Investigation of Drug–Excipient Interactions in ... - ACS Publications

ACS2GO © 2019. ← → → ←. loading. To add this web app to the home screen open the browser option menu and tap on Add to homescreen...
0 downloads 0 Views 4MB Size
Subscriber access provided by READING UNIV

Article

Investigation of Drug-Excipient Interactions in Biclotymol Amorphous Solid Dispersions Benjamin Schammé, Nicolas Couvrat, Vincent Tognetti, Laurent Delbreilh, Valérie Dupray, Eric Dargent, and Gérard Coquerel Mol. Pharmaceutics, Just Accepted Manuscript • DOI: 10.1021/acs.molpharmaceut.7b00993 • Publication Date (Web): 12 Jan 2018 Downloaded from http://pubs.acs.org on January 14, 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 free 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 accessible to all readers and 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.

Molecular Pharmaceutics 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 50 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

Investigation of Drug-Excipient Interactions in

2

Biclotymol Amorphous Solid Dispersions

3

Benjamin Schammé † §, Nicolas Couvrat †, Vincent Tognetti ‡, Laurent Delbreilh § *, Valérie

4

Dupray † *, Éric Dargent §, Gérard Coquerel †

5

† Normandie Univ, UNIROUEN, Sciences et Méthodes Séparatives, 76000, Rouen, France

6

§ Normandie Univ, UNIROUEN, INSA Rouen, CNRS, Groupe de Physique des Matériaux,

7

76000 Rouen, France

8

‡ Normandie Univ, UNIROUEN, INSA Rouen, CNRS, COBRA UMR 6014, 76821 Mont-Saint-

9

Aignan, France

10 11

* Corresponding authors: [email protected]. Tel: +33 2 32 39 90 82 (V.D.) [email protected]. Tel: +33 2 32 95 50 84 (L.D.)

12 13

ACS Paragon Plus Environment

1

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

14

Page 2 of 50

Table of Contents Graphic

15 16

Abstract

17

The effect of low molecular weight excipients on drug-excipient interactions, molecular mobility

18

and propensity to recrystallization of an amorphous active pharmaceutical ingredient is

19

investigated.

20

Pentaacetylglucose), five different drug:excipient ratios (1:5, 1:2, 1:1, 2:1 and 5:1, w/w) and

21

three different solid state characterization tools (Differential Scanning Calorimetry, X-Ray

22

Powder Diffraction and Dielectric Relaxation Spectroscopy) were selected for the present

23

research. Our investigation has shown that the excipient concentration likewise with its

24

molecular structure reveal quasi-identical molecular dynamic behavior of solid dispersions

25

above and below the glass transition temperature. Across to complementary quantum

26

mechanical simulations, we pointed out a clear indication of a strong interaction between

27

Biclotymol and the acetylated saccharides. Moreover, the thermodynamic study on these

28

amorphous solid dispersions highlighted a stabilizing effect of α-Pentaacetylglucose regardless

29

its quantity while an excessive concentration of β-Pentaacetylglucose reveal a poor

30

crystallization inhibition. Finally, through long-term stability studies, we also showed the

31

limiting excipient concentration needed to stabilize our amorphous API. Herewith, the developed

32

procedure in this paper appears to be a promising tool for solid-state characterization of

33

complex pharmaceutical formulations.

Two

structurally

related

excipients

(α-Pentaacetylglucose

and

β-

34 35

Keywords

36

Amorphous Solid Dispersion; Molecular Mobility; Dielectric Spectroscopy; Physical Stability;

37

Biclotymol; Excipients; Density Functional Theory

38

ACS Paragon Plus Environment

2

Page 3 of 50 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

39

Introduction

40

Formulation procedures involve the presence of crystalline matter, in particular in the pharmaceutical

41

industry.1 Active substances currently developed have become more and more complex, with a

42

decreasing solubility. Indeed, over half of potential new pharmaceuticals exhibits poor water

43

solubility leading to a limited bioavailability.2 As this proportion is likely to increase, improving the

44

biodisponibility of these new drugs presents a significant technological challenge for pharmaceutical

45

scientists.3–6 Numerous strategies have been devised and put forward in order to transform a

46

crystalline drug into a more soluble counterpart.7,8 Among them, formulation of amorphous solids is

47

a promising one.9–12 In opposition to the three-dimensional order of crystalline matter, amorphous

48

solids are the most energetic solid state of a material and result in higher dissolution rate.13 However,

49

the higher free energy of amorphous compounds is accountable for inherent thermodynamic

50

instability14 including structural relaxation and nucleation/crystal growth that may arise during

51

storage or dissolution in the human organism.15,16 Indeed, during their storage, amorphous states (as

52

well as metastable states) may undergo unintended transformations to stable crystalline forms,

53

altering their therapeutic properties in an uncontrolled manner.17 Consequently, keeping the benefits

54

and promoting the use of amorphous compounds in pharmaceutical formulations requires a correct

55

mastering of the amorphous state along with innovative approaches to extend its lifetime.18

56 57

In this sense, stabilization of amorphous active substances is of particular interest to the

58

pharmaceutical industry.19–26 Among suppression of crystallization of labile Active

59

Pharmaceutical Ingredients (APIs), preparation of homogeneous dispersions are often conducted

60

with a thorough complex mix of an active substance and excipient.27–31 A miscible solid

61

dispersion of a drug and an additive is set as a unique chemically homogeneous phase where all

62

components are narrowly mixed at the molecular level.32 From there, properties of the formulation

63

are different from the properties of the pure components. Crystallization of the pharmaceutical

ACS Paragon Plus Environment

3

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 4 of 50

64

ingredient into the blend can generate non-miscibility as the drug is in a separate crystalline state.

65

Moreover, a resulting phase separation could be demonstrated by two existing amorphous phases.

66

Felodipine, for instance, is completely miscible with polyvinylpyrrolidone (PVP)33 allowing an

67

inhibition of crystallization while with poly acrylic acid, the system tends to effortlessly crystallize

68

due to a partial miscibility.34

69 70

Experimental studies reported that association of an active pharmaceutical ingredient with specific

71

excipients such as antiplasticizers, surfactants or polymers could yield to an amorphous solid

72

dispersion with a greater resilience to crystallization during downstream operations and storage. Over

73

several years, polymers have been shown to modify solid-state crystallization of amorphous APIs

74

arising from their stabilizing effect.35,36 This stabilization stems from several possible explanations

75

including specific polymer-drug interactions, destabilization of drug interactions or reduction of

76

molecular mobility. Since crystallization has a complex character, all these mechanisms together

77

must be viewed as a whole: stabilization effect cannot be attributed to a single mechanism. Among

78

the polymer class, plenty of them own a high glass transition temperature typically greater than

79

100 °C. Thus, a high Tg polymer combined with a drug with a low Tg, ensures a higher glass

80

transition temperature as it obeys to the Gordon Taylor law.37 Moreover, a higher glass transition

81

temperature might result in a lower tendency to recrystallize. Indeed, addition of a polymer can act as

82

an antiplasticizer; molecular mobility of the amorphous drug would therefore be limited. Restrained

83

drug molecules would not be able to generate a stable nucleus, limiting crystallization outcomes. Yet,

84

increase of Tg of the solid dispersion could not be considered as the main factor preventing

85

crystallization since studies indicated that inhibiting crystallization could be accomplished even when

86

Tg is not decreased or affected.38,39 It therefore appears that specific molecular interactions between

87

APIs and excipients exist and could be accountable for stabilization of amorphous active

88

pharmaceuticals.

ACS Paragon Plus Environment

4

Page 5 of 50 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

89

The significance of specific interactions between the amorphous state of a drug and an additive has

90

also been broadly discussed in the background of crystallization from an amorphous matrix. On the

91

basis of Fourier Transform Infrared (FTIR) and Raman spectroscopies, Indomethacin and

92

Polyvinylpyrrolidone were shown to interoperate through intermolecular hydrogen bonding.40 It was

93

postulated that drug / polymer hydrogen bonding interactions were relevant for stabilizing

94

amorphous Indomethacin by disorganizing Indomethacin dimer’s self-association, even at low

95

polymer concentrations where the antiplasticizing effect of the polymer is minimal. Likewise, FTIR

96

analyses along with theoretical calculations have attested the presence of molecular interactions of

97

hydrogen bonding type between Celecoxib and Polyvinylpyrrolidone.41

98 99

More recently, the research work of Paluch and co-workers evidenced the considerable potential of

100

low molecular weight excipients as carbohydrate derivatives with acetate groups (i.e. acetylated

101

saccharides) to enhance the physical and chemical stability of amorphous drugs.42–52 Although

102

several parameters are relevant for stabilization of amorphous dispersions, the group of Prof. Paluch

103

indicated that the glass transition temperature might not be the only criterion for ensuring long-term

104

stability of amorphous API in binary systems and highlighted the formation of inter and

105

intramolecular interactions in the formed solid dispersions. However, despite examples of promising

106

results between possible interaction among excipients and an amorphous drug, there can be specific

107

cases where no particular interaction is observed between the amorphous drug substance and the

108

excipient; physical stability could therefore be accounted to kinetic factors.

109 110

As of now, a rationalization of interactions for APIs/excipients was still not reached due to the low

111

number of molecular compounds analyzed. With this in mind, this current study investigates the

112

effect of low molecular weight excipients on stabilization mechanism, molecular mobility and

113

the inclination to recrystallization of amorphous Biclotymol. This active pharmaceutical

ACS Paragon Plus Environment

5

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 50

114

ingredient which is used for the treatment of otolaryngology infections has been subject to

115

several studies regarding its crystallization kinetics and stability.53,54 To investigate Biclotymol

116

amorphous solid dispersions, we selected series of acetylated saccharides namely α-

117

Pentaacetylglucose (α-acGLU) and β-Pentaacetylglucose (β-acGLU) (Figure 1). A specific

118

feature of these two low molecular weight excipients is their chemical structure. When compared

119

to non-modified saccharides, acetylated derivatives are found to possess acetyl groups instead of

120

hydrogen atoms. As a matter of fact, this substitution towards acetyl groups has a significant

121

influence on intermolecular interactions: non-modified glucose is a strongly associating liquid

122

whereas acetylic systems are typical Van der Waals liquids. Acetylated derivatives are rich in

123

oxygen which is a good acceptor for hydrogen bonds, but do not have proton donors. Therefore,

124

acetylated derivatives cannot create hydrogen bonds on their own.

125 126

To do so, impact of components concentration, molecular structure and size of the excipient and

127

attainable interactions onto the molecular dynamics of the amorphous drug dosage were

128

regarded. Then, for the purpose of selecting the appropriate API-excipient blend, critical

129

characteristics of components are examined to provide insights into stabilization mechanisms as

130

well as the impact of intermolecular interactions on the life expectancy of amorphous

131

formulations.

132 133

Experimental Methods

134

Materials

135

Biclotymol (C21H26Cl2O2, Mw = 381.32 g/mol) received as a crystalline white powder was kindly

136

provided by Pharmasynthese (Inabata Group). No impurity was detected under high performance

137

liquid chromatography-sustained conditions. The compound was consequently used without

ACS Paragon Plus Environment

6

Page 7 of 50 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

138

prepurification. An X-ray pattern of commercial Biclotymol was recorded and revealed a

139

completely anhydrous crystalline form, in agreement with Rantsordas et al. (Cambridge

140

Structural Database (CSD-MYCMPP).55 This Active Pharmaceutical Ingredient (API) exists in

141

two crystalline varieties, one stable (Form I) and one metastable (Form II) in a monotropic

142

relationship.56,53,57,54 α-Pentaacetylglucose (α-acGLU, C16H22O11, Mw = 390.34 g/mol) with

143

purity greater than 99% was purchased from Sigma-Aldrich whereas β-Pentaacetylglucose (β-

144

acGLU, C16H22O11, Mw = 390.34 g/mol) with purity greater than 98% was purchased from Alfa

145

Aesar. These two compounds were used without prepurification. Developed formulae of

146

Biclotymol, α-Pentaacetylglucose and β-Pentaacetylglucose are shown in Figure 1.

147 148

Figure 1. Developed formulae of the Biclotymol molecule (A) and Haworth projection of

149

chemical structures of α-Pentaacetylglucose (B) and β-Pentaacetylglucose (β-acGLU) (C).

150

Preparation of Amorphous Systems

151

The amorphous Biclotymol, α-Pentaacetylglucose (α-acGLU) and β-Pentaacetylglucose (β-

152

acGLU) as well as binary systems with different weight fractions of acetylated saccharides were

153

prepared by the quench cooling technique. In order to obtain homogeneous Biclotymol-

154

acetylated saccharide solid dispersions, crystalline powders of both compounds were first

ACS Paragon Plus Environment

7

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 8 of 50

155

thoroughly mixed by hand mortar and pestle milling for five minutes (operation was repeated

156

twice). Crystalline mixtures were fully melted (Tm Biclotymol = 125.6 °C, Tm α-acGLU = 112.6 °C, Tm β-

157

acGLU

158

metal crucible to a very cold metal plate. Heating rates were only of 10 K/min while cooling

159

rates were always greater than 50 K/min, with a time spent in the melt of two minutes.

160

No separation, cracks or heterogeneity of the sample have been detected. Only completely

161

amorphous samples obtained in this way were analyzed immediately after preparation to protect

162

them from atmospheric moisture and any structural changes. A total of ten mixtures of

163

Biclotymol with acetylated saccharides were investigated. For each excipient (α-acGLU and β-

164

acGLU), various weight ratios as 5:1, 2:1, 1:1, 1:2 and 1:5 were prepared. The first number

165

corresponds to the weight ratio of Biclotymol whereas second number corresponds to the weight

166

ratio of excipients (α-acGLU or β-acGLU).

167

= 131.2 °C) and quenched either in-situ inside the DSC pan or ex-situ by a transfer of a

Dielectric Relaxation Spectroscopy (DRS)

168

Measurements were carried out in a frequency range of 10−2 – 2.106 Hz with a Novocontrol

169

Alpha Analyzer. Dielectric spectra were collected over a wide temperature range from -140 to

170

100 °C, with appropriate successive steps. Accurate temperature control was implemented using

171

the Quatro system (Novocontrol Technologies), allowing a temperature stability of ± 0.2 °C.

172

Amorphous solid dispersions were analyzed using high-quality interdigitated electrodes from

173

Novocontrol Technologies (sensor diameter 20 mm, gold-plated copper combs) as previously

174

outlined in other papers.58–60

175

X-ray Powder Diffraction (XRPD)

176

XRPD analyses of solid mixtures were recorded on a benchtop INEL Equinox 100

177

diffractometer, mounted with a copper micro focus tube (λ = 1.541 Å), equipped with a curved

178

detector. The acquisitions are performed by reflexion in real time up to 110° in 2θ, at 40 kV, 10

ACS Paragon Plus Environment

8

Page 9 of 50 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

179

mA with a metal rotating sample holder. A spinning sample holder has been used to overcome

180

preferential crystallization effects (Ø 15 mm in diameter and Ø 0.5 mm in thickness). XRPD

181

duration was of 10 min and performed directly after the production of samples.

182

Temperature-Modulated Differential Scanning Calorimetry (TM-DSC)

183

The thermodynamic properties of solid dispersions were investigated by the TM-DSC technique.

184

Calorimetric measurements were performed on a Q2000 DSC instrument (TA Instruments)

185

coupled with a liquid-nitrogen cooling system. Temperature and enthalpy calibrations were

186

carried out with both indium and benzophenone standards. Specific heat capacities were

187

measured using sapphire as a reference. Baseline was calibrated from -20 to 200 °C, with an

188

oscillation amplitude of ± 0.318 K, an oscillation period of 60 s, and a heating rate of 2 K/min

189

used in the experiments. The samples (5.00 ± 0.05 mg) were enclosed in sealed hermetic

190

aluminum pans Tzero, and the atmosphere was regulated by a nitrogen flow (50 mL/min).

191

Raman Spectroscopy

192

Raman analyzes were carried out with a Confocal Raman microscope (LabRam HR by Jobin-

193

Yvon Horiba) coupled to an optical microscope (model BX41, Olympus) with xyz mapping

194

stage. The excitation of Raman scattering was operated by a He-Ne laser at a wavelength of

195

632.8 nm. The laser beam was focused on the sample by a microscope objective X50. The

196

confocal pinhole was of 200 µm diameter. The Raman signal was analyzed using a 600 lines per

197

mm grating and a spectrometer slit of 100 µm. A 25 µm x 25 µm area was analyzed by steps of 1

198

µm. The resolution was estimated to 0.5 µm in (X,Y) and 9 µm in (Z). An exposure time of 15 s

199

was used to record the Raman Spectra over the wavenumber range 2700-3200 cm-1 with a one-

200

time accumulation. The duration of the data collection was adjusted in order to minimize the

201

background signal.

202

ACS Paragon Plus Environment

9

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

203

Page 10 of 50

Computational details

204

Theoretical calculations for molecular interactions of Biclotymol-α-acGLU and Biclotymol-β-

205

acGLU were performed within the framework of density functional theory (DFT) using the

206

Gaussian 09 package.61 The initial geometries for the isolated molecules were obtained from the

207

respective crystalline structures of Biclotymol, α-Pentaacetylglucose (α-acGLU) and β-

208

Pentaacetylglucose (β-acGLU), and were optimized with the dispersion corrected B97D exchange-

209

correlation functional.62 Forty structures for the bimolecular complexes (twenty for the Biclotymol-

210

α-acGLU and twenty for Biclotymol-β-acGLU) were generated using our own fully quantum

211

simulated annealing63, based on high temperature semi-empirical PM6 molecular dynamics

212

simulations, followed by root-mean-square deviation (RMSD) filtering and a subsequent DFT

213

optimization with small basis sets, finally completed by a second RMSD filtering before final

214

optimization at the high level of theory, for which the following basis set combination was

215

chosen: 6-31+G(d) for Cl and O, 6-31G(d) for C and H. The nature of stationary points were

216

checked by vibrational analysis (harmonic frequencies). Complexation energies, Ecomplexation,

217

were calculated as Ecomplexation = E(AB) - E(A)opt - E(B)opt with E(AB) the energy of the complex

218

A…B, and E(A)opt that for the optimized fragment A with the basis functions centered on A.

219

Interaction energies, Einteraction, were corrected for basis set superposition error using the standard

220

Boys−Bernardi counterpoise method64: Einteraction = E(AB) - EA…B (A) - EA…B (B) with EA…B (A)

221

the self-consistent field (SCF) energy of fragment A with the full basis of A and B at the

222

geometry A takes in the complex. Gibbs energies were evaluated using standard statistical

223

physics formula based on the ideal gas partition function. However, in order to better describe

224

entropy in condensed phase, all contributions for rotational and translational entropies were

225

removed since mobility in such phases is considerably reduced.

226

ACS Paragon Plus Environment

10

Page 11 of 50 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

227

Results and Discussion

228

Structural and Thermodynamic Characterization of Amorphous Solid Dispersions

229

The first part of this section is dedicated to the safe preparation (without thermal decomposition)

230

of amorphous acetylated saccharides alone and solid dispersions formed between acetylated

231

saccharides and Biclotymol by quench cooling from the melt. Thermogravimetric measurements

232

were carried on the commercial crystalline forms to check whether these compounds undergo

233

thermal degradation during the melting process. We checked that the melting endotherms of the

234

crystalline forms of α-acGLU (Tm α-acGLU = 112.6 °C) and β-acGLU (Tm β-acGLU = 131.2 °C) are

235

located far from the range of their thermal decomposition. TGA curves indicate that the onset of

236

thermal decomposition of α-acGLU and β-acGLU begins above 155 °C (see Figure S1,

237

Supporting Information). Besides, Biclotymol (Tm Biclotymol = 127 °C) exhibits no degradation up

238

to 180 °C.53 Therefore, these results show that melting the crystalline forms of Biclotymol as

239

well as acetylated saccharides and preparing various amorphous compositions based on both

240

compounds by vitrification from cooling the melt is safe. Moreover, it should be noted that the

241

melting temperatures of Biclotymol and derived saccharides are very similar. Mixing of

242

compounds in their liquid states could thus be promptly performed without the hazard of

243

components overheating. A merely simple observation shows the creation of homogeneous

244

solutions, in any weight ratio. It is worth noting that preparation of the solid dispersions by

245

quench cooling of the melt does not consume a lot of time and does not require using any

246

additional solvents in contrast to the other amorphous binary mixtures prepared for instance by

247

the solvent evaporation and precipitation techniques. In order to confirm the amorphous nature of

248

quench-cooled acetylated saccharides, the XRPD technique was applied. Figure 2 demonstrates

249

that Biclotymol, α-acGLU and β-acGLU can be transformed into amorphous solids with the

250

selected procedure.

ACS Paragon Plus Environment

11

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 12 of 50

251 252

Figure 2. X-Ray powder diffraction patterns obtained for α-acGLU (blue line), β-acGLU (red

253

line) and Biclotymol (purple line) (A) before melt-quenching and (B) after melt-quenching.

254

Beam cut-off was located close to 6.5° (2θ).

255

Amorphous nature of these excipients was also confirmed using TM-DSC analyses. Figure 3

256

(upper black curves) shows the sharp endothermic events corresponding to the melting process

257

of the commercial stable crystalline forms. After quenching from the melt, a subsequent

258

temperature-modulated DSC heating scan from the glassy state reveals well-defined Cp jumps.

259

For α-acGLU and β-acGLU, exothermic events are noticed corresponding to the emerging cold

260

crystallization. At higher temperatures, it is followed by sharp endothermic event corresponding

261

to the melting of the crystalline form obtained beforehand.

ACS Paragon Plus Environment

12

Page 13 of 50 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

262 263

Figure 3. TM-DSC measurements of crystalline and amorphous forms of Biclotymol (top panel),

264

α-acGLU (left panel) and β-acGLU (right panel). (1) Melting of the commercial stable form, (2)

265

glass transition region, (3) crystallization of the supercooled melt to the crystalline form, (4)

266

melting of the crystalline form.

267

It is remarkable that both pentaacetyl derivatives (differing only in location of the acetyl group

268

attached to the first carbon in the sugar ring), α-acGLU and β-acGLU, possess different

269

crystallization tendencies but practically the same values of Tg, fragilities and molecular

270

dynamics as highlighted by a recent study of Kaminski and coworkers.42 This enabled these

271

authors to postulate that, except for kinetic and thermodynamic factors, molecular conformation

272

may play an important role in controlling crystallization abilities of glass-formers.

ACS Paragon Plus Environment

13

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 14 of 50

273

To confirm the homogeneity of prepared amorphous solid dispersions, confocal Raman

274

spectroscopy has been applied. In recent years, Raman mapping has been effectively applied to

275

amorphous drug systems thanks to its high-spatial resolution.65,66 In this sense, spatial

276

distribution of chemical species can be attained. In the framework of Biclotymol and acetylated

277

saccharides solid dispersions, Raman spectrum analysis for characteristic bands associated with

278

Biclotymol and α-acGLU / β-acGLU was performed. We selected the characteristic band of

279

amorphous Biclotymol located at 2871 cm-1 as well as the one of acGLU and β-acGLU at 2939

280

cm-1 (see Figure S2, Supporting Information). Raman mapping for all ten mixtures, measured at

281

room temperature, showed homogeneous distribution in the selected wavenumber region,

282

attesting the uniformity of our amorphous solid dispersions.

283 284

Figure 4 presents diffraction patterns of the physical mixtures after melt-quenching. In the case

285

of the Biclotymol - α-acGLU mixtures no Bragg peaks are visible. Instead, broad amorphous

286

halo patterns characteristic of materials having no long-range three-dimensional molecular order,

287

were obtained. The crystalline Biclotymol fraction is not visible for compositions 2:1 and 5:1

288

(i.e. enriched in Biclotymol) while it was clearly observed in the XRPD patterns of the

289

crystalline mixtures (see Figure S3, Supporting Information). On the other hand, a totally

290

different scenario was noticed for Biclotymol - β-acGLU mixtures. In Figure 4 (right panel),

291

sharp Bragg peaks are present indicating the presence of crystals in the investigated samples.

292

However, Bragg peaks are only observable for physical mixtures of 1:2 and 1:5 ratios of

293

Biclotymol - β-acGLU. With a closer scrutiny on Figure 4 (right panel), characteristic Bragg

294

peaks of crystalline β-acGLU can be retrieved while no Bragg peaks of crystalline Form I or

295

Form II of Biclotymol can be seen (see Figure S4, Supporting Information). However, one

296

cannot rule out the potential presence of a low fraction of crystallized Biclotymol in the solid

ACS Paragon Plus Environment

14

Page 15 of 50 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

297

dispersions, non-detectable by XRPD. Therefore, for enriched compositions in β-acGLU, the

298

resulting mixture seems to possess fast kinetics of crystallization towards the stable crystalline

299

form of β-acGLU, although XRPD have been recorded immediately after quenching.

300 301

Figure 4. X-ray Powder Diffraction patterns for melt-quenched representative mixtures: (left

302

panel) melt-quenched Biclotymol-α-acGLU and (right panel) melt-quenched Biclotymol-β-

303

acGLU. XRPD analysis was performed directly after the production of samples.

304

One prominent feature in the use of stabilizers is the possibility to suppress a potential

305

crystallization phenomenon. Figure 5 shows TM-DSC measurements recorded upon heating

306

fused-quenched mixtures. For Biclotymol - α-acGLU system, only one endothermic event

307

associated with the glass transition is noticed within the whole examined temperature range. No

308

additional event that could indicate an exothermic cold crystallization or the presence of

309

heterogeneity (segregation) in the sample was observed. Therefore, the non-occurrence of cold-

310

crystallization whatever the temperature, is an indication of the increased stability of amorphous

311

Biclotymol. Conversely, in Figure 5 (right panel), solid dispersions of 1:2 and 1:5 ratios of

312

Biclotymol - β-acGLU are characterized by a strong tendency toward cold-crystallization to the

ACS Paragon Plus Environment

15

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 50

313

crystalline form of β-acGLU. The enthalpy of the exothermal processes increases with the

314

percentage of β-acGLU (∆Hc 1:2 = 25.4 J/g and ∆Hc 1:5 = 42.1 J/g) which seems to be in line with

315

our previous analyses. Solid dispersions of Biclotymol with the two investigated acetylated

316

disaccharides attest the presence of a single glass transition event, regardless the excipient’s

317

weight ratio. Besides, values of the glass transition temperatures for all investigated solid

318

dispersions are nearly the same, as attested by reversing heat capacity values of solid dispersions

319

(see Figures S5 and S6, Supporting Information). This finding is not surprising because even

320

tough amorphous solid dispersions conform to a Gordon-Taylor law, the Tg of pure components

321

are very similar.

322 323

Figure 5. TM-DSC thermograms obtained for melt-quenched Biclotymol-α-acGLU (left panel)

324

and melt-quenched Biclotymol-β-acGLU (right panel) solid dispersions during heating.

325

The structural (XRPD) and thermal (TM-DSC) findings presented above seem to indicate that

326

the amorphous solid dispersions exhibit a very different behaviour between the two selected

327

excipients in the presence of Biclotymol. In one case, there is no crystallization of either

328

Biclotymol and α-acGLU in the Biclotymol- α-acGLU solid dispersions while for Biclotymol-β-

329

acGLU, crystallization of β-acGLU occurs for 1:2 and 1:5 ratios. However, even if no trace of

ACS Paragon Plus Environment

16

Page 17 of 50 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

330

crystalline Biclotymol has been seen in above experiments, its presence cannot be totally

331

excluded. It might be argued that the difference between Tm and Tg could have a potential impact

332

on the evolution of the crystallization driving forces. In the same way, position of the maxima of

333

nucleation and growth of the solid dispersions made with either α-acGLU or β-acGLU will be

334

constrained differently due to the Tm - Tg difference.

335

To monitor the stability of amorphous solid dispersions, new freshly samples were prepared,

336

stored under normal conditions of use and structurally analyzed (Figure 6). It has been found that

337

for the highest concentration of acetylated saccharides (i.e. 1:2 and 1:5), presence of crystalline

338

α-acGLU (Figure 6, left panel) and β-acGLU (Figure 6, right panel) is still noticeable after four

339

months of storage. This is also the case for solid dispersions enriched in Biclotymol (i.e. 5:1)

340

where crystalline Form I of Biclotymol is detectable. In these two specific cases, one can note

341

the valuable information that recrystallization of Biclotymol toward its metastable Form II could

342

be inhibited by addition of modified saccharides. Thus, the right mixing ratio would be a drug

343

loading of Biclotymol two times larger than the selected excipient (ratio of 2:1) as no

344

crystallization was highlighted for both excipients.

ACS Paragon Plus Environment

17

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 18 of 50

345 346

Figure 6. X-ray Powder Diffraction patterns for melt-quenched representative mixtures after 4

347

months of storage at 293 K, 30% RH: (left panel) melt-quenched Biclotymol-α-acGLU and

348

(right panel) melt-quenched Biclotymol-β-acGLU.

349

Molecular Dynamics of Amorphous Solid Dispersions

350

Various features can play a role on the stabilization mechanisms of amorphous materials. Among

351

them molecular mobility is widely considered.67 To characterize the molecular dynamics of our

352

amorphous solid dispersions at several temperatures, dielectric measurements were completed. It

353

has been proven in recent years that this experimental procedure provides a precise investigation

354

of glass-forming systems by looking at relaxation processes at temperatures lower and higher

355

than the glass transition temperature.68–71

356

Figure 7, Figure 8 and Figure 9 display selected dielectric spectra for pure Biclotymol, α-acGLU,

357

β-acGLU and as well as their representative amorphous solid dispersions.

ACS Paragon Plus Environment

18

Page 19 of 50 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

358 359

Figure 7. Imaginary (ε″) parts of the complex dielectric permittivity vs frequency in the glassy

360

(T < Tg) and supercooled liquid states (T > Tg) for (a) pure α-acGLU, (b) β-acGLU and (c) pure

361

Biclotymol. Left vertical axe corresponds to the imaginary permittivity of the α-process while

362

right vertical axe represents the γ-process.

ACS Paragon Plus Environment

19

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 50

363 364

Figure 8. Dielectric loss spectra of representative amorphous solid dispersions between

365

Biclotymol and α-acGLU, analyzed above and below Tg. (a) Biclotymol - α-acGLU 1:5 (b)

366

Biclotymol - α-acGLU 1:2 (c) Biclotymol - α-acGLU 1:1 (d) Biclotymol - α-acGLU 2:1 (e)

367

Biclotymol - α-acGLU 5:1. Left vertical axe corresponds to the imaginary permittivity of the α-

368

process while right vertical axe represents the γ-process.

ACS Paragon Plus Environment

20

Page 21 of 50 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

369 370

Figure 9. Dielectric loss spectra of representative amorphous solid dispersions between

371

Biclotymol and β-acGLU, analyzed above and below Tg. (a) Biclotymol - β-acGLU 1:5 (b)

372

Biclotymol - β-acGLU 1:2 (c) Biclotymol - β-acGLU 1:1 (d) Biclotymol - β-acGLU 2:1 (e)

373

Biclotymol - β-acGLU 5:1. Left vertical axe corresponds to the imaginary permittivity of the α-

374

process while right vertical axe represents the γ-process.

ACS Paragon Plus Environment

21

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 22 of 50

375

Key characteristics can be monitored: at low frequencies, dc-conductivity can be identified.

376

Moving toward higher frequencies, the structural α-relaxation associated with the dynamic glass

377

transition can be observed. These two phenomena were found to shift toward lower frequencies

378

upon cooling. In addition, below Tg, in the glassy state, one distinct secondary relaxation

379

(denoted γ-relaxation) was noticed for pure Biclotymol, α-acGLU, β-acGLU as well as for the

380

amorphous solid dispersions. As emphasized in Figure 7, Figure 8 and Figure 9, dielectric

381

spectra obtained illustrate the uniformity of all of our amorphous solid dispersions since one

382

single clearly discernible α-relaxation was highlighted. In addition, no Maxwell-Wagner

383

relaxation process (often emerging in phases separating systems) was detected along all our

384

blends, further confirming the overall homogeneity of as-prepared solid dispersions.72 Besides,

385

additional analysis of structural relaxation loss peaks across the one-sided Fourier transform of

386

the Kohlrausch−Williams−Watts (KWW)73 function revealed that stretching parameter βKWW

387

remains steady over the entire temperature range (βKWW

388

temperature has nearly no impact on the shape of the α-relaxation process but once more confirm

389

the uniformity of our amorphous blends (see Figure S7, Supporting Information).

mean value

= 0.62). This highlights that

390 391

Primary and secondary relaxations of pure compounds and amorphous solid dispersions were

392

analyzed by means of a single Vogel-Fulcher-Tammann74–76 (VTF) and an Arrhenius equation,

393

respectively (detailed description of the data fitting can be found in Supporting Information).

394

Determined relaxation times were plotted as a function of the inverse of the temperature as

395

displayed in the relaxation map in Figure 10 (temperature dependence of the fit parameters αHN

396

and βHN as well as enlarged view of the caption can be seen in Figures S8, S9 and S10). From the

397

estimated parameters of VTF fits, Tg can be readily estimated at dielectric relaxation time τα =

398

100 s (Table 1). It should be emphasized that all primary α-relaxation processes lie in the same

ACS Paragon Plus Environment

22

Page 23 of 50 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

399

temperature region, highlighting almost identical Tg values as well as the fragility component,

400

regardless of the selected excipient. Taking into account errors on fragility values, we could

401

consider that such a parameter is concentration-independent. Moreover, Tg values obtained by

402

DRS are somewhat lower compared to those obtained from TM-DSC measurements but are in

403

accordance with the value of Tg for τα = 10 s (corresponding to the equivalent frequency of the

404

TM-DSC analysis).77

405 406

Figure 10. Relaxation map of pure Biclotymol, pure α-acGLU, pure β-acGLU and amorphous

407

solid dispersions. Solid gray lines are the VTF fits to the of structural α-relaxation times,

408

illustrated by filled symbols. Temperature dependence of the secondary γ-relaxations is

409

illustrated by open symbols and was fitted by the Arrhenius equation.

410

Table 1. Glass transition temperatures (Tg) of pure Biclotymol and pure excipients (α-acGLU

411

and β-acGLU) determined from TM-DSC and DRS, fragilities obtained from DRS data and

ACS Paragon Plus Environment

23

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 24 of 50

412

activation energy barriers Ea γ of the secondary relaxation observed in solid dispersion blends.

413

The uncertainty of the glass transition temperatures is ± 1 K while errors for fragility and

414

activation energies is of 10 %, standard error for polymers.78

Sample

Tg from TM-DSC (in K)

Tg from DRS (in K) (τα = 10 s)

Tg from DRS (in K) (τα = 100 s)

Fragility (from DRS)

Pure Biclotymol Pure α-acGLU Pure β-acGLU

293 290 294

293 286 292

290 283 288

88 97 85

Activation energy barriers of γrelaxation (in kJ/mol) 75.3 45.3 42.5

Biclotymol- α-acGLU 1:5 Biclotymol- α-acGLU 1:2 Biclotymol- α-acGLU 1:1 Biclotymol- α-acGLU 2:1 Biclotymol- α-acGLU 5:1

289 291 293 295 297

287 289 292 294 295

284 286 289 290 291

91 89 94 91 84

41.2 43.7 39.2 49.2 35.3

Biclotymol- β-acGLU 1:5 Biclotymol- β-acGLU 1:2 Biclotymol- β-acGLU 1:1 Biclotymol- β-acGLU 2:1 Biclotymol- β-acGLU 5:1

293 292 293 295 293

290 290 291 294 294

287 287 288 291 291

86 87 92 87 85

41.9 38.5 48.4 43.2 31.3

415 416

In order to analyze the progress of crystallization, the temperature dependence of the dielectric

417

strength (∆εα) of the α-relaxation versus temperature for all examined amorphous solid

418

dispersions as well as pure components was outlined in Figure 11. As defined by Onsager,

419

Fröhlich, and Kirkwood equation72, by increasing the temperature, the dielectric relaxation

420

strength ∆εα decreases for all samples.72,79 Similar trends can be seen for other glass-forming

421

systems as polymers80 and pharmaceutical substances.81

422 423

Interestingly, by contrast to similar glass transition temperatures obtained for all mixtures,

424

dielectric strength values for solid dispersions do not follow a classical mixing rule. As shown in

425

Figure 11 a), the dielectric strength of Biclotymol-α-acGLU amorphous solid dispersions is

426

relatively high and clustered around ∆εα = 8 to 11, except for the blend with an enriched

ACS Paragon Plus Environment

24

Page 25 of 50 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

427

Biclotymol composition which is consistently closer to pure Biclotymol. Relating to Biclotymol-

428

β-acGLU amorphous solid dispersions in Figure 11 b), blends made of 1:1, 2:1 and 5:1 possess

429

the same behavior pattern. In contrast, compositions enriched with excipients (1:2 and 1:5) have

430

lower ∆εα against other compositions. This certainly seems reasonable as the density of dipoles

431

involved in the relaxation process (i.e. amount of relaxing amorphous phase) should be falling

432

due to the onset of recrystallization of these formulations as displayed in Figure 5. This

433

difference in terms of ∆εα may originate from the spatial distribution of acetylated carbohydrates

434

with respect to Biclotymol molecules and may lead to a change of the strength of the

435

intermolecular interactions in amorphous solid dispersions.

436 437

Figure 11. Temperature dependence of the dielectric strength for α- relaxation (∆εα) versus

438

temperature for all investigated amorphous solid dispersions. Left panel a) corresponds to

439

Biclotymol-α-acGLU blends while right panel b) matches with Biclotymol-β-acGLU mixtures.

440

With reference to the glassy state of solid dispersions, pure Biclotymol demonstrates the highest

441

activation energy barrier Ea

442

Tripathi et al. (Ea = 75.5 kJ/mol).57 As for this API, the molecular origin of its secondary

443

relaxation was previously discussed. With the use of dielectric spectroscopy, Tripathi et al.

444

suggested a likely intramolecular secondary dynamic involving the hydrogen bonds formed by

γ-Biclotymol,

close to 75.3 kJ/mol in accordance with that found by

ACS Paragon Plus Environment

25

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 26 of 50

445

the hydroxyl groups of Biclotymol. Pure acetylated saccharides, α-acGLU and β-acGLU, own

446

lower activation energy barriers of Ea

447

respectively. Those values are in the range determined by Kaminski et al.42 Addition of modified

448

saccharides, in whatever amount, give rise to a single secondary relaxation process. The

449

activation energy barriers ascertained for the γ-relaxations are comparable, within the range of 35

450

- 49 kJ/mol for Biclotymol-α-acGLU and 31 - 48 kJ/mol for Biclotymol-β-acGLU (Table 1).

451

These values are similar to the ones of pure α-acGLU and β-acGLU, suggesting the involvement

452

of the same molecular motions. Owing to their relatively low Ea values, secondary processes in

453

our solid dispersions originate most likely from an intramolecular origin. According to an earlier

454

study, this stems from the motion of acetyl sequences in modified carbohydrates.82 We can

455

therefore consider that the secondary relaxation process occurring in solid dispersions is linked

456

to intramolecular mechanisms associated with short-range intermolecular interactions within

457

Biclotymol and the selected excipient. This is supported by the relaxation map in Figure 10,

458

depicting that the secondary γ-relaxations show nearly no energy variations but rather

459

temperature changes, supporting the potential impact of intermolecular interactions in

460

amorphous mixtures. In addition to this, by taking into account the possible formation of

461

hydrogen-bonds between Biclotymol and the acetylated saccharides, one would think that the

462

intramolecular γ-relaxation could be inhibited if some these H-bonds are related to the molecular

463

parts able to generate the relaxation (at least for some of the pure components in the mixture).

γ-α-acGLU =

45.3 kJ/mol and Ea

γ-β-acGLU =

42.5 kJ/mol,

464 465

It seems clear from the above findings that molecular mobility of investigated amorphous solid

466

dispersions made of Biclotymol and pentaacetyl derivatives are very alike. Investigated

467

formulations reveal a homogenous character with a single structural mechanism along with a

468

unique noticeable secondary relaxation process for all blends. In this way, the observed

ACS Paragon Plus Environment

26

Page 27 of 50 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

469

discrepancies by XRPD and DSC between solid dispersions made with α-acGLU and β-acGLU

470

were not spotlighted by dielectric measurements, known to finely characterize the amorphous

471

state. Hence, the excipient concentration likewise with its molecular structure tends to reveal

472

quasi-identical molecular dynamic behavior of solid dispersions above and below the glass

473

transition temperature. This indicates that a crystalline environment composed of β-acGLU, in

474

the case of unstable solid dispersions 1:2 and 1:5 of Biclotymol - β-acGLU, does not impact the

475

resulting amorphous fraction in terms of molecular mobility. On the other hand, the

476

thermodynamic study on solid dispersions highlighted a stabilizing effect of α-acGLU regardless

477

its quantity while an excessive concentration of β-acGLU (blends 1:2 and 1:5) would promote an

478

inclination to recrystallization to the crystalline form β-acGLU and physical instability. The

479

difference here with regard to a classical stabilization protocol83 lies in the fact that selected API

480

(i.e. Biclotymol) and excipients (i.e. α-acGLU and β-acGLU) possess quasi-identical glass

481

transition temperature, very close to room temperature. Therefore, the obtained results would

482

suggest that in solid dispersions, the glass transition temperature is not the only factor for the

483

stabilization and that molecular interactions between API and excipient could yield to greater

484

resilience to devitrification. With this viewpoint, complementary quantum mechanical

485

simulations have been conducted.

486 487

Modelling of Amorphous Solid Dispersions

488

Quantum simulations happen to become increasingly popular in the pharmaceutical field,

489

especially due to its usefulness in the interpretation of vibrational spectra (for experimental band

490

assignment mainly)84 or for the validation of molecular structures by means of identification of

491

conformational minima.85 Indeed, a significant increase in the use of density functional theory

492

(DFT) has been recorded over the past decades in computational chemistry, from medium to

ACS Paragon Plus Environment

27

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 28 of 50

493

large molecules, including pharmaceuticals.86–89 Therefore, we performed DFT calculations in

494

order to provide a more complete picture of the interactions between Biclotymol and selected

495

acetylated saccharides.

496

To this aim, an extensive sampling of the potential energy surface was performed, leading to the

497

location of many energy minima, which are collected in the supplementary information file

498

(Figures S11 and S12). The most stable optimized structures of Biclotymol-α-acGLU and

499

Biclotymol-β-acGLU systems are displayed in Figure 12. The most negative values for the

500

calculated complexation energies for both systems were very alike, with values of -91.1 kJ/mol

501

for Biclotymol-α-acGLU and -95.8 kJ/mol for Biclotymol-β-acGLU. Besides, interaction

502

energies (which do not take into account the deformation energy necessary to form the complex

503

from the isolated molecules at their energy minimum) are even closer since they equal -110.70

504

kJ/mol for Biclotymol-α-acGLU and -110.13 kJ/mol Biclotymol-β-acGLU.

505

These results provide a clear indication of a strong interaction between Biclotymol and the

506

acetylated saccharides. With respect to non-covalent interactions, a single hydrogen bond (C40-

507

O43...H65-C64) was noticed with a distance of 1.97 Å for Biclotymol-α-acGLU complexes (Figure

508

12 (A)), while two hydrogen bonds were found for Biclotymol-β-acGLU: one of intramolecular

509

origin (C54-O64…H67-O66 = 1.75 Å) and one of intermolecular nature (C44-O45…H65-O64) with a

510

distance of 1.85 Å (Figure 12 (B)), which could account for the very slight higher stability of the

511

dimer formed by β-acGLU. However, on account of their complexity, it has proved difficult to

512

achieve detection of Raman bands related to the new hydrogen bonds in the recorded spectra.

ACS Paragon Plus Environment

28

Page 29 of 50 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

513 514

Figure 12. Optimized structures of Biclotymol-α-acGLU (upper panel) and Biclotymol-β-

515

acGLU (lower panel) and corresponding complexation energies.

516

However, one should keep in mind that our DFT calculations, consistent with those in the

517

literature, were made on a ratio of one molecule of API with one molecule of excipient. Actually,

518

the hydrogen bonds network revealed by the DFT calculations suggests that collective effects

519

involving many molecules could be at stake, a behavior that fully unravel the drug loading effect

520

in the amorphous solid dispersions. Temperature certainly is an important factor to account for

521

the dynamics of these complexes. Even if our DFT calculations describe static properties,

522

temperature effects can actually be incorporated through the evaluation of standard Gibbs

523

complexation energies. Indeed, the most negative values were found to be equal to -93.7 kJ/mol

524

(Biclotymol-β-acGLU) and -109.9 kJ/mol (Biclotymol-α-acGLU). This not-negligible difference

525

can be ascribed to vibrational entropies (Svib), the –TSvib(AB) values amounting to -278.1 kJ/mol

526

and -287.5 kJ/mol, respectively. This indicates that even if interaction and complexation energies

ACS Paragon Plus Environment

29

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 30 of 50

527

are almost identical, temperature-dependent properties could differ for the two mixings. This

528

could explain why the two systems could appear similar or different depending on the

529

experimental technics used to study them.

530

Conclusions

531

In this paper, amorphous solid dispersions formed by a series of low molecular weight

532

excipients, i.e. acetylated saccharides, and Biclotymol were investigated. Effect of these

533

excipients on drug-excipient interactions, molecular mobility and recrystallization outcomes

534

were considered. We have shown that homogeneous amorphous solid dispersions can be safely

535

prepared and analyzed by means of various experimental techniques and also investigated by

536

molecular modelling techniques. Our findings show that the molecular mobility above and below

537

the glass transition temperature cannot be considered as the only factor for stabilization of solid

538

dispersions. Indeed, the glass transition temperatures of the solid dispersions, irrespective of the

539

ratio, are very close, a feature that was related to very similar interaction and complexation

540

energies (close to 100 kJ/mol in absolute value) as revealed by density functional theory

541

calculations. In addition, stability study showed the limiting excipient concentration needed to

542

stabilize our amorphous API for several months under classical storage conditions. The results of

543

the above study acted as the basis for further analyses, in particular toward a deeper

544

characterization of other complex formulations of molecular alloys, structured by a network of

545

intramolecular and intermolecular hydrogen bonds, before and after the cold-crystallization for

546

instance.

547

Supporting Information Description

548

TGA analysis, Raman spectra, XRPD patterns, TM-DSC analyses, DRS spectra and fitting

549

procedure, molecular complexes simulated by DFT.

ACS Paragon Plus Environment

30

Page 31 of 50 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

550

Acknowledgement

551

The Region Normandie is acknowledged for financial support to B. S. via the E.D. No. 591

552

(PSIME). The authors would like also to acknowledge the high-performance computing facility

553

(CRIANN) funded by the Region Haute Normandie.

554

Author Contributions

555

The manuscript was written through contributions of all authors. All authors have given approval

556

to the final version of the manuscript.

557

ACS Paragon Plus Environment

31

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

558

References

559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604

(1) (2)

(3) (4)

(5)

(6)

(7)

(8) (9)

(10)

(11)

(12)

(13)

(14)

(15) (16)

Page 32 of 50

Desiraju, G. R. Crystal Engineering: From Molecule to Crystal. J. Am. Chem. Soc. 2013, 135 (27), 9952–9967. Amidon, G. L.; Lennernäs, H.; Shah, V. P.; Crison, J. R. A Theoretical Basis for a Biopharmaceutic Drug Classification: The Correlation of in Vitro Drug Product Dissolution and in Vivo Bioavailability. Pharm. Res. 1995, 12 (3), 413–420. Babu, N. J.; Nangia, A. Solubility Advantage of Amorphous Drugs and Pharmaceutical Cocrystals. Cryst. Growth Des. 2011, 11 (7), 2662–2679. Löbmann, K.; Grohganz, H.; Laitinen, R.; Strachan, C.; Rades, T. Amino Acids as CoAmorphous Stabilizers for Poorly Water Soluble Drugs – Part 1: Preparation, Stability and Dissolution Enhancement. Eur. J. Pharm. Biopharm. 2013, 85 (3, Part B), 873–881. Wlodarski, K.; Sawicki, W.; Paluch, K. J.; Tajber, L.; Grembecka, M.; Hawelek, L.; Wojnarowska, Z.; Grzybowska, K.; Talik, E.; Paluch, M. The Influence of Amorphization Methods on the Apparent Solubility and Dissolution Rate of Tadalafil. Eur. J. Pharm. Sci. 2014, 62, 132–140. Xie, T.; Taylor, L. S. Dissolution Performance of High Drug Loading Celecoxib Amorphous Solid Dispersions Formulated with Polymer Combinations. Pharm. Res. 2016, 33 (3), 739–750. Serajuddin, A. T. M. Solid Dispersion of Poorly Water-Soluble Drugs: Early Promises, Subsequent Problems, and Recent Breakthroughs. J. Pharm. Sci. 1999, 88 (10), 1058– 1066. Descamps, M.; Willart, J. F. Perspectives on the Amorphisation/Milling Relationship in Pharmaceutical Materials. Adv. Drug Deliv. Rev. 2016, 100, 51–66. Brough, C.; Williams III, R. O. Amorphous Solid Dispersions and Nano-Crystal Technologies for Poorly Water-Soluble Drug Delivery. Int. J. Pharm. 2013, 453 (1), 157– 166. Kothari, K.; Suryanarayanan, R. Influence of Disorder on Dissolution. In Disordered Pharmaceutical Materials; Descamps, rc, Ed.; Wiley-VCH Verlag GmbH & Co. KGaA, 2016; pp 57–84. Viel, Q.; Brandel, C.; Cartigny, Y.; Eusébio, M. E. S.; Canotilho, J.; Dupray, V.; Dargent, E.; Coquerel, G.; Petit, S. Crystallization from the Amorphous State of a Pharmaceutical Compound: Impact of Chirality and Chemical Purity. Cryst. Growth Des. 2017, 17 (1), 337–346. Schammé, B.; Monnier, X.; Couvrat, N.; Delbreilh, L.; Dupray, V.; Dargent, É.; Coquerel, G. Insights on the Physical State Reached by an Active Pharmaceutical Ingredient upon High-Energy Milling. J. Phys. Chem. B 2017, 121 (19), 5142–5150. Murdande, S. B.; Pikal, M. J.; Shanker, R. M.; Bogner, R. H. Solubility Advantage of Amorphous Pharmaceuticals: I. A Thermodynamic Analysis. J. Pharm. Sci. 2010, 99 (3), 1254–1264. Zhou, D.; Zhang, G. G. Z.; Law, D.; Grant, D. J. W.; Schmitt, E. A. Physical Stability of Amorphous Pharmaceuticals: Importance of Configurational Thermodynamic Quantities and Molecular Mobility. J. Pharm. Sci. 2002, 91 (8), 1863–1872. Healy, A. M.; Worku, Z. A.; Kumar, D.; Madi, A. M. Pharmaceutical Solvates, Hydrates and Amorphous Forms: A Special Emphasis on Cocrystals. Adv. Drug Deliv. Rev. Lemercier, A.; Viel, Q.; Brandel, C.; Cartigny, Y.; Dargent, E.; Petit, S.; Coquerel, G. Optimization of Experimental Conditions for the Monitoring of Nucleation and Growth of Racemic Diprophylline from the Supercooled Melt. J. Cryst. Growth 2017, 472, 11–17.

ACS Paragon Plus Environment

32

Page 33 of 50 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

605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652

Molecular Pharmaceutics

(17) Morris, K. R.; Griesser, U. J.; Eckhardt, C. J.; Stowell, J. G. Theoretical Approaches to Physical Transformations of Active Pharmaceutical Ingredients during Manufacturing Processes. Adv. Drug Deliv. Rev. 2001, 48 (1), 91–114. (18) Edueng, K.; Mahlin, D.; Bergström, C. A. S. The Need for Restructuring the Disordered Science of Amorphous Drug Formulations. Pharm. Res. 2017, 34 (9), 1754–1772. (19) Laitinen, R.; Löbmann, K.; Strachan, C. J.; Grohganz, H.; Rades, T. Emerging Trends in the Stabilization of Amorphous Drugs. Int. J. Pharm. 2013, 453 (1), 65–79. (20) Löbmann, K.; Jensen, K. T.; Laitinen, R.; Rades, T.; Strachan, C. J.; Grohganz, H. Stabilized Amorphous Solid Dispersions with Small Molecule Excipients. In Amorphous Solid Dispersions; Springer, New York, NY, 2014; pp 613–636. (21) Martínez, L. M.; Videa, M.; López-Silva, G. A.; de los Reyes, C. A.; Cruz-Angeles, J.; González, N. Stabilization of Amorphous Paracetamol Based Systems Using Traditional and Novel Strategies. Int. J. Pharm. 2014, 477 (1–2), 294–305. (22) Martínez, L. M.; Videa, M.; Sosa, N. G.; Ramírez, J. H.; Castro, S. Long-Term Stability of New Co-Amorphous Drug Binary Systems: Study of Glass Transitions as a Function of Composition and Shelf Time. Molecules 2016, 21 (12), 1712. (23) Theil, F.; Anantharaman, S.; Kyeremateng, S. O.; van Lishaut, H.; Dreis-Kühne, S. H.; Rosenberg, J.; Mägerlein, M.; Woehrle, G. H. Frozen in Time: Kinetically Stabilized Amorphous Solid Dispersions of Nifedipine Stable after a Quarter Century of Storage. Mol. Pharm. 2017, 14 (1), 183–192. (24) Tominaka, S.; Kawakami, K.; Fukushima, M.; Miyazaki, A. Physical Stabilization of Pharmaceutical Glasses Based on Hydrogen Bond Reorganization under Sub-Tg Temperature. Mol. Pharm. 2017, 14 (1), 264–273. (25) Szczurek, J.; Rams-Baron, M.; Knapik-Kowalczuk, J.; Antosik, A.; Szafraniec, J.; Jamróz, W.; Dulski, M.; Jachowicz, R.; Paluch, M. Molecular Dynamics, Recrystallization Behavior, and Water Solubility of the Amorphous Anticancer Agent Bicalutamide and Its Polyvinylpyrrolidone Mixtures. Mol. Pharm. 2017, 14 (4), 1071–1081. (26) Knapik-Kowalczuk, J.; Wojnarowska, Z.; Rams-Baron, M.; Jurkiewicz, K.; CieleckaPiontek, J.; Ngai, K. L.; Paluch, M. Atorvastatin as a Promising Crystallization Inhibitor of Amorphous Probucol: Dielectric Studies at Ambient and Elevated Pressure. Mol. Pharm. 2017, 14 (8), 2670–2680. (27) Leuner, C.; Dressman, J. Improving Drug Solubility for Oral Delivery Using Solid Dispersions. Eur. J. Pharm. Biopharm. 2000, 50 (1), 47–60. (28) Van den Mooter, G. The Use of Amorphous Solid Dispersions: A Formulation Strategy to Overcome Poor Solubility and Dissolution Rate. Drug Discov. Today Technol. 2012, 9 (2), e79–e85. (29) Song, Y.; Yang, X.; Chen, X.; Nie, H.; Byrn, S.; Lubach, J. W. Investigation of Drug– Excipient Interactions in Lapatinib Amorphous Solid Dispersions Using Solid-State NMR Spectroscopy. Mol. Pharm. 2015, 12 (3), 857–866. (30) Dengale, S. J.; Grohganz, H.; Rades, T.; Löbmann, K. Recent Advances in Co-Amorphous Drug Formulations. Adv. Drug Deliv. Rev. 2016, 100, 116–125. (31) Martínez, L. M.; Videa, M.; López Silva, T.; Castro, S.; Caballero, A.; Lara-Díaz, V. J.; Castorena-Torres, F. Two-Phase Amorphous-Amorphous Solid Drug Dispersion with Enhanced Stability, Solubility and Bioavailability Resulting from Ultrasonic Dispersion of an Immiscible System. Eur. J. Pharm. Biopharm. 2017, 119, 243–252. (32) Qian, F.; Huang, J.; Hussain, M. A. Drug–polymer Solubility and Miscibility: Stability Consideration and Practical Challenges in Amorphous Solid Dispersion Development. J. Pharm. Sci. 2010, 99 (7), 2941–2947.

ACS Paragon Plus Environment

33

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

653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699

Page 34 of 50

(33) Konno, H.; Taylor, L. S. Influence of Different Polymers on the Crystallization Tendency of Molecularly Dispersed Amorphous Felodipine. J. Pharm. Sci. 2006, 95 (12), 2692– 2705. (34) Rumondor, A. C. F.; Ivanisevic, I.; Bates, S.; Alonzo, D. E.; Taylor, L. S. Evaluation of Drug-Polymer Miscibility in Amorphous Solid Dispersion Systems. Pharm. Res. 2009, 26 (11), 2523–2534. (35) Gue, E.; Willart, J. F.; Muschert, S.; Danede, F.; Delcourt, E.; Descamps, M.; Siepmann, J. Accelerated Ketoprofen Release from Polymeric Matrices: Importance of the Homogeneity/Heterogeneity of Excipient Distribution. Int. J. Pharm. 2013, 457 (1), 298– 307. (36) Chmiel, K.; Knapik-Kowalczuk, J.; Jurkiewicz, K.; Sawicki, W.; Jachowicz, R.; Paluch, M. A New Method To Identify Physically Stable Concentration of Amorphous Solid Dispersions (I): Case of Flutamide + Kollidon VA64. Mol. Pharm. 2017, 14 (10), 3370– 3380. (37) Gordon, J. M.; Rouse, G. B.; Gibbs, J. H.; Jr, W. M. R. The Composition Dependence of Glass Transition Properties. J. Chem. Phys. 1977, 66 (11), 4971–4976. (38) Matsumoto, T.; Zografi, G. Physical Properties of Solid Molecular Dispersions of Indomethacin with Poly(vinylpyrrolidone) and Poly(vinylpyrrolidone-Co-Vinyl-Acetate) in Relation to Indomethacin Crystallization. Pharm. Res. 1999, 16 (11), 1722–1728. (39) Khougaz, K.; Clas, S.-D. Crystallization Inhibition in Solid Dispersions of MK-0591 and Poly(vinylpyrrolidone) Polymers. J. Pharm. Sci. 2000, 89 (10), 1325–1334. (40) Taylor, L. S.; Zografi, G. Spectroscopic Characterization of Interactions Between PVP and Indomethacin in Amorphous Molecular Dispersions. Pharm. Res. 1997, 14 (12), 1691– 1698. (41) Gupta, P.; Thilagavathi, R.; Chakraborti, A. K.; Bansal, A. K. Role of Molecular Interaction in Stability of Celecoxib−PVP Amorphous Systems. Mol. Pharm. 2005, 2 (5), 384–391. (42) Kaminski, K.; Adrjanowicz, K.; Wojnarowska, Z.; Dulski, M.; Wrzalik, R.; Paluch, M.; Kaminska, E.; Kasprzycka, A. Do Intermolecular Interactions Control Crystallization Abilities of Glass-Forming Liquids? J. Phys. Chem. B 2011, 115 (40), 11537–11547. (43) Grzybowska, K.; Paluch, M.; Wlodarczyk, P.; Grzybowski, A.; Kaminski, K.; Hawelek, L.; Zakowiecki, D.; Kasprzycka, A.; Jankowska-Sumara, I. Enhancement of Amorphous Celecoxib Stability by Mixing It with Octaacetylmaltose: The Molecular Dynamics Study. Mol. Pharm. 2012, 9 (4), 894–904. (44) Kaminska, E.; Adrjanowicz, K.; Kaminski, K.; Wlodarczyk, P.; Hawelek, L.; Kolodziejczyk, K.; Tarnacka, M.; Zakowiecki, D.; Kaczmarczyk-Sedlak, I.; Pilch, J.; Paluch, M. A New Way of Stabilization of Furosemide upon Cryogenic Grinding by Using Acylated Saccharides Matrices. The Role of Hydrogen Bonds in Decomposition Mechanism. Mol. Pharm. 2013, 10 (5), 1824–1835. (45) Kossack, W.; Adrjanowicz, K.; Tarnacka, M.; Kipnusu, W. K.; Dulski, M.; Mapesa, E. U.; Kaminski, K.; Pawlus, S.; Paluch, M.; Kremer, F. Glassy Dynamics and Physical Aging in Fucose Saccharides as Studied by Infrared- and Broadband Dielectric Spectroscopy. Phys. Chem. Chem. Phys. 2013, 15 (47), 20641–20650. (46) Kaminska, E.; Adrjanowicz, K.; Tarnacka, M.; Kolodziejczyk, K.; Dulski, M.; Mapesa, E. U.; Zakowiecki, D.; Hawelek, L.; Kaczmarczyk-Sedlak, I.; Kaminski, K. Impact of Interand Intramolecular Interactions on the Physical Stability of Indomethacin Dispersed in Acetylated Saccharides. Mol. Pharm. 2014, 11 (8), 2935–2947.

ACS Paragon Plus Environment

34

Page 35 of 50 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

700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746

Molecular Pharmaceutics

(47) Kaminska, E.; Tarnacka, M.; Kolodziejczyk, K.; Dulski, M.; Zakowiecki, D.; Hawelek, L.; Adrjanowicz, K.; Zych, M.; Garbacz, G.; Kaminski, K. Impact of Low Molecular Weight Excipient Octaacetylmaltose on the Liquid Crystalline Ordering and Molecular Dynamics in the Supercooled Liquid and Glassy State of Itraconazole. Eur. J. Pharm. Biopharm. 2014, 88 (3), 1094–1104. (48) Kaminska, E.; Adrjanowicz, K.; Zakowiecki, D.; Milanowski, B.; Tarnacka, M.; Hawelek, L.; Dulski, M.; Pilch, J.; Smolka, W.; Kaczmarczyk-Sedlak, I.; Kaminski, K. Enhancement of the Physical Stability of Amorphous Indomethacin by Mixing It with Octaacetylmaltose. Inter and Intra Molecular Studies. Pharm. Res. 2014, 31 (10), 2887– 2903. (49) Kaminska, E.; Tarnacka, M.; Wlodarczyk, P.; Jurkiewicz, K.; Kolodziejczyk, K.; Dulski, M.; Haznar-Garbacz, D.; Hawelek, L.; Kaminski, K.; Wlodarczyk, A.; Paluch, M. Studying the Impact of Modified Saccharides on the Molecular Dynamics and Crystallization Tendencies of Model API Nifedipine. Mol. Pharm. 2015, 12 (8), 3007– 3019. (50) Kaminska, E.; Madejczyk, O.; Tarnacka, M.; Jurkiewicz, K.; Kaminski, K.; Paluch, M. Studying of Crystal Growth and Overall Crystallization of Naproxen from Binary Mixtures. Eur. J. Pharm. Biopharm. 2017, 113, 75–87. (51) Grzybowska, K.; Chmiel, K.; Knapik-Kowalczuk, J.; Grzybowski, A.; Jurkiewicz, K.; Paluch, M. Molecular Factors Governing the Liquid and Glassy States Recrystallization of Celecoxib in Binary Mixtures with Excipients of Different Molecular Weights. Mol. Pharm. 2017, 14 (4), 1154–1168. (52) Madejczyk, O.; Kaminska, E.; Tarnacka, M.; Dulski, M.; Jurkiewicz, K.; Kaminski, K.; Paluch, M. Studying the Crystallization of Various Polymorphic Forms of Nifedipine from Binary Mixtures with the Use of Different Experimental Techniques. Mol. Pharm. 2017, 14 (6), 2116–2125. (53) Schammé, B.; Couvrat, N.; Malpeli, P.; Delbreilh, L.; Dupray, V.; Dargent, É.; Coquerel, G. Crystallization Kinetics and Molecular Mobility of an Amorphous Active Pharmaceutical Ingredient: A Case Study with Biclotymol. Int. J. Pharm. 2015, 490 (1–2), 248–257. (54) Schammé, B.; Couvrat, N.; Malpeli, P.; Dudognon, E.; Delbreilh, L.; Dupray, V.; Dargent, É.; Coquerel, G. Transformation of an Active Pharmaceutical Ingredient upon HighEnergy Milling: A Process-Induced Disorder in Biclotymol. Int. J. Pharm. 2016, 499 (1– 2), 67–73. (55) Rantsordas, S.; Perrin, M.; Thozet, A. Crystal and Molecular Structures of 2,2’methylenebis(4-Chloro-3-Methyl-6-Isopropylphenol). Acta Crystallogr. B 1978, 34 (4), 1198–1203. (56) Céolin, R.; Tamarit, J.-L.; Barrio, M.; López, D. O.; Nicolaï, B.; Veglio, N.; Perrin, M.-A.; Espeau, P. Overall Monotropic Behavior of a Metastable Phase of Biclotymol, 2,2′Methylenebis(4-Chloro-3-Methyl-Isopropylphenol), Inferred From Experimental and Topological Construction of the Related P-T State Diagram. J. Pharm. Sci. 2008, 97 (9), 3927–3941. (57) Tripathi, P.; Romanini, M.; Tamarit, J. L.; Macovez, R. Collective Relaxation Dynamics and Crystallization Kinetics of the Amorphous Biclotymol Antiseptic. Int. J. Pharm. 2015, 495 (1), 420–427. (58) Carpentier, L.; Decressain, R.; Desprez, S. Dynamics of the Amorphous and Crystalline α, γ-Phases of Indomethacin. J. Phys. Chem. B 2006, 110 (1), 457–464.

ACS Paragon Plus Environment

35

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

747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792

Page 36 of 50

(59) Willart, J.-F.; Carpentier, L.; Danède, F.; Descamps, M. Solid-State Vitrification of Crystalline Griseofulvin by Mechanical Milling. J. Pharm. Sci. 2012, 101 (4), 1570–1577. (60) Schammé, B.; Mignot, M.; Couvrat, N.; Tognetti, V.; Joubert, L.; Dupray, V.; Delbreilh, L.; Dargent, E.; Coquerel, G. Molecular Relaxations in Supercooled Liquid and Glassy States of Amorphous Quinidine: Dielectric Spectroscopy and Density Functional Theory Approaches. J. Phys. Chem. B 2016, 120 (30), 7579–7592. (61) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. (62) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom–atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10 (44), 6615–6620. (63) Poyer, S.; Loutelier-Bourhis, C.; Tognetti, V.; Joubert, L.; Enche, J.; Bossée, A.; Mondeguer, F.; Hess, P.; Afonso, C. Differentiation of Gonyautoxins by Ion Mobility– mass Spectrometry: A Cationization Study. Int. J. Mass Spectrom. 2016, 402, 20–28. (64) Boys, S. F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19 (4), 553–566. (65) Nanubolu, J. B.; Burley, J. C. Investigating the Recrystallization Behavior of Amorphous Paracetamol by Variable Temperature Raman Studies and Surface Raman Mapping. Mol. Pharm. 2012, 9 (6), 1544–1558. (66) Hédoux, A. Recent Developments in the Raman and Infrared Investigations of Amorphous Pharmaceuticals and Protein Formulations: A Review. Adv. Drug Deliv. Rev. 2016, 100, 133–146. (67) Bhattacharya, S.; Suryanarayanan, R. Local Mobility in Amorphous PharmaceuticalsCharacterization and Implications on Stability. J. Pharm. Sci. 2009, 98 (9), 2935–2953. (68) Grzybowska, K.; Capaccioli, S.; Paluch, M. Recent Developments in the Experimental Investigations of Relaxations in Pharmaceuticals by Dielectric Techniques at Ambient and Elevated Pressure. Adv. Drug Deliv. Rev. 2016, 100, 158–182. (69) Minecka, A.; Kamińska, E.; Tarnacka, M.; Dzienia, A.; Madejczyk, O.; Waliłko, P.; Kasprzycka, A.; Kamiński, K.; Paluch, M. High Pressure Studies on Structural and Secondary Relaxation Dynamics in Silyl Derivative of D-Glucose. J. Chem. Phys. 2017, 147 (6), 064502. (70) Viel, Q.; Delbreilh, L.; Coquerel, G.; Petit, S.; Dargent, E. Molecular Mobility of an Amorphous Chiral Pharmaceutical Compound: Impact of Chirality and Chemical Purity. J. Phys. Chem. B 2017, 121 (32), 7729–7740.

ACS Paragon Plus Environment

36

Page 37 of 50 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

793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839

Molecular Pharmaceutics

(71) Ruiz, G. N.; Romanini, M.; Barrio, M.; Tamarit, J. L.; Pardo, L. C.; Macovez, R. Relaxation Dynamics vs Crystallization Kinetics in the Amorphous State: The Case of Stiripentol. Mol. Pharm. 2017. (72) Broadband Dielectric Spectroscopy; Kremer, F., Schönhals, A., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2003. (73) Williams, G.; Watts, D. C. Non-Symmetrical Dielectric Relaxation Behaviour Arising from a Simple Empirical Decay Function. Trans. Faraday Soc. 1970, 66 (0), 80–85. (74) Vogel, H. The Law of the Relation between the Viscosity of Liquids and the Temperature. Phys. Z. 1921, 22, 645. (75) Fulcher, G. S. Analysis of Recent Measurements of the Viscosity of Glasses. J. Am. Ceram. Soc. 1925, 8 (6), 339–355. (76) Tammann, G.; Hesse, W. The Dependence of Viscosity upon the Temperature of Supercooled Liquids. Z Anorg Allg Chem 1926, 156, 245−257. (77) Dhotel, A.; Rijal, B.; Delbreilh, L.; Dargent, E.; Saiter, A. Combining Flash DSC, DSC and Broadband Dielectric Spectroscopy to Determine Fragility. J. Therm. Anal. Calorim. 2015, 121 (1), 453–461. (78) Puente, J. A. S.; Rijal, B.; Delbreilh, L.; Fatyeyeva, K.; Saiter, A.; Dargent, E. Segmental Mobility and Glass Transition of Poly(ethylene-Vinyl Acetate) Copolymers: Is There a Continuum in the Dynamic Glass Transitions from PVAc to PE? Polymer 2015, 76, 213– 219. (79) Rijal, B.; Delbreilh, L.; Saiter, A. Dynamic Heterogeneity and Cooperative Length Scale at Dynamic Glass Transition in Glass Forming Liquids. Macromolecules 2015, 48 (22), 8219–8231. (80) Esposito, A.; Delpouve, N.; Causin, V.; Dhotel, A.; Delbreilh, L.; Dargent, E. From a Three-Phase Model to a Continuous Description of Molecular Mobility in Semicrystalline Poly(hydroxybutyrate-Co-Hydroxyvalerate). Macromolecules 2016, 49 (13), 4850–4861. (81) Rodrigues, A. C.; Viciosa, M. T.; Danède, F.; Affouard, F.; Correia, N. T. Molecular Mobility of Amorphous S-Flurbiprofen: A Dielectric Relaxation Spectroscopy Approach. Mol. Pharm. 2014, 11 (1), 112–130. (82) Kaminski, K.; Wlodarczyk, P.; Hawelek, L.; Adrjanowicz, K.; Wojnarowska, Z.; Paluch, M.; Kaminska, E. Comparative Dielectric Studies on Two Hydrogen-Bonded and van Der Waals Liquids. Phys. Rev. E 2011, 83 (6). (83) Baghel, S.; Cathcart, H.; O’Reilly, N. J. Polymeric Amorphous Solid Dispersions: A Review of Amorphization, Crystallization, Stabilization, Solid-State Characterization, and Aqueous Solubilization of Biopharmaceutical Classification System Class II Drugs. J. Pharm. Sci. 2016, 105 (9), 2527–2544. (84) Gordon, K. C.; McGoverin, C. M.; Strachan, C. J.; Rades, T. The Use of Quantum Chemistry in Pharmaceutical Research as Illustrated by Case Studies of Indometacin and Carbamazepine. J. Pharm. Pharmacol. 2007, 59 (2), 271–277. (85) Borba, A.; Gómez-Zavaglia, A.; Fausto, R. Molecular Structure, Infrared Spectra, and Photochemistry of Isoniazid under Cryogenic Conditions. J. Phys. Chem. A 2009, 113 (32), 9220–9230. (86) Young, D. C. Density Functional Theory. In Computational Chemistry; John Wiley & Sons, Inc., 2001; pp 42–48. (87) Maniruzzaman, M.; Morgan, D. J.; Mendham, A. P.; Pang, J.; Snowden, M. J.; Douroumis, D. Drug?polymer Intermolecular Interactions in Hot-Melt Extruded Solid Dispersions. Int. J. Pharm. 2013, 443 (1–2), 199–208.

ACS Paragon Plus Environment

37

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

840 841 842 843 844 845 846

Page 38 of 50

(88) Meng, F.; Trivino, A.; Prasad, D.; Chauhan, H. Investigation and Correlation of Drug Polymer Miscibility and Molecular Interactions by Various Approaches for the Preparation of Amorphous Solid Dispersions. Eur. J. Pharm. Sci. 2015, 71, 12–24. (89) Wang, B.; Wang, D.; Zhao, S.; Huang, X.; Zhang, J.; Lv, Y.; Liu, X.; Lv, G.; Ma, X. Evaluate the Ability of PVP to Inhibit Crystallization of Amorphous Solid Dispersions by Density Functional Theory and Experimental Verify. Eur. J. Pharm. Sci. 2017, 96, 45–52.

ACS Paragon Plus Environment

38

Page 39 of 50 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 1. Developed formulae of the Biclotymol molecule (A) and Haworth projection of chemical structures of α-Pentaacetylglucose (B) and β-Pentaacetylglucose (β-acGLU) (C). 163x115mm (150 x 150 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 2. X-Ray powder diffraction patterns obtained for α-acGLU (blue line), β-acGLU (red line) and Biclotymol (purple line) (A) before melt-quenching and (B) after melt-quenching. Beam cut-off was located close to 6.5° (2θ). 272x208mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 40 of 50

Page 41 of 50 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 3. TM-DSC measurements of crystalline and amorphous forms of Biclotymol (top panel), α-acGLU (left panel) and β-acGLU (right panel). (1) Melting of the commercial stable form, (2) glass transition region, (3) crystallization of the supercooled melt to the crystalline form, (4) melting of the crystalline form. 272x208mm (300 x 300 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 4. X-ray Powder Diffraction patterns for melt-quenched representative mixtures: (left panel) meltquenched Biclotymol-α-acGLU and (right panel) melt-quenched Biclotymol-β-acGLU. XRPD analysis was performed directly after the production of samples. 272x208mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 50

Page 43 of 50 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 5. TM-DSC thermograms obtained for melt-quenched Biclotymol-α-acGLU (left panel) and meltquenched Biclotymol-β-acGLU (right panel) solid dispersions during heating. 272x208mm (300 x 300 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 6. X-ray Powder Diffraction patterns for melt-quenched representative mixtures after 4 months of storage at 293 K, 30% RH: (left panel) melt-quenched Biclotymol-α-acGLU and (right panel) melt-quenched Biclotymol-β-acGLU. 272x208mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 44 of 50

Page 45 of 50 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 7. Imaginary (ε″) parts of the complex dielectric permittivity vs frequency in the glassy (T < Tg) and supercooled liquid states (T > Tg) for (a) pure α-acGLU, (b) β-acGLU and (c) pure Biclotymol. Left vertical axe corresponds to the imaginary permittivity of the α-process while right vertical axe represents the γprocess. 272x208mm (300 x 300 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 8. Dielectric loss spectra of representative amorphous solid dispersions between Biclotymol and αacGLU, analyzed above and below Tg. (a) Biclotymol - α-acGLU 1:5 (b) Biclotymol - α-acGLU 1:2 (c) Biclotymol - α-acGLU 1:1 (d) Biclotymol - α-acGLU 2:1 (e) Biclotymol - α-acGLU 5:1. Left vertical axe corresponds to the imaginary permittivity of the α-process while right vertical axe represents the γ-process. 272x208mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 46 of 50

Page 47 of 50 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 9. Dielectric loss spectra of representative amorphous solid dispersions between Biclotymol and βacGLU, analyzed above and below Tg. (a) Biclotymol - β-acGLU 1:5 (b) Biclotymol - β-acGLU 1:2 (c) Biclotymol - β-acGLU 1:1 (d) Biclotymol - β-acGLU 2:1 (e) Biclotymol - β-acGLU 5:1. Left vertical axe corresponds to the imaginary permittivity of the α-process while right vertical axe represents the γ-process. 272x208mm (300 x 300 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 10. Relaxation map of pure Biclotymol, pure α-acGLU, pure β-acGLU and amorphous solid dispersions. Solid gray lines are the VTF fits to the of structural α-relaxation times, illustrated by filled symbols. Temperature dependence of the secondary γ-relaxations is illustrated by open symbols and was fitted by the Arrhenius equation. 288x201mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 48 of 50

Page 49 of 50 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 11. Temperature dependence of the dielectric strength for α- relaxation (∆εα) versus temperature for all investigated amorphous solid dispersions. Left panel a) corresponds to Biclotymol-α-acGLU blends while right panel b) matches with Biclotymol-β-acGLU mixtures. 272x208mm (300 x 300 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 12. Optimized structures of Biclotymol-α-acGLU (upper panel) and Biclotymol-β-acGLU (lower panel) and corresponding complexation energies. 272x208mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 50 of 50