You are on page 1of 16

Article

pubs.acs.org/molecularpharmaceutics

Molecular Dynamics, Physical Stability and Solubility Advantage


from Amorphous Indapamide Drug
Z. Wojnarowska,*,† K. Grzybowska,† L. Hawelek,†,‡ M. Dulski,† R. Wrzalik,† I. Gruszka,† and M. Paluch†

Institute of Physics, University of Silesia, ul. Uniwersytecka 4, 40-007 Katowice, Poland

Institute of Non-Ferrous Metals, ul. Sowinskiego 5, 44-100 Gliwice, Poland

K. Pienkowska and W. Sawicki


Department of Physical Chemistry, Medical University of Gdansk, Hallera 107, 80-416, Gdansk, Poland

P. Bujak
Faculty of Chemistry, Warsaw University of Technology, Noakowskiego 3, 00664 Warszawa, Poland

K. J. Paluch and L. Tajber


School of Pharmacy and Pharmaceutical Sciences, Trinity College Dublin, Dublin 2, Ireland

J. Markowski
ENT Department, Silesian Medical University, ul. Francuska 20, Katowice, Poland

ABSTRACT: This study for the first time investigates physicochemical properties
of amorphous indapamide drug (IND), which is a known diuretic agent commonly
used in the treatment of hypertension. The solid-state properties of the vitrified,
cryomilled and ball-milled IND samples were analyzed using X-ray powder
diffraction (XRD), mass spectrometry, nuclear magnetic resonance (NMR),
infrared spectroscopy (FT-IR), differential scanning calorimetry (DSC) and
broadband dielectric spectroscopy (BDS). These analytical techniques enabled
us (i) to confirm the purity of obtained amorphous samples, (ii) to describe the
molecular mobility of IND in the liquid and glassy state, (iii) to determine the
parameters describing the liquid-glass transition i.e. Tg and dynamic fragility, (iv)
to test the chemical stability of amorphous IND in various temperature conditions
and finally (v) to confirm the long-term physical stability of the amorphous
samples. These studies were supplemented by density functional theory (DFT)
calculations and apparent solubility studies of the amorphous IND in 0.1 M HCl, phosphate buffer (pH = 6.8), and water (25 and 37 °C).
KEYWORDS: dielectric spectroscopy, molecular dynamics, glass transition, diuretic agents,
amorphous active pharmaceutical ingredients, tautomerization

■ INTRODUCTION
Indapamide (IND) is a sulfonamide diuretic drug used in the
During the past decade improvement of solubility of poorly
water-soluble medicines has become one of the most important
treatment of hypertension. Despite having a slightly different aspects of pharmaceutical research. In the literature one can
chemical structure than thiazides (e.g., hydrochlorothiazide), its find very few different approaches as to how to realize this task.
mechanism of action remains similar. It increases the urine volume One of the methods is to modify the active component by
by increasing the renal excretion of sodium, chlorine, potassium and means of salification. It means that the pure drug is transferred
magnesium ions. Apart from the diuretic action indapamide exerts into nitrate, hydrochloric or other salts which usually reveal
also spasmolytic effects on blood vessels, consequently reducing the
blood pressure.1 However, the commercial form of IND available on Received: February 28, 2013
the market works very weakly. The most probable reason is its low Revised: August 22, 2013
solubility (75 mg/L) and consequently poor bioavailability.2 Thus, Accepted: September 5, 2013
the question arises how to improve the solubility of indapamide? Published: September 5, 2013

© 2013 American Chemical Society 3612 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics


Article

better solubility than the original compound.3,4 Additionally, EXPERIMENTAL SECTION


one can also apply several formulation strategies such as
Material. The tested indapamide hemihydrate 4-chloro-N-
micronization,5 molecular dispersion,6 incorporation of surfac-
tants, comilling with sugars,7 cyclodextrins or PVP and (2-methyl-2,3-dihydro-1H-indol-1-yl)-3-sulfamoylbenzamide
amorphization.8,9 The last method is a relatively simple (C16H16ClN3O3S·(1/2)H2O), MW = 365.83 g/mol with the
approach and thus widely employed to enhance the apparent chemical structure presented in Figure 1, was supplied from
solubility of pharmaceutical materials. As examples one can
mention amorphous verapamil hydrochloride,10 glipizide11 or
celecoxib.12 It should be noted that apart from affording a
higher apparent solubility (and consequently higher bioavail-
ability) an amorphous form can also improve processability of
drugs.13 For instance, by tableting an amorphous pharmaceu-
tical as opposed to the crystalline form one can reduce the
amount of additives in the tablet and thus simplify the
formulation. Unfortunately, preparation of amorphous drugs
may be also quite complex. First, an amorphous phase is
thermodynamically unstable, which in consequence may lead to
crystallization of the drug under storage conditions and
subsequent loss of the benefits arising from the amorphous Figure 1. Chemical structure of indapamide.
state.14 Second, as a result of higher chemical reactivity,
amorphous pharmaceuticals are prone to chemical degradation Quimica Sintetica (Madrid, Spain) as a white crystalline powder
as well as isomeric transformations during manufacturing and and was used without further purification. The powder X-ray
storage. Tautomerization, reversible proton migration in an diffractogram of crystalline IND (see blue line in Figure 2) was
organic molecule, is of a great importance in pharmaceutical
science since it has been shown that preparation of an
amorphous pharmaceutical, in which tautomerization reaction
takes place, can lead to a state which is chemically different
from its crystalline counterpart.15 This phenomenon can have
positive as well as negative impacts on the quality of amorphous
drugs. On the one hand, one can expect that the presence of
tautomerization products may affect the drug solubility, and
consequently it may enhance the bioavailability of the active
substance. Moreover, the occurrence of two or more tautomers
in the sample may reduce its crystallization tendency.34 Also,
tautomerization of drugs might lead to serious medical
consequences, since one cannot exclude that the biological
activity of the given tautomers may be different from that of the
parent drug. All these factors are reasons why in recent years
one of the main challenges of developing amorphous
pharmaceuticals is to understand the molecular interactions in
Figure 2. Thermal analysis of crystalline and amorphous forms of
the glassy and supercooled liquid phases and subsequently to IND. All DSC thermograms were obtained during heating at a rate of
predict the physical and chemical stability of amorphous drugs. 10 K/min. The inset panel shows the glass transition regions of
In this study we investigate the amorphous indapamide drug. quenched, cryomilled and ball-milled IND samples.
Polymorphic and pseudopolymorphic forms of indapamide16
have been identified and characterized, however no reports are
available on its amorphous form. Three different preparation in good agreement with that of the commercial material as
techniques, such as quench cooling of the melt, ball milling published recently by Ghugare et al.16 Indapamide is soluble in
and cryogrinding, were applied to convert the crystalline acetonitrile, ethyl acetate, glacial acetic acid, methanol, ethanol
indapamide into its amorphous form. To confirm the and other alcoholic solvents. It is very slightly soluble in ether
amorphous nature of the examined materials, X-ray powder and chloroform, while practically insoluble in water.
diffraction (XRD) and differential scanning calorimetry (DSC) Amorphous Samples Preparation. Quench Cooling of
were used. The DSC was also applied to determine the glass the Melt. Crystalline IND hemihydrate was placed on a
transition temperature (Tg) of amorphous IND. The results of stainless steel plate and heated on a hot plate until a complete
FT-IR experimental technique give us information about the melt was achieved (the sample becomes a yellow, transparent
chemical stability of indapamide drug. Furthermore, dielectric film). The molten sample was then quenched by being put onto
spectroscopy (BDS), which is often applied to study liquid− a cold tile.
glass transitions in various pharmaceuticals, was employed to Ball Milling. The room temperature ball milling was
characterize the molecular dynamics of the analyzed diuretic performed using a Planetary Ball Mill (Retsch, Germany). A
agent. Experimental results obtained were supplemented by zirconium jar was filled with the examined material and 6
atomic density functional theory (DFT) calculations. Finally, zirconia balls (20 mm in diameter). The rotation speed was set
the apparent solubility studies of amorphous IND samples were to 400 rpm. We have performed three separate tests with the
performed highlighting crystallization tendency of amorphous same amount of material (6 g) applying different grinding times.
phase in aqueous media. Each milling cycle lasted 15 min and was followed by a 5 min break.
3613 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

Total milling times of indapamide were 1 h, 2.5 h and 5 h. After Sealed aluminum crucibles (40 μL) with lids containing five
milling, indapamide becomes a yellow powder. punctures to enable water to volatilize during heating were used
Cryogenic Grinding. Cryogenic grinding of indapamide was for all samples. Standard DSC measurements were performed at a
carried out by means of a 6770 SPEX freezer/mill. The total heating rate of 10 K/min under a nitrogen purge (60 mL/min).
mass of the milled indapamide was 3 g. The sample was placed Moreover, using a stochastic temperature-modulated differ-
in a stainless steel vessel and was immersed in liquid nitrogen. ential scanning calorimetry (TMDSC) technique implemented
The stainless steel rod present in the vessel is vibrated by by Mettler-Toledo, TOPEM , the dynamic behavior of the
means of a magnetic coil. Prior to the start of grinding, the glass−liquid transition of the studied materials has been
sample was subjected to 10 min of precooling. The mill was set analyzed in the frequency range from 4 mHz to 40 mHz in
to function at an impact frequency of 15 Hz. Five minute one single measurement at a heating rate of 0.5 K/min. In the
grinding intervals were separated by three minute cool-down experiment, the temperature amplitude of the pulses of 0.5 K
periods. The effective grinding times were 15, 30, and 45 min. was selected. The calorimetric structural relaxation times τα =
After the cryogenic grinding, the vessel with the ground sample 1/2πf were determined from the temperature dependences of
was equilibrated in a vacuum oven at 25 °C, until room the real part of the complex heat capacity cp′(T) obtained at
temperature was reached. After milling, indapamide becomes a different frequencies in the glass transition region. The glass
yellow powder. transition temperature Tg was determined for each frequency as
Analytical Techniques. XRD measurements. XRD experi- the temperature of the half step height of Cp′(T).
ments were performed at ambient temperature on a Rigaku- Infrared Measurements (FT-IR). Infrared measurements
Denki D/MAX RAPID II-R diffractometer (Rigaku Corpo- were performed using a Bio-Rad FTS-6000 spectrometer
ration, Tokyo, Japan) equipped with a rotating anode Ag KR equipped with a KBr beam splitter, a standard source and a
tube (λ = 0.5608 Å), an incident beam (002) graphite DTGS Peltier-cooled detector. The single crystalline and glassy
monochromator and an image plate in the Debye−Scherrer spectra of indapamide have been collected using GladiATR
geometry. The pixel size was 100 μm × 100 μm. Studied diamond accessory (Pike Technologies) in the range from 400
samples were placed inside Lindemann glass capillaries (1.5 mm to 4000 cm−1. The temperature measurements were carried out
in diameter). Then, the measurements were performed on using a heated plate of the GladiATR diamond accessory (Pike
sample-filled and empty capillaries, and the background Technologies) equipped with a temperature controller with
intensity of the empty capillary was subtracted from the sample accuracy ±0.1 K. Time-dependent spectra were recorded every
signal. The beam width at the sample was 0.1 mm. The two- 10 min. All spectra were recorded with a spectral resolution of
dimensional diffraction patterns were converted into one- 2 cm−1. Presented results are an average of 16 scans in order to
dimensional intensity data using suitable software. ensure good quality spectra.
Nuclear Magnetic Resonance (NMR). The 1H and 13C Mass Spectrometry. Liquid chromatography−mass spec-
NMR spectra were recorded at 400 and 100 MHz, respectively, trometry (LC−MS) was carried out using a Thermo Accela
on a Bruker 400 (Germany). Liquid chromatography coupled to a Thermo LTQ-XL-
1
H NMR (400 MHz, DMSO-d6): δ = 1.30 ppm (d, J = Orbitrap Discovery mass spectrometer as the detector. The
6.1 Hz, 3H, CH3); 2.59 (dd, J = 15.3 Hz, J = 11.1 Hz, 1H, CH2); column used for chromatographic separation was a Waters
3.17 (dd, J = 15.3 Hz, J = 8.1 Hz, 1H, CH2); 3.93−3.99 (m, 1H, Acquity HSS T3 C18, 2.1 mm × 150 mm 1.8 μm at operating
CH); 6.50 (d, J = 7.8 Hz, 1H, CH); 6.75−6.78 (m, 1H, CH); at 30 °C. A flow rate of 100 μL/min was used with an injection
7.02−7.06 (m, 1H, CH); 7.11 (d, J = 7.2 Hz, 1H, CH); 7.74 volume of 4 μL. The mobile phase consisted of two solutions:
(s, 2H, NH2); 7.80 (d, J = 8.3 Hz, 1H, CH); 8.12 (dd, J = 0.1% v/v formic acid in HPLC grade water as phase A and 0.1%
8.3 Hz, J = 2.2 Hz, 1H, CH); 8.51 (d, J = 2.2 Hz, 1H, CH); v/v formic acid in HPLC grade acetonitrile as phase B. The
10.53 (s, 1H, NH). total run time was 10 min, and the following gradient method
13
C NMR (100 MHz, DMSO-d6): δ = 18.5 ppm (CH3); 35.4 was used: 20% phase B to 80% phase B over 8.00 min, hold
(CH2); 62.9 (CH); 108.7 (CH), 119.9 (CH); 124.4 (CH); until 7.00 min, then 20% phase B at 7.01 min and equilibrate
127.0 (CH); 127.1 (C); 128.3 (CH); 131.6 (C); 131.7 (CH); for 3 min. The LTQ-XL ion trap mass spectrometer was
132.1 (CH); 133.6 (C); 141.3 (C); 151.4 (C); 163.9 (CO) coupled to the Accela LC system via an electrospray ionization
Broadband Dielectric Spectroscopy (BDS). Ambient (ESI) probe. The capillary temperature was maintained at
pressure dielectric measurements of glassy and supercooled 310 °C, sheath gas flow rate 60 arbitrary units, auxiliary gas flow
indapamide were performed over a wide frequency range from rate 5 arbitrary units, sweep gas flow rate 0 arbitrary units,
10−1 to 106 Hz using a Novo-Control GMBH Alpha dielectric source voltage 3.20 kV, source current 100 μA, capillary voltage
spectrometer. For the isobaric measurements, the sample was 48 V and tube lens 82 V. Indapamide was detected in negative
placed between two stainless steel electrodes of the capacitor ion mode, and its retention time was 6.6 min. Full FTMS
(diameter 20 mm) with a gap of 0.1 mm. The dielectric spectra scanning (m/z 80−1000) was used to detect degradation
were collected in a wide temperature range from 133 to 441 K. products. The samples as methanolic solutions were also
The temperature was controlled by the Novo-Control Quattro infused directly into the ion chamber at a flow rate of 5 μL/min
system, with the use of a nitrogen gas cryostat. Temperature using a Hamilton syringe.
stability of the samples was better than 0.1 K. Apparent Solubility Measurements. HPLC Equipment
Differential Scanning Calorimetry (DSC). Calorimetric for the Solubility Assay. In this study, a high performance
measurements of crystalline and amorphous indapamide were liquid chromatography HPLC system (Shimazu LCsolution
carried out by Mettler-Toledo DSC apparatus equipped with a Chromatography System, model Prominence LC-20A) was
liquid nitrogen cooling accessory and a HSS8 ceramic sensor used to quantify the amount of indapamide dissolved. The
(heat flux sensor with 120 thermocouples). Temperature and system was equipped with a UV/vis detector (model SPD-20 A),
enthalpy calibrations were performed using indium and zinc a quaternary gradient pump (model LC-20 AD) with a parallel-
standards. type double plunger, a degasser (model DGU-20 A5) with a
3614 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

fluoroethylene membrane, an autosampler (model Sil-20 A HT) theory, relaxed geometry scan by changing proper dihedral
with two trays for standard 1.5 mL vials and a column oven angle was performed. Thereafter, structures characterized by
(model CTO-1-AS VP). A Microsorb 100-5 C8 (S150 × 4.6, the highest value of electronic energy were optimized as
Varian USA) column was used. The column was thermostatted transition states at the same level by means of the eigenvector
at 313 K. The mobile phase was composed of water/methanol/ following method. Transition states and minima were
acetic acid (55:45:0.1 v/v) and used at flow rate of 1.0 mL/min. subsequently confirmed by performing vibrational analysis.
20 μL sample volumes were injected into the column, and Frequencies were calculated numerically at the same level of
elution was monitored at 240 nm. Shimadzu LC solution theory. Molecules were visualized using Avogadro package.18
software (ver. 1.25) data processing software was used to record
and integrate the chromatograms.
Apparent Solubility Study of IND. To determine the
■ RESULTS AND DISCUSSION
Part A: Study of the Amorphous IND Obtained by the
solubility of crystalline and amorphous forms of indapamide Vitrification Method. Various methods for preparation of
in purified water, phosphate buffer pH 6.8 and 0.1 M pharmaceutical systems in their amorphous forms have been
hydrochloric acid, about 100 mg of each sample was added reported, such as freeze-drying, spray drying or mechanical
to 50 mL of each medium. The resulting slurry was strirred for milling, however the simplest approach is based on the rapid
24 h at 25 °C ± 0.5 °C (I series) and 37 °C ± 0.5 °C (II cooling of the molten sample, typically referred as vitrification.
series). The samples were filtered through a sterile syringe filter However, this method carries a risk associated with chemical
(25 mm, PET-polyester, pore size 0.45 μm) from Macherey- decomposition of the drug during melting of the crystalline
Nagel (Bioanalytic, Poland). Each filtrate was diluted 1:200 v/v material. It should be emphasized that if thermal degradation
(0.25 mL of the filtrate solution was made up to 50 mL in a occurs, the decomposed amorphous material obtained is
volumetric flask). In order to prevent precipitation of disqualified as a therapeutic agent. This is the reason why
indapamide from the saturated solutions, all the flasks were chemical purity of drugs subjected to thermal stress conditions
heated to 25 or 37 °C. in a manufacturing process should be addressed with particular
Apparent solubility of the various forms of indapamide was care.
determined by HPLC (n = 9) as described above. The Chemical Stability Analysis. To determine the melting
concentration of indapamide was calculated from a rectilinear point of IND, differential scanning calorimetry was applied.
regression equation obtained from measurements of 10 standard The DSC curve obtained during heating of the crystalline
solutions in the concentration range of 0.025−2.0 μg/mL. The compound up to 473 K is depicted in Figure 2. An endothermic
calibration curve, y = 72615x − 355.8 (R2 = 0.9999), was peak, with an onset at 437 K, indicates melting of the sample.
determined from the relationship between the peak area of the The melting point (Tm) value of IND obtained in this study is
chromatograms and the concentration of standard solutions. The in good agreement with that determined recently by Ghugare
retention time of indapamide was 6.5 min. et al.16 Since the commercial form of the examined compound
Nonsink Dissolution Studies. Nonsink dissolution studies contains 2.5% water, in the thermogram presented in Figure 2
were performed in deionized water at 37 °C. 50 mg of the one can also observe a broad endotherm related to evaporation
tested material was placed in a glass vial (diameter, 22 mm; of water. To produce an amorphous form of the drug,
height, 70 mm) containing 25 mL of the medium and stirred indapamide was molten and then cooled to room temperature.
(Stuart SD162, U.K.) using a magnetic bar (diameter, 4 mm; The subsequent heating from the glassy state shows the
length, 12 mm) at a rate of 1300 rpm. The vial was placed in a characteristic signature for the glass transition in the temper-
double walled jacketed beaker connected to a water bath ature dependence of the heat flow at 377 K. Moreover, there
(Haake F3, Germany). Sampling of the liquid medium was was no thermal effect associated with evaporation of water,
carried out using a circulation system composed of an LSMatec indicating that indapamide obtained by vitrification is
peristaltic pump fitted with a 10 μm filter (HMWPE, Porex anhydrous. To confirm the disordered nature of the amorphous
Technologies, Germany). The filtered solution was passed form of IND, the XRD technique, being one of the most
through a 2 mm flow-through UV quartz cuvette placed in a definitive methods for detecting and quantifying molecular
Shimadzu UV-1700 PharmaSpec spectrophotometer (Japan) order in a system, was applied. As illustrated in Figure 3, the
set to 279 nm. The results are an average of three measure- very broad halo presented as a black line, compared to the
ments, and error bars present standard deviations. For statistical sharp peaks typical of the crystalline state (blue line), indicates
comparisons, one-way analysis of variance (ANOVA) followed that the vitrified sample is indeed an amorphous material.
by the post hoc Tukey’s test was used. Differences were It should be noted that the amorphous indapamide obtained
considered significant at p < 0.05.


by means of vitrification becomes yellow in contrast to the
white crystalline form. Such observation might indicate that the
CALCULATION METHODS examined drug undergoes thermal decomposition upon heating.
All quantum chemistry calculations for indapamide (IND) To verify this supposition FT-IR was employed, which is able to
molecule were performed with the use of density functional detect chemical changes occurring in a sample. First, we have
theory in the ORCA package.17 In the first step we have performed the analysis of crystalline IND as a reference. As
performed geometry optimizations of a dozen random IND illustrated in Figure 4, in the FT-IR spectrum of the crystalline
structures with the use of hybrid B3LYP functional and the compound one can distinguish several characteristic regions.
6-31g* basis set. Three the most stable structures were then The first one is located between 3700 and 3300 cm−1. There are
reoptimized using the same functional and 6-31++g(2d,2p) two sharp absorption bands associated with the symmetric
basis set. The structure presented is in the minimum (3499 cm−1) and asymmetric (3655 cm−1) stretching vibrations
determined by the position of the sulfonamide (−SO2NH2) for the hydroxy functional group of water molecules observed
group, which is the most flexible part of the molecule. In order in the commercial IND sample. The sulfonamide group has
to simulate the rotation at the B3LYP/6-31+g(2d,2p) level of strong bands due to its NH2 stretching vibrations at 3431 and
3615 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

In the next step, FT-IR spectra were recorded in a wide


temperature range, from 293 up to 453 K. As one can see in
Figure 4, with increasing temperature, broadening of bands and
the loss of spectral resolution are observed. Both of these effects
are typical for the melting process. Furthermore, there are a few
distinct differences in FT-IR spectra between the crystalline and
molten samples. The most significant changes are observed in
the 3700−3200 cm−1 region where two well resolved bands
assigned to vibrations of water molecules are detected i.e. at
3655 and 3499 cm−1. As illustrated in Figure 4, their intensity
decreases with increasing temperature, and finally, close to the
melting point, the bands are undetectable. It means that the
amount of water in the sample decreases with increasing
temperature. To better illustrate this process, the integral intensity
of two water bands was calculated and plotted versus temperature
(Figure 5). From Figure 5 it can be clearly seen that in fact the
whole water evaporates below 373 K. This result is in good
agreement with calorimetric measurements presented previously.
The other noticeable change in the temperature dependent IR
spectra, presented in Figure 4, is associated with the N−H band
at 3312 cm−1 of the crystalline sample. As it can be seen, with
increasing temperature the band becomes less intense and
disappears at 453 K. Simultaneously with rearrangement of the
N−H environment, the stretching band of the carbonyl moiety
shifts toward higher wavenumbers and a new band at 1458 cm−1 is
found to appear. According to the theoretical calculation of IND
Figure 3. X-ray diffraction patterns for various solid-state forms of
IND performed at room temperature T = 298 K.
spectrum, this new band is attributed to the bending C−OH
vibrations. All these changes indicate that the increase of
temperature induces the proton transfer from the nitrogen to
the oxygen atom of the carbonyl group. Consequently, the double
bond formerly placed between carbon and oxygen atoms shifts to
the carbon−nitrogen position (νCN at 1662 cm−1). It means
that indapamide molecule may exist in equilibrium with its two
easily interconvertible constitutional isomers (see Figure 1).
Taking into account the types of bonds involved in this proton
transfer, the transition observed in IND may be called an amide−
imidic acid tautomerism.
Since the FT-IR spectra recorded showed only the
tautomeric transformation and did not reveal any decom-
position at Tm, in the next step of our experiments we have
annealed the examined sample at 453 K i.e. 12 K above its
melting point. The spectra collected after one and two hours of
heating are presented in Figure 4. It is clearly seen that there is
Figure 4. FT-IR spectra collected during annealing of IND from 293 no difference between the isothermally annealed samples.
to 453 K. As a reference the spectrum of the starting, crystalline Taking into account the previous IR results one can expect that
material recorded at room temperature is presented as a bold red line.
Dashed lines indicate bands undergoing changes during annealing.
the indapamide is a thermally stable compound. The thermal
The upper panel presents FT-IR spectra of IND heated to 16 K above decomposition of the examined material was also excluded by
the melting point for one and two hours, respectively. HPLC−MS and NMR measurements performed in this study
(a detailed NMR description of this sample is presented in the
3342 cm−1 in the same infrared region. Moreover, a peak at Experimental Section). The above findings are additionally
3312 cm−1 of NH associated with the amide moiety of the supported by the thermogravimetric analysis (TGA) presented
indapamide molecule is observed. The second characteristic in a previous study.16 The TGA pattern indicates that the onset
region in the FT-IR spectrum is observed between 1700 and of thermal decomposition of IND, related to the onset of the
1000 cm−1. One of the strongest bands visible in this range is mass loss of IND, begins just above 560 K, well above the
located at 1655 cm−1 and arises from the carbonyl stretching melting point temperature. Consequently, one can conclude
vibration. Other band assignments with their wavenumber that it is possible to prepare the pure amorphous IND by
positions for crystalline IND phase are summarized in Table 1. means of vitrification. At the same time the change in color of
It should also be noted that infrared spectroscopy has recently IND drug during melting cannot be associated with chemical
been applied by Ghugare et al. to analyze the crystalline form of decomposition of the sample. It is worth noting that dissolution
indapamide.16 Despite the fact that the authors16 dispersed of indapamide in DMSO or ethanol at room temperature also
analyzed diuretic drug in KBr, the published diffuse reflectance results in the change of solution color from colorless to yellow.
infrared Fourier transform spectrum of IND is in good agree- Thus, the peculiar behavior of IND changing color may be an
ment with our results. effect of IND transition from crystalline to the liquid phase. It is
3616 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

Table 1. FT-IR Band Assignments with Their Wavenumber Positions for Crystalline and Glassy IND Samplesa
crystalline IND amorphous IND
vibration wavenumber [cm−1] vibration wavenumber [cm−1]
νasym H2O 3655 νOH 3604
νsym H2O 3499
νasym NH2 3431 νasymNH2 3387
νsym NH2 3342 νsymNH2 3277
νNH 3312
νCN; νOH 1662
νCO; νNH 1655
νasymCC(benzene); νCH 1611 νasymCC(benzene); νCH 1607
δNH2 1563 δNH2 1561
βCH; βNH 1460 βCH; βCOH 1458
βCH3; γNH2; νasym(SO)2; βCH 1323
νasymCC at Cl; γNH2; νasym(SO)2 1297 γNH2; νasym(SO)2; βCH 1295
νNN; βNH; βCH 1281
βNH; βCH 1242 βCOH; βCH; βCH2 1242
γNH2; νsymCCN(pyrrole ring) 1067 γCH2; γNH2; γCH3; γNH2; νsym(SO)2 1090
ωNH2; νNS 847 ωNH2; νNS; βCOH; νNC 841
ωNH2; νNS; βCOH; νCH 800
a
δ, twisting; ω, wagging; γ, rocking; β, bending; ν, stretching; asym, asymmetric; sym, symmetric vibrations.

well as theoretical studies of the tautomerization reaction.21−25


This is due to likely serious consequences of the proton transfer.
Especially, isomerization of pharmaceuticals becomes an
important problem as various tautomers of one drug may reveal
different pharmacological activity. A frequently used antibiotic,
erythromycin, exhibits as many as three isomers,26 however only
one of them, the ketone form, reveals the antibacterial activity.
Moreover, it cannot be excluded that tautomers with no desired
therapeutic activity may cause a number of side effects.
Therefore, the following question arises: Is the activity of the
imidic acid form of IND the same as that of the amide isomer?
Unfortunately, to answer this question thorough bioactivity
studies are required. However, one cannot exclude that the weak
diuretic activity of indapamide drug is due to the tautomeric
forms that appear after dissolving the tablet in the gastrointestinal
tract.
Figure 5. The temperature dependence of integral intensity for two Analysis of Molecular Dynamics in Supercooled and
bands, at 3655 and 3499 cm−1, respectively, recorded during heating of Glassy State of IND. In the literature one can find many
crystalline IND from room temperature to 453 K. examples that demonstrate that studies of the molecular
mobility in amorphous pharmaceuticals help to control the
noteworthy that there are several compounds that change their behavior of these systems at the macroscopic scale.27−30 The
color after melting or dissolution. As an example one can most frequently used experimental methods for monitoring the
mention indomethacin19 or piroxicam,20 which like indapamide molecular dynamics of amorphous medicines are dielectric
become yellow after amorphization. However, it should be spectroscopy, depolarized light scattering, dynamic mechanical
emphasized that in the latter case the tautomerization reaction analysis, proton correlation spectroscopy or nuclear magnetic
was found to be responsible for changing the color of sample. resonance. Here, to provide the essential information about
Since indapamide also reveals the isomerization tendency, one the molecular mobility in the supercooled and glassy states of
cannot exclude that the yellow color of the sample is due to the indapamide, broadband dielectric spectroscopy (BDS) was
tautomerisation, which occurs during melting. applied. Additionally, temperature-modulated differential scan-
To provide additional information about the tautomerization ning calorimetry (TMDSC) was implemented to analyze the
reaction of amorphous IND, FT-IR spectroscopy was employed dynamic behavior of IND as well as to determine its glass
once again. The infrared spectrum of vitrified indapamide transition temperature.
sample is depicted as a top green line in Figure 6. Since this The imaginary part of complex dielectric permittivity spectra
spectrum practically overlaps that which was recorded for the measured at ambient pressure over seven decades of frequency
molten sample i.e. there are bands corresponding to νCN in the supercooled liquid state and the amorphous phase of
and βC−OH vibrations, while the NH band associated with the IND are shown in Figure 7. Panel a of this plot presents
amide moiety of indapamide molecule is almost absent, one can dielectric loss spectra recorded above Tg: at temperatures from
conclude that the tautomeric equilibrium achieved at the 385 to 441 K with a uniform increment of 4 K. In this
melting temperature is conveyed into the amorphous material. In temperature range two dielectric processes can be distinguished:
recent years, there has been a growing interest in experimental as (i) conductivity relaxation at low frequencies which is related to
3617 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

Figure 6. FT-IR spectra of IND: crystalline, obtained by quench-cooling, obtained by cryogenic grinding and prepared by room temperature milling.
All spectra were collected at room temperature.

the α-relaxation peak toward lower frequencies is related to a


decrease of temperature. This behavior reflects the slowing down
of cooperative molecular motions of IND. Below Tg the
structural relaxation moves out of the frequency measurement
window, and then a secondary relaxation, called the γ-process,
becomes apparent (see Figure 7b). Similar to the mentioned
α-relaxation, the γ-mode also slows down during cooling, but it is
far less sensitive to the change of temperature. To move it over
6 decades in frequency, it was necessary to decrease the
temperature by 140 K, from 273 to 133 K. The nature of this
secondary mode will be discussed latter.
One of the most important parts of the dielectric spectra
analysis is to investigate the shape of the structural relaxation
process. From the pharmaceutical point of view, this parameter
is especially important as it was found to be directly related to
the physical stability of the amorphous material. As suggested
by Shamblin et al.,31 the stability of various amorphous drugs
should increase as the α-peak becomes narrower, i.e. with
increasing βKWW parameter. Examples of active substances
obeying this rule are e.g. nonivamide,32 telmisartan33 or
glibenclamide.34 To verify whether this relationship is valid
also in the case of indapamide, we first constructed a so-called
“masterplot” in which a number of dielectric curves measured at
different temperatures (both above and below Tg) have been
superimposed. As the reference one we have used the spectrum
recorded at T = 385 K. It can be clearly seen in Figure 8 that
the shape of the α-process is practically invariant upon cooling.
Thus, in the case of IND, similarly to many other glass-forming
liquids, the time−temperature superposition principle is
fulfilled. Additionally, it can be seen that the α-relaxation
Figure 7. Dielectric loss spectra of IND obtained by quench cooling of process of IND is asymmetric and broader than the classical
the melt. Panel a presents dielectric loss above the glass transition Debye response. To parametrize the shape of this mode we
temperature, whereas in panel b dielectric loss spectra collected below have applied the Kohlrausch−Williams−Watts (KWW)
Tg are presented.
function:35,36
the translational motion of ions and (ii) the structural relaxation ⎡ ⎛ ⎞ βKWW ⎤
t
process related to the cooperative rearrangements of indapamide ϕ(t ) = exp⎢ −⎜ ⎟ ⎥
molecules. In contrast to other pharmaceuticals, like glibencla- ⎢⎣ ⎝ τα ⎠ ⎥⎦
(1)
mide or indomethacin, the dc contribution for indapamide is not
significant, resulting in the maximum of structural relaxation peak where t is the time, τσ is the characteristic structural relaxation
being well resolved despite close proximity to the glass transition time and βKWW (0 < βKWW ≤ 1) denotes the stretching
region. As illustrated in the upper panel of Figure 7 the shift of parameter, which is related to the width of the relaxation peak.
3618 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

This procedure has been applied several times in the past to


predict the maximum of the hidden β-process. Using the values
of βKWW = 0.72 and the relaxation time determined from the
maximum of the α-peak recorded at 385 K, we have calculated
τ0 for indapamide. The τ0 was found to be equal to 2.2 × 10−4 s,
which corresponds to the frequency f 0 = 714 Hz. As illustrated in
Figure 8, the value of f 0 determined lies within the frequency
range where the excess wing is observed (green arrow).
In order to provide more information about the nature of
this excess wing visible in dielectric spectra of indapamide, we
have also performed an aging experiment. It is well-known that,
except for the high pressure measurements,48,49 this is one of
the most frequently used methods to explore more thoroughly
the origin of excess wing.50 If upon physical aging the excess
wing transforms into a well-separated peak (or shoulder), this
provides evidence that the excess wing may be classified as a
JG-type relaxation. During the aging experiment the examined
diuretic sample was annealed isothermally at 363 K, close to Tg.
Figure 8. Superimposed dielectric spectra of IND taken at ambient
pressure (p = 0.1 MPa), at seven different temperatures above and
Both the first and the last dielectric spectra (recorded at the
below Tg. The solid red line is the KWW function with βKWW = 0.72, beginning and after 22 h of annealing) are depicted in the inset
while the blue line is a Cole−Davidson fit (α = 1, β = 0.63). The green panel of Figure 8. As one can see, the excess wing becomes more
arrow denotes the position of the primitive relaxation calculated from pronounced with time. This experimental result suggests that the
eq 3 at 385 K. The inset panel presents the spectra collected at 363 K observed process may be attributed to the JG-relaxation.
during physical aging of IND . The dielectric α-relaxation peak of IND can be also analyzed
by means of the Havriliak−Negami function with the dc-
conductivity term added that fits the experimental data more
Since the above equation refers to the time not the frequency precisely than the KWW function.51
domain, to describe the ε″( f) peaks the one-sided Fourier
transform of the KWW function has to be applied: σdc Δε
ε″ (ϖ ) = + ε∞ +
∞⎛ dΦ ⎟⎞ ε0 ϖ [1 + (iωτHN)αHN ]βHN (4)
ε″(ϖ) = Δε ∫0 ⎜


dt ⎠
sin(ϖt ) dt
(2)
where ε∞ is the limiting high-frequency permittivity, τHN
where Δε is the dielectric strength of the α-relaxation. The denotes a characteristic relaxation time; exponents αHN and
representative fit of the KWW function to the experimental βHN characterize symmetric and asymmetric broadenings of
data, with the βKWW exponent equal to 0.72, is shown as a red line the dielectric loss curve; Δε is relaxation strength and ω is an
in Figure 8. Since the obtained value of βKWW is quite high, one angular frequency. A representative fit of eq 4 to the dielectric loss
can suppose that amorphous indapamide should not crystallize. curve recorded above Tg is depicted as a blue line in Figure 8. To
This assumption is supported by experimental data because parametrize the dielectric loss spectra recorded above Tg, the HN
neither dielectric nor calorimetric measurements reveal crystal- function with αHN =1, which is usually called the Cole−Davidson
lization signs during heating of the amorphous IND sample. (CD) function,52 was applied. As one can see, the CD equation
From the dielectric spectra analysis presented in Figure 8 it is describes the α-peak better than the previously used KWW
also evident that the KWW function describes the data well function. However, to obtain the perfect fit in whole measured
only in the vicinity of the α-peak maximum. It can be easily range, the effect of the unresolved β-mode should be taken into
seen that at frequencies about two decades higher than the account. Consequently, the superposition of the Cole−Davidson
maximum of the α-process the experimental data systematically and Cole−Cole function has to be applied. Herein, it should be
deviate from the KWW fit. The origin of such a trend, usually noted that the Cole−Cole function, i.e. when the parameter βHN is
called “excess wing”, has remained a matter of hot debate in the set to one, describes well symmetric and broad relaxation modes.
past decade.37−41 Some scientists interpreted this deviation as Thus, it can also be successfully used to characterize the secondary
an inherent part of the α-relaxation42 while others said that the γ-relaxation process of IND.
excess wing is a high frequency flank of the secondary relaxation In the next step of the dielectric data analysis, based on the
process hidden under the dominant α-peak.43,44 According to CC and CD fit parameters determined previously, the character-
the second explanation, for the first time postulated by Ngai45 istic relaxation times of both modes, α and γ, observed in
and experimentally verified by Lunkenheimer46 and other permittivity spectra of IND were calculated according to eq 5:53
researchers, this unresolved secondary mode originates from
some local motions of the entire molecule, and it is believed to ⎡ ⎛ α ·π ⎞⎤−1/ αHN ⎡ ⎛ α ·β ·π ⎞⎤1/ αHN
be a precursor of α-relaxation. To determine the maximum of τ = τHN⎢sin⎜⎜ HN ⎟⎟⎥ ⎢sin⎜⎜ HN HN ⎟⎟⎥
this process, usually called the Johari−Goldstein relaxation (JG), ⎢⎣ ⎝ 2 + 2βHN ⎠⎥⎦ ⎢⎣ ⎝ 2 + 2βHN ⎠⎥⎦
the coupling model (CM) approach is generally applied,47
(5)
τ0 = (tc)1 − βKWW (τα)βKWW (3)
The experimentally determined variations in log τα and log τγ with
where τ0, usually called the primitive relaxation time, denotes the temperature are depicted as solid squares and circles, respectively,
possible relaxation time of the hidden secondary mode while the in the form of a so-called Arrhenius plot shown in Figure 9.
parameters τα and βKWW characterize the α-relaxation, and tc = 2 ps. As illustrated in this figure the log τα(1/T) dependence can be
3619 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

Table 2. The Parameters of Fitting to the VFT and


Arrhenius Equations for Anhydrous and Hydrated Lidocaine
HCl Samples
sample vitrified IND cryomilled IND ball-milled IND
α-Process VFT Parameters
Tg (τ = 100 s) [K] 374.6
mP (τ = 100 s) 76 ± 2
log τ0 −16.9 ± 0.2
D 13.6 ± 0.6
T0 [K] 285 ± 2
υ-Process Arrhenius Dependence
log τ∞−υ −15.05 ± 0.06 −15.45 ± 0.13
Ea−υ [kJ/mol] 52.4 ± 0.3 54 ± 0.6
γ-Process Arrhenius Dependence
log τ∞−γ −13.33 ± 0.5
Ea−γ [kJ/mol] 32.9 ± 0.2

Figure 9. Relaxations map of vitrified IND. Temperature dependence give m = 76. This value is close to those found for other recently
of structural and secondary γ-relaxation relaxation times are depicted investigated amorphous pharmaceuticals such as verapamil
as solid squares and circles, respectively. Solid lines are VFT and hydrochloride (m = 88),10 telmisartan (m = 87),33 glibenclamide
Arrhenius fits to the experimental data. Fitting parameters are (m = 78)34 or indomethacin (m = 83).19 Thus, indapamide
presented in Table 2. Additionally, solid triangles are the structural investigated in this work can be classified as an intermediate glass-
relaxation times determined from TMDSC measurements for
former drug.
anhydrous IND.
Now we consider the molecular mobility in the glassy state of
anhydrous indapamide that is reflected in secondary relaxation
reasonably well described over the entire measured range by processes. As one can see in Figure 7, except for the excess
means of the Vogel−Fulcher−Tammann equation,54−56 wing, identified previously as a hidden β-process, in the
dielectric spectra of IND only one secondary relaxation (γ) can
⎛ DT0 ⎞ be found. Since log τγ is linearly related to the inverse of
τα = τ0 exp⎜ ⎟
⎝ T − T0 ⎠ (6) temperature, it is possible to fit the Arrhenius equation (eq 8)
to the experimental data and determine the activation energy
From the fit of the VFT equation to τα(T) dependence one can barrier of this mode.
easy determine the glass transition temperature of IND. Applying
⎛E ⎞
the most frequently used definition, which describes Tg as the τγ = τ∞ exp⎜ a ⎟
temperature at which the dielectric relaxation time τα reaches ⎝ RT ⎠ (8)
100 s, the glass transition temperature of IND was found to be The activation energy barrier was found to be equal to
equal to 374 K, and this value well agrees with the one found from 34 kJ/mol. This quite low value of Ea, together with the
DSC measurements (TgDSC = 377 K). Additionally, in this paper symmetrical shape of the γ-peak, suggests that the motions of a
we have determined the structural relaxation times of indapamide small part of indapamide molecule are responsible for this
from the temperature dependences of the real part of the complex secondary relaxation. In order to confirm this supposition we
heat capacity Cp′ measured by the stochastic temperature have performed the DFT calculations. Since the sulfonamide
modulated DSC technique. As illustrated in Figure 9, DSC data group seems to be the most interactive part of the indapamide
are in good agreement with the structural relaxation times molecule, we have studied the conformational changes of this
determined from the dielectric measurements (Table 2). moiety. Initially we have checked whether the dipole moment is
The most characteristic feature of molecular dynamics in changing during the conversion or not. It should be emphasized
vitrifying liquids is non-Arrhenius dependence of the τα(T) that the variation of the dipole moment is necessary to observe
curve. However, the degree of deviation from Arrhenius intramolecular motions as a secondary mode in the dielectric
behavior varies from one compound to another. To assess spectrum. As illustrated in Figure 10, the rotation of the
whether or not this divergence is significant, the “steepness sulfonamide group of indapamide causes a change in the dipole
index” or “fragility” defined by Bohmer et al.,57 moment. That is why this movement may be considered as an
d log τα origin of the γ-relaxation. The value of activation energy of the
m≡ sulfonamide group rotation obtained in the B3LYP/6-31+
d(Tg /T ) +g(2d,2p) model is equal to 25 kJ/mol, and it is slightly lower
T = Tg (7)
than the value of Ea determined experimentally.
is usually calculated. Using this parameter we can classify Theoretical Prediction and Experimental Verification
supercooled liquids into two types: (i) “strong”, if the of the Physical Stability of Quenched Indapamide.
temperature dependence of structural relaxation times is close Understanding of the factors affecting crystallization from the
to Arrhenius behavior (m ≤ 30), and (ii) “fragile”, if log τα(1/T) glassy state not only is important from a scientific perspective but
deviates significantly from the straight line (m ≥ 100). When m also has many practical applications e.g. in the pharmaceutical
falls within the 30 < m < 100 range, the liquid is classified as industry. However, in contrast to the chemical stability
intermediate glass-former. Fragility calculations performed for IND determination, where the standard analytical techniques are usually
3620 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

stable imidic acid forms are dispersed in the amide matrix. In


the literature one can find some reports that show that this kind
of equilibrium detected in the glassy state may strongly reduce
the tendency of such a system to crystallize. For example,
glibenclamide60 is a hypoglycemic agent with tautomerization
tendency or binary mixtures containing active pharmaceutical
substance and various excipients like polymers (PVP, PEG or
HPMC)61−63 or sugars (octaacetylmaltose)64 which are
characterized by highly physical stability in the amorphous
state. The above results indicate that factors influencing the
stability of an amorphous system should not be considered
separately, because only when studied together can they
provide full information about the physical stability of glassy
pharmaceuticals. Thus, in the context of molecular dynamics
investigations as well as the proton transfer in IND one can
Figure 10. Diagram representing changes of the dipole moment and suppose that the IND system examined should not reveal the
energy during rotation of the sulfonamide group of IND molecule. crystallization propensity from the amorphous state. To verify
these theoretical predictions, X-ray powder diffraction was used.
applied, there is no certain methodology to predict the physical XRD measurements performed after one year of storage of the
stability of an amorphous material. Until now crystallization of the sample at room temperature showed no sign of crystallization.
glassy sample has been many times correlated to the degree of its Part B: Study of Amorphous Indapamide Samples
molecular mobility, thermodynamic properties (such as heat of Obtained by Cryogenic Grinding and Room Temper-
fusion, heat capacity, configurational entropy) or the methods of ature Milling. Mechanical milling of crystalline solids is a very
amorphization. One of the parameters, considered as a key factor popular technique applied in many fields of industry to reduce
affecting stability of an amorphous system, is dynamic fragility. the particle size. It is also used in the pharmaceutical industry since
This is because the steepness index is related to an average degree micronization was proven to enhance the dissolution properties
of molecular mobility reflected in a structural relaxation near the of drugs through an increase in the surface area of micro- or
glass transition. According to the TOP (two order parameter) nanoparticulates. As shown in the literature, the mechanical
model proposed by Tanaka,58 near the glass transition region treatment may also lead to amorphization of crystalline materials.
competition between long-range density, being a driving force For instance, amorphization by milling was observed for
toward formation of the crystals, and short-range ordering favoring piroxicam,20 glibenclamide,34 sucrose or ziprasidone.65
formation of the locally preferred structures is postulated. If the Physicochemical Characterization of Milled IND
second effect dominates, the system becomes “frustrated”. Samples. Herein, to produce the amorphous form of IND
Consequently, the crystallization is expected to be inhibited. It drug we have applied two types of grinding: cryomilling, being
has been shown that the degree of frustration can be related to operated at very low temperatures using cryogenic media such
fragility, i.e. the lower frustrationthe greater fragilitythe as liquid nitrogen, and ball-milling that is carried out at room
greater tendency to crystallize. Taking into account this temperature. In the first case the IND sample (3 g) was milled
assumption, IND with the fragility index equal to 76 should be for 15, 30, and 45 min. On the other hand, the ball-milling
classified rather as a system with high crystallization tendency. On lasted for 1, 2.5, and 5 h. However, contrary to cryogrinding the
the other hand, as it was stated previously, we can distinguish only amount of milled sample was two times higher (6 g). The
one, well-pronounced secondary relaxation process of IND. It diffraction patterns of all micronized materials are presented in
means that the molecular mobility of the examined drug is Figure 3. Significant broadening and decrease of intensity of
significantly retarded below the liquid−glass transition, which is diffraction peaks attributed to reduction of sample crystallinity
observed at a relatively high temperature (377 K). Consequently, was observed with increasing milling time. After 15 min of
at room temperature conditions, i.e. almost 80 K below the Tg, cryogrinding and 1 h of ball-milling the degree of IND
motions of IND molecules are deeply frozen, which prevents crystallinity, calculated as a ratio of X-ray diffraction peak areas
crystallization of IND. Moreover, there is a general rule that says for the partially crystalline sample to that being 100%
that the long-term stability of amorphous drugs can be obtained by crystalline, was found to be equal to 34% and 36%, respectively.
storage at a temperature where molecular mobility associated with However, when the milling time was extended to 30 min and
the structural relaxation time approaches zero, i.e. at T0 in the 2.5 h, respectively, indapamide became almost completely
Vogel−Fulcher−Tamman (VFT) equation. This temperature was amorphous. The degrees of crystallinity were calculated as 5%
found to be close to the so-called Kauzmann temperature,59 where (cryomilled IND) and 8% (ball-milled IND). A complete
entropy of the supercooled liquid and crystal would be equal. In crystal−glass conversion was observed after 45 min of
the case of IND T0 is equal to 285 K. Thus, storage of IND at cryomilling while at room temperature 5 h of grinding was
such temperature conditions should guarantee its physical stability required to obtain fully amorphous IND. As illustrated in
for a typical shelf life. Figure 3, Bragg peaks of these samples, characteristic for
To better understand the physical stability of the examined materials with a long-range three-dimensional molecular order,
amorphous diuretic agent we have to consider other factors were replaced by broad halos identical to that recorded for
that may affect its crystallization. Specifically, the possibility molten-quenched indapamide. Interestingly, amorphous IND
of isomerization should be also taken into account. As it was powders obtained using the mechanical treatment were yellow,
mentioned in the previous part of this paper there are two just like the quenched material.
isomers coexisting in the glassy state of IND. Consequently, the The experimental data presented clearly indicate that cryomilling
amorphous sample becomes a binary mixture in which less is much more efficient in preparation of the glassy IND than
3621 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

grinding at room temperature. The time needed to fully amorphize by ball-milling. On the other hand, cryomilling was believed not to
crystalline IND by cryomilling was approximately seven times induce chemical changes in the milled material because liquid
shorter compared to with traditional room temperature milling. nitrogen prevents local overheating of the sample. However,
This result agrees with the general rule that a decrease of the milling Adrjanowicz et al.68 recently have shown that even cryomilling is
temperature brings similar effects as an increase in milling intensity. able to modify the chemical structure of ground materials.
Additionally, it is worth pointing out that mechanical amorphization Therefore, one can suppose that both room temperature milling
of crystalline IND is much more efficient than milling of other and cryogrinding can induce changes in the structure of
organic compounds reported in the literature. For example, full indapamide. To characterize the structure of cryomilled and ball-
amorphization of indomethacin19 and furosemide66 was achieved milled materials in detail, FT-IR spectroscopy was exploited. The
after 60 and 120 min of cryomilling, respectively, while in the case FT-IR spectra of milled IND are presented in Figure 6 together
of IND only 45 min was necessary. Moreover, it should be with the data obtained for the crystalline and vitrified samples. It is
stressed that in our experiments the amount of milled material was clearly visible that the data pertaining to all amorphous materials
three times higher than those for mentioned pharmaceutics. On are almost the same. The only difference in the high wavenumber
the other hand, the mechanical treatment of IND at room temp- region 3700−3200 cm−1 was the vibrations of water molecules
erature brings the complete amorphization only after 300 min, detected for the milled IND. This is because the milled samples,
while the amorphous state of indomethacin and trehalose was in contrast to the quenched IND, contain water. On the other
achieved after more than 1200 and 1380 min of traditional ball hand, similarly to the quenched indapamide, significant loss in
grinding. Hence, it can be stated that the mechanical treatment of intensity of the N−H stretching vibration at 3312 cm−1 was found.
IND at both room and liquid nitrogen temperatures works very Furthermore, new bands attributed to CN and C−OH
effectively. Thus, these methods can be successfully applied in the vibrations also appear in the spectra. These changes in the FT-IR
pharmaceutical industry to produce IND in the amorphous form. spectra indicate that milling of IND both at room temperature and
In the next step we have performed the DSC analysis of in cryogenic conditions leads to the conversion of the amide form
milled indapamide samples. In Figure 2 DSC heating scans of of drug to its imidic acid isomer.
cryomilled and ball-milled IND are illustrated. In both cases, two There are contradictory literature reports on the tautome-
thermal events were observed within the examined temperature rization reaction of milled materials. On the one hand, there are
range: the first one associated with water evaporation and the several examples when the less stable isomer appears after
second one ascribed to the glass transition. As seen in the inset grinding, like in the case of glibenclamide or piroxicam. In the
panel of Figure 2, a heat flow jump characteristic for a glass former case tautomerism is observed independently of the
transition detected in the thermograms of milled IND occurs grinding conditions while for the second compound the
exactly at the same temperature as that of the quenched liquid, i.e. transformation was detected after milling at the liquid nitrogen
377 K. However, in contrast to the anhydrous vitrified sample, the temperature. On the other hand, these findings are in contrast
cryomilled and ball-milled materials contain 1.3% and 2.5% water, with the experiments performed with sugars. Descamps et al.
respectively. The reduced water content in cryomilled IND is due have shown that mutarotation of saccharides, being a ring−
to the fact that before grinding indapamide hemihydrate has been chain tautomerization, does not occur in amorphous trehalose
dried at 363 K for 12 h. On the other hand, amorphous sample or lactose samples obtained by milling at liquid nitrogen
obtained by means of room temperature milling contains the same temperature.68 It means that after grinding there is only one,
amount of water as the starting hemihydrate. Nevertheless, in both initial anomeric form. However, it should be stressed that in the
cases the whole amount of water evaporates before the glass case of anhydrous lactose 96% of milled material was still in the
transition temperature is achieved. It is interesting that in the DSC α-anomeric form. Thus, one can suppose that after milling 4%
thermograms of the tested samples there is no exothermic peak of the investigated sample was converted into the β-form.
ascribed to so-called “cold crystallization”, which is frequently Molecular Mobility of Milled IND Samples. In Figure 11
observed during the slow heating of milled materials. This result we present the imaginary parts of dielectric permittivity spectra
suggests high stability of ground IND against crystallization. plotted as a function of frequency during heating of cryomilled
Indeed, X-ray diffraction measurements performed one year after and ball-milled IND from 153 K up to 443 K. To show the
sample preparation did not reveal any Bragg reflections. It means whole sets of data more clearly, dielectric spectra of each
that there is no difference in terms of the physical stability between sample were divided into two panels presenting the relaxation
amorphous IND samples prepared by milling and vitrification dynamics above and below the glass transition temperature. As
techniques. This result is interesting in the context of recent can be seen in panels c and d of Figure 11, the imaginary part of
reports of tri-o-methyl-β-cyclodextrin,66 nucleosides67 or gliben- the complex dielectric permittivity, recorded for both milled
clamide34 which demonstrate the opposite behavior. Thus, one materials, reveals a well-resolved secondary relaxation process,
can formulate the following question: What factors af fect the called υ-peak, which was not observed in dielectric spectra
stability of milled IND? As shown in part A of this paper, in the of the melt-quenched sample. Similar to the γ-relaxation, the
case of the vitrified sample the proton transfer was found to be υ-process shifts toward higher frequencies on heating, but in
one of the key factors determining the stability of quenched IND. contrast to the γ-mode it is more sensitive to temperature
Accordingly, a further important question to pose is: Does the changes. This behavior suggests that the activation energy of
milling induce the amide−imidic acid transformation of this υ-process is greater than that for the γ-relaxation. The close
indapamide or not? As shown in the literature, the high level of inspection of the dielectric loss spectra collected at the same
mechanical energy used during milling may cause significant temperature conditions (T = 216 K), for milled and quench-
changes in the structure of the ground material. These may be cooled samples, shows that the amplitude of the new dielectric
observed especially when the sample is milled above the liquid to mode is higher than that of the γ-relaxation (see Figure 12).
glass transition temperature. Such a scenario was shown for a Moreover, both processes, γ and υ, appear in the dielectric
number of pharmaceuticals such as chloramphenicol, cimetidine or spectrum at the same temperature range. This explains why the
indomethacin, where polymorphic transformations were induced γ-process becomes almost invisible in the dielectric spectra of
3622 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

Figure 11. Dielectric spectra of IND samples obtained by cryogenic grinding (panels a and c) and room temperature milling (panels b and d). The
dielectric data were collected above and below the glass transition temperature.

Figure 12. Comparison of dielectric spectra measured at the same


temperature T = 216 K for anhydrous IND obtained by vitrification
and hydrated samples prepared using cryomilling and ball milling.

the cryomilled and ball-milled IND samples. The only evidence


for the existence of this γ-mode is the poorly noticeable high Figure 13. Temperature dependence of α (diamond symbols), υ (solid
frequency flank of this process visible in Figure 11d. Therefore, and crossed circle symbols), γ (open symbols), relaxation times
the question arises: What is the molecular origin of the υ-mode? determined for anhydrous IND and both examined hydrated samples.
Solid lines indicate VFT and Arrhenius fits to the α-relaxation times
As it was mentioned in the previous part of this paper the only and secondary γ-, and υ-relaxation times, respectively.
difference between the milled and vitrified IND samples is the
water content. While the quench-cooled sample is anhydrous, containing IND samples are exactly the same at 216 K. Since the
the cryomilled and ball-milled materials contain small amounts relaxation times of the υ-process plotted as an inverse of
of water. Thus, one can suppose that the υ-process is related to temperature exhibit the Arrhenius behavior from fitting of eq 8
the mobility of water molecules present in milled IND. In order to the experimental data, the activation energy of the υ-mode was
to characterize the υ-modes observed for both water-containing determined to be 52.4 ± 0.3 kJ/mol. It is worth noting that this
IND materials, the υ-relaxation times have been calculated as value is in good agreement with the activation energy of water
the inverse of the frequency of the maximum peak position, i.e. relaxation found for other water mixtures such as e.g. DPG + H2O
τ = 1/2πf max. As clearly seen in Figure 13 the temperature (43 kJ/mol),69 or hydrated lidocaine HCl (49 kJ/mol).70 Moreover
dependences of log τυ, determined for the amorphous samples it is also close to the value of Ea for supercooled water examined in
obtained by various mechanical treatments, overlay each other. some confined water systems (Ea ≈ 44.4 ± 3.7 kJ/mol).71
This result is not surprising in the light of the data displayed in As it was demonstrated above, the effect of water on the
Figure 12, where the maximum of υ-modes for both water- molecular dynamics of IND in the glassy state is significant.
3623 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics Article

Table 3. Apparent Solubility of Crystalline and Amorphous Forms of IND Obtained by Vitrification, Ball Milling and
Cryomilling Determined after 24 h of the Experiment
solubility of indapamide mean value n = 9
amorphous form
crystalline form vitrified ball milled cryomilled
media temp [°C] mg/L RSD [%] mg/L RSD [%] mg/L RSD [%] mg/L RSD [%]
H2O 25 81 2.55 96 2.50 107 2.30 100 2.60
37 109 2.40 125 1.90 144 0.70 144 1.38
0.1 M HCl 25 88 2.63 102 2.50 125 2.20 122 3.10
37 106 2.25 140 2.50 149 1.00 151 1.23
phosphate buffer pH 6.8 25 81 2.92 91 2.10 103 2.70 96 2.30
37 106 2.57 120 2.10 136 1.50 140 0.40

Thus, the question is: How does water af fect the structural
relaxation process and consequently the values of glass transition
temperature and f ragility of the examined systems? In the
literature one can find abundance of experimental data showing
that water is a common plasticizer and an increase of water
content leads to a drop of Tg.68 To verify this statement for
IND, we analyzed the dielectric spectra of cryomilled and room
temperature milled samples collected above Tg. As illustrated in
panels a and b of Figure 11, in the supercooled region of the
milled materials the well resolved structural relaxation process
is observed. Moreover, in both cases the maximum of the α-
peak appears in our frequency window at the same temperature
conditions as for the quenched IND sample. Additionally,
comparison of the shape and position of the structural
relaxation processes collected at the same temperature for all
examined materials reveals only one difference between
dielectric spectra. Namely, the contribution of the dc Figure 14. Dynamic solubility studies of crystalline and amorphous
conductivity is more pronounced in the case of ball-milled (vitrified) indapamide (n = 3).
IND than for the melt-quenched and cryomilled samples. Thus,
one can expect the same temperature dependences of τα for all greater concentration of the drug in comparison to that of the
examined IND materials. Indeed, as depicted in Figure 13, crystalline material (Figure 14). Following the supersaturation
regardless of the amorphization method, the α-process detected peak, concentration of indapamide decreased to 129.4 ± 4.3 mg/L
in the dielectric loss spectra of IND exhibits the same pattern of at 10 min of the experiment. The drop in the drug concentration
behavior. Consequently, the values of the glass transition was related to solution mediated crystallization of the amorphous
temperature and dynamic fragility of IND do not depend on phase to indapamide hemihydrate, however the drug levels in
the applied amorphization technique. This result is consistent solution remained significantly higher (p < 0.05) than the “steady-
with the DSC thermograms, which show the characteristic state” concentration for crystalline indapamide hemihydrate. From
signature for the Tg in the heat flow with an onset at 377 K for Table 3 it can be seen that irrespective of the amorphization
each examined IND sample (see inset in Figure 2). technique, in each medium, the apparent solubility of amorphous
Part C: Solubility Advantage from an Amorphous IND IND, determined at 24 h of the experiment, is 13% to 42% greater
Drug. As a final point we would like to present results of than that of the crystalline form. The greatest value (151 mg/L i.e.
apparent solubility measurements of crystalline and all examined 42% higher than for crystalline IND) was obtained for the
amorphous forms of indapamide. The obtained values are cryomilled sample in 0.1 M hydrochloric acid at 37 °C.
collected in Table 3. As a reference, crystalline IND was examined. Additionally, it was found that there was only a slight difference
Crystalline indapamide subjected to nonsink dissolution studies at between apparent solubility of the cryomilled and ball-milled
37 °C dissolved rapidly to a concentration of 106.8 ± 12.1 mg/L samples. The solubility of IND samples was determined with high
within the first minute of the experiment. Solubility of crystalline precision. The values of the relative standard deviation (RSD),
indapamide reached a plateau at 109.2 ± 9.6 mg/L within 30 min. calculated on the basis of nine replicates for each sample, were in
the range of 0.4−3.1%.


PXRD analysis confirmed that the solid residue recovered at the
end of the experiment was indapamide hemihydrate, indicating
that this crystalline form was physically stable and did not convert CONCLUSIONS
in water to other forms of indapamide. Additionally, the solubility From our studies on amorphous IND we can draw the
of crystalline IND in water was found to be equal to 81 mg/L following conclusions:
when the experiment temperature was decreased to 25 °C. These 1. Quench-cooling of the melt, cryogenic grinding and
results are in good agreement with the literature data (75 mg/L).72 room temperature milling can be successfully applied to
On the other hand, amorphous indapamide, obtained using produce IND in the amorphous state. Especially
vitrification, reached a concentration of 394.6 ± 24.0 mg/L within important for commercial application seems to be
the first two minutes of the experiment, resulting in over 3-fold cryogenic milling, which was found to be very effective.
3624 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627
Molecular Pharmaceutics


Article

NMR and HPLC measurements combined with mass ACKNOWLEDGMENTS


spectrometry have shown that all examined amorphous The authors Z.W., K.G., M.P. and W.S. are deeply grateful for
samples were of purity satisfying the strict pharmaceutical
the financial support by the National Science Centre within the
requirements.
framework of the Opus3 project (Grant No. DEC-2012/05/B/
2. Investigations by means of FT-IR in a wide temperature range
1127NZ3/03233). L.T. and K.J.P. wish to acknowledge funding
have shown that conversion to the glassy state of IND leads to
an amide−imidic acid tautomerism, which is a reversible for this research from Solid State Pharmaceutical Cluster
proton transfer between different locations in the molecule. (SSPC), supported by Science Foundation Ireland under Grant
Interestingly, the less stable imidic acid form of IND appears No. 07/ 1130SRC/B1158. Z.W. acknowledges the financial
independent of the applied amorphization technique. assistance from FNP START (2013). M.D. is thankful for
3. Using BDS spectroscopy, we have studied the molecular support from PL-Grid Infrastructure.
dynamics of vitrified and milled IND in the supercooled
liquid and glassy states. In all cases, at a given temperature, the
position as well as shape of the structural relaxation process
■ REFERENCES
(1) Legorburu, M. J.; Alonso, R. M.; Jiménez, R. M.; Ortiz, E.
was found to be the same (the only difference between the Quantitative Determination of Indapamide in Pharmaceuticals and
spectra was the higher contribution of dc conductivity in the Urine by High-Performance Liquid Chromatography with Ampero-
case of the ball-milled sample). Thus, there was also no metric Detection. J. Chromatogr. Sci. 1999, 37 (8), 283−287.
difference between the temperature dependence of structural (2) http://www.drugbank.ca/drugs/DB00808
relaxation times determined for all examined samples. The (3) Wojnarowska, Z.; Roland, C. M.; Swiety-Pospiech, A.;
glass transition temperature calculated on the basis of BDS Grzybowska, K.; Paluch, M. Anomalous Electrical Conductivity
measurements was found to be equal to 374 K (τα = 100 s) Behavior at Elevated Pressure in the Protic Ionic Liquid Procainamide
for the vitrified and milled materials, and it is in good Hydrochloride. Phys. Rev. Lett. 2012, 108, 015701.
agreement with the value determined from calorimetric (4) Wojnarowska, Z.; Roland, C. M.; Kolodziejczyk, K.; Swiety-
Pospiech, A.; Grzybowska, K.; Paluch, M. Quantifying the Structural
measurements (TgDSC = 377 K).
Dynamics of Pharmaceuticals in the Glassy State. J. Phys. Chem. Lett.
On the other hand, the molecular dynamics of milled 2012, 3 (10), 1238−1241.
and melt-quench samples below Tg differ from each other (5) Bolten, D.; Türk, M. Micronisation of carbamazepine through
significantly. In all cases, the dielectric measurements reveal rapid expansion of supercritical solution (RESS). J. Supercrit. Fluids
only one, well-pronounced secondary relaxation process, 2012, 62, 32−40.
however its molecular origin is completely different. In the (6) Gupta, P.; Kakumanu, V. K.; Bansal, A. K. Stability and Solubility
case of the vitrified sample it was labeled as γ-mode. This of Celecoxib-PVP Amorphous Dispersions: A Molecular Perspective.
relaxation process with a low value of the activation energy Pharm Res. 2004, 21 (10), 1762−1769.
(34 kJ/mol) was found to be due to rotation of the (7) Descamps, M.; Willart, J. F.; Dudognon, E.; Caron, V.
sulfonamide group. Also, for milled samples the υ-relaxation Transformation of pharmaceutical compounds upon milling and
with Ea = 54 kJ/mol, related to mobility of water molecules, comilling: The role of Tg. J. Pharm. Sci. 2007, 96, 1398−1407.
was observed. Additionally, for each examined sample, a (8) Hancock, B. C.; Parks, M. What is the True Solubility Advantage
slower secondary mode (β), identified as a hidden JG for Amorphous Pharmaceuticals? Pharm. Res. 2000, 17 (4), 397−404.
relaxation, which originates from intermolecular interac- (9) Murdande, S. B.; Pikal, M. J.; Shanker, R. M.; Bogner, R. H.
tions, was visible in the dielectric spectra. Solubility advantage of amorphous pharmaceuticals: I. A thermody-
namic analysis. J. Pharm. Sci. 2010, 99 (3), 1254−1264.
4. It has been presented, using BDS, DSC and XRD
(10) Adrjanowicz, K.; Kaminski, K.; Paluch, M.; Wlodarczyk, P.;
techniques, that amorphous indapamide prepared by Grzybowska, K.; Wojnarowska, Z.; Hawelek, L.; Sawicki, W.; Lepek,
quench-cooling, cryogrinding and ball grinding is a P.; Lunio, R. Dielectric relaxation studies and dissolution behavior of
physically stable system in a wide temperature range, amorphous verapamil hydrochloride. J. Pharm. Sci. 2010, 99, 828.
above as well as below Tg. We did not notice any signs of (11) Murdande, S.; Pikal, M.; Shanker, R.; Bogner, R. Solubility
crystallization even one year after amorphization (storage Advantage of Amorphous Pharmaceuticals: II. Application of
conditions: T = 293 K, humidity 10%). Quantitative Thermodynamic Relationships for Prediction of
5. Amorphous IND crystallizes in aqueous media to indapamide Solubility Enhancement in Structurally Diverse Insoluble Pharmaceut-
hemihydrate. The supersaturation concentration of amor- icals. Pharm. Res. 2010, 27 (12), 2704−2714.
phous IND undergoing liquid-mediated crystallization at 37 (12) Grzybowska, K.; Paluch, M.; Grzybowski, A.; Wojnarowska, Z.;
°C in water was found to be nearly four times greater than the Hawelek, L.; Kolodziejczyk, K.; Ngai, K. Molecular Dynamics and
equilibrium solubility of crystalline IND. It was found that, Physical Stability of Amorphous Anti-Inflammatory Drug: Celecoxib. J.
irrespective of the amorphization technique, the apparent Phys. Chem. B 2010, 114 (40), 12792−12801.
(13) Sawicki, W.; Lepek, P.; Wlodarski, K.; Wojnarowska, Z.; Paluch,
solubility of amorphous indapamide was greater by 13−42%
M.; Guzik, L. Effect of amorphization method on telmisartan solubility
than that of the crystalline form. The greatest value (151 mg/ and the tableting process. Eur. J. Pharm. Biopharm. 2013, 83, 114−121.
L i.e. 42% greater than for crystalline IND) was obtained for (14) Luthra, S. A.; Shalaev, E. Y.; Medec, A.; Hong, J.; Pikal, M. J.
cryomilled sample in 0.1 M hydrochloric acid at 37 °C at 24 h Chemical stability of amorphous materials: Specific and general media
of the apparent solubility studies. effects in the role of water in the degradation of freeze-dried

■ AUTHOR INFORMATION
Corresponding Author
zoniporide. J. Pharm. Sci. 2012, 101 (9), 3110−3123.
(15) Smith, M. B.; March, J. Advanced Organic Chemistry, 5th ed;
Wiley Interscience: New York, 2001; pp 1218−1223.
(16) Ghugare, P.; Dongre, V.; Karmuse, P.; Rana, R.; Singh, D.;
*E-mail: zaneta.wojnarowska@us.edu.pl. Kumar, A.; Filmwala, Z. Solid state investigation and characterisation
Notes of the polymorphic and pseudopollymorphic forms of indapamide. J.
The authors declare no competing financial interest. Pharm. Biomed. Anal. 2010, 51 (3), 532−540.

3625 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627


Molecular Pharmaceutics Article

(17) Neese, F., ORCAan ab initio density functional and Quenched and Cryomilled Material. Mol. Pharmaceutics 2010, 7,
semiempirical program package, Version 2.8; University of Bonn: 1692−1707.
Bonn, 2006. (35) Kohlrausch, R. Nachtrag uber die elastiche Nachwirkung beim
(18) Avogadro: an open-source molecular builder and visualization tool, Cocon und Glasladen. Ann. Phys. (Leipzig) 1847, 72, 393.
Version 1.0.3; http://avogadro.openmolecules.net/. (36) Williams, G.; Watts, D. C. Non-symmetrical dielectric relaxation
(19) Wojnarowska, Z.; Adrjanowicz, K.; Wlodarczyk, P.; Kaminska, behavior arising from a simple empirical decay function. Trans.
E.; Kaminski, K.; Grzybowska, K.; Wrzalik, R.; Paluch, M.; Ngai, K. L. Faraday Soc. 1970, 66, 80−85.
Broadband dielectric relaxation study at ambient and elevated pressure (37) Hensel-Bielowka, S.; Wlodarczyk, P.; Mierzwa, M.; Paluch, M.;
of molecular dynamics of pharmaceutical: indomethacin. J. Phys. Chem. Ngai, K. L. Dynamics of α-Tetralone at Elevated Pressure and in
B 2009, 113, 12536. Mixture with Oligostyrene. J. Phys. Chem. B 2012, 116 (1), 22−29.
(20) Sheth, A. R.; Lubach, J. W.; Munson, E. J.; Muller, F. X.; Grant, (38) Kaminska, E.; Kaminski, K.; Hensel-Bielowka, S.; Paluch, M.;
D. J. W. Mechanochromism of Piroxicam Accompanied by Ngai, K. L. Characterization and identification of the nature of two
Intermolecular Proton Transfer Probed by Spectroscopic Methods different kinds of secondary relaxation in one glass-former. J. Non-
and Solid-Phase Changes. J. Am. Chem. Soc. 2005, 127 (18), 6641− Cryst. Solids 2006, 352 (42−49), 4672−4678.
6651. (39) Hensel-Bielowka, S.; Paluch, M. Origin of the High-Frequency
(21) Colominas, C.; Luque, F. J.; Orozco, M. Tautomerism and Contributions to the Dielectric Loss in Supercooled Liquids. Phys. Rev.
Protonation of Guanine and Cytosine. Implications in the Formation Lett. 2002, 89, 025704.
of Hydrogen-Bonded Complexes. J. Am. Chem. Soc. 1996, 118 (29), (40) Kaminski, K.; Kaminska, E.; Hensel-Bielowka, S.; Pawlus, S.;
6811−6821. Paluch, M.; Ziolo, J. High pressure study on molecular mobility of
(22) Wojnarowska, Z.; Wang, Y.; Sokolov, A. P.; Paluch, M. leucrose. J. Chem. Phys. 2008, 129, 084501.
Rheological studies of tautomerization kinetics in supercooled (41) Paluch, M.; Ziolo, J.; Rzoska, S. J.; Habdas, P. High-pressure and
glibenclamide drug. Phys. Rev. E 2012, 86, 067104. temperature dependence of dielectric relaxation in supercooled di-
(23) Wojnarowska, Z.; Paluch, M.; Pionteck, J. The tautomerization isobutyl phthalate. Phys. Rev E 1996, 54 (4), 4008−4010.
phenomenon of glibenclamide drug monitored by means of volumetric (42) Ngai, K. L.; Paluch, M. Classification of secondary relaxation in
measurements. J. Chem. Phys. 2011, 135, 214506. glass-formers based on dynamic properties. J. Chem. Phys. 2004, 122
(24) Wojnarowska, Z.; Paluch, M.; Wlodarczyk, P.; Dulski, M.; (2), 857−873.
Wrzalik, R.; Roland C.M., X. Tracking of Proton Transfer Reaction in (43) Paluch, M.; Ngai, K. L.; Hensel-Bielowka, S. Pressure and
Supercooled RNA Nucleoside. J. Phys. Chem. Lett. 2012, 3 (16), temperature dependences of the relaxation dynamics of cresolph-
2288−2292. thalein-dimethylether: Evidence of contributions from thermodynam-
(25) Gorb, L.; Kaczmarek, A.; Gorb, A.; Sadlej, A. J.; Leszczynski, J.
ics and molecular interactions. J. Chem. Phys. 2001, 114, 10872.
Thermodynamics and Kinetics of Intramolecular Proton Transfer in (44) Casalini, R.; Snow, A. W.; Roland, C. M. Temperature
Guanine. Post Hartree−Fock Study. J. Phys. Chem. B 2005, 109,
Dependence of the Johari−Goldstein Relaxation in Poly(methyl
13770−13776.
methacrylate) and Poly(thiomethyl methacrylate). Macromolecules
(26) Iuliucci, R. J.; Clawson, J.; Hu, J. Z.; Solum, M. S.; Barich, D.;
2013, 46 (1), 330−334.
Grant, D. M.; Taylor, C. M. Ring-chain tautomerism in solid-phase
(45) Ngai, K. L.; Paluch, M. Classification of secondary relaxation in
erythromycin A: evidence by solid-state NMR. Solid State Nucl. Magn.
glass-formers based on dynamic properties. J. Chem. Phys. 2004, 120,
Reson. 2003, 24 (1), 23−38.
(27) Adrjanowicz, K.; Zakowiecki, D.; Kaminski, K.; Hawelek, L.; 857.
Grzybowska, K.; Tarnacka, M.; Paluch, M.; Cal, K. Molecular (46) Schneider, U.; Brand, R.; Lunkenheimer, P.; Loidl, A. Excess
Dynamics in Supercooled Liquid and Glassy States of Antibiotics: Wing in the Dielectric Loss of Glass Formers: A Johari-Goldstein β
Azithromycin, Clarithromycin and Roxithromycin Studied by Dielec- Relaxation? Phys. Rev. Lett. 2000, 84, 5560−5563.
tric Spectroscopy. Advantages Given by the Amorphous State. Mol. (47) Ngai, K. L. An extended coupling model description of the
Pharmaceutics 2012, 9, 1748−1763. evolution of dynamics with time in supercooled liquids and ionic
(28) Adrjanowicz, K.; Kaminski, K.; Wojnarowska, Z.; Dulski, M.; conductors. J Phys: Condens. Matter 2003, 15, S1107.
Hawelek, L.; Pawlus, S.; Paluch, M. Dielectric Relaxation and (48) Roland, C. M.; Hensel-Bielowka, S.; Paluch, M.; Casalini, R.
Crystallization Kinetics of Ibuprofen at Ambient and Elevated Supercooled dynamics of glass-forming liquids and polymers under
Pressure. J. Phys. Chem. B 2010, 114 (19), 6579−6593. hydrostatic pressure. Rep. Prog. Phys. 2005, 68 (6), 1405−1478.
(29) Wlodarczyk, P.; Paluch, M.; Wojnarowska, Z.; Hawelek, L.; (49) Paluch, M.; Rzoska, S. J.; Habdas, P.; Ziolo, J. On the isothermal
Kaminski, K.; Pilch, J. Theoretical and experimental studies on the pressure behaviour of the relaxation times for supercooled glass-
internal mobility of two sulfonylurea agents: glibenclamide and forming liquids. J. Chem. Phys. 1999, 110 (22), 10978−10981.
glimepiride. J. Phys. Condens. Matter 2011, 23 (42), 425901. (50) Hensel-Bielowka, S.; Ziolo, J.; Paluch, M.; Roland, C. M. The
(30) Gupta, P.; Thilagavathi, R.; Chakraborti, A. K.; Bansal, A. K. effect of pressure on the structural and secondary relaxations in 1,1
Role of molecular interaction in stability of celecoxib-PVP amorphous ′-bis (p-methoxyphenyl) cyclohexane. J. Chem. Phys. 2002, 17 (5),
systems. Mol. Pharmaceutics 2005, 2 (5), 384−391. 2317−2323.
(31) Shamblin, S. L.; Hancock, B. C.; Dupuis, Y.; Pikal, M. J. (51) Havriliak, S.; Negami, S. A complex plane representation of
Interpretation of relaxation time constants for amorphous pharma- dielectric and mechanical relaxation processes in some polymers.
ceutical system. J. Pharm. Sci. 1999, 89, 417−427. Polymer 1967, 8 (4), 161.
(32) Wojnarowska, Z.; Hawelek, L.; Paluch, M.; Sawicki, W.; Ngai, K. (52) Kremer, F., Schönhals, A., Eds. Broadband Dielectric Spectros-
L. Molecular dynamics at ambient and elevated pressure of the copy; Springer: Berlin, 2003.
amorphous pharmaceutical: Nonivamide (pelargonic acid vanillyla- (53) Schönhals, A.; Kremer, F. Analysis of Dielectric Spectra. In
mide). J. Chem. Phys. 2011, 28 (134(4)), 044517. Broadband Dielectric Spectroscopy; Kremer, F., Schönhals, A., Eds.;
(33) Adrjanowicz, K.; Wojnarowska, Z.; Wlodarczyk, P.; Kaminski, Springer Verlag: Berlin, 2003; Chapter 3.
K.; Paluch, M.; Mazgalski, J. Molecular mobility in liquid and glassy (54) Vogel, H. The law of the relation between the viscosity of
states of Telmisartan (TEL) studied by Broadband Dielectric liquids and the temperature. Phys. Z. 1921, 22, 645.
Spectroscopy. Eur. J. Pharm. Sci. 2009, 38, 395. (55) Fulcher, G. S. Analysis of recent measurements of the viscosity
(34) Wojnarowska, Z.; Grzybowska, K.; Adrjanowicz, K.; Kaminski, of glasses. J. Am. Ceram. Soc. 1925, 8, 339.
K.; Paluch, M.; Hawelek, L.; Wrzalik, R.; Dulski, M.; Sawicki, W.; (56) Tammann, G.; Hesse, W. Abhangigkeit der Viscosität von der
Mazgalski, J.; Tukalska, A.; Bieg, T. Study of the Amorphous Temperatur bie unterkuhlten Flussigkeiten. Z. Anorg. Allg. Chem. 1926,
Glibenclamide Drug: Analysis of the Molecular Dynamics of 156, 245.

3626 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627


Molecular Pharmaceutics Article

(57) Bohmer, R.; Ngai, K. L.; Angell, C. A.; Plazek, D. J.


Nonexponential relaxations in strong and fragile glass formers. J.
Chem. Phys. 1993, 99, 4201.
(58) Tanaka, H. Two-order-parameter description of liquids. I. A
general model of glass transition covering its strong to fragile limit. J.
Chem. Phys. 1999, 111 (3163), 3175.
(59) Kauzmann, K. The Nature of the Glassy State and the Behavior
of Liquids at Low Temperatures. Chem. Rev. 1948, 43 (2), 219−256.
(60) Wojnarowska, Z.; Wlodarczyk, P.; Kaminski, K.; Grzybowska,
K.; Hawelek, L.; Paluch, M. On the Kinetics of Tautomerism in Drugs
- New Application of Broadband Dielectric Spectroscopy. J. Chem.
Phys. 2010, 133, 094507.
(61) Sharma, A.; Jain, C. P. Preparation and characterization of solid
dispersions of carvedilol with PVP K30. Res. Pharm. Sci. 2010, 5 (1),
49−56.
(62) Alanazi, F. K.; El-Badry, M.; Alsarra, I. A. Spray-dried HPMC
microparticles of indomethacin impact of drug-polymer ratio and
viscosity of the polymeric solution on dissolution. Saudi Pharm. J.
2006, 14 (2), 100−107.
(63) Mesnukul, A.; Yodkhum, K.; Phaechamud, T. Solid Dispersion
Matrix Tablet Comprising Indomethacin-PEG-HPMC Fabricated with
Fusion and Mold Technique. Indian J. Pharm. Sci. 2009, 71 (4), 413−
20.
(64) Grzybowska, K.; Paluch, M.; Wlodarczyk, P.; Grzybowski, A.;
Kaminski, K.; Hawelek, L.; Zakowiecki, D.; Kasprzycka, A.; Jankowska-
Sumara, I. Enhancement of Amorphous Celecoxib Stability by Mixing
It with Octaacetylmaltose: The Molecular Dynamics Study. Mol.
Pharmaceutics 2012, 9 (4), 894−904.
(65) Kaminski, K.; Adrjanowicz, K.; Wojnarowska, Z.; Grzybowska,
K.; Hawelek, L.; Paluch, M.; Zakowiecki, D.; Mazgalski, J. Molecular
dynamics of the cryomilled base and hydrochloride ziprasidones by
means of dielectric spectroscopy. J. Pharm. Sci. 2011, 100 (7), 2642−
57.
(66) Tsukushi, I.; Yamamuro, O.; Suga, H. Solid state amorphization
and glass transition of tri-O-methyl-β-cyclodextrin. J. Therm. Anal.
1991, 37 (7), 1359−1371.
(67) Adrjanowicz, K.; Wojnarowska, Z.; Grzybowska, K.; Hawelek,
L.; Kaminski, K.; Paluch, M.; Kasprzycka, A.; Walczak, K. Molecular
dynamics and crystallization phenomenon of supercooled and glassy
DNA and RNA nucleosides: β-adenosine, β-thymidine, and β-uridine.
Phys. Rev. E 2011, 84, 051507.
(68) Adrjanowicz, K.; Kaminski, K.; Grzybowska, K.; Hawelek, L.;
Paluch, M.; Gruszka, I.; Zakowiecki, D.; Sawicki, W.; Lepek, P.;
Kamysz, W.; Guzik, L. Effect of Cryogrinding on Chemical Stability of
the Sparingly Water-Soluble Drug Furosemide. Pharm. Res. 2011, 28
(12), 3220−3236.
(69) Grzybowska, K.; Grzybowski, A.; Zioło, J.; Paluch, M.;
Capaccioli, S. Dielectric secondary relaxations in polypropylene
glycols. J. Chem. Phys. 2006, 125, 044904.
(70) Wojnarowska, Z.; Grzybowska, K.; Hawelek, L.; Swiety-
Pospiech, A.; Masiewicz, E.; Paluch, M.; Sawicki, W.; Chmielewska,
A.; Bujak, P.; Markowski, J. Molecular Dynamics Studies on the Water
Mixtures of Pharmaceutically Important Ionic Liquid Lidocaine HCl.
Mol. Pharmaceutics 2012, 9, 1250−1261.
(71) Cerveny, S.; Schwartz, G. A.; Alegría, A.; Bergman, R.; Swenson,
J. Water dynamics in n-propylene glycol aqueous solutions. J. Chem.
Phys. 2006, 124, 194501.
(72) Youssef, N. F. Spectrophotometric, Spectrofluorimetric and
Densitometric Methods for the Determination of Indapamide. J.
AOAC Int. 2003, 86 (5), 935−940.

3627 dx.doi.org/10.1021/mp400116q | Mol. Pharmaceutics 2013, 10, 3612−3627

You might also like