Identifying Slip Planes in Organic Polymorphs by Combined Energy

Feb 14, 2018 - Here, we explored the feasibility of more reliably identifying slip planes by the energy framework approach, combined with analysis of ...
0 downloads 5 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Article

Identifying slip planes in organic polymorphs by combined energy framework calculations and topology analysis Chenguang Wang, and Changquan Calvin Sun Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.8b00202 • Publication Date (Web): 14 Feb 2018 Downloaded from http://pubs.acs.org on February 15, 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.

Crystal Growth & Design 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 22 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

Crystal Growth & Design

1 2

4

Identifying slip planes in organic polymorphs by combined energy framework calculations and topology analysis

5 6 7

Chenguang Wang and Changquan Calvin Sun*

3

8 9 10

Pharmaceutical Materials Science and Engineering Laboratory, Department of Pharmaceutics, College of Pharmacy, University of Minnesota, Minneapolis, MN 55455, USA

11 12 13 14 15 16 17 18

*Corresponding author

19

Changquan Calvin Sun, Ph.D.

20

9-127B Weaver-Densford Hall

21

308 Harvard Street S.E.

22

Minneapolis, MN 55455

23

Email: [email protected]

24

Tel: 612-624-3722

25

Fax: 612-626-2125

26

1 ACS Paragon Plus Environment

Crystal Growth & Design 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 2 of 22

27

Abstract

28

The relationship between crystal structure and mechanical properties is commonly studied by

29

identifying slip planes through inspecting crystal structures visually, with a focus on the hydrogen

30

bonding interactions. While useful, the visualization method lacks quantitative insight and the

31

identification of slip planes is sometimes subjective. Sometimes, crystal plasticity predicted from

32

structure visualization does not match experimental crystal plasticity and powder tabletability as

33

observed in three polymorphic systems, i.e., 6-chloro-2,4-dinitroaniline, indomethacin, and

34

febuxostat. Here, we explored the feasibility of more reliably identifying slip planes by the energy

35

framework approach, combined with the analysis of potential slip layer topology. In all three cases,

36

this new approach identified slip planes that are consistent with the observed mechanical plasticity

37

and compaction behavior. Thus, it is superior to the visualization method for crystal structure

38

analysis aimed at identifying active slip planes in organic crystals.

39

40

KEY WORDS: :structure – mechanical property, polymorph, slip plane, intermolecular interaction

41

2 ACS Paragon Plus Environment

Page 3 of 22 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

42

Crystal Growth & Design

Introduction

43

Polymorphism, the phenomenon of crystalline materials with identical molecular

44

composition but different internal crystal structures,1 plays an important role in the development

45

and commercialization of fine chemicals, including pharmaceuticals, pigments, herbicides, foods,

46

and explosives.2,

47

distinct properties, such as solubility, dissolution rate, bioavailability, stability, and mechanical

48

properties, which impact drug product development and manufacturing.2, 4 Polymorphs are also

49

suitable for probing crystal structure-mechanical property relationship, which is an active area of

50

research for crystal engineers, chemists, and pharmaceutical scientists. 5-12

3

Polymorphs are of interest to pharmaceutical industry because they display

51

The bonding area and bonding strength (BABS) model suggests that higher crystal plasticity

52

(i.e., lower hardness) favors larger bonding area and, therefore, stronger tablets during powder

53

compaction.13, 14 Under an external compaction stress, crystals initially always undergo reversible

54

elastic deformation but irreversible deformation, such as brittle fracture or plastic yield, takes place

55

once the elastic limit is exceeded.15 High crystal plasticity, i.e., low hardness, corresponds to facile

56

plastic deformation.11, 16 The presence of slip planes is necessary for plastic deformation to occur in

57

crystals, where the lower interaction energy and smoother surfaces of slip planes favor more facile

58

slip.17

59

hydrogen bonding patterns and molecular packing density of layers.18, 19 Stacking flat layers or

60

parallel columns with smooth surfaces that interact weakly are crystal packing features that

61

correspond to facile plastic deformation of crystals (Figure 1). Crystals with such structures were

62

observed to exhibit good compressibility and tabletability, if bonding strength (collective interaction

63

strength between two adjacent particles over unit bonding area) is comparable among crystals.6, 20-22

Slip planes are commonly identified by visualizing crystal structures, which relies on

3 ACS Paragon Plus Environment

Crystal Growth & Design 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

(a)

Page 4 of 22

(b)

64 65 66 67

Figure 1. Crystal packing features that favor facile plastic deformation. a) hexachlorobenzene (refcode, HCLBNZ14) with parallel columns and 2) acetaminophen form II (refcode, HXACAN) with flat layered structure. Possible slip directions are indicated by arrows.

68 69

Crystal structure analysis based on visualization is qualitative and objective when obvious

70

layers cannot be unambiguously identified. Therefore, crystal structure visualization has led to

71

some difficulties in explaining crystal mechanical properties and tabletability in several polymorph

72

systems.7,

73

dimensional (2D) layers without hydrogen bonds between them. However, the tabletability of Form

74

I is unexpectedly much lower than Form II.7 For indomethacin, crystal structure visualization

75

positively identified slip planes (0 1 1) in γ form but not in α form.

76

to exhibit better tabletability than γ form, which is inconsistent with the expected behavior based on

77

the BABS model if γ form were indeed more plastic.23 For febuxostat polymorphs, an active slip

78

plane was identified in Form Q but absent in Form H1.24 This is inconsistent with the high

79

plasticity of both crystals suggested by their low nanoindentation hardness (0.310 GPa for Form H1

23, 24

Both Forms I and II of 6-chloro-2,4-dinitroaniline (CDNA) exhibited flat two

23

However, α form was found

4 ACS Paragon Plus Environment

Page 5 of 22 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

Crystal Growth & Design

80

and 0.214 GPa for Form Q).

The discrepancies between experimentally observed

81

plasticity/tabletability and that expected from structure visualization of these polymorphs were

82

either not discussed7 or rationalized by considering factors that influence layer slip, e.g., presence of

83

weak hydrogen bonds,24, 25 or different bonding strength inferred from true density.23

84

An alternative approach to address these discrepancies is to re-examine the accuracy of slip

85

plane identification by the visualization method. One main drawback in the visualization method is

86

its reliance on sorting molecules based on the conventional hydrogen bonds.17 To identify slip

87

planes by visualization, the following assumptions are made: 1) layers with the largest separation

88

distance exhibit weakest interlayer interactions, hence, easier molecular slip along the layer; 2) 2D

89

hydrogen bonded layers are rigid and slipping along them is energetically more favored than

90

crossing them; 3) rough layer topology hinders facile interlayer slip, hence, lowers crystal plasticity.

91

While the last assumption about layer topology is correct, the first two assumptions are questionable

92

because weak hydrogen bonds, such as C-H···O, are usually not considered during visualization.25

93

In addition, because of geometrical sensitivity, the interaction strength cannot be accurately

94

assessed by visualization even when they are considered. Moreover, other important interactions,

95

e.g., halogen bonding, CH-π, and π-π interactions, are usually not graphically presented when

96

identifying slip planes by visualization. Consequently, visualization may not always accurately

97

identify slip planes in a crystal, especially if molecular layers in some crystals cannot be easily

98

visualized. Another common method for identifying slip planes is based on attachment energy, Eatt,

99

which is defined as energy per mole of molecules released when one slice of thickness dhkl attaches

100

onto a crystal face (hkl).26 The primary slip plane is assumed to be the crystallographic planes with

101

the lowest Eatt, i.e., the interactions between slip planes are the weakest. The major drawback of

102

this approach is that it does not consider layer topology. Therefore, planes with the lowest Eatt may

103

not easily slip when they are corrugated due to physical hindrance by adjacent planes.11 Obviously, 5 ACS Paragon Plus Environment

Crystal Growth & Design 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 22

104

simultaneous consideration of Eatt along with structure visualization is better than either method

105

alone. However, slip planes predicted by the two methods sometimes are at odds,18 which makes it

106

difficult to finalize slip plane assignment in those cases. Crystal packing may also be described

107

using “energy framework” based on pairwise intermolecular interaction energy by density

108

functional theory (DFT) calculation, which is graphically represented by cylinders connecting the

109

centers of mass of the two molecules.27, 28 The calculated intermolecular interaction energy may be

110

divided into the electrostatic, polarization, dispersion and exchange-repulsion components and

111

properly scaled based on a large training set.29 It has been successfully utilized to rationalize the

112

mechanical behavior of several molecular crystals, such as bending hexachlorobenzene and 3,4-

113

dichlorophenol crystals, as well as the remarkably different tabletability of two paracetamol

114

polymorphs.27 This method yields more accurate energy calculations than that based on force

115

fields, which is commonly used in calculating Eatt. A drawback of the DFT calculation is the longer

116

calculation time, which can be compensated by using faster computers. Another useful method that

117

also uses DFT calculation is Gavezzotti’s PIXEL method,30 which can also be graphically presented

118

to characterize intermolecular interaction topology. 31

119

Herein, we explored the potential application of the “energy framework” approach in

120

identifying slip planes using CDNA, indomethacin, and febuxostat polymorphs (Figure 2). We

121

assumed that slip planes are molecular layers exhibiting the lowest interlayer interaction energy and

122

ease of sliding is allowed by layer topology. The correctness of the predicted slip planes was

123

checked against the known mechanical properties and/or tableting behavior of those organic

124

crystals.

6 ACS Paragon Plus Environment

Page 7 of 22 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

Crystal Growth & Design

(a)

Form I

Form II

Form III

c

c

c

a

a a

Form α (b)

Form γ

b a

b c Form Q

(c)

Form H1

b

b

c

c

125 126

Figure 2. Crystal unit cell of a) CDNA, b) indomethacin, and c) febuxostat polymorphs.

127

128

Method

129

Crystal structure visualization

130

CIFs of selected crystal structures were downloaded from Cambridge Crystallographic

131

Structural Database using ConQuest (V. 1.19, CCDC, Cambridge, UK) and WebCSD. Crystal

132

structures were visually examined to identify crystallographic planes using Mercury CSD (V. 3.9,

7 ACS Paragon Plus Environment

Crystal Growth & Design 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 22

133

CCDC, Cambridge, UK). This effort was aided by showing hydrogen bonds that meet the default

134

criteria given in the software.

135

Attachment energy calculation

136

Attachment energy was calculated using the Dreiding force field and Qeq charges at fine

137

quality available in the crystal growth module of the Materials Studio 7.0. (Accelrys Software Inc.,

138

San Diego, CA, USA). The “Ewald” electrostatic summation method and “atom based” van der

139

Waals summation was chosen. In addition, a minimum dhkl was set at 1.0 Å.18

140

Energy framework

141

The intermolecular interaction energy calculation was performed using dispersion-corrected

142

DFT at the B3LYP-D2/6-31G(d,p) level of theory (CrystalExplorer V.17). During the calculation,

143

hydrogen positions were normalized to standard neutron diffraction values.

144

energies of a chosen molecule with all molecules having any atom within its 3.8 Å were calculated.

145

Subsequently, all interaction energies within and between an identified layer or column were

146

manually added to arrive at the total intra-layer, inter-layer, or inter-column interaction energies.

147

For a crystal structure containing multiple molecules in the asymmetric unit, different total

148

interaction energies were obtained. In that case, an arithmetic mean value is used for ease of

149

comparison. The “energy framework” was constructed based on the crystal symmetry and total

150

intermolecular interaction energy, which included electrostatic, polarization, dispersion, and

151

exchange-repulsion components with scale factors of 1.057, 0.740, 0.871, and 0.618, respectively.29

152

The interactions energies below a certain energy threshold are omitted for clarity and cylinder

153

thickness is proportional to the intermolecular interaction energies. It should be pointed out that,

154

although used to quantify intermolecular interaction energy in a periodic structure, the DFT

155

calculations are not periodic structure calculations.

32

Then, the interaction

Where needed, interlayer or intralayer 8

ACS Paragon Plus Environment

Page 9 of 22 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

Crystal Growth & Design

156

interaction energies were calculated by manually adding interaction energies between a given

157

molecule in one layer and all interacting molecules in a neighboring layer or within the same layer,

158

respectively (Figure S1).

159

160

Results and discussion

161

6-chloro-2,4-dinitroaniline (CDNA) polymorphs

162

Three CDNA polymorphs exhibited distinctive shear (Form I, UCECAG01), bending (Form

163

II, UCECAG02), and brittle (Form III, UCECAG03) behavior.12 The slip planes predicted by

164

visualization, attachment energy calculation, and energy framework approaches, as well as the

165

mechanical behavior of three CDNA polymorphs are summarized in Table 1. The monoclinic Form

166

I CDNA consists of stacking flat (1 0 -1) planes (Figure 3a). Molecules in (1 0 -1) plane are

167

hydrogen bonded (N-H···O=N, 3.032 Å; 3.054 Å), whereas no hydrogen bonds are present between

168

these planes. Therefore, (1 0 -1) plane is identified as the primary slip plane based on visualization.

169

However, the (1 0 0) plane is found to have the lowest attachment energy and corresponded to the

170

largest surface in the predicted crystal morphology (Table S2). For the (1 0 -1) plane, the energy

171

framework analysis shows comparable intra- and inter- layer intermolecular bonding energies

172

(Table S1, Figure 3b). In fact, the total intralayer intermolecular bonding energies (-84.0 kJ/mol)

173

are actually lower than that between layers (-106.4 kJ/mol). Therefore, sliding along (1 0 -1) plane

174

is energetically unfavorable. In contrast, for (1 0 0) plane, the total interlayer interaction energy (-

175

84.8 kJ/mol) is significantly lower than the total intralayer interaction energies (-106.1 kJ/mol),

176

implying the molecule slide along the (1 0 0) is energetically more favorable (Figure 3b).

177

Molecules sliding along the (1 0 0) plane is also topologically feasible. Thus, (1 0 0) is the primary

178

slip plane for CNDA Form I based on the energy framework approach. This is consistent with the 9 ACS Paragon Plus Environment

Crystal Growth & Design 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 10 of 22

179

experimental observation that the block morphology of Form I crystal is sheared along the

180

dominating (1 0 0) facet instead of (1 0 -1) facet.

181

The structure of monoclinic Form II CDNA consists of 2D thick layers parallel to the (0 0 1)

182

plane (Figure 3c). Molecules within the layers are connected through two N-H···O=N (2.938 Å;

183

3.066 Å) hydrogen bonds. Layers interact mainly through weaker Cl···Cl (3.996 Å) halogen

184

bonding and C-H···O=N (3.382 Å) hydrogen bonds. The notably larger intralayer (-148.8 kJ/mol)

185

interaction energy than the interlayer (-30.2 kJ/mol) interaction energy indicates that (0 0 1) is the

186

active slip plane (Table S1), which is consistent with its lowest attachment energy (Tables S2).

187

The structure of triclinic Form III CDNA consists of 2D interlocked layers (Figure 3e).

188

Both visualization and attachment energy methods suggested (0 0 1) as the slip plane (Table S2).

189

This was confirmed by the much larger total intralayer interaction energies (-148.6 kJ/mol) than

190

interlayer total interaction energies (-69.6 kJ/mol) (Figure 3f, Table S1). However, the topological

191

feature of (0 0 1) plane restricted the slip direction to be along the b direction only. This means that

192

slip in Form III CDNA is not facile.

193

The best tabletability of Form II among the three polymorphs is attributed to the ease of

194

molecules sliding along the slip planes, which exhibit much weaker interlayer interaction energies

195

compared with other polymorphs. Although flat slip planes are identified in both Forms I and II,

196

the lower interlayer bonding energies of Form II, which corresponds to higher plasticity, explains its

197

superior tabletability than Form I. The worst tabletability of Form III in this polymorph system is

198

attributed to the restricted mobility of the primary slip planes due to unfavorable layer topology.7

199

200

10 ACS Paragon Plus Environment

Page 11 of 22 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

Crystal Growth & Design

(a)

(10-1)

(c) (001)

(e)

(100)

(b)

(d)

(f) (001)

201 202 203 204 205

Figure 3. Crystal packing patterns (left) and corresponding energy frameworks (right) for CDNA polymorphs, a, b) Form I, c, d) Form II, and e, f) Form III. A likely slip layer by visualization is shaded in blue and that by energy framework is shaded in pink. The thickness of each cylinder (in blue) represents the relative strength of interaction. The energy threshold for the energy framework is set at -15 kJ/mol.

11 ACS Paragon Plus Environment

Crystal Growth & Design 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 22

206 207 208

Table 1. Summary of mechanical behavior and slip planes predicted by visualization, attachment energy, and energy framework approaches of three CNDA polymorphs. Slip planes

Form I

CSD

Visualization

refcode

(ref. 32)

UCECAG01

(1 0 -1)

Energy

Crystal Mechanical

framework

behavior

(1 0 0)

sheared along the largest

(hkl)/Eatt (kCal/mol)

(1 0 0)/-71.44

facet Form II

UCECAG02

(0 0 1)

(0 0 1)/ -14.15

(0 0 1)

Bending

Form III

UCECAG03

(0 0 1)

(0 0 1)/-45.12

(0 0 1)

Brittle

209 210 211

Indomethacin polymorphs

212

Molecules in indomethacin monoclinic Form α (INDMET02) are connected through O-

213

H···O (2.593 Å, 2.704 Å, and 2.735 Å) hydrogen bonds. Visualization suggests a lack of obvious

214

crystallographic slip planes (Figure 4a).23 However, a column structure could be identified with the

215

aid of the energy framework (Figure 4b, Table S1). The inter-columnar molecular interaction

216

energies (-41.7 kJ/mol) is significantly lower than intra-columnar interaction energies (-254.7

217

kJ/mol). Therefore, in contrast to the lack of slip planes predicted based on visualization, multiple

218

active slip planes parallel to a axis are identified by the energy framework approach. Among those

219

potential slip planes, (0 0 1) has the lowest interlayer interaction energy, which makes it the most

220

energetically favorable plane for molecules to slide under an external shear stress. This is also

221

consistent with the lowest attachment energy of (0 0 1) (Table S2). 12 ACS Paragon Plus Environment

Page 13 of 22 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

Crystal Growth & Design

222

For indomethacin triclinic Form γ (INDMET03), visualization suggested (0 1 1) as the

223

primary slip plane (Figure 4c).23 Within the (0 1 1) layer, indomethacin molecules formed dimers

224

through O-H···O (2.651 Å) hydrogen bonds (Figure 4c).

225

interaction energy between adjacent (0 1 1) layers (-189.6 kJ/mol) is much higher than that within

226

the layers (-54.9 kJ/mol) (Figure 4d), suggesting that molecules slipping along the (0 1 1) plane is

227

energetically unfavorable. An inspection of the energy components revealed that C-H···O (3.397

228

Å), CH- π, and π-π interactions contributed significantly to the total interaction energy between (0 1

229

1). Such contributions are neglected when identifying slip planes by the visualization method.

230

Instead, the (0 0 1) plane is likely the primary slip plane because the total intralayer bonding energy

231

(-164.8 kJ/mol) is much higher than the interlayer bonding energy (-86.0 kJ/mol). This assignment

232

of slip plane is consistent with the lowest attachment energy of (0 0 1) planes (Table S2). However,

233

molecular slip along (0 0 1) is difficult due to the rough topology of the (0 0 1) layers (Figure 4c).

234

Table 2 shows a summary of the slip planes predicted by visualization, attachment energy, and

235

energy framework approaches, as well as the tableting behavior of two indomethacin polymorphs.

236

Therefore, the analysis of crystal structures using energy framework combined with layer topology

237

consideration led to the conclusion that indomethacin Form α is more plastic than Form γ, a

238

prediction that is consistent with the superior tabletability of Form α.23

239 240

Table 2. Summary of tableting behavior and slip planes, predicted by visualization, attachment energy, and energy framework approaches, of two indomethacin polymorphs.

Surprisingly, strong intermolecular

Slip plane Visualization (ref. 23) (hkl)/Eatt (kCal/mol) Energy framework CSD refcode Tabletability Form α

None

(0 0 1)/-44.67

(0 0 1)

INDMET02

Higher

Form γ

(0 1 1)

(0 0 1)/-54.41

(0 0 1)

INDMET03

Lower

241 13 ACS Paragon Plus Environment

Crystal Growth & Design 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 22

242

(a)

(b)

(c)

(d)

(011)

(001)

243 244 245 246 247

Figure 4. Crystal packing patterns (left) and corresponding energy framework (right) for indomethacin polymorphs, a, b) α form, and c, d) γ form viewed along the a axis. Molecular layers based on visualization method are shaded in blue and those by energy framework are in pink. The green lines in (c) outline a molecular layer identified from the energy framework. The energy threshold for the energy framework is set at -25 kJ/mol.

248 249 250

Febuxostat polymorphs

251

Visualization suggested (-1 0 4) as slip plane in monoclinic febuxostate Form Q (HIQQAB).

252

Molecules within this layer are connected through C-H···N≡C (2.782 Å) hydrogen bonds while no

253

hydrogen bond is present between layers (Figure 5a).24 However, the energy framework suggests 14 ACS Paragon Plus Environment

Page 15 of 22 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

Crystal Growth & Design

254

the molecules stacking cross (-1 0 4) exhibit stronger interaction energy (-180.4 kJ/mol) than the

255

bonding energy within the (-1 0 4) plane (-114.1 kJ/mol) (Figure 5b, Table S1). Therefore, sliding

256

along the (-1 0 4) plane is energetically unfavorable even (-1 0 4) planes are smooth. The energy

257

framework of Form Q revealed a columnar structure, instead of layer structure, where the intra-

258

columnar interaction energy (-155.4 kJ/mol) is higher than inter-columnar interaction energy (-

259

139.1 kJ/mol) (Figure 5c).

260

planes based on energy consideration.

Therefore, planes (0 1 1), (0 1 -1), (0 0 1) and (0 1 0) are likely slip

261

In triclinic febuxostat Form H1 (HIQQAB02), molecules formed dimers via O-H···O (2.624

262

Å) hydrogen bonds, which are further connected through C-H···N≡C (3.536 Å) and C-H···O (3.408

263

Å) hydrogen bonds along (2 0 1) (Figure 5d). In an effort to explain the inferior plasticity of Form

264

H1, it was suggested that the sulfonamide groups participate in the interlayer interaction, thus

265

immobilizing the layers.24 However, the energy framework results showed significant dispersive

266

energy due to π- π interaction stacked along the (2 0 1) planes (Table S1). Consequently, the

267

intralayer interaction energies (-115.4 kJ/mol) are weaker than the interlayer interaction energies (-

268

183.2 kJ/mol), suggesting the molecules slip along the (2 0 1) plane is energetically unfavorable

269

(Figure 5e). Instead, the (0 1 0) plane is energetically more favored slip plane (Figure 5f). This is

270

consistent with the lowest attachment energy of this crystal facet (Table S2). The intermolecular

271

bonding energies within (0 1 0) (-257.7 kJ/mol) are significantly higher than those between (0 1 0)

272

(-31.5 kJ/mol), which suggests that it is energetically much more favorable for molecules to slid

273

along (0 1 0). Thus, the Form H1 consists of stacking thick 2D (0 1 0) layers, which could serve as

274

slip planes. The slip planes predicted by visualization, attachment energy calculation, and energy

275

framework approaches, as well as the mechanical properties of the two febuxostat polymorphs are

276

summarized in Table 2. Therefore, both Forms Q and H1 are expected to be plastic owing to the

277

slip planes identified by the energy framework based approach. This is in accordance with the 15 ACS Paragon Plus Environment

Crystal Growth & Design 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 22

278

higher plasticity (i.e., low hardness) and excellent compressibility and tabletability of both forms.24

279

The higher plasticity of Form Q than Form H1 corresponds to its columnar structure, which leads to

280

multiple slip systems and more facile plastic deformation than Form H1 featured with 2D thick

281

layers.

282

283 284

Table 3. Summary of mechanical properties and slip planes, predicted by visualization, attachment energy, and energy framework approaches, of two febuxostat polymorphs Slip plane

Form Q

Form H1 285 286

a.

CSD

Visuali-

(hkl)/Eatt

refcode

zation

(kCal/mol)

HIQQAB

(-1 0 4)

(0 1 1)/-32.81

HIQQAB02

(2 0 1) a

(0 1 0)/-22.10

Energy framework

Mechanical properties

(0 1 1), (0 1 -1), (0 0 1),

Plastic crystal with lower H

(0 1 0)

and higher tabletability

(0 1 0)

Plastic crystal

No active slip plane was identified in Form H1 in ref. 24.

287

16 ACS Paragon Plus Environment

Page 17 of 22 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

Crystal Growth & Design

(a)

(c)

(-104)

(001) (011)

(01-1)

(b) (010)

(d) (201)

(f)

(010)

(e)

288 289 290 291 292 293

Figure 5. Crystal packing patterns (a, d) and corresponding energy framework (b, c, e, f) for febuxostat polymorphs. a, b, c) Form Q, and d, e, f) Form H1. Molecular layers based on visualization are shaded in light blue and molecular columns and layers based on energy framework are shaded in pink. The thickness of each cylinder (in blue) represents the relative strength of interaction. The energy threshold for the energy framework is set at -20 kJ/mol.

294 295

296 297

Conclusions Qualitative analysis of crystal structure based on hydrogen bond analysis faced difficulty in explaining the mechanical properties and tableting behavior of three polymorph systems.

In

17 ACS Paragon Plus Environment

Crystal Growth & Design 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 22

298

contrast, the quantitative crystal structure by the energy framework approach successfully explained

299

experimental observations in all cases.

300

The higher plasticity of indomethacin Form α than γ and febuxostat Form Q than H1 is

301

explained by multiple active slip planes corresponding to their parallel columnar structures. Slip

302

planes with smooth layer topology were identified in CDNA Forms I and II, febuxostat Forms Q

303

and H1 forms. This is consistent with their low hardness and absence of tableting problem. The

304

relatively lower tabletability of CDNA Form I than Form II corresponds to its lower plasticity

305

because of the higher interlayer bonding energies between flat layers. The rough topology of

306

CDNA Form III and indomethacin Form γ hinders facile slip, which led to poor plasticity and

307

tabletability.

308

The visualization method accounts for hydrogen bonding but overlooks the large variability

309

of hydrogen bonds strength and contributions by other types of intermolecular interactions, e.g.,

310

dispersive interactions. Consequently, slip planes can be erroneously identified by visualization

311

alone. This, in turn, leads to difficulty in explaining mechanical properties and tableting behavior

312

of crystals in the three model polymorph systems. In contrast, the energy framework approach led

313

to correct identification of slip planes in all model crystals. Along with the consideration of slip

314

plane topology, such information was successfully used to clearly explain experimentally observed

315

mechanical behavior and tableting performance. The attachment energy approach could correctly

316

identify slip planes but inherently incapable of revealing structural features that promote molecular

317

slip, such as the slip columns, in certain crystals. Overall, the approach of combined energy

318

framework and topological analysis led to more accurate identification of slip planes and more

319

reliable predictions of crystal mechanical properties. If proven robust in a much larger set of

320

crystals, this approach is expected to be an important tool for the future research aimed at

321

establishing crystal structure – mechanical property - tabletability relationship. 18 ACS Paragon Plus Environment

Page 19 of 22 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

Crystal Growth & Design

322

323

Supporting Information

324

Intermolecular interaction energies, attachment energies and crystallographic parameters of CDNA,

325

indomethacin, and febuxostat polymorphs.

326

327

Acknowledgments

328

We thank the Minnesota Supercomputing Institute (MSI) at the University of Minnesota for

329

providing resources that contributed to the research results reported in this paper.

330

331

References

332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352

1. Vippagunta, S. R.; Brittain, H. G.; Grant, D. J. Crystalline solids. Adv. Drug Deliv. Rev. 2001, 48, (1), 3-26. 2. Datta, S.; Grant, D. J. Crystal structures of drugs: advances in determination, prediction and engineering. Nat. Rev. Drug Discov. 2004, 3, (1), 42. 3. Yu, L. Polymorphism in molecular solids: an extraordinary system of red, orange, and yellow crystals. Acc. Chem. Res. 2010, 43, (9), 1257-1266. 4. Gardner, C. R.; Walsh, C. T.; Almarsson, Ö. Drugs as materials: valuing physical form in drug discovery. Nat. Rev. Drug Discov. 2004, 3, (11), 926. 5. Maahs, A. C.; Ignacio, M. G.; Ghazzali, M.; Soldatov, D. V.; Preuss, K. E. Chiral crystals of an achiral molecule exhibit plastic bending and a crystal-to-crystal phase transition. Cryst. Growth Des. 2017, 17, (3), 1390-1395. 6. Joiris, E.; Di Martino, P.; Berneron, C.; Guyot-Hermann, A.-M.; Guyot, J.-C. Compression behavior of orthorhombic paracetamol. Pharm. Res. 1998, 15, (7), 1122-1130. 7. Bag, P. P.; Chen, M.; Sun, C. C.; Reddy, C. M. Direct correlation among crystal structure, mechanical behaviour and tabletability in a trimorphic molecular compound. CrystEngComm 2012, 14, (11), 3865-3867. 8. Sun, C.; Grant, D. J. Influence of crystal structure on the tableting properties of sulfamerazine polymorphs. Pharm. Res. 2001, 18, (3), 274-280. 9. Mondal, P. K.; Kiran, M.; Ramamurty, U.; Chopra, D. Quantitative investigation of the structural, thermal, and mechanical properties of polymorphs of a fluorinated amide. Chem. Eur. J. 2017, 23, (5), 1023-1027.

19 ACS Paragon Plus Environment

Crystal Growth & Design 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

353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398

Page 20 of 22

10. Upadhyay, P.; Khomane, K. S.; Kumar, L.; Bansal, A. K. Relationship between crystal structure and mechanical properties of ranitidine hydrochloride polymorphs. CrystEngComm 2013, 15, (19), 3959-3964. 11. Mishra, M. K.; Ramamurty, U.; Desiraju, G. R. Solid solution hardening of molecular crystals: tautomeric polymorphs of omeprazole. J. Am. Chem. Soc. 2015, 137, (5), 1794-1797. 12. Reddy, C. M.; Basavoju, S.; Desiraju, G. R. Sorting of polymorphs based on mechanical properties. Trimorphs of 6-chloro-2, 4-dinitroaniline. Chem. Comm. 2005, (19), 2439-2441. 13. Sun, C. C. Decoding powder tabletability: roles of particle adhesion and plasticity. J. Adhes. Sci. Technol. 2011, 25, (4-5), 483-499. 14. Osei-Yeboah, F.; Chang, S.-Y.; Sun, C. C. A critical examination of the phenomenon of bonding area-bonding strength interplay in powder tableting. Pharm. Res. 2016, 33, (5), 1126-1132. 15. Varughese, S.; Kiran, M.; Ramamurty, U.; Desiraju, G. R. Nanoindentation in crystal engineering: quantifying mechanical properties of molecular crystals. Angew. Chem. Int. Ed. 2013, 52, (10), 2701-2712. 16. Mishra, M. K.; Ramamurty, U.; Desiraju, G. R. Mechanical property design of molecular solids. Curr. Opin. Solid State Mater. Sci. 2016, 20, (6), 361-370. 17. Zolotarev, P. N.; Moret, M.; Rizzato, S.; Proserpio, D. M. Searching new crystalline substrates for OMBE: topological and energetic aspects of cleavable organic crystals. Cryst. Growth Des. 2016, 16, (3), 1572-1582. 18. Sun, C. C.; Kiang, Y. H. On the identification of slip planes in organic crystals based on attachment energy calculation. J. Pharm. Sci. 2008, 97, (8), 3456-3461. 19. Sun, C. C., Role of surface free energy in powder behavior and tablet strength. In Adhesion in Pharmaceutical, Biomedical, and Dental Fields, 2017; pp 75-88. 20. Chattoraj, S.; Shi, L.; Sun, C. C. Understanding the relationship between crystal structure, plasticity and compaction behaviour of theophylline, methyl gallate, and their 1: 1 co-crystal. CrystEngComm 2010, 12, (8), 2466-2472. 21. Perumalla, S. R.; Shi, L.; Sun, C. C. Ionized form of acetaminophen with improved compaction properties. CrystEngComm 2012, 14, (7), 2389-2390. 22. Sun, C. C.; Hou, H. Improving mechanical properties of caffeine and methyl gallate crystals by cocrystallization. Cryst. Growth Des. 2008, 8, (5), 1575-1579. 23. Khomane, K. S.; More, P. K.; Raghavendra, G.; Bansal, A. K. Molecular understanding of the compaction behavior of indomethacin polymorphs. Mol. Pharm. 2013, 10, (2), 631-639. 24. Yadav, J. A.; Khomane, K. S.; Modi, S. R.; Ugale, B.; Yadav, R. N.; Nagaraja, C.; Kumar, N.; Bansal, A. K. Correlating single crystal structure, nanomechanical, and bulk compaction behavior of febuxostat polymorphs. Mol. Pharm. 2017, 14, (3), 866-874. 25. Khomane, K. S.; Bansal, A. K. Weak hydrogen bonding interactions influence slip system activity and compaction behavior of pharmaceutical powders. J. Pharm. Sci. 2013, 102, (12), 42424245. 26. Hartman, P.; Bennema, P. The attachment energy as a habit controlling factor: I. Theoretical considerations. J. Cryst. Growth 1980, 49, (1), 145-156. 27. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy frameworks: insights into interaction anisotropy and the mechanical properties of molecular crystals. Chem. Comm. 2015, 51, (18), 3735-3738. 28. Mackenzie, C. F.; Spackman, P. R.; Jayatilaka, D.; Spackman, M. A. CrystalExplorer model energies and energy frameworks: extension to metal coordination compounds, organic salts, solvates and open-shell systems. IUCrJ 2017, 4, (5), 575-587.

20 ACS Paragon Plus Environment

Page 21 of 22 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

399 400 401 402 403 404 405 406 407 408 409

Crystal Growth & Design

29. Turner, M. J.; Grabowsky, S.; Jayatilaka, D.; Spackman, M. A. Accurate and efficient model energies for exploring intermolecular interactions in molecular crystals. J. Phys. Chem. Lett. 2014, 5, (24), 4249-4255. 30. Maschio, L.; Civalleri, B.; Ugliengo, P.; Gavezzotti, A. Intermolecular interaction energies in molecular crystals: comparison and agreement of localized møller–plesset 2, dispersion-corrected density functional, and classical empirical two-body calculations. J. Phys. Chem. A 2011, 115, (41), 11179-11186. 31. Bond, A. processPIXEL: a program to generate energy-vector models from Gavezzotti's PIXEL calculations. Journal of Applied Crystallography 2014, 47, (5), 1777-1780. 32. M. J. Turner, J. J. M., S. K. Wolff, D. J. Grimwood, P. R. Spackman, D. Jayatilaka and M. A. Spackman CrystalExplorer17. http://hirshfeldsurface.net

410 411

21 ACS Paragon Plus Environment

Crystal Growth & Design 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

412 413

414 415 416 417 418 419 420 421 422 423 424

Page 22 of 22

For Table of Contents Use Only

Synopsis Analysis of crystal structures using the energy framework approach correctly identified slip planes that reconciled inconsistency between observed mechanical properties and predicted plasticity based on slip plane predicted from crystal structure visualization.

22 ACS Paragon Plus Environment