Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

NJC

PAPER

Avoiding irreversible 5-fluorocytosine hydration


via supramolecular synthesis of pharmaceutical
Cite this: New J. Chem., 2018,
42, 14994 cocrystals†
a
Matheus S. Souza, Luan F. Diniz, a Lautaro Vogt, a
Paulo S. Carvalho Jr., a
ab
Richard F. D’vries and Javier Ellena *a

The antimetabolite 5-fluorocytosine (5-FC) was used to form pharmaceutical cocrystals in order to modulate
its poor physicochemical stability in humid environments, which leads to the irreversible incorporation of a
water molecule at the structural level under storage conditions. The anhydrous form of 5-FC is a well-known
fluorinated analog of cytosine with antifungal activity and it has become one of the most used medications
for cancer treatment via the gene therapy approach. In this study, novel 5-FC cocrystals were obtained from
the reaction of 5-FC with three nontoxic coformers: caffeine (CAF), p-aminobenzoic acid (PABA) and caprylic
acid (CA). These cocrystals, namely 5FC–CAF, 5FC–PABA and 5FC–CA, were characterized by single-crystal
and powder X-ray diffraction (SCXRD and PXRD), spectroscopic (FT-IR and FT-Raman) and thermal
(thermogravimetric analysis, differential scanning calorimetry, and hot-stage microscopy) techniques. The
physical stabilities of 5-FC and its cocrystals were evaluated in environments with high relative humidity
and the equilibrium solubility was measured in a pH 1.2 buffer medium. These studies show that the
prodrug 5-FC is able to form different homo and heterosynthons that lead to cocrystal formation.
Additionally, the solubility profiles of the novel multicomponent solid forms were found to be similar to
Received 30th May 2018, the API raw material, a BCS class I drug, exhibiting a high solubility profile. The hydration stabilities of 5-FC
Accepted 26th July 2018 and its cocrystals were evaluated in humid environments to confirm the irreversible hydration of 5-FC in
DOI: 10.1039/c8nj02647e contrast with the absence of phase transitions in its cocrystal forms. In this way all 5-FC cocrystals
reported herein maintained to a large degree the API solubility and do not undergo the hydration process
rsc.li/njc or phase transition under extreme storage conditions, being more stable than the parent 5-FC.

a
Instituto de Fı́sica de São Carlos, Universidade de São Paulo, CP 369, 13.560-970 – São Carlos, SP, Brazil. E-mail: javiere@ifsc.usp.br;
Tel: +55 (016)3373-8096/3373-9876
b
Facultad de Ciencias Básicas, Universidad Santiago de Cali, Calle 5 # 62-00, Cali, Valle del Cauca, Colombia
† Electronic supplementary information (ESI) available: Fig. S1: experimental powder X-ray diffraction patterns of 5-FC and its cocrystals that remain non-solubilized after
the solubility test; Fig. S2: absorption spectra of 5-FC 0.0055 mg mL1, caffeine 0.0086 mg mL1, p-aminobenzoic acid 0.0065 mg mL1; and the cocrystals 5FC–CAF
0.016 mg mL1, 5FC–PABA 0.013 mg mL1 and 5FC–CA 0.016 mg mL1 in buffer pH = 1.2; Fig. S3: heterosynthons observed in the 5-FC cocrystals; Fig. S4: homosynthons
observed in the 5-FC cocrystals; Fig. S5(a) and (c): microscope images of crystalline phases with single-crystals of different habits obtained directly from the SES method for
5FC–CAF and 5FC–PABA cocrystals, respectively, (b) and (d): grinding products recrystallized through SES as a unique crystalline phase; Fig. S6: simulated and
experimental powder X-ray diffraction patterns of 5-FC cocrystals. All diffractograms are in good agreement indicating that the samples present high purity and
representativeness; Fig. S7: experimental powder X-ray diffraction patterns of 5-FC cocrystals obtained from solvent-drop grinding (SDG) compared to simulated ones. Only
the diffractograms of 5FC–PABA are not in good agreement which suggest a possible physical mixture between the starting materials; Fig. S8: TGA and DSC curves of the
5-FC cocrystals; Fig. S9: DSC curves of: (i) the raw material 5-fluorocytosine (5-FC anhydrous); (ii) 5-FC after one week in an environment with high relative humidity (5-FC
monohydrate) and (iii) the two solid coformers, namely: CAF (caffeine) and PABA (p-aminobenzoic acid); Fig. S10: FT-Raman spectra of 5-FC and its cocrystals; Fig. S11:
powder X-ray diffraction patterns of 5-FC confirming that this API undergoes a phase transition from the anhydrous to the monohydrate form when exposed to humid
environments; Fig. S12: 3D surfaces generated through a script in Python from the 5-FC monohydrate pattern overlap collected during a VT-PXRD experiment; Fig. S13: the
DSC curve of the resulting material after VT-PXRD; Fig. S14: powder X-ray diffraction patterns of 5-FC after the solubility study (upper) and different solid forms of 5-FC
anhydrous and hydrated reported in the literature; Fig. S15: DSC profiles of 5-FC before (black) and after (purple) the solubility study; Fig. S16: TGA profiles of 5-FC
monohydrate (blue) and 5-FC anhydrous (purple) after the solubility study; Fig. S17: PXRD patterns of 5-FC (anhydrous, monohydrate, hydrate and hemihydrate) structures
confirming that the 5-FC monohydrate sample gives rise under heating to a totally different crystalline phase that does not correspond to any other similar crystal structure
deposited in the Cambridge Structure Database; Table S1: composition of the solution used for preparing the calibration curves and solubility determinations; Table S2:
standard solutions of 5-FC, caffeine and p-aminobenzoic acid used to construct the calibration curves; Table S3: regression coefficients for the calibration curves of 5-FC
and the cocrystal formers in buffer media, pH = 1.2; Table S4: pH values measured after solubility; Table S5: geometric parameters of the intermolecular interactions for
5-FC organic cocrystals; Table S6: main FT-IR and FT-Raman bands (cm1) for 5-FC and 5-FC cocrystals; and Table S7: solubility values of the four compounds. CCDC
1817618, 1817630 and 1817635. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c8nj02647e

14994 | New J. Chem., 2018, 42, 14994--15005 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018
Paper NJC

Introduction of genetic material that triggers the development of diseases such


as fungal or some cancers. In this way great effort has been
The concepts of crystal engineering remain an established way devoted to the development of new strategies to treat these
to develop and formulate improved pharmaceuticals.1–3 The diseases.20,21 Under this global scenario, the 5-FC API emerged
design and synthesis of multicomponent crystals4 is increasing, as a promising prodrug for cancer treatment via the gene therapy
allowing us to rearrange the supramolecular assembly and approach,22–29 exhibiting high solubility and permeability pro-
modulate the physicochemical properties of pharmaceuticals. files – making it belong to Biopharmaceutics Classification
This occurs based on the inclusion of chosen molecule(s) within System (BCS) class I.6,30 The anhydrous form of 5-FC is the thermo-
the API crystal lattice without the need to establish or break dynamically most stable phase at room temperature but when
covalent bonds5 – while preserving the inherent activity of the exposed to higher relative humidity, the physical stability becomes
parent drugs.6 In the case of drug-coformer cocrystals (DCC) these very questionable and the conversion to the monohydrate form
guest molecules must belong to the GRAS (Generally Recognized happens.31 Such a hydration event leads to serious problems during
as Safe) list.7 Crystal engineering methodologies become necessary processing and storage, making 5-FC tableting not trivial. Generally,
since not all marketed drugs achieved the parameters required to the API’s physical instability17 could affect a range of factors that are
be a safe and effective solid medicine. The advantages of obtaining important in view of successful formulation, quality, safety and
a homogeneous crystalline entity stabilized by noncovalent inter- efficacy of pharmaceuticals such as solubility, permeability, dissolu-
actions such as hydrogen bonds, ionic and p–p interactions, and tion rate, bioavailability and extended ‘‘Shelf Life’’.32
van der Waals forces8 are linked to the fine control of important In order to expand the range of new 5-FC solid forms, an
pharmaceutical questions such as manufacturability, storage and extensive number of structures have been reported in the last
final clinical performance.9–16 few years.33–36 Recently, our group synthesized and depicted six
5-Fluorocytosine (4-amino-5-fluoro-1,2-dihydropyrimidin- 5-FC cocrystals using as coformers adipic, succinic, terephtalic,
2-one, 5-FC, Scheme 1) is a high-dose prodrug17 which is benzoic and malic acids as well as the antineoplastic drug
established at the forefront of the treatment of various invasive 5-FU.35 Following this approach, herein we report a standardized
infections caused mainly by the fungal biological species protocol for supramolecular synthesis and the main physical and
Candida spp. and Cryptococcus neoformans.18,19 In the last chemical properties of three novel cocrystals obtained from reac-
few years, renowned researchers have dedicated themselves to tion of 5-FC with the coformers caffeine (CAF), p-aminobenzoic
obtain the Cytosine Deaminase (CD) enzyme in mammalian acid (PABA) and caprylic acid (CA) (Scheme 1).
tumor cells to study the intracellular route of 5-FC, its conver- Initially, the samples were prepared by slow evaporation of
sion to 5-FU (5-fluoro-2,4-(1H,3H)-pyrimidinedione, Scheme 1) solvent (SES) or manual mechanochemistry (see the Experi-
APIs and its influence in the inhibition of uncontrolled replication mental section), but in an attempt to develop more suitable

Scheme 1 Chemical structures of the drugs 5-fluorocytosine (5-FC), 5-fluorouracil (5-FU) and the GRAS coformers caffeine (CAF), p-aminobenzoic
acid (PABA) and caprylic acid (CA).

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018 New J. Chem., 2018, 42, 14994--15005 | 14995
NJC Paper

Table 1 Comparison between hydrothermal (SES) and automatic mechanochemistry (SDG) techniques for 5FC–CAF cocrystal supramolecular synthesis

Method SES SDG


1
MW 5-FC 129.09 g mol 129.09 g mol1
MW CAF 194.19 g mol1 194.19 g mol1
Initial mass (5-FC) 5 mg 25 mg
n n=1 n=1
Formula used for CAF mass calculation [(5-FC mass/MW 5-FC)  n  MW CAF] [(5-FC mass/MW 5-FC)  n  MW CAF]
Initial mass (CAF) 7.5 mg 37.6 mg
Total mass used 12.5 mg 62.6 mg
Amount of solvent used 1.5 mL (1500 mL) 0.025 mL (25 mL)
Duration of the experiment 3 to 5 days 1 hour
Productivity 62.6/12.5 = B5 times more
Solvent use 1500/25 = 60 times less
SES: slow evaporation of solvent; SDG: solvent-drop grinding; MW: molecular weight; 5-FC: 5-fluorocytosine; CAF: caffeine and n: mol number.

productivity methods based on the principles of green chemistry37 recrystallization through the slow evaporation method. Colorless
we also establish protocols involving automatic mechano- prismatic crystals were obtained within 3–5 days.
chemistry (solvent-drop grinding, SDG) for the production of 5-Fluorocytosine p-aminobenzoic acid cocrystal (5FC–PABA).
the 5FC–CAF cocrystal sample (Table 1).38–43 This technique Suitable single-crystals were obtained by manual mechano-
provides considerable benefits in terms of economy, efficiency chemistry. 5.00 mg of 5-FC (0.039 mmol) and 5.31 mg of PABA
and environmentally sustainable by reducing total processing (0.039 mmol) were macerated together in a pestle and mortar
times from days/weeks to hours, and minimizing exposure to by adding 300 mL of a hot acetonitrile/Milli-Q water (1 : 1, v/v,
aqueous conditions and also the use of large quantities of 100 1C) mixture until a homogeneous system was formed. The
solvent, which reduce or eliminate side-reactions.44 However, in resulting ground material was dissolved in the same hot mixture
spite of our efforts it was not possible to obtain the 5FC–PABA for its recrystallization through the slow evaporation method.
cocrystal by this method. Finally, we also report here the Colorless plate crystals were obtained within 3–5 days.
hydration stability study of the cocrystals as well as the 5-FC 5-Fluorocytosine caprylic acid cocrystal (5FC–CA). These
raw material under controlled humidity conditions, and their cocrystals were obtained by dissolving 20 mg of 5-FC (0.155 mmol)
solubility profiles at stomach pH. in an acetonitrile/Milli-Q water (5 : 1, v/v, 150 1C) mixture. Next,
to this hot homogeneous solution was added 500 mL of CA. The
system was stirred at 100 1C for approximately 15 minutes.
Experimental details Then, the crystallization batches were allowed to cool down
5-FC anhydrous was purchased from Sigma-Aldrich Brazils and slowly until room temperature and covered with Parafilms for
used without any further purification. All coformers were obtained slow evaporation of the solvent. All these systems were main-
from commercial sources and used directly. Milli-Q water and tained at 5 1C until the appearance of crystals, which occurs
acetonitrile were used as solvents. within 10–15 days.
Additionally, the same 5FC–CAF phase was also obtained
Preparation of cocrystals by an automatic mechanochemistry method (SDG) using an
Cocrystallization of equimolar amounts of 5-FC and the co- oscillatory mixer ball mill MM400 RETSCH. The sample powder
formers was performed by slow evaporation as well as by was placed in a 1.5 mL volume stainless steel milling jar
mechanochemical methods. Manual mechanochemistry experi- containing two 7 mm diameter stainless steel balls. The opti-
ments were carried out using a mortar, a pestle and stoichio- mized final condition to obtain this new solid form was
metric quantities of molecular solvents. 5-FC and the cocrystal achieved by milling the system at a frequency of 25 Hz, at room
formers CAF and PABA were ground for about 10–15 minutes. temperature, and for 1 hour. This experiment was performed by
A small amount of solvent (300 mL) was added to the grinding the addition of a 50 mg of 5-FC (0.39 mmol), 75 mg of CAF
mixture. After that, solvent evaporation was used to obtain (0.39 mmol) and only 25 mL of acetonitrile/Milli-Q water (3 : 1, v/v)
single crystals for X-ray diffraction experiments. For the mixture.
case of the CA coformer the cocrystallization solutions were
made using the SES method from mixture solvents like Single crystal X-ray diffraction (SCXRD)
acetonitrile–water. The crystallographic data for the cocrystals 5FC–CAF and 5FC–
5-Fluorocytosine caffeine cocrystal (5FC–CAF). Suitable PABA were collected at room temperature (298  2 K) using an
single-crystals were obtained by manual mechanochemistry. Agilent Super Nova diffractometer equipped with a CCD detec-
5.00 mg of 5-FC (0.039 mmol) and 7.50 mg of CAF (0.039 mmol) tor system and a Mo source (l = 0.71073 Å). Cell determination,
were ground in a pestle and mortar by adding 300 mL of a hot data integration and reduction and final parameters were
acetonitrile/Milli-Q water (3 : 1, v/v, 100 1C) mixture until a obtained using the software CrysAlisPro45 (version 171.38.43b).
homogeneous mass was formed. The resulting ground material Using Olex2,46 the structures were solved by direct methods and
was dissolved in the same composition hot mixture for its the models obtained were refined by full-matrix least squares

14996 | New J. Chem., 2018, 42, 14994--15005 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018
Paper NJC

on F2 (SHELXTL).47,48 The aromatic ring of the PABA coformer Optical microscopy


is disordered over two positions with relative occupancies Cocrystal images (5FC–CAF and 5FC–PABA crystallization medium)
of 45 : 55%. were taken using an Olympus SZ61 microscope connected to the
For the 5FC–CA compound, the data collection was carried Leica EC3 camera device (zoom: 4.5; magnification: 10).
out at 298  2 K using a Bruker Apex-II CCD diffractometer
using graphite-monochromated MoKa radiation (0.71073 Å). Thermal analysis
The SAINT software49 was used for strategy planning and data
Thermogravimetric analysis (TGA) was performed using a
reduction. Using Olex2,46 the structure was solved with the
Shimadzu TGA-50 thermobalance. An amount of approximately
ShelXT47 structure solution program using intrinsic phasing
5.0 mg  0.001 mg of each sample was placed in an alumina
and refined with the ShelXL48 refinement package using least
crucible and heated at 10 1C min1 under a N2 atmosphere
squares minimization.
(50 mL min1). For the samples 5FC–CAF and 5FC–PABA the
For all structures, the H atoms bonded to N and O atoms
temperature range analyzed was from 50 to 500 1C. For the
were located from a difference Fourier map and freely refined.
sample 5FC–CA, however, this range was from 50 to 200 1C,
Methyl H atoms were located in a Fourier calculation and refined
since at this temperature the sample has already degraded.
as riding atoms, with Uiso(H) = 1.5Ueq(C). The remaining H atoms
Differential Scanning Calorimetry (DSC) data acquisition was
bonded to C atoms were geometrically positioned (aromatic C–H =
carried out using a Shimadzu DSC-60 calorimeter until the
0.93 Å) and were refined using a riding model, with Uiso(H) =
degradation temperature of each compound. The samples
1.2Ueq(C). Molecular representations, tables and pictures were
(2.0 mg  0.01 mg) 5FC–CAF and 5FC–PABA were heated from
generated by MERCURY 3.10.250 and Olex2.46 For the 5FC–CAF and
50 to 350 1C and 5FC–CA (2.0 mg  0.01 mg) from 50 to 200 1C
5FC–PABA cocrystals one patent application was generated and
with a heating rate of 10 1C min1 in a crimped sealed aluminum
deposited on November 16, 2017 under the code BR1020170245764
pan. The purge gas was nitrogen under a flow of 50 mL min1.
(title of the invention: pharmaceutical cocrystals and their uses). All
All the values reported here were processed using the Shimadzu
CIFs were deposited in the Cambridge Structural Data Base51 under
TA-60 thermal data analysis software (version 2.2).
the codes CCDC 1817618 (5FC–CAF), 1817630 (5FC–PABA) and
1817635 (5FC–CA).
Vibrational spectra
Powder X-ray diffraction (PXRD) Fourier transform infrared (FT-IR) spectra were recorded using
The milled samples were analyzed at room temperature (293  an Alpha Bruker FT-IR spectrophotometer, using KBr pellets,
2 K) using a Rigaku Rotaflex RU200B diffractometer, in the in the range 3600–600 cm1, with an average of 64 scans and a
Bragg–Brentano reflective geometry, with CuKa radiation 2 cm1 spectral resolution. FT-Raman spectroscopy was per-
(l = 1.54 Å) at 40 kV–60 mA and a Ni filter. The diffractograms formed using a Bruker RFS 100 instrument with a Nd3+/YAG
of the 5-FC raw material and the novel cocrystals were acquired laser operating at 1064 nm in the near-infrared and a CCD
in the 5–501 2y range with a step width of 0.021 and a constant detector cooled with liquid nitrogen using a spectral resolution
counting time of 3 s per step. In addition, the 5-FC mono- of 4 cm1.
hydrate sample was also measured at different temperatures
in the 25–150 1C range with a heating rate of 2.5 1C min1. Spectrophotometric measurements and calibration curves
Variable temperature powder X-ray diffraction (VT-PXRD) A UV-1800 Shimadzu spectrophotometer was used to determine
experiments were performed using a Rigaku Ultima IV powder the absorbance of standard solutions of the raw precursors of
diffractometer coupled to a variable temperature control device the cocrystals and construct pattern curves. These curves were
using CuKa radiation (l = 1.54 Å) at 40 kV–30 mA with a step used to determine the unknown concentration of solutions
scan width of 0.021 in an interval of 5–501 in 2y and 3 s per step. of 5-FC, 5FC–CAF, 5FC–PABA or 5FC–CA. The spectra were
The 3D surfaces were generated through a script in Python measured in the range from 200 to 400 nm using 1 cm quartz
from the VT-PXRD pattern overlap. The points of the corres- cuvettes at a medium scan speed at a 1.0 nm data interval and
ponding angular positions were connected for a better observa- 1 nm bandwidth. Standard stock solutions of 5-FC, caffeine and
tion of the diffractogram evolution and, in this case, for the p-aminobenzoic acid were prepared separately in a pH 1.2
phase transition display. hydrochloric buffer (Table S1, see ESI†). 5-FC standard solu-
tions were prepared by dissolving 10.00 mg of the raw material
Hot-stage polarized optical microscopy in 1.5 mL of buffer in a 10 mL beaker (24 h agitation) and after that
Polarized microscopy studies were performed using a Leica a series of dissolutions from 0.006 mg mL1 to 0.024 mg mL1
DM2500P microscope connected to the Linkam T95-PE hot- were carried out. CAF solutions were prepared by dissolving
stage equipment. Data were visualized with the Linksys 32 10.00 mg of the solid drug in 4 mL of buffer in a 50 mL volumetric
software for hot-stage control. Each crystal was placed on an flask. After complete dissolution (24 h agitation) different concen-
individual 13 mm glass coverslip, placed on a 22 mm diameter tration points were made by appropriate dilutions at concentra-
pure silver heating block inside of the stage. The samples were tions ranging from 0.007 to 0.02 mg mL1. PABA solutions were
heated at a ramp rate of 10 1C min1 until the beginning of the prepared by dissolving 15.00 mg of the solid drug in 4 mL
degradation. of buffer in a 50 mL volumetric flask. After 24 h agitation,

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018 New J. Chem., 2018, 42, 14994--15005 | 14997
NJC Paper

different concentration points were prepared by appropriate and CA coformers give DpKa values within the ones required for
dilutions at concentrations ranging from 0.004 to 0.012 mg mL1 cocrystal formation (DpKa o 0). These values correspond,
(Table S2, see ESI†). UV-vis spectra of these solutions were used to respectively, to 10.74 and 1.63. The only exception was
build calibration curves of the API (5-FC) and the cocrystal formers the PABA coformer (DpKa = 0.88), for which the DpKa value fell
(CAF and PABA) individually (Table S3, see ESI†). Caprylic acid into the range from 0 to 3. Within this range, stated as the
shows a very low spectra signal at the range studied (200–400 nm). continuum region in the literature, a cocrystal or salt can be
Because of this, the calibration curve of CA was not constructed equally obtained.54
and the 5-FC calibration curve was used to determinate the Fig. S3 and S4 (ESI†) show the main hydrogen bonding
concentration of 5FC–CA unknown solutions. supramolecular hetero and homosynthons, respectively, found
for the 5-FC cocrystals reported here. The cocrystals were
Equilibrium solubility studies designed considering the trend of 5-FC to associate with amine,
Equilibrium solubilities of 5-FC and its cocrystals were deter- amide and carboxylic functional groups through amine  amine
mined by the shake-flask method at 25 1C in a pH 1.2 buffer (heterosynthon I, Fig. S3, ESI†), amide  carboxyl (heterosynthon
medium.52 Saturated solutions of the compounds were prepared II, Fig. S3, ESI†) and carbonyl  carboxyl (heterosynthon III,
by stirring an excess amount of 5-FC, 5FC–CAF, 5FC–PABA and Fig. S3, ESI†) classical synthons. Interestingly, the non-classical
5FC–CA, sufficient to reach saturation, into 2 mL of the dissolu- C–H  F interactions connects neighbor 5-FC and CAF mole-
tion medium for a 48 h period. These solutions were prepared cules helping in the stabilization of the heterosynthon I form
in triplicate according to the method outlined in the BCS by classical N–H  N hydrogen bonds (Fig. S3, ESI†). In a
guidance.53 After 48 h of sedimentation, the solutions were Cambridge Structural Database (CSD)55 survey none of the
filtered through a 0.45 mm PTFE hydrophilic filter (Millipores). 5-FC structures reported to date presented heterosynthon I.
The solid sediment identity was checked by PXRD analysis In addition, heterosynthons II and III appeared in only 10%
(Fig. S1, see ESI†). UV-vis spectroscopy was employed to analyze of the 5-FC structures. In turn, homosynthons I, II and III
the supernatant concentration of the compounds. The samples (Fig. S4, ESI†) were found in 38%, 22% and 16% of the structures,
were diluted in the pH 1.2 buffer medium before we started respectively.
measuring. 5-FC and cocrystals showed similar spectra that In all cases, obtaining single crystals was not a straight-
could be resolved with the addition of absorbances (Fig. S2, see forward step. The crystallization attempts of 5FC–CAF and
ESI†). Solubilities of 5-FC, 5FC–CAF, 5FC–PABA and 5FC–CA 5FC–PABA from solution gave very poor-quality crystals and/or
were measured by interpolating their maximum absorbance even mixtures of phases (5-FC raw and the respective coformers;
readings to the corresponding calibration curves constructed see Fig. S5, ESI†). In order to obtain pure single crystals for the
previously. After the equilibrium solubility measurements, the X-ray diffraction experiments, manual mechanochemistry methods
pH value in each dissolution medium was determined (Table S4, were applied only for these two cocrystals. For 5FC–CA, however,
see ESI†) using a pH meter, QX 1500 Plus Qualxtron. suitable single crystals were obtained directly by evaporation
methods. Thus, 1 : 1 5-FC : CAF, 1 : 1 5-FC : PABA and 1 : 1
Stability study in relative humidity (RH) 5-FC : CA cocrystals were all formed. Crystalline purities of
5-FC cocrystal samples were assessed by PXRD (see Fig. S6,
Powder samples of the anhydrous 5-FC and 5-FC cocrystals
ESI†). Interestingly, cocrystal 5FC–CAF was also obtained as a
(30 mg) were stored in a sealed glass desiccator chamber
powder pure phase by the SDG method (see Fig. S7, ESI†).
containing a saturated K2SO4 salt solution (distilled water) for
The ORTEP56 diagrams of the asymmetric unit (ASU) of 5-FC
B100% RH. At any temperature, the concentration of a satu-
cocrystals are shown in Fig. 1. Details of the data collection,
rated solution is fixed and does not have to be determined.
refinement and crystallographic parameters for the novel com-
Further, the milled materials were subjected to PXRD measure-
pounds are summarized in Table 2. The geometric parameters
ments as a characterization tool to monitor possible phase
of H-bonds present in the crystal structures are listed in Table S5
transitions.
(see ESI†).

Structural description
Results and discussion
5FC–CAF. The ASU comprises two 5-FC as well as two CAF
5-FC is a weak basic molecule (pKa = 3.26) that structurally molecules. As expected from the planar conformation of raw
presents two donors: NH(pirimidinic ring) and NH2(amine group), and materials, 5FC–CAF crystallizes in the monoclinic P21/c space
two acceptors: Nring and carbonyl (CQO). It cocrystallizes with group (Z 0 = 2, Z00 = 4, Fig. 1a), forming a layered structure
coformers containing COOH and NH2 functional groups. By the stabilized by classical (N–H  N) and non-classical (C–H  F)
DpKa approach, ie. DpKa = pKa (conjugate acid of the base)  interactions (Fig. 2a). In 5FC–CAF, the symmetry-independent
pKa (acid), cocrystals of 5-FC can be obtained using weak acids, 5-FC molecules are connected via N–H  O (2.983(4) Å, 169.3(2)1)
ca. DpKa o 0. Based on this perspective, caffeine (CAF, pKa = and N–H  N (2.761(4) Å, 176.5(2)1) H-bonds to form a dimeric
14.00), p-aminobenzoic acid (PABA, pKa = 2.38) and caprylic R22(8) synthon (Fig. 2b). As a result, these homodimers form
acid (CA, pKa = 4.89) were selected for the cocrystallization zigzag 1D chains (represented by the black arrows in Fig. 2a)
experiments. The differences between the pKas of 5-FC and CAF running along the a axis (highlighted in green in Fig. 2a).

14998 | New J. Chem., 2018, 42, 14994--15005 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018
Paper NJC

Fig. 1 ORTEP type diagrams showing atomic numbering schemes and 50% probability ellipsoids for: (a) 5FC–CAF, (b) 5FC–PABA and (c) 5FC–CA
compounds.

Table 2 Crystal data and structure refinement of the 5-fluorocytosine cocrystals

Identification code 5FC–CAF 5FC–PABA 5FC–CA


Empirical formula C24H28F2N14O6 C11H11FN4O3 C12H20FN3O3
Molecular weight 646.6 266.24 273.31
Temperature/K 298(2) 298(2) 298(2)
Crystal system Monoclinic Monoclinic Triclinic
Space group P21/c P21/c P1%
a/Å 9.1890(6) 4.8695(3) 5.2617(8)
b/Å 22.4128(17) 24.4911(16) 8.7066(14)
c/Å 13.3425(13) 9.8792(6) 15.710(2)
a/1 90 90 81.892(3)
b/1 93.485(8) 97.567(6) 85.884(3)
g/1 90 90 84.460(3)
Volume/Å3 2742.8(4) 1167.92(13) 707.96(19)
Z/Z 0 4/2 4/1 2/1
rcalc/g cm3 1.566 1.514 1.282
m/mm1 0.126 0.124 0.101
F(000) 1344.0 552.0 292.0
Reflections collected 17 471 19 318 24 416
Independent reflections 6188 2671 5518
Unique reflections 3818 2204 3804
Data/restraints/parameters 6188/0/421 2671/36/215 5518/0/174
R1 [I Z 2s(I)] 0.0671 0.0565 0.0560
wR2 [all data] 0.2286 0.1500 0.2134
Goodness-of-fitness on F2 1.036 1.074 1.078

By the N–H  N (3.047(4) Å, 169.4(2)1) H-bonds and C–H  F disorder with 45 : 55 occupancies. In the ASU, the 5-FC and
(3.067(2) Å, 128.64(2)1) intermolecular interactions the CAF PABA molecules interact by a N–H  O (2.745(2) Å, 169.9(2)1)
molecules are inserted on both sides of the 5-FC chain in the H-bond through an acid–pyrimidine heterosynthon (Fig. 3a).
ab plane. The dominant driving force in the 3D network Like other cocrystals, the 5-FC molecules form a robust R22(8)
assemblies is the p  p interactions along the [010] direction homodimer (highlighted in blue, Fig. 3a) through N–H  O
connecting the aromatic centers of CAF and 5-FC from adjacent (2.772(2) Å, 177.4(2)1) H-bonds (Fig. 3a). Along the [112] direc-
planes. Furthermore, C–H  O interactions are observed between tion, these homodimers connect two PABA molecules through
these planes with distances of 3.498, 3.422 and 3.514 Å, forming N–H  O (2.892(2) Å, 157.0(2)1) and N–H  O (2.942(2) Å,
supramolecular ‘‘roof tile-like’’ layers (Fig. 2c). 163.1(2)1) H-bonds to form a R24(8) motif (highlighted in light brown,
5FC–PABA. 5FC–PABA crystallizes in the monoclinic P21/c Fig. 3a) that gives rise to chains (highlighted in green in Fig. 3b). Two
space group (Z 0 = 1, Z00 = 2; 1 : 1 5-FC : PABA stoichiometry), additional PABA molecules (color in orange in Fig. 3c) cross link
Fig. 1b. The aromatic rings of the PABA molecules show two site parallel chains through N–H  O (2.745(2) Å, 169.9(2)1) H-bonds.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018 New J. Chem., 2018, 42, 14994--15005 | 14999
NJC Paper

Fig. 2 (a) Partial view down the [001] direction of 5FC–CAF crystal packing highlighting the formation of the layers. Black dotted lines indicate hydrogen
bonds. (b) Robust homosynthon formed between 5-FC’s structures. (c) Three-dimensional arrangement of crystalline 5FC–CAF showing the C–H  O
and p  p interactions that give rise to supramolecular ‘‘roof tile-like’’ layers.

Fig. 3 (a) Partial view down the [100] direction of 5FC–PABA crystal packing highlighting the formation of R24(8) and R22(8) graph set motifs via N–H  O
interactions. Black dotted lines indicate hydrogen bonds. (b) Propagation of 5-FC dimer chains attached to two PABA molecules through N–H  O
H-bonds. (c) Three-dimensional arrangement of crystalline 5FC–PABA showing the insertion of two PABA molecules via N–H  O interactions.
As a consequence, two adjacent chains are placed together along the [1% 01] direction.

As a consequence, two adjacent chains are placed together along interaction (Fig. 4a). The 5-FC molecules form 1D chains
the [1% 01] direction (highlighted in red in Fig. 3c). This arrange- through the robust amide  carbonyl (highlighted in red, Fig. 4b)
ment is also stabilized by p  p interactions with an inter- and amide  amine (highlighted in green, Fig. 4b) homosynthons
centroid distance of 3.937(2) Å between 5-FC molecules of (N–H  O: 2.792(2) Å, 173.78(2)1 and N–H  N: 2.981(2) Å,
adjacent chains. 170.58(2)1). Along this 5-FC chain, the CA molecules are con-
5FC–CA. The 5FC–CA cocrystal crystallizes in the triclinic P1% nected via the R23(9) synthon [O–H  O: 2.617(2) Å, 171.49(2)1;
space group with Z0 = 1. The ASU comprises a 5-FC molecule asso- N–H  O: 2.792(1) Å, 173.78(2)1 and C–H  O: 3.150(2) Å,
ciated with a CA one (Fig. 1c) via a O–H  O [2.617(2) Å, 171.49(2)1] 149.83(2)1, highlighted in yellow, Fig. 4b] and the R44(16) one

15000 | New J. Chem., 2018, 42, 14994--15005 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018
Paper NJC

Fig. 4 (a) Asymmetric unit of the cocrystal 5FC–CA stabilized by O–H  O interactions. (b) Partial view down the [100] direction of 5FC–CA crystal
packing highlighting the formation of amide  carbonyl and amine  amide homosynthons as well as R23(9) and R44(16) graph set motifs stabilized by
N–H  O, O–H  O and C–H  O interactions. Black dotted lines indicate hydrogen bonds. (c) Stacking of 5FC–CA molecules showing the p  p
interactions between 5-FC molecules of adjacent chains that give rise to a three-dimensional arrangement in which parallel hydrophilic channels
(the spacefill representation colored in light blue) of 5-FC are intercalated by hydrophobic channels of CA.

[N–H  O: 2.810(2) Å, 125.95(2)1 and O–H  O: 2.617(2) Å, For the 5FC–CA sample, however, this temperature is about
171.49(2)1, highlighted in grey, Fig. 4b] forming a plane. Such 87.49 1C. After these temperatures, the respective material
planes are stacked through slip-stacked p  p interactions shows a significant weight loss as can be observed in its corre-
(highlighted in red, Fig. 4c) between adjacent 5-FC molecules sponding TGA curve.
from different planes (centroid–centroid distance: 3.975(2) Å, Moreover, the DSC curve of 5FC–CAF shows three endo-
Fig. 4c). Parallel hydrophilic channels of 5-FC (the spacefill repre- thermic peaks at 252.28 1C, 272.80 1C and 294.63 1C, followed by
sentation colored in light blue) and CA hydrophobic molecules an exothermic one at 306.08 1C. For the 5FC–PABA cocrystal, two
characterize the crystal packing assembly (Fig. 4c). endothermic peaks at 221.93 1C and 288.03 1C were detected.
Finally, for the sample 5FC–CA, the DSC curve exhibits a single
Thermal analysis very well defined endothermic peak at 116 1C followed by
undefined and coupled peaks. For all cases, these events were
The thermal analysis and the phase purity of 5-FC cocrystals
attributed to the cocrystal degradation since a gradual mass
were assessed by a combination of DSC, TGA and HSM techni-
loss is observed in the corresponding TGA curves.
ques. DSC and TGA curves are shown in Fig. S8 (see ESI†). For
The thermal behavior of the cocrystals was also observed in
comparative purposes, the DSC curve of 5-FC (anhydrous and
the HSM experiments (Fig. 5) and was successfully confirmed
monohydrate) and those of all solid coformers were also measured
by them. The crystals gradually get dark as the temperature
(see Fig. S9, ESI†).
rises becoming opaque at B195.0 1C (5FC–CAF), B150.0 1C
The DSC curve of the 5-FC raw material (anhydrous) is
(5FC–PABA) and B95 1C (5FC–CA). From these temperatures
characterized by an endothermic peak at B299.62 1C coupled
forward the samples continue reducing their mass in a degra-
to an exothermic one, both attributed to the degradation
dation process.
process. The DSC curve of 5-FC exposed to a humid environ-
ment (B100% relative humidity and 25 1C), in turn, is very
different from the anhydrous form, since it is characterized by Spectroscopy analysis
one extended and endothermic peak in the 70.19–131.91 1C Fourier transform infrared (FT-IR) and Raman (FT-Raman)
range, which corresponds to the dehydration process, and one spectroscopy are popular analytical methods that provide
endothermic peak at 237.80 1C, related to the melting process. crucial information about vibrational modes of the molecules
All the novel phases exhibited lower thermal stability when as well as information about hydrogen bonds, molecular con-
compared to the anhydrous 5-FC raw material. The cocrystal formations and crystal packing of the API.57,58 For this purpose,
degradation temperatures showed a direct dependence on the a comparative analysis of 5-FC and its cocrystals’ FT-IR and
coformer melting point, i.e. a lower degradation temperature FT-Raman spectra (Fig. 6 and Fig. S10, respectively; see ESI†)
cocrystal was formed by a lower melting coformer. According to was carried out. This study shows slight dislocations of cocrystal
the TGA data, the cocrystals 5FC–CAF and 5FC–PABA are thermally 5-FC bands when compared to the bands present in the
stable from 50 1C up to about 196.76 1C and 155.75 1C, respectively. 5-FC spectra. Generally speaking, peak shifts of about 10 to

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018 New J. Chem., 2018, 42, 14994--15005 | 15001
NJC Paper

in the 3500–3400 cm1 and 1750–1670 cm1 ranges which are


related to the hydroxyl and carbonyl stretches of the COOH group,
respectively. These bands are observed in the FT-IR cocrystal spectra
at 3476 cm1 and 1675 cm1 (5FC–PABA); and 3358 cm1 and
1720 cm1 (5FC–CA). The same findings are observed in the 5FC–
PABA and 5FC–CA Raman spectra, where the CQO absorption
band appears at 1703 and 1722 cm1, respectively. Additionally, in
the case of 5FC–CA, new bands associated with the aliphatic chain
vibrations could be observed. The FT-IR and FT-Raman spectra
show characteristic bands in the region 3000–2800 cm1 assigned
to the C–H stretching modes.

Phase transition and stability test


PXRD as a function of temperature was used to show the
irreversible hydration process of 5-FC (Fig. 7) and to study the
Fig. 5 Crystal behavior as a function of temperature visually checked by
physical stability of the reported cocrystals (Fig. 8). Variable
HSM of the forms (a) 5FC–CAF, (b) 5FC–PABA and (c) 5FC–CA. temperature powder X-ray diffraction (VT-PXRD) is a method
where XRD experiments are performed at different tempera-
tures and it provides information about the solid-state changes
in materials during heating. This direct information is neces-
sary to understand the structure of pharmaceutical solids in the

Fig. 6 FT-IR spectra of 5-FC raw, 5FC–CAF, 5FC–PABA and 5FC–CA.

30 cm1 in the spectra are indicative of changes in hydrogen


bonding patterns.59,60 Fig. 7 Variable temperature PXRD patterns for 5-FC monohydrate.
The spectra interpretation and band assignments (Table S6,
see ESI†) were performed taking into account the crystallo-
graphic study and the spectroscopic data available for related
5-FC compounds.36,61,62 5-FC molecules exhibit IR and Raman
stretching frequencies at 3375 cm1 (amine N–H stretch),
1682 cm1 and 1677 cm1 (amide CQO stretches), 1337 cm1
and 1340 cm1 (pyrimidine ring C–N stretches) and 1227 cm1
and 1236 cm1 (C–F stretches). Some of these stretching frequen-
cies appear shifted (10–30 cm1) in the FT-IR and FT-Raman
spectra (Table S6, see ESI†), indicating cocrystal formation.
According to the literature, amide groups normally display a
characteristic absorption band at about 1700–1600 cm1 related
to the carbonyl stretch. Thus, two absorption bands are observed
in the 5FC–CAF spectra at 1650 cm1 in the FT-IR and 1630 cm1
in the FT-Raman, and are attributed to the carbonyl stretching
modes of caffeine.
As expected, carboxylic p-aminobenzoic and caprylic acids, Fig. 8 PXRD patterns of 5-FC cocrystals showing the maintenance of the
used as coformers, display two characteristic absorption bands crystalline phase after one week in a humid environment.

15002 | New J. Chem., 2018, 42, 14994--15005 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018
Paper NJC

development process of new formulations.63–67 The VT-PXRD might correspond to the phase transition and/or dehydration
studies of 5-FC monohydrate show no change until B75 1C. process; (ii) the DSC plot (also in Fig. S15, see ESI†) shows the
After this temperature the highly defined diffractions peaks of same signal at B300 1C (highlighted with a red asterisk) corres-
the starting material start to disappear showing its amorphiza- ponding to the decomposition of 5-FC anhydrous and finally,
tion due to water molecule loss (colored in green, Fig. 7). (iii) the TGA comparison between 5-FC monohydrate and the 5-FC
Since the water molecules are strongly bound into the crystal solid form obtained after the solubility studies (Fig. S16, see ESI†)
structure (corroborated by the extended peak in the DSC curve, reveals that the amount of water in the crystal structure is reduced
Fig. S9, ESI†) their loss destabilizes the crystal packing. However, when compared to the monohydrate. The summary of these facts
in the 125 1C to 150 1C range the crystallinity of the sample seems to indicate that solid 5-FC anhydrous, in the presence of
reappears but with another identity, that is, it does not correspond the buffer solution pH 1.2, produces a different kind of hydrates
to the anhydrous 5-FC solid phase (colored in red, Fig. S11, see and also suffers a phase transition. The DSC profiles (Fig. S15,
ESI†). This phase transition could also be observed in the 3D ESI†) show that all these changes are reversible.
surfaces (Fig. S12, see ESI†). This information is also supported by
the DSC curve of the resulting material after VT-PXRD (Fig. S13,
see ESI†). The DSC curve shows four endothermic peaks at
Conclusions
88.99 1C, 235.22 1C, 252.11 1C and 265.75 1C. Is interesting to Single-crystal X-ray diffraction studies show that 5-FC is able to
note that this new anhydrous phase turns back to the mono- form different homo and heterosynthons, which is a notable
hydrate form after one week of exposure to a highly humid feature of N-heterocyclic derivatives, such as pyridines and
atmosphere, as can be seen in the PXRD patterns of Fig. S11 pyrimidines. These results corroborate the instability of the
(see ESI†). The degrees of crystallinity of cocrystal forms were 5-FC anhydrous raw material and the efficacy of methodologies
reduced after the accelerated stability test (100% RH atmosphere), such as rational planning and supramolecular synthesis in the
but the main characteristic peaks of the calculated patterns development of more stable API solid forms. The vibrational
remain present in the experimental ones after the same exposure spectroscopic studies were found to be congruent with the
time, confirming no phase transition as can be observed in Fig. 8. chemical structures of the three novel cocrystals. VT-PXRD
shows that it is not possible to recover anhydrous 5-FC by
Equilibrium solubility
heating the monohydrate form since this hydration process is
The solubility study of the cocrystals was performed in a buffer irreversible and the heating gives rise to a non-reported crystal-
mimicking the stomach pH. These tests show that the new solid line phase (see Fig. S17, ESI†). Since the hydration process does
forms retain solubility values statistically similar to the one of the not occur in the new cocrystals reported here they can be used
5-FC raw material (Fig. 9 and Table S7, see ESI†). PXRD also shows by the pharmaceutical industry as a very attractive strategy to
that the crystal structures of the solids obtained after the solubility increase the physical stability of the API, avoiding in this way
test remain the same as the original ones (Fig. S1, see ESI†). the phase transitions in pharmaceuticals containing 5-FC in
However, diffractograms before and after the solubility study of their formulations. The solubility profiles of 5-FC cocrystals in
5-FC do not show the same results. In this case, the remnant solid buffer pH 1.2 show solubility values very close to the 5-FC raw
shows additional peaks that correspond to other 5-FC polymorphs material. This prodrug is classified as a class I drug due to its
as well as different hydrates (Fig. S14, see ESI†). When thermal considerable solubilization in the stomach at low pH. This fact
analyses of these solids are performed three important facts are strengthens the establishment of cocrystallization as an effec-
exposed: (i) the DSC contrast between the solids before and after tive tool for the stabilization of an API with respect to hydration
the solubility study shows an endothermic signal at 52.19 1C and without severe impairment of solubility.
another at 104.42 1C (highlighted in gray, Fig. S15, see ESI†) that

Author contributions
The manuscript was written through contributions of all authors.
All authors have given approval to the final version of the
manuscript.

Conflicts of interest
The authors declare no competing financial interest.

Acknowledgements
The authors acknowledge the Brazilian funding agencies CAPES
Fig. 9 Solubility concentration of 5-FC and the different novel cocrystal (M. S. S.), FAPESP (L. F. D. grant 15/25694-0) and CNPq (J. E.
forms in pH 1.2 dissolution media. grant #305190/2017-2) for financial support. The authors would

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018 New J. Chem., 2018, 42, 14994--15005 | 15003
NJC Paper

also like to thank Dr Charlane C. Correa, and Dr Luiz F. C. 21 H. M. Pinedo and G. F. Peters, J. Clin. Oncol., 1988, 6,
de Oliveira (Federal University of Juiz de Fora) for allowing access 1653–1664.
to the Raman spectroscopy (measurements for all cocrystals 22 C. J. Springer, I. Niculescu-duvaz and D. H. Kirn, J. Clin.
reported here) and single-crystal X-ray facilities (data collection Invest., 2000, 105, 1161–1167.
of the cocrystals 5FC–CAF and 5FC–PABA). L. Vogt thanks IFSC/ 23 C. Altaner, Cancer Lett., 2008, 270, 191–201.
USP for being granted with a Salazar Scholarship which made 24 M. H. Amer, Mol. Cell. Ther., 2014, 2, 1–19.
possible the active collaboration in this work. Moreover, the 25 O. Greco and G. U. Dachs, J. Cell. Physiol., 2001, 187, 22–36.
authors would like to thank Prof. Dr Christian W. Lehmann 26 B. Yi, S. U. Kim, Y. Kim, H. J. U. N. Lee, M. Cho and K. Choi,
(Max-Planck-Institut für Kohlenforschung) for allowing access to Oncol. Rep., 2012, 27, 1823–1828.
the X-ray diffraction facilities, which made possible the 5FC–CA 27 A. V. Patterson, Molecules, 2009, 14, 4517–4545.
data collection. Finally, the authors would like to thank Igor R. dos 28 W. Touati, P. Beaune and I. De Waziers, Cancer Gene
Santos (São Carlos Institute of Physics) for generating the PXRD 3D Therapy: The New Targeting Challenge, 2010.
surfaces through a script in Python. 29 M. Nyati, Z. Symon, E. Kievit, K. Dornfeld, S. Rynkiewicz,
B. Ross and A. Rehemtulla, Gene Ther., 2002, 9, 844–849.
30 L. B. Arrunátegui, N. M. Silva-barcellos, K. R. Bellavinha,
References S. Ev and J. De Souza, Braz. J. Pharm. Sci., 2015, 51, 143–154.
31 L. Wang, X. Wen, P. Li, J. Wang, P. Yang, H. Zhang and
1 P. Sanphui, V. K. Devi, D. Clara, N. Malviya, S. Ganguly and Z. Deng, CrystEngComm, 2014, 16, 8537–8545.
G. R. Desiraju, Mol. Pharmaceutics, 2015, 12, 1615–1622. 32 R. Zelkó and G. Szakonyi, Int. J. Pharm. Invest., 2012, 2, 18.
2 S. Tothadi, P. Sanphui and G. R. Desiraju, Cryst. Growth 33 M. Tutughamiarso, G. Wagner and E. Egert, Acta Crystallogr.,
Des., 2014, 14, 5293–5302. Sect. B: Struct. Sci., 2012, 68, 431–443.
3 S. K. Nechipadappu and D. R. Trivedi, J. Mol. Struct., 2017, 34 M. Tutughamiarso and E. Egert, Acta Crystallogr., Sect. B:
1141, 64–74. Struct. Sci., 2012, 68, 444–452.
4 N. Schultheiss and A. Newman, Cryst. Growth Des., 2009, 9, 35 C. C. P. Da Silva, R. D. O. Pepino, C. C. De Melo, J. C. Tenorio
2950–2967. and J. Ellena, Cryst. Growth Des., 2014, 14, 4383–4393.
5 C. B. Aakeröy, S. Forbes and J. Desper, J. Am. Chem. Soc., 36 Y. Du, Q. Cai, J. Xue, Q. Zhang and D. Qin, Spectrochim. Acta,
2009, 131, 17048–17049. Part A, 2017, 178, 251–257.
6 A. Samie, G. R. Desiraju and M. Banik, Cryst. Growth Des., 37 D. K. Leahy, J. L. Tucker, I. Mergelsberg, P. J. Dunn, M. E.
2017, 17, 2406–2417. Kopach and V. C. Purohit, Org. Process Res. Dev., 2013, 17,
7 Food and Drug Administration, Database of Select Committee 1099–1109.
on GRAS Substances (SCOGS) Reviews. 38 C. C. De Melo, C. P. Cecilia, C. C. S. S. Pereira, P. C. P. Rosa
8 K. Chadha, M. Karan, Y. Bhalla, R. Chadha, S. Khullar, S. Mandal and J. Ellena, PHASCI, 2016, 81, 149–156.
and K. Vasisht, Cryst. Growth Des., 2017, 17, 2386–2405. 39 Y. Zhou, F. Guo, C. E. Hughes, D. L. Browne, T. R. Peskett
9 C. C. Sun, Expert Opin. Drug Delivery, 2012, 10, 1–13. and K. D. M. Harris, Cryst. Growth Des., 2015, 15, 2901–2907.
10 S. F. Chow, M. Chen, L. Shi, A. H. L. Chow and C. C. Sun, 40 A. Delori and T. Fri, CrystEngComm, 2012, 14, 2350–2362.
Pharm. Res., 2012, 29, 1854–1865. 41 J.-L. Do and T. Friščić, ACS Cent. Sci., 2017, 3, 13–19.
11 Y. Chen, L. Li, J. Yao, Y. Y. Ma, J. M. Chen and T. B. Lu, 42 G. Kaupp, CrystEngComm, 2009, 11, 388–403.
Cryst. Growth Des., 2016, 16, 2923–2930. 43 S. Bordignon, P. Cerreia Vioglio, E. Priola, D. Voinovich,
12 X. L. Dai, S. Li, J. M. Chen and T. B. Lu, Cryst. Growth Des., R. Gobetto, Y. Nishiyama and M. R. Chierotti, Cryst. Growth
2016, 16, 4430–4438. Des., 2017, 17, 5744–5752.
13 R. Z. Lin, P. J. Sun, Q. Tao, J. Yao, J. M. Chen and T. B. Lu, 44 O. Eguaogie, J. S. Vyle, P. F. Conlon, M. A. Gı̂lea and
Eur. J. Pharm. Sci., 2016, 85, 141–148. Y. Liang, Beilstein J. Org. Chem., 2018, 14, 955–970.
14 S. Li, J.-M. Chen and T.-B. Lu, CrystEngComm, 2014, 16, 45 CrysAlisPRO Oxford Diffraction/Agilent Technologies UK
6450–6458. Ltd, 2014.
15 N. Geng, J. M. Chen, Z. J. Li, L. Jiang and T. B. Lu, Cryst. 46 O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K.
Growth Des., 2013, 13, 3546–3553. Howard and H. Puschmann, J. Appl. Crystallogr., 2009, 42,
16 Q. Tao, J. M. Chen, L. Ma and T. B. Lu, Cryst. Growth Des., 339–341.
2012, 12, 3144–3152. 47 G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Adv., 2014,
17 S. R. Perumalla, S. Paul and C. C. Sun, J. Pharm. Sci., 2016, 71, 3–8.
105, 1960–1966. 48 G. M. Sheldrick, Acta Crystallogr., Sect. C: Struct. Chem.,
18 E. R. Block, A. E. Jennings and J. E. Bennett, Antimicrob. 2015, 71, 3–8.
Agents Chemother., 1973, 3, 649–656. 49 SAINT, Version 7.60a, Bruker AXS Inc, Madison, WI, USA,
19 L. Wang, X. Wen, P. Li, J. Wang, P. Yang, H. Zhang and 2006.
Z. Deng, CrystEngComm, 2014, 16, 8537–8545. 50 C. F. Macrae, et al., J. Appl. Crystallogr., 2006, 39, 453–457.
20 D. B. Longley, D. P. Harkin and P. G. Johnston, Nat. Rev. 51 F. Allen, Acta Crystallogr., Sect. B: Struct. Sci., 2002, 58,
Cancer, 2003, 3, 330–338. 380–388.

15004 | New J. Chem., 2018, 42, 14994--15005 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018
Paper NJC

52 A. Glomme, J. März and J. B. Dressman, J. Pharm. Sci., 2005, 60 A. Mukherjee, S. Tothadi, S. Chakraborty, S. Ganguly and
94, 1–16. G. Desiraju, CrystEngComm, 2013, 15, 4640–4654.
53 O. M. M. Santos, M. E. D. Reis, J. T. Jacon, M. E. de, S. Lino, 61 G. Ślósarek and R. Zamboni, Spectrochim. Acta, Part A, 1991,
J. S. Simões and A. C. Doriguetto, Braz. J. Pharm. Sci., 2009, 47, 863–874.
50, 1–24. 62 A. Jaworski, M. Szczesniak, K. Kubulat and W. B. Person,
54 C. C. P. Da Silva, R. De Oliveira, J. C. Tenorio, S. B. Honorato, J. Mol. Struct., 1990, 223, 63–92.
A. P. Ayala and J. Ellena, Cryst. Growth Des., 2013, 13, 4315–4322. 63 M. Karjalainen, S. Airaksinen, J. Rantanen, J. Aaltonen and
55 C. R. Groom, I. J. Bruno, M. P. Lightfoot and S. C. Ward, Acta J. Yliruusi, J. Pharm. Biomed. Anal., 2005, 39, 27–32.
Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2016, 72, 64 J. Aaltonen, M. Allesø, S. Mirza, V. Koradia, K. C. Gordon
171–179. and J. Rantanen, Eur. J. Pharm. Biopharm., 2009, 71, 23–37.
56 L. J. Farrugia, J. Appl. Crystallogr., 2012, 45, 849–854. 65 V. R. Vangala, P. S. Chow, M. Schreyer, G. Lau and R. B. H.
57 A. Heinz, C. J. Strachan, K. C. Gordon and T. Rades, Tan, Cryst. Growth Des., 2016, 16, 578–586.
J. Pharm. Pharmacol., 2009, 61, 971–988. 66 J. X. Song, Y. Yan, J. Yao, J. M. Chen and T. B. Lu, Cryst.
58 A. P. Ayala, H. W. Siesler, R. Boese, G. G. Hoffmann, G. I. Growth Des., 2014, 14, 3069–3077.
Polla and D. R. Vega, Int. J. Pharm., 2006, 326, 69–79. 67 M. F. Pina, M. Zhao, J. F. Pinto, J. J. Sousa, C. S. Frampton,
59 S. Gunasekaran, G. Sankari and S. Ponnusamy, Spectrochim. V. Diaz, O. Suleiman, L. Fábián and D. Q. M. Craig, Cryst.
Acta, Part A, 2005, 61, 117–127. Growth Des., 2014, 14, 3774–3782.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2018 New J. Chem., 2018, 42, 14994--15005 | 15005

You might also like