You are on page 1of 11

http://informahealthcare.

com/phd
ISSN: 1083-7450 (print), 1097-9867 (electronic)

Pharm Dev Technol, 2014; 19(8): 976–986


! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/10837450.2013.846374

RESEARCH ARTICLE

Curcumin amorphous solid dispersions: the influence of intra and


intermolecular bonding on physical stability
Lindsay A. Wegiel1, Yuhong Zhao2,3, Lisa J. Mauer2, Kevin J. Edgar4, and Lynne S. Taylor1
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

1
Department of Industrial and Physical Pharmacy, College of Pharmacy, Purdue University, West Lafayette, IN, USA, 2Department of Food Sciences,
Purdue University, West Lafayette, IN, USA, 3Department of Food Sciences, Forestry College, Northeast Forestry University, Harbin, China, and
4
Department of Sustainable Biomaterials, Virginia Polytechnic Institute and State University, Blacksburg, VA, USA

Abstract Keywords
We have investigated the physical stability of amorphous curcumin dispersions and the role of Amorphous, curcumin, FTIR, hydrogen
curcumin–polymer intermolecular interactions in delaying crystallization. Curcumin is an bonding, physical stability, polymers, solid
interesting model compound as it forms both intra and intermolecular hydrogen bonds in the dispersion
crystal. A structurally diverse set of amorphous dispersion polymers was investigated;
poly(vinylpyrrolidone), Eudragit E100, carboxymethyl cellulose acetate butyrate, hydroxypropyl History
methyl cellulose (HPMC) and HPMC-acetate succinate. Mid-infrared spectroscopy was used to
determine and quantify the extent of curcumin–polymer interactions. Physical stability under Received 5 July 2013
different environmental conditions was monitored by powder X-ray diffraction. Curcumin Revised 12 September 2013
chemical stability was monitored by UV-Vis spectroscopy. Isolation of stable amorphous Accepted 13 September 2013
Published online 4 November 2013
For personal use only.

curcumin was difficult in the absence of polymers. Polymers proved to be effective curcumin
crystallization inhibitors enabling the production of amorphous solid dispersions; however, the
polymers showed very different abilities to inhibit crystallization during long-term storage.
Curcumin intramolecular hydrogen bonding reduced the extent of its hydrogen bonding with
polymers; hence most polymers were not highly effective crystallization inhibitors. Overall,
polymers proved to be crystallization inhibitors, but inhibition was limited due to the
intramolecular hydrogen bonding in curcumin, which leads to a decrease in the ability of the
polymers to interact at a molecular level.

Introduction can be expected upon disrupting its crystalline structure.


The value of amorphous solid dispersions may be limited by the
Curcumin is a yellow–orange pigment derived from the rhizome
potential physical instability of these systems over pharmaceut-
of Curcuma longa. Curcumin exhibits keto–enol tautomerism: the
ically relevant time scales, in particular, drug crystallization,
di-keto form is shown in Figure 1a, and the keto–enol form is
which will negate the solubility advantage offered by using an
shown in Figure 1b. This compound has been of great interest
amorphous formulation.
recently due to its promising and diverse pharmacological and
The tendency of a drug to crystallize depends on both
biological properties, including anti-oxidant1,2, anti-inflamma-
environmental and formulation parameters. Factors such as
tory2, anti-fungal3, anti-tumor4, anti-cancer5, anti-viral6, anti-
temperature and relative humidity (RH) can have a pronounced
13
obesity7 and cardioprotective effects8. Curcumin is moderately
20
effect on the physical stability of solid dispersions16,17. Polymers
hydrophobic (logP 2.5) and has very low bioavailability due, in
are typically added to inhibit drug crystallization during processing
part, to its low aqueous solubility (11 ng/mL at pH 5.0)9.
and storage. The physical stability of a solid dispersion will not
In addition, curcumin degrades very quickly at neutral or alkaline
only depend on the inherent crystallization tendency of the drug but
pH, which can further decrease its bioavailability10,11. Numerous
also on the ability of the polymer to inhibit crystallization18.
attempts to increase curcumin solubility have been reported in the
Curcumin is an intermediate crystallizer according to the classi-
literature, including complexation with cyclodextrins9, formation
fication system of Van Eerdenbrugh et al.19. It is therefore crucial
of nanoparticles12 and formulation into solid dispersions13. One
to understand polymer properties important for crystallization
of the most promising strategies for curcumin is formulation
inhibition of curcumin if stable amorphous dispersions are to be
as a solid dispersion, as the amorphous state of the drug has a
formed.
significant solubility advantage over its crystalline counter-
Recent studies from our group have suggested correlations
part9,14,15. This is due to the high melting point of curcumin
between crystallization rates and the extent to which different
(183  C), which reflects strong crystal lattice energy; substantial
polymers form drug–polymer intermolecular interactions20–23.
improvements in solution concentration and biological exposure
Furthermore, using techniques such as infrared (IR) spectroscopy
to monitor drug polymer interactions appears to be a reliable
method for predicting amorphous solid dispersion physical
Address for correspondence: Lynne S. Taylor. Tel: þ1-755-496-6614. stability23. Curcumin is an interesting model compound as it
Fax: þ1-755-494-6545. E-mail: lstaylor@purdue.edu has considerable intramolecular hydrogen bonding; therefore,
DOI: 10.3109/10837450.2013.846374 Curcumin solid dispersions – role of interactions on physical stability 977

there is less opportunity for this compound to form intermolecular Materials


hydrogen bonds with polymers (Figure 1). In this study, disper-
Dichloromethane (ChromAR grade) and acetone (ChromAR
sions of curcumin with select polymers were formed. IR
grade) were obtained from Mallinckrodt Baker, Inc., Paris, KY.
spectroscopy was used to evaluate the extent of curcumin–
Absolute ethanol was obtained from Pharmaco Products, Inc.,
polymer intermolecular interactions, and the impact of these
Brookfield, CT and Aaper, Shelbyville, KY. Curcumin was
interactions upon the rate of crystallization was evaluated. Model
obtained from Spectrum Chemical Manufacturing Corp., New
polymers were chosen for diversity in hydrogen bonding potential,
Brunswick, NJ. PVP (K29/32, Mw 58 000) was purchased from
including poly(vinylpyrrolidone) (PVP), Eudragit E100 (E100),
Sigma-Aldrich (St. Louis, MO). HPMC (grade 606, hypromellose
carboxymethyl cellulose acetate butyrate (CMCAB), hydroxypro-
United States Pharmacopeia substitution type 2910, Mw 35 600)
pyl methyl cellulose (HPMC) and HPMC-acetate succinate
and HPMCAS (grade AS-MF, Mw 17 000) were provided by Shin-
(HPMCAS) (Figure 2).
Etsu Chemical Co., Ltd., Tokyo, Japan. Eudragit E100 (Mw
47 000) was obtained from Rohm GmbH, Darmstadt, Germany.
CMCAB (641-0.5, Mw 35 000) was obtained from Eastman
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

Chemical Co., Kingsport, TN. The polymers were kept in


desiccators with P2O5 for at least 1 week prior to use to remove
any residual moisture.

Methods
Formation of solid dispersions
Attempts to make curcumin amorphous by rapid solvent evapor-
ation were only successful by using the small scale spincoating
technique described below. Attempts to obtain larger quantities of
neat amorphous curcumin by rotary evaporation (technique
described below) were not successful. However, a solvent-free
process (cryomilling in a liquid nitrogen bath) permitted prepar-
ation of a larger quantity of amorphous curcumin powder.

Spincoating
For personal use only.

Solutions were prepared by dissolving curcumin and polymer at


different dry weight ratios in a 1:1 (by weight) mixture of
dichloromethane and ethanol. All mixtures were visually
inspected to confirm that both curcumin and polymer were fully
Figure 1. Chemical structure of (a) curcumin di-keto form, (b) curcumin
dissolved, forming a uniform one-phase solution. Two or three
keto–enol form and (c) curcumin keto–enol form showing resonance drops of each solution were placed on a thallium bromoiodide
assisted intra-molecular hydrogen bond. *Denotes atoms involved in (KRS-5) optical crystal, which was immediately rotated on a
intermolecular hydrogen bonding in the crystalline form I polymorph.

Figure 2. Chemical structures of (a) Eudragit E100, (b) PVP, (c) CMCAB, (d) HPMC and (e) HPMCAS.
978 L. A. Wegiel et al. Pharm Dev Technol, 2014; 19(8): 976–986

KW-4A spin coater at 500/2500 rpm for 18 and 30 s, respectively, labeled X-ray amorphous. The detection limit was found to be
to produce a thin film. about 5–10 wt% crystalline material by adding small amounts
crystalline material to a solid dispersion until the powder X-ray
Rotary evaporation diffraction (PXRD) showed visible peaks in the diffractogram.
Solutions were prepared by weighing out curcumin and polymer
Differential scanning calorimetry
at a dry weight ratio of 3:1, then dissolving the mixture in a 1:1
(by weight) mixture of dichloromethane and ethanol. All mixtures Thermal analysis was performed on the samples prepared by
were visually inspected to confirm that the curcumin and the rotary evaporation using a TA Q2000 differential scanning
polymers were fully dissolved, forming uniform one-phase calorimeter (DSC) equipped with a refrigerated cooling accessory
solutions. The solvent was removed using a rotary evaporator (TA Instruments, New Castle, DE). Nitrogen, 50 mL/min, served
(Brinkman Instruments, Westbury, NY) with a water bath as the purge gas. Samples (3–7 mg) were weighed into aluminum
maintained at 60  C. The samples were then placed under T-zero sample pans with pin holes (TA instruments) and sealed.
vacuum for 24 h at room temperature to remove any residual Samples were heated from 0  C to 120  C at 10  C/min, quickly
cooled to 0  C at the maximum instrument cooling rate (15  C/
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

solvents. The obtained material was subsequently cryomilled in a


liquid nitrogen bath as described below to obtain a uniform min), then heated from 0  C to 200  C at 10  C/min; transitions
powder sample. are reported from this second heating scan. For the PVP solid
dispersions, the Tg could not be resolved without modulation.
Cryomilling These samples were heated from 25  C to 120  C at 10  C/min
then quickly cooled to 0  C at the maximum cooling rate of the
The rotary evaporated samples were added to the cryomill in a
instrument, held isothermally at 0  C for 5 min, then heated at
liquid nitrogen bath and milled for 2 min at a rate of two strikes
2  C/min to 200  C with a modulation of  1  C every 60 s. DSC
per second then cooled for 2 min. This process was repeated once.
heating curves were analyzed using Universal Analysis 2000
For the dry powdered curcumin, for which the goal was to obtain
software (TA Instruments).
an amorphous material, the sample was added to the cryomill in a
liquid nitrogen bath. The sample was milled for 2 min at a rate of
Ultraviolet-visible spectra
10 strikes per second then cooled for 2 min; this was repeated for
30 cycles (1 h of total milling time). Effects of formulation and storage treatments on curcumin content
of solid dispersions prepared by the rotary evaporation technique
IR spectroscopy were determined using an ultraviolet-visible spectrometer
(Beckman DU800 Spectrophotometer, Fullerton, CA) at max
For personal use only.

IR spectra of the spin-coated thin films were obtained in


424 nm using a method adapted from Jasim and Ali24. Curcumin
absorbance mode using a Bio-Rad FTS 6000 spectrophotometer
concentration in the sample was calculated using a standard curve
(Bio-Rad Laboratories, Hercules, CA) equipped with globar IR
constructed over the concentration range of interest. The results
source, KBr beamsplitter, and DTGS detector. The scan range was
were treated statistically using SPSS software version 17 (SPSS
set from 4000 to 500 cm–1 with 4 cm–1 resolution, and 128 scans
Inc., Chicago, IL), and data were expressed as means  standard
were co-added. The absorbance intensity of the spectral region of
deviations. One-way analysis of variance with post hoc Tukey
interest was between 0.6 and 1.2. During measurements, the spin-
(HSD) test was employed to identify significant differences
coated samples, the sample and optical compartments of the
(p50.05) between data sets.
spectrophotometer were flushed with dry air23.
Results
Storage treatments
Curcumin is an intermediate crystallizer according to the
Amorphous curcumin and dispersions (25% polymer and 75% classification system of Van Eerdenbrugh et al.19. Therefore, it
curcumin) were stored at six different environmental conditions was possible to make the compound amorphous by spin coating,
for up to 350 d and sampled periodically. The storage conditions but a larger amorphous sample could not be prepared via rotary
were: room temperature (22–25  C) in a desiccator with P2O5 (0% evaporation. However, cryomilling enabled preparation of a larger
RH), room temperature (22–25  C) in a desiccator with potassium mass of amorphous powder.
chloride (84% RH), 40  C in a desiccator with sodium chloride
(75% RH), 60  C in a desiccator with potassium sulfate (98% RH) IR spectroscopic investigation of curcumin–polymer
as well as 50  C and 90  C in desiccators with DrieriteÔ (W.A. interactions
Hammond Drierite Company, Ltd., Xenia, OH) (anhydrous
calcium sulfate, 0% RH). Drug–polymer interactions in solid dispersions can be explored
using IR spectroscopy23, whereby the presence of specific
interactions is often used to infer miscibility25,26. The IR spectrum
Powder X-ray diffraction
of a miscible solid dispersion typically contains peaks that are
Immediately after cryomilling, the samples were analyzed using a shifted relative to those found in the spectrum of the physical
Shimadzu XRD-6000 diffractometer (Shimadzu Corporation, mixture of amorphous drug and polymer. This is due to specific
Kyoto, Japan) equipped with a Cu-Ka source and set in Bragg- interactions between the chemical entities of the drug and
Brentano geometry, scanning between 5 and 35 2 at 8 /min with polymer, usually hydrogen bonding. If these interactions differ
a 0.04 step size. Samples were also analyzed as a function of in strength or extent relative to self-associations in these systems,
time following the storage treatments listed above. Before each then there are changes in IR peak positions and shape for the
day of measurement, the accuracy of the 2 angle was checked by functional groups involved.
verifying that the [111] peak of a Si-standard sample was between In order to understand the origin of any spectroscopic changes,
28.423 and 28.463 . The measured photon intensity of the [111] it is first important to consider the inter- and intra-molecular
peak was used to normalize the data collected from samples on interactions formed in pure curcumin. Curcumin exhibits keto–
that day. Samples where sharp peaks were observed were deemed enol tautomerism: the di-keto form is shown in Figure 1a, and the
crystalline, and those samples that lacked sharp peaks were keto–enol form is shown in Figure 1b. Single crystal studies have
DOI: 10.3109/10837450.2013.846374 Curcumin solid dispersions – role of interactions on physical stability 979
Figure 3. IR spectra of amorphous and
crystalline curcumin.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

shown that the keto–enol form is favored in all crystal forms27,28. Table 1. OH peak position in the IR spectra of amorphous curcumin and
The keto–enol tautomer is further stabilized in the most stable curcumin–polymer dispersions at 80% w/w curcumin.
polymorphic form by a resonance-assisted intramolecular hydro-
For personal use only.

gen bond (Figure 1c) where the hydrogen atom is symmetrically Sample OH peak position (cm–1)
positioned between the two oxygen atoms27. Formation of an Amorphous curcumin 3400
intramolecular hydrogen bond leads to increased delocalization of HPMC–curcumin 3386
the -conjugate electrons as depicted in Figure 1c29. Figure 3 HPMCAS–curcumin 3394
shows the IR spectrum of crystalline curcumin (stable polymorph CMCAB–curcumin 3399
and monoclinic form) whose structure is known to be similar to E100–curcumin 3396
PVP–curcumin 3187
that depicted in Figure 1c27. As expected, the spectrum is missing
the strong carbonyl stretching band that would be anticipated for
the di-keto form at around 1700 cm–1. Rather, there is a prominent
stretching band at 1626 cm–1 consistent with formation of a keto– Table 2. C¼O stretching vibration peak positions in IR spectra of the
enol tautomer stabilized by a strong intramolecular hydrogen model polymers, and observed shifts in curcumin–polymer dispersions at
bond. The spectrum of amorphous curcumin (Figure 3) is very 80% w/w curcumin.
similar to that of crystalline curcumin with regard to the pattern of
peaks in this spectral region, indicating that the structure of C¼O peak C¼O peak
position in position with 80% Shift
curcumin is predominantly present as the intramolecular hydrogen
Sample polymer (cm–1) curcumin (cm–1) observed (cm–1)
bonded keto–enol tautomer in the amorphous form. This is to be
expected since ab initio calculations indicated that the keto–enol PVP 1682 1655 27
tautomers are lower energy than the corresponding di-keto form30. HPMC N/A N/A N/A
HPMCAS 1743 1735 8
Strong intramolecular hydrogen bonding makes the keto–enol
CMCAB 1749 1745 4
group less available for intermolecular hydrogen bonding. E100 1730 1728 2
Curcumin also has two phenol groups that have the potential for
intermolecular hydrogen bonding. However, in crystalline curcu-
min, intramolecular hydrogen bonds are formed between the wavenumbers would be expected upon formation of these
phenolic groups and methoxy substituents, oriented ortho to one interactions31. The C¼O stretching vibrations assigned to the
another on the phenyl ring27. Therefore, it might be anticipated carbonyls for the relevant polymers are listed in Table 2, along
that curcumin hydrogen bonding with a polymer will be weak due with the solid dispersion shifts observed in the curcumin–polymer
to limited possibilities for intermolecular hydrogen bonding. dispersion at 80% w/w curcumin (larger shift indicates stronger/
Changes in the OH region can be used to provide information more extensive hydrogen bonding).
about curcumin–polymer intermolecular hydrogen-bonding inter- For all of the solid dispersions of curcumin with cellulosic
actions. The OH peak positions of amorphous curcumin and polymers (HPMC, HPMCAS or CMCAB), curcumin–polymer
curcumin–polymer dispersions at 80% w/w curcumin (lower interactions were observed but they appeared to be minor
wavenumber indicates stronger hydrogen bonding) are listed in (Figures 4–6). For example, in the presence of 20% polymer,
Table 1. In addition, all of the polymers, except HPMC, have the OH peak occurred at 3386 cm–1, 3394 cm–1 and 3399 cm–1 for
C¼O groups. The carbonyl group of the polymer can potentially HPMC, HPMCAS and CMCAB, respectively, compared to
hydrogen bond with the phenolic OH groups in curcumin, and 3400 cm–1 for the pure amorphous material (Table 1). Based on
resultant shifts of the carbonyl stretching frequency to lower the relative extent of the OH peak shift, it can be inferred that
980 L. A. Wegiel et al. Pharm Dev Technol, 2014; 19(8): 976–986

Figure 6. IR spectra of curcumin–HPMC solid dispersions showing the


Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

OH stretching region (3600–2700 cm–1). Percentage indicates the weight


percent of curcumin in the solid dispersion.

Figure 4. IR spectra of curcumin–CMCAB solid dispersions showing (a)


the OH stretching region (3600–2800 cm–1) and (b) the carbonyl
stretching region (1775–1580 cm–1). Percentage indicates the weight
For personal use only.

percent of curcumin in the solid dispersion.

Figure 7. IR spectra of curcumin–PVP solid dispersions showing (a) the


OH stretching region (3600–2650 cm–1) and (b) the carbonyl stretching
region (1720–1480 cm–1). Percentage indicates weight % curcumin in
solid dispersion.

confirmed that CMCAB has a weaker hydrogen bonding inter-


action with curcumin based on the observation that the polymer
carbonyl peak shifts only by about 4 cm–1 in the presence of
curcumin, versus an 8 cm–1 wavenumber shift with HPMCAS
(HPMC does not have a carbonyl group). Overall the IR
spectroscopy results indicate weak hydrogen bonding between
the cellulosic polymers and curcumin.
PVP contains an amide carbonyl on the pyrrolidone monomer,
which is a strong hydrogen bond acceptor, thus stronger
interactions would be anticipated relative to the cellulosic
polymers, which have with their weaker acceptor groups22.
Stronger interactions were indeed observed for the curcumin–
PVP system, based on the IR data. In the hydroxyl region of the
Figure 5. IR spectra of curcumin–HPMCAS solid dispersions showing
(a) the OH stretching region (3700 cm–1–2700 cm–1) and (b) the carbonyl IR spectrum (Figure 7a), at 80% curcumin, the curcumin OH peak
stretching region (1800–1580 cm–1). Percentage indicates the weight had a maximum at around 3187 cm–1. This is considerably lower
percent of curcumin in the solid dispersion. than the peak maximum in pure curcumin (3400 cm–1) and is the
lowest value seen for this peak out of all of the curcumin–polymer
curcumin formed the weakest hydrogen bonds with CMCAB and dispersions studied. This suggests that the hydrogen bonds formed
the strongest with HPMC with regard to the cellulosic polymers. in this system were the strongest/most extensive. In the carbonyl
In the carbonyl region, where the polymer acceptor group spectral region (Figure 7b), a new shoulder at 1655 cm–1 emerges
provides additional information about hydrogen bonding, it was due to the presence of a hydrogen bonded PVP carbonyl. This is a
DOI: 10.3109/10837450.2013.846374 Curcumin solid dispersions – role of interactions on physical stability 981

wavenumber shift of 27 cm–1 relative to the non-hydrogen bonded (Figure 8a). Thus, there is only about a 4 cm–1 shift relative to
carbonyl in pure PVP (1682 cm–1) and confirms curcumin– amorphous curcumin, indicating that interspecies hydrogen
polymer hydrogen bonding. As expected, the ratio of the free bonding is limited or weak. Consistent with this interpretation,
carbonyl peak to the hydrogen bonded carbonyl peak increased as little to no evidence of hydrogen bonding can be inferred from the
the concentration of the polymer increased. The large shift in the carbonyl peak position where only a 2 cm–1 wavenumber shift is
carbonyl peak relative to that observed for the cellulosic polymers observed. This is the smallest shift of all the polymers studied.
provides further evidence that the strongest hydrogen bonds are However, E100 contains dimethylamino groups, which can also
formed in the PVP–curcumin dispersion. potentially interact with curcumin through ionic interactions
In the case of the E100-curcumin system, the curcumin OH (Figure 8b). In the E100 spectrum, there are two peaks at
peak occurred at 3396 cm–1 in the presence of 20% polymer 2820 cm–1 and 2770 cm–1 that are reported to correspond to the
non-ionized dimethylamino groups32. As the concentration of
curcumin was increased, these peaks quickly decreased in
intensity and were fully eradicated when 60% curcumin was
present in the dispersion. The disappearance of these peaks
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

provides evidence for an acid–base interaction32 between the


acidic phenol hydroxyls of curcumin and the amino groups of
E100, as it indicates the presence of ionized dimethylamino
groups. This acid–base interaction potentially can lead to an
increase in physical stability33–35 and will be discussed later.

Crystallization tendency of curcumin solid dispersions


Storage stability was tested at 0% RH (25  C, 50  C and 90  C),
84% RH (25  C), 75% RH (40  C) and 98% RH (60  C).
Interestingly, the nature of the dispersion polymer had an
impact on color (Figure 9). The E100 solid dispersions were
rich dark red in color, while the other solid dispersions all had an
orange color similar to that of pure curcumin. Chemical
degradation did not appear to be the cause of the color change,
as the curcumin-E100 dispersions did not show a decreased level
For personal use only.

of curcumin content compared to the other polymer dispersions.


In fact, all of the dispersions were found to be chemically stable
following storage for up to 200 d for all stability conditions apart
from the most extreme storage condition of 98% RH (60  C)
(Figure 10). Since the latter system was physically unstable and
crystallized after 1 d, its chemical instability is not of relevance to
Figure 8. IR spectra of curcumin-E100 solid dispersions showing: (a) OH this study. Pure crystalline curcumin did not show evidence of
stretching region (3600–2700 cm–1) and (b) carbonyl stretching region degradation at any environmental condition and proved to be
(1770–1450 cm–1). Percentage indicates weight % curcumin in solid chemically stable over the time frame of this study. The color
dispersion.

Figure 9. Images of curcumin solid dispersions with 75% curcumin and 25% polymer. (a) Curcumin, (b) curcumin–E100, (c) curcumin–PVP, (d)
curcumin–CMCAB, (e) curcumin–HPMCAS and (f) curcumin–HPMC.
982 L. A. Wegiel et al. Pharm Dev Technol, 2014; 19(8): 976–986
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

Figure 11. PXRD of amorphous curcumin versus storage time at 50  C,


0% RH.
Figure 10. Chemical stability of curcumin-E100 solid dispersion (75%
curcumin w/w) versus storage temperature, relative humidity.
between the polymers could be made. At all other storage
Table 3. Long-term physical stability of amorphous curcumin-polymer conditions, the physical stability of the solid dispersions was
dispersion at 75% w/w curcumin by PXRD (days before crystalline highly dependent on the nature of the matrix polymer.
material is observed).
The poorest crystallization inhibitor was PVP. At 50  C, the
PVP solid dispersion started to crystallize after 6 d, as compared
Solid Tg 0% RH 0% RH 0% RH 84% RH 75 % RH 98% RH
dispersions ( C) 25  C 50  C 90  C 25  C 40  C 60  C
to amorphous curcumin, which exhibited crystallization at day 3.
In addition, at high RH conditions (84% RH) and room
Cur amorphous 43.6 4100 3 1 14 1 1 temperature, crystallization occurred by 22 d for the PVP
E100 62.4 4350 4350 1 4350 4200 1 dispersion versus 10 d for amorphous curcumin alone. At 40  C
PVP 62.8 4350 6 1 22 16 1
HPMC 68.3 4350 4350 1 75a 100a 1b
(75% RH), crystallization of the PVP solid dispersion was
HPMCAS 65.3 4350 75a 1b 75a 30a 1b inhibited for 16 d versus amorphous curcumin, which crystallized
CMCAB 61.7 4350 75a 1 75a 30a 1b after only 1 d. Overall, PVP inhibited crystallization to some
extent; however, the PVP system crystallized faster than disper-
For personal use only.

4X indicates monitoring was stopped after this time point. sions containing any of the other polymers at comparable storage
a
System only partially crystallized. conditions.
b
Crystallized into a metastable polymorph.
The dispersions containing E100 were the most stable against
crystallization at the intermediate temperature and RH storage
change observed in the presence of E100 can be attributed to a conditions. The E100 dispersion was stable at 0% RH (25  C and
change in the ionization state of curcumin. E100 is a basic 50  C) and was the only dispersion stable at both 84% RH (25  C)
polymer and therefore can cause ionization of curcumin if an and 75% RH (40  C). For the latter conditions, the dispersions
acid–base reaction occurs. It is well known that curcumin changes were stable to crystallization for up to 350 d and 200 d,
color as a function of pH and has a darker color at high pH36. This respectively. E100 was unable to prevent crystallization at the
observation is in good agreement with the inferences made from harsh conditions of 60  C (98% RH) and 90  C (0% RH) where
the IR analysis. complete crystallization was seen after 1 d. However, this system
Curcumin-polymer solid dispersion physical stability as a was X-ray amorphous for considerably longer time than any of the
function of time was found to depend greatly on the polymer used other dispersions stored at the same condition, showing much
to form the dispersion, as well as on the storage conditions better physical stability.
(Table 3). The difference in crystallization tendency could not be The cellulosic polymers provided some enhancement of
correlated with the glass transition temperature (Tg) as all of the physical stability, with HPMC proving to be the most effective
dry solid dispersions had Tg values close to 65  C (Table 3), while cellulosic crystallization inhibitor. However, the curcumin–
pure amorphous curcumin has a Tg of 44  C. Thus, although the HPMC dispersion only had long-term stability at 0% RH (25  C
Tg in the solid dispersions is increased relative to curcumin alone, and 50  C). Crystallization occurred after 75 d for the sample at
the significant differences in physical stability of the solid 84% RH (25  C) and after 100 d when stored at 75% RH (40  C).
dispersions containing different polymers could not be explained At all other storage conditions, the sample showed evidence of
by small differences in Tg values. In addition, there was no trend crystalline material after 1 d. Interestingly, complete crystalliza-
with pure polymer Tg values as E100 has a Tg of 41  C, and tion was never reached at some of the storage conditions. The
CMCAB has a Tg of 137  C. system instead underwent partial crystallization, and then no
Neat amorphous curcumin was used as a negative control and additional crystal growth was observed over the time scale of the
was found to crystallize at all storage conditions (Figure 11), apart study (Figure 12). The two other cellulosic polymers (CMCAB
from 0% RH and 25  C, where it was physically stable for more and HPMCAS) were similar in their ability to inhibit crystalliza-
than 100 d. The crystal form was the form I polymorph, listed in tion at a given storage condition. Both systems were stable at 0%
the Cambridge Crystallographic Data Center under reference code RH (25  C) for up to 350 d; however, after 75 d at 0% RH (50  C)
BINMEQ, with characteristic reflections at 8.9, 17.4, 23.5 and crystalline material was observed. In addition, it only took 30 d
25.1 2. It is thus unsurprising that all of the solid dispersions and 75 d to observe crystallization at 75% RH (40  C) and 84%
were also physically stable at this storage condition, with no RH (25  C), respectively. Again, complete crystallization was not
crystallinity detected for storage periods approaching 1 year. At observed at these intermediate storage conditions. In addition, for
the harshest storage conditions of 60  C (98% RH) and 90  C (0% dispersions prepared with the cellulosic polymers, a different
RH), all of the samples crystallized within a day. Thus for the polymorphic form of curcumin was found to crystallize from the
mildest and harshest crystallization conditions, no differentiation polymeric dispersion following storage at some conditions
DOI: 10.3109/10837450.2013.846374 Curcumin solid dispersions – role of interactions on physical stability 983

(Figure 13). At 98% RH (60  C), curcumin crystallized to form II the more stable polymorphic form over the time frame of the
from all of the cellulose dispersions. This polymorph (CCDC study.
reference code of BINMEQ), could be identified from the
characteristic reflections at 13.9, 26.0 and 27.1 2 (the 26.0 Discussion
and 27.1 2 peaks are not resolved by our instrument). In
Curcumin has a relatively high inherent crystallization tendency,
addition, form II was seen in the HPMCAS dispersion following
so it is difficult to prepare pure, amorphous curcumin using
storage at 0% RH (90  C). Furthermore, form II did not convert to
methods such as rapid solvent evaporation. When made amorph-
ous using cryomilling, it remains amorphous only when stored at
room temperature and extremely low RH, but crystallizes rapidly
when exposed to higher humidity or temperature. The rapid
crystallization behavior and the extensive intramolecular hydro-
gen bonds present in both crystalline and amorphous curcumin
make this an interesting model compound with which to evaluate
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

the crystallization inhibition ability of different polymers. This is


because the presence of extensive intramolecular hydrogen
bonding in curcumin might be expected to hinder curcumin–
polymer interactions in the solid dispersions that are thought to be
important for inhibition of crystallization20–23.
Prior to 2011, only one polymorph of curcumin had been
Figure 12. PXRD of curcumin–CMCAB solid dispersion (75% curcumin reported. However, recently, Sanphui et al.28 reported two new
w/w) after 250 d at different conditions (only partial crystallization
observed).
polymorphs of curcumin, named forms II and III, both of which
are metastable with respect to the original form I polymorph. In
form I, curcumin is present as a non-planar molecule with the
benzene rings at the end of the heptane chain slightly twisted with
respect to one another, but curcumin is planar in forms II and III
(Figure 14)28. Extensive intramolecular hydrogen bonding persists
in all polymorphic forms of curcumin, suggesting that such H-
bonding is energetically favorable. The keto–enol tautomer forms
an intramolecular hydrogen bond in each of the crystal structures.
For personal use only.

In form I, the hydrogen is shared equally between the two oxygen


atoms, while in forms II and III, the hydrogen atom is not centered
and favors one oxygen over the other28. All forms also have strong
intramolecular hydrogen bonds between the ortho methoxy and
hydroxyl benzene substituents27. Form I has intermolecular
hydrogen bonds formed between the phenolic OH and the enol
OH group. There are also a number of much weaker CHO
intermolecular interactions. Forms II and III have more extensive
intermolecular hydrogen bonding but are less stable than Form I.
Figure 13. PXRD of curcumin–polymer solid dispersions (98% RH, Considering the strong and extensive intramolecular inter-
60  C and 1 d). The form II polymorph could be identified by the presence actions in curcumin, it is reasonable to expect that polymers that
of the characteristic peaks between 26.0 and 27.1 2 and the absence of interact only through hydrogen bonds might be poor curcumin
the form I peak at 17.4 2.
crystallization inhibitors. This supposition arises from the

Figure 14. IR spectra of the carbonyl stretching region (1710 cm–1–1630 cm–1) of (a) curcumin–PVP solid dispersion and (b) resveratrol–PVP solid
dispersion23. Percentage indicates the weight percent of polyphenol in the solid dispersion.
984 L. A. Wegiel et al. Pharm Dev Technol, 2014; 19(8): 976–986

expectation that there will be less opportunity for strong/extensive Table 4. The onset glass transition temperatures of polymers and
hydrogen bonding after intramolecular interactions are accounted polymer–curcumin solid dispersions at 25% drug loading.
for. Furthermore, polymers that contain only hydrogen bond
acceptors, such as PVP, suffer the additional disadvantage of only Solid Polymer Calculated Observed Deviation
dispersions Tg ( C) Tg ( C) Tg ( C) ( C)
being able to interact with a donor group on curcumin. In contrast,
polymers that contain both donor and acceptor groups, such as the Cur amorphous 44
cellulosic polymers, may gain an advantage by forming hydrogen E100 41 42 62 20
bonds with both curcumin donors and acceptor moieties, and may PVP 159 96 63 –33
HPMC 145 92 68 –24
thus be able to better disrupt curcumin self-associations important HPMCAS 111 80 65 –15
to crystal formation. This may well explain why curcumin–PVP CMCAB 137 89 62 –27
was the least stable of all the dispersions against crystallization.
This result is in direct contrast to the PVP-resveratrol system, Calculated Tg values were based on the Fox equation, and deviations are
where PVP was found to be an extremely effective crystallization taken as the difference between the calculated and observed values.
inhibitor23. Indeed, comparison of curcumin–PVP and resvera-
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

trol–PVP IR spectra shows that much less hydrogen bonding


occurs between PVP and the polyphenol in the case of curcumin extremely good physical stability of curcumin–E100 dispersions
(Figure 11). Given the similar phenol H-bonding groups in to crystallization, since little evidence of curcumin–polymer
resveratrol and curcumin, it can be supposed that the intramo- hydrogen bonding was found in this system.
lecular hydrogen bond formed by curcumin is indeed the critical It also important to consider the Tg values of the dispersions
factor in diminishing its ability to interact with PVP. The investigated. It has been argued that the dispersions with the
cellulosic polymers are slightly better crystallization inhibitors for highest Tg will be the most stable43. However, all of the dispersion
curcumin, presumably because they can form both donor and Tg values are around 62–68  C (Table 4). Certainly, Tg does not
acceptor hydrogen bond interactions with curcumin. However, explain the higher stability of the E100 and the lower stability of
these interactions appear to be fairly weak based on the IR the PVP dispersions against crystallization relative to the other
spectra; hence the cellulosic polymers fail to prevent curcumin polymers, since all the dispersions have similar Tg values. It has
crystallization under challenging humidity and temperature con- been suggested that deviations in Tg predictions reflect the nature
ditions. Therefore, it appears that for a compound like curcumin, and extent of interactions formed between drug and the
which forms intramolecular hydrogen bonds, hydrogen bonding additive44–46. The Fox equation is one of the simplest models to
with polymers is impeded and it is more difficult to inhibit predict the Tg. In this model, the Tg of a system comprised of two
crystallization. A similar observation was made with nimesulide amorphous components is calculated from the weight-averaged
For personal use only.

where it was seen that there was little differentiation between contribution of each component47:
chemically dissimilar polymers in terms of their ability to reduce 1 w1 w2
the crystal growth rate of this compound because of the inability ¼ þ : ð1Þ
Tg, mix Tg1 Tg2
of the polymer to compete with intramolecular interactions21.
These observations are also in line with the work of Etter37,38, where wi and Tgi are the weight fractions and glass transition of
whereby it is suggested that intramolecular hydrogen bonds that component i. This model is based on the free-volume theory. In
complete a six-membered ring structure are typically favored over the presence of specific interactions, such as hydrogen bonding or
intermolecular hydrogen bonds in crystals. ion-dipole or ionic interactions, the mathematical prediction of Tg
Given the reluctance of curcumin to form intermolecular for mixed systems are no longer valid, since the assumptions of
hydrogen bonds, it is of interest to explore the origin of the free-volume additivity and ideal mixing are likely to have been
extremely good resistance of the E100-curcumin dispersions to violated. For example, in the presence of strong specific
crystallization. E100 is a polybase and proton acceptor. The interactions between the different components, the free volume
average pKa of the basic monomer is 8.439. Curcumin has pKa of the mixed system can be expected to be smaller than the
values of 10.51, 9.88 and 8.3840. Based on the IR data, it is predicted free volume. This would in turn lead to positive
apparent that E100 was able to form ionic interactions with deviations of Tg. Conversely, the net loss of specific interactions
curcumin. Such interactions with E100 have been seen with a would lead to a negative deviation of Tg. Table 4 lists the
number of pharmaceutical compounds, including resveratrol23, calculated and observed Tg as calculated by the Fox equation.
dexamethasone phosphate41, and indomethacin42. There are a Interestingly it can be seen that the E100 dispersion has a higher
number of studies in the literature that have investigated Tg than that observed for either the pure polymer or pure
miscibility and physical stability of solid dispersion systems, curcumin. This leads to a large deviation from the Fox equation
which have ionic interactions. Weuts et al.34 showed salt and is consistent with the formation of ionic interaction observed
formation between a basic drug (loperamide) and an acidic between curcumin and E100. However, for all the other disper-
polymer (poly(acrylic acid); PAA), and correlated the salt sions, negative deviations are observed, with PVP showing the
formation with good miscibility and physical stability. Miyazaki largest negative deviation. This may suggest a net loss of
et al.33 reported the inhibition of drug recrystallization by proton interactions for these systems resulting from hydrogen bonding of
transfer between the acid functional groups of the polymer PAA curcumin with the polymer. Similar results have been observed for
and basic functional groups of the drug (p-aminoacetanilide). other systems showing hydrogen bonding interactions48. Clearly,
In addition, Yoo et al.35 reported that ionic interactions play Table 4 shows that it is hard to predict the stability of the system
a significant role for both drug–polymer miscibility and drug to crystallization based on consideration of Tg values alone.
physical stability in that systems with ionic interactions were Interestingly, the cellulosic polymers were able to influence
stable for up to 50 d at 60% RH and 25  C, while other non-ionic the crystalline form of curcumin that recrystallizes from the solid
systems tended to recrystallize. These reports demonstrate the dispersion. The PXRD patterns obtained following crystallization
significant influence of acid–base ionic interactions on the of curcumin from dispersions containing cellulose derivatives
physical stability of amorphous solid dispersions. The results match that of the polymorph form II reported by Sanphui et al.28.
from this study are consistent with these previous reports in that This metastable polymorph has a higher dissolution rate
ionic drug–polymer interactions appear to be the reason for the and increased solubility relative to the stable polymorph28.
DOI: 10.3109/10837450.2013.846374 Curcumin solid dispersions – role of interactions on physical stability 985

Clearly, recrystallization to any form would be undesirable, but it 12. Anand P, Nair HB, Sung BK, et al. Design of curcumin-loaded
interesting that this uncommon polymorph of curcumin appears PLGA nanoparticles formulation with enhanced cellular uptake, and
upon crystallization of some solid dispersion formulations. increased bioactivity in vitro and superior bioavailability in vivo.
Biochem Pharmacol 2010;79:330–338.
Therefore, even if the amorphous phase cannot be stabilized for 13. Paradkar A, Ambike AA, Jadhav BK, Mahadik KR. Characterization
a pharmaceutically relevant time scale, there might still be of curcumin-PVP solid dispersion obtained by spray drying. Int J
benefits from formulating with a cellulosic polymer that could Pharm 2004;271:281–286.
influence and stabilize the polymorphic form II, thus leading to a 14. Leuner C, Dressman J. Improving drug solubility for oral delivery
possible increase in bioavailability. However, this clearly would using solid dispersions. Eur J Pharm Biopharm 2000;50:47–60.
not be ideal from a regulatory perspective. 15. Serajuddin ATM. Solid dispersion of poorly water-soluble drugs:
early promises, subsequent problems, and recent breakthroughs.
J Pharm Sci 1999;88:1058–1066.
Conclusions 16. Rumondor ACF, Stanford LA, Taylor LS. Effects of polymer type
By studying the crystallization behavior of curcumin, additional and storage relative humidity on the kinetics of felodipine crystal-
insight into the role of compound–polymer interaction on lization from amorphous solid dispersions. Pharm Res 2009;26:
2599–2606.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

crystallization tendency has been gained. Intramolecular bonding 17. Rumondor ACF, Marsac PJ, Stanford LA, Taylor LS. Phase behavior
in curcumin reduced the extent of hydrogen bonding between the of poly(vinylpyrrolidone) containing amorphous solid dispersions in
polyphenol and polymers. Thus, for a compound like curcumin, the presence of moisture. Mol Pharmaceutics 2009;6:1492–1505.
hydrogen bonding with polymers is impeded, and it is more 18. Van Eerdenbrugh B, Taylor LS. Small scale screening to determine
difficult to inhibit crystallization. In contrast, ionic interactions the ability of different polymers to inhibit drug crystallization upon
with E100 were not hindered, and when formed, resulted in rapid solvent evaporation. Mol Pharm 2010;7:1328–1337.
increased physical stability. This suggests that the best polymers 19. Van Eerdenbrugh B, Baird JA, Taylor LS. Crystallization tendency
of active pharmaceutical ingredients following rapid solvent evap-
for amorphous solid dispersion of compounds prone to intramo-
oration – classification and comparison with crystallization tendency
lecular hydrogen bonding are those that allow the formation of from undercooled melts. J Pharm Sci 2010;99:3826–3838.
interactions other then hydrogen bonds (e.g. ionic interactions). 20. Kestur US, Taylor LS. Role of polymer chemistry in influencing
crystal growth rates from amorphous felodipine. Cryst Eng Comm
Acknowledgements 2010;12:2390–2397.
21. Kestur US, Van Eerdenbrugh B, Taylor LS. Influence of polymer
We gratefully acknowledge Alicia Winters for her help in preparing chemistry on crystal growth inhibition of two chemically diverse
samples and conducting PXRD and IR measurements. organic molecules. Cryst Eng Comm 2011;13:6712–6718.
22. Van Eerdenbrugh B, Taylor LS. An ab initio polymer selection
methodology to prevent crystallization in amorphous solid disper-
For personal use only.

Declaration of interest sions by application of crystal engineering principles. Cryst Eng


Comm 2011;13:6171–6178.
The authors report no declarations of interest. 23. Wegiel LA, Mauer LJ, Edgar KJ, Taylor LS. Crystallization of
We thank the USDA (Grant number 09-35603-05068) and Purdue amorphous solid dispersions of resveratrol during preparation and
Research Foundation for financial support. We thank Eastman Chemical storage – impact of different polymers. J Pharm Sci 2013;102:
Company and Shin-Etsu Ltd. for their gracious donations of CMCAB and 171–184.
HPMCAS, respectively. 24. Jasim F, Ali F. A novel and rapid method for the spectrofluorometric
determination of curcumin in curcumin spices and flavors.
References Microchem J 1992;46:209–214.
25. Marsac PJ, Shamblin SL, Taylor LS. Theoretical and practical
1. Sharma OP. Antioxidant activity of curcumin and related com- approaches for prediction of drug-polymer miscibility and solubility.
pounds. Biochem Pharmacol 1976;25:1811–1812. Pharm Res 2006;23:2417–2426.
2. Motterlini R, Foresti R, Bassi R, Green CJ. Curcumin, an 26. Taylor LS, Zografi G. Spectroscopic characterization of interactions
antioxidant and anti-inflammatory agent, induces heme oxygenase-
between PVP and indomethacin in amorphous molecular disper-
1 and protects endothelial cells against oxidative stress. Free Radical
sions. Pharm Res 1997;14:1691–1698.
Biol Med 2000;28:1303–1312.
27. Parimita SP, Ramshankar YV, Suresh S, Row TNG.
3. Apisariyakul A, Vanittanakom N, Buddhasukh D. Antifungal
Redetermination of curcumin: (1E,4Z,6E)-5-hydroxy-1,7-
activity of turmeric oil extracted from Curcuma longa
bis(4-hydroxy-3-methoxyphenyl)hepta-1,4,6-trien-3-one. Acta
(Zingiberaceae). J Ethnopharmacol 1995;49:163–169.
4. Sadzuka Y, Nagamine M, Toyooka T, et al. Beneficial effects of Crystallographica Section E-Structure Reports Online 2007;63:
curcumin on antitumor activity and adverse reactions of doxorubi- O860–O862.
cin. Int J Pharm 2012;432:42–49. 28. Sanphui P, Goud NR, Khandavilli UBR, et al. New polymorphs of
5. Aggarwal BB, Kumar A, Bharti AC. Anticancer potential of curcumin. Chem Commun 2011;47:5013–5015.
curcumin: preclinical and clinical studies. Anticancer Res 2003; 29. Gilli G, Bellucci F, Ferretti V, Bertolasi V. Evidence for resonance-
23:363–398. assisted hydrogen-bonding from crystal-structure correlations on the
6. Vlietinck AJ, De Bruyne T, Apers S, Pieters LA. Plant-derived enol form of the beta-diketone fragment. J Am Chem Soc 1989;111:
leading compounds for chemotherapy of human immunodeficiency 1023–1028.
virus (HIV) infection. Planta Med 1998;64:97–109. 30. Benassi R, Ferrari E, Lazzari S, et al. Theoretical study on
7. Sergent T, Vanderstraeten J, Winand J, et al. Phenolic compounds curcumin: a comparison of calculated spectroscopic properties with
and plant extracts as potential natural anti-obesity substances. Food NMR, UV-vis and IR experimental data. J Mol Struct 2008;892:
Chem 2012;135:68–73. 168–176.
8. Miriyala S, Panchatcharam M, Rengarajulu P. Cardioprotective 31. Tang XC, Pikal MJ, Taylor LS. A spectroscopic investigation of
effects of curcumin. Adv Exp Med Biol 2007;595:359–377. hydrogen bond patterns in crystalline and amorphous phases in
9. Tonnesen HH, Masson M, Loftsson T. Studies of curcumin and dihydropyridine calcium channel blockers. Pharm Res 2002;19:
curcuminoids. XXVII. Cyclodextrin complexation: solubility, chem- 477–483.
ical and photochemical stability. Int J Pharm 2002;244:127–135. 32. Moustafine RI, Zaharov IM, Kemenova VA. Physicochemical
10. Wang Y-J, Pan M-H, Cheng A-L, et al. Stability of curcumin in characterization and drug release properties of Eudragit E PO/
buffer solutions and characterization of its degradation products. Eudragit L 100-55 interpolyelectrolyte complexes. Eur J Pharm
J Pharm Biomed Anal 1997;15:1867–1876. Biopharm 2006;63:26–36.
11. Priyadarsini KI. Photophysics, photochemistry and photobiology of 33. Miyazaki T, Yoshioka S, Aso Y. Physical stability of amorphous
curcumin: studies from organic solutions, bio-mimetics and living acetanilide derivatives improved by polymer excipients. Chem
cells. J Photochem Photobiol C-Photochem Rev 2009;10:81–95. Pharmaceut Bull 2006;54:1207–1210.
986 L. A. Wegiel et al. Pharm Dev Technol, 2014; 19(8): 976–986

34. Weuts I, Kempen D, Verreck G, et al. Salt formation in solid 41. Guzman ML, Manzo RH, Olivera ME. Eudragit E100 as a drug
dispersions consisting of polyacrylic acid as a carrier and three basic carrier: the remarkable affinity of phosphate ester for dimethyla-
model compounds resulting in very high glass transition tempera- mine. Mol Pharm 2012;9:2424–2433.
tures and constant dissolution properties upon storage. Eur J Pharm 42. Quinteros DA, Rigo VR, Kairuz AEJ, et al. Interaction between
Sci 2005;25:387–393. a cationic polymethacrylate (Eudragit E100) and anionic drugs.
35. Yoo SU, Krill SL, Wang Z, Telang C. Miscibility/stability consid- Eur J Pharm Sci 2008;33:72–79.
erations in binary solid dispersion systems composed of functional 43. Van den Mooter G, Wuyts M, Blaton N, et al. Physical stabilization
excipients towards the design of multi-component amorphous of amorphous ketoconazole in solid dispersions with polyvinylpyr-
systems. J Pharm Sci 2009;98:4711–4723. rolidone K25. Eur J Pharm Sci 2001;12:261–269.
36. Kuswandi B, Jayus TS, Larasati A, et al. Real-time monitoring of 44. Lu Q, Zografi G. Phase behavior of binary and ternary amorphous
shrimp spoilage using on-package sticker sensor based on natural mixtures containing indomethacin, citric acid, and PVP. Pharm Res
dye of curcumin. Food Anal Methods 2012;5:881–889. 1998;15:1202–1206.
37. Etter MC. Encoding and decoding hydrogen-bond patterns of 45. Tong P, Zografi G. Solid-state characteristics of amorphous
organic compounds. Acc Chem Res 1990;23:120–126. sodium indomethacin relative to its free acid. Pharm Res 1999;16:
38. Etter MC. Hydrogen bonds as design elements in organic chemistry. 1186–1192.
J Phys Chem 1991;95:4601–4610. 46. Tong P, Zografi G. A study of amorphous molecular dispersions of
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Maastricht on 06/09/14

39. Menjoge AR, Kulkarni MG. Mechanistic investigations of phase indomethacin and its sodium salt. J Pharm Sci 2001;90:1991–2004.
behavior in Eudragit (R) blends. Int J Pharm 2007;343:106–121. 47. Fox TG. Influence of diluent and of copolymer composition on
40. Bernabe-Pineda M, Ramirez-Silva MT, Romero-Romo M, et al. the glass temperature of a polymer system. Bull Am Phys Soc 1956;
Determination of acidity constants of curcumin in aqueous solution 1:123.
and apparent rate constant of its decomposition. Spectrochim Acta A 48. Shamblin SL, Taylor LS, Zografi G. Mixing behavior of colyophi-
Mol Biomol Spectrosc 2004;60:1091–1097. lized binary systems. J Pharm Sci 1998;87:694–701.
For personal use only.

You might also like