Investigation into the Solid-State Properties and Dissolution Profile of

Publication Date (Web): July 18, 2018. Copyright © 2018 American ... ASDs adversely affected the physical stability and dissolution by promoting crys...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Article

An investigation into the solid-state properties and dissolution profile of spray dried ternary amorphous solid dispersions: A rational step towards the design and development of multi-component amorphous system Shrawan Baghel, Helen Cathcart, and Niall J. O'Reilly Mol. Pharmaceutics, Just Accepted Manuscript • DOI: 10.1021/acs.molpharmaceut.8b00306 • Publication Date (Web): 18 Jul 2018 Downloaded from http://pubs.acs.org on July 19, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

1 2 3 4 5 6 7 8 9 10 11 12 13 14

Title: An investigation into the solid-state properties and dissolution profile of spray

15

Shrawan Baghel, Helen Cathcart and Niall J. O’Reilly

16 17 18

Synthesis and Solid State Pharmaceutical Centre (SSPC), Pharmaceutical and Molecular Biotechnology Research Centre (PMBRC), Waterford Institute of Technology, Cork Road, Waterford, Ireland.

dried ternary amorphous solid dispersions: A rational step towards the design and development of multi-component amorphous system

19 20 21 22 23 24 25 26 27 28 29 30 31 32

1 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

33

ABSTRACT:

34

The formulation of oral amorphous solid dispersions (ASD) includes the use of excipients to

35

improve physical stability and enhance bioavailability. Combinations of excipients (polymers

36

and surfactants) are often employed in pharmaceutical products to improve the delivery of

37

poorly water-soluble drugs. However, additive interactions in multi-component ASD systems

38

have not been extensively studied and may promote crystallization in an unpredictable

39

manner, which in turn may affect the physical stability and dissolution profile of the product.

40

The main aim of this study was to understand the effect of different surfactant and polymer

41

combinations on the solid-state properties and dissolution behaviour of ternary spray dried

42

solid dispersions of dipyridamole and cinnarizine. The surfactants chosen for this study were

43

sodium dodecyl sulphate and poloxamer 188 and the model polymers used are polyvinyl

44

pyrrolidone K30 and hydroxypropyl methylcellulose K100. The spray dried ternary

45

dispersions maintained higher supersaturation compared to either the crystalline drug

46

equilibrium solubility or their respective physical mixtures. However, rapid and variable

47

dissolution behaviour was observed for different formulations. The maximum supersaturation

48

level was observed with drug-polymer-polymer ternary dispersions. On the other hand,

49

incorporating the surfactant into binary (drug-polymer) and ternary (drug-polymer-polymer)

50

ASDs adversely affected the physical stability and dissolution by promoting crystallization.

51

Based on these observations, a thorough investigation into the impact of combinations of

52

additives on amorphous drug crystallization during dissolution and stability studies is

53

recommended in order to develop optimized formulations of supersaturating dosage forms.

54

Keywords: Amorphous solid dispersion, spray drying, dissolution, supersaturation, drug-

55

polymer interaction, nuclear magnetic resonance, crystallization, surfactant, synergism

56 57 58 59 60

2 ACS Paragon Plus Environment

Page 2 of 45

Page 3 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

61

1. INTRODUCTION

62

The majority of new chemical entities in the development pipeline are poorly water

63

soluble, thereby posing a challenge for oral formulation development.1 Various strategies,

64

such as formation of prodrug,2 lipid formulations,3 micronization,4 surfactants,5 salt

65

formation,6 and the development of amorphous solid dispersions (ASDs)7 have recently been

66

utilized to enhance the solubility and dissolution rate of poorly soluble active pharmaceutical

67

ingredients (APIs). Amorphization of crystalline drugs is a promising strategy to overcome

68

the solubility challenge.8 Being a thermodynamic high-energy form (random arrays of

69

molecules), the apparent solubility and dissolution rate of amorphous APIs are higher than

70

their crystalline counterparts.9, 10 However, the high free energy of the amorphous form also

71

renders it prone to crystallization and gives rise to physical instability.11 Crystallization of

72

APIs is a two-step process involving nucleation followed by crystal growth.11 Nucleation is a

73

process in which the drug molecules form small stable clusters or aggregates of a critical size.

74

This is a rate limiting step in the crystallization process and the rate of nucleation depends on

75

the crystallization activation energy or solid-liquid interfacial energy.12,

76

crystal growth phase is governed by the diffusion of solute molecules to the nuclei interface

77

and addition to the crystal lattice. Thus, one of the main challenges in formulating ASDs is

78

the prevention of nucleation and crystal growth in APIs which results in poor physical

79

stability and compromised oral bioavailability.9, 14

13

The subsequent

80

In an attempt to delay crystallization and improve physical stability, different

81

polymers (which interfere with either nucleation or crystal growth or both) have been used to

82

prepare binary ASDs.15 Polymers, usually used at high concentrations, reduce the molecular

83

mobility of the amorphous API in a binary dispersion by increasing the glass transition

84

temperature (Tg) of the system. While it is important to improve the physical stability of ASD

85

formulations, generating and maintaining drug supersaturation during dissolution is equally

86

important.16 When the intraluminal concentration of an orally administered drug exceeds its

87

thermodynamic equilibrium solubility, the drug is supersaturated in the gastric media.

88

Recently, binary ASDs (drug-polymer) have gained attention as an effective method of

89

improving the supersaturation concentration of poorly soluble drugs during dissolution.17, 18

90

Different mechanisms have been proposed to explain the improvement in drug dissolution

91

and stability from binary ASDs including the formation of fine particles, drug-polymer

92

interaction in solution, formation of a physical barrier to crystallization (adsorption of the

3 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

93

polymer molecules onto the crystal surfaces), viscosity enhancement, reduction in Gibb’s free

94

energy and an anti-plasticization effect of polymer on the drug.9, 19

95

Conceptually, multi-component ASDs are considered to be an option to further

96

improve the solubility of poorly water-soluble drugs. A multi-component system,

97

incorporating other pharmaceutical excipients in addition to the drug and amorphous polymer

98

matrix, modifies the drug particles’ microenvironment at the dissolution front leading to the

99

enhancement of API solubility.20 The physical stability of the amorphous dispersions can be

100

improved by the selection of appropriate matrix components for multi-component system.21

101

Therefore, multi-component ternary ASDs composed of API, amorphous polymer matrix and

102

functional excipients are considered to be a propitious step towards the improvement of

103

physical stability and enhancement of dissolution of poorly water-soluble drugs. However, in

104

formulating such systems several issues, such as the formation of homogenous dispersions

105

and the physical stability, should be considered. It has been observed that in some cases the

106

solution concentration of free drug during the dissolution of ternary ASDs (drug, polymer and

107

an additive) declined due to the recrystallization of an amorphized additive.22 It has also been

108

reported that multi-component systems of drug and polymer may undergo phase separation

109

due to the addition of an additive to the binary ASD system.23 There is a paucity of literature

110

on the application of multi-component spray dried solid dispersions and further research is

111

required to understand and optimize ternary ASDs. Thus it would seem that the choice and

112

quantity of functional additive may be critical to formulating stable, multi-component ASDs.

113

Therefore, a rational starting point towards the design of a multi-component ternary ASDs,

114

where the API is one of the amorphous state components, would be a ternary mixture of

115

amorphous polymer and functional excipients forming drug-polymer-excipient system.

116

Generally the third component is either a polymer or surfactant to form ternary ASDs.

117

To date, very few studies have examined the effect of a third component on the stability and

118

dissolution of ASDs.23, 24, 25, 26 The addition of a third component commonly aims to further

119

improve the physical stability and increase the dissolution by inhibiting drug recrystallization

120

during shelf life and in solution state, respectively. Moreover, it would be interesting to see

121

the effect of combining two different polymers which have different types of interactions

122

with the drug on the physical stability and dissolution behaviour of ternary ASDs. Although

123

there are many publications on amorphous drug crystallization inhibition in the presence of a

124

single polymer (binary ASDs), there is a need to understand drug crystallization (both in the

125

solid-state and during dissolution) in the presence of two polymers (ternary ASDs). Due to an 4 ACS Paragon Plus Environment

Page 4 of 45

Page 5 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

126

increase in the overall number of poorly soluble APIs, ternary ASDs having two different

127

polymers may be utilised to further enhance their physical stability and apparent solubility.

128

Another class of pharmaceutical excipient used to improve the aqueous solubility and

129

dissolution performance of poorly water-soluble drugs are surfactants.27, 28 As a consequence,

130

it may be expected that incorporating surfactants into ASDs to form drug-polymer-surfactant

131

ternary ASDs may further enhance the dissolution rate of the poorly water-soluble drug by

132

promoting wettability of the particles. In this work, the main reason for choosing surfactants

133

as third component in addition to binary drug-polymer combinations was based on the fact

134

that polymer may dissolve the amorphous drug in the solid state, leading to a stable system

135

without crystallization or phase separation, while the surfactant may promote wettability in

136

solution which can lead to an increase in solubility and dissolution rate of the drug.

137

Previously, surfactants have been used as plasticizers for polymers during hot melt extrusion

138

processes to enhance drug-polymer miscibility.29, 30 However, currently there is very limited

139

literature available on the preparation and characterization of spray dried drug-polymer-

140

surfactant dispersions and therefore, it is of interest to study the effect of surfactants on ASD

141

properties when incorporated within spray dried ASDs.31,

142

surfactants on drug release from ASDs has also been reviewed, with some studies finding a

143

positive impact on the release kinetics of hydrophobic APIs and others reporting a negative

144

effect of surfactants on the dissolution of amorphous drugs from solid dispersions.33, 34 It is

145

therefore reasonable to suggest that surfactants, like polymers, may have an impact on the

146

physical stability and dissolution behaviour of ASDs and it is of interest to explore this

147

possibility. Consequently, the physical stability and dissolution behaviour of spray dried

148

ASDs with and without surfactants as a ternary agent will also be investigated in this study.

32

Furthermore, the role of

149

We have chosen polyvinyl pyrrolidone K30 (PVP), hydroxypropyl methylcellulose

150

K100 (HPMC) and a combination of both as a carrier matrix for ASDs. The surfactants

151

selected for this study are poloxamer 188 (P188) and sodium dodecyl sulphate (SDS). P188,

152

which is a non-ionic, triblock copolymer, comprised of a hydrophobic block of polypropylene

153

glycol and two hydrophilic blocks of polyethylene glycol, is categorised as a surfactant in this

154

study as the use of this excipient in the pharmaceutical area is related to its surfactant

155

properties. The other surfactant investigated in this study is SDS, which is an anionic

156

surfactant. The selection of these excipients was based on the anticipation that differences in

157

their physicochemical characteristics may allow one to distinguish the critical factors

158

affecting the drug-polymer interactions and miscibility more clearly. As a control, we have 5 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 45

159

also studied the drug-polymer-polymer-surfactant (quaternary ASDs) combinations to probe

160

the effect of surfactant on drug-polymer-polymer system.

161

Dipyridamole (DPM) and cinnarizine (CNZ) were selected as model drugs because of

162

their poorly water-soluble properties. Although weakly basic and therefore soluble in gastric

163

media in the crystalline state, DPM and CNZ have been chosen as model compounds for

164

amorphous solid dispersion studies to understand the mechanism involved in improving

165

physical stability and aqueous solubility. They are particularly suited because:

166

1. They are typical BCS class II drugs with practical poor aqueous solubility (~6.8

167

µg/mL at pH 6.8 for DPM and ~2 µg/mL at pH 6.5 for CNZ) and high GI membrane

168

permeability (Log P values for DPM and CNZ are 3.71 and 5.71, respectively).11, 19

169

2. The glass transition temperatures of DPM and CNZ are ~38°C and ~6°C,

170

respectively.11 It would be interesting to see the behaviour of two different compound, one

171

having Tg above room temperature (DPM) and the other having Tg below room temperature

172

(CNZ).

173

3. DPM and CNZ have a different number of H-bond donor (DPM – 4 and CNZ – 0)

174

and acceptor groups (DPM – 12 and CNZ – 2).19 Thus, it would be interesting to compare the

175

recrystallization tendency of DPM and CNZ in a polymer matrix since both the drugs have

176

different nature and strength of forming H-bonds.

177

Recently, we have reported that the DPM and CNZ are very unstable in their

178

amorphous forms and have high crystallization tendency.11 We also reported the role of drug-

179

polymer interaction and anti-plasticization in improving the solid-state stability and in-vitro

180

dissolution of DPM and CNZ binary ASDs.19 14 This is a continuation of our previous work

181

in order to understand the role of polymer-surfactant and polymer-polymer combinations on

182

the physical stability and dissolution profile of ASDs by preparing and characterizing spray

183

dried binary (drug-polymer), ternary (drug-polymer-surfactant or drug-polymer-polymer) and

184

quaternary

185

characterized by differential scanning calorimetry (DSC), modulated DSC, Fourier-transform

186

infrared (FTIR), in-vitro dissolution, 1H nuclear magnetic resonance (NMR) and dynamic

187

vapour sorption (DVS). This study will add to the overall understanding of the use of

188

excipients in ASDs and will help in future multi-component amorphous product

189

development.

(drug-polymer-polymer-surfactant)

solid

dispersions.

6 ACS Paragon Plus Environment

The

ASDs

were

Page 7 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

190

2. MATERIALS AND METHODS

191

2.1. Materials

192

Dipyridamole (DPM), cinnarizine (CNZ), polyvinyl pyrrolidone K30 (PVP) and

193

poloxamer 188 were supplied by Sigma Aldrich, Ireland. Hydroxypropyl methylcellulose

194

K100 (HPMC) was obtained from Colorcon, England, and sodium dodecyl sulphate (SDS)

195

(>99%) was obtained from Fisher Scientific, U.K. All reagents were of analytical grade and

196

used without further purification. The chemical structure and key physicochemical properties

197

of the model drugs and polymers used in this study are shown in Figure 1 and Table 1,

198

respectively.

199

Table 1. Physicochemical properties of model drugs and polymers

Log P b

pKa

37.44

3.71

6.435

121.25

5.86

5.71

8.436

-

-

164.13

-

-

450,000

-

-

153.83

-

-

SDS

288.37

-

-

-

-

-

P188

102.13

14.43

48.67

-

-

-

Molecular

ᅪࡴࢌ

ࢀ࢓

ࢀࢍ

weight (g/mol)

(kJ/mol)a

(°C)a

(°C)a

DPM

504.63

29.06

168.11

CNZ

368.51

38.33

PVP

40,000

HPMC

200

a

b

201

2.2. Preparation of amorphous solid dispersions

Values obtained from DSC (n=3), Values obtained from literature

19

202

The spray dried solid dispersions were prepared using a Pro-C-epT 4M8-TriX spray

203

drier (Zelzate, Belgium) with a bifluid nozzle. The spray drying solutions were prepared by

204

dissolving the component mixtures in dichloromethane-ethanol mixture (50:50 v/v). The

205

drug-polymer proportion in the binary mixture was 20:80 w/w. For the ternary solid

206

dispersion systems, the proportion of drug-polymer-surfactant was 20:75:5 and drug-

207

polymer-polymer was 20:40:40. The amount of surfactant in ternary ASDs was kept low as

208

they are not well tolerated in the body at high concentration.9 The proportion of drug-

209

polymer-polymer-surfactant in quaternary systems was 20:37.5:37.5:5 w/w. The key

210

processing parameters were feed concentration (10% w/v in dichloromethane-ethanol

211

mixture), spray rate (6 mL/min), nozzle size (0.2 mm), atomization air flow rate (5 L/min),

212

inlet temperature (80 ± 5°C), outlet drying temperature (50 ± 5°C) and drying air rate (100

213

± 10 L/min). The spray dried solid dispersions were collected using a small cyclone separator

214

and stored in a vacuum desiccator at room temperature for further analysis. The powder yield

215

obtained was approximately 60-70% w/w. 7 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

216

2.3. Preparation of physical mixtures

217

Physical mixtures for in-vitro dissolution testing and melting point depression (MPD)

218

analysis were prepared by manually mixing the components by mortar and pestle. The weight

219

fraction of drug, polymer and surfactant for in-vitro dissolution testing was identical to the

220

weight fraction used for preparing the spray dried solid dispersion. The ratio of drug and

221

polymer chosen to perform MPD analysis for binary systems was 95:5, 90:10, 85:15 and

222

80:20 w/w and for ternary systems (dug-polymer-polymer) was 95:2.5:2.5, 90:5:5, 85:7.5:7.5

223

and 80:10:10 w/w.

224

2.4. Thermogravimetric analysis (TGA)

225

The residual solvent content of the spray dried solid dispersions was assessed using a

226

TGA Q50 (TA Instruments Corp., Elstree, Herts, U.K.). Samples were heated at 10°C/min

227

from 25 to 200 °C. During all TGA experiments nitrogen was used as the purging gas at 50

228

mL/min. All analyses were performed in duplicate.

229

2.5. Differential scanning calorimetry (DSC)

230

Thermal behaviour of the ASDs was analysed using a TA instrument Q2000 DSC.

231

Temperature and heat flow calibration was carried out using a high-purity indium standard.

232

Nitrogen was used as the purge gas at a flow rate of 50 mL/min. 3-5 mg of samples were

233

accurately weighed into aluminium pans with pin holes. The samples were heated from 0 to

234

200 ºC at a heating rate of 10 ºC/min. MPD analysis was carried out at a heating rate of 2

235

ºC/min from 0 to 200 ºC. The glass transition temperature (Tg) of the freshly prepared and

236

aged samples was analysed using modulated DSC (mDSC). 3-5 mg of the samples were

237

accurately weighed into aluminium pans with pin holes. The experiment was performed

238

under nitrogen gas flow (50 mL/min) using an underlying heating rate of 5 ºC/min from 0 to

239

200 ºC and a modulation amplitude of ± 0.53 ºC over a period of 40 s for all the samples. All

240

measurements were performed in triplicate and data analysis was performed using Universal

241

Analysis Software (TA instruments).

242

2.6. Fourier-Transform Infrared Spectroscopy

243

Fourier-transform Infrared Spectroscopy (FTIR) was performed using a Varian 660-

244

IR FT-IR Spectrometer (32 scans at 4 cm-1 resolution). Test samples were mixed with KBr

245

and then compressed into a disk and analysed immediately.

8 ACS Paragon Plus Environment

Page 8 of 45

Page 9 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

246

2.7. In-vitro dissolution study of solid dispersions

247

The equilibrium solubility of crystalline drugs and dissolution rate of amorphous

248

DPM and CNZ from various powdered physical mixtures and ASDs, respectively, were

249

measured using the USP II paddle method. Samples containing 25 mg of DPM and CNZ

250

were added to 500 mL of phosphate buffer solution pH 6.8 (PBS 6.8) at 37.0 ± 0.2 °C. The

251

solution was stirred at 100 rpm using a USP II paddle apparatus. 2 mL samples were

252

withdrawn from each vessel at pre-defined intervals (5, 10, 15, 30, 60, 120, 180, 240, 300,

253

360 and 1440 min) and centrifuged for 5 min at 15000 rpm. At each time-point the same

254

volume of fresh medium was replaced. The concentration of DPM and CNZ was determined

255

using a UV-Vis spectrophotometer (Varian VK 7010 L1168, Santa Clara, USA). All

256

measurements were carried out in duplicate.

257

2.8. Solution state Nuclear Magnetic Resonance (NMR)

258

In order to understand the molecular mechanism of drug-polymer and drug-polymer-

259

polymer interactions, 1H NMR spectra were recorded on a Jeol ECX-400 spectrometer

260

operating at 400 MHz. The measurements were performed at 40°C in dimethylsulfoxide

261

(DMSO-d6) using tetramethylsilane (TMS) as an internal standard. Initial trials were

262

performed using D2O. However the poor solubility of the model drugs in D2O leads to weak

263

signals and the chemical shifts were not clear. Therefore, DMSO-d6 has been selected due to

264

its high dielectric constant and hydrophilic nature. All the model drugs and polymers were

265

completely soluble in DMSO-d6. Furthermore, the molecular interaction between polymer

266

and surfactant were characterised by two-dimensional Nuclear Overhauser Effect

267

spectroscopy (2D-NOESY) in DMSO-d6. The polymer to surfactant ratio was kept the same

268

as that of the dissolution studies. For 2D-NOESY, the mixing time was set to 0.5 s and the

269

experiment was performed with a 1.5 s relaxation delay and 0.32 s acquisition time.

270

2.9. Moisture sorption analysis and Stability studies

271

Moisture sorption analysis of the ASDs was carried out using dynamic vapour

272

sorption (DVS)-1000 Intrinsic, Surface Measurement Systems, U.K. Amorphous drug

273

samples (5-10 mg) were analysed using a double ramp method from 0-90-0% RH (2 cycles)

274

in 10% increments (dm/dt = 0.001 at each step) at 40 °C. Solid dispersions were also stored at

275

40 °C/75% RH. The physical stability of all the samples was evaluated after 4 weeks.

9 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

276

Page 10 of 45

3. RESULTS AND DISCUSSION

277

The association of polymer-polymer or polymer-surfactant combinations in aqueous

278

solution have attracted considerable fundamental and technological interest. The resultant

279

properties of these combinations possess unique properties that are significantly different

280

from those of the individual components. Consequently, they have important applications in

281

biological and industrial processes.37 For example, synthetic water soluble polymer-polymer

282

combinations have been used in operations such as flocculation during mineral processing,38

283

while polymer-surfactant combinations have been used for immobilization of enzymes in

284

polyelectrolyte complexes,39 rheology control,40 improving the solubility of drugs with low

285

aqueous solubility and to control the drug release rate.13

286

polymer-surfactant combinations are highly relevant, since pharmaceutical formulations often

287

contains both excipients types.25 41 42 Formulations containing high-energy amorphous drugs

288

can enhance drug dissolution by generating supersaturated solutions.9 Polymers are

289

frequently added to delay crystallization, and surfactants are typically employed to improve

290

processing properties or dissolution profiles.14

25 41

Thus, polymer-polymer or

291

In the past, several publications have investigated the association between polymer-

292

polymer and polymer-surfactant mixtures in solution.13 25 41 42 The presence of hydrophilic

293

excipients, such as polymer or surfactant, in a formulation containing a polymeric matrix may

294

help prevent amorphous drug crystallization and/or prevent agglomeration of a fine

295

crystalline precipitate into larger aggregates.9 However, there has been limited systematic

296

investigation of the interplay between surfactant or polymeric additive combinations in terms

297

of the impact on solid-state properties and dissolution behaviour, although such an interplay

298

could be very influential on the performance of the dosage form. Excipients such as polymers

299

or surfactants are well-known to impact nucleation and crystal growth and may accelerate or

300

inhibit crystallization depending on their effect on crystallizing solute and solution

301

properties.41 42 This is important in the context of formulation of dosage forms where the goal

302

is to prevent crystallization of amorphous drug during the shelf-life or dissolution. It is thus

303

crucial to probe into the effect of excipient combinations on the solid-state properties and

304

dissolution profile of amorphous drugs within solid dispersions, since the interaction can

305

result in change in dispersion properties which may result in favourable or adverse effect on

306

the amorphous drug crystallization inhibition. In this study, we have investigated the effect of

307

polymer-polymer and polymer-surfactant combinations on the physical stability and

308

supersaturation behaviour during dissolution of poorly water-soluble model drugs. 10 ACS Paragon Plus Environment

Page 11 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

309

Molecular Pharmaceutics

3.1. Drug-polymer and drug-polymer-polymer miscibility

310

The first step towards developing a multi-component amorphous systems is to assess

311

the miscibility of the drug within the polymer matrix. Drug-polymer miscibility is crucial

312

requirement for the formulation of stable ASDs.19 Solubility parameters (δ) are used widely

313

to quickly and effectively determine the drug-polymer miscibility when screening different

314

systems.43 Previously, we reported the δ values of DPM (28.57 MPa1/2), CNZ (21.00 MPa1/2)

315

and PVP (26.28 MPa1/2).19 The δ value of HPMC, calculated in this study, is 22.48 MPa1/2. It

316

has been reported that compounds with a ∆δ of less than 7.0 MPa1/2 are likely to be miscible

317

because the energy released by the interaction between the components balances the enthalpy

318

of mixing within the components.25 On the other hand, compounds with ∆δ greater than 10.0

319

MPa1/2 are likely to be immiscible. In this study, both the model polymers (PVP and HPMC)

320

were expected to be miscible with both drugs (DPM and CNZ), as they exhibited a ∆δ of less

321

than 7.0 MPa1/2.

322

To further investigate the drug-polymer and drug-polymer-polymer interaction and

323

the likelihood of the formation of a miscible system after spray drying, we employed melting

324

point depression (MPD) analysis. The chemical potential of a crystalline material becomes

325

equal to the chemical potential of its molten material at its melting point. However, drug-

326

polymer miscibility reduces the chemical potential of the drug in the solution (molten drug

327

and polymer mixture), compared to the pure molten drug, which leads to a depression in

328

melting point of the drug dissolved in the polymer.44 Strong exothermic mixing of the drug

329

and polymer produces a large depression in melting point of the drug, whereas weakly

330

exothermic mixing or immiscible drug-polymer system should result in less significant MPD

331

or no depression at all. We have employed this method previously to investigate drug-

332

polymer mixing thermodynamics of DPM-PVP and CNZ-PVP systems, which showed clear

333

evidence of a depression of the melting point of both drugs, indicating a significant degree of

334

drug-polymer mixing at the melting temperature of the drug.19 In this work, we have

335

investigated the DSC thermogram for physical mixtures of DPM and CNZ at different weight

336

ratios with HPMC and PVP-HPMC combinations. As shown in Figure 2 and 3, significant

337

mixing of model drugs with HPMC and PVP-HPMC were observed. To compare the

338

depression in melting point due to the individual polymers and their combination, theoretical

339

depression values for the ternary mixtures were proposed based on the contribution of each

340

polymer within the binary systems. Since the ternary mixture contains two polymers (each

341

polymer in half proportion compared to binary system), the depression from both the 11 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

342

polymers was averaged. If the experimental values showed a greater effect compared to this

343

average value, then that would indicate a synergistic benefit. It was found that the depression

344

in melting point of DPM and CNZ in the ternary systems (drug-PVP-HPMC) was higher than

345

the average values from the binary PVP and HPMC systems suggesting synergism (Figure 3).

346

This suggests that ternary ASDs may perform better than binary ASDs in increasing physical

347

stability of amorphous drug within dispersions.

348

3.2. Solid State Characterization of freshly prepared Spray Dried dispersions

349

The spray dried binary, ternary and quaternary solid dispersions were prepared and

350

examined using DSC and MDSC to determine extent of amorphization and glass transition.

351

The spray dried samples were stored in the vacuum desiccator for 48 hrs to remove the

352

residual solvent. In a standard TGA ramp test, all the dispersions showed a residual solvent

353

content of ≤2% w/w.

354 355

3.2.1. Thermal analysis: PVP-based system: As reported in Table 2, all PVP based dispersions were found to

356

be completely amorphous except for the CNZ-PVP-P188 system, where a crystalline CNZ

357

melting peak was observed (see supplementary information for full DSC thermograms). As

358

shown in Figure 4, the DPM-PVP and CNZ-PVP dispersions displayed the highest Tg values

359

(single Tg indicating molecular level mixing of the drug and polymer) when compared to

360

other PVP-surfactant based dispersions (Table 2). The surfactant based PVP dispersions of

361

DPM and CNZ (Figure 4) displayed two Tg’s (DPM-PVP-SDS, DPM-PVP-P188 and CNZ-

362

PVP-SDS) or one Tg along with a melting peak (CNZ-PVP-P188). The ASDs displaying two

363

Tg’s represent amorphous system with phase separation leading to the formation of drug-rich

364

(lower Tg region) and polymer-rich domains (higher Tg region). These results suggest that

365

surfactants have a negative effect on the stability of PVP based DPM and CNZ spray dried

366

ASDs.

367

HPMC-based system: Binary ASDs of DPM and CNZ with HPMC remained

368

completely amorphous as reported in Table 2 (see supplementary information for full DSC

369

thermograms). The dispersions displayed a single Tg (Figure 5) indicating complete mixing

370

and amorphization of the drug with HPMC. The Tg of the DPM-HPMC and CNZ-HPMC

371

systems is higher than the Tg of their PVP counterparts, suggesting that HPMC has a greater

372

antiplactization effect on the model drugs (Figure 5 and Table 2). On the other hand,

373

surfactants were found to have a negative effect on the stability of DPM-HPMC ASDs with a

374

drug melting peak, indicating the presence of crystalline drug. The negative effects of SDS 12 ACS Paragon Plus Environment

Page 12 of 45

Page 13 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

375

and P188 were more pronounced in CNZ-HPMC based dispersions where a CNZ

376

recrystallization peak was also present in addition to drug melting peak. These results

377

indicate that, in the absence of surfactants, HPMC is a better antiplasticizing agent for DPM

378

and CNZ dispersions compared to PVP whereas in the presence of surfactants PVP is a more

379

effective stabilizer of amorphous DPM and CNZ. The presence of SDS and P188 in HPMC

380

based dispersions of DPM or CNZ increased the molecular mobility of the amorphous drugs

381

(lower Tg values compared to drug-polymer system without surfactant) leading to phase

382

separation in DPM dispersions and crystallization in CNZ systems. Surfactants interfered

383

more with the stabilizing efficiency (drug-polymer interaction and miscibility) of HPMC

384

compared to PVP.

385

Table 2: Physicochemical properties of binary, ternary and quaternary ASDs

Formulations

Fresh (n = 3)

Solubility after 24 hr

Aged (n = 3)

(µg/mL) (n = 2)

Tg (°C)

∆Hm (J/g)

Tg (°C)

∆Hm (J/g)

PM

SD

DPM-PVP

147.9

-

138.7/55.3

-

7.04

22.32

DPM-PVP-SDS

139.3/100.2

-

125.8/48.2

-

7.19

19.18

DPM-PVP-P188

145.9/99.2

-

137.6/98.6

-

6.04

17.18

CNZ-PVP

106.5

-

50.8

10.84

2.43

12.84

CNZ-PVP-SDS

100.1/148.2

-

52.1

12.29

2.51

12.20

CNZ-PVP-P188

100.2

2.84

21.7

11.02

2.89

12.53

DPM-HPMC

150.2

-

134.4

1.75

6.93

26.45

DPM-HPMC-SDS

99.9

0.29

98.8

6.95

6.31

18.94

DPM-HPMC-P188

99.3

2.54

98.3

3.43

6.04

21.12

CNZ-HPMC

145.6

-

64.2

2.49

2.38

13.24

CNZ-HPMC-SDS

65.7

6.45

98.4

9.35

2.34

13.23

CNZ-HPMC-P188

41.09

8.53

111.9

10.62

2.09

13.12

DPM-PVP-HPMC

157.7

-

148.6

-

6.66

29.76

DPM-PVP-HPMC-SDS

124.2

-

109.5

-

6.41

25.48

DPM-PVP-HPMC-P188

139.1

-

117.8

-

6.95

21.41

CNZ-PVP-HPMC

66.1/143.5

-

88.2

-

2.11

16.79

CNZ-PVP-HPMC-SDS

9.7

5.78

82.2

16.96

2.27

14.57

CNZ-PVP-HPMC-P188

10.9

5.27

67.9

16.34

2.09

13.49

386 387

PM and SD represents physical mixtures containing crystalline drug and solid dispersions containing amorphous drug, respectively;

388

PVP-HPMC based system: The mDSC thermogram of all the PVP-HPMC based

389

systems are shown in Figure 6 and the Tg values are reported in Table 2. DPM-PVP-HPMC

390

ternary dispersions were found to be amorphous with the Tg of the system significantly higher 13 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

391

than the Tg of the binary DPM-PVP or DPM-HPMC based systems (see supplementary

392

information for full DSC thermograms). This confirms the improved efficiency of the

393

polymer combination in the formation of stable ternary DPM ASDs compared with binary

394

dispersions. Adding surfactants to DPM-PVP-HPMC significantly reduced the Tg of the

395

system. CNZ-PVP-HPMC dispersions displayed two Tg’s suggesting the presence of two

396

different amorphous domains. We have previously reported that CNZ exhibits no interaction

397

with PVP and formed a solid solution with PVP such that the amorphous CNZ was

398

kinetically frozen within the PVP matrix rather than being in a thermodynamic equilibrium.19

399

We found that if the concentration of CNZ within PVP is increased then it will lead to self-

400

association of CNZ molecules leading to phase separation and crystallization. Furthermore,

401

as reported previously (section 3.1 (Figure 2) and section 3.2 (HPMC based systems)),

402

HPMC has a greater stabilizing effect on amorphous CNZ due to strong drug-polymer

403

interaction compared to PVP. Therefore, it could be generalised that the lower concentration

404

of HPMC (40% w/w) in the CNZ ternary dispersion has led to the instability of the system.

405

The two Tg’s of in the CNZ-PVP-HPMC system is in accordance with these observations.

406

Furthermore, the inclusion of surfactant has negatively affected the stability of CNZ-PVP-

407

HPMC system.

408

Based on the DSC results of freshly spray dried ASDs of DPM and CNZ, it was found

409

that the surfactants promote the phase separation and crystallization of amorphous drugs

410

within spray dried ASDs which may be due to the interference with the drug-polymer

411

interaction or miscibility. In addition, the synergistic effect of PVP-HPMC combinations was

412

observed to be stronger in the DPM dispersions than CNZ dispersions.

413 414

3.2.2. Spectroscopic Analysis FTIR spectroscopy was used to assess the molecular interactions within the samples.

415

The FTIR spectra of DPM, CNZ, PVP, HPMC and their solid dispersions are shown in

416

Figure 7. The characteristic peaks of DPM are present at 1359 cm-1 (C – N bonds), 1533 cm-1

417

(C = N ring), 2851 cm-1 (symmetrical stretching of CH2 group), 2923 cm-1 (asymmetrical

418

stretching of CH2 group) and 3303/3377 cm-1 (OH stretching vibration). The CNZ spectra

419

displayed characteristic peaks at 963 cm-1 (aromatic out-of-plane), 1001 cm-1 (= C – H

420

alkene), 1141 cm-1 (C – N stretching peak), 2956 cm-1 (aliphatic CH stretching peak), 3021

421

cm-1 (alkene CH stretching) and 3066 cm-1 (aromatic CH stretching). The spectra of the

422

physical mixtures of DPM and CNZ with PVP, HPMC and surfactants are not shown here as

423

they did not show any apparent changes in the FTIR spectra. 14 ACS Paragon Plus Environment

Page 14 of 45

Page 15 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

424

The formation of DPM ASD with PVP, HPMC and PVP-HPMC resulted in the

425

disappearance of DPM OH stretching vibration indicating the formation of hydrogen bonds

426

with the polymers (Figure 7). Broadening (DPM-PVP, DPM-HPMC and DPM-PVP-HPMC)

427

and shifting (DPM-PVP-HPMC) of symmetrical (2851 cm-1) and asymmetrical (2923 cm-1)

428

stretching of DPM CH2 group was also observed suggesting drug-polymer-polymer

429

interaction was stronger compared to drug-polymer interaction. Shifts were also recorded for

430

C = N and C-N bands at 1533 and 1359 cm-1 in DPM-PVP-HPMC system. CNZ, on the other

431

hand, displayed no interaction with PVP but HPMC based CNZ dispersions demonstrated the

432

presence of drug-polymer interaction (shifts in CNZ C – N peak). These shifts and

433

broadening of peaks were observed in CNZ-PVP-HPMC dispersion which could be due to

434

the synergistic interaction between CNZ and PVP-HPMC combination. This helps explain

435

the better synergistic effect of PVP-HPMC on DPM compared to CNZ; both the PVP and the

436

HPMC interact with DPM individually whereas only the HPMC interacts with the CNZ. This

437

highlights the importance of the nature and strength of drug-polymer interactions while

438

choosing a polymer combination for stabilizing amorphous drug. Furthermore, on comparing

439

drug-PVP-HPMC spectra of DPM and CNZ with drug-PVP-HPMC-SDS displayed

440

significant differences with reference to peak broadening and shifts suggesting the presence

441

of surfactant affects drug-polymer-polymer interactions (Figure 7). It is important to mention

442

here that FTIR spectroscopic studies conducted here are only for qualitative purpose and do

443

not provide any quantitative understanding of the molecular interactions. Nevertheless, when

444

combined with the DSC results, they help build a better understanding of molecular

445

interactions.

446

3.3. Supersaturation generation and maintenance

447 448

3.3.1. Equilibrium solubility The effect of various compositions of physical mixtures (containing crystalline drug)

449

and solid dispersions (containing amorphous drug) on the equilibrium solubility of DPM and

450

CNZ in PBS 6.8 are reported in Table 2. For direct comparison, the composition of physical

451

mixtures has been kept the same as that of respective solid dispersion. The physical mixtures

452

of DPM and CNZ with polymers and/or surfactants did not result in a significant change in

453

the equilibrium solubility of crystalline DPM and CNZ. The reported CMC of poloxamer

454

188 and SDS are 0.743 mg/mL and 0.29 mg/mL.45 46 The presence of surfactants (P188 and

455

SDS) in physical mixtures with concentration below their respective critical micelle

456

concentration (CMC) in PBS 6.8 did not result in any significant rise in the solubility of 15 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 45

457

crystalline DPM and CNZ. Moreover, PVP, HPMC and their combination also have no effect

458

on the equilibrium solubility of the crystalline DPM and CNZ (Table 2). On the other hand,

459

the equilibrium solubility of amorphous drugs in the solid dispersions after 24 hr was much

460

higher than their physical mixture (crystalline drug) counterparts. Thus, the higher drug

461

concentration during dissolution was not due to the solubilising effect of polymers and

462

surfactant on the crystalline drug. It is due to crystallization inhibition efficiency of the

463

polymer in solution to maintain apparent higher solubility of amorphous drug. Furthermore,

464

the maintenance of amorphous DPM and CNZ supersaturation in PBS 6.8 was due to the

465

polymers rather than the surfactants. This was confirmed when the amorphous drug

466

equilibrium solubility from surfactant based ASDs (higher amorphous drug solubility) was

467

compared to their surfactant free counterpart (lower amorphous drug solubility).

468

maximum supersaturation concentration of DPM and CNZ was observed in drug-polymer-

469

polymer combinations which again suggests a synergistic effect of this polymer-combination

470

on drug supersaturation. By comparison, surfactants were found to have negative impact on

471

the amorphous drug supersaturation maintenance (Table 2) with lower concentrations

472

observed after 24 hrs for ASDs containing surfactants. The difference in drug concentration

473

was found to be significant for drug-polymer or drug-polymer-polymer systems compared to

474

surfactant incorporated dispersions.

475 476

3.3.2. In-vitro dissolution studies To further understand how the molecular interactions in the multi-component solid

477

dispersions affect the DPM and CNZ release profile from spray dried ASDs in pH 6.8 buffer,

478

in-vitro dissolution experiments were performed. Of particular interest was whether the

479

synergistic effect of the polymer combinations and the incorporation of surfactants into the

480

solid dispersions affected the generation and maintenance of DPM and CNZ supersaturation.

481

It is important to note that pH 6.8 buffer was chosen only to understand the effects of

482

excipients (drug-excipient interactions, antiplasticization etc.) on the apparent solubility of

483

amorphous drugs from solid dispersions. In order to isolate these effects we have chosen

484

simple discriminatory media. More complex simulated body fluids also contain surfactants

485

and other components which may directly affect the dissolution behaviour of the

486

formulations. Further experimental work would be required to study this in detail which is

487

outside the scope of this article.

The

488

The dissolution profile of spray dried dispersions of DPM in PBS 6.8 (Figure 8)

489

demonstrated that supersaturation was generated rapidly with a maximum concentration 16 ACS Paragon Plus Environment

Page 17 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

490

(Cmax) achieved within 5 min. Then, a rapid drop in the free DPM concentration was

491

observed, suggesting that water-mediated crystallization predominated in this stage. Once the

492

concentration of DPM reached a plateau, the dissolution process was in equilibrium with the

493

crystallization process. Significant differences were found between the Cmax for different

494

systems (PVP, HPMC and PVP-HPMC based dispersions). The descending order of Cmax is

495

DPM-PVP-HPMC > DPM-PVP > DPM-HPMC.

496

combination has a synergistic effect on generating and maintaining DPM supersaturation.

497

The presence of 5% w/w SDS or P188 in the formulations have variable effects on the

498

dissolution profile of DPM. SDS or P188 had a minimal effect on the dissolution profile of

499

DPM in PVP based dispersions (Figure 8 (a)). This suggests that the surfactants did not have

500

any significant influence on the DPM dissolution profile in PVP based dispersions. In DPM-

501

HPMC systems, both SDS and P188 had a significant effect on the supersaturation level of

502

DPM (Figure 8 (b)), adversely affecting the release of DPM as indicated by the lower Cmax

503

attained in the presence of SDS or P188. The level of supersaturation achieved and the

504

duration of supersaturation maintained was found to be maximum for DPM-HPMC system

505

followed by DPM-HPMC-P188 dispersion and the lowest for DPM-HPMC-SDS system.

506

These results suggest that the presence of the surfactant did not inhibit DPM crystallization

507

during dissolution to the same extent as their surfactant free counterparts in the HPMC-based

508

dispersions. Similar results have been observed for DPM-PVP-HPMC based systems (Figure

509

8 (c)) where the dissolution profile of the DPM dispersions without SDS or P188 was

510

significantly better than the dispersions with surfactants. The decreasing order of dissolution

511

performance is as follows: DPM-PVP-HPMC > DPM-PVP-HPMC-P188 > DPM-PVP-

512

HPMC-SDS.

Thus, it was found that PVP-HPMC

513

The effect of binary, ternary and quaternary combinations of drug, polymer and

514

surfactant on the dissolution profile of spray dried CNZ dispersions was also investigated.

515

The dissolution profiles of CNZ solid dispersions in PBS 6.8 containing a combination of

516

polymer and surfactants are illustrated in Figure 9. CNZ, in PVP and PVP-HPMC based

517

formulations, achieved a Cmax within 5 min followed by a decline in free CNZ concentration,

518

indicating that CNZ crystallizes quickly in the presence of water. This could be due to strong

519

plasticization effect of water on the amorphous drug. The CNZ-HPMC system, on the other

520

hand, achieved a maximum concentration after 120 mins followed by decline in free drug

521

concentration. The dissolution of ternary CNZ-PVP-HPMC resulted in significantly higher

17 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

522

CNZ concentration in solution compared to CNZ-PVP and CNZ-HPMC binary dispersions.

523

The effect of surfactants on CNZ dissolution is similar to that of DPM dispersions.

524

The dissolution results suggests that the surfactants interfere with the crystallization

525

inhibitory efficiency of the polymers. The effect of SDS and P188 on HPMC based

526

dispersions of DPM and CNZ is more pronounced than in the PVP based systems. The effect

527

is even more pronounced in PVP-HPMC based dispersions of DPM and CNZ. There are two

528

possible explanations for the lower drug supersaturation in the presence of surfactants. The

529

first possibility is that the surfactant and polymer molecules could be competing to interact

530

with the drug molecules.41 Due to the hydrophobic nature of DPM and CNZ, there was a

531

higher tendency for the hydrophobic parts of SDS and P188 molecules to interact to DPM or

532

CNZ molecules than for the hydrophilic HPMC or PVP-HPMC polymers. The adsorption of

533

surfactant molecules on to the drug molecules would lead to a reduced interfacial tension

534

which eventually reduces the nucleation activation energy, finally leading to a higher

535

nucleation rate. 9, 20, 41 Another possible explanation for the reduced supersaturation level is

536

that the polymer and surfactant molecules interact to form clusters and therefore the number

537

of freely available polymer molecules to inhibit the crystallization of dissolved DPM and

538

CNZ molecules is reduced.47,

539

performance of spray dried solid dispersion formulations may be counterproductive.

540 541

3.3.3. Solution 1H NMR studies Solution 1H NMR spectroscopic studies were performed to probe the molecular

542

mechanism behind the improved supersaturation performance of ternary drug-polymer-

543

polymer ASDs of DPM and CNZ. This study was performed on the best performing ASDs

544

(generating and maintaining highest level of supersaturation) screened during the dissolution

545

studies and compared to the binary drug-polymer ASDs. This study was performed to study

546

the site of interaction between the drug and polymer in solution to help explain the results

547

obtained in the dissolution studies. The peak shifts in 1H NMR spectra of DPM, CNZ and

548

their binary and ternary ASDs (without surfactants) are shown in Figure 10 and 11. The full

549

NMR spectra of drug, polymer and their solid dispersions are provided in supplementary

550

information.

48

Thus, the use of surfactant to enhance the dissolution

551

As shown in Figure 10, the chemical shift of protons (1) and (3) shifted downfield to

552

higher values in all the systems. This suggests the deshielding of the protons in the presence

553

of polymers. A chemical shift for proton (2) and (4) was not observed in DPM-PVP system

554

whereas a deshielding effect was observed in the DPM-HPMC and DPM-PVP-HPMC 18 ACS Paragon Plus Environment

Page 18 of 45

Page 19 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

555

systems. In case of CNZ, the shifts were most pronounced for protons (1), (2), (6) and (7) of

556

CNZ in PVP dispersions and for protons (1), (2), (3), (4), (5), (6) and (7) in HPMC and PVP-

557

HPMC systems (Figure 11). The shifts shown in Figure 10 and 11 suggests DPM and CNZ

558

interact with PVP, HPMC and PVP-HPMC systems in a different way which goes some way

559

towards explaining the different polymeric effect, particularly in terms of inhibition of drug

560

recrystallization during dissolution of DPM and CNZ ASDS. The larger peak shifts

561

corresponds to the possibility of stronger drug-polymer interaction in ternary system

562

compared to individual PVP and HPMC binary systems. The significantly higher shifts in

563

ternary systems again suggests a synergistic effect of polymer on the drug-polymer

564

interaction in solution. The relative extent of shift, in the decreasing order, was drug-PVP-

565

HPMC > drug-HPMC > drug-PVP (for both the drugs) which explains the dissolution results

566

where the ternary drug-PVP-HPMC generated and maintained highest level of

567

supersaturation compared to other systems studied. This also matches the rank order observed

568

for MPD analysis.

569

It is interesting to observe that the thermal characterization results (Figure 4, 5 and 6)

570

have been negatively affected by the presence of SDS and P188. Also, as shown in Figure 8

571

and 9, the increased crystallization rate of amorphous drugs during dissolution incorporating

572

surfactants could be attributed to their higher affinity for water that could act as plasticizer.

573

The negative impact of surfactant is believed to be due to reduction in glass transition

574

temperature (Table 2) and also may be due to molecular effects possibly promoting polymer

575

chain entanglements and polymeric globule formation.51 The globules formed in the spray

576

drying solution seemed to be present in the solid state, and therefore, diffusion of the drug is

577

reduced. This indicates the presence of a memory where incorporation of surfactant in the

578

liquid state (spray drying solution) in polymer chain remained even after the solid particles

579

have had formed.

580

Towards this end, we have conducted 2D-NOESY experiments to understand the

581

surfactant-polymer interaction in the solution state (spray drying drug-polymer-surfactant

582

solution and also during dissolution). It provides information on the distance between protons

583

that are spatially close to each other.52 We observe the difference in 2D-NOESY spectra of

584

polymer compared to the spectra of polymer-surfactant combination (see supplementary

585

information). However, due to very small concentration of surfactant in the solution (to keep

586

the surfactant concentration same as used for dissolution experiments) the magnitude of

587

differences observed are very small. Similar effect of surfactant on PVP and HPMC have 19 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 45

588

been reported previously.52 53 54 Nevertheless, these considerations are in agreement with the

589

results obtained from solid-state thermal analysis and in-vitro dissolution which suggests that

590

surfactant is competing with the drug molecules for polymer functional groups and thereby

591

reducing the crystallization inhibition efficiency of the polymer.

592 593

3.4. Moisture sorption analysis and Stability studies To further understand the behaviour of multi-component (binary, ternary and

594

quaternary) ASDs in the presence of moisture under stress conditions, the freshly prepared

595

spray dried dispersions were exposed to DVS analysis using double ramp method from 0-90-

596

0% RH (2 cycles) in 10% increments (dm/dt = 0.001 at each step) at 40 °C. As can be seen in

597

Figure 12, the amount of water uptake was greater in PVP based dispersions compared to

598

HPMC or PVP-HPMC based systems. This could be due to the more hydrophilic nature of

599

PVP compared to HPMC. Furthermore, the effect of surfactants on the moisture sorption

600

profile was greater in PVP based dispersions compared to HPMC or PVP-HPMC based

601

systems (Figure 12). Also, the time required to complete the double sorption-desorption cycle

602

varied significantly in different systems which suggests incorporating surfactant within the

603

dispersion significantly affects the moisture sorption ability of drug-polymer or drug-

604

polymer-polymer systems. This could be due to differences in the surface energy and

605

relaxation of amorphous API at the surface.53 It has been reported that structural relaxation at

606

the surface can vary significantly from relaxation in the bulk of amorphous materials.54

607

Surfactants get localised at the surface of the particles, affecting the relaxation energy and

608

consequently the crystallization rate.51 Furthermore, the Tg of the systems decreased when the

609

surfactants were incorporated within the dispersions (freshly prepared) which could suggest a

610

plasticization effect (refer supplementary information for

611

Crystallization was evident (post DVS cycles, Figure 12) in DPM-PVP-SDS, DPM-HPMC,

612

DPM-HPMC-SDS and DPM-HPMC-P188 dispersions signifying altered drug-polymer

613

interaction due to moisture sorption and surfactant incorporation. Other dispersions showed

614

no signs of crystallization. Similar results were observed for CNZ ASDs (data not shown).

615

These observations are in accordance with the previous DSC analysis and dissolution results

616

of freshly prepared dispersions.

mDSC thermograms).

617

We have previously reported that the amorphous forms of pure DPM and CNZ tend to

618

recrystallize rapidly in the absence of stabilizing polymers.11 We have also shown that

619

recrystallization of amorphous drug is delayed within ASDs.19 In this study, the physical

620

stability of multi-component ASDs was assessed under storage conditions of 40 °C/75% RH 20 ACS Paragon Plus Environment

Page 21 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

621

for 4 weeks (Figure 13, 14 and 15). The melting enthalpies of the crystallized samples are

622

provided in Table 2. The full DSC thermograms of stability samples are provided in

623

supplementary information. As shown in Figure 13, DPM-PVP systems were found to be

624

completely amorphous after 4 weeks, both with and without surfactant. However,

625

amorphous-amorphous phase separation was observed in DPM-PVP, DPM-PVP SDS and

626

DPM-PVP-P188 systems. All HPMC based DPM dispersions (Figure 14 and Table 2)

627

underwent crystallization suggesting PVP is better in maintaining solid-state stability of

628

amorphous DPM compared to HPMC. The solid dispersions of ternary DPM-PVP-HPMC

629

remained amorphous with significantly higher Tg values compared to binary dispersions

630

(Figure 15 and Table 2). This suggests a synergistic effect of polymer combination in

631

improving the solid-state stability of amorphous DPM. Further evidence of the synergistic

632

effect of PVP-HPMC was observed in case of CNZ dispersions where all CNZ systems

633

showed crystallization after 4 weeks of storage at 40 °C/75 % RH except CNZ-PVP-HPMC

634

dispersion (Figure 13, 14, 15 and Table 2).

635

Similar observations have been made in the past where surfactants had a negative

636

impact on the solid dispersion performance.51 Difference in the physicochemical properties of

637

SDS and P188 suggest that these events are controlled by the type of surfactant used. Thus, in

638

this study, we observed surfactants (when they are incorporated within the dispersions)

639

neither improved the solid state stability (Figure 4, 5, 6, 13, 14, 15 and Table 2) nor increased

640

the supersaturation level during dissolution (Figure 8 and 9). We found, in this case, that

641

drug-excipient and excipient-excipient interaction could have considerable impact on the

642

crystallization of amorphous API. Different systems behaved differently when observed

643

under different conditions (DSC, dissolution, DVS or stability). Further investigations are

644

required to confirm the localization of surfactants within spray dried particles to have a better

645

understanding of how surfactants affect drug-polymer interactions. Nevertheless, the use of

646

surfactants (as solubilizing or wetting agents) in spray dried ASDs requires prudent

647

consideration.

648

On the other hand, drug-polymer-polymer ternary ASDs showed higher stability and

649

dissolution compared with binary ASDs. PVP K30 and HPMC K100 complemented each

650

other to provide enhanced stability as well as improved drug supersaturation. The synergistic

651

efficiency of these polymers when used in combination in ternary ASDs could be due to

652

greater DPM and CNZ miscibility in the ternary systems compared with the binary system as

653

confirmed by MPD analysis and also stronger drug-polymer interaction in the ternary 21 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

654

dispersions, as confirmed by Tg measurements (Table 2) and FTIR results. In solution, the

655

presence of molecular interactions between the drug and polymer is responsible for

656

generating and maintaining drug supersaturation for a prolonged period of time (Figure 10

657

and 11). Similarly, crystallization of amorphous drug in solid state requires favourable

658

orientation for the crystal nucleation. It has been confirmed that surface crystallization occurs

659

at relatively faster rate compared to bulk crystallization.25 In ternary systems, the presence of

660

two polymers might increase overall entropy of the system by providing a barrier for

661

crystallization in surface and bulk. This is confirmed by the stability studies of CNZ-PVP-

662

HPMC system (Figure 15 and Table 2) which showed it was stable after exposure to moisture

663

for 4 weeks whereas all other CNZ system showed drug crystallization. Ternary dispersions

664

are complicated systems where weak drug-polymer interactions are involved in the

665

amorphous drug solubilisation and stabilization. Further molecular level investigations are

666

required to obtain a mechanistic understanding of the synergistic effects reported in this

667

study. Future work should investigate the interactions between additives and drug nuclei or

668

crystal surfaces, which had previously been reported as a critical factor of crystallization

669

inhibition or morphology modification.41 Also, more studies are required to examine more

670

drug-polymer combinations or the combinations of different additives including other

671

surfactants at different concentrations for synergistic effects using advanced characterization

672

techniques such as solid-state NMR, dielectric spectroscopy, inverse gas chromatography,

673

nucleation induction studies in the presence of moisture or during dissolution, and others.

674

These studies are currently underway in our lab and will form the basis of next article.

675

5. CONCLUSION

676

The incorporation of surfactants and polymer combinations seems to have a

677

significant effect on the properties of the resulting ASDs. Surfactants, when incorporated in

678

ASDs, resulted in altered dissolution rates, reduced stability and altered drug-polymer

679

interactions. The combination of PVP K30 and HPMC K100, two polymers which displayed

680

drug-polymer interaction, resulted in significant crystallization inhibition (both in solid-state

681

and dissolution) of the poorly soluble drugs DPM and CNZ. This enhanced crystallization

682

inhibition efficiency can be correlated to the synergistic effect of PVP K30 and HPMC K100.

683

This study also highlights the importance of utilizing ternary drug-polymer-polymer

684

interactions by combining polymers with different crystallization inhibition mechanisms for

685

improved solid-state stability and dissolution enhancement of poorly soluble drugs.

22 ACS Paragon Plus Environment

Page 22 of 45

Page 23 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

686 687

688

Molecular Pharmaceutics

NOTE The authors declare no competing financial interest.

ACKNOWLEDGEMENT

689

This publication has emanated from research conducted with the financial support of

690

the Synthesis and Solid State Pharmaceutical Centre, funded by Science Foundation Ireland

691

under Grant Numbers (SB, grant 12/RC/2275) as well as support from Waterford Institute of

692

Technology.

693 694 695

Supporting Information:

696

MDSC thermograms (fresh, stability and post-DVS samples), full NMR spectra and XRD

697

spectra have been provided in the supporting information.

698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714

23 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

715

REFERENCES

716

1. Babu, N. J.; Nangia, A., Solubility advantage of amorphous drug and pharmaceutical

717

cocrystals. Crystal Growth and Design 2011, 11, 2662-2679.

718

2. Jornada, D. H.; dos Santos Fernandes, G. F.; Chiba, D. E.; de Melo, T. R.; dos Santos, J.

719

L.; Chung, M. C., The prodrug approach: A successful tool for improving drug solubility.

720

Molecules 2015, 21, 1-31.

721

3. Liu, Y.; Wang, L.; Zhao, Y.; He, M.; Zhang, X.; Niu, M.; Feng, N., Nanostructured lipid

722

carriers versus microemulsions for delivery of the poorly water-soluble drug-luteolin.

723

International Journal of Pharmaceutics 2014, 476, 169-177.

724

4. Karashima, M.; Sano, N.; Yamamoto, S.; Arai, Y.; Yamamoto, K.; Amano, N.; Ikeda, Y.,

725

Enhanced pulmonary absorption of poorly soluble itraconazole by micronized cocrystal dry

726

powder formulation. European Journal of Pharmaceutics and Biopharmaceutics 2017, 115,

727

65-72.

728

5. Buckley, S. T.; Frank, K. J.; Fricker, G.; Brandl, M., Biopharmaceutical classification of

729

poorly soluble drugs with respect to “enabling formulations”. European Journal of

730

Pharmaceutical Science 2013, 50, 8-16.

731

6. Samie, A.; Desiraju, G. R.; Banik, M., Salts and cocrystals of the antidiabetic drugs

732

gliclazide tolbutamide and glipizide: Solubility enhancements through drug-coformer

733

interactions. Crystal Growth and Design 2017, 17, 2406-2417.

734

7. Sun, D. D.; Lee, P. I., Probing the mechanism of drug release from amorphous solid

735

dispersions in medium-soluble and medium-insoluble carriers. Journal of Controlled Release

736

2015, 211, 85-93.

737

8. Shah, T.; Laaksonen, T.; Rades, T.; Aaltonen, J.; Peltonen, L.; Strachan, C. J., Unravelling

738

the relationship between degree of disorder and the dissolution behaviour of milled

739

glibenclamide. Molecular Pharmaceutics 2013, 11, 234-242.

740

9. Baghel, S.; Cathcart, H.; O’Reilly, N. J., Polymeric amorphous solid dispersions: A review

741

of amorphization, crystallization, stabilization, solid-state characterization and aqueous

742

solubilisation of biopharmaceutical classification system class II drugs. Journal of

743

Pharmaceutical Sciences 2016, 105, 2527-2544.

744

10. Brouwers, J.; Brewster, M. E.; Augustijns, P., Supersaturating drug delivery systems: the

745

answer to solubility-limited oral bioavailability? Journal of Pharmaceutical Sciences 2009,

746

98, 2549-2572.

747

11. Baghel, S.; Cathcart, H.; Redington, W.; O’Reilly, N. J., An investigation into the

748

crystallization tendency/kinetics of amorphous active pharmaceutical ingredients: A case 24 ACS Paragon Plus Environment

Page 24 of 45

Page 25 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

749

study with dipyridamole and cinnarizine. European Journal of Pharmaceutics and

750

Biopharmaceutics 2016, 104, 59-71.

751

12. Amstad, E.; Spaepen, F.; Weitz, D. A., Stabilization of the amorphous structure of spray-

752

dried drug nanoparticles. Journal of Physical Chemistry B 2016, 120, 9161-9165.

753

13. Xie, T.; Taylor, L. S., Dissolution performance of high drug loading celecoxib amorphous

754

solid dispersions formulated with polymer combinations. Pharmaceutical Research 2016, 33,

755

739-750.

756

14. Baghel, S.; Cathcart, H.; O’Reilly, N. J., Understanding the generation and maintenance

757

of supersaturation during the dissolution of amorphous solid dispersions using modulated

758

DSC and 1H NMR. International Journal of Pharmaceutics 2018, 536, 414-425.

759

15. Murdande, S. B.; Pikal, M. J.; Shanker, R. M.; Bogner, R. H., Solubility advantage of

760

amorphous pharmaceuticals, Part 3: I maximum solubility advantage experimentally

761

attainable and sustainable? Journal of Pharmaceutical Sciences 2011, 100, 4349-4356.

762

16. Kalepu, S.; Nekkanti, V., Insoluble drug delivery strategies: review of recent advances

763

and business prospects. Acta Pharma Sinica B 2015, 5, 442-453.

764

17. Vasconcelos, T.; Marques, S.; das Neves, J.; Sarmento, B., Amorphous solid dispersions:

765

Rational selection of manufacturing process. Advanced Drug Delivery Reviews 2016, 100,

766

85-101.

767

18. Vo, C. L.; Park, C.; Lee, B. J., Current trends and future perspective of solid dispersions

768

containing poorly water-soluble drugs. European Journal of Pharmaceutics and

769

Biopharmaceutics 2013, 85, 799-813.

770

19. Baghel, S.; Cathcart, H.; O’Reilly, N. J., Theoretical and experimental investigation of

771

drug-polymer interaction and miscibility and its impact on drug supersaturation in aqueous

772

medium. European Journal of Pharmaceutics and Biopharmaceutics 2016, 107, 16-31.

773

20. Yoo, S.; Krill, S. L.; Wang, Z.; Telang, C., Miscibility/stability consideration in binary

774

solid dispersion systems composed of functional excipients towards the design of multi-

775

component amorphous systems. Journal of Pharmaceutical Sciences 2009, 98, 4711-4723.

776

21. van den Mooter, M. G.; Wuyts, M.; Blaton, N.; Busson, R.; Grobet, P.; Augustijns,

777

Kinget, R., Physical stabilization of amorphous ketoconazole in solid dispersions with

778

polyvinyl pyrrolidone K25. European Journal of Pharmaceutical Sciences 2001, 12, 261-

779

269.

780

22. Damian, F.; Blaton, N.; Kinget, R.; van den Mooter, M. G., Physical stability of solid

781

dispersions of the antiviral agent UC-781 with PEG 6000, Gelucire 44/14 and PVP K30.

782

International Journal of Pharmaceutics 2002, 244, 87-89. 25 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

783

23. Wang, X.; Michoel, A.; van den Mooter, M. G., Solid state characteristics of ternary solid

784

dispersions composed of PVP VA64, Myrj 52 and itraconazole. International Journal of

785

Pharmaceutics 2005, 303, 54-61.

786

24. Karavas, E.; Geogarakis, E.; Sigalas, M. P.; Avgoustakis, K.; Bikiaris, D., Investigation

787

of the release mechanism of a sparingly water-soluble drug from solid dispersions in

788

hydrophilic carriers based on physical state of drug, particle size distribution and drug-

789

polymer interactions. European Journal of Pharmaceutics and Biopharmaceutics 2007, 66,

790

334-347.

791

25. Prasad, D.; Chauhan, H.; Atef, E., Amorphous stabilization and dissolution enhancement

792

of amorphous ternary solid dispersions: Combination of polymers showing drug-polymer

793

interaction for synergistic effects. International Journal of Pharmaceutical Science 2014,

794

103, 3511-3523.

795

26. Ghebremeskel, A. N.; Vemavarapu, C.; Lodaya, M., Use of surfactant as plasticizers in

796

preparing solid dispersions of poorly soluble API: Selection of polymer-surfactant

797

combinations using solubility parameters and testing the processability. International Journal

798

of Pharmaceutics 2007, 328, 119-129.

799

27. Ghebremeskel, A. N.; Vemavarapu, C.; Lodaya, M., Use of surfactants as plasticizers in

800

preparing solid dispersions of poorly soluble API: stability testing of selected solid

801

dispersions. Pharmaceutical Research 2006, 23, 1928-1936.

802

28. Jain, S. K.; Shukla, M.; Shrivastava, V., Development and in vitro evaluation of

803

ibuprofen mouth dissolving tablets using solid dispersion technique. Chemical and

804

Pharmaceutical Bulletin 2010, 58, 1037-1042.

805

29. Agrawal, A. M.; Dudhedia, M. S.; Zimny, E., Hot melt extrusion: Development of an

806

amorphous solid dispersion for an insoluble drug from mini-scale to clinical scale. AAPS

807

PharmSciTech 2016, 17, 133-147.

808

30. Ghebremeskel, A. N.; Vemavarapu, C.; Lodaya, M., Use of surfactant as plasticizers in

809

preparing solid dispersions of poorly soluble API: Selection of polymer-surfactant

810

combinations using solubility parameters and testing the processability. International Journal

811

of Pharmaceutics 2007, 328, 119-129.

812

31. Paudal, A.; Worku, Z. A.; Meeus, J.; Guns, S.; van den Mooter, G., Manufacturing of

813

solid dispersions of poorly water soluble drugs by spray drying: Formulation and process

814

consideration. International Journal of Pharmaceutics 2013, 453, 253-284.

26 ACS Paragon Plus Environment

Page 26 of 45

Page 27 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

815

32. Al-Obaidi, H.; Lawrence, M. J.; Buckton, G., Atypical effects of incorporated surfactants

816

on stability and dissolution properties of amorphous polymeric dispersions. Journal of

817

Pharmacy and Pharmacology 2016, 68, 1373-1383.

818

33. Rashid, R.; Kim, D. W.; ud Din, F.; Mostapha, O.; Yousaf, A. M.; Park, J. H.; Kim, J. O.;

819

Yong, C. S.; Choi, H. G., Effect of hydroxypropylcellulose and Tween80 on the

820

physicochemical properties and bioavailability of ezetimibe-loaded solid dispersions.

821

Carbohydrate Polymers 2015, 130, 26-31.

822

34. Mah, P. T.; Peltonen, L.; Novakovic, D.; Rades, T.; Strachan, C. J.; Laaksonen, T., The

823

effect of surfactants on the dissolution behaviour of amorphous formulations. European

824

Journal of Pharmaceutics and Biopharmaceutics 2016, 103, 13-22.

825

35. Vora, A.; Patadia, R.; Mittal, K.; Mashru, R., Preparation and characterization of

826

dipyridamole solid dispersions for stabilization of supersaturation: effect of precipitation

827

inhibitors type and molecular weight. Pharmaceutical Development and Technology 2015, 3,

828

1-9.

829

36. http://www.drugbank.ca/drugs/DB00568

830

37. Zheng, X.; Yang, R.; Tang, X.; Zheng, L., Part 1: Characterization of solid dispersions of

831

nimodipine prepared by hot-melt extrusion. Drug Development and Industrial Pharmacy

832

2007, 33, 791-802.

833

37. Dubin, P. L.; Li, Y., Structure and flow in surfactant solutions: Polymer-surfactant

834

complexes. ACS Symposium Series 1994, 578, 320-336.

835

38. Yu, X.; Somasundaran, P., Role of polymer conformation in interparticle-bridging

836

dominated flocculation. Journal of Colloid and Interface Science 1996, 177, 283-287.

837

39. Margolin, A.; Sherstyuk, S. F.; Izumdov, V. A.; Zexin, A. B.; Kabanov, V. A., Enzymes

838

in polyelectrolyte complexes. The effect of phase transition on thermal stability. European

839

Journal of Biochemistry 1985, 146, 625-632.

840

40. Brackman, J. C., Sodium dodecyl sulfate-induced enhancement of the viscosity and

841

viscoelasticity of aqueous solutions of poly (ethylene oxide), A rheological study on

842

polymer-micelle interaction. Langmuir 1991, 7, 469-472.

843

41. Ileebare, G. A.; Liu, H.; Edgar, K. J.; Taylor, L. S., Effect of binary additive combination

844

on solution crystal growth of the poorly water-soluble drug, ritonavir. Crystal Growth 2012,

845

12, 6050-6060.

846

42. Deshpande, T. M.; Shi, H.; Pietryka, J.; Hoag, S. W.; Medek, A., Investigation of

847

polymer-surfactant interactions and their impact on itraconazole solubility and precipitation

27 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

848

kinetics for developing spray dried amorphous solid dispersions. Molecular Pharmaceutics

849

2018, doi: 10.1021/acs.molpharmaceut.7b00902

850

43. Zheng, X.; Yang, R.; Tang, X.; Zheng, L., Part 1: Characterization of solid dispersions of

851

nimodipine prepared by hot-melt extrusion. Drug Development and Industrial Pharmacy

852

2007, 33, 791-802.

853

44. Marsac, P. J.; Shamblin, S. L.; Taylor, L. S., Theoretical and practical approaches for

854

prediction of drug-polymer miscibility and solubility. Pharmaceutical Research 2010, 23,

855

2417-2426.

856

45. Tian, J.; Zhao, Y. Z.; Jin, Z.; Lu, C. T.; Tang, Q. Q.; Xiang, Q.; Sun, C. Z.; Zhang, L.;

857

Xu, Y. Y.; Gao, H. S.; Zhou, Z. C.; Li, X. K.; Zhang, Y., Synthesis and characterization of

858

poloxamer 188-grafted heparin copolymer. Drug Development and Industrial Pharmacy

859

2010, 36, 832-838.

860

46. Khan, A. M.; Shah, S. S., Determination of critical micelle concentration (CMC) of

861

sodium dodecyl sulfate (SDS) and the effect of low concentration of pyrene on its cmc using

862

ORIGIN software. Journal of Chemical Society of Pakistan 2008, 30, 186-191.

863

47. De Martins, R.; Da Silva, C.; Becker, C.; Samios, D.; Christoff, M.; Bica, C. D.,

864

Interaction of (hydroxypropyl) cellulose with anionic surfactants in dilute regime. Colloid

865

and Polymer Science 2006, 284, 1353-1361.

866

48. Wesley, R. D.; Cosgrove, T.; Thompson, L.; Armes, S. P.; Baines, F. L., Structure of

867

polymer/surfactant complexes formed by poly(2-(dimethylamino)ethyl methacrylate) and

868

sodium dodecyl sulphate. Langmuir 2002, 18, 5704-5707.

869

49. http://sdbs.db.aist.go.jp/sdbs/cgi-bin/cre_index.cgi

870

50. Sassene, P. J.; Mosgaard, M. D.; Loebmann, K.; Mu, H.; Larsen, F. H.; Rades, T.;

871

Mullertz, A., Elucidating the molecular interactions occurring during drug precipitation of

872

weak bases from lipid based formulations – a case study with cinnarizine and a long chain

873

self nano-emulsifying drug delivery system. Molecular Pharmaceutics 2015, 12, 4067-4076.

874

51. Al-Obaidi, H.; Lawremce, M. J.; Buckton, G., Atypical effects of incorporated surfactants

875

on the dissolution properties of amorphous polymeric dispersions. Journal of Pharmacy and

876

Pharmacology 2016, 68, 1373-1383.

877

52. Chen, Y.; Wang, S.; Wang, S.; Liu, C.; Su, C.; Hageman, M.; Haskell, R.; Stefanski,

878

K.; Qian, F., Sodium lauryl sulphate competitively interacts with HPMC-AS and

879

consequently reduces oral bioavailability of Posaconazole/HPMC-AS amorphous solid

880

dispersion. Molecular Pharmaceutics 2016, 13, 2787-2795.

28 ACS Paragon Plus Environment

Page 28 of 45

Page 29 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

881

53. Hasegawa, S.; Ke, P.; Buckton, G., Determination of the structural relaxation at the

882

surface of amorphous solid dispersions using inverse gas chromatography. Journal of

883

Pharmaceutical Sciences 2009, 98, 2133-2139.

884

54. Ke, P.; Hasegawa, S.; Al-Obaidi, H.; Buckton, G., Investigation of preparation methods

885

on surface/bulk structural relaxation and glass fragility of amorphous solid dispersions.

886

International Journal of Pharmaceutics 2012, 422, 170-178.

29 ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. Chemical structure of DPM, CNZ, PVP K30, SDS, P188 and HPMC K100 (clockwise from top) 270x148mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 30 of 45

Page 31 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 2. DSC thermograms of physical mixtures of model drugs and polymers at various drug loads; n = 3. Values in parentheses are % weight fraction. 215x140mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. Experimental and theoretical depression in melting point of model drugs at various %w/w drug loading. 238x83mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 32 of 45

Page 33 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 4. mDSC thermograms of freshly prepared DPM and CNZ ASDs with PVP; n = 3; Full DSC thermograms are provided in the supplementary information. 267x132mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5. mDSC thermograms of freshly prepared DPM and CNZ ASDs with HPMC; n = 3; Full DSC thermograms are provided in the supplementary information. 262x138mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 34 of 45

Page 35 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 6. mDSC thermograms of freshly prepared DPM and CNZ ASDs with PVP-HPMC; n = 3; Full DSC thermograms are provided in the supplementary information. 261x138mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 7. FTIR spectra of freshly prepared DPM ASDs 213x141mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 36 of 45

Page 37 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 8. In-vitro dissolution profile of DPM ASDs in PBS 6.8 at 37°C with PVP (a), HPMC (b) and PVP-HPMC (c); samples equivalent to 25 mg of drug was added to PBS 6.8; Each point represents mean ± SD; n = 2 267x124mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9. In-vitro dissolution profile of CNZ ASDs in PBS 6.8 at 37°C with PVP (a), HPMC (b) and PVP-HPMC (c); samples equivalent to 25 mg of drug was added to PBS 6.8; Each point represents mean ± SD; n = 2 264x123mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 38 of 45

Page 39 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 10. Solution 1H NMR spectra of DPM (1000 µg/mL) in DMSO-d6. Figures in inset represent the chemical shift of DPM (a), DPM-PVP (b), DPM-HPMC (c), and DPM-PVP-HPMC (d). The numbers at top of each inset figure represent the peak number in DPM spectra. Assignments of DPM resonance is based on spectral database [49]. Numbers are assigned arbitrarily 253x142mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 11. Solution 1H NMR spectra of CNZ (1000 µg/mL) in DMSO-d6. Figures in inset represent the chemical shift of CNZ (a), CNZ-PVP (b), CNZ-HPMC (c), and CNZ-PVP-HPMC (d). The numbers at top of each inset figure represent the peak number in CNZ spectra. Assignments of CNZ resonance is based on Sassene et al.[50]. Numbers are assigned arbitrarily 234x144mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 40 of 45

Page 41 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 12. DVS analysis of freshly prepared DPM ASDs with PVP (a), HPMC (b) and PVP-HPMC (c) using double ramp method from 0-90-0% RH (2 cycles) in 10% increment (dm/dt=0.001 at each step) at 40 °C; DSC thermograms (d) of DPM ASDs post DVS double cycle 212x141mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 13. mDSC thermograms of stressed DPM and CNZ ASDs with PVP; n = 3; Full DSC thermograms are provided in the supplementary information 254x139mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 42 of 45

Page 43 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Figure 14. mDSC thermograms of stressed DPM and CNZ ASDs with HPMC; n = 3; Full DSC thermograms are provided in the supplementary information 256x140mm (96 x 96 DPI)

ACS Paragon Plus Environment

Molecular Pharmaceutics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 15. mDSC thermograms of stressed DPM and CNZ ASDs with PVP-HPMC; n = 3; Full DSC thermograms are provided in the supplementary information 252x137mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 44 of 45

Page 45 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Molecular Pharmaceutics

Graphical Abstract 211x141mm (96 x 96 DPI)

ACS Paragon Plus Environment