhGs by RSM. The optimal formulation
(pH 7.4) and hydrochloric acid buffer (pH 1.2) in Fig. 5, respectively. obtained was drug/soybean phospholipid/poloxamer-188/sodium
The particle size showed less difference between hydrated CPhGsP deoxycholate/propylene glycol.
and CPhGsPD in phosphate buffer (pH 7.4). The size in hydrochloric 1:22.38:3.52:0.84:80 (w/w/w/w/v). The proliposome encapsulation
acid buffer was a little smaller than that in hydrochloric acid buffer. efficiency reached 51.97%. In addition, the stability study indicates that
The reason for this might be that some liposomes were disrupted in the proliposome stability is better than that of ordinary liposomes at
the acidic environment (Brocks and Betageri, 2002). In addition, the the same temperature, and a lower temperature of 4 °C is preferred
EE% and zeta potential, were 49.56 ± 0.75% and − 20.71 ± 2.1 mV in for storage. The CPhGsPD release is good in gastrointestinal solution in
phosphate buffer (pH 7.4), and 47.89 ± 0.84% and − 21.71 ± 1.7 mV theory as shown by in vitro release experiments and the structure of li-
in hydrochloric acid buffer (pH 1.2), respectively, close the hydrated posomes did not destroy the prolipospme dropping pills by hydration.
proliposomes. These results showed that the structure of the liposomes This liquid proliposome dripping pills formulation improve the oral bio-
did not destroy the proliposome dropping pills following hydration. For availability of CPhGs in New Zealand rabbits and offers a new approach
research involving water-soluble drug proliposomes and CPhGsPD the to preparing liposomes for oral administration.
oral route of administration provides a theoretical basis.
Acknowledgements

3.6. Pharmacokinetic studies in rabbits The authors would like to thank Dr. J. Han and Dr. N. Liu for valuable
discussions and interest in the work. Thanks also to Shenyang Science
The mean plasma concentration-time profiles of CPhGs, hydrated and Technology Bureau (F16-205-1-45) for supporting this work.
proliposomes and CPhGsPD in rabbits are shown in Fig. 9. The plasma
echinacoside concentrations were higher in rabbits given hydrated
proliposomes and CPhGsPD than in those given CPhGs, at all time References
points. The peak concentration (Cmax) and time to reach the peak con-
Ariën, A., Goigoux, C., Baquey, C., et al., 1993. Study of in vitro and in vivo stability of lipo-
centration (Tmax) were obtained directly from the individual plasma somes loaded with calcitonin or indium in the gastrointestinal tract. Life Sci. 53,
echinacoside vs. time profiles. The Cmax values of the CPhGsPD group 1279–1290.
and the hydrated proliposomes group were higher than that obtained Bangham, A.D., Standish, M.M., Watkins, J.C., 1965. Diffusion of univalent ions across the
lamellae of swollen phospholipids. J. Mol. Biol. 13, 238–252.
with CPhGs. Using DAS 2.0 pharmacokinetic software, according to the Brocks, D.R., Betageri, G.V., 2002. Enhanced oral absorption of halofantrine enantiomers
second chamber model, the statistical moment calculations of the after encapsulation in a proliposomal formulation. J. Pharm. Pharmacol. 54,
main pharmacokinetic parameters are summarized in Table 6. CPhGsPD 1049–1053.
Cai, R.L., Yang, M.H., Shi, Y., et al., 2010. Antifatigue activity of phenylethanoid-rich extract
and hydrated proliposomes gave a Tmax of 30 min, compared with from Cistanche deserticola. Phytother. Res. 24, 313–315.
15 min for CPhGs, the delayed Tmax with CPhGsPD and hydrated Chen, H., Jing, F.C., Li, C.L., et al., 2007. Echinacoside prevents the striatal extracellular
proliposomes, were possibly due to the slow release of drug from the levels of monoamine neurotransmitters from diminution in 6-hydroxydopamine le-
sion rats. J. Ethnopharmacol. 114, 285–289.
hydrated liposomes. Chiang, W.-D., Chang, S.-W., Shieh, C.-J., 2003. Studies on the optimized lipase-catalyzed
When formulations containing liposomal CPhGs were administered, biosynthesis of cis-3-hexen-1-yl acetate in n-hexane. Process Biochem. 38,
some pharmacokinetic parameters were completely different from 1193–1199.
Cui, Q., Pan, Y., Xu, X., et al., 2016. The metabolic profile of acteoside produced by human
CPhGs. Compared with CPhGs, the area under the plasma level-time
or rat intestinal bacteria or intestinal enzyme in vitro employed UPLC-Q-TOF–MS.
curve (AUC0 − ∞) for the CPhGsPD group and hydrated proliposomes Fitoterapia 109, 67–74.
group was higher, confirming slower CPhGs distribution and elimina- Dayan, N., Touitou, E., 2000. Carriers for skin delivery of trihexyphenidyl HCl: ethosomes
tion from the plasma of CPhGs encapsulated liposomes. The CPhGs vs. liposomes. Biomaterials 21, 1879–1885.
Gabizon, A., Dagan, A., Goren, D., et al., 1982. Liposomes as in vivo carriers of adriamycin:
group mean residence time (MRT0 − ∞) was higher than that of the reduced cardiac uptake and preserved antitumor activity in mice. Cancer Res. 42,
other two, possibly because of the CPhGs low absorption in vivo and a 4734–4739.
232 M. Li et al. / European Journal of Pharmaceutical Sciences 93 (2016) 224–232

Geng, X., Song, L., Pu, X., et al., 2004. Neuroprotective effects of phenylethanoid glycosides Maruyama, S., Hashizume, S., Tanji, T., et al., 2008. Cistanche salsa extract enhanced anti-
from Cistanches salsa against 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine body production in human lymph node lymphocytes. Pharmacologyonline 2,
(MPTP)-induced dopaminergic toxicity in C57 mice. Biol. Pharm. Bull. 27, 797–801. 341–348.
Gregoriadis, G., 1976. The carrier potential of liposomes in biology and medicine. N. Engl. Mi, Z., Burke, T.G., 1994. Marked interspecies variations concerning the interactions of
J. Med. 295, 704–710. camptothecin with serum albumins: a frequency-domain fluorescence spectroscopic
Grit, M., de Smidt, J.H., A., S., et al., 1989. Hydrolysis of phosphatidylcholine in aqueous li- study. Biochemistry 33, 12540–12545.
posome dispersions. Int. J. Pharm. 50, 1–6. Morikawa, T., Pan, Y., Ninomiya, K., et al., 2010. Acylated phenylethanoid oligoglycosides
Hamsaveni, D.R., Prapulla, S.G., Divakar, S., 2001. Response surface methodological ap- with hepatoprotective activity from the desert plant Cistanche tubulosa. Bioorg. Med.
proach for the synthesis of isobutyl isobutyrate. Process Biochem. 36, 1103–1109. Chem. 18, 1882–1890.
Hunt, C.A., Tsang, S., 1981. α-Tocopherol retards autoxidation and prolongs the shelf-life Nan, Z., Zeng, K., Shi, S., et al., 2013. Phenylethanoid glycosides with anti-inflammatory ac-
of liposomes. Int. J. Pharm. 8, 101–110. tivities from the stems of Cistanche deserticola cultured in Tarim desert. Fitoterapia
Jia, C., Shi, H., Wu, X., et al., 2006. Determination of echinacoside in rat serum by reversed- 89, 167–174.
phase high-performance liquid chromatography with ultraviolet detection and its ap- Nguyen, T.T.T.N., Østergaard, J., Stürup, S., et al., 2013. Determination of platinum drug re-
plication to pharmacokinetics and bioavailability. J. Chromatogr. B 844, 308–313. lease and liposome stability in human plasma by CE-ICP-MS. Int. J. Pharm. 449,
Jia, C., Shi, H., Jin, W., et al., 2009. Metabolism of echinacoside, a good antioxidant, in rats: 95–102.
isolation and identification of its biliary metabolites. Drug Metab. Dispos. 37 (2), Park, S.M., Kim, M.S., Park, S.-J., et al., 2013. Novel temperature-triggered liposome with
431–438. high stability: formulation, in vitro evaluation, and in vivo study combined with
Jiang, Y., Tu, P.F., 2009. Analysis of chemical constituents in Cistanche species. high-intensity focused ultrasound (HIFU). J. Control. Release 170, 373–379.
J. Chromatogr. A 1216 (11), 1970–1979. Payne, N.I., Timmins, P., Ambrose, C.V., et al., 1986. Proliposomes: a novel solution to an
Junping, W., Maitani, Y., Takayama, K., et al., 2000. In vivo evaluation of doxorubicin car- old problem. J. Pharm. Sci. US 75, 325–329.
ried with long circulating and remote loading proliposome. Int. J. Pharm. 203, 61–69. Stuart, D.D., Allen, T.M., 2000. A new liposomal formulation for antisense
Kumar, R., Gupta, R.B., Betageri, G.V., 2001. Formulation, characterization, and in vitro re- oligodeoxynucleotides with small size, high incorporation efficiency and good stabil-
lease of glyburide from proliposomal beads. Drug Deliv. 8, 25–27. ity. Biochim. Biophys. Acta Biomembr. 1463, 219–229.
Laridi, R., Kheadr, E.E., Benech, R.O., et al., 2003. Liposome encapsulated nisin Z: optimiza- Vicente, G., Coteron, A., Martinez, M., et al., 1998. Application of the factorial design of ex-
tion, stability and release during milk fermentation. Int. Dairy J. 13, 325–336. periments and response surface methodology to optimize biodiesel production. Ind.
Lee, W.C., Yusof, S., Hamid, N.S.A., et al., 2006. Optimizing conditions for hot water extrac- Crop. Prod. 8, 29–35.
tion of banana juice using response surface methodology (RSM). J. Food Eng. 75, Xiong, Y., Guo, D., Wang, L., et al., 2009. Development of nobiliside A loaded liposomal for-
473–479. mulation using response surface methodology. Int. J. Pharm. 371, 197–203.
Li, F., Yang, X., Yang, Y., et al., 2015. Phospholipid complex as an approach for bioavailabil- Yadav, A., Murthy, M.S., Shete, A.S., et al., 2011. Stability aspects of liposomes. Ind.
ity enhancement of echinacoside. Drug Dev. Ind. Pharm. 41, 1777–1784. J. Pharm. Educ. 45, 402–413.
Lin, L.W., Hsieh, M.T., Tsai, F.H., et al., 2002. Anti-nociceptive and anti-inflammatory activ- Zhang, Z.-S., Li, D., Wang, L.-J., et al., 2007. Optimization of ethanol–water extraction of
ity caused by Cistanche deserticola in rodents. J. Ethnopharmacol. 83, 177–182. lignans from flaxseed. Sep. Purif. Technol. 57, 17–24.
Liyana-Pathirana, C., Shahidi, F., 2005. Optimization of extraction of phenolic compounds Zhang, D., Li, H., Wang, J.B., 2015. Echinacoside inhibits amyloid fibrillization of HEWL and
from wheat using response surface methodology. Food Chem. 93, 47–56. protects against Aβ-induced neurotoxicity. Int. J. Biol. Macromol. 72, 243–253.

You might also like