Dudognon Et Al 2021 Milling Induced Phase Transformations Underlying Mechanisms and Resulting Physical States in An

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

pubs.acs.

org/crystal Article

Milling-Induced Phase Transformations, Underlying Mechanisms,


and Resulting Physical States in an Enantiotropic System: The Case
of Bezafibrate
Emeline Dudognon,* Florence Danède, and Mathieu Guerain
Cite This: Cryst. Growth Des. 2022, 22, 363−378 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The mechanisms driving the milling-induced transformations of an


enantiotropic polymorph system, bezafibrate, are depicted. The phase trans-
Downloaded via UNIV DI PISA on November 29, 2023 at 13:54:00 (UTC).

formations promoted by milling at two different temperatures and the resulting


physical states were carefully investigated by the cross-use of structural (X-rays
powder diffraction), thermodynamic (differential scanning calorimetry), and
dynamic (dielectric relaxation spectroscopy) techniques. Our results highlight that
milling of the commercial α phase at Tg50 °C (−10 °C) leads to a complete
amorphization (Tg ∼ 40 °C), whereas at Tg15 °C (25 °C), it leads to the stable
β phase. We establish that, as for monotropic situations, the solid−solid conversion
is mediated by a transient amorphous state resulting from a milling-induced
disordering of the crystalline structure partly counterbalanced by a slower re-crystallization. However, the monitoring of the
transformation kinetics (phase ratio and crystallites’ size) reveals that at least 10% amorphous phase is required to trigger the re-
crystallization to the stable β form instead of the metastable α form. For the first time, the molecular mobility of the physical states
resulting from milling is finely investigated by dielectric relaxation spectroscopy. Strikingly, it evidenced, for the crystalline phases
produced by milling, a residual mobility (detection of localized intra-molecular motions but absence of wide-amplitude motions
characterizing the amorphous state), which originates through the mobility of part of molecules on the surface of crystallites. This
outstanding result emphasizes and unravels the highly defective nature of the crystalline phases generated by milling.

1. INTRODUCTION understood despite milling-induced phase transitions, that is,


Milling is an important stage in the processing of all types of amorphization or polymorphic transformations, have been
materials. Indeed, in addition to reducing the size of particles, extensively studied over the years.
For molecular compounds, it has been shown that a
high-energy milling is able to trigger physical transitions and
preponderant parameter is the milling temperature with regard
chemical reactions.1 As it allows to reach nonequilibrium
to the glass transition (Tg) of the compound:15 amorphization
states, called driven materials by Martin and Bellon,2 it is used
is observed when milling is performed well below Tg whereas
for the formation of alloys (from intermetallics3−5 till
solid−solid conversion between polymorphic crystalline forms
molecular crystals6), or on a more fundamental point of
is observed when milling is performed above Tg. Regarding the
view, it can be used to explore the different physical states of a
involved mechanisms, the amorphization appears to result
compound and the screening of polymorphs and their relative
from the increasing disorder generated by the mechanical
stabilities.7 Moreover, milling avoids the use of solvents or
perturbation.16 However, as far as solid−solid transformations
high-temperature processing. It is thus an interesting
are concerned, the situation is blurrier as it refers to the relative
technology that develops in many different fields from
thermodynamic stability of the different crystalline phases of
metallurgy to biology.1 In particular, in the pharmaceutical the compound. On the one hand, the two crystalline
field, originally used to decrease the size of particles in order to polymorphic phases can form a monotropic set: one of the
enhance the dissolution, the powder flowability, or the two phases is stable in the entire temperature range whereas
tableting,8 it becomes more and more a tool for mechanosyn- the other one is always metastable (the Gibbs free energy of
thesis of promising multi-components systems aiming to
overcome solubility issues of drugs such as nanocrystals,9 co-
crystals,10,11 amorphous solid dispersions in early-stage Received: August 30, 2021
research,12 co-amorphous system,13 or the loading of drugs Revised: November 12, 2021
into mesoporous matrices.14 However, in order to use Published: November 25, 2021
efficiently this technology, the mechanisms underlying these
transformations have to be controlled. However, even the
simple case of the milling of a pure compound is still not fully

© 2021 American Chemical Society https://doi.org/10.1021/acs.cgd.1c00997


363 Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

the first is always lower than that of the latter).17,18 Then, it to the wide accessible frequency range (mHz to MHz). The
seems from previous investigations of the mechanisms ruling cross-use of these techniques allows us to fully describe the
the solid−solid conversion under milling that it proceeds nature of the physical states of bezafibrate generated by the
through an intermediate amorphous state: milling leads to an mechanical perturbation.
amorphization of the compound but the mobility enables a
rapid re-crystallization.19 On the other hand, the system can be 2. EXPERIMENTAL SECTION
enantiotropic: each phase is stable in a temperature range, and 2.1. Material. Bezafibrate (2-[4-[2-(4-chlorobenzamide)ethyl]-
the temperature at which the inversion of stability occurs phenoxy]-2-methylpropanoic acid, C19H20ClNO4, cf. Figure 1, CAS:
(crossing of the two corresponding Gibbs free energy curves) 41859-67-0, and Mw = 361.82 g mol−1) was furnished by Sigma-
being the theoretical temperature of the solid−solid trans- Aldrich and was used without further purification (purity ≥ 98%).
formation.17,18 Then, because this transition point exists, a The crystalline form is the α form.
solid−solid conversion may directly occur under milling
without the requirement of a transient amorphous state.
However, unfortunately, too few cases of such enantiotropic
sets are reported to solve this issue. It has been shown that the
milling at room temperature of the stable phase (II) (at this
temperature) of sulindac leads to a complete amorphization of
this compound.20 In the case of sulfamerazine, milling of both
the stable and the metastable phases leads to the crystalline Figure 1. Bezafibrate molecule.
stable phase.21 It is also the case of sulfathiazole but a complex
behavior is observed for the milling of the stable form (III) (at 2.2. Methods. 2.2.1. Milling. Milling of bezafibrate was performed
the milling temperature). Indeed, at the initial stage of milling, in a high-energy planetary mill (Pulverisette 7 from Fritsch GmbH). 1
the metastable form (I) appears, presumably assisted by the g of material was placed in ZrO2 milling jars of 43 cm3 containing
seven ZrO2 balls (Ø = 15 mm), corresponding to a ball/sample
development of a transient amorphous fraction. This results
weight ratio of 75:1. The milling was performed at two different
after 10 min milling in a mixture of stable (III) and metastable temperatures: −10 °C and room temperature. In the former case, the
(I) crystalline phases; however, upon further milling, the milling device was placed in a cold room with a dehumidifier so that a
transformation reverses back.22,23 Therefore, this kind of temperature of −10 °C and a null relative humidity percentage could
milling-induced transformations and their mechanism still be maintained during milling. In order to limit the overheating of the
remain unclear. compound, cycles of 20 min milling (rotation speed of the solar disk
Bezafibrate is an organic compound used for its efficiency as set to 400 rpm) and 10 min pause periods were applied. The total
an antihyperlipoproteinemic drug. It can appear in two effective milling time was tmill = 14 h at −10 °C and ranged from tmill =
crystalline forms: α and β, commercial bezafibrate being in 5 min to tmill = 14 h at room temperature.
2.2.2. XRPD. The XRPD experiments were performed with an
the α form. From a detailed investigation of both crystalline XPERT PRO MPD diffractometer in Debye−Scherrer geometry
phases, Lemmerer and co-workers evidenced that they form an (λCu Kα = 1.5406 Å) equipped with an X’celerator linear position-
enantiotropic set: the β form is the stable form at room sensitive detector. Samples were enclosed in Lindemann glass
temperature and solid−solid transition to the α form occurs on capillaries (Ø = 0.7 mm) that were mounted on a rotating sample
heating at around 160 °C.24 The crystalline structure of the α holder in order to avoid effects of preferential orientations of
form, which melts at 185 °C (ΔHm = 52 kJ mol−1),24 is in the crystallites if any. X-rays diagrams were recorded from 5 to 70° (2θ)
orthorhombic system with the space group P212121,25 whereas with 50 s per step and a step size of 0.0167° (2θ), mostly at room
the β form crystallizes in the centrosymmetric space group temperature. For X-rays diagrams recorded at higher temperatures (in
P21/c.24 Besides, this active pharmaceutical ingredient is poorly the range of 54−190 °C), an Huber furnace was used. Microstructural
analyses and phase quantification were performed thanks to Rietveld
water soluble (0.016 ± 0.003 mg mL−126 ). Because the use of refinements using MAUD software29 from the data obtained by
the amorphous state can improve the bioavailability, the glass- Lemmerer et al. for α and β polymorphs (CSD Reference
forming ability of bezafibrate was investigated.27 ,28 Wytten- VAMBOA01 and VAMBOA02).24
bach and co-workers28 showed that a glass of bezafibrate could 2.2.3. Differential Scanning Calorimetry. DSC measurements
be obtained by the quenching of the liquid (Tg = 37 °C) and were performed in a differential scanning calorimeter Q200 from TA
did not crystallize on subsequent heating. However, to the best Instrument equipped with a refrigerated cooling system. Samples were
of our knowledge, the glassy state of bezafibrate, in particular encapsulated in a standard aluminum pan T0 (2 mg < m < 3 mg) and
the dynamic, has not been further explored. measurements were conducted under nitrogen flow (50 mL min−1).
In order to separate when needed the events depending on Cp change
In this article, we report a detailed investigation of the
(reversing heat flow), in particular to evidence the Cp-jump
behavior of bezafibrate under high-energy milling. The aim is characteristic of a glass transition, from those depending on physical
to take advantage of the thermodynamic features of this transformations (non-reversing heat flow), a modulated temperature
compound to enhance the knowledge of the mechanisms protocol (MDSC) was followed. A ramp of 5 °C min−1 with a
ruling the milling-induced transitions. For this purpose and temperature modulation of ±0.663 °C per 50 s or ±0.530 °C per 40 s
given the key role of the milling temperature on triggered was applied. Temperatures and enthalpies were calibrated using
phase transformations, milling is performed at two different indium at the same heating rate and under the same environmental
temperatures, −10 °C and the room temperature (∼25 °C). In conditions as the experiments. Specific heat capacity was measured
both cases, milled bezafibrate is then deeply characterized by using sapphire as a reference.
2.2.4. Dielectric Relaxation Spectroscopy. Dielectric relaxation
three complementary techniques: X-rays powder diffraction spectroscopy (DRS) measurements were performed using a
(XRPD) to probe the structure and microstructure, differential Novocontrol Technologies GmbH Alpha analyzer. A small amount
scanning calorimetry (DSC) to probe the thermodynamic of the milled sample powder was put between two gold-coated
behavior, and the dynamic relaxation spectroscopy (DRS) that electrodes (diameter of 10 mm), and inert quartz spacers (50 μm
allows to probe the global and localized molecular motions due thickness) were added to avoid short circuits between electrodes. The

364 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

powder was packed as much as possible to reduce the noise due to the The thermogram of the milled compound recorded by
non-homogeneity of the sample. However, it should be noted that, modulated DSC at 5 °C min−1 ± 0.663 °C per 50 s from 0 to
due to the unavoidably entrapped air between grains, the exact 200 °C is reported in Figure 3. It exhibits four events. First, a
volume of the sample could not be measured. Nevertheless, as shown
by Kothary and co-workers,30 even if the increase in porosity leads to
a decrease in the measured dielectric constant, it has no influence on
the relaxation time. In spite of the lack of precision, we preferred this
protocol instead of the application of a pressure for the tableting of
the powder because it would not modify the physical state resulting
from milling. The complex permittivity ε*(ω) = ε′(ω) − iε″(ω) was
recorded while a sinusoidal electrical field of 1 V magnitude and of
frequency, F, varying from 106 to 10−2 Hz was applied to the samples.
Temperature was controlled through a Quatro cryosystem (supplied
by Novocontrol Technologies), and spectra were collected during
isotherms ranging from −100 to 70 °C using a nitrogen gas cryostat
with a stability condition of 0.5 °C. Data were analyzed using the
Havriliak−Negami (HN) model function31 (or a sum of HN
functions when several relaxations appeared) as expressed using the
following equation
Δε
ε*(ω) = ε∞ +
[1 + (iωτHN)αHN ]βHN (1)
where Δε = εs − ε∞ is the difference between the real permittivity
values at respectively low and high frequency limits, τHN is the
Havriliak−Negami characteristic relaxation time of the process, and Figure 3. Thermogram of bezafibrate milled for 14 h at −10 °C
αHN and βHN are shape parameters taking into account the broadening recorded using MDSC (total heat flow) at 5 °C min−1 ± 0.663 °C per
and asymmetry (0 < αHN ≤ 1 and 0 < βHN ≤ 1). From the HN fitting 50 s. The inset shows the reversing heat flow recorded between 20
parameters values, the model independent relaxation time τmax = 1/ and 50 °C for the milled sample and the glass obtained from the

ÄÅ É Ä É
2πFmax was calculated using the following equation

ÅÅ i α π yÑÑÑ−1/ αHN ÅÅÅ i α β π yÑÑÑ−1/ αHN


quenching of the liquid.

ÅÅ jj zzÑÑ ÅÅ jj HN HN zzÑÑ
τmax = τHNÅÅÅsinjjj zzÑÑ
zÑÑ
ÅÅsinjj zÑ
ÅÅ j 2 + 2β zzÑÑÑ
ÅÅ
ÅÇ k 2 + 2βHN {ÑÑÖ ÅÅÇ k HN {ÑÑÖ
HN small and broad endothermic bump can be seen between 10
(2) and 50 °C, both on the total heat flow and the non-reversing
heat flow (not shown). Due to its location, it corresponds to
the loss of some adsorbed water. Indeed, such water sorption is
3. RESULTS usually observed for milled compounds which are more
3.1. Milling at −10 °C (Tmill = Tg50 °C). The reactive, owing to an increased specific area. However, an
commercial α form of bezafibrate was milled for 14 h at −10 inspection of the reversing heat flow in this temperature range
°C. The X-ray diffraction pattern, recorded at room temper- (cf. the inset in Figure 3) reveals that the loss of water masks
ature just after milling, can be seen in Figure 2 (bottom). The the Cp-jump characteristic of a glass transition, which
absence of Bragg peaks indicates that the compound has been corroborates that the sample has been amorphized by milling.
fully amorphized by milling. Successive recordings on It is located at Tg = 39 ± 1 °C, and we obtained the ΔCp
increasing temperature show that the sample re-crystallizes value, ΔCp = 0.41 ± 0.01 J g−1 °C−1 (cf. the inset in Figure 3).
around 50 °C in the β form, which converts into the α form At higher temperatures, an exothermic peak can be seen
between 130 and 150 °C, and then, it melts at 180 °C (cf. between 60 and 70 °C. It corresponds to the re-crystallization
Figure 2). of the sample in agreement with X-ray diffraction results.
Between 140 and 160 °C, a small endothermic peak is
observed (ΔHconv = 3.5 kJ mol−1). As reported by Lemmerer
and co-workers,24 it is due to the solid−solid conversion of the
re-crystallized β form to the α form. Finally, an endothermic
peak can be seen between 180 and 190 °C. It corresponds to
the melting of the α form (Tm = 181.5 ± 0.1 °C and ΔHm =
55.4 kJ mol−1). In addition, commercial bezafibrate has been
heated at 5 °C min−1 until 200 °C, and the melt has been
immediately quenched to −50 °C to obtain a glass. On re-
heating, a Cp jump can be seen in the same temperature range
as for the compound amorphized by milling. However, a close
inspection of the reversing heat flow (cf. the inset in Figure 3)
shows that the glass transition temperature is a little bit lower
(Tg = 37 ± 1 °C and ΔCp = 0.45 ± 0.01 J g−1 °C−1) but in
nice agreement with the value reported by Wyttenbach et al.28
(Tg = 36.9 ± 0.2 °C at 10 °C min−1). It should be added that
no re-crystallization exotherm is observed at higher temper-
ature.
Figure 2. Successive XRPD patterns recorded on increasing To check the chemical nature of the glasses obtained from
temperatures of Bezafibrate milled for 14 h at −10 °C. these two different routes, analysis by 1H NMR has been
365 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

Figure 4. Evolutions of tan δ = ε″/ε′ vs the frequency F of the applied electrical field for bezafibrate milled for 14 h at −10 °C. (a) In the glassy
state, at four representative temperatures. Solid lines: first recordings (freshly milled compound) dashed lines: recordings after dehydration at 25
°C. (b) In the supercooled liquid state, on isotherms ranging from 38 to 52 °C by steps of 2 °C (solid lines). At 52 °C, two recordings are added
after isotherms of 300 s (dashed line) and 600 s (dotted line).

performed. The spectra were compared to the spectrum of the location of the maximum of the γ mode in the frequency
commercial bezafibrate (see Supporting Information S1). range where γw becomes asymmetric and the similarities of
Spectra of the commercial and milled compounds are identical spectra on the high-frequency side suggest that the γ and δ
whereas some differences appear in the spectrum of bezafibrate modes are already present in the first recording but are masked
amorphized by quenching of the melt (in particular between 6 at low frequency by the more intense γw. As an example, a
and 7 ppm). It reveals first that some thermal degradation focus on spectra recorded at −70 °C can be seen in Supporting
occurs as bezafibrate melts. According to Shen and Zhou,32 it Information (S2). The spectra recorded in the second scan
seems to correspond to the breaking of the ether bond. Such have been analyzed using two Cole−Cole functions (HN
thermal degradation, even in small quantity, could explain why functions with βHN = 1, see Experimental Section) (the fit at
the quenched liquid does not re-crystallize on heating. Second, −70 °C of the two corresponding modes γ and δ can also be
1
H NMR analysis highlights that the milling of bezafibrate at seen in S2). In a first approximation, the spectra recorded in
−10 °C permits to obtain a glass of bezafibrate, which is the first scan have been fitted with three Cole−Cole functions,
chemically pure and, in the authors’ knowledge, it is the only the γ and δ modes being fixed from fit of the second scan (cf.
way to obtain it. S2, e.g., at −70 °C). It permits to obtain an evolution versus
The molecular mobility of milled bezafibrate has been temperature of the frequency of the γw peak maximum similar
investigated by DRS. First, in order to detect motions ascribed but more regular than by a direct reading. Experiments were
to local mobility, experiments have been performed in the also performed at higher temperatures: the recording of the
glassy state, below Tg. The evolution of the permittivity was permittivity between 10−2 and 106 Hz from 25 to 70 °C shows
recorded between 10−1 and 106 Hz in isotherms ranging from a peak that enters the frequency window at 38 °C (cf. Figure
−100 to 25 °C, every 5 °C. As an example, some tan δ spectra 4b) and shifts toward high frequency with increasing
(tan δ = ε″/ε′) obtained in this temperature range are reported temperature. It corresponds to the main relaxation, α, due to
in Figure 4a (solid lines). They show the shift of a peak, γw, large amplitude motions that are frozen below Tg and thus
toward higher frequencies as the temperature increases. Its associated with the dynamic glass transition. From 50 °C, its
magnitude decreases from −10 °C. Moreover, it is magnitude suddenly decreases and vanishes as the sample re-
dissymmetric, and a broad shoulder can be seen on its high- crystallizes. One can notice in Figure 4b that, during
frequency flank. A second recording in the same temperature crystallization at 52 °C, the maximum of the relaxation slightly
range has been performed. The spectra obtained at the shifts toward high frequency, which indicates a slight
previous temperatures are added in Figure 4a (dashed lines). acceleration of the dynamics of the remaining amorphous
The γw peak has disappeared revealing at higher frequencies a fraction. This prevents a proper analysis of the α mode of
broad peak, γ, less intense in addition to a shoulder, δ, that is partly amorphous samples at higher temperature. In addition,
visible at high frequency for the lowest temperature. The at low frequencies, a tail of conductivity and Maxwell−
decrease in the magnitude of the γw mode from −10 °C, in the Wagner−Sillars effects appear when the sample starts to
first recordings to 25 °C, and its disappearance in the second crystallize.
recordings indicate that the corresponding mobility is linked to The relaxation times characteristic of the main relaxation
traces of adsorbed water. This kind of mobility related to and the different secondary relaxations extracted from the HN
adsorbed water is usually found for compounds amorphized by fits are reported in an Arrhenius diagram in Figure 5.
milling, owing to their enhanced reactivity.33,34 Here, this Concerning the main relaxation, α, one can see that the
phenomenon could also be reinforced by the fact that the evolution of the relaxation time with the reciprocal temper-
milled powder has been directly put on cold electrodes to ature is not linear, even if few points are available, owing to the
avoid an overheating at room temperature of the sample, which re-crystallization. It is usually described using the Vogel−
could catch some moisture of the surrounding air. Moreover, Tammann−Fulcher−Hesse (VTFH) law35−37
366 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

where τα is the relaxation time and βKWW is a parameter varying


from 1 to 0 as the relaxation mode departs from the
exponential (Debye) decay. This parameter can be calculated
from the HN shape parameters, αHN and βHN, through the
relation50
βKWW = (αHNβHN)1/1.23 (6)
We found, in the vicinity of Tg, βKWW = 0.56 ± 0.01. This
value indicates a relatively large degree of non-exponentiality.
It is in agreement with the high fragility and the anti-
correlation law established by Böhmer and co-workers.51
Concerning the different secondary relaxations, the Ar-
rhenius diagram in Figure 5 shows that the associated
relaxation times all obey an Arrhenius law
Ea
log τmax = log τ∞ +
ln(10)RT (7)
Figure 5. Relaxation map of bezafibrate showing the α, γw, γ, and δ
relaxations observed for the sample amorphized by milling at −10 °C where Ea is the activation energy of the involved process and R
(full symbols). The relaxations observed for the samples milled at is the perfect gas constant. The fitting parameters are reported
room temperature are also added (open symbols): the γ mode for tmill in Table 1. The activation energy of δ relaxation is 30 ± 1 kJ
= 1 h (green open circles) and the γ′ mode for tmill = 4 h (red open
circles). The lines represent the corresponding fits (see text).
Table 1. Parameters of the Fit with the Arrhenius Law of the
B Relaxation Times Associated With the Different Secondary
log τmax = log τ∞ + Relaxations Observed for the Material Milled for 14 h at
ln(10)(T − T0) (3) −10 °C (Glassy: δ, γ, and γw Modes) and at Room
Temperature (β Crystalline: γ′ Mode)
where τ∞ would be the relaxation time at infinite temperatures,
B is a parameter that accounts for the curvature of the mode log τ∞ Ea (kJ mol−1)
evolution, and T0 is the so-called Vogel temperature at which δ −14.0 ± 0.3 30 ± 1
the relaxation time would diverge. In a first approximation, log γ −14.91 ± 0.09 48 ± 1
τ∞ was fixed at −14 to reduce uncertainties in the fit of the γw −15.6 ± 0.2 57 ± 1
data with the VTFH law (due to the lack of points at high γ′ −14.0 ± 0.2 50 ± 1
temperature). It leads to B = 1656 ± 30 K, T0 = 264 ± 1 K,
and an extrapolated value of T = 36 ± 1 °C for τ = 100 s. This
latter value is in nice agreement with the glass transition found mol−1. If we refer to the work of Brás and co-workers on
by DSC (Tgonset = 35 ± 1 °C and Tgmid = 39 ± 1 °C). The ibuprofen,52 the corresponding motion of the δ mode could be
Vogel temperature is 45 °C below this extrapolated Tg, in fluctuations of the carboxylic group as the same activation
agreement with the 40−50 °C commonly observed.38 The energy was found for this motion. For γ relaxation, observed
degree of departure from the Arrhenius law is usually after dehydration, the value is higher (48 ± 1 kJ mol−1). The
quantified through the “strength parameter”, D, related to B mode is quite broad and spreads over more than five frequency

ÄÅ É
by D = B/T039 or to the fragility, m, defined as40−42 decades. The HN shape parameter αHN slowly increases from
ÅÅ d log τmax(T ) ÑÑÑ
Å ÑÑ
0.24 at −80 °C to 0.45 at 20 °C (βHN = 1). Taking into
m = ÅÅÅ ÑÑ
ÅÅÇ d(Tg/T ) ÑÑÖ
BTg account the higher activation energy and the broadness of the
= mode, the corresponding motion could involve the central part
T = Tg
ln(10)(Tg − T0)2 (4) of the molecule carrying the amide group (cf. Figure 1). For
the γw relaxation appeared in the first scan and related to a
The VTFH fitting parameters lead to D = 6.3 ± 0.2 and m = small amount of adsorbed water, we found an activation energy
110 ± 12. Values of m can vary from 16, when no deviation of 57 ± 1 kJ mol−1. Because it appears close from the γ mode,
from the Arrhenius law is observed for the compounds it could be due to the same type of intramolecular motion but
qualified as strong,39 to 17043 for the compounds qualified as slowed down by some hydrogen interactions between the
fragile for which the deviation is the most important.39 water molecules and the oxygen and nitrogen atoms of the
Bezafibrate thus appears as a quite fragile glass-former because amide group.
its value is comparable to that of compounds considered as These secondary relaxations are of intramolecular type. It
very fragile materials like flurbiprofen (m = 113),44 terfenadine should be noted that another type of secondary relaxation can
(m = 112),45 N-acetyl-α-methylbenzylamine (m = 111),46 or be found: those of intermolecular origin (so-called β relaxation
celecoxib (m = 110).47 of Johari-Goldstein type)53,54 that involve the motion of the
The non-exponential character of the main relaxation is entire molecule and are regarded as a precursor of the
usually described in the time domain using the Kohlrausch− cooperative α relaxation according to Ngai’s coupling

ÄÅ É
Williams−Watts function48,49
ÅÅ i y βKWW ÑÑÑ
model.55−57 Such relaxation appears either as a well-resolved

ÅÅ jj t zz ÑÑ
φ(t ) = expÅÅÅ−jj zz ÑÑ
peak or as a change in slope (excess wing) on the high

ÅÅ jk τα z{ ÑÑ
frequency flank of the α relaxation.58 In the case of bezafibrate,
ÑÑÑ
ÇÅÅ Ö
this type of relaxation could not be observed (possibly because
(5) of the rapid re-crystallization of the sample).
367 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

3.2. Milling at Room Temperature (Tmill = Tg15 °C). peaks characteristic of the β form are observed. Therefore, the
3.2.1. Structural and Thermodynamic Analyses. Bezafibrate milling of the α form at room temperature for at least 2 h
has been milled at room temperature for a time, tmill, varying allows to obtain the β form of bezafibrate. In the authors’
from 5 min to 14 h. Some amounts of milled powder were knowledge, this is the only way to obtain this form directly
regularly sampled from the milling jars and immediately from the commercial product, without the use of solvents. A
analyzed by XRPD and MDSC. deeper analysis of XRPD spectra was performed using Rietveld
Some of the XRPD patterns recorded at room temperature refinements. It permits to determine from the width of the
are reported in Figure 6. Before milling, only the Bragg peaks Bragg peaks the evolution upon milling of the size of
crystallites and, from the peaks’ relative intensity evolution
and the background evolution, the proportion of each phase
and amorphous content. Both evolutions are reported in
Figure 7. One can see (in the inset) that, after only 5 min of
milling, the size of the α crystallites decreases from
approximately 2000 Å to less than 500 Å and it reaches a
plateau value of 250 Å after around tmill = 50 min. An
interesting point is the evidence of some amount of the
amorphous phase that appears from the start of the milling
process. The exact proportion of amorphous phase is difficult
to be precisely determined but it seems to be of the order of
10%. Then, after around 1 h of milling, the β phase appears
and develops for approximately another 1 h of milling, whereas
the amounts of the crystalline α phase and of the amorphous
phase decrease. After tmill = 2 h, the milled powder is 100% in
the β form and no change occurs until the end of the milling
(tmill = 14 h). As the β phase appears, the corresponding
Figure 6. XRPD patterns recorded at room temperature for crystallite size first starts to increase to around 1000 Å and then
bezafibrate milled for increasing milling times (indicated on the rapidly decreases to reach a plateau value of 500 Å after tmill = 2
right) at room temperature, showing the conversion of the h and does not evolve anymore (cf. inset).
commercial α form (tmill = 0 min) to the β form (tmill = 850 min) Milled powder was also analyzed using MDSC. Some
upon milling. thermograms are reported in Figure 8. Three different types of
behavior are observed in the total heat flow. First, for milling
corresponding to the commercial α form are observed. From times shorter than 1 h, an exothermic peak appears between 45
the first minutes to 70 min milling, these Bragg peaks broaden and 70 °C and tends to develop with milling time. This peak is
and their magnitude decreases. Then, from 90 to 130 min, new located just above the glass transition range, and it indicates a
Bragg peaks characteristic of the β form appear and develop re-crystallization of the sample. The appearance of such an
whereas those characteristic of the α form disappear. From 130 exotherm of re-crystallization is an indirect proof of
to 850 min milling, there is no noticeable evolution and only amorphization. Nevertheless, the Cp jump characteristic of a

Figure 7. Evolution vs the milling time at room temperature of the fraction of the different phases: α phase (full blue squares), β phase (full red
circles), and amorphous phase (full green triangles). The evolutions of the size of α (open blue squares) and β (open red circles) crystallites are
represented in the inset. Lines are guide for the eyes.

368 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

flow signal shows no event. This indicates that no amorphous


fraction is detected and that the milled compound is crystalline
(in β-form), in agreement with XRPD results. At higher
temperatures, from 145−150 to 170 °C, an endothermic peak
appears shifting toward high temperature with milling time. It
corresponds to the solid−solid conversion of the β form to the
α form, and above approximately 175 °C, the endotherm of
melting of this α form can be seen. One can notice on the
high-temperature side of the endotherm of conversion, a
shoulder that develops into a peak as the milling time increases
and an exothermic peak at the foot of the endotherm of the α
melting. The shoulder could be due to the melting of β
crystallites that are not yet converted (more numerous because
the conversion is delayed) but the corresponding endotherm of
melting could be partly masked by a rapid re-crystallization to
the α form at the origin of the exothermic event. It should be
noted that the α melting temperature onset is the same as that
of commercial bezafibrate, which indicates that α crystallites
Figure 8. MDSC thermograms recorded at 5 °C min−1 ± 0.530 °C resulting from the conversion have larger sizes than those
per 40 s (total heat flow) of bezafibrate milled at room temperature directly resulting from shorter milling times.
for increasing milling times (indicated on the left). Finally, for milling times ranging from 90 min (cf. Figure 8)
up to 120 min, thermograms present both a thin exothermic
glass transition cannot clearly be observed on the reversing peak of re-crystallization between 40 and 60 °C, which
heat flow: the re-crystallization occurs as soon as the Cp jump indicates a partial amorphization of the sample upon milling,
begins and the mobility is released. For most of the recordings and an endothermic peak of conversion of the β form to the α
and despite the use of different modulation parameters, only form, which also reveals the existence of β crystallites, in
the start of the Cp jump is observed, but then, the signal is agreement with XRPD data.
disturbed by the rapid re-crystallization. The Cp jump is all the It should be pointed out that whatever be the milling time
more difficult to observe as it is quite small, which indicates longer than 50 min, a close inspection of thermograms reveals,
that only a small fraction of the sample has been amorphized. from around 60 °C until the endotherms of the α form melting
In the cases where the reversing heat flow signal was the least or the β → α transition, a shallow but very broad exothermic
disturbed (tmill = 60 min), tentative analysis of the Cp jump basin. Its origin will be discussed later (see the discussion part
leads to values of amorphous content around 15%, which is of 4.2).
the order of the amount extracted from XRPD analysis. At 3.2.2. Dynamic Analysis. 3.2.2.1. Milling Time of 1 H. The
higher temperatures, the endotherm of melting of the α form sample milled for 1 h at room temperature was analyzed by
clearly appears between 180 and 190 °C. It confirms that, after DRS. As for the sample milled at −10 °C, the evolutions of the
milling, the crystalline part of the sample is still in the α form real and imaginary parts of the permittivity were first recorded
and it indicates that a small amorphous content has re- at low temperature between −90 and 25 °C upon two
crystallized to this form. It should be noted that the melting successive scans. In spite of a very noisy and weak response, a
temperature onset is depressed by 3 °C from that of the relaxation mode can be observed as illustrated in Figure 9a.
commercial product, which can be explained by the smaller This reveals the existence of some kind of residual mobility.
size of crystallites leading to a Gibbs−Thomson effect. It should be noted that the two successive scans are a bit
Second, for milling times longer than 120 min (cf. Figure 8), different: in the second scan, the low-frequency flank of the
the previous exothermic peak located in the glass transition mode is lower as can be seen in Figure 10 where, as an
temperature range is no longer visible and the reversing heat example, the two spectra recorded at −20 °C are reported (in

Figure 9. 3D representation of tan δ recordings: (a) for bezafibrate milled for 60 min at room temperature on isotherms between −90 and 25 °C
(every 5 °C) showing the γ relaxation; (b) for bezafibrate milled for 4 h at room temperature on isotherms between −120 and 50 °C (every 5 °C)
showing the γ′ relaxation.

369 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

Figure 10. Recordings of tan δ (F) at −20 °C. In black (left axis), bezafibrate amorphized by milling at −10 °C: freshly milled (full symbols) and
after heating until 25 °C (open symbols). In red (right axis), bezafibrate milled for 60 min at room temperature: freshly milled (full symbols) and
after heating until 25 °C (open squares), 50 °C (open circles), and 110 °C (open triangles).

Figure 11. Evolution of tan δ vs temperature between 10 and 75 °C for different frequencies of the applied electrical field in the case of bezafibrate
freshly milled for 60 min at room temperature. It shows the collapse of the signal for the lowest frequencies (0.1, 1, and 11 Hz) between 42 and 50
°C.

red, full and open squares). For comparison, the two scans function. Because of the noisy recording, this could be carried
recorded at the same temperature in the case of the sample out only between −55 and −15 °C. The evolution of the
amorphized by milling at −10 °C are also reported (in black, extracted relaxation time is reported in the Arrhenius diagram
full and open squares). The vertical scale is not the same as in Figure 5 (green open circles). As expected, it appears very
these are far more intense, but the position of the mode is close to that of the γ mode obtained for the glass of bezafibrate.
similar. This suggests that the detected mobility has the same In the second step, the evolution of the permittivity was
origin as for the glassy sample. Therefore, the relaxation mode recorded for isotherms above 25 °C, in particular, between 35
observed in the second recording would be the γ relaxation, °C and 50 °C, that is, the temperature range where the α mode
and the difference between the two successive scans would could be detected at low frequencies in the case of the
result from the loss of some adsorbed water. It should be noted amorphous sample. In the obtained scans, at high frequencies,
that, due to the very small magnitude of the signal, it is difficult the γ relaxation mode can still be detected, and as the
to determine whether the observed mode is due to only one frequency decreases, both the real and imaginary parts of the
relaxation process (γ) or two (γ and δ) like for the glass. In a permittivity increase. This increase is certainly due to
first approximation, the mode was fitted with one Cole−Cole Maxwell−Wagner−Sillars effects resulting from the hetero-
370 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

geneity of the sample. However, a close inspection of the scans intense and noisy in the first recording, which is certainly due
recorded between 42 and 48 °C reveals at low frequency some to traces of adsorbed water. Therefore, a relaxation related to
incidents of the increase in the ε″ value. To clear things up, the some adsorbed water appears at lower frequency (longer
evolution of tan δ with temperature was analyzed (cf. Figure relaxation time) than the γ′ mode. The evolution of the γ′
11). One can observe that, at 0.1 Hz, tan δ increases from 30 relaxation time follows an Arrhenius behavior, and an
to 42 °C and then decreases till approximately 50 °C, resulting activation energy of 50 ± 1 kJ mol−1 is found (log τ∞ =
in a peak. At 1 and 11 Hz, the same behavior is observed but −14.0 ± 0.2) for tmill = 4 h. It should be underlined that this
the magnitude of the peak decreases with increasing frequency result does not seem to depend on milling times (see
and is slightly shifted toward higher temperature. Above 500 Supporting Information S3), taking into account the small
Hz, this phenomenon disappears. The considered temperature fluctuations that can result from the analysis of the noisy
range is the zone where, in the case of the amorphous sample, recordings.
the α mode appears before the re-crystallization occurs (from Concerning the evolution of the permittivity recorded for
50 °C). Thus, the sudden increase in tan δ observed in Figure isotherms above 25 °C, for all the studied milling times, in the
11 from 30 °C for the lowest frequency can reasonably be temperature range 35−60 °C, the γ′ relaxation is observed at
attributed to the appearance of the α mode, leading to a peak. high frequency and, like in the case of the sample milled for 1
The anomalous decrease in its magnitude with increasing h, for decreasing frequencies, both the real and the imaginary
frequency could be explained by re-crystallization of the parts of the permittivity increase due to Maxwell−Wagner−
sample, which agrees with the exotherm of crystallization Sillars effects. However, this time, no accident appears at low
observed by DSC. This would result in the disappearance of frequencies, in particular between 40 and 50 °C and no trace of
the α mode at higher temperature that should be observed for an α relaxation or of a sudden re-crystallization can be detected
higher frequencies. So, even if the main relaxation assigned to (even in tan δ vs the temperature evolution), in agreement
large-amplitude motions cannot clearly be evidenced as a well- with DSC results. Therefore, in the case of bezafibrate milled
resolved relaxation, its existence masked by a rapid re- for more than 2 h at room temperature, only a localized
crystallization is highly suspected. It corroborates the DSC mobility is evidenced. It should be added that, like for the
and X-ray results indicating a partial amorphization of the sample milled for 60 min, this mobility is still detected,
sample. although decreasing, after further heating of the sample to 110
It should be added that, interestingly, after heating to 50 °C °C (data not shown).
leading to a complete re-crystallization of the sample (α-form)
and the disappearance of the α relaxation, the recordings of the 4. DISCUSSION
permittivity on isotherms below the room temperature 4.1. Mechanisms of Transformations Induced by
(between −90 and 25 °C) still exhibit the secondary γ Milling. The results presented in part 3.1. have shown that a
relaxation. This is illustrated in Figure 10 for the spectrum at 14 h milling performed at −10 °C leads to the amorphization
−20 °C (red open circles). As can be seen, the magnitude of of bezafibrate. This temperature is about 50 °C below the glass
the mode is lower but the mode is still clearly detected. It does transition temperature of the compound. Therefore, the
not disappear even after further heating to 110 °C (cf. Figure observed behavior is in agreement with results previously
10, red open triangle). reported on other materials: when milling is performed well
3.2.2.2. Milling Time Longer than 2 H. Some samples below the glass transition of the compound, it amorphizes.15
milled for a time longer than 2 h were analyzed by DRS. Like Indeed, the physical state reached by milling results from a
before, the evolutions of the real and the imaginary parts of the disorganization of the crystalline network due to ballistic
permittivity were recorded from 106 to 10−1 Hz at low shocks, the re-crystallization being blocked by the too weak
temperature for isotherms ranging from −120 to 50 °C (two molecular mobility of the glassy state.59
successive series of recordings). As an example, the tan δ In part 3.2, it was shown that bezafibrate resulting from a 14
evolution (second series of scans) obtained for bezafibrate h milling at room temperature is still crystalline but
milled for 4 h at room temperature can be seen in Figure 9b. polymorphic transformation to the β form has occurred. In
As for the sample milled for 1 h, the response is very noisy and this case, although the milling is performed 15° below Tg,
weak but the striking point is that a relaxation mode, γ′, is there is no effective amorphization even for long milling times.
detected in spite of the sample being fully crystalline (β form) Such a behaviour has been previously reported for compounds
according to XRPD and DSC analyses. This, again, emphasizes such as indomethacin,60 biclotymol,61 or glucose.59 However,
the existence of some kind of residual mobility. The same for glucose, it was reported that a milling at Tg13 °C leads
response (with comparable magnitude) is obtained for other to non-amorphization, which is only apparent because it results
milling times. This relaxation process, γ′, has been fitted with from a re-crystallization process made more efficient by the
the Cole−Cole function, and the evolution of the associated increase in temperature that counterbalances the amorphiza-
relaxation time is reported in the Arrhenius diagram in Figure tion due to ballistic shocks.59 Therefore, obtaining a
5 (red open circles). It appears very close from the γw mode completely crystalline bezafibrate after a 14 h milling at
observed in the case of glassy bezafibrate, only in the first room temperature does not mean that transient amorphization
recording, and assigned to movement of a small part of the does not occur, and the analysis of the milling kinetics permits
molecule linked through hydrogen bonds to some water to better understand the mechanisms of the transformation.
molecules. However, for bezafibrate milled at room temper- Indeed, the XRPD study of the beginning of the milling
ature for tmill > 2 h, it is quite unlikely that the origin of the process kinetics (cf. Figure 7) reveals not only a rapid decrease
relaxation is the same. Indeed, no significant difference was in the average size of crystallites, which is due to their
observed between the first and the second recordings except fragmentation by the impacts of milling balls, but also the
for samples that were stored in a freezer after milling: for these appearance of some amount of the amorphous phase. The
samples, the low-frequency flank of the γ′ mode is more existence of this small amorphous content was confirmed by
371 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

the occurrence of a re-crystallization exotherm in DSC between 110 and 130 min of milling and the plateau value
thermograms on consecutive heating (cf. Figure 8). For a obtained for longer milling times suggest that, like for the α-
sample milled for 60 min, a small Cp jump can even be crystallites, the milling tends to amorphize the β phase but the
observed, and from a close inspection of DRS spectra, the mobility allows the rapid re-crystallization (faster than the
existence of the α relaxation associated to the dynamic glass amorphization), and the β-crystallite size decreases to a steady-
transition is highly suspected (cf. Figure 11). Therefore, it state plateau value, resulting from a balance between
appears that, like at −10 °C (Tg50 °C), milling at 25 °C fragmentation by mechanical shocks and a cold welding
(Tg15 °C) induces a real amorphization of bezafibrate. process.76
However, the fact that after 50 min of milling, the crystallite It is stunning that the milling process has to last more than 1
size reaches a plateau value of 250 Å (cf. Figure 7), a value h to induce the polymorphic transformation of the α form to
close to 100−200 Å corresponding to the critical size for the β form. Indeed, the α and β phases form an enantiotropic
spontaneous amorphization,62 indicates the simultaneous re- system, and the β phase is the stable one below the transition
crystallization in the initial α form, which is induced by an temperature (∼150 °C) and thus at the milling temperature.
enhanced mobility. It leads to a stationary state composed of Moreover, with the temperature of the transition and the
approximately 10% amorphous phase and 90% crystalline α associated enthalpy change, it is possible to calculate the
form. The type of mobility involved in the re-crystallization of difference in Gibbs free energy of the two phases at room
glassy materials (below Tg) is still an issue under debate in temperature assuming that the entropy change, ΔSconv,
spite of the extensive studies devoted to the subject. The associated with the transition is equal to the ratio of the
structural relaxation (α relaxation) appears to be the main transition enthalpy, ΔHconv, and the temperature, Tconv. With
factor because, in the last years, studies on different the values found by Lemmerer and co-workers (Tconv = 157 °C
compounds47,63−65 have highlighted, based on predictions of = 430 K and ΔHconv = 4.3 kJ mol−124 ), it leads to ΔSconv = 10 J
τα in the glassy state, correlations between this relaxation and mol−1. Therefore, at room temperature (T = 25 °C = 298 K),
the kinetics of re-crystallization. Other studies indicate that the the difference in Gibbs free energy is Gα − Gβ = ΔHconv −
implication of more localized motions of intermolecular origin TΔSconv = 1.32 kJ mol−1. This value is very small, so one could
(β relaxation of Johari-Goldstein type) cannot be excluded.66,67 think that the α phase would easily transform to the more
Moreover, Descamps et al. have shown that milling produces stable β phase as soon as the milling process begins, but it
glasses of high energy with enhanced sub-Tg mobility.68 needs more than 1 h. One should note however that the fact
Unfortunately, in our case, the evolution of τα and τβJG below that commercial bezafibrate is in the metastable α form
Tg could not be extracted from the DRS experiments, which indicates that the transformation to the β phase does not occur
prevents us to conclude which of these mobilities allows, so easily one could expect. This means that, although the
during milling at room temperature, the re-crystallization of difference in Gibbs free energy between the two forms is small,
bezafibrate counterbalancing the amorphization due to the energy barrier that has to be overcome so that the
mechanical shocks. transformation occurs is high.
Concerning the re-crystallization, it should be underlined As shown by Lemmerer and co-workers,24 the propanoic
that it occurs to the initial metastable α phase. The existence of acid groups of bezafibrate molecules of the α phase are in syn-
a transient amorphous state and the re-crystallization to the conformation whereas they are in anti-conformation for the β
metastable crystalline state have been highlighted in several phase. This different conformations of the propanoic acid
studies of polymorphic transformation under milling at a groups induce different H bonding of the carboxylic acid
temperature close to Tg (in monotropic situations).69−72 It is groups, which results in different conformations (different
usually explained by Ostwald’s rule of stages:73,74 re- torsion angles) of the propanoic acid group bonds. A change in
crystallization proceeds through the fastest mechanism each of these torsion angles requires a large amount of energy
(which does not lead to the most stable form). It could be because of repulsive interactions with the rest of the molecules.
underlined for the following discussion that exception seems to Therefore, it implies, according to Lemmerer et al.,24
be the case of D-mannitol, for which no intermediate correlated rotations of the different bonds of the propanoic
amorphous phase was detected during milling at Tg + 7 °C acid group. Therefore, due to the different and complex H
(in spite of in situ XRPD measurements).75 bonding networks of the α and β crystalline phases, the
However, in the case of bezafibrate, what is puzzling is that polymorphic transformation appears to be of the reconstruc-
the XRPD analyses of the milling process kinetics reveal a tive type. Under milling at 25 °C, the existence of a transient
second stage in the behavior under milling at room amorphous phase highlighted in the present work could allow
temperature: after more than 1 h of milling, polymorphic this transformation to occur with the release of some
transformation to the stable β form occurs because for sample cooperative large amplitude motions (main relaxation,
milled for 90 min, the β form is detected for the first time (cf. evidenced by DRS for tmill = 60 min). Besides, one can note
Figure 7). Then, for an additional short milling time (∼20 that when the α, β, and amorphous phases co-exist (after a
min), the two crystalline forms and the glassy form co-exist. milling of 90 or 110 min, e.g.), a spontaneous transformation
However, during this time, the content of the β form increases of the remaining α fraction to the β form is not observed. For
at the expense of the contents of the α and glassy forms, with example, after a 2 day storage at room temperature of the
the size of β-crystallites being around 1000 Å: with milling, the sample milled for 110 min, a XRPD recording shows that the
amorphous content and the remaining α-crystallites, which amorphous fraction has re-crystallized but without a noticeable
keep on amorphizing, re-crystallize to the β form, which increase in the β fraction at the expense of the α fraction.
explains the quite high value of β-crystallite size. After 130 min Another example is given by the analysis of the endothermic
of milling, when the sample is completely transformed to the β peak of the β → α conversion observed in DSC around 140 °C
form, the size of β-crystallites falls to a value of 500 Å that does when the sample milled for 90 min is heated (cf. Figure 8). On
not evolve for the following 12 h milling: the decrease in size comparing the area of this peak with the one observed in the
372 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

Figure 12. Milling of bezafibrate at room temperature for 70, 110, and 290 min: evolution vs the temperature of the size of crystallites for the α
form (blue squares) and the β forms (red circles). Lines are guide for the eyes. The dashed areas indicate the β → α conversion temperature range.

case of bezafibrate re-crystallized in the β form after complete associated with large-amplitude motions that are frozen
amorphization at −10 °C (assumed to represent the below Tg, only for the sample milled for 60 min, in spectra
transformation of a 100% β crystalline bezafibrate), we found recorded just after milling (before it disappears when the
an amount of β phase corresponding to the sum of the β and sample re-crystallizes on heating to 50 °C). These results are in
amorphous fractions determined at the end of the milling agreement with the DSC analysis showing a re-crystallization
process from the XRPD analysis (ΔHconv/ΔHconv 100% = 16%). exotherm around 60 °C that partly masks a Cp jump and the
This suggests that after milling and upon heating till 140 °C, XRPD analysis reporting an amorphous content of around
the amorphous fraction has re-crystallized in the β form but the 10%. For longer milling times (4−11 h), the α mode cannot be
α phase did not transform further to the β form. This detected, which emphasizes the absence of any amorphous
corroborates that the conversion under milling of the α form to fraction, in agreement with DSC and XRPD results. Indeed,
the β form is not direct and that a transient amorphous phase the sensitivity of DRS is such that it can reveal very low
(reached by further milling) is required to continue the content of the amorphous phase as in the case of a recent
transformation of the α form to the β form due to the required dielectric study, where an α relaxation due to 1 to 2% of the
energy. It could be remarked, by the way, that in the case remaining amorphous phase confined by surrounding crystal-
previously mentioned of D-mannitol, presenting no transient lites could be detected, whereas no sign was visible by DSC or
amorphous state, the polymorphic transformation is of XRPD.45
displacive type, which certainly requires less energy to occur.75 Concerning the more localized intramolecular mobility, the
Then, one can wonder why, although an amorphous fraction DRS study reveals interesting original results for bezafibrate
is detected in the sample already after 10 min milling, this milled at room temperature. For the sample milled for 60 min,
fraction does not re-crystallize in the β form but in the α form. a weak relaxation, the γ mode, is detected even after the re-
When we consider the amorphous content of milled samples crystallization of the sample (heating at 50 °C resulting in the
deduced from XRPD (cf. Figure 7) or DSC (cf. Figure 8, disappearance of any trace of an α mode). The activation
development of the exotherm of re-crystallization around 50 energy and the temperature dependence of the associated
°C), we observe that it reaches its maximum when the β phase relaxation time are in agreement with those of the γ relaxation
appears, just after in XRPD or just before in DSC, depending if observed in the case of completely amorphous bezafibrate
the technique probes the structural disorder (XRPD) or the (resulting from milling at Tg50 °C). For samples milled for
dynamic disorder (DSC). This points out that the trans- 4 h or 11 h (completely crystalline), a weak γ′ relaxation is also
formation under milling of the α form to the β form requires detected. These remaining localized mobilities cannot thus be
not only a transient amorphous phase but also in sufficient assigned to the remaining of the, strictly speaking, amorphous
quantity for the remaining α phase to be defective enough phase. Moreover, such localized residual mobilities are
(coherent domains small enough, network of H bonding detected neither for commercial α crystalline bezafibrate nor
enough disorganized) so that the transformation from the for β crystalline bezafibrate obtained from re-crystallization on
metastable α phase to the stable β form occurs and the new H heating of the amorphous state or for α crystalline bezafibrate
bonding network rebuilds. obtained on heating from the conversion of the β form
4.2. Origin of the Remaining Mobility in the resulting from long milling times. Thus the localized mobilities,
Crystalline Milled Samples. In the case of bezafibrate γ or γ′, detected in crystalline milled samples are not due to an
milled at room temperature, the results of the DRS intrinsic mobility of the crystal but seem to be rather generated
experiments suggested the existence of the α relaxation by the milling. Besides, when the samples milled at room
373 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

temperature are heated, these residual mobilities decrease and γ′ relaxations strongly suggest that the residual localized
because the magnitude of the γ and γ′ modes clearly decreases mobilities are due to motions of part of molecules located at
(cf. Figure 10, heating up to 110 °C for the γ mode). the surface of crystallites.
To better understand the origin of the residual localized DRS results help us to be more precise concerning the
mobility, samples of bezafibrate milled at room temperature involved motions associated with the γ and γ′ relaxations. We
have been analyzed by XRPD when heated. The evolution with have seen in the dielectric study of bezafibrate amorphized by
temperature of the size of crystallites extracted from Rietveld milling (part 3.1) that, taking into account the activation
refinements is reported in Figure 12 for milling times of 70, energy and the broadness of the γ mode, this relaxation could
110, and 290 min. It shows a continuous increase in the result from the motion of the central part of the molecule
crystallites’ size until temperatures closed to the melting (not linked to the amide group. The development on the low-
shown) of the α phase (tmill = 70 min) or until the conversion frequency side of the γw mode, only visible during first heating
of the β phase to the α phase (tmill = 110 and 290 min) occurs. to 50 °C, was assigned to the same motion slowed down by
Moreover, as mentioned in part 3.2.1, a close inspection of some hydrogen bonds between the amide group and a small
DSC thermograms reveals a broad exotherm in this temper- amount of adsorbed water molecules. Concerning the γ′ mode
ature range. As shown in Figure 13, this exotherm starts observed for bezafibrate samples converted by milling to the β
form, the fit with the Arrhenius law of the associated relaxation
times gives an activation energy of the corresponding
movement, which is roughly the same as that of the γ
relaxation (Eaγ′ = 50 ± 1 kJ mol−1 and Eaγ = 48 ± 1 kJ mol−1).
Besides the γ′ relaxation times are of the order of those found
for the γw mode. However, in this case, water molecules cannot
be involved as a slight adsorption of water has an effect at even
lower frequencies. It is therefore tempting to assign the γ′
mode to the motion of the central part of bezafibrate
molecules, as for the γ mode, but slowed down by the
development of hydrogen bonds linking the amide group, not
with water molecules, but with other bezafibrate molecules. To
go further, it is necessary to consider the possible hydrogen
bonds between bezafibrate molecules that could involve the
amide group. Lemmerer et al.24 have shown that, in the
crystalline state, the differences in conformations of the
carboxylic hydrogen atoms (syn-conformation in the α form
and anti-conformation in the β form) result for the two
polymorphs in different and complex H bonding networks. For
the α form, bezafibrate molecules form cyclic dimers through
two OH···O hydrogen bonds between the carboxylic acid
group of one molecule and the oxygen atom of the amide
group of the other molecule. These dimers are inter-connected
to two adjacent dimers by NH···O hydrogen bond between the
Figure 13. MDSC thermograms recorded at 5 °C min−1 ± 0.530 °C amide group and the carboxylic O atom, forming tetrameric
per 40 s (total heat flow) of bezafibrate milled at room temperature
for 1 h (top) and 4 h (bottom). Top: scans recorded at the very end enclosed rings. This leads to a packing structure with one-
of milling (in blue), after a few minute storage at room temperature dimensional tube-shaped chains.24 For the β form, bezafibrate
(in green), and after heating until 110 °C (in red). Bottom: scans molecules are linked through NH···O hydrogen bonds
recorded at the end of milling (in green) and after heating until 110 between the amide group and the carboxylic O atom, forming
°C (in red). In both figures, the thermogram of the crystalline chains connected to adjacent ones by OH···O hydrogen bonds
unmilled sample is added for comparison (in black). between the carboxylic acid group and the oxygen atom of the
amide group. This results in tetramers forming a two-
around 60−70 °C and extends till the melting of the α phase dimensional hydrogen-bonded sheet network.24 The oxygen
(case of tmill = 60 min) or until the β → α transformation (case atom and the NH function of the amide group are both
of tmill = 4 h). It is interesting to note that, when the heating of involved in these hydrogen bonds, irrespective of the
the sample is stopped at a temperature in this temperature crystalline polymorphs. One can therefore suspect that it
range (e.g., 110 °C in Figure 13), the exotherm starts from this blocks the mobility of the amide group and the linked −CH2−,
temperature on the subsequent heating. This exotherm is which is consistent with the fact that no localized mobility is
characteristic of the coarsening of crystallites. The heat given detected by DRS in the nondefective crystalline forms. When
off corresponds to a decrease in the interfacial enthalpy that is the crystalline structure is made highly defective by milling (α
induced, upon heating when the mobility is high enough, by or β forms), molecules on the surface of crystallites are
the cure of crystalline defects and a coalescence of crystallites. certainly only partially involved in the network of hydrogen
This leads to an increase in the average crystallite size observed bonds, which would partially release the local mobility,
by XRPD (cf. Figure 12). These results show that the resulting in the detection of the γ and γ′ modes. The
crystalline phases generated by milling (α or β forms) are difference between these two modes could be due to a
highly defective. Besides, when heated, the highlighting of the difference in the hydrogen bonding in which the oxygen atom
coarsening of crystallites (which induces a decrease in the total and the NH function of the amide group are involved, its
surface area) and a coupled decrease in the magnitude of the γ motion coupled to that of the linked −CH2− being more or
374 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

less hindered. When the material is amorphized by milling, the form (Tg = 39 ± 1 °C and ΔCp = 0.41 ± 0.01 J g−1 °C−1),
crystalline network and thus the hydrogen bonding network which is chemically pure unlike the glass obtained from the
that stabilizes it are destroyed and the mobility of the central quenching of the melt, whereas the same milling performed at
part of the molecule is fully released. It is therefore intriguing 25 °C (Tg15 °C) enables to obtain the β polymorph (stable
that the same relaxation mode is found for the sample milled form at room temperature). This nicely illustrates how varying
for 60 min (defective α crystalline form) and the glassy sample. the milling temperature permits to play with the physical state
An explanation could arise from the calculation of the of compounds. Moreover, the monitoring under milling of the
hydrogen bond energies for the different conformations kinetics of this latter transformation allowed us to shed some
made by Lemmerer and co-workers24 showing that, for the α light on the mechanisms of the induced conversion of the
form, the largest stabilization effect comes from the two OH··· metastable to the stable form of this enantiotropic system. It
O hydrogen bonds between the carboxylic acid group and the was evidenced that at first, two antagonist processes occur: an
oxygen atom of the amide group, whereas the NH···O amorphization (due to mechanical shocks) partly counter-
interaction contribution is small. It is thus likely that, when balanced by a slower re-crystallization (enabled by the
milled, the NH···O hydrogen bonds are first destroyed, while mobility) to the starting α form (metastable at room
the cyclic dimers resulting from the OH···O interactions temperature). This prevails until the crystalline structure is
persist. This could then indicate that these dimers also remain disorganized enough (mean α crystallite size of 250 Å) and the
in the glassy sample. For the β form, the energy difference amorphous content is high enough (around 10%) after
between the two types of hydrogen bonds is not so approximately 1 h of milling to trigger the transformation to
important.24 Consequently, for the defective β crystalline the stable (at room temperature) β form. Once the conversion
sample, the probability that, on the surface of crystallites, the is achieved (after 2 h of milling), a further milling leads to β
NH group of the bezafibrate molecule remains H-bonded crystallites with a mean size of 500 Å. It thus appears that the
could be more important than for the α polymorph, leading to polymorphic transformation induced by milling in the case of
a more hindered local motion of the central part of the
bezafibrate requires going through an intermediate amorphous
molecule. It should nevertheless be noted that although the
state, certainly because the conversion is of the reconstructive
relaxation times extracted from the maxima of the γ modes for
type. Such transformation to the stable form triggered after 1 h
the glassy bezafibrate and the defective α form correspond, the
of milling has already been observed for sulfamerazine,21
relaxation found in the amorphous sample is broader. It could
be due to its much higher intensity but one cannot exclude that another compound for which the conversion between the two
it could result from the convolution of the preponderant γ polymorphic phases that form an enantiotropic set is also of
relaxation and also to a small extent of the γ′ relaxation. In any reconstructive type.79 Milling-induced transformation in this
case, these results raise the question of the remaining of some kind of systems thus seems to be ruled, as for monotropic
peculiar arrangements of bezafibrate molecules (cyclic dimers situations, by a transient amorphous state. However, due to the
or linear dimers, trimers...) in the glassy state. One can wonder very short lifetime of this state, only the use of complementary
whether these hydrogen-bonded arrangements, if any, would techniques probing the structure, the thermodynamic and the
be the same depending that the glassy state is obtained from dynamic, as in this work, allows to reveal its existence. These
the milling of the crystalline material, undercooling of the outcomes also underline the complexity of such trans-
liquid, or another way. It would be interesting to determine by formations and the need to follow the kinetics to fully
experimental techniques such as solid-state NMR or FTIR or understand which crystalline form can be reached and the
by molecular dynamics simulation (as degradation exper- underlying mechanisms.
imentally occurs on melting) the different types of bezafibrate Moreover, the detailed characterization by DRS of the
molecule associations that could remain in the glassy state molecular mobility of the amorphized bezafibrate has
obtained through different routes. Besides, the link between permitted us to establish that it belongs to the class of fragile
the milled polymorphic form and the conformational and liquids (m = 110 ± 12), and to highlight, in the glassy state,
molecular packing of the amorphous form obtained by milling two secondary relaxations of intramolecular origin, in addition
is still an open question. For example, in the case of piroxicam to a relaxation, γw, originating through a small amount of water
that can appear in two polymorphic phases that form in the first scans: the faster δ mode probably due to small
hydrogen-bonded dimers (form I) or continuous hydrogen fluctuations of the carboxylic group and the γ mode (Ea = 48 kJ
bond chains (form II),77 it has been shown by pair distribution mol−1) that could be assigned to movements of the central part
function analysis of XRPD78 that, under milling, the molecular of the molecule carrying the amide group. Furthermore, a
packing of form I evolves continuously to the amorphous one, striking result arose from the investigation by DRS of the
whereas in the case of the milling of form II, a transient mobility of the milled crystalline forms (α or β): a residual
disordered state that maintained the local packing of the localized mobility was evidenced that is neither due to an
crystalline polymorph was observed. Such analysis of the intrinsic mobility of the crystal nor to a strictly speaking
evolution of the local packing of bezafibrate molecules under amorphous phase but to a residual mobility at the surface of
milling of each polymorph could help to better understand the crystallites, which diminishes as the coarsening of crystallites
origin of the difference in localized mobilities evidenced in this occurs with heating and defects are cured. A difference in this
work by DRS. dynamics for the two crystalline forms could arise from a
difference in the hydrogen bonding network between
5. CONCLUSIONS bezafibrate molecules. Owing to its high sensitivity, the
In this article, the complementary use of XRPD, DSC, and dielectric relaxation spectroscopy thus appears as an interesting
DRS allowed us to carefully investigate the behavior upon tool for distinguishing the threshold between a defective
milling of bezafibrate (α form). It was shown that a 14 h crystalline state and an amorphous one. The more frequent use
milling at −10 °C (Tg50 °C) enables to obtain the glassy of this technique to characterize milled compounds could allow
375 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

a better understanding of the process-induced disorder in (6) Nagahama, M.; Suga, H.; Andersson, O. Formation of molecular
crystalline structures of material. alloys by solid-state vitrification. Thermochim. Acta 2000, 363, 165−


174.
ASSOCIATED CONTENT (7) Oliveira, P. F. M.; Willart, J.-F.; Siepmann, J.; Siepmann, F.;
Descamps, M. Using Ball Milling To Explore Physical States: The
*
sı Supporting Information
Amorphous and Polymorphic Forms of Dexamethasone. Cryst.
The Supporting Information is available free of charge at Growth Des. 2018, 18, 1748−1757.
https://pubs.acs.org/doi/10.1021/acs.cgd.1c00997. (8) Brittain, H. G. Effects of Mechanical Processing on Phase
1 Composition. J. Pharm. Sci. 2002, 91, 1573−1580.
H NMR spectra of commercial Bezafibrate, sample (9) Zhou, Y.; Du, J.; Wang, L.; Wang, Y. State of the art of
milled for 14 h at −10 °C and sample obtained from the nanocrystals technology for delivery of poorly soluble drugs. J.
quenching of the liquid, DRS recording of ε″(F) at −70 Nanopart. Res. 2016, 18, 257.
°C for bezafibrate milled for 14 h at −10 °C, and (10) Braga, D.; Maini, L.; Grepioni, F. Mechanochemical
Arrhenius diagram showing the γ′ mode observed for preparation of co-crystals. Chem. Soc. Rev. 2013, 42, 7638−7648.
bezafibrate milled at room temperature for 4 and 11 h (11) Frišcǐ ć, T.; Jones, W. Recent Advances in Understanding the
(PDF) Mechanism of Cocrystal Formation via Grinding. Cryst. Growth Des.


2009, 9, 1621−1637.
(12) Mahieu, A.; Willart, J.-F.; Dudognon, E.; Danède, F.;
AUTHOR INFORMATION Descamps, M. A New Protocol To Determine the Solubility of
Corresponding Author Drugs into Polymer Matrixes. Mol. Pharm. 2013, 10, 560−566.
Emeline Dudognon − Univ. Lille, CNRS, INRAE, Centrale (13) Dengale, S. J.; Grohganz, H.; Rades, T.; Löbmann, K. Recent
Lille, UMR 8207UMETUnité Matériaux et advances in co-amorphous drug formulations. Adv. Drug Delivery Rev.
2016, 100, 116−125.
Transformations, Lille F-59000, France; orcid.org/0000- (14) Malfait, B.; Correia, N. T.; Mussi, A.; Paccou, L.; Guinet, Y.;
0003-4176-1925; Phone: +33 (0)3 20 43 69 90; Hédoux, A. Solid-state loading of organic molecular materials within
Email: emeline.dudognon@univ-lille.fr; Fax: +33 (0)3 20 mesoporous silica matrix: Application to Ibuprofen. Microporous
43 40 84 Mesoporous Mater. 2019, 277, 203−207.
(15) Descamps, M.; Willart, J. F.; Dudognon, E.; Caron, V.
Authors Transformation of Pharmaceutical Compounds upon Milling and
Florence Danède − Univ. Lille, CNRS, INRAE, Centrale Lille, Comilling: The Role of Tg. J. Pharm. Sci. 2007, 96, 1398−1407.
UMR 8207UMETUnité Matériaux et (16) Willart, J. F.; Descamps, M. Solid State Amorphization of
Transformations, Lille F-59000, France; orcid.org/0000- Pharmaceuticals. Mol. Pharm. 2008, 5, 905−920.
0001-5139-9187 (17) Burger, A.; Ramberger, R. On the polymorphism of
Mathieu Guerain − Univ. Lille, CNRS, INRAE, Centrale Lille, pharmaceuticals and other molecular crystals. I. Mikrochim. Acta
UMR 8207UMETUnité Matériaux et 1979, 72, 259−271. Theory of thermodynamic rules
Transformations, Lille F-59000, France; orcid.org/0000- (18) Burger, A.; Ramberger, R. On the polymorphism of
0002-7254-9014 pharmaceuticals and other molecular crystals. II. Mikrochim. Acta
1979, 72, 273−316. Applicability of thermodynamic rules
Complete contact information is available at: (19) Descamps, M.; Willart, J. F. Perspectives on the amorphisation/
https://pubs.acs.org/10.1021/acs.cgd.1c00997 milling relationship in pharmaceutical materials. Adv. Drug Delivery
Rev. 2016, 100, 51−66.
Notes (20) Latreche, M.; Willart, J.-F.; Guerain, M.; Hédoux, A.; Danède,
The authors declare no competing financial interest. F. Using Milling to Explore Physical States: The Amorphous and


Polymorphic Forms of Sulindac. J. Pharm. Sci. 2019, 108, 2635−2642.
(21) Zhang, G. G. Z.; Gu, C.; Zell, M. T.; Burkhardt, R. T.; Munson,
ACKNOWLEDGMENTS E. J.; Grant, D. J. W. Crystallization and Transitions of Sulfamerazine
This project has received funding from the Interreg 2 Seas Polymorphs. J. Pharm. Sci. 2002, 91, 1089−1100.
programme 2014−2020 co-funded by the European Regional (22) Shakhtshneider, T. P.; Boldyrev, V. V. Phase transformations in
Development Fund (FEDER) under the subsidy contract Sulfathiazole during mechanical activation. Drug Dev. Ind. Pharm.
2S01-059_IMODE. The FEDER, the French National Centre 1993, 19, 2055−2067.
for Scientific Research (CNRS), the Nord-Pas de Calais (23) Shakhtshneider, T. Phase transformations and stabilization of
region, and the French Ministry of National Education, Higher metastable states of molecular crystals under mechanical activation.
Solid State Ionics 1997, 101−103, 851−856.
Education and Research, are also acknowledged for funding of
(24) Lemmerer, A.; Báthori, N. B.; Esterhuysen, C.; Bourne, S. A.;
the X-ray diffractometer.


Caira, M. R. Concomitant Polymorphs of the Antihyperlipoproteine-
mic Bezafibrate. Cryst. Growth Des. 2009, 9, 2646−2655.
REFERENCES (25) Djinović, K.; Globokar, M.; Zupet, P. Structure of Bezafibrate
(1) Boldyreva, E. Mechanochemistry of inorganic and organic (2-{ p-[2-(p-Chlorobenzamide)ethyl]phenoxy}-2-methylpropanoic
systems: what is similar, what is different? Chem. Soc. Rev. 2013, 42, Acid). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1989, 45,
7719−7738. 772−775.
(2) Martin, G.; Bellon, P. Driven Alloys. Solid State Phys. 1996, 50, (26) Saal, W.; Ross, A.; Wyttenbach, N.; Alsenz, J.; Kuentz, M. A
189−331. Systematic Study of Molecular Interactions of Anionic Drugs with a
(3) Bakker, H.; Zhou, G. F.; Yang, H. Mechanically driven disorder Dimethylaminoethyl Methacrylate Copolymer Regarding solubility
and phase transformations in alloys. Prog. Mater. Sci. 1995, 39, 159− Enhancement. Mol. Pharm. 2017, 14, 1243−1250.
241. (27) Mahlin, D.; Bergström, C. A. S. Early drug development
(4) Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. predictions of glass-forming ability and physical stability of drugs. Eur.
Sci. 2001, 46, 1−184. J. Pharm. Sci. 2013, 49, 323−332.
(5) Ma, E. Alloys created between immiscible elements. Prog. Mater. (28) Wyttenbach, N.; Kirchmeyer, W.; Alsenz, J.; Kuentz, M.
Sci. 2005, 50, 413−509. Theoretical Considerations of the Prigogine-Defay Ratio with Regard

376 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

to the Glass-Forming Ability of Drugs from Undercooled Melts. Mol. (49) Williams, G.; Watts, D. C. Non-Symmetrical Dielectric
Pharm. 2016, 13, 241−250. Relaxation Behaviour Arising from a Simple Empirical Decay
(29) Lutterotti, L. Total pattern fitting for the combined size-strain- Function. Trans. Faraday Soc. 1970, 66, 80−85.
stress-texture determination in thin film diffraction. Nucl. Instrum. (50) Alvarez, F.; Alegra, A.; Colmenero, J. Relationship between the
Methods Phys. Res., Sect. B 2010, 268, 334−340. time-domain Kohlrausch-Williams-Watts and frequency-domain
(30) Kothari, K.; Ragoonanan, V.; Suryanarayanan, R. Dielectric Havriliak-Negami relaxation functions. Phys. Rev. B: Condens. Matter
Spectroscopy of Small Molecule PharmaceuticalsEffect of Sample Mater. Phys. 1991, 44, 7306−7312.
Configuration. J. Pharm. Sci. 2014, 103, 3190−3196. (51) Böhmer, R.; Ngai, K. L.; Angell, C. A.; Plazek, D. J.
(31) Havriliak, S.; Negami, S. A complex plane analysis of α- Nonexponential relaxations in strong and fragile glass formers. J.
dispersions in some polymer systems. J. Polym. Sci., Part C: Polym. Chem. Phys. 1993, 99, 4201−4209.
Symp. 1966, 14, 99−117. (52) Brás, A. R.; Noronha, J. P.; Antunes, A. M. M.; Cardoso, M. M.;
(32) Shen, C.-S.; Zhou, C.-R. Investigation of the thermal Schönhals, A.; Affouard, F.; Dionísio, M.; Correia, N. T. Molecular
decomposition kinetics of bezafibrate. J. Therm. Anal. Calorim. Motions in Amorphous Ibuprofen As Studied by Broadband
2016, 126, 959−967. Dielectric Spectroscopy. J. Phys. Chem. B 2008, 112, 11087−11099.
(33) Adrjanowicz, K.; Grzybowska, K.; Kaminski, K.; Hawelek, L.; (53) Johari, G. P.; Goldstein, M. Viscous Liquids and the Glass
Paluch, M.; Zakowiecki, D. Comprehensive studies on physical and Transition. II. Secondary Relaxations in Glasses of Rigid Molecules. J.
chemical stability in liquid and glassy states of telmisartan (TEL): Chem. Phys. 1970, 53, 2372−2388.
solubility advantages given by cryomilled and quenched material. (54) Johari, G. P.; Goldstein, M. Viscous Liquids and the Glass
Philos. Mag. 2011, 91, 1926−1948. Transition. III. Secondary Relaxations in Aliphatic Alcohols and Other
(34) Wojnarowska, Z.; Grzybowska, K.; Hawelek, L.; Dulski, M.; Nonrigid Molecules. J. Chem. Phys. 1971, 55, 4245−4252.
Wrzalik, R.; Gruszka, I.; Paluch, M.; Pienkowska, K.; Sawicki, W.; (55) Ngai, K. L. An extended coupling model description of the
Bujak, P.; Paluch, K. J.; Tajber, L.; Markowski, J. Molecular Dynamics, evolution of dynamics with time in supercooled liquids and ionic
Physical Stability and Solubility Advantage from Amorphous conductors. J. Phys. Condens. Matter 2003, 15, S1107−S1125.
Indapamide Drug. Mol. Pharm. 2013, 10, 3612−3627. (56) Ngai, K. L.; Rendell, R. W. Basic Physics of the Coupling
(35) Vogel, H. The law of the relation between the viscosity of Model: Direct Experimental Evidences. In Supercooled liquids,
liquids and the temperature. Phys. Z. 1921, 22, 645−646. advances and novel applications; Fourkas, J. T., Kivelson, D.,
(36) Fulcher, G. S. Analysis of Recent Measurements of the Mohanty, U., Nelson, K., Eds.; ACS Symposium Series; American
Viscosity of Glasses. J. Am. Ceram. Soc. 1925, 8, 339−355. Chemical Society: Washington DC, 1997; Vol. 676, pp 45−66.
(37) Tammann, G.; Hesse, W. Die Abhängigkeit der Viskosität von (57) Ngai, K. L.; Tsang, K. Y. Similarity of relaxation in supercooled
der Temperatur bei Unterkühlten Flüssigkeiten. Z. Anorg. Allg. Chem. liquids and interacting arrays of oscillators. Phys. Rev. E: Stat.,
Nonlinear, Soft Matter Phys. 1999, 60, 4511−4517.
1926, 156, 245−257.
(58) Lukenheimer, P.; Wehn, R.; Riegger, T.; Loidl, A. Excess wing
(38) Angell, C. A. Why the C1 = 16-17 in the WLF equation is
in the dielectric loss of glass formers: further evidence for a β-
physicaland the fragility of polymers. Polymer 1997, 38, 6261−
relaxation. J. Non-Cryst. Solids 2002, 307−310, 336−344.
6266.
(59) Dujardin, N.; Willart, J. F.; Dudognon, E.; Hédoux, A.; Guinet,
(39) Angell, C. A. Strong and Fragile Liquids. In Relaxations in
Y.; Paccou, L.; Chazallon, B.; Descamps, M. Solid state vitrification of
complex systems; Ngai, K. L., Wright, G. B., Eds.; Naval Research
crystalline α and β-D-glucose by mechanical milling. Solid State
Laboratory: Washington DC, 1985; pp 3−12.
Commun. 2008, 148, 78−82.
(40) Plazek, D. J.; Ngai, K. L. Correlation of polymer segmental
(60) Otsuka, M.; Matsumoto, T.; Kaneniwa, N. Effect of
chain dynamics with temperature-dependent time-scale shifts. Macro- environmental temperature on polymorphic solid-state transformation
molecules 1991, 24, 1222−1224. of indomethacin during grinding. Chem. Pharm. Bull. 1986, 34, 1784−
(41) Böhmer, R.; Angell, C. A. Correlations of the nonexponentiality
1793.
and state dependence of mechanical relaxations with bond (61) Schammé, B.; Couvrat, N.; Malpeli, P.; Dudognon, E.;
connectivity in Ge-As-Se supercooled liquids. Phys. Rev. B: Condens. Delbreilh, L.; Dupray, V.; Dargent, E.; Coquerel, G. Transformation
Matter Mater. Phys. 1992, 45, 10091−10094. of an active pharmaceutical ingredient upon high-energy milling: A
(42) Böhmer, R. Non-linearity and non-exponentiality of primary process-induced disorder in Biclotymol. Int. J. Pharm. 2016, 499, 67−
relaxations. J. Non-Cryst. Solids 1994, 172−174, 628−634. 73.
(43) Wang, L.-M.; Angell, C. A.; Richert, R. Fragility and (62) Dujardin, N.; Willart, J. F.; Dudognon, E.; Danède, F.;
thermodynamics in nonpolymeric glass-forming liquids. J. Chem. Descamps, M. Mechanism of Solid State Amorphisation of Glucose
Phys. 2006, 125, 074505. upon milling. J. Phys. Chem. B 2013, 117, 1437−1443.
(44) Rodrigues, A. C.; Viciosa, M. T.; Danède, F.; Affouard, F.; (63) Bhugra, C.; Shmeis, R.; Krill, S. L.; Pikal, M. J. Prediction of
Correia, N. T. Molecular Mobility of Amorphous S-Flurbiprofen: A Onset of Crystallization from Experimental Relaxation Times. II
dielectric Relaxation Spectroscopy Approach. Mol. Pharm. 2014, 11, Comparison between Predicted and Experimental Onset Times. J.
112−130. Pharm. Sci. 2008, 97, 455−472.
(45) Dudognon, E.; Bama, J.-A.; Affouard, F. Molecular Mobility of (64) Kothari, K.; Ragoonanan, V.; Suryanarayanan, R. Influence of
Terfenadine: Investigation by Dielectric Relaxation Spectroscopy and Molecular Mobility on the Physical Stability of Amorphous
Molecular Dynamics Simulation. Mol. Pharm. 2019, 16, 4711−4724. Pharmaceuticals in the Supercooled and Glassy States. Mol. Pharm.
(46) Atawa, B.; Correia, N. T.; Couvrat, N.; Affouard, F.; Coquerel, 2014, 11, 3048−3055.
G.; Dargent, E.; Saiter, A. Molecular mobility of amorphous N-acetyl- (65) Knapik, J.; Wojnarowska, Z.; Grzybowska, K.; Hawelek, L.;
α-methylbenzylamine and Debye relaxation evidenced by dielectric Sawicki, W.; Wlodarski, K.; Markowski, J.; Paluch, M. Physical
relaxation spectroscopy and molecular dynamics simulations. Phys. stability of the amorphous anti-cholesterol agent (Ezetimibe): the role
Chem. Chem. Phys. 2019, 21, 702−717. of molecular mobility. Mol. Pharm. 2014, 11, 4280−4290.
(47) Grzybowska, K.; Paluch, M.; Grzybowski, A.; Wojnarowska, Z.; (66) Okamoto, N.; Oguni, M. Discovery of crystal nucleation
Hawelek, L.; Kolodziejczyk, K.; Ngai, K. L. Molecular Dynamics and proceeding much below the glass transition temperature in a
Physical Stability of Amorphous Anti-Inflammatory Drug: Celecoxib. supercooled liquid. Solid State Commun. 1996, 99, 53−56.
J. Phys. Chem. B 2010, 114, 12792−12801. (67) Carpentier, L.; Decressain, R.; Desprez, S.; Descamps, M.
(48) Kohlrausch, R., II Theorie des elektrischen Rückstandes in der Dynamics of the amorphous and crystalline α-, γ- phases of
Leidener Flasche. Ann. Phys. 1854, 167, 179−214. Indomethacin. J. Phys. Chem. B 2006, 110, 457−464.

377 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378
Crystal Growth & Design pubs.acs.org/crystal Article

(68) Descamps, M.; Aumelas, A.; Desprez, S.; Willart, J. F. The


amorphous state of pharmaceuticals obtained or transformed by
milling: Sub-Tg features and rejuvenation. J. Non-Cryst. Solids 2015,
407, 72−80.
(69) Shakhtshneider, T. P.; Boldyrev, V. V. Mechanochemical
synthesis and mechanical activation of drugs. In Reactivity of Molecular
Solids; Boldyreva, E., Boldyrev, V., Eds.; John Wiley & Sons: New
York, 1999; pp 271−311.
(70) De Gusseme, A.; Neves, C.; Willart, J. F.; Rameau, A.;
Descamps, M. Ordering and disordering of molecular solids upon
mechanical milling: the case of fananserine. J. Pharm. Sci. 2008, 97,
5000−5012.
(71) Descamps, M.; Willart, J.-F.; Dudognon, E.; Lefort, R.; Desprez,
S.; Caron, V. Phase transformations induced by grinding: what is
revealed by molecular materials. Mater. Res. Soc. Symp. Proc. 2006,
979, 606.
(72) Willart, J.-F.; Lefebvre, J.; Danède, F.; Comini, S.; Looten, P.;
Descamps, M. Polymorphic transformation of the Γ-form of D-
Sorbitol upon milling: structural and nanostructural analysis. Solid
State Commun. 2005, 135, 519−524.
(73) Ostwald, W. Studien über die Bildung und Umwandlung fester
Körper. 1. Abhandlung: Ü bersättigung und Ü berkaltung. Z. Phys.
Chem. 1897, 22U, 289−330.
(74) Threlfall, T. Structural and Thermodynamic Explanations of
Ostwald’s Rule. Org. Process Res. Dev. 2003, 7, 1017−1027.
(75) Martinetto, P.; Bordet, P.; Descamps, M.; Dudognon, E.;
Pagnoux, W.; Willart, J.-F. Structural Transformations of D-Mannitol
Induced by in situ Milling Using Real Time Powder Synchrotron
Radiation Diffraction. Cryst. Growth Des. 2017, 17, 6111−6122.
(76) Mørup, S.; Jiang, J. Z.; Bødker, F.; Horsewell, A. Crystal growth
and the steady-state grain size during high-energy ball-milling.
Europhys. Lett. 2001, 56, 441−446.
(77) Sheth, A. R.; Bates, S.; Muller, F. X.; Grant, D. J. W.
Polymorphism in Piroxicam. Cryst. Growth Des. 2004, 4, 1091−1098.
(78) Bates, S.; Zografi, G.; Engers, D.; Morris, K.; Crowley, K.;
Newman, A. Analysis of Amorphous and Nanocrystalline Solids from
Their X-Ray Diffraction Patterns. Pharm. Res. 2006, 23, 2333−2349.
(79) Caira, M. R.; Mohamed, R. Positive Identification of Two
Orthorhombic Polymorphs of Sulfamerazine (C11H12N4O2S), their
Thermal Analyses and Structural Comparison. Acta Crystallogr., Sect.
B: Struct. Sci. 1992, 48, 492−498.

378 https://doi.org/10.1021/acs.cgd.1c00997
Cryst. Growth Des. 2022, 22, 363−378

You might also like