Accepted Manuscript Title: PREPARATION AND CHARACTERIZATION OF QUERCETIN-LOADED LIPID LIQUID CRYSTALLINE SYSTEMS Author: A. Linkeviˇci¯ut˙e A. Misi¯unas E. Naujalis J. Barauskas PII: DOI: Reference:

S0927-7765(15)00076-4 http://dx.doi.org/doi:10.1016/j.colsurfb.2015.02.001 COLSUB 6891

To appear in:

Colloids and Surfaces B: Biointerfaces

Received date: Revised date: Accepted date:

15-10-2014 22-1-2015 1-2-2015

Please cite this article as: A. Linkeviˇci¯ut˙e, A. Misi¯unas, E. Naujalis, J. Barauskas, PREPARATION AND CHARACTERIZATION OF QUERCETIN-LOADED LIPID LIQUID CRYSTALLINE SYSTEMS, Colloids and Surfaces B: Biointerfaces (2015), http://dx.doi.org/10.1016/j.colsurfb.2015.02.001 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1

PREPARATION AND CHARACTERIZATION OF QUERCETIN-LOADED

2

LIPID LIQUID CRYSTALLINE SYSTEMS

3

A. Linkevičiūtė1,2,*, A. Misiūnas2, E. Naujalis1,2, J. Barauskas3

5

1

6

LT-03225 Vilnius, Lithuania

7

2

9

cr

Department of Analytical and Environmental Chemistry, Vilnius University, Naugarduko 24,

State Research Institute, Center for Physical Sciences and Technology, A. Goštauto 11,

LT-01108 Vilnius, Lithuania 3

us

8

ip t

4

Biomedical Science, Faculty of Health and Society, Malmö University, SE-20506 Malmö,

Sweden

11

E-mail addresses: [email protected], [email protected], [email protected],

12

[email protected]

an

10

M

13

*Corresponding author: Ausra Linkeviciute, State Research Institute, Center for Physical

15

Sciences and Technology, A. Gostauto 11, LT-01108 Vilnius, Lithuania. Tel.: +370 5 2612758;

16

Fax. +370 5 2627123. E-mail addresses: [email protected]

17

te

d

14

Abstract

19

The aim of the present study was to investigate mixtures of soy phosphatidylcholine (SPC) and

20

glycerol dioleate (GDO) as encapsulation matrices for antioxidant quercetin. The effects of

21

quercetin loading into non-aqueous formulations, non-lamellar liquid crystalline phases and their

22

colloidal dispersions were studied by using synchrotron small angle X-ray diffraction, dynamic

23

light scattering, cryogenic electron microscopy and high performance liquid chromatography.

24

Quercetin incorporation is discussed in the context of lipid aggregation behavior, self-assembled

25

nanostructure and chemical stability. The obtained results show that SPC/GDO-based

26

formulations can incorporate relatively high amounts of quercetin and serve as liquid crystalline

27

delivery vehicles in the form of bulk phases or colloidal dispersions.

Ac ce p

18

28 29

Keywords:

Page 1 of 27

30

Quercetin, lipid liquid crystals, non-lamellar phases, lipid liquid crystalline nanoparticles, small

31

angle X-ray diffraction, cryo-TEM, HPLC.

Ac ce p

te

d

M

an

us

cr

ip t

32

Page 2 of 27

32

1.

Introduction

34

Flavonoids belong to the large and diverse group of polyphenolic compounds characterized by a

35

common benzo-γ-pyrone structure and are ubiquitously present in plants [1, 2]. During recent

36

years, these compounds are of growing interest due to broad pharmacological features, including

37

anti-thrombotic, anti-inflammatory, anticancer and immunostimulator activities [3-5]. Moreover,

38

in the human diet quercetin (QU) (3,3´,4´,5,7-pentahydroxyflavone) is one of the most abundant

39

natural flavonoid, which antioxidant activity is higher than for the other well-known antioxidant

40

molecules, i.e. ascorbylpalmitate, trolox and rutin [6, 7]. Besides antioxidant properties,

41

quercetin also possesses anticancer, antiviral and antiallergic features [8, 9]. The anticancer

42

properties of this natural antioxidant have been proved by in vivo and in vitro experiments which

43

demonstrated that QU has a significant role in inhibition of breast, colon, prostate, ovary,

44

endometrium, and lung tumor cancer cells [10-12].

45

The main limitation in the therapeutic employment of QU is its low water solubility and

46

instability in physiological medium which restricts its use to oral route of administration [13]. In

47

order to extend clinical applications and to solve the solubility, instability, and bioavailability

48

issues of QU it is necessary to develop an appropriate flavonoid delivery system by

49

entrapping/adsorbing QU into the more favorable environment. A variety of different QU

50

immobilization/incorporation strategies, such as using emulsions [14-16], nanosuspensions [17],

51

microcapsules [18, 19], solid lipid nanoparticles [20-24], have been suggested to overcome at

52

least some of these limitations. Much attention has been also made to a various types

53

phospholipid-based vesicles and liposomes as potential vehicles for entrapment and delivery of

54

QU [3, 25-30]. An advantage to use phospholipids is their amphiphilic nature that can modify the

55

solubility behavior and rate of drug release and enhance drug absorption across the biological

56

barriers [3, 31, 32]. However, complex preparation routines and low encapsulation capability of

57

poor water soluble molecules restricts the potential application of vesicles and liposomes for the

58

entrapment and delivery of QU.

59

It was known for years that many phospholipids, such as phosphatidylcholines, spontaneously

60

form flat lipid bilayers and fragmented vesicular particles in aqueous dispersions. It was early on

61

also observed that a variety of polar lipids and their mixtures can self-assemble into different

Ac ce p

te

d

M

an

us

cr

ip t

33

Page 3 of 27

non-lamellar liquid crystalline (LC) phases in water under solution conditions similar to that of

63

biological systems [33]. Non-lamellar LC phases, such as sponge, hexagonal, and cubic,

64

typically comprise of hydrophilic and hydrophobic, which may be continuous or discrete, thus

65

forming either mono- or bicontinuous networks, depending on the molecular nature of the lipid

66

or lipid mixture [34]. They have generally much higher surface area per volume than lamellar

67

structures, may better solubilize hydrophobic, hydrophilic and also amphiphilic molecules [35-

68

37], and can be used as delivery systems for peptides, proteins and food bioactives [38, 39].

69

Recently, a versatile drug delivery lipid system based on soy phosphatidylcholine (SPC) and

70

glycerol dioleate (GDO) has been developed. It offers optimal functional properties, such as

71

bioadhesion and controlled release, good drug loading ability, enhanced flexibility of self-

72

assembly that can be tuned in the range from lamellar to various reversed non-lamellar LC

73

phases [40, 41]. Furthermore, in the presence of polymeric particle stabilizer Polysorbate 80

74

(P80) SPC/GDO mixtures can be easily dispersed into non-lamellar liquid crystalline

75

nanoparticles (LCNPs) with controllable small size and inner morphology making such

76

nanoparticles suitable for parenteral administration routes [42-44].

77

Here we investigated the potential of SPC/GDO-based non-aqueous lipid formulations, non-

78

lamellar LC phases and LCNPs as delivery vehicles of QU. To our best knowledge, only one

79

study of QU entrapment into non-lamellar lipid-based LC system has been previously reported.

80

Murgia and co-workers [45] have been studied the incorporation of QU into glycerol monooleate

81

(GMO)-based bicontinuous cubic LCNPs and demonstrated that small amounts of QU does not

82

alter size, charge and inner nanostructure of the nanoparticles. In this study several mono-, di-

83

glyceride and SPC/GDO-based non-lamellar LC phase forming lipid compositions were explored

84

in the context of their ability to solubilize QU. The main objective of the study was to elucidate

85

the effects of QU solubilization on the nanostructure of lipid-based non-lamellar LC phases and

86

their LCNPs as a function of lipid composition and QU concentration by means of polarized

87

light microscopy, synchrotron small-angle X-ray diffraction (SAXD), dynamic light scattering

88

(DLS) and cryogenic transmission electron microscopy (cryo-TEM). Finally, chemical stability

89

of entrapped QU in different LC systems has been evaluated by using HPLC.

Ac ce p

te

d

M

an

us

cr

ip t

62

90

Page 4 of 27

90 91

2.

Materials and methods

2.1.

Materials

93

ip t

92

94

Quercetin (QU) (CAS-No.117-39-5; purity ≥ 98.0%) was purchased from Sigma-Aldrich.

96

Diglycerolmonooleate (DGMO) (Rylo PG 29) and glycerol dioleate (GDO) (Rylo PG 19) from

97

Danisco, commercial mixture of mono-, di- and triglycerides denoted as Capmul GMO-50 from

98

Abitec, polysorbate P80 (P80) from Croda, glycerylmonooleyl ether (GME) from Niko

99

Chemicals, and soy phosphatidylcholine (SPC) denoted as SPC S100 from Lipoid were used as

100

received. Methanol (MeOH) and acetonitrile (ACN) was purchased from Sigma-Aldrich, 99.7 %

101

ethanol (EtOH) from either Merck or Solveco, and formic acid from Scharlau Chemie. All

102

chemicals were of analytical grade. The water used was passed through the NANOpure Infinity

103

(Branstead) water purification system.

104

2.2.

Preparation of lipid-based formulations

107

Non-aqueous lipid formulations were prepared by mixing appropriate amounts of lipid

108

components (DGMO, GDO, Campul GMO-50, GME and SPC) without and in the presence of

109

co-solvent ethanol to facilitate mixing. Lipid mixtures were then placed on a roller mixer at room

110

temperature for 24 h until mixed completely. Prepared lipid formulations were kept at room

111

temperature (RT) until further use. QU containing formulations were prepared by weighing

112

appropriate amounts of QU and non-aqueous lipid formulations and placing on a roller mixer for

113

24–48 h at RT until homogeneous mixture was obtained. Visual inspection between cross-

114

polarizers and light microscopy were used to inspect samples for the presence of undissolved

115

QU. In this manuscript, the concentration of QU is always expressed as wt% of the total

116

formulation weight. Component ratios of lipid formulations are also always expressed in wt%.

Ac ce p

te

106

d

105

M

an

us

cr

95

117 118

2.3.

Preparation of bulk lipid liquid crystalline (LC) phases

119

Page 5 of 27

120

LC phases were prepared by injection of QU containing non-aqueous lipid formulations (about

121

300 mg) into water at the formulation/water weight ratio of 5/95. Samples were immediately

122

sealed and left to equilibrate at RT in still standing vials for at least 10 days before experiments.

123

2.4.

Preparation of lipid liquid crystalline nanoparticle (LCNP) dispersions

ip t

124 125

Non-aqueous lipid formulations for LCNP preparation were prepared by mixing appropriate

127

amounts of lipids (SPC/GDO = 35/65, 60/40 wt%), polymeric particle stabilizer P80, and ethanol

128

at fixed (SPC+GDO)/P80/EtOH weight ratio of 75/15/10. QU containing formulations were

129

prepared by weighing appropriate amounts of QU and mixing with non-aqueous formulations at

130

QU concentration of 0.5, 1.0, 2.0 and 4.0 wt%. Lipid mixtures were placed on a roller mixer for

131

24 h until mixing was complete, and were then dispersed in 82.5 wt% water (17.5 wt% LCNPs).

132

The aqueous dispersions were immediately sealed, shaken, and left to vortex for 72 h on a

133

mechanical mixing table at 300 rpm at RT. Prepared dispersions were stored at RT until further

134

use.

M

an

us

cr

126

136

2.5.

Polarizing light microscopy (PLM)

te

137

d

135

The texture and temperature-induced phase transitions of LC phases were examined by a

139

polarizing microscope Optiphot equipped with a digital camera DS-2Mv (Nikon) and a heating

140

table Analysa LTS350 (Linkam). A small specimen of sample was placed between two

141

microscope slides and their edges were immediately sealed with a thermo stable silicon grease to

142

prevent evaporation of water from the specimen. A stepwise increase of temperature (typically

143

5 °C in every step with a heating rate of 1 °C/min) was used to induce phase transitions. QU

144

loaded lipid LCNPs were examined at RT by a polarizing microscope DM 750 (Leica) equipped

145

with a digital camera MC170 (Leica).

Ac ce p

138

146 147

2.6.

Small-angle X-ray diffraction (SAXD)

148 149 150

The nanostructure of LC phases was studied using synchrotron SAXD measurements, which were performed at the I911-4beamline at MAX-lab (Lund University, Sweden), using a 1M

Page 6 of 27

PILATUS 2D detector containing a total of 981 x 1043 pixels. Bulk lipid LC samples were

152

mounted between kapton windows in a steel sample holder at the sample to detector distance of

153

1917 mm. Diffractograms were recorded with a wavelength of 0.91 Å and the beam size of 0.25

154

× 0.25 mm (full width at the half-maximum) at the sample. Temperature control within 0.1 °C

155

was achieved using computer controlled Julabo heating circulator F12-MC (Julabo Labortechnik

156

GMBH, Seelbach, Germany). The experiments were performed successively at 25, 35, 45, 55,

157

and 65 °C with a 60 s exposure time at each temperature and a wait of 10 minutes between

158

temperature steps. The resulting CCD images were integrated and analyzed using the Fit2D

159

software provided by Dr. A. Hammersley [http://www.esrf.fr/computing/scientific/FIT2D].

160

Silver behenate calibrated sample-to-detector distance and detector positions were used.

us

cr

ip t

151

162

2.7.

an

161

Particle size and zeta potential

163

LCNP particle size distributions and zeta potentials were measured using Zetasizer Nano ZS

165

analyzer from Malvern Instruments. For particle size distributions the disposable cuvette filled

166

with 1 mL of LCNP dispersion, which was first diluted to 99.5 wt% of water. The obtained data

167

were averaged from 30 measurements (10 s each). The refractive indices used for lipid particles

168

and water were 1.48 and 1.33, respectively. The particle size distributions were reported as

169

intensity-averaged. The surface charge of the particles was measured using disposable zeta cells

170

filled with 1 mL of LCNP dispersion which was first diluted to 99.5 wt% of water. The zeta

171

potential was calculated using the Smoluchowski approximation for dispersion in water with

172

viscosity of 0.8872 cP, refractive index of 1.33, and dielectric constant of 78.5.

173 174 175

2.8.

Ac ce p

te

d

M

164

Cryogenic transmission electron microscopy (Cryo-TEM)

176

Lipid LCNP dispersions for electron microscopy were prepared in a controlled environment

177

vitrification system to ensure stable temperature and to avoid loss of water during sample

178

preparation. The climate chamber temperature was kept at 25−28 °C, and the relative humidity

179

was kept close to saturation to prevent sample evaporation. The samples were prepared by

180

placing 5 μL of LCNP dispersion on lacey carbon filmed copper grids and gently blotted with

181

filter paper to obtain a thin liquid film (20−400 nm) on the grid. Immediately after blotting, the

Page 7 of 27

grids were rapidly plunged into the liquid ethane at −180 °C to vitrify the water-rich samples to

183

prevent ice crystal formation and to preserve the internal crystalline structure. The vitrified

184

specimens were stored in liquid nitrogen (−196 °C) until measurements. An Oxford CT3500

185

cryo-holder and its work station were used to transfer the samples into the electron microscope

186

(Philips CM120 BioTWINCryo) equipped with a post-column energy filter (Gatan GIF100). The

187

acceleration voltage was 120 kV, and the working temperature was kept below −180 °C. The

188

images were recorded digitally with a CCD camera under low electron dose conditions.

cr

ip t

182

189

2.9.

QU stability study

us

190 191

The chemical stability of QU entrapped in different lipid non-aqueous formulations, LC phases

193

and LCNP dispersions were monitored for 3 months. During stability study all samples were

194

kept at RT in the darkness. For the evaluation of the residual QU concentration a portion of the

195

sample (about 10-15 mg) was collected at predetermined time points (1, 15, 30, 60 and 90 days),

196

dissolved in ACN:MeOH (1:1 v/v) solvent mixture at the lipid sample to solvent weight ratio of

197

1:100 and immediately analyzed using HPLC. Each sample was analyzed in triplicate.

199 200

2.10.

High performance liquid chromatography (HPLC)

te

198

d

M

an

192

QU concentration was determined using an Agilent HPLC 1100 Series (Agilent Technologies,

202

USA) chromatography system equipped with a quaternary pump, a vacuum degasser module, a

203

manual injector with a 20 µL sample loop, a temperature controlled column compartment and a

204

diode array detector (DAD) set at 370 nm. Chromatographic separation was achieved using a

205

Zorbax Eclipse XDB–C18 (5 µm, 150×4.60 mm, Agilent Technologies, USA) reversed phase

206

column coupled with an Eclipse XDB–C18 guard column (5 µm, 12.6×4.6 mm, Agilent

207

Technologies, USA). Determination was carried out in a solvent system of methanol–formic

208

acid–water as previously described, with minor modifications [46]. Data were collected and

209

processed using a ChemStation software version B.01.03. The obtained values with standard

210

methanolic QU solutions showed linearity over the concentration range of 0.1–100 μg/g with a

211

correlation coefficient (r2) of 0.999. The quantification limit in the HPLC assay was 0.1 μg/g and

212

standard deviation under repeatability conditions was no more than 5.6% in all concentrations

Ac ce p

201

Page 8 of 27

213

tested. The calibration curves for quantification of QU in QU-loaded non-aqueous lipid

214

formulations, LC phases and LCNP dispersions were prepared using QU standard with

215

respective lipid compositions. The obtained results of the quantification of various formulations

216

are summarized in Table S1 (Supporting information).

Ac ce p

te

d

M

an

us

cr

ip t

217

Page 9 of 27

217 218

3.

Results and discussions

3.1.

Solubility of QU in lipid formulations

220

ip t

219

221

It is known that some poorly water soluble compounds like QU may also have low solubility in

223

lipid excipients due to different physical chemical reasons [31]. As shown in other lipid- and

224

amphiphile-based systems, such as microemulsions [19], liposomes [26], lecithin-chitosan

225

nanoparticles [23] and solid lipid nanoparticles [24], the maximum solubility of QU is quite low

226

and often is in the order of 0.5 wt%, which can be somewhat improved by using additional

227

polymeric surfactants and/or co-solvents [16].

228

Therefore, before further experiments several non-lamellar LC phase forming non-aqueous lipid

229

mixtures were investigated with respect to their ability to solubilize QU. First, binary mixtures of

230

DGMO/GDO (85/15 and 60/40) (here and everywhere in the text formulation compositions are

231

expressed in wt%), DGMO/Capmul GMO-50 (85/15 and 60/40), DGMO/GME (85/15 and

232

60/40) and SPC/GDO (50/50) were explored. The results have shown that independently from

233

lipid ratio the solubility of QU in all tested mono- and di-glyceride-based mixtures was very low

234

and varies between 0.3 and 0.4 wt%. The addition of up to 10 wt% of ethanol (in respect to total

235

lipid) did not improve the solubility of QU in neither of the formulations. The observed solubility

236

in mixtures of mono- and di-glycerides was in line with previous results in similar systems. The

237

solubility of QU in commercial mixtures of long chain monoglycerides (i.e., Miglyol 812,

238

Capmul MCM, Labrafil 1944) is usually well below 1 wt% [16].

239

Much higher solubility of QU was found for the mixtures of SPC and GDO. Since solvent free

240

SPC/GDO mixtures prepared at equal lipid weight ratio were difficult to mix and were very

241

viscous their ability to solubilize QU were only tested in the presence of small amount of ethanol

242

at fixed Lipid/EtOH weight ratio of 90/10. SPC/GDO composition in the formulations was

243

varied between 60/40 and 35/65. Independently on lipid ratio about 5 wt% of QU was soluble in

244

all formulations. From the solubility test it may be concluded that lipid mixtures composed of

245

exclusively mono- and di-glycerides with and without EtOH have very limited ability to

246

solubilize QU which may be increased by about 10-20 times with the introduction of SPC into

Ac ce p

te

d

M

an

us

cr

222

Page 10 of 27

247

the mixture. Considering these results, SPC/GDO-based formulations containing up to 4 wt% of

248

QU and 10 wt% EtOH were selected for further studies.

249 250

3.2.

SPC/GDO-based LC phases with entrapped QU

ip t

251

The aim of this part was to investigate the effects of QU on the aggregation behavior of fully

253

hydrated non-lamellar bulk LC phases of SPC/GDO. As shown in recent study, the aqueous

254

phase behavior of mixtures of SPC and GDO is rather complex [40]. At 25 °C with increasing

255

GDO content fully hydrated SPC/GDO mixtures in water form the following LC phase

256

sequence: lamellar (Lα)  reversed 2D hexagonal (H2, up to 62.5/37.5)  reversed micellar

257

cubic of Fd m space group (50/50 – 45/55)  reversed 3D hexagonal of P63/mmc space group

258

(42/58 – 40/60)  unresolved “intermediate” (39/61 – 37/63)  Fd m (35/65 – 22.5/77.5) 

259

reversed micellar solution (L2, from 20/80). In this study, the effects of QU on the nanostructure

260

of different LC phases were investigated at four fixed SPC/GDO weight ratios (60/40, 50/50,

261

40/60 and 35/65) and four temperatures (25, 35, 45 and 55 °C).

262

Figure 1 shows SAXD data of fully hydrated SPC/GDO (60/40 and 35/65) LC phases with

263

entrapped 0.0, 0.5, 1 and 4 wt% of QU as a function of temperature.

264

As seen from Figure 1a, at weight ratio of 60/40 and at 25 °C SPC/GDO forms H2 LC phase,

265

which features three strong Bragg reflections positioned in ratios 1: 3: 4. At 45 °C another

266

coexisting LC phase, most likely Fd m cubic, also starts to form what is evidenced by the

267

appearance of additional peaks. However, independently on temperature the appearance of all

268

diffractograms is practically unaffected by the presence of up to 4 wt% of QU (Figures 1b-1d).

269

In addition, the calculated lattice parameter (a) for this LC phase is constant regardless QU

270

concentration and temperature and only slightly varies between 6.6 and 6.7 nm. This shows that

271

H2 phase prepared at high SPC content is quite robust and can accommodate relatively high

272

amounts of hydrophobic QU without changing nanostructure.

273

At weight ratio of 35/65 and 25 °C SPC/GDO self-assembles into ordered reversed micellar

274

Fd m cubic phase which is clearly characterized by the appearance of first 9 Bragg peaks

275

located at relative positions in ratios 3: 8: 11: 12: 16: 19: 24: 27: 32 (Figure 1e).

276

Contrary to H2 phase, the structure of Fd m cubic phase prepared at low SPC content is rather

277

sensitive to QU entrapment and temperature. At 45 °C it starts to “melt” and at 55 °C completely

Ac ce p

te

d

M

an

us

cr

252

Page 11 of 27

transforms into unordered reversed micelles (L2), which is corroborated by the disappearance of

279

the Bragg diffraction peaks and their smearing into two broad diffuse diffraction features. The

280

entrapment of QU even further disturbs ordered structure of the Fd m cubic phase. At 0.5 and 1

281

wt% of QU ordered cubic phase is almost fully transformed into unordered phase at 45 °C

282

(Figures 1f and 1g). Moreover, only a fraction of the Fd m cubic arrangement with entrapped 4

283

wt% of QU is present already at 35 °C (Figure 1h). As shown in Figure S1 (SI), both the

284

entrapment of QU and temperature considerably enlarge the unit cell dimensions of the cubic LC

285

phase at the SPC/GDO ratio of 35/65. With increasing QU concentration from 0 to 4 wt% the

286

calculated a value increases from 14.9 to 15.6 nm at 25 °C. At elevated temperatures a further

287

increases and reaches maximum value of about 15.9 nm at 1 wt% of QU and 45 °C before

288

complete transformation of ordered Fd m structure into unordered L2 phase.

289

The observed disordering effect of QU may be explained by reasonable assumption that in the

290

LC phase water insoluble hydrophobic QU molecule is preferentially located between lipid

291

hydrocarbon chains. Accommodation of QU in the hydrophobic regions between ordered

292

reversed micelles increases lipid chain packing stress and distance between micelles resulting in

293

slightly larger a values. At some point, further accommodation of QU and unit cell dimension

294

increase is not possible without transformation into unordered L2 phase. At higher temperatures

295

this process occurs at lower QU concentrations due to additional temperature induced lipid chain

296

packing disorder. Similar phase behavior trends were also observed for the SPC/GDO Fd m

297

cubic phases with entrapped benzydamine, lidocaine and granisetron [41].

298

In addition, less extensive but similar disordering effects of QU were also observed for the LC

299

phases prepared at intermediate SPC/GDO weight ratios of 50/50 and 40/60 (Figure S2, SI).

300

Here, the entrapment of QU has little effect on the unit cell dimensions of the LC phases.

301

However, in both cases QU decreases phase transition temperatures of ordered LC phases into L2

302

phase, which is observed for temperatures higher than 45 °C. Here it may be concluded that

303

effect of QU on the nanostructure of the bulk SPC/GDO LC phases is lipid weight ratio-

304

dependent. Thus, at weight ratio of 60/40, H2 LC structure remains unaffected in the presence of

305

QU. In contrast, at weight ratio of 35/65 the structure of the cubic Fd m phase is very sensitive

306

to QU concentration increase and starts transformation into L2 phase already at 35 °C and 4 wt%

307

of QU. Finally, LC phases at the SPC/GDO weight ratios of 50/50 and 40/60 show moderate

308

sensitivity to QU.

Ac ce p

te

d

M

an

us

cr

ip t

278

Page 12 of 27

309 310

3.3.

SPC/GDO/P80-based LCNPs with entrapped QU

311

The aim of this part was to investigate the effects of QU on size and stability of the dispersed

313

SPC/GDO/P80-based liquid crystalline nanoparticles (LCNPs). Dispersions were prepared at

314

SPC/GDO weight ratios of 60/40 and 35/65 in the presence of polymeric stabilizer P80 at fixed

315

lipid to polymer ratio. Dispersions containing up to 4 wt% of QU were prepared and

316

characterized with regards to particle size, charge and morphology.

317

Figure 2 shows obtained particle size distributions as a function of QU concentration. As seen

318

from Figures 2a and 2b, at low QU concentrations all LCNP dispersions are well-defined

319

displaying monomodal size distributions with polydispersity indices ranging between 0.13 and

320

0.16 µm. At 2 and 4 wt% of QU additional larger aggregates of about 5 m appear in the

321

dispersions prepared at both SPC/GDO ratios. Sample examination under polarized light

322

microscope reveals that these larger aggregates are small needle-like crystals phase separated

323

from LCNPs (Figure S3, SI). Since crystals have distinct yellow color (seen in microscope under

324

nonpolarized light), there is good reason to assume that they are formed exclusively by QU. This

325

show that SPC/GDO/P80 LCNPs can homogeneously entrap only up to 2 wt% of QU when

326

compared to 4-5 wt% of QU in the bulk LC phases. This difference may be attributed to a very

327

large surface-to-volume ratio of the particles and exposure to the excess aqueous phase creating

328

more defects and crystallization centers for QU. In addition, one cannot exclude that presence of

329

P80 may also change QU solubilization properties and induce precipitation. Note however, that

330

QU crystallization and phase separation does not affect LCNP own size distribution

331

characteristics. The colloidal stability of the LCNP dispersions with entrapped QU is also good.

332

Only minor changes in the size distributions are observed after 3 months of storage at RT (Figure

333

2, dotted lines).

334

Figure 3 demonstrates that the obtained mean particle size clearly depends on both lipid ratio and

335

QU concentration. Thus, LCNPs at SPC/GDO ratio 60/40 are slightly larger with the mean

336

particle size increasing from 140 to 210 nm as QU concentration is increased from 0 to 4 wt%. In

337

contrast, LCNPs at SPC/GDO ratio 35/65 are smaller with the mean particle size ranging from

338

about 80 to 110 nm. Unfortunately, we cannot explain such an effect of QU on the particle size

339

considering only bulk phase behavior and nanostructural features of the LC phases. Since LCNP

Ac ce p

te

d

M

an

us

cr

ip t

312

Page 13 of 27

dispersions were prepared by mechanical agitation we believe that the entrapment of QU may

341

slightly increase cohesion forces within LC structure and/or influence monocrystalline domain

342

size. Therefore, more energy is required to brake LC phase with entrapped QU into smaller

343

particles. In addition, LCNP zeta potential remains unaffected by the entrapment of QU and is

344

about -16 and -10 mV for particles at SPC/GDO ratio 60/40 and 35/65, respectively. Considering

345

size and surface to volume ratio differences the surface charge density of the particles is similar

346

at both SPC/GDO ratios.

cr

347

ip t

340

Representative cryo-TEM images shown in Figure 4 together with the measured SAXD profiles (Figure S4, SI) give further insights into QU-loaded particle morphology and

349

nanostructure. Overall, the observed particle sizes are consistent with DLS measurements. At

350

SPC/GDO ratio 60/40 particles are larger and characterized by a denser inner core and a less

351

dense shell comprising lamellar and “sponge”-like petals (Figures 4a and 4b). The observed

352

particle morphology is comparable with those previously prepared at similar SPC/GDO weight

353

ratio and P80 content without QU [47]. The obtained LCNP SAXD profiles as a function of QU

354

concentration in Figure S4a (SI) are almost identical showing negligible effect of QU on particle

355

nanostructure and are in line with the bulk phase behavior observations (Figures 1a-1d). The data

356

show only weak diffuse scattering which is not surprising since it is know that dispersion agent

357

P80 has a disordering effect on the internal SPC/GDO LCNP nanostructure [43]. Only small

358

shoulder appearing at q value of about 1.1 nm-1 can be observed which position may be related to

359

the first reflection of the undispersed H2 phase (Figures 1a-1d). Regardless of this relation SAXD

360

data suggest that in dispersed state the long-range order of the H2 LC arrangement in the particle

361

core is almost or completely lost.

362

As shown in Figures 4c and 4d, particles prepared at SPC/GDO 35/65 ratio are smaller in size

363

with more or less even density and hardly any noticeable features in the interior. This is also

364

reflected in the SAXD profiles shown in Figure S4b where only weak scattering pattern is

365

observed without any indications of the Bragg diffraction. From inspection of several dozen of

366

cryo-TEM images it may be also concluded that GDO-rich particles, with entrapped QU do not

367

possess pronounced swollen coronas of multiply connected bilayers which are always present for

368

SPC/GDO LCNPs prepared at high fraction of P80 [47, 48]. The observed difference may be

369

related to the bulk phase behavior where even low concentrations of QU were able to induce

370

Fd m phase transformation towards more reversed L2 phase (Figures 1e-1h). Most likely, in the

Ac ce p

te

d

M

an

us

348

Page 14 of 27

371

dispersed LC state the addition of small amount of QU is enough to allow more homogeneous

372

distribution of components within the particle and to prevent the formation of lamellar-like

373

corona.

374

3.4.

376

LCNPs

QU stability in SPC/GDO non-aqueous formulations, bulk LC phases and dispersed

ip t

375

cr

377

One of the important aspects of delivery formulations is the chemical stability of active

379

substance. Therefore, HPLC was used to evaluate the chemical stability of QU entrapped in all

380

three studied SPC/GDO-based systems: non-aqueous formulations, LC phases and LCNP

381

dispersions. Samples were prepared at two different SPC/GDO weight ratios containing 0.5 and

382

1 wt% of QU in respect to dry formulation weight.

383

For up to 90 days QU showed very good chemical stability when solubilized in non-aqueous

384

SPC/GDO-based formulations (Figures 5a and 5d). About 85–90% of QU is retained

385

independently on lipid composition and QU concentration. Introduction of water and hydration

386

of the formulations into bulk LC phases (Figures 5b and 5e) and LCNP dispersions (Figures 5c

387

and 5f) had more pronounced effect on the stability of QU. Gradual but significant loss of QU

388

content upon storage was observed in both systems. In the LC phases prepared at SPC/GDO ratio

389

35/65 QU content decreased to about 65%, whereas only 25–40% of QU was found after the end

390

of the study in the LC phases prepared at lipid weight ratio of 60/40. Such substantial difference

391

may be explained by the fact that SPC used in this study contain considerable amounts of linoleic

392

(C18:2) fatty acid chains. It is known that degree of unsaturation plays positive role in the lipid

393

autoxidation processes, in which highly reactive primary and secondary lipid oxidation products

394

are produced [49]. In SPC-rich LC phases larger amounts of reactive species may be formed,

395

which interact and chemically degrade QU when compared to the GDO-rich LC phases where

396

oleic fatty acid residues are dominant.

397

The lowest stability of QU was found in the LCNP dispersions where only 30–45% was

398

recovered after 90 days. Here, GDO-rich LCNPs showed only slight improvement over SPC-rich

399

dispersions. Most likely, P80 which was used to stabilize LCNPs dispersions also influence lipid

400

autoxidation complex processes and further decrease stability of QU. On the other hand, water

Ac ce p

te

d

M

an

us

378

Page 15 of 27

penetration ability and diffusion of lipid components is less restricted within small particle when

402

compared to macroscopic bulk LC phase and may accelerate degradation of QU.

403

Overall, QU shows good chemical stability when loaded in SPC/GDO-based non-aqueous

404

formulations for up to 3 months of storage at room temperature. On the other hand, hydrated

405

formulations (bulk LC phases and LCNP dispersions) show reduced QU chemical stability and

406

therefore are not suited for a long-term shelf storage but can still serve as suitable delivery

407

systems if are formed via in situ hydration of non-aqueous formulation (in case of LC phases) or

408

freshly produced just prior application/administration (in case of LCNP dispersions). For real-life

409

application, further studies on parameter optimization (e.g., sterilization, oxygen concentration,

410

temperature, pH, additives and other) in order to extend long-term stability of the formulations

411

will be needed.

412

3.

an

us

cr

ip t

401

Conclusion

M

413

The present study demonstrates the potential of liquid crystalline structures forming SPC/GDO

415

lipid formulations as encapsulation and delivery matrices of QU. The results have shown that up

416

to 5 wt% of QU can be effectively solubilized in non-aqueous SPC/GDO formulations by using

417

small amounts of EtOH as solvent. Prepared formulations are stable and have minimal effect on

418

the chemical stability of QU for few months. Upon hydration non-aqueous formulations can

419

easily self-assemble into non-lamellar bulk LC matrices with different nanostructures. The effect

420

of QU on the nanostructure of LC phases is lipid composition dependent. At high SPC content,

421

the entrapment of QU has practically no effect on the nanostructure of H2 phase at

422

physiologically relevant temperatures. At low SPC content, QU slightly increases the unit cell

423

dimensions of the reversed micellar Fd m cubic phase and promotes the formation of reversed

424

micellar solution at elevated temperatures. Finally, concentrated and colloidally stable non-

425

lamellar SPC/GDO-based LCNP dispersions containing up to 2 wt% of entrapped QU can be

426

easily prepared in the presence of stabilizer P80. Dispersion particle size can be tuned in the

427

range of about 80 – 210 nm by changing lipid composition and entrapped QU concentration.

Ac ce p

te

d

414

428 429

Author’s contributions

430

Page 16 of 27

431

Linkeviciute A. executed the experiment, extracted the data and wrote the initial draft of the

432

manuscript. Barauskas J. supervised the research, analysis of data and interpreted the results

433

obtained from the models. Misiunas A. technical assistance, while Naujalis E. was co-

434

supervising the research.

436

ip t

435

Conflict of interest

438

cr

437

The authors declare that they have no conflict of interest.

us

439

Acknowledgements

441

The authors acknowledge the Swedish synchrotron X-ray facility MAX IV Laboratory for

442

allocated beamtime at the I911-4beamline and Ana Labrador for technical support during

443

experiments. Authors also thank Gunnel Karlsson and Viveka Alfredsson at Lund University for

444

their help with the cryo-TEM imaging. AL acknowledges financial support from the Research

445

Council of Lithuania "Promotion of Student Scientific Activities" (VP1-3.1-ŠMM-01-V-02-003).

M

d te Ac ce p

446

an

440

Page 17 of 27

446 447

FIGURES

448

Figure 1. SAXD profiles of SPC/GDO mixtures in excess water prepared at lipid weight ratios

450

of 60/40 (a-d) and 35/65 (e-h) as a function of QU concentration (0, 0.5, 1.0 and 4.0 wt%) and

451

temperature (25, 35, 45 and 55 °C). Arrows in (e) and (h) show indexing of the reflections from

452

the reversed Fd m cubic phase. An explanation is given in the text.

us

453

cr

ip t

449

Figure 2. Particle size distributions of freshly prepared (solid lines) and 3 months old (dotted

455

lines) SPC/GDO/P80 LCNPs at SPC/GDO weight ratios of 60/40 (a) and 35/65 (b) as a function

456

of QU concentration. Dispersions were prepared at fixed lipid/P80 weight ratio of 75/15 in

457

82.5% water.

M

an

454

458

Figure 3. Dependence of mean particle size of the SPC/GDO/P80 LCNPs on QU concentration.

460

Dispersions were prepared at SPC/GDO weight ratios of 60/40 (open circles) and 35/65 (filled

461

circles). The lines are drawn to guide the eye.

te

Ac ce p

462

d

459

463

Figure 4. Representative cryo-TEM images of SPC/GDO/P80 LCNPs with entrapped 1 wt% QU

464

prepared at SPC/GDO weight ratios of 60/40 (a and b) and 35/65 (c and d). Dispersions were

465

prepared at fixed lipid/P80 weight ratio of 75/15 in 82.5% water.

466 467

Figure 5. Chemical stability of QU (normalized to initial) entrapped in SPC/GDO-based non-

468

aqueous formulations (a and d), LC phases (b and e) and LCNP dispersions (c and f). Samples

469

were prepared at fixed SPC/GDO weight ratios of 60/40 (open symbols) and 35/65 (filled

470

symbols) containing 0.5 (a-c) and 1.0 wt% (d-f) of QU in respect to dry weight. Results are

471

represented as mean values ± standard deviation (n=3).

472

Page 18 of 27

472 473

References

474

te

d

M

an

us

cr

ip t

[1] W. Vermerris and R. Nicholson, Phenolic Compound Biochemistry, Springer, 2006. [2] S. Kumar and A.K. Pandey, Sci World J, (2013). [3] D. Singh, M.S.M. Rawat, A. Semalty and M. Semalty, Curr Drug Deliv, 9 (2012) 305-314. [4] F. Bonina, M. Lanza, L. Montenegro, C. Puglisi, A. Tomaino, D. Trombetta, F. Castelli and A. Saija, Int J Pharmaceut, 145 (1996) 87-94. [5] A.C. Santos, S.A. Uyemura, J.L.C. Lopes, J.N. Bazon, F.E. Mingatto and C. Curti, Free Radical Bio Med, 24 (1998) 1455-1461. [6] M.I. Kaldas, U.K. Walle, H. Van der Woude, J.M. McMillan and T. Walle, J Agr Food Chem, 53 (2005) 4194-4197. [7] N. Nuengchamnong, A. Hermans-Lokkerbol and K. Ingkaninan, Naresuan University Journal, 12 (2004) 25-37. [8] J.V. Formica and W. Regelson, Food Chem Toxicol, 33 (1995) 1061-1080. [9] Y. Zheng, I.S. Haworth, Z. Zuo, M.S.S. Chow and A.H.L. Chow, J Pharm Sci-Us, 94 (2005) 1079-1089. [10] R.K. Hansen, S. Oesterreich, P. Lemieux, K.D. Sarge and S.A.W. Fuqua, Biochem Bioph Res Co, 239 (1997) 851-856. [11] G. Elia, C. Amici, A. Rossi and M.G. Santoro, Cancer Res, 56 (1996) 210-217. [12] M. Koishi, N. Hosokawa, M. Sato, A. Nakai, K. Hirayoshi, M. Hiraoka, M. Abe and K. Nagata, Jpn J Cancer Res, 83 (1992) 1216-1222. [13] T. Pralhad and K. Rajendrakumar, J Pharmaceut Biomed, 34 (2004) 333-339. [14] R. Casagrande, S.R. Georgetti, W.A. Verri, M.F. Borin, R.F.V. Lopez and M.J.V. Fonseca, Int J Pharmaceut, 328 (2007) 183-190. [15] S. Jain, A.K. Jain, M. Pohekar and K. Thanki, Free Radical Bio Med, 65 (2013) 117-130. [16] T.H. Tran, Y. Guo, D.H. Song, R.S. Bruno and X.L. Lu, J Pharm Sci-Us, 103 (2014) 840-852. [17] L. Gao, G.Y. Liu, X.Q. Wang, F. Liu, Y.F. Xu and J. Ma, Int J Pharmaceut, 404 (2011) 231-237. [18] D.H. Lee, G.S. Sim, J.H. Kim, G.S. Lee, H.B. Pyo and B.C. Lee, J Pharm Pharmacol, 59 (2007) 16111620. [19] F.T.M.C. Vicentini, T.R.M. Simi, J.O. Del Ciampo, N.O. Wolga, D.L. Pitol, M.M. Iyomasa, M.V.L.B. Bentley and M.J.V. Fonseca, Eur J Pharm Biopharm, 69 (2008) 948-957. [20] Guo Chen-yua, Yang Chun-fena, Li Qi-luc, Tan Qia, Xi Yan-weia, Liu Wei-nad and Z. Guang-xia, Int J Pharmaceut, 430 (2012) 292– 298. [21] S. Bose and B. Michniak-Kohn, Eur J Pharm Sci, 48 (2013) 442-452. [22] Y.Y. Zhang, Y. Yang, K. Tang, X. Hu and G.L. Zou, J Appl Polym Sci, 107 (2008) 891-897. [23] Q. Tan, W.D. Liu, C.Y. Guo and G.X. Zhai, Int J Nanomed, 6 (2011) 1621-1630. [24] H.L. Li, X.B. Zhao, Y.K. Ma, G.X. Zhai, L.B. Li and H.X. Lou, J Control Release, 133 (2009) 238-244. [25] E. Alexopoulou, A. Georgopoulos, K.A. Kagkadis and C. Demetzos, J Liposome Res, 16 (2006) 17-25. [26] A. Priprem, J. Watanatorn, S. Sutthiparinyanont, W. Phachonpai and S. Muchimapura, NanomedNanotechnol, 4 (2008) 70-78. [27] A.A. Date, M.S. Nagarsenker, S. Patere, V. Dhawan, R.P. Gude, P.A. Hassan, V. Aswal, F. Steiniger, J. Thamm and A. Fahr, Mol Pharmaceut, 8 (2011) 716-726. [28] A.P. Landi-Librandi, T.N. Chrysostomo, A.E.C.S. Azzolini, C.M. Marzocchi-Machado, C.A. de Oliveira and Y.M. Lucisano-Valim, J Liposome Res, 22 (2012) 89-99.

Ac ce p

475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515

Page 19 of 27

545 546 547

te

d

M

an

us

cr

ip t

[29] P.G. Cadena, M.A. Pereira, R.B.S. Cordeiro, I.M.F. Cavalcanti, B.B. Neto, M.D.C.B. Pimentel, J.L. Lima, V.L. Silva and N.S. Santos-Magalhaes, Bba-Biomembranes, 1828 (2013) 309-316. [30] C. Caddeo, O. Diez-Sales, R. Pons, X. Fernandez-Busquets, A.M. Fadda and M. Manconi, Pharm ResDordr, 31 (2014) 959-968. [31] H.L. Mu, R. Holm and A. Mullertz, Int J Pharmaceut, 453 (2013) 215-224. [32] K. Sandeep, M. Mohanvarma and P. Veerabhadhraswamy, Acta Pharmaceutica Sinica B, 3 (2013) 361-372. [33] V. Luzzati, R. Vargas, P. Mariani, A. Gulik and H. Delacroix, J Mol Biol, 229 (1993) 540-551. [34] K. Larsson, P. Quinn, K. Sato and F. Tiberg, Vol. 19 (2006). [35] L. Sagalowicz and M.E. Leser, Curr Opin Colloid In, 15 (2010) 61-72. [36] K. Larsson and F. Tiberg, Curr Opin Colloid In, 9 (2005) 365-369. [37] A. Angelova, B. Angelov, R. Mutafchieva, S. Lesieur and P. Couvreur, Accounts Chem Res, 44 (2011) 147-156. [38] I. Amar-Yuli, D. Libster, A. Aserin and N. Garti, Curr Opin Colloid In, 14 (2009) 21-32. [39] S.B. Rizwan, B.J. Boyd, T. Rades and S. Hook, Expert Opin Drug Del, 7 (2010) 1133-1144. [40] F. Tiberg, M. Johnsson, M. Jankunec and J. Barauskas, Chem Lett, 41 (2012) 1090-1092. [41] J. Barauskas, L. Christerson, M. Wadsater, F. Lindstrom, A.K. Lindqvist and F. Tiberg, Mol Pharmaceut, 11 (2014) 895-903. [42] F. Tiberg and M. Johnsson, J Drug Deliv Sci Tec, 21 (2011) 101-109. [43] M. Wadsater, J. Barauskas, T. Nylander and F. Tiberg, Acs Appl Mater Inter, 6 (2014) 7063-7069. [44] M. Wadsater, J. Barauskas, T. Nylander and F. Tiberga, Soft Matter, 9 (2013) 8815-8819. [45] S. Murgia, S. Bonacchi, A.M. Falch, S. Lampis, V. Lippolis, V. Meli, M. Monduzzi, L. Prodi, J. Schmidt, Y. Talmon and C. Caltagirone, Langmuir, 29 (2013) 6673-6679. [46] A. Linkeviciute, R. Butkiene and E. Naujalis, Chemija, 24 (2013) 217-222. [47] D.P. Chang, M. Jankunec, J. Barauskas, F. Tiberg and T. Nylander, Langmuir, 28 (2012) 10688-10696. [48] D.P. Chang, M. Jankunec, J. Barauskas, F. Tiberg and T. Nylander, Acs Appl Mater Inter, 4 (2012) 2643-2651. [49] A. Morales, C. Dobarganes, G. Marquez-Ruiz and J. Velasco, J Am Oil Chem Soc, 87 (2010) 12711279.

Ac ce p

516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544

Page 20 of 27

547 548 549 550 551

 We examine the mixtures of SPC and GDO as encapsulation matrices for quercetin (QU).  SPC/GDO-based formulations can incorporate relatively high amounts of QU.  At high SPC content, the loaded QU has no effect on the nanostructure of H2 phase.  At low SPC content, QU slightly increases the unit cell dimensions of Fd m phase.

Ac ce p

te

d

M

an

us

cr

ip t

552

Page 21 of 27

Ac

ce

pt

ed

M

an

us

cr

i

Figure(s)

Page 22 of 27

Ac ce p

te

d

M

an

us

cr

ip t

Figure(s)

Page 23 of 27

Ac

ce

pt

ed

M

an

us

cr

i

Figure(s)

Page 24 of 27

Ac ce p

te

d

M

an

us

cr

ip t

Figure(s)

Page 25 of 27

Ac

ce

pt

ed

M

an

us

cr

i

Figure(s)

Page 26 of 27

Ac

ce

pt

ed

M

an

us

cr

i

*Graphical Abstract (for review)

Page 27 of 27

Preparation and characterization of quercetin-loaded lipid liquid crystalline systems.

The aim of the present study was to investigate mixtures of soy phosphatidylcholine (SPC) and glycerol dioleate (GDO) as encapsulation matrices for an...
465KB Sizes 1 Downloads 8 Views