You are on page 1of 14

ADR-12886; No of Pages 14

Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Advanced Drug Delivery Reviews

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

Recent developments in the Raman and infrared investigations of


amorphous pharmaceuticals and protein formulations: A review☆
Alain Hédoux
Université Lille Nord de France, F-59000 Lille, France
USTL UMET UMR, 8207F-59655 Villeneuve d'Ascq, France

a r t i c l e i n f o a b s t r a c t

Article history: The success rate for drug discovery and the development of innovative therapeutic strategies are intimately re-
Received 15 October 2015 lated to the physical properties of the solid-state condensed matter, which have direct influence on the bioavail-
Received in revised form 26 November 2015 ability of Active Pharmaceutical Ingredients. In order to transform a new molecule in efficient drug, the material is
Accepted 30 November 2015
brought into an amorphous state using various manufacturing processes including freeze drying, spray drying,
Available online xxxx
hot melt extrusion and loading in different delivery devices. The infrared and Raman spectroscopic analyses
Keywords:
used for exploring disordered and amorphous states, for the monitoring of the drug physical stability in drug de-
Low-frequency Raman spectroscopy livery systems are described in this review.
Chemometrics © 2015 Elsevier B.V. All rights reserved.
Physical stability
Drug delivery systems

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Description of the different kinds of motions in molecular systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.1. Raman spectrometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.2. Infrared spectrometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Spectrum analysis of molecular compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.1. Analysis of the low-frequency spectrum (LFRS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.2. Analysis of the high frequency IR or Raman spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Spectroscopic signatures of amorphous states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.1. Amorphous solid state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.2. Amorphous liquid state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6. Using LFRS to analyze disordered and amorphous solid states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.1. Structural description of disordered states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.2. Evidence of disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
7. Using IR and Raman spectroscopy for the in-process monitoring and for analyzing solid dispersions . . . . . . . . . . . . . . . . . . . . . . . 0
8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction knowledge about its stability conditions are crucial to respect the drug
dosage. Numerous new drug candidates are characterized by a poor sol-
It is well recognized that the physical state of an active pharmaceu- ubility in crystalline states and require amorphization in special dosage
tical ingredient (API) is closely related to their bioavailability, and forms for enhanced bioavailability [1]. However, the amorphous state is
metastable with respect to the crystalline state, and may spontaneously
☆ This review is part of the Advanced Drug Delivery Reviews theme issue on “SI: recrystallizes under stress induced by manufacturing processes, or by
Amorphous pharmaceutical solids”. storage conditions [2–4] (humidity, temperature). The determination

http://dx.doi.org/10.1016/j.addr.2015.11.021
0169-409X/© 2015 Elsevier B.V. All rights reserved.

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
2 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

of the stability conditions of amorphous formulations, requires to un- disadvantages of one of two techniques, with interesting use in the do-
derstand excipient — API interactions, and also the monitoring through main of pharmaceutical applications. Water vibrations give an intense
the several stages of processing as solid-dosage forms. contribution to the IR spectra, which forces to eliminate the presence
Raman and near infrared (NIR) spectroscopies are recognized as of moisture even in low proportion. Additionally, the analysis of aque-
non-destructive tools for both process analytical technology (PAT) [5, ous solution by IR spectroscopy is highly perturbed by the stretching
6] applications and in quality control laboratories for stability analysis. and bending bands of water. On the other hand, Raman scattering can
Both spectroscopic techniques enable the enhancement of process generate specific excitations when the virtual states are overlapping
knowledge for solid, liquid and biopharmaceutical forms [7, 8]. with excited states (Fig. 1). This can induce interesting enhanced
Combined with chemometric tools, they are very sensitive and com- Raman intensity of specific vibrational bands (resonant Raman scatter-
plementary indirect structural probes, which provide information on ing) or intense interference signal (fluorescence) depending on the
molecular conformation and molecular packing in disordered and com- timescale of the virtual state (~10−14 s for scattering and ~ 10−8 s for
plex systems. fluorescence). Polarizability is proportional to the number of electrons,
Interaction between light and matter can result in absorption, scat- to the molecular size, and inversely proportional to the bond strength.
tering or no interaction of the incident photon with the material Generally, Raman scattering intensity is much larger for π systems
which may pass through it without any change. Absorption and scatter- than σ bonded structures, making Raman spectroscopy very sensitive
ing processes are presented in Fig. 1. Absorption requires that the ener- for analyzing APIs compared with excipients.
gy of the incident photon corresponds to the energy gap between the For many years the IR spectroscopy has been widely used with re-
ground state and a higher energy level of the molecule. The process of spect to the Raman spectroscopy, largely due to the cost of the lasers.
absorption spectroscopy is used in various techniques including infra- Nowadays, the advances in notch-filter and diode laser technologies
red spectroscopy. In this technique, infrared energy covering a range have resulted in new Raman spectrometers accessible to non-expert
of frequencies is used as radiation. By contrast, the scattering phenom- users, and this technique becomes industry standard with advantages
enon does not require that the photon energy matches the difference with respect to IR spectroscopy. The most significant advantage of
between two energy levels of the molecule. Two types (elastic and in- Raman spectroscopy is that no specific sample preparation is required
elastic) of scattering can occur, in which the energy photon is trans- without care of moisture, making possible the analysis of tablets or
ferred to the molecule, promoted in a short-lived (called virtual) powders as obtained from manufacturing process, or almost in-situ
vibrational energy level. In the case of elastic scattering, the photon en- from in line manufacturing process. Raman spectroscopy gives the op-
ergy is spontaneously re-emitted without energy change; this phenom- portunity to investigate low-frequency domains down to ~ 2 cm−1,
enon is called Rayleigh scattering. In the inelastic scattering process the while far-IR spectroscopy is limited to the 100–400 cm−1 range. It is
energy photon is re-emitted with an energy lower or a higher than the an undeniable advantage of Raman spectroscopy which makes it possi-
incident radiation energy, corresponding to the anti-stokes or stokes ble the analysis of low-frequency excitations distinctive of the amor-
Raman scattering. Raman scattering requires a monochromatic radia- phous state. Selecting the use of one of the two spectroscopic
tion from a laser source. Infrared absorption and Raman scattering are techniques was discussed by De Beer et al. [5].
two vibrational spectroscopies providing similar information on molec- In this review, the capabilities of Raman and NIR spectroscopy to
ular vibrations. However, the use of two different types of radiation has carefully describe the physical state of different kinds of disordered sys-
important consequence on the selection rules of the vibrational modes. tems, especially amorphous states and mixed amorphous — (macro,
Infrared absorption requires that a vibrational mode of the molecule in- micro, nano) crystalline states, will be shown. The contribution of IR
duces a change in the dipole moment, i.e. in the charge distribution and Raman spectroscopy to the analysis of protein formulations in solu-
within the molecule, to be IR active. In contrast the electromagnetic ra- tions and in the solid state will be presented. The description of the THz
diation associated with the laser source momentary distorts the elec- spectroscopy methods was deliberately omitted since it's the subject of
tronic cloud distributed around a bond within a molecule. As a result, another review.
the vibrational mode of the bond will be Raman active if the vibration
induces a distortion of the electronic distribution around the bond, 2. Description of the different kinds of motions in molecular systems
and causes large polarization change, i.e. a polarizability change. The
Raman activity of vibrational modes is dependent on parameters Considering a molecule as a rigid body, three different kinds of mo-
influencing polarizability, and marking pronounced differences with tions can exist in disordered organic solids, and only the access of the
IR spectroscopy. Symmetric vibrations induce the largest polarizability low-frequency region (below 100 cm−1) gives the opportunity to ana-
change and give the greatest Raman scattering, while asymmetric vibra- lyze these different kinds of motions.
tions cause more significant change in the charge distribution and then The internal motions correspond to vibrations of interatomic bonds
IR absorption activity. This difference is responsible for advantages or within the molecule. They are mainly distinctive of the molecular con-
formation in the fingerprint region, usually lying between 500 and
1800 cm−1, and corresponding to the region where the vibrations of
the molecular skeleton are active. Generally, a molecule composed of
N atoms has 3 N-6 vibrations, 3 N-5 for linear molecules. If the molecule
is placed in crystalline lattice, the 3 N-6 vibrations generates a number
of Raman and IR vibrational bands higher than 3 N-6, depending both
on the crystalline and the molecular symmetries, i.e. the so-called site
symmetry. The increase of the degree of disorder in a crystal induces
higher lattice symmetry, and thus less and broader vibrational bands.
Consequently, the spectrum of internal vibrations can be used as a sen-
sitive probe of the local molecular environment, i.e. to the degree of dis-
order, by counting the vibrational modes, or/and by analyzing the band-
shape of Raman or IR active modes. The spectrum of an amorphous state
corresponds to the spectrum of the isolated molecule, and then is
usually composed of 3 N-6 bands. Internal Raman or IR spectrum
Fig. 1. Description of Raman effect, fluorescence, resonant Raman effect, IR and NIR makes it possible the evidence of molecular associations via hydrogen
absorption. bonds (H-bonds). Polymorphism in molecular compounds including

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx 3

pharmaceuticals is characterized by different types of molecular associ- use a Nd:YAG laser (1.064 μm) with a near-infrared interferometer
ations via H-bonding. Raman or IR spectroscopy can be used to deter- coupled to either a liquid nitrogen cooled germanium or indium gallium
mine molecular associations existing in the amorphous state in order arsenide detector. The main advantages are (i) the reduction in the
to predict and to analyze the crystallization mechanism. Additionally, number of samples that exhibit laser-induced fluorescence, (ii) the
spectroscopy is a powerful tool to reveal the existence of H-bonding, ease of operation, (iii) and the high spectral resolution with good wave-
taking into account the limit of X-ray diffraction to accurately determine length accuracy. However, the use of FT-Raman spectrometers presents
H-atom positions. the disadvantage of being relatively unsuitable for sample environment
At lower frequencies, two types of bands can be observed which are (temperature, pressure, humidity,…). Dispersive Raman spectrometers
mainly analyzed by Raman spectroscopy, since far-infrared spectrosco- use multi-channel (CCD) detectors, with extremely low intrinsic noise
py is commonly limited down to frequencies around 100 cm−1. and high quantum efficiency. The benefits for dispersive Raman are far
The external motions between a molecule and its nearest molecular higher sensitivities and far lower detection limits than FT-Raman. The
neighboring. In ordered crystalline phases, external motions give rise shorter wavelengths of lasers associated with dispersive Raman in-
to the phonon spectrum corresponding to the lattice vibrations. crease the sensitivity of the spectrometer, because of the ν4 scattering
Disordering induces an inhomogeneous broadening of the phonon efficiency dependency. As a consequence, the data acquisition times
spectrum, and the low-frequency spectrum of the amorphous state is are much shorter than for FT-Raman methods, that is important for
observed to be a very close representation of the vibrational density of the study of fast transformation kinetics. The performances of dispersive
states, usually determined by inelastic neutron scattering. Molecular Raman spectrometers result from the adequacy between a good resolu-
compounds are characterized by the pronounced contrast between tion and rejection of the laser line, and a sufficient scattered intensity,
the strong covalent intramolecular interactions and soft van der Waals imposed by the dispersive system. Modern instrumentation almost uni-
intermolecular interactions or hydrogen bond associations. This con- versally employs notch or edge filters for laser rejection, and spectro-
trast is responsible for the frequency gap between internal and external graph (monochromator or grating). However, in this case the analysis
vibrations. Intermolecular interactions are responsible for specific phys- of the low-frequency range (b150 cm−1) is generally prohibited. The
ical properties of molecular compounds (polymorphism, low melting use of multiple dispersive stages (double and triple monochromators)
temperature,…) while covalent bonds maintain the cohesion within mounted with a long focal length allow the detection of Raman-active
the molecule. Molecular flexibility favors disorder and reduces the gap modes with frequencies as low as 2–5 cm−1. New advances in ultra-
between external and internal vibration modes. narrow band notch filter technology [10] with holographic gratings en-
The semi-internal (or semi-external) motions mainly exist in disor- able measurements of low-frequency spectra from a relative compact,
dered states. These motions usually correspond to large amplitude rota- easy to use and cost effective system [11]. The development of this
tions of the molecule or a group of atoms within the molecule, giving a kind of dispersive system composed of a single stage spectrograph
Raman signature in the very low-frequency range (2–~ 70 cm− 1). opens new perspectives for the qualitative and quantitative analysis of
This low-frequency component corresponds to fast local motions, called the physical state of various pharmaceutical forms by Raman spectros-
β-fast relaxational motions, typically on the picosecond timescale. copy with respect to IR spectroscopy.
These motions are thermally activated and their temperature depen- Depending on the nature of the analysis (large volumes or micro-
dence is closely related with the mechanism of phase transitions. scopic samples, or heterogeneities in multi-component systems) two
The simultaneous investigation of these three types of motions pro- types of installation can be distinguished; the conventional installation
vides structural description of disordered states (crystalline, micro- or the macroanalysis, and the installation micro Raman suitable for mi-
nano crystalline, and amorphous states), and the understanding of the croanalyses. Detailed information on spatial resolution accessible by
mechanism of phase transformation in relation with the molecular con- micro-Raman investigations can be found in a recent review of Smith
formation. This type of analysis requires investigations down to the very et al. [12]. Raman investigations under pressure are typically performed
low frequencies (~ 5–2 cm−1) and is only possible by Raman scattering on microscopic samples located in diamond anvil cells [13, 14] (DACs),
and Terahertz pulsed spectroscopy well detailed in a previous review and then micro-Raman installation is systematically used for pressure
[9]. The development of a new generation of filters makes it possible analyses. For micro Raman, the sample holder is a XYZ table equipped
the access to the low-frequency range with routine dispersive Raman with a microscope composed of high numerical aperture objectives. Mi-
spectrometers. This feature opens important perspectives to Raman croscopic investigations are also very suited to analyze different kinds of
spectroscopy for analyzing the physical state in molecular systems. A heterogeneities in samples with pharmaceutical applications: tablets
significant part of this review will focus on the experimental con- [15], extrudates [16], freeze-dried [17–19] formulations. Online analysis
figuration and on the method of data analysis required to perform can be also performed, by using a fiber optic probe, during the
low-frequency investigations in disordered states. Additionally, new manufacturing process such as holt-melt extrusion [20, 21] or freeze-
technical advances in Raman spectroscopy will be presented with drying [22]. Micro Raman analyses can also be performed using a fiber
their expected contribution to the analysis of the physical state of phar- optic Raman probe to investigate the solid state transformation of hy-
maceutical materials. drate forming drugs, in-situ during dissolution tests [23].

3. Equipment 3.2. Infrared spectrometers

3.1. Raman spectrometers Infrared spectrometers can be schematically described by the basic
components including a light source, a monochromator, a sample and
Raman spectrometers are usually composed of four main compo- a detector. Tungsten lamp, or more generally an inert solid that are heat-
nents: (1) the laser source of monochromatic radiation, (2) a sample il- ed to generate thermal emission of radiation in the infrared region of
lumination system and light connection optics, (3) a device for the the electromagnetic spectrum, are commonly used. The monochroma-
analysis of the scattered light (spectrometer), (4) and a sensitive detec- tor is used to separate the polychromatic broad IR spectrum into narrow
tor such as a charge-coupled device (CCD). Two kinds of spectrometers monochromatic lines. Different optical configurations exist, using grat-
can be used for collecting Raman spectra, the Fourier transform (FT- ing or interferometer corresponding to dispersive or FT spectrometers
Raman) or dispersive Raman spectrometers. The difference between respectively. Detectors convert the analog spectral output into an elec-
both types of spectrometer bases essentially on the device (3): a grating trical signal. Different types of detectors are used characterized by dif-
dispersing system for dispersive Raman spectrometers, or an interfer- ferent spectral domains of sensitivity and acquisition rates. The more
ometer for FT-Raman spectrometers. Most of FT-Raman spectrometers commonly used are silicon, and InGaAS, PbSe, InSb which are more

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
4 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

sensitive in the NIR domain. Detailed descriptions of the different types systems and (ii) on the mechanism of (order–disorder) phase transfor-
of IR spectrometers can be found in specific literature dedicated to infra- mation, confined in a narrow frequency range (3–200 cm−1). The
red spectroscopy [5, 24, 25]. Raman intensity (IRaman(ν, T)) in the LFRS is related to the vibrational
density of states (VDOS, G(ν)) according to [26, 27]
4. Spectrum analysis of molecular compounds
½nðν; TÞ þ 1
IRaman ðν; TÞ ¼ CðνÞGðνÞ ð1Þ
4.1. Analysis of the low-frequency spectrum (LFRS) ν

In a crystalline state characterized by a unit cell composed of Z mol- and


ecules, 6Z external motions are expected including 3 acoustic vibrations
corresponding to collective translation motions of molecules, not de- CðνÞ
χ}ðνÞ ¼ GðνÞ ð2Þ
tectable by Raman spectroscopy. The number of Raman active external ν
motions is usually lower than 6Z-3 and is depending on the crystalline
where C(ν) is the light-vibration coupling coefficient, and χ″(ν) is the
symmetry. LFRS is an indirect structural probe of the long-range order
Raman susceptibility. To obtain the χ″(ν) spectrum (plotted in
in crystalline states. Disorder in the molecular environment of a mole-
Fig. 2b), which contains the structural information, the Raman intensity
cule induces an inhomogeneous broadening of phonon peaks, and the
is firstly transformed into reduced intensity Ir(ν) plotted in Fig. 2a using
LFRS of amorphous states results in the envelope of phonon peaks of
the underlying crystalline state. This broadening is related to the exis-
IRaman ðν; TÞ
tence of different nearest molecular neighborings, i.e. different molecu- Ir ðνÞ ¼ : ð3Þ
½nðν; TÞ þ 1ν
lar cages.
The analysis of the low-frequency spectrum (LFRS) of disordered
molecular materials is very difficult, because of the overlapping contri- In this representation the QES intensity is the dominant contribution
butions of the quasi-elastic intensity (IQES) and the vibrational intensity of the LFRS. This component is usually determined from a fitting proce-
(IVIB), as shown in Fig. 2. However, it can be considered as containing dure using a Lorentzian shape centered at ν = 0. The vibrational compo-
rich information (i) about the molecular organization in disordered nent is usually well described with a lognormal function. After
subtracting the QES component, Ir(ν) can be converted into Raman sus-
ceptibility by:

χ}ðνÞ ¼ ν  Ir ðνÞ: ð4Þ

The frequency dependence of C(ν) has been hotly debated [27–31].


Several studies have converged into a linear dependence of C(ν) sug-
gesting that χ " (ν) ∝ G(ν) [32–34]. This relation was corroborated by
the similarity between the χ″(ν) and G(ν) spectra plotted in Fig. 2b
the molecular glass-forming liquid TriPhenyl Phosphite (TPP) [33].
This figure confirm that χ″(ν) spectra can be considered as a close rep-
resentation of the VDOS, and demonstrates that the spectral resolution
of χ″(ν) spectra is significantly higher than that of G(ν) spectra. Fig. 3a
shows the temperature dependence of χ″(ν) spectra in TPP, plotted in
the liquid, undercooled liquid and glassy states. The temperature de-
pendence over about 100 K is relatively weak, reflecting the quasi-
harmonic behavior of collective motions.
The temperature dependence of the QES component is analyzed
using the fitting procedure described in Fig. 2. This contribution to the
LFRS corresponds to fast anharmonic motions, the so-called β-fast
relaxational motions. The Ir(ν) spectra are normalized by the integrated
intensity of the vibrational component, which has a weak temperature
dependence, and plotted at various temperatures (in the liquid,
undercooled liquid and glassy states) in Fig. 3b. This figure indicates a
strong temperature dependence distinctive of anharmonic motions,
contrasting with Fig. 3a. IQES(T) is similar to the temperature depen-
dence of the mean-square displacement (MSD) [35] bu2N (T), usually
determined by neutron scattering experiments. The temperature de-
pendence of these kinds of anharmonic motions is commonly deter-
mined as the driving force in phase transitions. Fragile molecular
liquids can be distinguished from strong liquids by comparing the tem-
perature dependence of MSD [35], and assuming the empirical relation


1= u2 ðTÞ∝ logηðTÞ ð5Þ

experimentally determined in different glass-forming systems [36, 37].


The temperature dependence of logη is plotted in the inset of Fig. 3b
with the fitting curve corresponding to the Vogel–Fulcher–Tammann
(VFT) function given below:
Fig. 2. Representations of the low-frequency spectrum of triphenyl phosphite: a) reduced
intensity and description of the fitting procedure of experimental spectrum. b) Raman sus-
ceptibility compared with the VDOS (G(ν)) obtained by neutron scattering. expðA  T0 =ðT−T0 ÞÞ ð6Þ

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx 5

the unfolding process in protein formulations, to identify polymorphism


and to distinguish amorphous from crystalline states. Univariate spec-
tral analyses are based on peak height at a specified frequency of the un-
treated Raman data. The selected spectral response must be not biased
by contributions from other sources (presence of water, buffer compo-
nents…). Multivariate data analysis (MVA) takes into account the
whole spectra of information, and is widely used to reveal very weak
changes in the molecular conformation or in the structure of the physi-
cal state.
Data pretreatments are the first stage in MVA. This preprocessing
stage consists to eliminate or to minimize the effect of various uncon-
trollable parameters on the intensity. The baseline correction is com-
monly performed, since its contribution may be different from one
spectrum to another. Many methods exist for baseline correction [39],
second-order derivatives and polynomial fitting, being the most com-
monly used. In a second step a normalization procedure is performed
in order to correct the overall intensity variations which could be in-
duced by weak differences in focusing distances because of the rough-
ness of the sample surface. This correction aims to remove intensity
differences unrelated to the sample composition. The most widely
used normalization procedure is the standard normal variate (SNV)
where a normalized spectrum is obtained by subtracting the mean of
its spectrum and divided by the standard deviation.
MVA consists to used newly defined variables that are linear combi-
nations of the original ones, and are mutually independent. The aim is to
reduce the number of correlated variables existing in IR or Raman spec-
tra in a few uncorrelated variables containing the relevant information
for a qualitative or/and quantitative analysis of inhomogeneous sam-
ples… The best known and probably most widely used method to deter-
mine the number of linear independent components in a given system
is the principal component analysis (PCA).
The basic idea of PCA is to represent the original data matrix D by a
product of two smaller matrices T and P, respectively the scores matrix
and the loadings matrix, according to:

D ¼ T  PT ð7aÞ
Fig. 3. Temperature dependence of the LFRS of TPP. a) In the representation of reduced in-
tensity; the inset shows the temperature behavior of bu2N versus Tg/T. b) In the represen-
tation of Raman susceptibility. schematically presented in Fig. 4a. PCA builds a new coordinate system,
called principal components (PC), as a linear combination of the old co-
ordinates (Fig. 5). D matrix has the dimensions of number of spectra
(n) by intensity at particular frequencies (m). The score matrix has
where A reflects the curvature of the plot distinctive of the fragility of the dimensions of the number of spectra (n) by principal components
the system. (p). The loading matrix has the dimensions of intensity (m) by principal
components (p). The most important information is contained in the
4.2. Analysis of the high frequency IR or Raman spectrum first two or three PCs. The rest of PCs mainly corresponds to noise, and
Fig. 4b describes the decomposition of D matrix depending on the num-
Raman investigations are commonly performed at frequencies ber of significant PCs, according to:
higher than 200 cm−1, i.e. in the domain which does not require any
specific high-dispersive equipment (triple monochromator) used for D ¼ T  PT þ E: ð7bÞ
the high rejection of the elastically scattered light. This region can be
also easily analyzed using routine IR spectrometers. For rigid molecules, Fig. 4b shows that the loading matrix can be decomposed into load-
internal modes are detected above 150–200 cm−1. Two regions are usu- ing spectra which look like spectra composed of negative and/or posi-
ally distinguished in the spectrum of internal vibrations. (i) the finger- tive bands representing spectral information negatively or positively
print region covering the 500–1800 cm−1 range, distinctive of the correlated through particular PC. The presence of negative bands
molecular conformation and (ii) the intramolecular C–H, N–H, O–H makes it difficult the interpretation of loading spectra in terms of chem-
stretching region, lying from 2200 up to 3800 cm−1, recognized to be ical information. In order to achieve loadings more representative of
very sensitive to molecular associations via H-bonding [38], OH chemical information, different MVA methods have been developed.
and NH groups being the most commonly involved in intermolecular The widely used MCR technique is a two-step multivariate regression
H-bonding. In these two regions, no intensity correction is required, method, in which PCA of the data is firstly carried out, in order to deter-
since the Bose–Einstein factor becomes close to 1 above 100 cm− 1. mine the number of components and their contribution in the sample,
However, to obtain information on the formation of H-bonds, related used to decompose the D matrix. In a second step various regression
to the Raman intensity of X–H stretching bands [38] (with XO, N, C), methods can be used, including alternating least squares (ALS) method
the spectrum must be normalized with respect to Raman bands which [40] which is probably the most widely used. Using the ALS algorithm, D
are temperature independent. matrix is decomposed in such a way that a component corresponds
The fingerprint region is widely used to characterize molecular con- chemically to relevant distinctive feature of the sample. Several con-
formations both in small API molecules and biomolecules, to monitor straints are applied, especially the non-negative constraint [40, 41]. No

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
6 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

Fig. 4. Schematic diagrams for describing the decomposition of data matrix: a) as a function of the score data matrix. b) as function of the significant PCs and noise.

prior information is required in this method. The T matrix represents 5. Spectroscopic signatures of amorphous states
the distribution of the components contained in the sample, whereas
the PT matrix corresponds to the spectra of components. The universal signature of a solid amorphous state is an excess of
When the pure components are known and their spectra can be col- VDOS in the low-frequency region, mainly detected by inelastic neutron
lected, direct classical least square (DCLS) method [41] can be applied. scattering and LFRS. This enhancement of the low-frequency intensity
In DCLS method the PT matrix (spectra of pure components) is available. of the VDOS in the amorphous states reflects the deviation from the ν2
DCLS method finds the linear combination of spectra from the pure behavior below 50 cm− 1, i.e. the deviation of the Debye behavior of
components that most closely matches the Raman spectra collected VDOS commonly observed in crystalline states. The excess of VDOS
during the mapping (D matrix). Quantitative information can be obtain- gives rise to a low-frequency peak, the so-called Boson peak (BP),
ed by calculating the concentration matrix (T matrix) according to: which is considered as the universal signature of a solid amorphous
state [30, 52]. Among the many controversial interpretations of the ori-
 −1 gin of the BP, an emerging hypothesis is related to the confinement of
T ¼ DP PT P : ð7cÞ molecular vibrations within domains formed by different nearest mo-
lecular neighborings [52–55]. The designation of BP is extended to the
low-frequency band detected in liquids, corresponding to the “caging
DCLS is undoubtedly the most unambiguous method to display the
effect”, i.e. the frequency distribution of vibrations that molecules expe-
distribution of components from a Raman mapping data collection.
rience within the cage formed by their neighbors. The frequency of BP is
Using Raman micro-spectroscopy for quantitative analyses in a
inversely proportional to the size of these domains [52] both in glasses
multi-components system may have some limitation because of sub-
and liquids. The origin of BP is only related to the disorder existing in
sampling [42]. The backscattered Raman signal is mainly dependent
the amorphous state, and should have no correspondence in the low-
on the size of a laser spot (Ø ~ 100 μm). Consequently, Raman spectra
frequency spectrum of the underlying crystalline state.
collected in this spot size are representative of only a small area of the
sample. Different methods have been developed to avoid subsampling,
described in two reviews [6, 43]. It was shown [44] that subsampling 5.1. Amorphous solid state
can be reduced by moving the sample over a larger area during the
data collection. This method was tested on the quantitative analysis of The detection of BP in the solid state is highly depending on the
tablets with satisfactory results [44]. A second method consists to in- physical properties of the liquid. It was shown that the LFRS can be
crease the spot size. The so-called wide area illumination (WAI) method used to identify the nature of the liquid state (fragile or strong in the
was used for a quantitative determination of a drug in liquid formula- Angel classification [56]), and estimate the degree of fragility of a liquid
tions by direct measurement through a plastic bottle [45], in solid for- [35]. The ratio between the maximum and minimum reduced intensity
mulations in an intact capsule [46] and in tablets [47]. In these ([Ir]min/[Ir]max) reflects the relative contributions of relaxation and vi-
methods only the surface of samples are mainly analyzed. Transmission bration to the low-frequency spectrum. Fig. 6, plotted from data pub-
Raman spectroscopy becomes widely used for investigation of drugs lished by Sokolov et al. [35], shows that this ratio is significantly
more deeply inside the different kinds of samples [48, 49] including cap- higher for a fragile liquid (see LFRS of m-tricresyl phosphate — m-TCP,
sules [49]. Spatially offset Raman scattering is also a technique leading Fig. 6b) than for stronger liquid (see LFRS of glycerol in Fig. 6a). This fea-
to more representative sampling volumes with pharmaceutical applica- ture is usually considered as an empirical criterion to characterize the
tions [50]. In this method the imaging optics collecting the scattered nature of the liquid. Comparing Fig. 6a and b clearly reveals that BP is
light are moved away from the illuminated area [6, 42, 43, 51]. However, easily detected for strong liquid whereas it is rarely observed for most
in these different configurations, the microscopic information about of fragile liquids, i.e. in most of organic molecular glasses. An alternative
heterogeneities in the sample are lost. to detect BP is to plot χ " (ν)/ν2 to point out the deviation of χ " (ν) from

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx 7

Fig. 5. Relation between spectroscopic data and spectrum projection in the PCs coordinates.

the ν2-behavior. The LFRS has much unexplored potential for sensitive the knowledge of the whole low-frequency spectra of crystalline and
detection of low crystal content in solid dosage forms. However, there amorphous states, about 1% of amorphous content can be detected in
are only a few studies performed in the domain of pharmaceutical sci- a crystalline matrix, idem for the degree of crystallinity detected by
ences. X-ray powder diffraction is often used, since amorphous material analyzing a ground powder of IMC. A chemometric analysis [59] from
is characterized by a halo background and crystal by Bragg peaks, pro- FT-Raman investigation in the CO stretching region, has led to the deter-
viding clear and direct information on the nature of the physical state. mination of 2% of amorphous or crystalline IMC. Such a small amount of
Low-frequency Raman spectroscopy also offers the same type of con- amorphous material can be determined only because of the high sensi-
trast between the spectrum of phonon peaks in the crystalline state tivity of 1697 cm−1 band to the lost of long-range order, especially in
and the very broad hump of the Raman susceptibility, as shown in the γ-form in which the band is very intense and sharp. Several other
Fig. 7 for indomethacin (IMC). At higher frequencies, Raman active groups have used MVA analysis in combination with high-frequency
bands in amorphous materials exhibit broadening and frequency shift Raman spectroscopy to quantify crystallinity in amorphous systems
of Raman bands active in crystals, and there is no signature distinctive [59–61], but a recent paper [62] reports the first study of a low-
of an amorphous state, making it distinguishable from a very disordered frequency investigation combined with a MVA compared with classical
crystalline state. Additionally, it was shown that the LFRS was much high-frequency methods to quantify crystallinity in amorphous
more sensitive to detect and to identify the very early stages of crystal- grisefulvin tablets. This paper clearly demonstrates that LFRS is very
lization from a ground amorphous powder of IMC than the high- useful in the quantification of crystallinity, regarding higher frequency
frequency range [57]. A quantitative analysis has led to determining a regions analyzed by Raman or IR spectroscopy.
very accurate kinetics curve, making possible the interpretation of the The LFRS was recently used [63], combined with broadband
crystallization mechanism as corresponding to a two-dimensional (sur- terahertz time domain spectroscopy to determine the type H-bonding
face) crystallization from pre-existing nuclei of γ-IMC. This result sug- in the glassy state of IMC from analyzing intermolecular vibrations.
gests that grinding the γ-phase reduces the size of crystallites into an This study has revealed the existence of H-bonded cyclic dimers in the
amorphous powder, preserving nuclei of γ-phase. The LFRS is also glassy state of IMC and the intermediate correlation length of the glassy
very suited to the detection of a very small amount of amorphous via structure was estimated to about 2.5 nm from analyzing BP in glassy
the detection of BP. Fig. 7 shows the different representation of the IMC.
LFRS of the different physical states in IMC. It is observed that BP is de- Different preparations (quenching the liquid below the temperature
tected in the very low-frequency region, where no phonon peak of the of glass transition (Tg), milling crystals below Tg, pressurizing, freeze-
crystalline states is Raman active. The analysis of LFRS at different stages drying) can produce amorphous forms with different energies resulting
of grinding γ-IMC [58] has revealed very small amount of amorphous in variable properties, including bioavailability and the stability condi-
material after only two minutes of grinding, via the detection of a tions. The classical experimental methods are relatively unsuitable for
clear enhancement of χ″(ν) below 20 cm−1 localized by an arrow in the direct characterization of different amorphous forms. PCA method
Fig. 7e. Applying an analytical method, which can be only used from was used by Karmwar et al. [64] to analyze spectral variation between
the different amorphous states of indomethacin prepared via different
transformation methods: transformation via melt quenching, spray dry-
ing and by ball- and cryo-milling the α and γ crystalline forms. Spectra
of each sample have been recorded in the 1000–1720 cm−1 and 2800–
3100 cm−1 regions including the CO stretching region containing a sig-
nature distinctive of each state of IMC. Karmwar et al. [64] have shown
the existence of three clusters in the scores plot resulting from PCA anal-
ysis, that mirror the differences observed in the diffractograms of the
differently prepared samples. Despite a strong similarity between spec-
tra collected during a Raman mapping procedure, this study demon-
strates the capability of PCA to point out subtle structural differences
in amorphous IMC samples prepared by different transformation
methods. Sample differences have been interpreted using the spectral
loading plots and the assignment of Raman bands [65] in the spectral
regions where spectral differences are observed, suggesting that differ-
ent amorphous states could be related to different intermolecular
Fig. 6. Description of the criterion used to characterize the degree of fragility of liquids in
associations.
two cases from data taken in Ref. [34]: a) for a relatively strong molecular liquid: glycerol. Similar analysis using PCA method was applied to IMC glassy states
b) for a fragile molecular liquid: m-TCP. prepared by cooling the liquid with different cooling rates [66]. This

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
8 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

Fig. 7. Representations of the low-frequency Raman spectrum of the different states of indomethacin: a) Reduced spectrum (Ir) of the liquid state at 443 K, the glassy state at room tem-
perature obtained by grinding (ground amorphous state, GAS) and from quenching the glassy state (thermal glass, TG). b) Raman susceptibility (χ″) of the liquid state, TG and GAS sam-
ples. c) Raman susceptibility divided by ν2 to point out the deviation from the ν2-behavior. d) Raman susceptibility of the solid states: glassy state, and α-form, δ-form, γ-form. e) Raman
spectra collected at various times of grinding from γ-phase; the insert displays the very low-frequency range of spectra collected at time t = 0, 2 min, 4 min, and 6 min to show the en-
hancement of the Raman susceptibility in the very low-frequency range; the arrows localize this low-frequency signature of the amorphous state.

analysis has revealed significant structural changes in the different Tg) and by pressurization [14]. This feature combined with the evidence
amorphous samples, in the fingerprint region and the high frequency of two different mechanisms of amorphization, revealed by Raman in-
spectrum, suggesting different local organizations depending on the vestigations performed during pressurization and at different stages of
cooling rate used to undercool the liquid until the glass. cryogrinding, suggests the existence of a polyamorphism [67] in IMC,
Pressure-induced amorphization, using MDAC can be easily moni- i.e. the existence of two different amorphous forms linked via a first-
tored using Raman micro-spectroscopy. It was shown from a low- and order transition, as observed in water. Application of high pressure
high-frequency investigation that two different amorphous forms of could place molecular compounds in an amorphous state characterized
IMC can be produced by a thermal route (quenching the liquid below by higher stability than an amorphous state obtained by simple quench

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx 9

of the liquid as observed in ibuprofen [68]. Raman analyses under high network [71–74]. Addition of a protein, e.g. lysozyme, has two conse-
pressure can provide interesting information about mechanisms re- quences on the χ″(ν) spectrum of water [75–78]. (i) A shift of band
sponsible for possible enhanced stability of amorphous states. (I) reflecting a change in the close neighboring of a water molecule,
and (ii) an intensity decrease of band (II) induced by the disruption of
5.2. Amorphous liquid state the H-bonding network of water molecules. Comparison of χ″(ν) spec-
tra of water, dry protein and protein solution [76] in Fig. 8 shows that
The Analysis the liquid state has important applications in the bio- band (I) mainly reflects the dynamics of the protein with a weaker con-
pharmaceutical area, since a considerable amount of vaccines including tribution of the coupling between dynamics of protein and solvent. On
influenza vaccines are formulated as liquids. In aqueous solution they the other hand, band (II) gives the opportunity to probe the organiza-
are subjected to physical and chemical degradation, and then were tion of the H-bonding network in the solvent. Moreover, it was shown
stored and distributed under refrigeration. One strategy is to bring liq- that the temperature dependence of low-frequency region of the Ir(ν)
uid formulations into the dry state using suitable bioprotectants and spectrum was similar to that of the mean square displacements (MSD,
drying biotechnologies [69, 70]. An alternative to the use of these very bu2N). The combination of the analysis of the low-frequency region
expensive drying processes is to stabilize the aqueous formulations at and that of the amide I band, distinctive of the secondary structure [79],
room temperature and against temperature variations during storage provide a detailed description of the thermal denaturation process [80].
and shipping. In this context deciphering the thermostabilization mech- An important result is the discrimination between the transformation of
anisms of protein in aqueous solutions is probably a necessary step to the native tertiary structure into a molten globule state, detected from a
determine innovative methods for stabilizing vaccines in liquid solu- break in bu2N (T) [35, 37, 38], i.e. in the temperature dependence of
tions. Raman and IR spectroscopy are widely recognized as suitable Ir(ν), and the unfolding process of the secondary structure detected by
techniques for monitoring protein unfolding and were used to analyze a frequency upshift of the amide I band. The combination of low and
the mechanism of protein denaturation/stabilization. high-frequency Raman investigations is the first experimental method
There is no difference between the spectra of internal modes in the providing a relation between the dynamic transition at high tempera-
liquid and solid amorphous states. In the low-frequency region, Fig. 7 ture and the thermal denaturation process of proteins in solution. The
clearly shows significant changes between the Ir spectra of liquid and analysis of the temperature dependence of χ″(ν) spectrum has shown
solid amorphous IMC, mainly resulting from the strong contribution of the influence of the H-bond network of water on the denaturation
relaxational motions in the liquid state. χ″(ν) spectra of liquid and mechanism. Same types of analysis in presence of denaturants [81]
solid amorphous states are very similar and only distinguishable be- (urea, hydrochloride guanidine) and sugars [77, 78, 80, 82] (disaccha-
cause of the slight temperature dependence of the VDOS. The spectra rides) have led to the understanding of protein stabilization mecha-
are dominated by a broad band reflecting the molecular interactions nisms in aqueous solution, against various kinds of stress [13, 80, 83]
of a molecule with its disordered nearest neighboring. This indicates (low and high temperature, high-pressure). The influence of the differ-
that the short-range order is similar in the liquid and glassy states. In ent types of solutes (urea, guanidine hydrochloride, disaccharides) on
this context, it can be considered that the caging effect in the liquid the H-bond network the solvent was clearly shown by Raman spectros-
state and BP in the solid amorphous state have probably the same origin. copy. The preferential hydration [84, 85] of proteins in presence of di-
Low-frequency Raman spectroscopy has been widely used to analyze saccharides [86] or other sugars [87] was clearly evidenced at room
solution of proteins [8]. These solutions are mainly composed of temperature [80]. The use of D2O instead of H2O and deuterated treha-
water, and the analysis of protein solutions is based on the description lose, has led to a detailed description of sugar – protein – water interac-
of the spectra of water and dry proteins. The analysis of the low- tions, suggesting that the preferential hydration hypothesis is not valid
frequency spectrum of a protein formulation is described in Fig. 8. Con- to explain biopreservation mechanisms of disaccharides against high
trasting with the spectrum of IMC, χ″(ν) spectrum of water is composed temperature [80]. Combining Raman spectroscopy with molecular dy-
of two broad bands [71]. The low-frequency band (I) was assigned to namics simulations [71, 75, 88] has bring out crucial information at
the caging effect [71–74] and the high frequency band (II) is associated the molecular level, about the influence of disaccharides on the
with the intermolecular O–H stretching vibrations in the H-bonded hydrogen-bond network of water, explaining the exceptional capabili-
ties of trehalose to stabilize proteins.
Investigations for characterization of therapeutic proteins at high
concentration were recently performed using a Raman spectrometer
and dynamic light scattering system combined in single platform [89].
This kind of study provides information on the protein secondary and
tertiary structures and hydrodynamic size in real time during heating,
making possible the rigorous comparison between structural changes
and protein aggregation.

6. Using LFRS to analyze disordered and amorphous solid states

6.1. Structural description of disordered states

Forms I and II of anhydrous caffeine (C8H10N4O2) constitute an


enantiotropic system [90]. The commercial form II is thermodynamical-
ly stable at room temperature and transforms upon heating in form I at
153 °C [91]. Form I is an orientationally [92, 93] and dynamically [93]
disordered phase, called rotator phase, where molecules are slowly ro-
tating around their molecular C6 axis. Such a rotator phase is considered
as dynamical analogous of under-cooled liquid state [94]. Refinements
of X-ray powder data collected in form I [92] have conducted to a struc-
tural description of the molecular packing of caffeine molecules, very
Fig. 8. Raman susceptibility of aqueous solution of lysozyme (LW) compare to that of similar to that determined in theophylline [12]. The main differences
water (W) and dry lysozyme (dry-L). between both compounds belonging to the family of methylxanthines,

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
10 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

is the absence of disorder and a tilt of molecules out of the molecular phonon peaks, suggesting the existence of a substantial disorder. Sec-
plane in theophylline. Form II is quite intriguing, since despite more ond, both polymorphic forms of caffeine can be only distinguished
than 50 years of investigations [95] its structural description remains below 100 cm− 1; the fingerprint regions of both forms, plotted in
unclear. Indeed, the last refinements from X-ray powder diffraction Fig. 9b, are very similar and does not display highlights with respect to
data [96, 97] performed without any consideration of disorder, that of the liquid. These considerations indicate similar local order in
have led to the determination of an unusually very large unit cell both forms, and then similar description of the disorder in forms I and
(~ 4277 Å3) composed of 20 molecules, suggesting that the existence II. The close relationship between the fingerprint regions of polymor-
of disorder similar to that existing in phase I [98] should be considered. phic phases and the liquid state confirm the high degree of orientational
After purification by cold sublimation, caffeine is in the metastable form disorder of caffeine molecules. χ″(ν) spectra of theophylline and form II
I at room temperature. The I → II transformation is very hindered of caffeine are plotted in the inset of Fig. 9a at room temperature. This
around room temperature and faster transformations are observed figure reveals that the Raman susceptibility of form II of caffeine corre-
around 90 °C [99, 100]. Consequently, Raman spectra of both caffeine sponds to the rigorous envelope of the low-frequency phonon peaks in
forms are directly plotted in Fig. 9 at 90 °C, after isothermal transforma- theophylline. The close relationship between the LFRS of caffeine and
tion of form I into form II. Fig. 9 provides original and important infor- the low-frequency region of the phonon spectrum of theophylline can
mation on structural description of disordered phases. First, the LFRS be related to the predominant contribution of librational motions in ro-
of form I is composed of a single and broad band and characterized by tator phases. The relationship between both spectra indicates that in
the absence of phonon peaks, contrasting with the observation of form II, caffeine molecules are located in positions similar to those
sharp Bragg peaks in the X-ray powder diffraction diagram of form I occupied by theophylline molecules, and rotate around their molecular
[92, 97]. Bragg peaks only result from the periodic organization of the axis, probably more slowly than in form I, in agreement with previous
molecular mass centers, while lattice modes are related to intermolecu- studies [98]. The I → II transformation can be interpreted as the tilt of
lar atom-atom potentials. Consequently, the orientational molecular molecules out of the hexagonal plane in the lattice of form I, giving mo-
disorder induces an inhomogeneous broadening of phonon peaks corre- lecular positions similar to those determined by last X-ray investiga-
sponding to a librational density of states [100]. χ″(ν)-spectrum of form tions [96, 97].
II is characterized by a splitting of the broad low-frequency band
into two components [100] which are too broad to be considered as 6.2. Evidence of disorder

It was shown that racemic ibuprofen (IBP) transforms into a meta-


stable crystalline phase (phase II) upon heating from a deeply quenched
state [101]. Only LFRS has revealed a substantial disorder in phase II
[102], not directly detected by X-ray diffraction [103]. The χ″(ν) spec-
trum of phase II was observed typical of an amorphous state. However,
Raman features distinctive of the transformation of undercooled IMC
into phase II correspond to the signatures of the stable phase I. The spec-
trum of phase II was interpreted as reflecting the VDOS of the amor-
phous state where the first trace of the long-range order of phase I are
detected. Phase II was then interpreted as an early stage in crystalliza-
tion process in agreement with the “Oswald rule of stages” [104].

7. Using IR and Raman spectroscopy for the in-process monitoring


and for analyzing solid dispersions

It is well known that pharmaceutical manufacturing may induce dis-


order and (expected or not) amorphization. Grinding, holt-melt extru-
sion, spray-drying, freeze-drying are often used in order to improve
bioavailability of poorly water soluble drugs. Freeze-drying is a standard
but costly procedure, and then mainly used for preserving biological
systems in the solid state, including various types of vaccines which
are marginally stable in solutions.
During a freeze drying cycle proteins are exposed to different ex-
treme conditions (low temperature, pH, dehydration), and protective
additives whose effects are still not completely understood are used to
alleviate these conditions. Several hypotheses have been proposed
to explain the preservation of protein functions in the dry state
[105–112]. Raman and IR spectroscopy can be used to get a better un-
derstanding on biopreservation mechanisms in the dry state. Dong
et al. [113] have analyzed by Raman microscopy frozen protein formu-
lations prepared using various conditions of supercooling the solvent,
to determine the distribution of the excipient in the frozen medium in
relation with secondary structural changes and aggregation. Both IR
and Raman spectroscopy are used as process analytical (PAT) tools [5,
114–116], especially to monitor the primary drying by determining
the endpoint from the analysis the sharp and intense OH stretching
band in ice. De Beer et al. [22] have clearly demonstrated the comple-
Fig. 9. Raman spectra of crystalline phases in caffeine in different representations.
mentarity of NIR and Raman spectroscopy as PAT tools during a freeze
a) Raman susceptibility; the inset shows the close relationship between Rama spectra of drying cycle. Pressure-temperature devices can be used to mimic a
phase II and theophylline. b) Raman intensity of crystalline and liquid phases of caffeine. freeze drying cycle on microscopic volume, analyzed by Raman

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx 11

microspectroscopy to monitor simultaneously the secondary structure amorphization mechanism by collecting Raman spectra at different
of the protein via the analysis of the amide I band and the freeze- grinding times. Using this method, it was shown that the amorphization
drying process [18, 19, 117]. This analysis has shown that the water de- of IMC by grinding directly results from the “easy” breaking of the
sorption during the primary stage of freeze drying was the stress re- H-bond network of the γ-form (after short time of grinding), while
sponsible for a distortion of the protein structure, in absence of it was observed that this H-bond network remains stable at very high
stabilizing agents. In presence of sugar (trehalose), Raman mapping hydrostatic pressures and collapses in a high density state prior
performed at the end of the first (congelation) and second (primary amorphization [14].
drying) stages of freeze drying, have shown that the rigidity of the It is recognized that amorphous drugs are significantly more soluble
amorphous water-trehalose matrix was the main stabilizing effect of than their crystalline counterparts [131]. However, a major issue
trehalose to preserve the secondary structure of the protein during the associated with amorphous drugs is their inherent low stability, leading
freeze-drying process, in agreement with the vitrification hypothesis to unexpected crystallization. It was found that co-grinding a drug
proposed by Green et al. [108]. The results of these investigations are with a polymer enhances the physical stability of the drug (e.g.
in agreement with IR experiments [118], suggesting that embedding indomethacin-PVP [65, 132–134]), without detecting intermolecular
the protein in a rigid amorphous water-disaccharide matrix mainly ex- interactions between drug and polymer [57, 135]. Polymers enhance
plain the stabilization of the protein, making it less flexible. Raman spec- the physical stability during the storage and also during dissolution
troscopy also gives the opportunity to monitor the physical state of when the drug is introduced to aqueous media. Raman spectroscopy
excipients during a freeze-drying cycle which can be crucial for the was used to detect phase transformation of amorphous pharmaceuticals
properties of the freeze-dried product. Cao et al. [17] have analyzed exposed to an aqueous environment, to understand the behavior of
the influence of the operating parameters on the transformations of amorphous pharmaceuticals during dissolution, in presence or not of
mannitol. The presence of remaining water at the end of the freeze- polymers [136]. More recently co-amorphous drugs were found to be
drying process is crucial because of the plasticizing effect of water. IR promising for both enhancing solubility and physical stability of the
spectroscopy is very sensitive to a small amount of moisture, and is amorphous solid state [137]. FTIR was used to determine the origin of
used to monitor the presence of water in the amorphous freeze-dried the stability enhancement [138, 139]. The use of density functional the-
product. IR spectroscopy has been widely used to analyze freed-dried ory calculations to analyze the IR spectra of naproxen-indomethacin
products since Raman spectroscopy requested high laser power, re- glassy systems has revealed the formation of heterodimers (naproxen-
sponsible for protein degradation. Nowadays, the development of de- indomethacin) in a single phase co-amorphous mixture [140].
tector technology makes possible analyses freeze-dried products with It is recognized that grinding usually induces disorder [130]. Intrigu-
low laser power and short time of laser exposure. Consequently, IR ingly, it was shown from LFRS investigations, that under grinding at
and Raman spectroscopy combined with MVA are widely used to deter- room temperature, metastable form I transforms into form II, and
mine slightly changes in the secondary structure of freeze-dried pro- form II transforms into form I [141]. After a long enough grinding time
teins [119]. A quantitative estimate of the protein denaturation can be of form I or form II, a metastable driven state is reached, only stabilized
obtained using both techniques, by calculating a degree of overlap under grinding. This metastable state was described by LFRS as corre-
area of the amide I band collected in the denatured and native states sponding to a partially transformed form I into form II or inversely as
[19, 120]. Raman spectroscopy was also used for in-situ secondary a partially transformed form II toward form I.
structure determination of proteins in ice for frozen storage [121].
The demand for drug delivery systems is gradually growing, and the
use of IR and micro Raman spectroscopy is crucial to understand the ki- 8. Conclusion
netics of drug release. Understanding the mechanisms of drug dissolu-
tion and drug release requires the knowledge of the component It was shown that IR and Raman spectroscopy are very suited for
distribution within delivery device. Raman mapping is widely used for characterizing the physical state, the distribution of pharmaceuticals
providing the characterization of distribution and the solid state of the and biopharmaceuticals in drug delivery systems, and then is a crucial
drug [122–125]. aid for predicting the kinetics of drug release and drug dissolution. Now-
Hot melt extrusion (HME) is now widely used for the rapid produc- adays, Raman spectroscopy also focuses on interactions of cells with
tion of solid dispersions. Wilson et al. [126] have used Rama spectrosco- drug molecules and carrier systems [142]. Interactions of drugs with
py as PAT tool for monitoring of the concentration of drugs in the cells in suspension in an aqueous media can be analyzed by using optical
extrudate, of the physical state of extrudates, and for analyzing drug — tweezers for cell trapping [143–145].
excipient interactions. Saerens et al. [127, 128] also used Raman spec- If the most significant signature of an amorphous state is localize in
troscopy for in-line monitoring of the physical state of extrudates, for the very low-frequency region, the internal vibration region is widely
various formulations, and various operating parameters. Raman spec- used to characterized the physical state of pharmaceuticals with rou-
troscopy gives direct and rapid information to choose operating condi- tine IR and Raman spectrometers, including to distinguish the amor-
tions of HME and formulation to obtain amorphous extrudates. phous state from crystalline phases. However, to obtain structural
Haaser et al. [129] have used Raman microscopy to investigate dif- information on very disordered states or to distinguish the amorphous
ferent extrudate formulations in terms of component distribution and state from disordered crystalline states or micro/nano crystallized
structural changes during dissolution testing. Raman maps of cross- states, low-frequency data are requested requiring the use of complex
section of the extrudate corresponding to a binary system, based on a multi-grating equipment. It is often requested to investigate the very
tripalmitin matrix containing equal amount of the model drug theoph- low-frequency region, to perform quantitative analyses, or separate
ylline anhydrate, were performed after a dissolution period of 120 min. quasielastic scattering from collective vibrations. It can be expected
It was shown that the distribution of the drug particles in some places that this type of investigation becomes possible using routine spec-
completely surrounded by lipid, made dissolution impossible. trometers with the development of a new generation of holographic
Grinding is basically used to enhance bioavailability of poorly water filters [11]. In this context, very precise structural information on
soluble drugs by reducing the particle size of powder. The resulting amorphous states will become easily accessible with routine Raman
ground powder can be amorphous or not, mainly depending on the spectrometers. Combining low-frequency investigations with chemo-
temperature of grinding (below or above Tg) and other operating pa- metric analyses could provide very detailed structural information on
rameters (intensity, time) [130]. Raman spectroscopy is a fast acquisi- amorphous, especially making possible to distinguish amorphous
tion technique and does not require any specific sample preparation, states which are characterized by different calorimetric signatures of
giving the opportunity to analyze almost in-line the solid-state the glass transition.

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
12 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

Acknowledgments [27] R. Shuker, R. Gammon, Raman-scattering selection-rule breaking and the density
of states in amorphous materials, Phys. Rev. Lett. 25 (1970) 222–225.
[28] A.J. Martin, W. Brenig, Phys. Status Solidi B 64 (1974) 163.
The author is indebted to MMT group in Unité Matériaux Et Trans- [29] A.P. Sokolov, A. Kisliuk, D. Quitmann, E. Duval, Evaluation of density of vibrational
formations (UMR CNRS 8207), especially to Professor Y. Guinet and states of glasses from low-frequency Raman spectra, Phys. Rev. B 48 (1993)
7692–7695.
Dr L. Paccou. [30] A. Fontana, F. Rocca, M.P. Fontana, B. Rossi, A.J. Dianoux, Low-frequency dynamics
in superionic borate glasses by coupled Raman and inelastic neutron scattering,
Phys. Rev. B 41 (1990) 3778–3785.
References [31] N.V. Surovtsev, S.V. Adichtchev, E. Rossler, M.A. Ramos, Density of vibrational states
and light-scattering coupling coefficient in the structural glass and glassy crystal of
[1] K. Kawakami, Current status of amorphous formulation and other special dosage ethanol, J. Phys. Condens. Matter 16 (2004) 223–230.
forms as formulation for early clinical phases, J. Pharm. Sci. 98 (2009) 2875–2885. [32] T. Achibat, A. Boukenter, E. Duval, Correlation effects on Raman scattering from
[2] D.Q.M. Craig, P.G. Royall, V.L. Kett, M.L. Hopton, The relevance of the amorphous low-energy vibrational modes in glasses. II. Experimental results, J. Chem. Phys.
state to pharmaceutical dosage forms, Int. J. Pharm. 179 (1999) 179–207. 99 (1993) 2046–2051.
[3] C. Ahlneck, G. Zografi, The molecular basis of moisture effects on the physical and [33] A. Hedoux, P. Derollez, Y. Guinet, A.J. Dianoux, M. Descamps, Low-frequency vibra-
chemical stability of drugs in the solid state, Int. J. Pharm. 62 (1990) 87–95. tional excitations in the amorphous and crystalline states of triphenyl phosphite: a
[4] M. Yoshioka, B.C. Hancock, G. Zografi, Crystallization of indomethacin from the neutron and Raman scattering investigation — art. no. 144202, Phys. Rev. B 6314
amorphous state below and above its glass transition temperature, J. Pharm. Sci. (2001) 4202 (NIL_4169-NIL_4175).
83 (1994) 1700–1705. [34] V.N. Novikov, E. Duval, A model of low-frequency Raman scattering in glasses:
[5] T. De Beer, A. Burggraeve, M. Fonteyne, L. Saerens, J. Remon, C. Vervaet, Near infra- comparison of Brillouin and Raman data, J. Chem. Phys. 102 (1994) 4691–4698.
red and Raman spectroscopy for the in-process monitoring of pharmaceutical pro- [35] A.P. Sokolov, E. Rossler, A. Kisliuk, D. Quitmann, Dynamics of strong and fragile
duction processes, Int. J. Pharm. 417 (2011) 32–47. glass formers: differences and correlation with low-temperature properties,
[6] A. Paudel, D. Raijada, J. Rantanen, Raman spectroscopy in pharmaceutical product Phys. Rev. Lett. 71 (1993) 2062–2065.
design, Adv. Drug Deliv. Rev. 89 (2015) 3–20. [36] U. Buchenau, R. Zorn, A relation between fast and slow motions in glassy and liquid
[7] A. Bunaciu, H. Aboul-Enein, V. Dang Hoang, Raman spectroscopy for protein anal- selenium, Europhys. Lett. 18 (1992) 523–528.
ysis, Appl. Spectrosc. Rev. 50 (2015) 377–386. [37] M.C.C. Ribeiro, Low-frequency Raman spectra and fragility of imidazolium ionic
[8] Z.-Q. Wen, Raman spectroscopy of protein pharmaceuticals, J. Pharm. Sci. 96 liquids, J. Chem. Phys. 133 (2010) 24503–24506.
(2007) 2861–2878. [38] A. Hédoux, Y. Guinet, L. Paccou, P. Derollez, F. Danede, Vibrational and structural
[9] J.A. Zeitler, P.F. Taday, D.A. Newnham, M. Pepper, K.C. Gordon, T. Rades, Terahertz properties of amorphous n-butanol: a complementary Raman spectroscopy and
pulsed and spectroscopy imaging in the pharmaceutical setting — a review, J. X-ray diffraction study, J. Chem. Phys. 138 (2013) (214506-214501 – 2145068).
Pharm. Pharmacol. 59 (2007) 209–223. [39] G. Qun, L. Yan, L.H. Chen, C. Yifeng, L. Feng, Comparison of several chemometric
[10] C. Moser, F. Havermeyer, Ultra-narrow band tunable laserline notch filter, Appl. methods of libraries and classifiers for the analysis of expired drugs based on
Phys. B Lasers Opt. 95 (2009) 597–601. Raman spectra, J. Pharm. Biomed. Anal. 94 (2014) 58–64.
[11] R. Heyler, A. Randy, J. Carriere, F. Havermeyer, THz-Raman — accessing molecular [40] J.H. Wang, P.K. Hopke, T.M. Hancewicz, S.L. Zhang, Application of modified alter-
structure with Raman spectroscopy for enhanced chemical identification, analysis nating least squares regression to spectroscopic image analysis, Anal. Chim. Acta
and monitoring, in: M. Druy, R. Crocombe (Eds.),Conference on Next-generation 476 (2003) 93–109.
Spectroscopic Technologies VI, Book Series: Proceedings of SPIE, Baltimore, 2013 [41] N. Scoutaris, V. Kapilkumar, I. Slipper, B. Chowdhry, SEM/EDX and confocal Raman
(UNSP 87260 J). microscopy as complementary tools for the characterization of pharmaceutical
[12] E.D.L. Smith, R.B. Hammond, M.J. Jones, K.J. Roberts, J.B.O. Mitchell, S.L. Price, R.K. tablets, Int. J. Pharm. 470 (2014) 88–98.
Harris, D.C. Apperley, J.C. Cherryman, R. Docherty, The determination of the crystal [42] S.J. Fraser, K.C. Gordon, Raman spectroscopy in the study of pharmaceuticals; the
structure of anhydrous theophylline by X-ray powder diffraction with a systematic problem and solutions of subsampling and data analysis, Eur. Pharm. Rev. 19
search algorithm, lattice energy calculations, and 13C and 15N solid-state NMR: a (2014) 3–8.
question of polymorphism in a given unit cell, J. Phys. Chem. B 105 (2001) [43] C.M. McGoverin, T. Rades, K.C. Gordon, Recent pharmaceutical applications of
5818–5826. Raman and Terahertz spectroscopies, J. Pharm. Sci. 97 (2008) 4598–4621.
[13] A. Hédoux, Y. Guinet, L. Paccou, Analysis of the mechanism of lysozyme pressure [44] J. Johansson, S. Pettersson, S. Folestad, Characterization of different laser irradiation
denaturation from Raman spectroscopy investigations, and comparison with ther- methods for quantitative Raman tablet assessment, J. Pharm. Biomed. Anal. 39
mal denaturation, J. Phys. Chem. B 115 (2011) 6740–6748. (2005) 510–516.
[14] A. Hedoux, Y. Guinet, F. Capet, L. Paccou, M. Descamps, Evidence for a high-density [45] M. Kim, H. Chung, Y. Woo, M.S. Kemper, A new non-invasive, quantitative Raman
amorphous form in indomethacin from Raman scattering investigations, Phys. Rev. technique for the determination of an active ingredient in pharmaceutical liquids
B 77 (2008) 094205-094201-094210. by direct measurement through a plastic bottle, Anal. Chim. Acta 587 (2007)
[15] S. Hubert, S. Briancon, A. Hédoux, Y. Guinet, L. Paccou, H. Fessi, F. Puel, Process in- 200–207.
duced transformations during tablet manufacturing: phase transition analysis of [46] M. Kim, J. Noh, H. Chung, Y. Woo, M.S. Kemper, Y. Lee, Direct, non-destructive
caffeine using DSC and low frequency micro-Raman spectroscopy, Int. J. Pharm. quantitative measurement of an active pharmaceutical ingredient in an intact
420 (2011) 76–83. capsule formulation using Raman spectroscopy, Anal. Chim. Acta 598 (2007)
[16] R. Fule, P. Amin, Hot-melt extruded amorphous solids dispersion of posaconazole 280–287.
with improved bioavailability: investigating drug-polymer miscibility with ad- [47] M. Kim, H. Chung, Y. Woo, M.S. Kemper, New reliable Raman collection system
vanced characterization, Biomed. Res. 2014 (2014) (146781-146716). using the wide area illumination (WAI) scheme combined with the synchroneous
[17] W. Cao, Y. Xie, S. Krishnan, H. Lin, M. Ricci, Influence of process conditions on the intensity correction standard for the analysis of pharmaceutical tablets, Anal. Chim.
crystallization and transition of metastable mannitol forms in protein formulations Acta 579 (2006) 209–216.
during lyophilization, Pharm. Res. 30 (2013) 131–139. [48] J. Villaumie, H. Jeffrey, Revolutionising Raman with the transmission technique,
[18] A. Hédoux, L. Paccou, S. Achir, Y. Guinet, In situ monitoring of proteins during ly- Eur. Pharm. Rev. 20 (2015) 41–45.
ophilization using micro-Raman spectroscopy: a description of structural changes [49] P. Matousek, A.W. Parker, Non-invasive probing of pharmaceutica capsules using
induced by dehydration, J. Pharm. Sci. 101 (2012) 2316–2326. transmission Raman spectroscopy, J. Raman Spectrosc. 38 (2007) 563–567.
[19] A. Hédoux, L. Paccou, S. Achir, Y. Guinet, Mechanism of protein stabilization by tre- [50] C. Eliasson, P. Matousek, Passive signal enhancement in spatially offset Raman
halose during freeze-drying analyzed by in-situ micro-Raman spectroscopy, J. spectroscopy, J. Raman Spectrosc. 39 (2008) 633–637.
Pharm. Sci. 102 (2013) 2484–2494. [51] P. Matousek, I.P. Clark, E.R. Draper, M.D. Morris, A.P. Goodship, N. Everall, M.
[20] V.S. Tumuluri, M.S. Kemper, I.R. Lewis, S. Prodduturi, S. Majumdar, B.A. Avery, M.A. Towrie, W.F. Finney, A.W. Parker, Subsurface probing in diffuselely Scattering
Repka, Off-line and on-line measurements of drug-loaded hot-melt extruded films media using spatially offset Raman Spectroscopy, Appl. Spectrosc. 59 (2005)
using Raman spectroscopy, Int. J. Pharm. 357 (2008) 77–84. 393–400.
[21] L. Dierickx, L. Saerens, A. Almeida, T. De Beer, J. Remon, C. Vervaet, Co-extrusion as [52] E. Duval, A. Boukenter, T. Achibat, Vibrational dynamics and the structure of
manufacturing technique for fixed-dose combination mini-matrices, Eur. J. Pharm. glasses, J. Phys. Condens. Matter 2 (1990) 10227–10234.
Biopharm. 81 (2012) 683–689. [53] V.K. Malinovski, V.N. Novikov, A.P. Sokolov, Log-normal spectrum of low-energy
[22] T. De Beer, P. Vercruysse, A. Burggraeve, T. Quinten, J. Ouyang, X. Zhang, C. Vervaet, vibrational excitations in glasses, Phys. Lett. A 153 (1991) 63–66.
J. Remon, W.R.G. Baeyens, In-line and real-time process monitoring of a freeze dry- [54] V.K. Malinovski, A.P. Sokolov, The nature of boson peak in Raman scattering in
ing process using Raman and NIR spectroscopy as complementary process analyt- glasses, Solid State Commun. 57 (1986) 757–761.
ical technology (PAT) tools, J. Pharm. Sci. 98 (2009) 3430–3446. [55] I. Pocsik, M. Koos, Cluster size determination in amorphous structures using the
[23] J. Aaltonen, P. Heinanen, L. Peltonen, H. Korteiarvi, V.P. Tanninen, L. Christiansen, J. Boson peak, Solid State Commun. 57 (1990) 1253–1256.
Hirvonen, J. Yliruusi, J. Rantanen, In-situ measurement of solvent-mediated phase [56] C.A. Angell, Relaxation in liquids, polymers and plastic crystals — strong/fragile pat-
transition during dissolution testing, J. Pharm. Sci. 95 (2006) 2730–2737. terns and problems, J. Non-Cryst. Solids 131–133 (1991) 13–31.
[24] G. Reich, Near-infrared spectroscopy and imaging: basic principles and pharma- [57] A. Hédoux, L. Paccou, Y. Guinet, J.-F. Willart, M. Descamps, Using the low-frequency
ceutical applications, Adv. Drug Deliv. Rev. 57 (2005) 1109–1143. Raman spectroscopy to analyze the crystallization of amorphous indomethacin,
[25] J.M. Chalmers, P.R. Griffiths, Handbook of Vibrational Spectroscopy, John Wiley & Eur. J. Pharm. Sci. 38 (2009) 156–164.
Sons, New York, 2002. [58] A. Hédoux, Solid state transformations of APIs during manufacturing by Raman
[26] F.L. Galeener, P.N. Sen, Theory for the first-order vibrational spectra of disordered analysis of pharmaceutical molecules and dosage forms, Eur. Pharm. Rev. 16
solids, Phys. Rev. B 17 (1978) 1928–1933. (2011) 31–38.

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx 13

[59] T. Okumura, M. Otsuka, Evaluation of the microcrystallinity of a drug substance, in- [87] G. Xie, S.N. Timasheff, Mechanism of the stabilization of ribonuclease A by sorbitol:
domethacin, in a pharmaceutical model tablet by chemometric FT-Raman spec- preferential hydration is greater for the denatured than for the native protein, Pro-
troscopy, Pharm. Res. 22 (2005) 1350–1357. tein Sci. 6 (1997) 211–221.
[60] W. Sinclair, M. Leane, G. Clarke, A. Dennis, M. Tobyn, P. Timmins, Physical stability [88] R. Ionov, A. Hédoux, Y. Guinet, P. Bordat, A. Lerbret, F. Affouard, D. Prevost, M.
and recrystallization kinetics of amorphous ibinabant drug product by Fourier Descamps, Sugar bioprotective effects on thermal denaturation of lysozyme: in-
Transform Raman spectroscopy, J. Pharm. Sci. 100 (2011) 4687–4699. sights from Raman scattering experiments and molecular dynamics simulation,
[61] A. Heinz, M. Savolainen, T. Rades, C.J. Strachan, Quantifying ternary mixtures of dif- J. Non-Cryst. Solids 352 (2006) 4430–4436.
ferent solid-state forms of indomethacin by Raman and near-infrared spectrosco- [89] C. Zhou, W. Qi, E.N. Lewis, J.F. Carpenter, Concomitant Raman spectroscopy and dy-
py, Eur. J. Pharm. Sci. 32 (2007) 182–192. namic light scattering for characterization of therapeutic proteins at high concen-
[62] P.T. Mah, S.J. Fraser, M.E. Reish, T. Rades, K.C. Gordon, C.J. Strachan, Use of low- trations, Anal. Biochem. 472 (2015) 7–20.
frequency Raman spectroscopy and chemometrics for the quantification of crystal- [90] H. Bothe, H.K. Cammenga, Phase transitions and thermodynamic properties of an-
linity in amorphous giseofulvin tablets, Vib. Spectrosc. 77 (2015) 10–16. hydrous caffeine, J. Therm. Anal. 16 (1979) 267–275.
[63] T. Shibata, T. Mori, S. Kojima, Low-frequency vibrational properties of crystalline [91] A. Cesaro, G. Starec, Thermodynamic properties of caffeine crystal forms, J. Phys.
and glassy indomethacin probed by terahertz time-domain spectroscopy and Chem. 84 (1980) 1345–1346.
low-frequency Raman scattering, Spectrochim. Acta A 150 (2015) 207–211. [92] P. Derollez, N. Correia, F. Danede, F. Affouard, J. Lefebvre, M. Descamps, Ab initio
[64] P. Karmwar, K.A. Graeser, K.C. Gordon, C.J. Strachan, T. Rades, Investigations of structure determination of the high-temperature phase of anhydrous caffeine by
properties and recrystallisation behaviour of amorphous indomethacin samples X-ray powder diffraction, Acta Crystallogr. B61 (2005) 329–334.
prepared by different methods, Int. J. Pharm. 417 (2011) 94–100. [93] M. Descamps, N. Correia, P. Derollez, F. Danede, F. Capet, Plastic and glassy crystal
[65] L.S. Taylor, G. Zografi, Spectroscopic characterization of interactions between PVP states of caffeine, J. Phys. Chem. B 109 (2005) 16092–16098.
and indomethacin in amorphous molecular dispersions, Pharm. Res. 14 (1997) [94] P. Lunkenheimer, U. Schneider, R. Brand, A. Loidl, Glassy dynamics, Contemp. Phys.
1691–1698. 41 (2000) 15–36.
[66] P. Karmwar, J.P. Boetker, K.A. Graeser, C.J. Strachan, J. Rantanen, T. Rades, Investiga- [95] D.J. Sutor, The structures of pyrimidines and purines. VII. The crystal structure of
tions on the effect of different cooling rates on the stability of amorphous indo- Caffeine, Acta Crystallogr. 11 (1958) 453–458.
methacin, Eur. J. Pharm. Sci. 44 (2011) 341–350. [96] C. Lehmann, F. Stowasser, The crystal structure of anhydrous beta-caffeine as de-
[67] M.C. Wilding, M. Wilson, P.F. McMillan, Structural studies and polymorphism in termined from x-ray powder-diffraction data, Chem. Eur. J. 13 (2007) 2908–2911.
amorphous solids and liquids at high pressure, Chem. Soc. Rev. 35 (2006) [97] G. Enright, V. Terskikh, D. Brouwer, J. Rpmeester, The structure of two anhydrous
964–986. polymorphs of caffeine from single-crystal diffraction and ultrahigh-field solid-
[68] K. Adrjanowicz, A. Grzybowski, K. Kaminski, M. Paluch, Temperature and volume state C NMR spectroscopy, Cryst. Growth Des. 7 (2007) 1406–1410.
effect on the molecular dynamics of superccooled Ibuprofen at ambient and elevat- [98] J. Moura Ramos, N. Correia, H. Diogo, M. Descamps, Dielectric study of the slow
ed pressure, Mol. Pharm. 8 (2011) 1975–1979. motional processes in the polymorphic states of anhydrous caffeine, J. Phys.
[69] J.-P. Amorij, A. Huckriede, J. Wilschut, H.W. Frijlink, W.L.J. Hinrichs, Development of Chem. B110 (2006) 8268–8273.
stable influenza vaccine powder formulations: challenges and possibilities, Pharm. [99] V.P. Lehto, E. Laine, A kinetic study of polymorphic transition of anhydrous caffeine
Res. 25 (2008) 1256–1273. with micocalorimeter, Thermochim. Acta 317 (1998) 47–58.
[70] Y.-F. Maa, M. Ameri, C. Shu, L.G. Payne, D. Chen, Influenza vaccine powder formu- [100] A. Hédoux, A.-A. Decroix, Y. Guinet, L. Paccou, P. Derollez, M. Descamps, Low- and
lation development: spray-drying and stability evaluation, J. Pharm. Sci. 93 (2004) high-frequency investigations on caffeine: polymorphism, disorder and phase
1912–1923. transformation, J. Phys. Chem. B 115 (2011) 5746–5753.
[71] A. Lerbret, F. Affouard, P. Bordat, A. Hédoux, Y. Guinet, M. Descamps, Slowing down [101] E. Dudognon, F. Danede, M. Descamps, N. Correia, Evidence for a new crystalline
of water dynamics in disaccharide aqueous solutions, J. Non-Cryst. Solids 357 phase of racemic ibuprofen, Pharm. Res. 25 (2008) 2853–2858.
(2011) 695–699. [102] A. Hédoux, Y. Guinet, P. Derollez, E. Dudognon, N. Correia, Raman spectroscopy of
[72] G.E. Walrafen, M.R. Fisher, M.S. Hokmabadi, W.H. Yang, Temperature dependence racemic ibuprofen: evidence of molecular disorder in phase II, Int. J. Pharm. 421
of the low- and high-frequency Raman scattering from liquid water, J. Chem. (2011) 45–52.
Phys. 85 (1986) 6970–6982. [103] P. Derollez, E. Dudognon, F. Affouard, F. Danede, N.T. Correia, M. Descamps, Ab
[73] G.E. Walrafen, Y.C. Chu, G.J. Piermarini, Low-frequency Raman scattering from initio structure determination of phase II of racemic ibuprofen by X-ray powder
water at high pressures and high temperatures, J. Phys. Chem. 100 (1996) diffraction, Acta Crystallogr. B 66 (2010) 76–80.
10363–10372. [104] W. Oswald, Studien uber die Bildung und Umwandlung fester Korper, Z. Phys.
[74] J.A. Padro, J. Marti, An interpretation of the low-frequency spectrum of liquid Chem. 22 (1897) 289–330.
water, J. Chem. Phys. 118 (2003) 452–453. [105] J.H. Crowe, S.B. Leslie, L.M. Crowe, Is vitrification sufficient to preserve liposomes
[75] A. Lerbret, F. Affouard, P. Bordat, A. Hédoux, Y. Guinet, M. Descamps, Low- during freeze-drying? Cryobiology 31 (1994) 355–366.
frequency vibrational properties of lysozyme in sugar aqueous solutions: a [106] J.H. Crowe, L.M. Crowe, J.F. Carpenter, C. Aurell Wistrom, Stabilization of dry phos-
Raman scattering and molecular dynamics simulation study, J. Chem. Phys. 131 pholipid bilayers and proteins by sugars, Biochem J. 242 (1987) 1–10.
(2009) 245103–245110. [107] J. Crowe, J.F. Carpenter, L. Crowe, The role of vitrification in anhydrobiosis, Annu.
[76] A. Hédoux, R. Ionov, J.F. Willart, A. Lerbret, F. Affouard, Y. Guinet, M. Descamps, D. Rev. Physiol. 60 (1998) 73–103.
Prevost, L. Paccou, F. Danède, Evidence of a two-stage thermal denaturation pro- [108] J.L. Green, J. Fan, C.A. Angell, the protein-glass analogy: some insights from
cess in lysozyme: a Raman scattering and differential scanning calorimetry inves- homopeptide comparisons, J. Phys. Chem. 98 (1994) 13780–13790.
tigation, J. Chem. Phys. 124 (2006) 14703–14709. [109] T. Arakawa, S.N. Timasheff, Stabilization of protein structure by sugars, Biochemis-
[77] J.-A. Seo, A. Hedoux, Y. Guinet, L. Paccou, F. Affouard, A. Lerbret, M. Descamps, Ther- try 21 (1982) 6536–6544.
mal denaturation of beta-lactoglobulin and stabilization mechanism by trehalose [110] T. Arakawa, S.N. Timasheff, Preferential interactions of proteins with salts in con-
analyzed from paman spectroscopy investigations, J. Phys. Chem. B114 (2010) centrated solutions, Biochemistry 21 (1982) 6545–6552.
6675–6684. [111] T. Arakawa, S.J. Prestrelski, W.C. Kenney, J.F. Carpenter, Factors affecting short-term
[78] A. Hedoux, J.F. Willart, L. Paccou, Y. Guinet, F. Affouard, A. Lerbret, M. Descamps, and long-term stabilities of proteins, Adv. Drug Deliv. Rev. 46 (2001) 307–326.
Thermostabilization mechanism of bovine serum albumin by trehalose, J. Phys. [112] L. Chang, M. Pikal, Mechanisms of protein stabilization in the solid state, J. Pharm.
Chem. B 113 (2009) 6119–6126. Sci. 98 (2009) 2886–2908.
[79] R. Williams, K. Dunjer, Determination of the secondary structure of proteins from [113] J. Dong, A. Hubel, J. Bischof, A. Aksan, Freezing-induced phase separation and spa-
the amide I band of the laser spectrum, J. Mol. Biol. 152 (1981) 783–813. tial microheterogeneity in protein solutions, J. Phys. Chem. B 113 (2009)
[80] A. Hédoux, L. Paccou, Y. Guinet, Relationship between β-relaxation and structural 10081–10087.
stability of lysozyme: microscopic insight on thermostabilization mechanism by [114] R.L. Remmele Jr., C. Stushnoff, J.F. Carpenter, Real-time in situ monitoring of lyso-
trehalose from Raman spectroscopy experiments, J. Chem. Phys. 140 (2014) zyme during lyophilization using infrared spectroscopy: dehydration stress in
225102–225107. the presence of sucrose, Pharm. Res. 14 (1997) 1548–1555.
[81] A. Hédoux, S. Krenzlin, L. Paccou, Y. Guinet, M.-P. Flament, J. Siepmann, Influence of [115] A. Kauppinen, M. Toiviainen, J. Aaltonen, O. Korhonen, K. Jarvinen, M. Juuti, R.
urea and guanidine hydrochloride on lysozyme stability and thermal denaturation; Pellinen, J. Ketolainen, Microscale freeze-drying with Raman spectroscopy as a
a correlation between activity, protein dynamics and conformational changes, tool for process development, Anal. Chem. 85 (2013) 2109–2116.
Phys. Chem. Chem. Phys. 12 (2010) 13189–13196. [116] T. De Beer, M. Alleso, F. Goethals, A. Coppens, Y. Van der Heyden, H. Lopez De Diego, J.
[82] A. Hédoux, J.F. Willart, R. Ionov, F. Affouard, Y. Guinet, L. Paccou, A. Lerbret, M. Rantanen, F. Verpoort, C. Vervaet, J. Remon, W. Baeyens, Implementation of a process
Descamps, Analysis of sugar bioprotective mechanisms on the thermal denatur- analytical technology system in a freeze-drying process using Raman spectroscopy for
ation of lysozyme from Raman scattering and differential scanning calorimetry in- in-line process monitoring, Anal. Chem. 79 (2007) 7992–8003.
vestigations, J. Phys. Chem. B 110 (2006) 22886–22893. [117] T. De Beer, M. Alleso, F. Goethals, A. Coppens, Y. Van der Heyden, H. Lopez De
[83] A. Hédoux, J.-A. Seo, Y. Guinet, L. Paccou, Analysis of cold denaturation mechanism Diego, J. Rantanen, F. Verpoort, C. Vervaet, J. Remon, W. Baeyens, Implementation
of β-lactoglobulin and comparison with thermal denaturation from Raman spec- of a process analytical technology system in a freeze-drying process using Raman
troscopy, J. Raman Spectrosc. 43 (2011) 16–23. spectroscopy for in-line process monitoring, Anal. Chem. 79 (2007) 7992–8003.
[84] S.N. Timasheff, Protein-solvent preferential interactions, protein hydration, and the [118] L. Cordone, G. Cottone, A. Cupane, S. Giuffrida, M. Levantino, Proteins in saccharides
modulation of biochemical reactions by solvent components, Proc. Natl. Acad. Sci. matrices and trehalose peculiarity: biochemical and biophysical properties, Curr.
U. S. A. 99 (2002) 9721–9726. Org. Chem. 19 (2015) 1684–1706.
[85] S.N. Timasheff, Protein hydration, thermodynamic binding, and preferential hydra- [119] S. Pieters, Y. Van der Heyden, J.-M. Roger, C. Vervaet, T. De Beer, Raman spectros-
tion, Biochemistry 41 (2002) 13473–13482. copy and multivariate analysis for the rapid discrimination between native-like
[86] G. Xie, S.N. Timasheff, The thermodynamic mechanism of protein stabilization by and non-native states in freeze-dried protein formulations, Eur. J. Pharm.
trehalose, Biophys. Chem. 64 (1997) 25–43. Biopharm. 85 (2013) 263–271.

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021
14 A. Hédoux / Advanced Drug Delivery Reviews xxx (2015) xxx–xxx

[120] B.S. Kendrick, A. Dong, S.D. Allison, M.C. Manning, J.F. Carpenter, Quantitation of [133] A.G. Schmidt, S. Wartewig, K.M. Picker, Polyethylene oxides: protection potential
the area overlap between second-derivative amide I infrared spectra to determine against polymorphic transitions of drugs? J. Raman Spectrosc. 35 (2004) 360–367.
the structural similarity of a protein in different states, J. Pharm. Sci. 85 (1996) [134] T. Matsumoto, G. Zografi, Physical properties of solid molecular dispersions of
155–158. indomethacin with poly(vinylpyrrolidone) and poly(vinylpyrrolidone-co-vinyl-
[121] U. Roessl, S. Leitgeb, S. Pieters, T. De Beer, B. Nidetzky, Spectroscopy-based process acetate) in relation to indomethacin crystallization, Pharm. Res. 16 (1999)
analytical tool for frozen storage of biopharmaceuticals, J. Pharm. Sci. 103 (2014) 1722–1728.
2287–2295. [135] A. Hédoux, Y. Guinet, M. Descamps, The contribution of Raman spectroscopy to the
[122] A. Balogh, G. Dravavoelgyi, K. Farago, A. Farkas, T. Vigh, P.L. Soti, I. Wagner, J. analysis of phase transformations in pharmaceutical compounds, Int. J. Pharm. 417
Madarasz, H. Pataki, G. Marosi, Z.K. Nagy, Plasticized drug-loaded melt electrospun (2011) 17–31.
polymer mats: characterization, thermal degradation, and release kinetics, [136] D.E. Alonzo, G.G. Zhang, D. Zhou, Y. Gao, L.S. Taylor, Understanding the behavior of
J. Pharm. Sci. 103 (2014) 1278–1287. amorphous pharmaceutical systems during dissolution, Pharm. Res. 27 (2010)
[123] E. Karavas, M. Georgarakis, A. Doscoslis, D. Bikiaris, Combining SEM, TEM, and 608–618.
micro-Raman techniques to differentite between the amorphous molecular level [137] M. Alleso, N. Chieng, S. Rehder, J. Rantanen, T. Rades, J. Aaltonen, Enhanced disso-
dispersions and nanodispersions of a poorly water-soluble drug within a polymer lution rate and synchronized release of drugs in binary systems through formula-
matrix, Int. J. Pharm. 340 (2007) 76–83. tion: amorphous naproxen-cimetidine mixtures prepared by mechanical
[124] B. Vajna, H. Pataki, Z.K. Nagy, I. Farkas, G. Marosi, Characterization of melt activation, J. Control. Release 136 (2009) 45–53.
extrudedand conventional isoptin formulations using Raman chemical imaging [138] K. Lobmann, R. Laitinen, H. Grohganz, K.C. Gordon, C.J. Strachan, T. Rades,
and chemometrics, Int. J. Pharm. 419 (2011) 107–113. Coamorphous drug systems: enhanced physical stability and dissolution rate of in-
[125] I. Sievens-Fegora, A. Bhakay, J.I. Jerez-Rozo, N. Pandya, R.J. Romanach, B. Michniak- domethacin and naproxen, Mol. Pharm. 8 (2011) 1919–1928.
Kohn, Z. Iqbal, E. Bilgili, R.N. Dave, Preparation and characterization of hydroxypro- [139] K. Lobmann, C.J. Strachan, H. Grohganz, T. Rades, O. Korhonen, R. Laitinen, Co-
pyl methyl cellulose films containing stable BCS class II drug nanoparticles for amorphous simvastatin and glipizide combinations show improved physical sta-
pharmaceutical applications, Int. J. Pharm. 423 (2012) 496–508. bility without evidence of intermolecular interactions, Eur. J. Pharm. Biopharm.
[126] M. Wilson, M.A. Williams, D.S. Jones, G.P. Andrew, Holt-melt extrusion technology 81 (2012) 159–169.
and pharmaceutical application, Ther. Deliv. 3 (2012) 787–797. [140] K. Lobmann, R. Laitinen, H. Grohganz, C.J. Strachan, T. Rades, K.C. Gordon, A theo-
[127] L. Saerens, D. Ghanam, C. Raemdonck, K. Francois, J. Manz, R. Kruger, S. Kruger, C. retical and spectroscopic study of co-amorphous naproxen and indomethacin,
Vervaet, J. Remon, T. De Beer, In-line solid state prediction during pharmaceutical Int. J. Pharm. 453 (2013) 80–87.
holt-melt extrusion in a 12 mm twin screw extruder using Raman spectroscopy, [141] A. Hédoux, Y. Guinet, L. Paccou, F. Danede, P. Derollez, Polymorphic transformation
Eur. J. Pharm. Biopharm. 87 (2014) 606–615. of anhydrous caffeine upon grinding and hydrostatic pressurizing analyzed by
[128] L. Saerens, L. Dierickx, B. Lenain, C. Vervaet, J. Remon, T. De Beer, Raman spectros- low-frequency Raman spectroscopy, J. Pharm. Sci. 102 (2013) 162–170.
copy for the in-line polymer-drug quantification and solid state characterization [142] B. Kann, H.L. Offerhaus, M. Windbergs, C. Otto, Raman microscopy for cellular in-
during a pharmaceutical holt-melt extrusion process, Eur. J. Pharm. Biopharm. 77 vestigations — from single cell imaging to drug carrier uptake visualization, Adv.
(2011) 158–163. Drug Deliv. Rev. 89 (2015) 71–90.
[129] M. Haaser, M. Windbergs, C.M. McGoverin, P. Kleinebudde, T. Rades, K.C. Gordon, [143] D.G. Grier, A revolution in optical manipulation, Nature 424 (2003) 810–816.
C.J. Strachan, Analysis of Matrix dosage forms during dissolution testing using [144] A. Ashkin, J.M. Dziedzic, T. Yamane, Optical trapping and manipulation of single
Raman Microscopy, J. Pharm. Sci. 100 (2011) 44524459. cells using infrared laser beams, Nature 30 (1987) 769–771.
[130] J.-F. Willart, M. Descamps, Solid state amorphization of pharmaceuticals, Mol. [145] S. Rao, S. Raj, S. Balint, C. Bardina Fons, S. Campoy, Single DNA molecule detection
Pharm. 5 (2008) 905–920. in an optical trap using surface-enhanced Raman scattering, Appl. Phys. Lett. 96
[131] B. Hancock, M. Parks, What is the true solubility advantage for amorphous pharma- (2010) 213701–213703.
ceuticals, Pharm. Res. 17 (2000) 397–404.
[132] L.S. Taylor, G. Zografi, The quantitative analysis of crystallinity using FT-Raman
spectroscopy, Pharm. Res. 15 (1998) 755–761.

Please cite this article as: A. Hédoux, Recent developments in the Raman and infrared investigations of amorphous pharmaceuticals and protein
formulations: A review, Adv. Drug Deliv. Rev. (2015), http://dx.doi.org/10.1016/j.addr.2015.11.021

You might also like