You are on page 1of 8

International Journal of Pharmaceutics 453 (2013) 8087

Contents lists available at SciVerse ScienceDirect

International Journal of Pharmaceutics


journal homepage: www.elsevier.com/locate/ijpharm

A theoretical and spectroscopic study of co-amorphous naproxen and


indomethacin
Korbinian Lbmann a , Riikka Laitinen a , Holger Grohganz b , Clare Strachan a ,
Thomas Rades a,b, , Keith C. Gordon c
a

School of Pharmacy, University of Otago, Dunedin, New Zealand


Department of Pharmacy, University of Copenhagen, Copenhagen, Denmark
c
Department of Chemistry and MacDiarmid Institute for Advanced Materials and Nanotechnology, University of Otago, Dunedin, New Zealand
b

a r t i c l e

i n f o

Article history:
Received 6 March 2012
Received in revised form 7 May 2012
Accepted 11 May 2012
Available online 18 May 2012
Keywords:
Density functional theory
Quantum mechanical chemistry
Co-amorphous
Naproxen-indomethacin
Infrared spectroscopy

a b s t r a c t
Co-amorphous drug systems were recently introduced as potential drug delivery systems for poorly
water soluble drugs in order to overcome problems associated with amorphous materials. The improved
physical stability and dissolution of these systems was attributed to molecular interactions between the
co-amorphous partners, such as hydrogen bonds. However, molecular level characterization with vibrational spectroscopy of even the amorphous drugs alone presents a signicant challenge. This becomes
even more complicated when more than one compound is present in the material under investigation.
In this study, the co-amorphous drug mixture containing naproxen (NAP) and indomethacin (IND) was
investigated using infrared spectroscopy (IR) and quantum mechanical calculations. The structures of
both drugs were optimized as monomer, homodimer and heterodimer using density functional theory
and used for the calculation of IR spectra. Conformational analysis conrmed that the optimized structures were suitable for the theoretical prediction of the spectra. Vibrational modes from the calculation
could be matched with experimentally observed spectra for crystalline and amorphous NAP and IND,
and it could be shown that both drugs exist as homodimers in their respective individual amorphous
form. With the results from the experimental single amorphous drugs and theoretical homodimers, a
detailed analysis of the experimental co-amorphous and theoretical heterodimer spectra was performed
and evaluated. It is suggested that NAP and IND exist as heterodimers in the co-amorphous mixture when
quench cooled together from the melt in a 1:1 molar ratio.
2012 Elsevier B.V. All rights reserved.

1. Introduction
Low aqueous solubility is a major concern for many drugs,
and this problem is likely to increase in the future (Aaltonen
and Rades, 2009; Mllertz et al., 2010). Poor water solubility is
of great importance, as it may result in a low bioavailability of
the drug especially when formulated as solid dosage forms (e.g.
tablets, capsules) since their bioavailability relies mainly on the
aqueous dissolution of the drug in the GI tract. Transforming a
crystalline drug into its more soluble amorphous counterpart is
one way to overcome this limitation (Aaltonen and Rades, 2009;
Hancock and Zogra, 1997). In comparison to the well-dened
three-dimensional order in the crystal lattice, the amorphous state
is characterized by the absence of a long range molecular order.
Nevertheless, molecules can still exhibit short range order, for

Corresponding author at: School of Pharmacy, University of Otago, P.O. Box 913,
Dunedin 9054, New Zealand. Tel.: +64 3 479 5410; fax: +64 3 479 7034.
E-mail addresses: thomas.rades@otago.ac.nz, thor@farma.ku.dk (T. Rades).
0378-5173/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ijpharm.2012.05.016

example by forming dimers through hydrogen bonding (Taylor and


Zogra, 1997).
Recently, co-amorphous systems have been found to be a
promising way to stabilize the amorphous state (Alles et al., 2009;
Chieng et al., 2009; Lbmann et al., 2011, 2012). By the combination
of two drugs to form a single phase amorphous system, the physical stability and aqueous dissolution could be improved over the
individual amorphous drugs. Intermolecular interactions between
the two drugs present in the system have been suggested to be
responsible for this improvement. In our recent study on the coamorphous drugdrug combination between the two non-steroidal
anti inammatory drugs indomethacin (IND) and naproxen (NAP)
it was shown that the co-amorphous 1:1 molar ratio exhibited a
signicant higher physical stability than the respective amorphous
individual drugs and co-amorphous mixtures at molar ratios of 2:1
and 1:2 (Lbmann et al., 2011). Furthermore, intrinsic dissolution
showed a synchronized release for the co-amorphous drugs at the
1:1 ratio and was signicantly faster in comparison to the individual
crystalline or amorphous (in case of IND) compounds. The dissolution rate of individual amorphous NAP could not be established

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087

due to the low physical stability of amorphous NAP. It was suggested that the changes in dissolution behavior might be explained
by the formation of a heterodimer between both drugs, and that
the heterodimer was responsible for the synchronized release and
the increased physical stability. The synchronized release might
also be responsible for the faster dissolution behavior of IND in the
co-amorphous mixture.
In order to obtain information on potential interactions in amorphous systems, vibrational spectroscopy provides insight into the
molecular arrangement in amorphous systems. As spectroscopic
methods probe the molecular level, changes in the molecular environment (e.g., H-bonding or interactions) due to solid state
changes may be detected as a result of altered vibrations of functional groups involved in these interactions (Chieng et al., 2011).
However, these changes might only be minor or lost in the complexity of the spectra.
Recently, quantum mechanical chemistry has become increasingly popular in the pharmaceutical eld, especially due to its
usefulness in the interpretation of vibrational spectra (Gordon et al.,
2007). In addition, quantum mechanical calculations have been
applied for example to determine the molecular structure with
respect to the conformational minima of a drug molecule (Borba
et al., 2009), however in most of the cases analysis has been simply applied for band assignment of the experimental infrared (IR)
and Raman spectra of drugs (Ali et al., 2007; Borba et al., 2009; Hu
et al., 2010; Iliescu et al., 2004; Jubert et al., 2006; Mishra et al.,
2008; Sagdinc et al., 2007; Srivastava et al., 2010). Other studies have focused on spectra of various crystalline, polymorphic
forms of drug molecules. Spectral differences in carbamazepine
polymorphs I and III, for example, could be correlated with differences in the hydrogen bonding in the respective polymorphs
(Strachan et al., 2004). Similar ndings were obtained for chlorpropamide polymorphs. Again, differences were associated mainly
to intermolecular interactions rather than to a different molecular
conformation (Chesalov et al., 2008). The formation of dimers in the
solid state has also been shown for other crystalline drugs, such
as ibuprofen (Vueba et al., 2008), indomethacin (Strachan et al.,
2007), ketoprofen (Vueba et al., 2006), and olanzapine (Ayala et al.,
2006). Another study on budesonide revealed the epimeric contribution of the R and S epimer in the IR spectra (Ali et al., 2010). For
theophylline, caffeine and theobromine, hydrogen bonding could
be determined in anhydrate and hydrate forms with the help of
computational chemistry (Nolasco et al., 2006). In addition to the
computational and spectroscopic approach, XRPD was used in most
of the studies to probe the crystal lattice as an additional tool. Thus,
an insight on the molecular arrangement and potential molecular
interactions within a crystal could be identied.
Especially in the case of amorphous materials however, quantum mechanical calculations can become an invaluable tool to
gain further understanding of the near range order of amorphous systems and to help interpret and support experimentally
obtained spectra. Strachan et al. for example have shown that
the dimer structure in crystalline -indomethacin undergoes only
little disruption upon transformation into its amorphous phase
(Strachan et al., 2007). In comparison, for fenobrate, which is a
non hydrogen-bonded drug molecule with only weak intermolecular interactions (i.e. interactions) in its crystalline state, Heinz
et al. (2009a) suggested that these weak interactions can easily be
disturbed, and the disruption of conformation and molecular orientation of fenobrate is likely to be more random in its amorphous
state.
All the above mentioned computational studies dealt only with
discrete drug molecules of one kind and their respective molecular
near range order in various polymorphic forms or in the amorphous
state. The situation becomes more complex in mixed systems, for
example in co-amorphous drug combinations. Little is known about

81

Fig. 1. Molecular structures with numbering depicted for IND (A) and NAP (B).
Subscript letters i and n are used to differentiate the atoms of both drugs.

the molecular arrangement of these systems and quantum mechanical calculations could provide a clearer insight.
In this study, quantum mechanical chemistry was used to gain
a better understanding on the near range order in co-amorphous
IND and NAP. The main interest was to investigate the possible
formation of a heterodimer. Firstly, a geometrical optimization of
the drug molecules as single molecules, homodimers (INDIND
and NAPNAP) and heterodimer (INDNAP) was conducted. Secondly, calculations of the vibrational bands and band assignment
were performed and compared the experimentally obtained FTIR
spectra. To our knowledge, this is the rst time that quantum
mechanical calculations in combination with FTIR spectroscopy are
used to interpret the molecular near range order of a co-amorphous
drug combination.
2. Materials and methods
2.1. Materials
(IND,

polymorph;
M = 357.79 g/mol;
Indomethacin
Tm = 162.0 C) and naproxen (NAP, M = 230.26 g/mol; Tm = 158.1 C)
were sourced from SigmaAldrich, USA and Divis Laboratories, Ltd.
USA, respectively. Molecular structures and numbering of the drug
molecules are depicted in Fig. 1. The numbering system follows

82

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087

that used by Kim and Song (1984) and Kistenmacher and Marsh
(1972) for NAP and IND, respectively. Subscript letters i (for IND)
and n (for NAP) are added in the numbering system in order to
differentiate the individual atoms of both drugs.

2.2. Methods
2.2.1. Sample preparation
Amorphous samples (IND, NAP and a 1:1 molar mixture of both
drugs) were prepared by melting them in aluminum dishes (volume approx. 100 mL; bottom diameter: 8 cm) in a preheated oven
at 168 C for 5 min and subsequent quench cooling by pouring liquid
nitrogen onto the samples. Briey, the sample dishes containing the
molten compounds were placed in a desiccator containing phosphorus pentoxide, followed by pouring liquid nitrogen into the
dishes. The desiccator was immediately closed afterwards in order
to avoid moisture sorption onto the sample. The liquid nitrogen was
allowed to evaporate before proceeding with the measurements of
the infrared spectra.
Before quench cooling of the 1:1 molar ratio, a physical mixture
was prepared by gently mixing 1000 mg of the drugs (1:1 molar
ratio; 608.1 mg IND and 391.6 mg NAP) with mortar and pestle for
60 s.

2.2.2. Attenuated total reectance FT-infrared (ATR-FTIR)


spectroscopy
IR spectroscopy was performed using a Varian 3100 FTIR (Excalibur series) attached with an attenuated total reectance accessory
(GladiATR, Piketech, WI, USA). ATR-FTIR is a suitable set-up for
structure analysis despite the fast recrystallization of NAP at room
temperature, as the ATR crystal and surrounding plate could be
cooled down with ice prior to the measurements to ensure experimental conditions below the glass transition temperature of NAP
(Tg = 5.04 C) (Lbmann et al., 2011). Spectra of the crystalline drugs
were obtained by using the powdered sample as received. Spectra
were recorded over a range of 4000400 cm1 and a mean of 64
scans at a resolution of 4 cm1 . Data analysis was performed on
Varian Resolution Pro, v4.1.0.101 software and graphs were plotted
using Origin (version 8.5.1, OriginLab Corporation, Northampton,
MA, USA).

2.2.3. Computational studies


Quantum mechanical modeling was performed on the individual drug monomers, homodimers and heterodimer (Fig. 2).
As a starting point, the single drug molecules were determined using the bond parameters, angles and dihedral angles
from the respective crystallographic data for NAP and -IND.
For the homodimer calculations, the single drug molecules
were duplicated and arranged to form dimers. For the heterodimer
a similar approach was performed by using one of the individual
drug molecules each.
Calculations were performed using the Gaussian 09 suite of
programs (Frisch et al., 2009). Vibrational modes were determined using density functional theory (DFT) calculations (B3-LYP
functional, 6-31G(d) basis set). The automated calculation by the
software rstly optimized the conformations of the monomer and
dimer structures. Secondly, the vibrational frequency calculations
were determined to obtain the IR intensities. The predicted vibrational modes were scaled with the recommended factor of 0.9614
(Scott and Radom, 1996) using GaussSum 2.2 (OBoyle et al., 2008)
software and visualized using GaussView (v. 5.0, Gaussian, Inc.,
Wallingford, CT, USA).

Fig. 2. Molecular structures with hydrogen bonding between the carboxylic groups
of the IND homodimer (A), NAP homodimer (B), and heterodimer of both drugs (C).

3. Results and discussion


3.1. Conformational analysis
Kim et al. (1987) and Kistenmacher and Marsh (1972) have carried out crystallographic studies on NAP and -IND, respectively.
Their results were taken as a starting point for the optimization of
the two drugs. No major differences in bond length and bond angles
were found in the optimized monomer and dimer structures of both
molecules compared to the respective experimental data obtained
by Kim et al. (1987) and Kistenmacher and Marsh (1972) (data
not shown). Therefore, these geometric parameters of crystalline
NAP and IND were close to the lowest energy as calculated for the
structures. This is not surprising since changes in bond length and
angles are associated with large amounts of energy (Bernstein and
Hagler, 1978), resulting in similar low energy conformation in both

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087

theoretical and experimental data. However, variations in dihedral


angles between the calculated and experimental structures were
observed. This can be explained by the low energy involved in torsional changes, i.e. single bond rotations (Bernstein and Hagler,
1978).
For NAP, the main difference was found in the dihedral angles
along the Cn 2 Cn 9 axis (Table 1), the connection between the
isopropionic acid moiety and the naphthalene structure of the
molecule. Here, the optimized structures were slightly rotated
along this axis when compared to the crystal conformation. The
molecular arrangement in the NAP crystal involves the incorporation in a rigid lattice where each molecule is surrounded by
neighbor molecules and the molecules are arranged to build a
chain through intermolecular hydrogen bonding (Kim and Song,
1984). In comparison, the calculated monomer was not involved
in any hydrogen bonding and the dimer structures only involved
one interaction, the cyclic hydrogen bonding. Further, in the rigid
lattice of the crystal, single bond rotations can occur in order to t
the molecules into the crystal network (Bar and Bernstein, 1985).
With other molecules similar observations with dihedral angles in
quantum mechanical calculations have been reported (Heinz et al.,
2009a; Strachan et al., 2004, 2007).
In our previous study, it was suggested that NAP forms a
homodimer once it is transferred into its amorphous form and a
heterodimer with IND in the co-amorphous form (Lbmann et al.,
2011). Therefore, in this study the focus was on the dimer formation. Unfortunately, it was not possible to optimize the NAP
molecules in a chain-like structure similar to the crystal structure.
The energetic minimum calculation for that conformation changed
the molecular chain arrangement since the software does not take
the crystal lattice into account. Therefore, prediction of an IR spectrum of naproxen arranged in a chain-like manner similar to its
crystal structure could not be modeled.
The torsional angles along the Cn 10 Cn 9 axis for the calculated
NAP monomer and the experimental data were nearly identical
(Table 1, line N5-8). Therefore, the molecules in the crystalline
structure are close to the energetic minimum for this particular
dihedral angle. Upon formation of the dimer structures these angles
change by approximately 5 . Again, this can be explained by the differences in the crystal lattice compared to the dimer model in this
study as mentioned above.
For IND, similar observations on the dihedral angles were made.
Again, the main difference was a single bond rotation, in particular
along the Ci 18 Ci 19 axis (Table 2, line I5 & 6), i.e. the conformation
of the carboxylic acid group. This group is involved in the cyclic
hydrogen bonding of the dimer and differences can be explained by
the same arguments as mentioned above. However, compared to
NAP, the molecules do exist as dimers in the crystal structure of IND (Kistenmacher and Marsh, 1972) and dimers are also thought
to be formed in the amorphous form (Taylor and Zogra, 1997).
Similar ndings are also reported in a computational study on IND
by Strachan et al. (2007).
In order to further investigate the accuracy of the calculations, also the predicted intermolecular distances in the modeled
hydrogen-bonded dimer structures were compared with the crystallographic data. The Oi 3(a) Oi 4(b) distance in the optimized IND
and the On 2(a) On 1(b) distance in the calhomodimer (2.682 A)
are close to the distances in
culated NAP homodimer (2.693 A)
and NAP (2.681 A),
respectively. Intercrystalline IND (2.669 A)
estingly, in the optimized heterodimer both values from the
calculated homodimers appear with molecular distances of 2.682 A
and 2.693 A for Oi 3(a) On 1(b) and On 2(b) Oi 4(a), respectively.
Overall, the geometry of the optimized monomer and dimer
structures of NAP and IND are similar to those in their respective crystalline forms. Small changes were found in the torsional
angles of single bonds for the optimized molecules in the monomer,

83

Fig. 3. IR spectra of experimental crystalline NAP (a), experimental amorphous NAP


(b), calculated NAP homodimer (c), and calculated NAP monomer (d) between 1000
and 1800 cm1 .

homodimer and heterodimers. These changes can be explained by


the difference of the discrete structures in the models and the
rigid molecular environment associated with the crystal lattice.
Conformational changes along single bonds can also be expected
once a crystalline compound is transformed into its amorphous
counterpart due to the loss of long range three dimensional order.
Therefore, the optimized structures are suitable for the calculation
of vibrational spectra.
3.2. Spectroscopic analysis
The DFT prediction of the IR spectra was carried out on the optimized geometry of each structure. The resulting calculated and the
experimental vibrational spectra are given in Figs. 36. Initially
each of the pure compounds was characterized. The calculated
monomer and homodimer structures of each drug were compared
to the experimental crystalline and amorphous spectra. Furthermore, differences in the amorphous and crystalline spectra were
analyzed. Then the experimental co-amorphous and calculated heterodimer spectra were investigated and compared to the individual
amorphous drugs and homodimer predicitions.
In general, the intensities for strong and weak bands in the
calculations are close to those observed in the experimental spectra showing that the predictions are a good model. Furthermore,
the experimental spectra of the amorphous form of each drug
show band shifts and band broadening when compared to the
individual crystalline spectrum. This observation is common with
the crystalline to amorphous transformation of compounds (Heinz
et al., 2009b). For both drugs in this study, specic changes are
detectable in the vibrational regions of the hydrogen bonded carboxylic acid moiety (1700 cm1 region) and the aromatic ring
systems (11001400 cm1 ). Therefore, the spectral region between
1000 and 1800 cm1 was chosen for a detailed analysis. From the
calculated spectra several bands in this region were matched to
bands in the experimental spectra and are given in Tables 3 and 4 for

84

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087

Table 1
Dihedral angles for NAP from literature and after optimization.
Line no.

Dihedral angles, ( )

Kim and Song (1984)

N1
N2
N3
N4
N5
N6
N7
N8

Cn 1
Cn 1
Cn 3
Cn 3
On 1
On 1
On 2
On 2

70.7
53.0
108.9
127.4
88.7
35.2
88.8
147.3

Cn 2 Cn 9 Cn 10
Cn 2 Cn 9 Cn 11
Cn 2 Cn 9 Cn 10
Cn 2 Cn 9 Cn 11
Cn 10 Cn 9 Cn 2
Cn 10 Cn 9 Cn 11
Cn 10 Cn 9 Cn 2
Cn 10 Cn 9 Cn 11

Calculated B3-LYP 6-31G(d)


Monomer

Homodimer

Heterodimer

56.4
66.5
125.0
112.2
89.1
35.0
89.5
146.4

58.2
65.2
123.1
113.4
84.9
39.6
94.2
141.4

59.2
64.4
122.2
114.3
85.2
39.4
93.8
141.7

Table 2
Dihedral angles for IND from literature and after optimization.
Line no.

Dihedral angles, ( )

Kistenmacher and Marsh (1972)

Calculated B3-LYP 6-31G(d)


Monomer

Homodimer

Heterodimer

I1
I2
I3
I4
I5
I6
I7

Ci 2
Ni 1
Oi 1
Ci 2
Ci 3
Ci 3
Ci 5

25.5
144.2
39.3
99.9
146.7
34.9
5.9

29.6
154.4
28.9
88.4
84.7
93.8
0.1

28.9
153.7
29.6
99.1
93.5
85.5
1.23

29.3
153.9
29.4
96.1
89.7
89.0
0.8

Ni 1 Ci 10 Oi 1
Ci 10 Ci 11 Ci 12
Ci 10 Ci 11 Ci 12
Ci 3 Ci 18 Ci 19
Ci 18 Ci 19 Oi 3
Ci 18 Ci 19 Oi 4
Ci 6 Oi 2 Ci 20

Table 3
Selected calculated and experimental wavenumbers of NAP.
Calculated
monomer 
(cm1 )

Calculated
homodimer
 (cm1 )

Calculated
heterodimer
 (cm1 )

Experimental
crystalline 
(cm1 )

Experimental
amorphous 
(cm1 )

Experimental
co-amorphous 
(cm1 )

Vibrational assignment

1135
1186

1227
1186

1224
1187

1193

1210
1194

1196

1220

1220

1218

1226

1228

1224

1256

1256

1256

1263

1263

1261

1602
1764

1601
1710

1603
1718

1603
1681

1604
1697

1605
1703

Out-of-plane On 2 Hn 2 bending, Cn 9H bending


In-plane-naphthalene breathing, out-of-phase
Cn 6 On 3 Cn 12 stretching,
Naphthalene deformation, out-of-phase
Cn 2 Cn 9 Cn 11 stretching, out-of-phase
Cn 6 On 3 Cn 12 stretching
Naphthalene deformation, out-of-phase
Cn 6 On 3 Cn 12 stretching
Cn 2 Cn 9 stretching, naphthalene deformation
Asymmetric On 1 Cn 10 On 2 stretching (carboxylic
acids out of phase in dimer)

NAP and IND, respectively. Since band assignment from quantum


mechanical calculations for NAP (monomer) (Jubert et al., 2006) and
IND (monomer and homodimer) (Strachan et al., 2007) can be found
in literature, only a selection of bands are stated. The vibrational
mode assignments are in good agreement with those mentioned in
literature.

The calculated spectra for NAP homodimer and heterodimer


are most different in the carboxylic acid vibration (Fig. 3). This is
not surprising because a hydrogen bonding interaction is involved
in the modeled dimer compared to the free carboxylic acid in
the monomer. The carboxylic acid vibration in the homodimer
has a 54 cm1 lower wavenumber than in the monomer. The

Table 4
Selected calculated and experimental wavenumbers of IND. These modes and assignments are in good agreement with those reported by Strachan et al. (2007).
Calculated
monomer 
(cm1 )

Calculated
homodimer
 (cm1 )

Calculated
heterodimer
 (cm1 )

Experimental
crystalline 
(cm1 )

Experimental
amorphous 
(cm1 )

Experimental
co-amorphous 
(cm1 )

Vibrational assignment

1111
1214

1271
1216

1269
1218

1291
1222

1289
1219

1288
1216

1246

1246

1246

1261

1259

1261

1298

1298

1298

1306

1314

1316

1604
1693
1779

1604
1692
1723

1603
1693
1718

1614
1689
1714

1608
1680
1708

1605
1680
1703

Out-of-plane Oi 3Hi 3 bending, Ci 12H2 bending


Indole ring deformation, Ci 10 Ci 11 stretching,
out-of-phase Ci 6 Oi 2 Ci 20 stretching, out-of-phase
Ci 2 Ni 1 Ci 9 stretching
Out-of-phase Ci 2 Ni 1 Ci 10 stretching and
out-of-phase Ni 1 Ci 10 Ci 11 stretching, out-of-phase
Ci 6 Oi 2 Ci 20 stretching, in-plane indole and
chlorobenzene ring deformations, Ci 19 Oi 3 stretching
(dimer)
Out-of-phase and in-phase Ci 9 Ni 1 Ci 10 stretching,
in-plane indole and chlorobenzene ring defomrations
Indole ring deformation
Ci 10 Oi 1 stretching
Asymmetric Oi 3 Ci 19 Oi 4 stretching (carboxylic
acids out-of-phase in dimer)

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087

85

Fig. 6. Calculated IR spectra of NAP homodimer (a), NAPIND heterodimer (b), and
IND homodimer (c) between 1000 and 1800 cm1 .

Fig. 4. IR spectra of experimental crystalline IND (a), experimental amorphous IND


(b), calculated IND homodimer (c), and calculated IND monomer (d) between 1000
and 1800 cm1 .

dimer vibration at 1710 cm1 is much closer to the experimentally assigned bands for the hydrogen-bonded crystalline and
amorphous NAP at 1681 cm1 and 1697 cm1 (Fig. 3 and Table 3),
respectively. Therefore, the calculated homodimer represents a
better model for comparison. The experimental spectra show
additional bands at 1725 cm1 and 1728 cm1 in crystalline and
amorphous NAP. These bands are also associated with the hydrogen bonded carbonyls but a likely to represent a different bonded
conformation; hence the spectral shift. It is unlikely that they can
be attributed to the symmetric carbonyl dimer vibration as this is

Fig. 5. Experimental IR spectra of amorphous NAP (a), co-amorphous NAPIND (b),


and amorphous IND (c) between 1000 and 1800 cm1 .

predicted to be at lower frequency and to have much less intensity


than the observed features.
When comparing the experimental crystalline and amorphous
NAP spectra, the solid state transformation into an amorphous
form results in peak shifts for the hydrogen bonded and carboxylic acid vibrations from 1681 cm1 and 1725 cm1 in the
crystal to 1697 cm1 and 1728 cm1 . In addition, the intensities of
these peaks differ greatly of each other. These differences strongly
indicate a change in hydrogen bonding upon amorphization and
rearrangement within the NAP molecules.
Upon the NAP homodimer formation, a vibrational mode at
1227 cm1 is predicted in the calculation (Fig. 3). This mode can be
attributed to hydrogen bonded out-of-plane On 2 Hn 2 and Cn 9H
bending (Table 3). The calculated mode for this vibration in the
monomer is at 1135 cm1 . Interestingly, the experimental spectrum of amorphous NAP also generates a new peak in this region
at 1210 cm1 compared to the crystalline spectrum. This further
supports the assumption that amorphous NAP undergoes a change
in hydrogen bonding with the formation of dimers.
Similar to NAP, the hydrogen bonded IND dimer gave a better prediction of the experimental spectra than the IND monomer
(Fig. 4 and Table 4). The calculated carboxylic acid vibration shifts
from 1779 cm1 in the monomer to 1723 cm1 in the homodimer.
In the experimental spectra for this particular vibration, a downward shift from 1714 cm1 (crystalline) to 1708 cm1 (amorphous)
was detected. Furthermore, a shoulder at 1734 cm1 appears in the
amorphous spectrum which represents a fraction of carbonyl in
a different hydrogen bonded environment and the benzoyl C O
vibration shifts from 1689 cm1 to 1680 cm1 upon amorphization.
Since the benzoyl C O is not involved any interactions in the model,
the predicted vibrations here are nearly identical in the monomer
and homodimer.
The 1271 cm1 vibration in the calculated IND homodimer corresponds to the hydrogen bonded Oi 3Hi 3 and Ci 18H2 out-of-plane
bending similar to that in NAP. This mode can be matched with
the experimental bands in crystalline (1291 cm1 ) and amorphous
(1289 cm1 ) IND. In comparison, the calculation for the monomer
results in a vibrational mode at 1111 cm1 . This further supports
the hypothesis by Taylor and Zogra (1997) and Strachan et al.
(2007) that IND is forming homodimers in its amorphous form.

86

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087

The experimental spectra for the individual amorphous drugs


and the respective co-amorphous single form are given in Fig. 5.
In general, peak shifts and changes in peak intensities for various bands can be detected when compared to the amorphous
individual drugs. The main differences can be found for the carboxylic acid vibrations. In the co-amorphous mixture this peak
appears, with increased intensity, at an intermediate wavenumber
of 1703 cm1 in between the vibrational modes for the carboxylic
acid in amorphous NAP (1697 cm1 ) and IND (1708 cm1 ). In our
previous study, this change in wavenumber was attributed to
the formation of a heterodimer between both drugs (Lbmann
et al., 2011). The quantum mechanical calculation predicted the
same trend in the change of the vibrational modes (Fig. 6). The
heterodimer carboxylic acid stretching (1718 cm1 ) has an intermediate wavenumber to those predicted for the NAP (1710 cm1 )
and IND homodimer (1723 cm1 ). It is worth mentioning that the
carboxylic acid stretching in the heterodimer resulted in a synchronized and single vibrational mode for both acid groups in NAP
and IND. This nding strongly supports the assumption of a heterodimer formation.
Furthermore, the calculated vibrations for the Oi 3Hi 3
(1271 cm1 ) and On 2Hn 2 (1227 cm1 ) bendings in the homodimers were predicted to shift slightly to lower wavenumbers in
the heterodimer, i.e. to 1269 cm1 and 1224 cm1 , respectively.
In the experimental spectra, the Oi 3Hi 3 band in amorphous IND
(1289 cm1 ) was found at 1288 cm1 in the co-amorphous form.
However, the change in wavenumber is below the spectral resolution. Unfortunately, the respective peak for the On 2Hn 2 bending
in amorphous NAP (1210 cm1 ) could not be matched with certainty to a peak in the co-amorphous spectrum. A broad peak
at 1216 cm1 with two shoulders at 1196 cm1 and 1224 cm1
appeared in the co-amorphous spectra. In general, major peak
shifts were detectable for the experimental spectra in the area from
1194 to 1263 cm1 . Since this area is also responsible for aromatic
ring vibrations of both drugs (Tables 3 and 4), this also suggests
a change in the molecular environment for the NAP naphthalene
and IND indol and chlorobenzene rings upon the formation of a
co-amorphous phase.
A similar, but small, change for the indol (1216 cm1 ) and
naphthalene (1220 cm1 ) vibrations was also detected in the calculations resulting in a vibrational mode at 1218 cm1 for the
heterodimer (Fig. 6).
The experimental vibration for the benzoyl C O in amorphous
IND (1680 cm1 ) is the same in the co-amorphous spectrum (Fig. 5).
The observation that the band has not shifted suggests that the C O
is not involved in any hydrogen bonding. The calculations are consistent with this; there is no predicted linkage involving the C O
group and the vibrational modes for the benzoyl C O in the IND
homodimer and heterodimer are at similar predicted wavenumbers.
In our previous work, IR spectroscopy was used to identify interactions between IND and NAP. It was suggested that homodimers
would form in the amorphous state of the single compounds. In
contrast, when NAP and IND are quench-cooled together, a heterodimer is formed in the single phase co-amorphous mixture. The
results presented here further support this hypothesis.
It can be discussed whether the observed results are due to
the formation of a heterodimer or simply due to a mixture of
homodimers, as the presence of a mixture of IND and NAP
homodimers will give spectra similar to the heterodimer. The calculation of the energetics upon the formation of the various homoand hetero-dimer forms is inconclusive in this point as it shows
that dimer formation is favorable in all cases both homodimer
and heterodimer are energetically favorable by about 20 kcal/mol.
The energetics of formation of the dimer were determined by
comparing the sum of electronic and thermal free energies in

the frequency output le for the respective dimer with that of


the reactant monomers (Ochterski, 2000). However, although the
experimental data depicted in Fig. 5 show many similarities, the
experimental data for the co-amorphous mixture is not a simple
linear combination of spectra of the homodimers. This is supportive
of the conclusion that the spectrum is indeed due to the formation
of a heterodimer.
The formation of a heterodimer in a single phase co-amorphous
mixture also suggests the possibility of a co-crystal formation in the
same molecular arrangement upon recrystallization. However, this
was not observed in the case of co-amorphous NAP and IND. Upon
recrystallization, X-ray powder diffraction revealed the presence of
both crystalline compounds indicating the inability of both drugs
to crystallise into a co-crystal and the separation back into single
crystalline compounds (data not shown). This nding suggests, that
even if the formation of a co-crystal is not successful, there is still
the possibility of creating a single phase co-amorphous mixture of
two small molecular compounds upon molecular interactions.
4. Conclusion
In this study, quantum mechanical calculations were successfully applied to investigate the structure and IR spectra of NAP
and IND as individual amorphous drugs and in a co-amorphous
(1:1 molar ratio) combination after quench cooling. The analysis
of the vibrational modes of the individual drugs suggested the formation of homodimers in the respective single amorphous solid
state forms. Furthermore, characterization of the experimental coamorphous IR spectrum and the quantum mechanical calculation
of the heterodimer provided insight into the molecular structure
in co-amorphous NAPIND. In particular, the formation of a heterodimer between NAP and IND when quench cooled together
could be conrmed by the use of density functional theory calculations. This study highlights the usefulness of quantum mechanical
calculations as a valuable tool to improve the understanding of
structural properties of amorphous and most importantly coamorphous systems.
Acknowledgements
The authors would like to thank Ms. A.B.S. Elliot (Dept. of Chemistry, University of Otago) for her help in conducting the DFT
calculations. The Academy of Finland, Magnus Ehrnrooth Foundation and Saastamoinen Foundation are also acknowledged for
funding (RL).
References
Aaltonen, J., Rades, T., 2009. Towards physico-relevant dissolution testing: the
importance of solid-state analysis in dissolution. Dissolution Technol. 16, 4754.
Ali, H.R.H., Edwards, H.G.M., Kendrick, J., Munshi, T., Scowen, I.J., 2007. Vibrational
spectroscopic study of budesonide. J. Raman Spectrosc. 38, 903908.
Ali, H.R.H., Edwards, H.G.M., Kendrick, J., Munshi, T., Scowen, I.J., 2010. An experimental and computational study on the epimeric contribution to the infrared
spectrum of budesonide. Drug Test. Anal. 2, 447451.
Alles, M., Chieng, N., Rehder, S., Rantanen, J., Rades, T., Aaltonen, J., 2009. Enhanced
dissolution rate and synchronized release of drugs in binary systems through
formulation: amorphous naproxen-cimetidine mixtures prepared by mechanical activation. J. Control. Release 136, 4553.
Ayala, A.P., Siesler, H.W., Boese, R., Hoffmann, G.G., Polla, G.I., Vega, D.R., 2006. Solid
state characterization of olanzapine polymorphs using vibrational spectroscopy.
Int. J. Pharm. 326, 6979.
Bar, I., Bernstein, J., 1985. Conformational polymorphism VI: the crystal and
molecular structures of form II, form III, and form V of 4-amino-N-2pyridinylbenzenesulfonamide (sulfapyridine). J. Pharm. Sci. 74, 255263.
Bernstein, J., Hagler, A.T., 1978. Conformational polymorphism. The inuence of
crystal structure on molecular conformation. J. Am. Chem. Soc. 100, 673681.
Borba, A., Gmez-Zavaglia, A., Fausto, R., 2009. Molecular structure, infrared spectra,
and photochemistry of isoniazid under cryogenic conditions. J Phys. Chem. A
113, 92209230.

K. Lbmann et al. / International Journal of Pharmaceutics 453 (2013) 8087


Chesalov, Y.A., Baltakhinov, V.P., Drebushchak, T.N., Boldyreva, E.V., Chukanov, N.V.,
Drebushchak, V.A., 2008. FT-IR and FT-Raman spectra of ve polymorphs of
chlorpropamide. Experimental study and ab initio calculations. J. Mol. Struct.
891, 7586.
Chieng, N., Aaltonen, J., Saville, D., Rades, T., 2009. Physical characterization and stability of amorphous indomethacin and ranitidine hydrochloride binary systems
prepared by mechanical activation. Eur. J. Pharm. Biopharm. 71, 4754.
Chieng, N., Rades, T., Aaltonen, J., 2011. An overview of recent studies on the analysis
of pharmaceutical polymorphs. J. Pharm. Biomed. Anal. 55, 618644.
Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R.,
Scalmani, G., Barone, V., Mennucci, B., Petersson, G.A., Nakatsuji, H., Caricato,
M., Li, X., Hratchian, H.P., Izmaylov, A.F., Bloino, J.Z., Zheng, G., Sonnenberg, J.L.,
Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T.,
Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J.J.A., Peralta, J.E., Ogliaro,
F., Bearpark, M., Heyd, J.J., Brothers, E., Kudin, K.N., Staroverov, V.N., Kobayashi,
R., Normand, J., Raghavachari, K., Rendell, A., Burant, J.C., Iyengar, S.S., Tomasi,
J., Cossi, M., Rega, N., Millam, N.J., Klene, M., Knox, J.E., Cross, J.B., Bakken, V.,
Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R.E., Yazyev, O., Austin, A.J.,
Cammi, R., Pomelli, C., Ochterski, J.W., Martin, R.L., Morokuma, K., Zakrzewski,
V.G., Voth, G.A., Salvador, P., Dannenberg, J.J., Dapprich, S., Daniels, A.D., Farkas,
., Foresman, J.B., Ortiz, J.V., Cioslowski, J., Fox, D.J., 2009. Gaussian 09, Revision
A.1. Gaussian, Inc., Wallingford, CT.
Gordon, K.C., McGoverin, C.M., Strachan, C.J., Rades, T., 2007. The use of quantum chemistry in pharmaceutical research as illustrated by case studies of
indometacin and carbamazepine. J. Pharm. Pharmacol. 59, 271277.
Hancock, B.C., Zogra, G., 1997. Characteristics and signicance of the amorphous
state in pharmaceutical systems. J. Pharm. Sci. 86, 112.
Heinz, A., Gordon, K.C., McGoverin, C.M., Rades, T., Strachan, C.J., 2009a. Understanding the solid-state forms of fenobrate a spectroscopic and computational
study. Eur. J. Pharm. Biopharm. 71, 100108.
Heinz, A., Strachan, C.J., Gordon, K.C., Rades, T., 2009b. Analysis of solid-state transformations of pharmaceutical compounds using vibrational spectroscopy. J.
Pharm. Pharmacol. 61, 971988.
Hu, Y., Erxleben, A., Ryder, A.G., McArdle, P., 2010. Quantitative analysis of sulfathiazole polymorphs in ternary mixtures by attenuated total reectance infrared,
near-infrared and Raman spectroscopy. J. Pharm. Biomed. Anal. 53, 412420.
Iliescu, T., Baia, M., Kiefer, W., 2004. FT-Raman, surface-enhanced Raman spectroscopy and theoretical investigations of diclofenac sodium. Chem. Phys. 298,
167174.
Jubert, A., Legarto, M.L., Massa, N.E., Tvez, L.L., Okulik, N.B., 2006. Vibrational
and theoretical studies of non-steroidal anti-inammatory drugs Ibuprofen
[2-(4-isobutylphenyl)propionic acid]; naproxen [6-methoxy-[alpha]-methyl-2naphthalene acetic acid] and tolmetin acids [1-methyl-5-(4-methylbenzoyl)1H-pyrrole-2-acetic acid]. J. Mol. Struct. 783, 3451.
Kim, Y., Song, H., 1984. The molecular structure of (+)-6-methoxy--methyl-2naphthaleneacetic acid determined by X-ray method. Arch. Pharmacal Res. 7,
137139.
Kim, Y., Song, H., Park, I., 1987. Renement of the structure of naproxen,
(+)-6-methoxy--methyl-2-naphthaleneacetic acid. Arch. Pharmacal Res. 10,
232238.

87

Kistenmacher, T.J., Marsh, R.E., 1972. Crystal and molecular structure of an


antiinammatory agent, indomethacin, 1-(p-chlorobenzoyl)-5-methoxy-2methylindole-3-acetic acid. J. Am. Chem. Soc. 94, 13401345.
Lbmann, K., Laitinen, R., Grohganz, H., Gordon, K.C., Strachan, C., Rades, T., 2011.
Coamorphous drug systems: enhanced physical stability and dissolution rate of
indomethacin and naproxen. Mol. Pharma. 8, 19191928.
Lbmann, K., Strachan, C., Grohganz, H., Rades, T., Korhonen, O., Laitinen, R., 2012.
Co-amorphous simvastatin and glipizide combinations show improved physical stability without evidence of intermolecular interactions. Eur. J. Pharm.
Biopharm. 81 (1), 159169.
Mishra, S., Chaturvedi, D., Tandon, P., Gupta, V.P., Ayala, A.P., Honorato, S.B., Siesler,
H.W., 2008. Molecular structure and vibrational spectroscopic investigation of
secnidazole using density functional theory. J. Phys. Chem. A 113, 273281.
Mllertz, A., Ogbonna, A., Ren, S., Rades, T., 2010. New perspectives on lipid and
surfactant based drug delivery systems for oral delivery of poorly soluble drugs.
J. Pharm. Pharmacol. 62, 16221636.
Nolasco, M.M., Amado, A.M., Ribeiro-Claro, P.J.A., 2006. Computationally-assisted
approach to the vibrational spectra of molecular crystals: study of hydrogenbonding and pseudo-polymorphism. ChemPhysChem 7, 21502161.
OBoyle, N.M., Tenderholt, A.L., Langner, K.M., 2008. cclib: a library for packageindependent computational chemistry algorithms. J. Comput. Chem. 29,
839845.
Ochterski, J.W., 2000. Thermochemistry in Gaussian. http://www.gaussian.com/
g whitepap/thermo.htm (accessed 25.02.2012).
Sagdinc, S., Kandemirli, F., Bayari, S.H., 2007. Ab initio and density functional computations of the vibrational spectrum, molecular geometry and some molecular
properties of the antidepressant drug sertraline (Zoloft) hydrochloride. Spectrochim. Acta Part A: Mol. Biomol. Spectrosc. 66, 405412.
Scott, A.P., Radom, L., 1996. Harmonic vibrational frequencies: an evaluation
of HartreeFock, MllerPlesset, quadratic conguration interaction, density functional theory, and semiempirical scale factors. J. Phys. Chem. 100,
1650216513.
Srivastava, A., Mishra, S., Tandon, P., Patel, S., Ayala, A.P., Bansal, A.K., Siesler,
H.W., 2010. Molecular structure and vibrational spectroscopic analysis of an
antiplatelet drug; clopidogrel hydrogen sulphate (form 2) a combined experimental and quantum chemical approach. J. Mol. Struct. 964, 8896.
Strachan, C.J., Howell, S.L., Rades, T., Gordon, K.C., 2004. A theoretical and spectroscopic study of carbamazepine polymorphs. J. Raman Spectrosc. 35, 401408.
Strachan, C.J., Rades, T., Gordon, K.C., 2007. A theoretical and spectroscopic
study of -crystalline and amorphous indometacin. J. Pharm. Pharmacol. 59,
261269.
Taylor, L.S., Zogra, G., 1997. Spectroscopic characterization of interactions between
pvp and indomethacin in amorphous molecular dispersions. Pharm. Res. 14,
16911698.
Vueba, M.L., Pina, M.E., Batista de Carvalho, L.A.E., 2008. Conformational stability of
ibuprofen: assessed by DFT calculations and optical vibrational spectroscopy. J.
Pharm. Sci. 97, 845859.
Vueba, M.L., Pina, M.E., Veiga, F., Sousa, J.J., de Carvalho, L.A.E.B., 2006. Conformational study of ketoprofen by combined DFT calculations and Raman
spectroscopy. Int. J. Pharm. 307, 5665.

You might also like