You are on page 1of 16

pubs.acs.

org/JPCB Article

Impact of Low Concentration of Strongly Hydrogen-Bonded Water


Molecules on the Dynamics of Amorphous Terfenadine: Insights
from Molecular Dynamics Simulations and Dielectric Relaxation
Spectroscopy
Jeanne-Annick Bama, Emeline Dudognon, and Frédéric Affouard*
Cite This: J. Phys. Chem. B 2021, 125, 11292−11307 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: The impact of low water concentration of strongly hydrogen-bonded water


Downloaded via UNIV DI PISA on November 28, 2023 at 20:47:39 (UTC).

molecules on the dynamical properties of amorphous terfenadine (TFD) is investigated


through complementary molecular dynamics (MD) simulations and dielectric relaxation
spectroscopy (DRS) experiments. In this article, we especially highlight the important role
played by some residual water molecules in the concentration of 1−2% (w/w) trapped in
the TFD glassy matrix, which are particularly difficult to remove experimentally without a
specific heating/drying process. From MD computations and analyses of the hydrogen
bonding (HB) interactions, different categories of water molecules are revealed and
particularly the presence of strongly HB water molecules. These latter localize themselves
in small pockets in empty spaces existing in between the TFD molecules due to the poor
packing of the glassy state and preferentially interact with the polar groups close to the
flexible central part of the TFD molecules. We present a simple model which rationalizes at
the molecular scale the effect of these strongly HB water molecules on dynamics and how
they give rise to a supplementary relaxation process (namely process S) which is detected for the first time in the glassy state of TFD
annealed at room temperature while this process is completely absent in a non-annealed glass. It also explains how this
supplementary relaxation is coupled with the intramolecular motion (namely process γ) of the very flexible central part of the TFD
molecule. The present findings help to understand more generally the microscopic origin of the secondary relaxations often detected
by DRS in the glassy states of molecular compounds for which the exact nature is still debated.

1. INTRODUCTION crystalline state due to strong inter- and intra-molecular HBs


Amorphous solids (glasses)1 composed of small and large can be highly hygroscopic in the amorphous solid state.9 The
organic molecules are ubiquitous in natural systems, and they amount of water content and water uptake are also highly
are commonly formed both intentionally and unintentionally dependent on the processes used to generate the amorphous
during technological processes. These materials have partic- solids. In practice, it can be produced by many different
ularly attracted strong interest in the development of routes:10 (i) the classical thermal quench from the liquid
pharmaceutical2 and food3 products for their remarkable state,11,12 (ii) the mechanical route (compression and milling)
advantages in terms of solubility, mechanical properties, or occurring entirely in the solid state,13−17 and (iii) the solvent
their capabilities for preserving biological agents and cells.4−7 removal processes (freeze-drying and spray-drying) requiring
In almost all cases, molecular amorphous solids contain a the dissolution of the compound in a solvent.18−20 It is well
certain amount of water and may also adsorb even more water recognized that water uptake can be particularly important for
upon storage since environmental moisture is generally amorphous solids obtained by milling since this technique
unavoidable.8 The reported values for the water content intrinsically increases the specific surface area of the
usually range from low concentrations of less than 1% (w/w) materials.13,14 Water content might also be important in
when specific heating and drying treatments have been some freeze-dried or spray-dried systems since they are first
employed, to high concentrations of about 30% (w/w) as
often found in protein systems mixed with sugars.8 These Received: July 8, 2021
different concentrations are actually very dependent on the Revised: September 15, 2021
hydrogen properties of the molecules that compose the Published: September 30, 2021
amorphous solids and particularly on the number of groups
that can donate or accept hydrogen bonds (HBs). It is
remarkable that the systems that are almost insoluble in the

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcb.1c06087


11292 J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Figure 1. Schematic representation of a molecule of TFD (α-[4-(1,1-dimethylethyl)phenyl]-4-(hydroxydiphenylmethyl)-1-piperidinebutanol).

dissolved in water before imposing subsequent freezing and can be strongly impacted by the presence of water because of
drying stages to systems.4−7 its high polarity.5,34−38 Traces of water that can be present
Even in small quantities, the presence of water has a major even in very dry products can be detected by DRS and not by
influence on the physicochemical properties of the amorphous some other techniques. Above Tg, in the liquid and
solids including their effective stability against recrystallization supercooled regime, dynamics as seen by DRS is often
or their chemical degradations.2,8,21,22 A significant piece of characterized by a single relaxation usually called the primary
work has been devoted to this topic aiming to understand the α-relaxation.32 This relaxation is associated with the global
exact mechanisms of action of water. Numerous studies have dynamics of the liquid or the supercooled liquid. It measures
particularly focused on the structural and dynamical properties the structural relaxation of the systems, and the characteristic
of water molecules themselves. Several investigations have time associated with the α-relaxation approximately equals 100
demonstrated the existence of different categories of water s at the glass-transition temperature. Addition of water makes
molecules, that is, free water, loosely bound water, or tightly the α-relaxation faster and thus lowers the glass-transition
bound water, depending on their positions and how they temperature.32 This effect is well known as plasticization and
interact with the molecules forming the amorphous solid.23,24 has been investigated in many works.22,26,35,39−42 It seems well
Evidence of the inhomogeneous organization of water and the accepted that this effect in polymers can be interpreted as due
formation of water clustering8 of various sizes depending on to an increased free volume and the consequent loosening of
the water content has also been obtained by various methods the interchain and/or intrasegmental motions.35 A similar
that include neutron and X-ray scattering, molecular dynamics interpretation has also been provided for systems made of low
(MD) simulation, and nuclear magnetic resonance. At very low molecular weight molecules. Oppositely, below Tg, dynamics as
water contents on various glassy carbohydrate matrices, based probed by DRS, is typically characterized by a number of faster
on the positron annihilation lifetime spectroscopy, it was also relaxations called secondary relaxations, which are far from
particularly proposed that water might act as a hole filler being so well understood.23,32,36,37 A deeper understanding on
between molecules.25 these local relaxations is important since they may have some
Major efforts have been particularly devoted to the impacts of different phenomena including sub-Tg crystalliza-
investigation of the impact of water on the dynamical tion processes or the stability of dried biological systems.5
properties of the amorphous solids for which water uptake These secondary relaxations originate from often unclear local
has been shown to generate greater molecular mobility and dynamics of molecules associated with some polar groups of
consequently a depression of the glass-transition temperature molecules (moiety and functional groups). The influence of
Tg.21,22,26 The possibility for water molecules to diffuse in the water on these processes has been studied, and it has been
glassy matrices has been demonstrated by several works from shown, by analogy with the primary α-relaxation, that water
NMR spectroscopy and neutron scattering experiments.27,64 may have some kind of plasticization or anti-plasticization
MD studies of water diffusion in a number of complex systems effect on them. Some other works have particularly suggested
have also been undertaken in which the mechanisms of that some secondary relaxations could directly originate for the
translational jumps for water diffusion were of particular orientation of water molecules themselves.34 The molecular
interest in this context. Based on computations, several origin behind these relaxations and the peculiar role of water
extremely complex processes were proposed to describe the are still under controversy.
“tortuous” paths, followed by the water molecules.28,29 The In the present study, the model pharmaceutical terfenadine
problem of diffusion of a small species in another frozen matrix (TFD) C32H41NO2 has been investigated. The literature on
is not restricted to molecular systems, and it is also seen in this system is relatively scarce. Calorimetric studies of TFD
metallic30 and ionic conductor glasses.31 have shown the existence of two polymorphic forms I and II
Dielectric relaxation spectroscopy (DRS) is a well-suited with melting temperatures TmI = 424.1 K and TmII = 416.5 K,
technique to follow rotational dynamics since it probes both respectively, but others forms are suspected.43,44 Details about
global and localized dipolar reorientation motions over a wide the crystalline structure of the two forms are lacking. The
accessible frequency range (mHz to GHz).32,33 DRS signals molecular mobility of TFD has been recently investigated by
11293 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Dudognon et al. by means of DRS experiments and MD short circuit between the two electrodes. A heating to 453 K/
simulations.45 TFD was shown to be a good glass former (Tg = cooling cycle at 5 K min−1 was applied while recording the
330 K) with fragile nature (m = 112),45 which also has the complex permittivity at some frequencies. This ensures that the
advantage to be easily vitrified by different amorphization melting of the sample is achieved and verifies that no re-
routes including melt-quenching or milling.46,47 However, it crystallization occurs on cooling. After cooling to room
was evidenced that these two routes of preparation weakly temperature, the electrodes were removed from the dielectric
influence its dynamics.45 TFD is a hydrophobic molecule with cell and immediately stored in the conditions previously
solubility in water equal to 10.2 μg mL−1,48 and consequently, described.
the interactions between TFD and water are poorly 2.3. Dielectric Relaxation Spectroscopy. Dielectric
documented. Interestingly, based on scanning electron measurements of amorphous annealed TFD were carried out
microscopy, a peculiar interaction of amorphous TFD with using a Novocontrol Technologies GmbH Alpha Analyzer in
water was shown to generate a self-limiting crystallization the frequency range from 10−1 to 106 Hz at temperatures
process.49 This work particularly demonstrated that when ranging from 173 to 453 K. The temperature was controlled
amorphous TFD is exposed to water vapor, a small amount of through the Quatro Cryosystem (supplied by Novocontrol
the surface crystallizes, which inhibits the whole crystallization Technologies) with a stability condition of 0.5 K. The complex
process. It was suggested that water molecules may localize in permittivity ε*(ω) = ε′(ω) − iε″(ω) was recorded while
different pores or cavities at the surface of the amorphous samples were submitted to a sinusoidal electrical field of 1 V
material and trigger the process of crystallization of TFD in magnitude.
these pores. However, these newly crystallized regions at the Dielectric relaxation data were analyzed using the Havriliak−
surface also seal the access of further water molecules, which Negami (HN) model function50 as expressed by
stops the continuation of the crystallization process.49
In this article, based on complementary DRS experiments Δε
ε*(ω) = ε∞ +
and MD simulations, we performed a thorough investigation of [1 + (iωτHN)αHN ]βHN (1)
the structural organization of water in TFD amorphous solids
and how it impacts dynamics and in particular secondary where Δε = εs − ε∞, defined as the difference between the
relaxations. dielectric constant measured at low and high frequencies, gives
the magnitude of the relaxation process. τHN is the Havriliak−
2. EXPERIMENTAL SECTION Negami characteristic relaxation time of the process, and αHN
and βHN are the shape parameters taking into account the
2.1. Materials. The racemic TFD (C32H41NO2) was
broadening and asymmetry (0 < αHN ≤ 1, 0 < βHN ≤ 1). In the
furnished by Sigma-Aldrich (purity ≥ 97.5%, Lt
case of the existence of dc-conductivity, we used an additional
MKBX6318V) and was used as received. TFD is a quite
term (−i σdc/(ε0 ω)), where σdc is the dc-conductivity and ε0 is
large molecule (Mw = 471.67 g mol−1) compared to the
the permittivity of the vacuum. From the HN fitting parameter
majority of active pharmaceutical ingredients (see Figure 1). It
values, the model-independent relaxation time τmax = 1/2πFmax
contains three polar sites (two hydroxyl groups and one

ÄÅ É Ä É
was calculated by

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


nitrogen) numbered 1, 2, and 3 as presented in Figure 1. At

ÅÅ jj zzÑÑ ÅÅ jj HN HN zzÑÑ
= τHNÅÅsinjjj zzÑÑ ÅÅsinjj zÑ
ÅÅ j 2 + 2β zzÑÑÑ
room temperature, TFD is in the form of a white crystalline

ÅÅ 2 + 2βHN z{ÑÑÑ
ÅÅÇ k ÅÅÇ k HN {Ñ
powder. TFD melting point is found at Tm = 425 K.49 It can be
ÑÖ ÑÖ
HN
τmax
vitrified by quenching from the melt, and the glass-transition
temperature reported in the literature is Tg = 327−332 K with
(2)
a heating rate of 10 K min−1.43,45
2.2. Preparation of Amorphous TFD. The amorphous 2.4. MD Simulations. In the present work, MD
TFD was prepared by quenching from the melt using a simulations were performed on pure TFD (see Figure 1)
differential scanning calorimetry (DSC) Q200 from TA and TFD/water mixtures using the all-atom OPLS and the
Instrument equipped with a Refrigerated Cooling System. TIP3P water force field.51,52 Four systems have been
About 5 mg of crystalline TFD was heated in an aluminum pan investigated: pure TFD and three TFD/water mixtures
until 453 K with a heating rate of 5 K min−1. This latter corresponding roughly to 1, 10, and 99% w/w water
temperature was chosen since it is about 30 K above the concentrations. To prepare the pure TFD system, a pseudo-
melting point and thus allows us to obtain a melt with no crystal made of 64 TFD molecules (4 × 4 × 4 simple cubic
crystalline germs. A cooling of 20 K min−1 to room cells) was constructed. Then, this system was rapidly melted at
temperature vitrified the melt. The obtained glass was a high temperature T = 500 K in order to generate an initial
immediately stored either at room temperature (RH ≈ 36%) disordered configuration. For TFD/water mixtures at 1 and
for 2−47 days or in a desiccator containing saturated sodium 10% water concentrations, the same procedure was used and
chloride solution (RH ≈ 76% at RT) for 3 weeks for further either 16 or 192 water molecules were just included in the
analysis. The annealing times were arbitrarily chosen. construction of the pseudo-crystal. For the TFD/water mixture
It should be noted that the annealed glass of TFD prepared at the 99% w/w water concentration, a single TFD molecule
by melt-quenching and stored either at room temperature or in was inserted into a large TIP3P water box pre-equilibrated at
a controlled humidity environment (RH ≈ 76%) was found to 300 K. Then, water molecules overlapping with the TFD
be X-ray amorphous at room temperature. Assessment of molecule were removed to generate a system made of a single
chemical degradation by 1H NMR analysis (data not shown) TFD molecule and 1338 water molecules. This latter system
revealed that the chemical integrity of TFD, is maintained. aims at investigating TFD in a fully hydrated situation.
For dielectric measurement, about 60 mg of crystalline TFD Computations have been performed at different temperatures
was placed between two gold-coated electrodes (diameter 10 by rapidly cooling the pure TFD and the 1 and 10% TFD/
or 20 mm) with quartz spacers (50 μm thickness) to avoid water mixtures from 500 to 300 K. The simulation times range
11294 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

from 10 ns at 500 K to 50 ns at 300 K. This procedure has


been repeated five times to generate five independent
trajectories in order to improve the statistics and estimate
the uncertainties. It should be noted that MD simulations have
been thus performed either on equilibrated or non-equilibrated
states. At the highest temperatures, 450 and 500 K, the pure
TFD and the 1 and 10% TFD/water mixtures systems can be
considered as fully equilibrated over the duration of the MD
simulations [see some examples of time-dependent correlation
function in Figure S1 (Supporting Information)]. However, at
the lowest temperatures, only the 99% TFD/water mixtures
can be considered as fully equilibrated. The three other
systems cannot be fully equilibrated and they thus enter the
glassy state.
All calculations have been carried out using the DL_POLY
package.53 Cubic periodic boundary conditions have been
Figure 2. First recordings of tan δ = ε″/ε′ vs F from 193 to 283 K of a
applied in all simulations. To integrate the equations of
glass of TFD annealed 17 days at room temperature (full symbols)
motion, we used the Verlet algorithm, with a time step of 2 fs, where the supplementary S-relaxation appears. For the sake of
and bond lengths were constrained by means of the SHAKE comparison, the spectra of a non-annealed glass of TFD (open
algorithm. A Lennard-Jones potential has been employed to symbols) recorded at four representative temperatures (193, 213, 233,
represent van der Waals interactions. For the electrostatic and 253 K) are added. The solid lines represent the fits of the data
interactions, a pairwise damped shifted method developed by (two and one Cole−Cole functions, respectively, for annealed and
Wolf was used. The same cut-off radius of 10 Å has been used non-annealed samples).
for both van der Waals and Coulombic interactions. Thermal-
ization was carried out in the isobaric−isothermal NPT to the high-frequency shoulder of the S-relaxation found in the
ensemble at a pressure P of 1.0 Atm. The Nosé−Hoover annealed glass. This is clearly visible in Figure 3, where, for the
thermostat and barostat relaxation times have been chosen as
0.2 and 2.0 ps, respectively. The stabilized volume of the
simulation box during the NPT simulation was considered to
compute the averaged volume of the system and used to
perform the subsequent production simulation in the NVT
ensemble (constant number of particles, volume, and temper-
ature).
Dipolar properties have been obtained from the time
dependence and statistics of the individual molecular dipole
moments computed in the different systems at the different
investigated temperatures. The time-dependent dipole mo-
ment of a given molecule at time t is particularly calculated
N
from the classical expression: μ⃗ (t ) = ∑α =a 1 qα rα⃗ (t ), where qα
and r⃗α(t) are, respectively, the fixed charge localized on atom α
and its position at time t and Na is the number of atoms in the
considered molecule (Na =3 and 64 for the water and TFD
molecule, respectively). Fixed charges are directly taken from Figure 3. tan δ = ε″/ε′ recorded at 228 K for the glass annealed 17
the OPLS and TIP3P force field.51,52 days at room temperature (black symbols). The green and red lines
represent the Cole−Cole functions of the S- and γ- processes,
3. RESULTS respectively, and the black line corresponds to the sum. For the sake
of comparison, the recording at the same temperature of the non-
3.1. Dynamic Behavior. The dielectric spectra of annealed glass of TFD (blue symbols) is added, as well as the fitting
amorphous TFD annealed 17 days at room temperature have by a Cole−Cole function (blue line).
been recorded in the glassy state between 10−1 and 106 Hz
from 173 to 283 K (increment of 5 K). Figure 2 represents tan sake of comparison, the tan δ dielectric spectra recorded at 228
δ as a function of frequency (full symbols). In the temperature K for TFD non-annealed (in blue) and annealed 17 days at
range from 193 to 283 K, we can observe one relaxation room temperature (in black) are reported (the spectra showing
crossing the frequency window. It is very intense and the evolution at 228 K of the imaginary part of the permittivity
characterized by a shoulder on the high-frequency side that and of the modulus can also be seen in Supporting Information
can be seen especially at the lowest temperatures. For S2). Therefore, in order to fit the data of annealed TFD, two
comparison, some dielectric spectra of a freshly obtained Cole−Cole functions (βHN = 1)54 were used (also reported in
glass recorded in the same temperature range are also reported Figure 3). It should be noted that, since the S-relaxation is
in Figure 2 (open symbols).45 It can be seen that this about 3 times more intense than the γ-one and the γ mode is
relaxation mode does not appear on the dielectric spectra of quite broad, the two modes are difficult to deconvolute. The
non-annealed glass. It will then be called S-relaxation (for magnitude of the fitting functions can thus be biased, which
supplementary relaxation). Besides, observed in more detail, could explain the slight shift toward the low frequency of the
the γ-relaxation of the non-annealed glass seems to correspond fitted γ mode observed for the annealed glass. However, based
11295 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

on raw data which are reliable, one can note that the added in Figure 4 (dashed lines). It can be seen that, contrary
magnitude of the shoulder on the spectra of annealed TFD to the freshly quenched sample, the annealed TFD
systematically appears slightly lower than the magnitude of the recrystallizes on heating, but the most striking feature is that
γ-relaxation of non-annealed TFD. the α-relaxation of TFD previously annealed at room
The fit of these data permitted us to extract the characteristic temperature (solid lines) is enhanced compared to the one
relaxation time τmax of the γ and S modes appearing for of the freshly melted-quenched sample (dashed lines). There is
annealed glass. The Arrhenius diagram can be seen in the a shift of about 1 decade toward high frequencies. This
Supporting Information (Figure S3). indicates that the relaxation time of the α-process which refers
The dependence of S-relaxation time versus the reciprocal to the global motions of the system decreases when this latter
temperature obeys an Arrhenius law is annealed a few days at room temperature.
Ea Upon further heating, the recrystallized sample melts at 423
log τmax = log τ∞ + K. The sample of TFD obtained after the quench of this liquid
Ln 10 R T (3)
was analyzed above and below Tg and compared to the freshly
where τ∞ would be the relaxation time at infinite temperatures, quenched one. Figure 5 is a plot of tan δ as a function of
Ea is the activation energy, and R is the perfect gas constant.
The best fit was obtained for log τ∞ = −14.48 ± 0.06 close to
the expected value (log τ∞ = −14) and an activation energy of
Ea = 50 ± 1 kJ mol−1. The shape parameter αHN slightly
increases with temperature from about 0.5 to 0.6 (βHN = 1).
This high value reflects the narrowness of this mode.
Regarding the γ-relaxation, the log τmax values extracted from
the Cole−Cole fit are close to the evolution of the relaxation
time obtained for the non-annealed glass by Dudognon and et
al.45 (log τ∞ = −14.1 ± 0.1 and Ea = 40 ± 1 kJ mol−1) (cf.
Figure S3). However, the dispersion of points prevented us to
precisely determine the behavior law of the relaxation time.
This dispersion comes from the difficulty to deconvolute the S
and γ modes, as previously mentioned. In any case, it seems
that the γ-relaxation time is not or only slightly influenced by
the appearance of the S-mode.
The results obtained at higher temperatures upon isotherms Figure 5. Recordings at 343 K of tan δ = ε″/ε′ vs F showing the α-
ranging from 343 to 368 K of TFD previously annealed 45 relaxation of TFD freshly melted (red line), after annealing of the
days at room temperature are reported in Figure 4 (solid glass 45 days at RT (green line), and after melting of that sample
recrystallized on previous heating (black line). The inset shows a
comparison in the glassy state at 253 K of the same samples.

frequency obtained upon isotherms at 343 and 253 K (cf. inset


Figure 5, black line). It can clearly be seen that below Tg, in
the glassy state, the S-relaxation does not appear anymore and
only the γ-process remains. The obtained spectrum is nearly
identical to the spectrum of the non-annealed glass (red line)
(cf. inset Figure 5). Additionally, the dielectric spectrum at 343
K of the obtained liquid (black line) is superimposed to that of
the freshly melt-quenched one (red line). Hence, it proves that
the phenomenon at the origin of the S-process and the shift of
the main relaxation is reversible: it is not due to a chemical
degradation of TFD glass that could have occurred during
annealing.
In order to understand the difference in dynamics that exists
between the annealed and non-annealed glasses of TFD,
Figure 4. Recordings between 343 and 368 K of tan δ = ε″/ε′ vs F
showing the α-relaxation of TFD annealed 45 days at room
progressive heatings of glass annealed 17 days at room
temperature (solid lines). From 363 K, the signal collapses due to temperature were carried out. In the temperature range of
crystallization. The increase of tan δ values at low frequencies is due 193−353 K, we performed cycles of successive isotherms
to σdc. For comparison, the recording at 348, 358, and 368 K for non- starting from 193 K with an increment of 5 K till an end
annealed TFD is added (dashed lines). temperature of the “n + 1” cycle 10 K higher than that of the
“n” cycle. At the end temperature of each cycle, an annealing of
40 min was then carried out before recording the next one.
lines). We can observe the α-relaxation which moves toward The dielectric spectra acquired at 218 K during these
higher frequencies with increasing temperature. Its magnitude successive measurements are presented in Figure 6. One can
vanishes above 363 K. It is due to the re-crystallization of the observe that, during these cycles of heatings, the S-process
sample. At low frequencies, a tail of dc-conductivity also magnitude decreases and it shifts toward the lower frequencies.
appears. For the sake of comparison, the spectra recorded in In the same time, the γ-relaxation arises and its intensity
the same temperature range for non-annealed TFD are also increases upon heating. This result indicates that the dynamics
11296 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

glasses containing 1.23% water (blue line) and 1.73% water


(red line) and for a non-annealed glass (black line) (free of
water). One can observe that the intensity of the S-process
increases strongly with increasing water content. Therefore, we
can definitely relate the S-process with the presence of water,
absorbed by the sample during annealing.
Moreover, as can be seen in Figure 7B, the α-relaxation
moves toward higher frequencies with increasing relative
humidity. This enhancement of mobility of the main relaxation
above Tg can indeed be assigned to a plasticization effect
because of the well-recognized ability of water to act as a
plasticizer in lowering the glass transition and increasing the
molecular mobility.7,9,39 This remarkable increase of the
mobility at 343 K is about 1 decade for the sample annealed
at 36% RH and 2 decades for the one annealed at 76% RH,
Figure 6. Recordings of tan δ = ε″/ε′ vs F, at 218 K, of a glass of TFD which is a clear indication that the TFD matrix mobility is
stored 17 days at room temperature during cycles of successive changing due to the incorporation of water molecules. In a
isotherms starting from 193 K with an increment of 5 K until an end detailed study of supercooled amorphous IMC, Andronis and
temperature of the “n + 1” cycle 10 K higher than the “n” cycle. An Zografi26 observed a similar behavior; that is, the increasing
annealing time of 40 min was carried out at the end temperature of water content significantly lowers the relaxation time of
each cycle before recording the next one. amorphous IMC. Indeed, they provided the evidence that for
every 1% water content in amorphous IMC, the Tg is lowered
of these two modes are coupled. The decrease of the S-process by ≈ 10 K.9 Besides, a simple rescaling of ε″/εmax″ vs F/Fmax
magnitude upon heatings added to the fact that the relaxation allows us to evaluate the effect of water on the shape
time of the α-process of TFD decreases when the glass is parameters of the α-relaxation peak. The spectra of TFD
annealed a few days at room temperature suggests that the previously annealed with different RH conditions superimpose
absorbed water could be at the origin of these two phenomena. with the spectrum of the freshly melted-quenched glass (cf.
Therefore, these progressive heatings could be assimilated to Figure S5). In other words, the shape parameters of α-
drying of the sample, which would result in a loss of absorbed relaxation are water independent. This suggests that the
water. In other words, the annealed glass of TFD could be distribution of the dielectric relaxation time remains the same
dehydrated during these progressive heatings, with water being as the water uptake increases. Hence, water does not generate
responsible of the observed changes of dynamics. more heterogeneity in the main dynamics of TFD.
To investigate a possible effect of water, we have annealed a As a final remark of this part, it is striking that traces of the
glass of TFD 21 days in a desiccator containing a saturated S-relaxation which have been associated with the presence of
sodium chloride solution (RH ≈ 76% at RT) and compared it absorbed water are still detected after heating and annealing at
to a glass annealed 45 days at ambient conditions (RH = 36%) 353 K (see Figure 6) and that the plasticization of the main
(cf. Table 1). Thermogravimetric analysis (TGA) shows that dynamic is still observed at 368 K (see Figure 4). Moreover, a
close inspection of the inset of Figure 5 shows that the
Table 1. Annealing Conditions at Ambient Temperature of spectrum of the glass obtained from the melting of TFD
TFD Glasses and Water Content Determined by TGA previously annealed 21 days at RH = 76% and recrystallized
annealing relative humidity water content upon subsequent heating (black line) does not perfectly
sample time (%) (%) superimpose with the non-annealed glass (red line). Indeed, a
glass of TFD 1 21 days 76 1.73 small discrepancy can be observed around 102 Hz that could
glass of TFD 2 45 days 36 1.26 be due to the presence of a “residual” S-process. To get more
insights concerning the origin of this discrepancy, recordings in
the glassy state were carried out after successive increasing
the weight loss of these two glasses occurs in two steps (cf. annealing times at 453 K (Tm +30 K) of TFD previously stored
Figure S4). First, there is a continuous loss from room 21 days at RH ≈ 76%. The spectra recorded at 253 K are
temperature to at least 400 K. Then, there is a second loss reported in Figure 8. In this figure, one can notice that, after a
starting from around 475 K, which is due to the thermal simple heating until 453 K and cooling, the spectrum of the
degradation of TFD. The first loss can then be assigned to loss obtained glass shows a γ-relaxation with a shoulder appearing
of absorbed water. The measurements of the corresponding clearly around 102 Hz. It indicates the existence of a residual S-
water content for both samples were done at 448 K in order to process. However, one can notice that the intensity of this
have results comparable to those of DRS, where the effective residual S-process decreases for increasing annealing time at
heating rate is much slower (about 0.9 K min−1) and the 453 K. Also, the recrystallized sample has to be heated up to
material quantity is 12 times greater. The sample annealed 21 the melting temperature Tm = 425 K in order to release the
days at RH ≈ 76% shows a water content of 1.73%, so about water molecules that were trapped as inclusion in the
0.5 point more than the glass annealed 45 days at RH ≈ 36%. crystalline structure. It is quite striking that the sample should
Thus, despite the hydrophobic characteristic of crystalline be heated up to 453 K (Tm +30 K) and annealed at this
TFD,55 the TGA experiment shows that the glass can absorb a temperature for about 1 h to find the signature of the non-
significant amount of water. The same behavior was observed annealed glass (back line). This would suggest that some water
in the case of indomethacin (IMC).26 Figure 7A shows the molecules absorbed during the storage of the glass at room
dielectric spectra obtained upon isotherm at 253 K for the temperature really remain embedded in the sample (till 453 K)
11297 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Figure 7. (A) Recordings at 253 K of tan δ = ε″/ε′ vs F showing the γ- and S-relaxation of TFD glasses containing 0, 1.26, and 1.73% of water. (B)
Recordings of ε″ vs F at 343 K of the same samples. The lines through the data points correspond to the HN fit of the α-relaxation after subtraction
of the σdc contribution.

two −OH hydroxyl groups and the nitrogen atom localized


close to the central part of the molecule (see Figure 9C). In
order to identify HBs in the different studied system, a
geometric criterion has been chosen. Two molecules are
considered to be H-bonded if (i) the nitrogen-oxygen or the
oxygen−oxygen distance is less than 3.4 Å and (ii) the (N−
H−O) or (O−H−O) angle is larger than 150°. This criterion
is in some way arbitrary due to the lack of information on the
electron density, but it has been successfully used in many MD
simulations and allows us to take into account a wide variety of
HBs including more deformed and weaker HBs in
statistics.56−58 Some snapshots of instantaneous configurations
of the pure TFD system and the TFD/water system at 1% w/w
water concentration at T = 300 K have been generated and are
displayed in Figure 9A,B, respectively. They illustrate the very
Figure 8. Recordings at 253 K of tan δ = ε″/ε′ vs F showing the γ- rich diversity of HB associations which can exist between the
relaxation in the case of the non-annealed TFD (black line) and after TFD molecules and TFD and water molecules in a dried
different annealing times (0, 5, 20, and 50 min, respectively) at 453 K sample in which some residual water is still present.
for the TFD sample previously stored at RT 21 days in a desiccator 3.2.1. Statistics of Intra-molecular and Inter-molecular
over a saturated sodium chloride solution. HBs. Statistics on the number of intra-molecular and inter-
molecular HBs formed on average in the four investigated
and are possibly highly hydrogen-bonded to TFD molecules. systems at T = 300 K are reported in Table 2. For pure TFD
In order to clarify this feature at the molecular scale, molecular and TFD/water mixtures at low water concentrations (1 and
modeling works have been performed that will be described in 10% w/w), a few intra-molecular HBs are clearly detected (see
the next section. Figure 9C). Since one TFD molecule can only form one intra-
3.2. MD Simulations. MD simulations have been molecular HB, it corresponds to a fraction of about 12.5% of
performed in order to investigate in more detail structural TFD molecules in the pure TFD system. This 12.5% fraction is
and dynamical properties of TFD/water mixtures (see simply calculated by dividing the number of intra-molecular
simulation details in the Experimental Section) and in HBs, 8, by the total number of TFD in the system, 64. This
particular the HB interactions between TFD and water fraction tends to slightly decrease to about 9.8% upon
molecules. Our aim was to shed some light on the behavior increasing the water concentration, but this trend is not
of TFD in the presence of water, which seems to be the completely clear due to the large uncertainties associated with
microscopic origin of the peculiar S dynamical process the number of HBs. Indeed, the pure TFD and the TFD/water
detected from DRS experiments. A thorough analysis of HB mixtures at the studied low water concentrations (1 and 10%
associations that can be formed in the pure TFD system and in w/w) at T = 300 K are actually in a glassy state obtained from
the different TFD/water mixtures has been thus performed. a rapid quench from high temperatures. Although MD
The TFD molecule possesses two −OH hydroxyl groups and simulations were repeated five times at each concentration,
one nitrogen atom from which it may form HBs with other the large uncertainties associated with the number of HBs
TFD molecules and/or with water molecules (see Figure 1). actually reflect the possibility for the different systems to be
The two −OH hydroxyl groups can participate to an HB either frozen in many different configurations with many different
as a donor or as an acceptor while the nitrogen atom can only kinds of HB associations between molecules. It should be thus
be an acceptor. It should be noted that TFD can also certainly necessary to perform a much larger number of
potentially form an intra-molecular HB involving one of the independent MD simulations for the different systems in order
11298 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Figure 9. (A) Snapshots of instantaneous configurations as obtained from MD simulations at T = 300 K in the pure TFD system. Only
configurations involving the hydroxyl group and the nitrogen atom localized close to the central part of the TFD molecule are shown. (a) Simple
HB association as a dimer where two TFD molecules (#TFD1 and #TFD2) are HB-bonded by a single HB involving two hydroxyl groups. (b):
More complex dimer for which two TFD molecules (#TFD1 and #TFD2) are HB-bonded by two HBs involving two hydroxyl groups and one
nitrogen atom. Because of the formation of two HBs, this association is clearly more stable than the previous one. (c): Simple trimer where three
TFD molecules (#TFD1, #TFD2, and #TFD3) are HB-bonded through single HBs between them. A short chain-like structure is thus formed. (d)
Another example of a trimer in which the nitrogen atom is involved. HBs are indicated by dotted lines. (B) Snapshots of instantaneous
configurations as obtained from MD simulations at T = 300 K in the TFD/water mixture at 1% w/w concentration. A single water molecule is
shown to be capable to form a HB with two (a), three (b), or four (c) TFD molecules. Figures (b,c) are actually zoom-centered on the water
molecule. HBs are indicated by dotted lines. (C) TFD molecule for which the possible intra-molecular HB is shown (dashed line).

Table 2. Number of Intra-molecular HBs Formed by TFD and Number of Inter-molecular HBs Formed between TFD
Molecules and TFD and Water Molecules at T = 300 K for the Four Investigated Systemsa
inter-molecular HB TFD−TFD inter-molecular HB TFD−water
intra- HB TFD
O2−H−N O1 O2 N total O1 O2 N total
pure TFD (0% w/w) 8.0 ± 1.0 5.7 ± 0.5 14.5 ± 1.0 2.3 ± 0.2 22.5 ± 1.6 - - - -
TFD/Water (1% w/w) 7.9 ± 1.0 5.0 ± 1.2 13.5 ± 1.5 2.1 ± 1.0 20.6 ± 1.7 5.1 ± 0.5 10.1 ± 0.5 4.0 ± 0.4 19.2 ± 1.1
TFD/Water (10% w/w) 6.3 ± 0.5 3.5 ± 0.5 10.5 ± 0.6 0.8 ± 0.2 14.8 ± 1.2 23.7 ± 0.7 41.4 ± 2.2 21.9 ± 0.8 87.0 ± 3.6
TFD/Water (99% w/w) 0 - - - - 71.7 ± 7.0 110.0 ± 5.1 53.8 ± 1.9 235.5 ± 14.1
a
For inter-molecular HBs, the number of HBs is also specifically given for the two hydroxyl groups (O1 and O2) and the nitrogen atom (N) (see
Figure 1) of the TFD molecule. For the TFD/water mixtures at 99 % w/w water concentration, the number of HBs have been multiplied by the
number of TFD molecules present in the other investigated systems, i.e., 64, to allow a direct comparison (see details in MD simulations).
Uncertainties have been calculated based on five independent MD runs and/or using block averaging.

to obtain a full description of the HB associations that could TFD/water mixture. Table 2 particularly shows that 64 TFD
exist in the glassy state of the different TFD/water systems. It molecules form on average 71.7, 110.0, and 53.8 HBs with
is clear that the formation of an intra-molecular HB for TFD water molecules through the −OH hydroxyl group (O1), the
tends to reduce its inter-molecular HB propensity. For −OH hydroxyl group (O2), and the nitrogen atom (N)
example, it might have strong consequences on the capability localized close to the central part of the molecule, respectively
of TFD to form large HB clusters as already demonstrated for (see Figure 1). It roughly corresponds to about 1.1, 1.7, and
disaccharide/water mixtures from MD simulations.56 Interest- 0.9 HBs per TFD molecule. From these results, the HB
ingly, no intra-molecular HBs are found for the fully hydrated propensity of the central part of the TFD molecule approaches
TFD system at 99% w/w water concentration. It shows that an optimal situation assuming theoretically that two water
the presence of a large amount of water molecules completely molecules can HB-bond to the −OH hydroxyl group (O2) and
inhibits the formation of this intra-molecular HB. It could be one water molecule to the nitrogen atom. Oppositely, the
suggested that water molecules, because of their quite small terminal −OH hydroxyl group (O1), localized in between
size and high mobility, can very easily access to the central three large rings, seems much less accessible and only about
−OH hydroxyl group (O2 in Figure 1) and the nitrogen atom one water molecule can form a HB with this group. Upon
and form HB bonds with them. This suggestion is confirmed decreasing the water concentration, this trend is still observed.
by a closer examination on the details of the HBs formed with Table 2 shows that the central −OH hydroxyl group (O2)
the different HB groups of the TFD molecules in the 99% w/w systematically forms from 2−3 times more HBs than the
11299 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Figure 10. (A) Fraction of HB aggregates formed by TFD molecules at 300 K calculated from MD simulations for pure TFD and TFD/water
mixtures at 1 and 10% w/w concentrations. n represents the number of TFD molecules in the HB aggregate. For example, n = 1 corresponds to
isolated molecules, n = 2 to dimers, n = 3 to trimers, and so forth. Dotted lines are a guide for the eye. (B) Fraction of HB aggregates formed by
water molecules at 300 K calculated from MD simulations for TFD/water mixtures at 1% (inset) and 10% w/w concentrations. n represents the
number of water molecules in the HB aggregate. Dotted lines are a guide for the eye.

terminal −OH hydroxyl group (O1) either considering the Interestingly, the obtained results suggest that the small
inter-molecular HBs between the TFD molecules or the inter- amount of water molecules that mainly localized close to the
molecular HBs between the TFD and water molecules. It central part of the TFD molecules through HB may potentially
should be noted that the same trends are also observed at have a stronger impact on its peculiar dynamics, which was
higher temperatures (data not shown). Overall, at all associated with the γ-mode as suggested in ref 45. More
investigated concentrations and temperatures, about 75% of discussions on the link between MD simulations results and
all inter-molecular HBs originate from the HB groups localized DRS experiments will be provided in the following.
close to the central part of the TFD molecule, that is, the −OH 3.2.2. TFD and Water HB Clusters. Based on the HB
hydroxyl group (O2) and the nitrogen atom. These results thus geometric criterion used in the present study (see above), the
clearly provide evidence of the importance of these HB groups distribution of clusters composed of n TFD molecules or n
through which TFD preferentially interacts. water molecules have been calculated from MD simulations at
For pure TFD, the total number of HBs formed is about T = 300 K. In the following, a cluster of size n is defined as a
22.5. This number seems relatively weak compared to the set of n molecules connected by at least one HB between them.
potential capability of the TFD molecule to form HBs. Indeed, The TFD cluster distribution is shown in Figure 10A,B.
theoretically, one could expect that five HBs could be formed Figure 10A shows that a large variety of HB aggregates
for one TFD molecule: two for each −OH hydroxyl group and composed of TFD molecules are theoretically possible (see
one for the nitrogen atom. In the present simulations, the pure Figure 9A,B for snapshots of HB associations). The fraction of
TFD system is composed of 64 molecules. It thus indicates HB aggregates mostly ranges from isolated molecule (n = 1) to
that each TFD molecule does form less than one HB on large aggregates made of n = 8 molecules. Additional even
average and potentially many TFD does not form HBs at all. larger (n > 8) associated structures could also be transiently
This result will be confirmed in the following by analyzing the observed over a very short time, but their fraction remains
TFD HB cluster distribution. Upon increasing water really very small at all investigated temperatures. It should be
concentration, Table 2 reveals that the number of HBs formed noted that the extension of the TFD HB clusters does not
between TFD molecules decreases, while at the same time, the seem very developed compared to similar studies performed by
number of HBs formed between TFD and water molecules some of the authors on other small organic molecules for
significantly increases. This result is obviously expected which clusters involving more than n = 10 molecules can be
because of the change in the concentration of water in the found.57,58 Figure 10A particularly shows that the maximum of
mixtures. However, overall, the total number of HBs increases the cluster size distribution corresponds to n = 1, that is,
upon increasing water concentration because of the higher isolated molecules. Then, the probability of forming a cluster
capability of water to access HB sites of TFD. Interestingly, for of size n monotonically decreases upon increasing n. It thus
the 1% w/w water concentration, one may particularly notice again indicates that a large fraction of TFD molecules (45−
in Table 2 that the number of HBs formed between TFD 62% depending on the water concentration) actually does not
molecules remains almost the same within uncertainties form HBs with other TFD molecules and about 20% of them
compared to the pure TFD systems while TFD also forms may just form HB dimers (see Figure 9A). Large HB
HBs with water. This result suggests that, at this very low water associations seem rare or very rare (10% or below).
concentration, the few water molecules in the system actually Nevertheless, snapshots displayed in Figure 9A(c,d) reveals
fill the pre-existing empty spaces between the TFD molecules. the possibility for TFD molecules to form short HB chains,
This feature has been already nicely shown from MD which remind the presence of HB linear chains in HB liquids
simulations in lysozyme/trehalose/glycerol and trehalose/ such as monohydroxy alcohols and amides.59 The cluster size
glycerol glasses at low hydration.7 distribution shows the inability for TFD to form very large
11300 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

clusters. It certainly stems from the TFD-specific molecular Oppositely, calculations performed at the 10% w/w water
architecture (see Figure 1) as already mentioned for which the concentration show the formation of one large water aggregate
three HB groups of the TFD molecule unfavor the HB inter- (see Figure S6b). It could be hypothesized that water
molecular associations: (i) the −OH hydroxyl group (O1) is molecules present in such a large water cluster (see Figure
strongly sterically screened because of its position surrounded S6b) would be much easy to remove by standard heating
by three large rings, (ii) the central −OH hydroxyl group (O2) processes than the small fraction of quite isolated water
is more accessible, but because of its “not terminal” position molecules embedded in the TFD matrix at the 1% w/w water
(as in monohydroxy alcohols), it is not possible to build very concentration (see Figure S6a). The presence of large
large supramolecular associations through it, and (iii) the percolating water volume which may facilitate dehydration
nitrogen atom can only be the HB acceptor. In addition, the was particularly suggested in freeze-drying processes.60 It
fact that an intra-molecular HB (see Table 2) may also form should also be mentioned that both situations encountered at
between the central −OH hydroxyl group (O2) and the 1% w/w and 10% w/w are perfect illustrations of “unclustered
nitrogen atom (see Figure 9C) is another strong factor which water” and “clustered water” found in pharmaceutical systems,
contributes to the limited extend of the TFD HB clusters. which have been extensively reviewed in ref 8. In order to
For the pure TFD system, on average, isolated molecules confirm this hypothesis, some snapshots of instantaneous
represent about 45% of all TFD molecules. This fraction configurations of the 1% w/w water concentration system at T
increases to 50 and 62% upon increasing the water = 300 K have been generated and are displayed in Figure 9B.
concentration because water molecules progressively inter- They illustrate the type of interactions which can exist between
calate between TFD molecules, and HBs between TFD TFD and water in a dried sample in which some residual water
molecules are progressively replaced by HBs between TFD and is still present. Interestingly, a single water molecule is
water molecules (see Table 2 and Figure 9B). At the first particularly shown to be capable of forming at least transiently
glance, the increase of the fraction of isolated TFD molecules HBs with two (Figure 9B(a)), three (Figure 9B(b)), or even
upon increasing water content may seem surprising. In this four (Figure 9B(c)) TFD molecules. Obviously, situations for
calculation, it should be reminded that only HB interactions which water molecules form only one HB or even no HB are
between TFD molecules were actually taken into consideration also frequently detected. It is particularly remarkable that a
excluding HB interactions between TFD and water molecules. single water molecule may form simultaneously four HBs with
Obviously, the fraction of TFD molecules forming no HB its surrounding made of non-water molecules. Excluding the
either with other TFD molecules or with water molecules deformed HBs, it is actually the maximum number of HBs that
a single water molecule can form. It should be noted that
decreases upon increasing water concentration. It roughly
several occurrences of this type of HB associations have been
corresponds to 35 and 14% at 1 and 10% w/w water
detected in MD runs. The snapshots displayed in Figure 9B
concentration, respectively. Interestingly, one may note that
suggest that the presence of water may contribute to reinforce
the TFD cluster size distribution for the pure TFD system and
the affinity between TFD molecules by creating some kind of
the 1% w/w TFD/water mixtures are actually very close.
bridges through HBs. This type of effects have been reported
Therefore, a very small amount of water does not seem to in another context related to the water anchorage hypothesis,
strongly impact on the HB network existing between TFD which assumes that residual water in freeze-dried biomaterial
molecules. As already indicated, it shows that, at this low water products anchors the protein dynamics to its surroundings.61,62
concentration, the few water molecules present in the system From MD simulations, it was also particularly demonstrated
preferentially localize in the empty holes in between TFD that the presence of water molecules reinforce molecular
molecules (see Figure 9B(b,c)). Upon a further increase of the coupling for some bioprotectants such as trehalose dis-
water concentration to 10% w/w, once most of the empty accharide and proteins.7
spaces between TFD molecules are actually filled by water In order to complete this analysis and to be more
molecules, the action of water molecules actually seems to quantitative, the fraction of water molecules forming either 0,
separate TFD molecules as proven by the sudden jump of the 1, 2, 3, or 4 HBs with TFD molecules at different temperatures
fraction of isolated TFD molecules in Figure 10A from 50 to has been calculated and is shown in Figure 11. This figure
65% and a decrease of the probability to form large associated clearly demonstrates that at the highest investigated temper-
HB structures such as trimers, tetramers, and so forth. atures, water molecules are weakly HB with TFD. At T = 500
Figure 10B shows the fraction of HB aggregates formed by K, about 68, 23, and 4% of water molecules form no HB, one
water molecules in the TFD/water mixtures at 1 and 10% w/w HB, and 2 HBs with TFD, respectively. Upon decreasing
concentrations. A snapshot of an instantaneous configuration temperature, at about 400 K, the number of water molecules
of the simulation box at 1 and 10% w/w concentrations is also forming no HB with TFD significantly drops below 50%, while
displayed in Figure S6. the number of water molecules forming 1 or more HB
The distribution of HB water cluster size at 1 and 10% w/w progressively increases. One may particularly notice the
water concentration is shown in Figure 10B. At 1% w/w water continuous increase of the small fraction of water molecules
concentration, the distribution of the cluster size does not forming three or four HBs with TFD molecules. It should be
indicate the presence of large water aggregates. It is confirmed noted that these features corroborate particularly well in line
from the snapshot displayed in Figure S6a which clearly shows with the present reported experimental results, which show
that water molecules are quite dispersed in the 1% w/w TFD/ that water molecules only leave the samples in reasonable
water mixtures. From the distribution shown in Figure 10B, conditions above 453 K. At T = 300 K, it remains only about
about 70% of water molecules do not form HBs with other 32% of water molecules forming no HB with TFD. This
water molecules and about 25% may only form dimers. The decrease seems to continue even at lower temperatures. In
remaining 5% of the water molecules may form trimers or even other words, at least at low temperatures and at low water
more rarely tetramers. contents, that is, 1% w/w in the present case, it seems that
11301 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

molecules forming none, one, or two HBs with TFD (see


particularly Figure 9B(a)), which could be easy to remove
from the sample by a simple heating/drying process and (ii) a
minority of much more highly HB water molecules embedded
in the dried TFD matrix (see particularly Figure 9B(b,c)) that
could be much more difficult to remove without a specific
heating/drying treatment.
3.2.3. Link between Inter-molecular HBs and Molecular
Mobility. A detailed investigation of the link between inter-
molecular HBs formed by water molecules with TFD
molecules and water-specific translational and rotational
molecular mobility has been performed. As already mentioned,
water is recognized as an efficient plasticizer.22,26,35,39−41 The
Figure 11. Fraction of water molecules forming either 0, 1, 2, 3, or 4 increase of molecular mobility upon increasing the water
HBs with TFD molecules at different temperatures in the TFD/water content has thus been checked from the present MD
mixture at 1% w/w water concentration. Uncertainties have been simulations by calculating the time-dependent dipolar
statistically estimated by performing five independent MD simulations correlation of the TFD molecules ϕT(t) as well as the time-
at the different investigated temperatures. dependent correlation function of the TFD−TFD HBs ϕHB(t).
Both functions are shown in Figure S1 (Supporting
water molecules because of their smaller size and higher Information) at the temperature T = 450 K for pure TFD
mobility tend to interact through HB with TFD more and TFD/water mixtures at 1% and 10% w/w concentrations.
preferentially than with other water molecules well in line A faster decay is found for both ϕT(t) and ϕHB(t) upon
with the data reported in Table 2. From the calculations increasing the water content, which clearly confirms the
performed at T = 300 K, water molecules could be thus plasticization effect of water on the TFD matrix.
classified into different categories depending on the average In the following, we will mostly focus on the 1% w/w TFD/
number of HBs they may form with TFD: weakly HB water water mixture which is more relevant for comparison with the
molecules forming either 0 or 1 HB with TFD, moderately HB experimental works performed in the present paper in which
water molecules forming two HBs with TFD, and strongly HB water removal was found to be problematic. Figure 12b
water molecules forming three or four HBs with TFD. It is represents the number of HB bond formed by some targeted
thus tempting to suggest that different categories of residual water molecules (among the 16 water molecules included in
water molecules could coexist in dried samples of TFD: (i) a the simulation box at this concentration) with TFD as a
majority of weakly or moderately hydrogen-bonded water function of temperature. Two targeted water molecules have

Figure 12. (a) Mean-squared displacement of water ⟨u2w⟩ and TFD ⟨u2T⟩ molecules as a function of temperature in the TFD/water mixture at 1%
w/w water concentration. The mean-squared displacements for water molecules are calculated at time t = 10 ns. The mean-squared displacement at
time t = 10 ns of bulk TIP3P water at room temperature is also indicated for comparison (dashed line) and (b) number of HBs formed between
water and TFD molecules. The curves labeled “max HB” or “min HB” represent the results obtained for the water molecule (among the 16 water
molecules included in the simulation box at this concentration) forming the maximum or the minimum number of HBs with TFD, respectively.
The curves labeled “avg HB” represent the average values obtained considering the 16 water molecules. Uncertainties have been statistically
estimated by performing five independent MD simulations at the different investigated temperatures. (c) Time-dependent dipolar correlation
function ϕw(t) at the temperature T = 300 K (full symbol) and T = 450 K (open symbol) for three targeted water molecules labeled #1(circle), #2
(square), and #3 (triangle up), which are strongly, moderately, or weakly HB to TFD molecules. Water molecules have been selected to match with
the definition max/min/avg defined in Figure 12a,b. On average, water molecules #1 and #2 form the highest and the lowest number of HBs with
TFD, while water molecule #3 forms an intermediate number of HBs with TFD. The function ϕw(t) is calculated from the normalized
autocorrelation function ⟨μ⃗ (t)·μ⃗ (0)⟩/⟨μ⃗ 2⟩, where μ⃗ (t) is the dipole moment of a given water molecule (either #1, #2, or #3) at time t. The brackets
mean averaging over the different time origins. Dashed lines (T = 300 K) and dotted lines (T = 450 K) are a guide for the eye.

11302 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

been specifically followed: the water molecule forming the number (max HB) of HBs with TFD (see illustrations
maximum (max HB) and minimum number (min HB) of HBs displayed in Figure 9B(b,c)), a marked increase of the number
with TFD, respectively. The average number of HBs (avg HB) of HBs with TFD upon decreasing temperature is seen. It may
formed by water molecules with TFD molecules is also shown. be speculated that the trend found in the investigated
It is actually the average value at each temperature of the data temperature range will continue somehow at lower temper-
represented in Figure 11. Figure 12a shows the mean-squared atures at which the separation between min HB and max HB
displacement ⟨u2w⟩ of the two targeted water molecules as well will become even more important. This difference on the
as the average value calculated including all water molecules in number of HBs is also clearly reflected on the mean-squared
the simulated system. Figure 12a also shows for comparison displacement ⟨u2w⟩ with a sudden drop of more than 2 decades
the average mean-squared displacement ⟨u2T⟩ of the TFD between 400 and 350 K of ⟨u2w⟩ for the water molecule forming
molecules. The mean-squared displacement ⟨u2w⟩ and ⟨u2T⟩ are the maximum (max HB) with TFD. Owing to the very low
actually shown instead of the usual diffusion coefficients value of ⟨u2w⟩ ≈ 1 Å2 at 300 and 350 K, the max HB water
because the diffusive regime cannot be actually reached over molecule could be considered as almost immobilized over the
the duration of the present MD simulations at the lowest duration of the MD simulations. This sudden drop of
investigated temperatures since the investigated systems molecular mobility for some water molecules actually coincides
progressively enter into a glassy state. The arbitrary time t = with the strong decrease of molecular mobility of TFD
10 ns is chosen in order to ensure that the system has left at molecules as measured by the average mean-squared displace-
least the short-time dynamics regime associated with very fast ment ⟨u2T⟩. At the lowest investigated temperatures T = 300
motions such as vibrations and librations and enters either and 350 K, ⟨u2T⟩ also decreases to values of about 1 Å2 or
some cage dynamics or some sub-diffusive or possibly diffusive below. It reinforces the idea of the presence of water molecules
regime depending on temperature. Indeed, previous inves- completely slaved of the TFD matrix. Interestingly, one may
tigations of the water mobility in glassy matrices have shown remark that the ⟨u2w⟩ calculated for the fastest water molecule
that the paths of water molecules followed during their and the average value continuously decreases upon decreasing
somehow “tortuous” trajectories could be extremely com- temperature without significantly decreasing unlike the ⟨u2w⟩
plex29,63 including several diffusive processes. A full description for the water molecule forming the maximum (max HB) with
of diffusion of water molecules in the TFD matrices thus goes TFD. It suggests that the dynamical properties of some water
beyond the present investigation. moleculesand possibly majority of themare decoupled
At the highest temperatures T = 450 and 500 K, Figure 12b from that of the TFD matrix. These findings are in good
shows that the two selected water molecules (min HB and max agreement with previous experimental studies on decoupling of
HB) form a similar number of HB with TFD, and values are the mobility of water molecules from that of the glassy matrix
actually quite close to the average value obtained considering in which they are embedded. For maltose,64 the authors have
all water molecules. Some small differences ranging from about shown that water diffusion could be remarkably fast and
0.2 to 0.4 HBs actually exist between the two selected water decouples from the matrix diffusion, and it is not strongly
molecules. Oppositely, Figure 12a reveals that, already at those affected by the glass transition. The same results were also
high temperatures, some significant differences exist on proposed for water mobility in glassy sucrose.60 This marked
translational molecular mobility between water molecules. A difference in dynamical behavior between selected water
difference of about 1−2 orders of magnitude is found on the molecules is particularly well in line with the ideas previously
mean-squared displacement ⟨u2w⟩ between the fastest water suggested in the present study of the two different categories of
molecule, which is forming the minimum number (min HB) of water molecules, that is, weakly and strongly HB to TFD, that
HBs with TFD and the slowest water molecule, which is may coexist in the 1% w/w water system (see above and also
forming the maximum number (max HB) of HBs with TFD. Figure 9B).
Therefore, although the number of inter-molecular HBs In addition to translational dynamics of water molecules
formed by those two water molecules with TFD only slightly probed by mean-squared displacements ⟨uw2 ⟩, rotational
differs by 0.2−0.4 HBs, a significant difference on mobility is dynamics have also been analyzed using time-dependent
found. Water molecules thus seem to behave quite differently dipolar correlation functions ⟨ϕw(t)⟩ based on autocorrelation
in the system. It is interesting to note that the ⟨u2w⟩ calculated functions of an individual water molecule dipole moment.
for the fastest water molecule and the average value are close to Results are shown in Figure 12c. Functions ϕw(t) have been
the one obtained for bulk water at room temperature. It computed at two temperatures, 300 and 450 K, for three
suggests that most water molecules are diffusing very fast in the targeted molecules that are weakly, moderately, or strongly HB
high-temperature range except possibly the few water to TFD molecules in the same manner as reported in Figure
molecules which are slightly more HBs to TFD. Upon 12a,b. At 450 K, the time-dependent dipolar correlation
decreasing the temperature, from about T = 400 K, a clear functions ϕw(t) of the three targeted molecules decay very
separation emerges between the water molecules (see Figure similarly. This behavior actually contrasts with the results
12b) demonstrated by a difference of more than two HBs obtained at the same temperature for the mean-squared
between the water molecule forming the maximum (max HB) displacements ⟨u2w⟩ (see Figure 12a) for which clear differences
and the minimum number (min HB) of HBs with TFD. The were noticed between water molecules weakly and strongly HB
behavior of both molecules clearly separates from the average to TFD molecules. This feature can be understood from the
value, which increases monotonically upon decreasing the fact that a given water molecule may rotate without to
temperature well in line with the densification of the system. necessarily break the inter-molecular HBs, while it is required
For the water molecule forming the minimum number (min for the water molecule to break those HBs to diffuse. In other
HB) of HBs with TFD, a quite constant value of about 0.5 HBs words, translational degrees of freedom may be more
is found, which seems completely independent of the influenced by small changes of the HB network than rotational
temperature. For the water molecule forming the maximum degrees of freedom. It might have an impact on how water
11303 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

molecules are actually removed in drying/heating processes. In the supercooled domain, the main α-relaxation of TFD
Indeed, small differences on HB inter-molecular interactions was found to move toward higher frequencies with increasing
between water molecules and TFD could explain why some relative humidity. This feature was usually assigned to a
water molecules may easily leave from the sample while some plasticization effect because of the well-recognized ability of
others may remain. Oppositely, at T = 300 K, major differences water to increase the molecular mobility.
between the three types of water molecules are observed well The S-process is detected in the same frequency range as a
in line with the results previously reported in Figure 12a,b. previously detected process (namely, processes γ) observed in
While the ϕw(t) function of the less HB water molecule to the non-annealed TFD glass.45 Taking into account the
TFD may still fully relax and decay to zero, the relaxation is important broadness of this mode that spreads over 7 decades
only partial for the moderately HB water molecule and ϕw(t) and its quite high activation energy, the origin of the γ-process
does not decay to zero but to a plateau value. For the strongly was attributed to the intramolecular motion of the very flexible
HB water molecule to TFD, rotational dynamics is almost central part of the TFD molecule well in line with MD
arrested and no long-time decay is seen over the duration of computations. During the cycles of heating of the annealed
the MD simulation. It could be hypothesized that the different glass, the intensity of the S-process and γ-process was shown to
behaviors seen for water molecules embedded in the TFD behave in an opposite trend, suggesting a peculiar coupling
matrix, either weakly, moderately, or strongly HB to TFD between these two sub-Tg relaxations. Moreover, it was also
molecules, may give rise to some additional dynamical observed that the sample had to be heated up to 453 K (Tm
processes, which could be observed in DRS experiments +30 K) and annealed at this temperature for about 1 h to
because of the very high polarity of the water molecule. completely remove the S-process and then find the signature of
It particularly reinforces the previously suggested ideas that the non-annealed glass. It was interpreted as an indication of
the origin of the supplementary S-process would originate from strongly HB-bonded water molecules trapped in the TFD
peculiar dynamics of residual water molecules. Since these matrix.
water molecules seem to preferentially localize through HB MD simulations performed on TFD samples at different
close to the central part of the TFD molecule, one could thus water concentrations allowed us to clarify these two unconven-
speculate on the existence of some kind of coupling between tional features (coupling between γ- and S-processes and
dynamics of these water molecules and dynamics of the very strongly HB water molecules) which allowed us to progress on
flexible central part of the TFD molecule, which is also the exact nature of secondary relaxations. From MD
suspected to give rise to the very broad γ-relaxation.45 It would computations and analyses of the HB interactions, two
thus explain why the S-process is exactly detected in the same categories of water molecules were revealed: i) a majority of
frequency range as the γ-relaxation and could be then weakly or moderately HB water molecules to TFD and ii) a
described as a component of the γ-relaxation of slower minority of much more highly HB water molecules strongly
mobility. interacting with the OH hydroxyl group and the nitrogen atom
localized close to the flexible central part of the TFD molecule.
4. CONCLUSIONS Weakly or moderately HB water molecules to TFD seem easier
to remove from the system by usual heating/drying processes.
In this article, the liquid, supercooled liquid, and glassy states Such water molecules particularly tend to form very large water
of TFD have been investigated by means of complementary clusters at the highest water concentration 10% w/w (see
DRS experiments and MD simulations. Figure 10A). Dynamics of this type of water molecules enter in
A new secondary relaxations process (namely process S) is the pico-nanosecond time domain (GHz to THz frequency
detected for the first time in the glassy state of TFD annealed range) accessible to the MD simulations. However, these
at room temperature (Tg30 °C). This process is completely dynamics are too fast to be observed since they are outside of
absent in a non-annealed glass of TFD. The molecular origins the present DRS investigated frequency range. Oppositely,
behind secondary relaxations are often under controversy, and highly HB water molecules to TFD are more dispersed in the
therefore, thorough investigations to understand the micro- TFD matrix and do not form large water clusters between
scopic origin of the S-process have been performed. We have them. Actually, very small water clusters made of few water
shown that this mode is clearly linked to the presence of molecules may still exist in the very dried sample at 1% w/w
residual water in the concentration of 1−2% (w/w). Several (see the inset in Figure 10B). These water molecules localize
results supporting this attribution were particularly obtained. themselves in small pockets in the pre-existing voids spaces
During the cycles of heating of the annealed glass, it was existing in between the TFD molecules due to the poor
shown that the S-process magnitude decreases and shifts packing. Because of their strong HB interactions with TFD and
toward lower frequencies. These progressive heatings could be weak mobility, they seem much more difficult to remove
assimilated to a form of drying of the sample, which would without a specific treatment. These highly HB water molecules
result in a loss of absorbed water. might particularly explain why it is necessary to heat the
From annealing studies in controlled humidity, based on sample at so high temperatures to completely remove the S-
TGA experiments, it was actually confirmed that the TFD glass process. Molecular mobilities of these water molecules which
can absorb a significant amount of water (about 2%) in are typically in the [102 to 105] Hz frequency domain are not
contrast to the known hydrophobic characteristic of crystalline accessible to MD simulations.
TFD. During the cycles of heating of the annealed glass or from
The temperature dependence of the S-process was shown to annealing studies in controlled humidity, it was shown that the
possess an Arrhenian behavior with an activation energy of S-process magnitude decreases and shifts toward lower
about 50 kJ·mol−1, which is the value reported for processes frequencies upon decreasing the water content. This can be
usually seen in DRS and associated with the influence of understood assuming that the S-process originates from the
water.34 reorientation of the remaining water molecules that are still
11304 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

present even in the dried samples and which are bound to TFD Author Contributions
molecules. At 1% w/w water concentration, a given water The manuscript was written through contributions of all
molecule belonging to these small clusters may interact both authors. All authors have given approval to the final version of
with a TFD molecule, which tends to reduce its mobility and the manuscript.
to other water molecules of the same cluster, which tend to Notes
accelerate its mobility. It seems obvious that water molecules The authors declare no competing financial interest.


are more mobile than the TFD molecule. Upon a further
decrease of the water content, the number of molecules in the ACKNOWLEDGMENTS
water clusters will be decreased. Accordingly, the given water
molecule will interact less with its surrounding neighboring This project has received funding from the Interreg 2 Seas
water molecules and more with TFD. Its mobility will be thus program 2014−2020 co-funded by the European Regional
decreased well in line with the frequency of the S-process, Development Fund (FEDER) under subsidy contract 2S01-
059_IMODE.


which decreases upon decreasing water content. Furthermore,
taking into account that the origin of the γ-process is
associated with molecular mobility of the central part, and REFERENCES
since the remaining water molecules tend to localize (1) Descamps, M. Disordered Pharmaceuticals Materials; Wiley, 2016.
preferentially close to this zone, it may explain the peculiar (2) Newman, A.; Zografi, G. What we need to know about solid-
coupling found between the S- and the γ-processes. state isothermal crystallization of organic molecules from the
The present study is an attempt to clarify the microscopic amorphous state below the glass transition temperature. Mol.
origin of secondary relaxations detected in TFD and the role of Pharm. 2020, 17, 1761−1777.
(3) Harry, L. Amorphous Food and Pharmaceutical Systems; NBN
water. It clearly motivates to perform investigation on other International: Cambridge, 2007.
systems to test the universality of our findings.


(4) Manning, M. C.; Chou, D. K.; Murphy, B. M.; Payne, R. W.;
Katayama, D. S. Stability of protein pharmaceuticals: An update.
ASSOCIATED CONTENT Pharm. Res. 2010, 27, 544−575.
*
sı Supporting Information (5) Cicerone, M. T.; Douglas, J. F. β-Relaxation governs protein
The Supporting Information is available free of charge at stability in sugar-glass matrices. Soft Matter 2012, 8, 2983−2991.
https://pubs.acs.org/doi/10.1021/acs.jpcb.1c06087. (6) Cicerone, M. T.; Pikal, M. J.; Qian, K. K. Stabilization of proteins
in solid form. Adv. Drug Delivery Rev. 2015, 93, 14−24.
Time-dependent dipolar autocorrelation function ϕT(t) (7) Lerbret, A.; Affouard, F. Molecular packing, hydrogen bonding,
of TFD molecules and time-dependent autocorrelation and fast dynamics in lysozyme/trehalose/glycerol and trehalose/
function ϕHB(t) of inter-molecular HBs formed between glycerol glasses at low hydration. J. Phys. Chem. B 2017, 121, 9437−
the TFD molecule; comparison of tan δ, ε″, and M″ 9451.
versus F at 228 K for the annealed (17 days) and non- (8) Authelin, J. R.; MacKenzie, A. P.; Rasmussen, D. H.; Shalaev, E.
annealed glasses of TFD; Arrhenius diagram of annealed Y. Water clusters in amorphous pharmaceuticals. J. Pharm. Sci. 2014,
103, 2663−2672.
and non-annealed glass of TFD; TGA description and
(9) Andronis, V.; Yoshioka, M.; Zografi, G. Effects of sorbed water
weight losses of crystalline TFD and TFD glasses on the crystallization of indomethacin from the amorphous state. J.
annealed 45 days at RH ≈ 36% and 21 days at RH ≈ Pharm. Sci. 1997, 86, 346−351.
76% as a function of temperature; invariance of the (10) Ngono, F.; Cuello, G. J.; Jiménez-Ruiz, M.; Willart, J.-F.;
shape of the α-relaxation of TFD samples containing 0, Guerain, M.; Wildes, A. R.; Stunault, A.; Hamoudi-Ben Yelles, C.-M.;
1.26, and 1.76% of water; snapshots of instantaneous Affouard, F. Morphological and structural properties of amorphous
configurations as obtained from MD simulations at T = lactulose studied by scanning electron microscopy, polarized neutron
300 K in the TFD/water mixture at 1% w/w and 10% scattering, and molecular dynamics simulations. Mol. Pharm. 2020, 17,
water concentration; and TGA description (PDF) 10−20.


(11) Debenedetti, P. G.; Stillinger, F. H. Supercooled liquids and the
glass transition. Nature 2001, 410, 259−267.
AUTHOR INFORMATION (12) Ediger, M. D.; Harrowell, P. Perspective: Supercooled liquids
Corresponding Author and glasses. J. Chem. Phys. 2012, 137, 080901.
Frédéric Affouard − University Lille, CNRS, INRAE, Centrale (13) Willart, J. F.; Descamps, M. Solid state amorphization of
Lille, UMR 8207UMETUnité Matériaux et pharmaceuticals. Mol. Pharm. 2008, 5, 905−920.
(14) Descamps, M.; Willart, J. F. Perspectives on the amorphisation/
Transformations, Lille F-59000, France; orcid.org/0000- milling relationship in pharmaceutical materials. Adv. Drug Delivery
0001-8429-6416; Phone: +33 (0)3 20 43 68 15; Rev. 2016, 100, 51−66.
Email: Frederic.Affouard@univ-lille.fr; Fax: +33 (0)3 20 (15) Brittain, H. G. Effects of mechanical processing on phase
43 40 84 composition. J. Pharm. Sci. 2002, 91, 1573−1580.
(16) Suryanarayana, C. Mechanical alloying and milling. Prog. Mater.
Authors Sci. 2001, 46, 1−184.
Jeanne-Annick Bama − University Lille, CNRS, INRAE, (17) Baláž, P.; Achimovičová, M.; Baláž, M.; Billik, P.; Cherkezova-
Centrale Lille, UMR 8207UMETUnité Matériaux et Zheleva, Z.; Criado, J. M.; Delogu, F.; Dutková, E.; Gaffet, E.; Gotor,
Transformations, Lille F-59000, France F. J.; Kumar, R.; Mitov, I.; Rojac, T.; Senna, M.; Streletskii, A.;
Emeline Dudognon − University Lille, CNRS, INRAE, Wieczorek-Ciurowa, K. Hallmarks of mechanochemistry: From
Centrale Lille, UMR 8207UMETUnité Matériaux et nanoparticles to technology. Chem. Soc. Rev. 2013, 42, 7571−7637.
Transformations, Lille F-59000, France; orcid.org/0000- (18) Tang, X.; Pikal, M. J. Design of freeze-drying processes for
0003-4176-1925 pharmaceuticals: Practical advice. Pharm. Res. 2004, 21, 191−200.
(19) Broadhead, J.; Edmond Rouan, S. K.; Rhodes, C. T. The spray
Complete contact information is available at: drying of pharmaceuticals. Drug Dev. Ind. Pharm. 1992, 18, 1169−
https://pubs.acs.org/10.1021/acs.jpcb.1c06087 1206.

11305 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

(20) Singh, A.; Van den Mooter, G. Spray drying formulation of (42) Zografi, G. States of water associated with solids. Drug Dev. Ind.
amorphous solid dispersions. Adv. Drug Delivery Rev. 2016, 100, 27− Pharm. 1988, 14, 1905−1926.
50. (43) Canotilho, J.; Costa, F. S.; Sousa, A. T.; Redinha, J. S.; Leitao,
(21) Bhattacharya, S.; Suryanarayanan, R. Local mobility in M. L. P. Calorimetric study of polymorphic forms of terfenadine.
amorphous pharmaceuticalscharacterization and implications on Thermochim. Acta 1997, 299, 1−6.
stability. J. Pharm. Sci. 2009, 98, 2935−2953. (44) Leităo, M. L. P.; Canotilho, J.; Cruz, M. S. C.; Pereira, J. C.;
(22) Mehta, M.; Kothari, K.; Ragoonanan, V.; Suryanarayanan, R. Sousa, A. T.; Redinha, J. S. Study of polymorphism from dsc melting
Effect of water on molecular mobility and physical stability of curves; polymorphs of terfenadine. J. Therm. Anal. Calorim. 2002, 68,
amorphous pharmaceuticals. Mol. Pharm. 2016, 13, 1339. 397−412.
(23) Zhao, H.; Chen, Z.; Du, X.; Chen, L. Contribution of different (45) Dudognon, E.; Bama, J.-A.; Affouard, F. Molecular mobility of
state of adsorbed water to the sub-tg dynamics of cellulose. Carbohydr. terfenadine: Investigation by dielectric relaxation spectroscopy and
Polym. 2019, 210, 322−331. molecular dynamics simulation. Mol. Pharm. 2019, 16, 4711.
(24) Mathesan, S.; Rath, A.; Ghosh, P. Insights on water dynamics in (46) Hakanen, A.; Laine, E. Characterization of two terfenadine
the hygromorphic phenomenon of biopolymer films. J. Phys. Chem. B polymorphs and a methanol solvate: Kinetic study of the thermal
2017, 121, 4273−4282. rearrangement of terfenadine from the methanol solvate to the lower
(25) Townrow, S.; Roussenova, M.; Giardiello, M.-I.; Alam, A.; melting polymorph. Thermochim. Acta 1995, 248, 217−227.
Ubbink, J. Specific volume−hole volume correlations in amorphous (47) Yoshihashi, Y.; Kitano, H.; Yonemochi, E.; Terada, K.
carbohydrates: Effect of temperature, molecular weight, and water Quantitative correlation between initial dissolution rate and heat of
content. J. Phys. Chem. B 2010, 114, 1568−1578. fusion of drug substance. Int. J. Pharm. 2000, 204, 1−6.
(26) Andronis, V.; Zografi, G. The molecular mobility of (48) Badwan, A. A.; Al Kaysi, H. N.; Owais, L. B.; Salem, M. S.;
supercooled amorphous indomethacin as a function of temperature Arafat, T. A. Terfenadine, in Analytical Profiles of Drug Substances.
and relative humidity. Pharm. Res. 1998, 15, 835−842. Elsevier B.V: 1990; Vol. 19, pp 627−662.
(27) Shalaev, E.; Soper, A. K. Water in a soft confinement: Structure (49) Samra, R.; Buckton, G. The crystallisation of a model
of water in amorphous sorbitol. J. Phys. Chem. B 2016, 120, 7289− hydrophobic drug (terfenadine) following exposure to humidity and
7296. organic vapours. Int. J. Pharm. 2004, 284, 53−60.
(28) Molinero, V.; Goddard, W. A., III Microscopic mechanism of (50) Havriliak, S.; Negami, S. A. A complex plane analysis of α-
water diffusion in glucose glasses. Phys. Rev. Lett. 2005, 95, 045701. dispersions in some polymers systems. J. Polym. Sci., Part C: Polym.
(29) Kulasinski, K.; Guyer, R.; Derome, D.; Carmeliet, J. Water Symp. 1966, 14, 99−117.
diffusion in amorphous hydrophilic systems: A stop and go process. (51) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R.
Langmuir 2015, 31, 10843−10849. W.; Klein, M. L. Comparison of simple potential functions for
(30) Faupel, F.; Frank, W.; Macht, M.-P.; Mehrer, H.; Naundorf, V.; simulating liquid water. J. Chem. Phys. 1983, 79, 926−935.
Rätzke, K.; Schober, H. R.; Sharma, S. K.; Teichler, H. Diffusion in (52) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development
metallic glasses and supercooled melts. Rev. Mod. Phys. 2003, 75, and testing of the opls all-atom force field on conformational
237−280. energetics and properties of organic liquids. J. Am. Chem. Soc. 1996,
(31) Angell, C. A. Dynamic processes in ionic glasses. Chem. Rev. 118, 11225−11236.
1990, 90, 523−542. (53) Smith, W.; Forester, T. R.; Todorov, I. T. The dl_poly Classic
(32) Ngai, K. L. Relaxation and Diffusion in Complex Systems; User Manual; Daresbury Laboratory: United Kingdom, 2012.
Springer: New York, 2011. (54) Cole, K. S.; Cole, R. H. Dispersion and absorption in dielectrics
(33) Schönhals, A. Dielectric Spectroscopy on the Dynamics of i. Alternating current characteristics. J. Chem. Phys. 1941, 9, 341−351.
Amorphous Polymeric Systems, 1998; Vol. 1. (55) Murdande, S. B.; Pikal, M. J.; Shanker, R. M.; Bogner, R. H.
(34) Cerveny, S.; Schwartz, G.; Bergman, R.; Swenson, J. Glass Solubility advantage of amorphous pharmaceuticals: II. Application of
transition and relaxation processes in supercooled water. Phys. Rev. quantitative thermodynamic relationships for prediction of solubility
Lett. 2005, 93, 245702. enhancement in structurally diverse insoluble pharmaceuticals. Pharm.
(35) Chan, R. K.; Pathmanathan, K.; Johari, G. P. Dielectric Res. 2010, 27, 2704−2714.
relaxations in the liquid and glassy states of glucose and its water (56) Lerbret, A.; Bordat, P.; Affouard, F.; Descamps, M.; Migliardo,
mixtures. J. Phys. Chem. 1986, 90, 6358−6362. F. How homogeneous are the trehalose, maltose, and sucrose water
(36) Kaminski, K.; Kaminska, E.; Ngai, K. L.; Paluch, M.; solutions? An insight from molecular dynamics simulations. J. Phys.
Wlodarczyk, P.; Kasprzycka, A.; Szeja, W. Identifying the origins of Chem. B 2005, 109, 11046−11057.
two secondary relaxations in polysaccharides. J. Phys. Chem. B 2009, (57) Atawa, B.; Correia, N. T.; Couvrat, N.; Affouard, F.; Coquerel,
113, 10088−10096. G.; Dargent, E.; Saiter, A. Molecular mobility of amorphous N-acetyl-
(37) Kaminski, K.; Adrjanowicz, K.; Kaminska, E.; Grzybowska, K.; α-methylbenzylamine and Debye relaxation evidenced by dielectric
Hawelek, L.; Paluch, M.; Tarnacka, M.; Gruszka, I.; Kasprzycka, A. relaxation spectroscopy and molecular dynamics simulations. Phys.
Impact of water on molecular dynamics of amorphous α-, β-, and γ- Chem. Chem. Phys. 2019, 21, 702−717.
cyclodextrins studied by dielectric spectroscopy. Phys. Rev. E 2012, 86, (58) Ottou Abe, M. T.; Viciosa, M. T.; Correia, N. T.; Affouard, F.
031506. Impact of chirality on peculiar ibuprofen molecular dynamics:
(38) Schartel, B.; Wendling, J.; Wendorff, J. H. Cellulose/poly(vinyl Hydrogen bonding organization and syn vs. Anti carboxylic group
alcohol) blends. 1. Influence of miscibility and water content on conformations. Phys. Chem. Chem. Phys. 2018, 20, 29528−29538.
relaxations. Macromolecules 1996, 29, 1521−1527. (59) Böhmer, R.; Gainaru, C.; Richert, R. Structure and dynamics of
(39) Hancock, B. C.; Zografi, G. The relationship between the glass monohydroxy alcohols-Milestones towards their microscopic under-
transition temperature and the water content of amorphous standing, 100 years after Debye. Phys. Rep. 2014, 545, 125−195.
pharmaceutical solids. Pharm. Res. 1994, 11, 471−477. (60) Shalaev, E.; Soper, A.; Zeitler, J. A.; Ohtake, S.; Roberts, C. J.;
(40) Blasi, P.; D’Souza, S. S.; Selmin, F.; Deluca, P. P. Plasticizing Pikal, M. J.; Wu, K.; Boldyreva, E. Freezing of aqueous solutions and
effect of water on poly(lactide-co-glycolide). J. Controlled Release chemical stability of amorphous pharmaceuticals: Water clusters
2005, 108, 1−9. hypothesis. J. Pharm. Sci. 2019, 108, 36−49.
(41) Aso, Y.; Yoshioka, S.; Kojima, S. Relationship between water (61) Cordone, L.; Cottone, G.; Giuffrida, S.; Palazzo, G.; Venturoli,
mobility, measured as nuclear magnetic relaxation time, and the G.; Viappiani, C. Internal dynamics and protein-matrix coupling in
crystallization rate of amorphous nifedipine in the presence of some trehalose-coated proteins. Biochim. Biophys. Acta, Proteins Proteomics
pharmaceutical excipients. Chem. Pharm. Bull. 1996, 44, 1065−1067. 2005, 1749, 252−281.

11306 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

(62) Francia, F.; Dezi, M.; Mallardi, A.; Palazzo, G.; Cordone, L.;
Venturoli, G. Protein−Matrix Coupling/Uncoupling in “Dry” Systems
of Photosynthetic Reaction Center Embedded in Trehalose/Sucrose:
The Origin of Trehalose Peculiarity. J. Am. Chem. Soc. 2008, 130,
10240−10246.
(63) Johnson, D. L.; Koplik, J.; Dashen, R. Theory of dynamic
permeability and tortuosity in fluid-saturated porous media. J. Fluid
Mech. 1987, 176, 379−402.
(64) Zhu, L.; Cai, T.; Huang, J.; Stringfellow, T. C.; Wall, M.; Yu, L.
Water self-diffusion in glassy and liquid maltose measured by raman
microscopy and nmr. J. Phys. Chem. B 2011, 115, 5849−5855.

11307 https://doi.org/10.1021/acs.jpcb.1c06087
J. Phys. Chem. B 2021, 125, 11292−11307

You might also like