Advanced Drug Delivery Reviews 89 (2015) 3–20
Contents lists available at ScienceDirect
Advanced Drug Delivery Reviews
journal homepage: www.elsevier.com/locate/addr
Raman spectroscopy in pharmaceutical product design☆
Amrit Paudel a,1, Dhara Raijada b,1,2, Jukka Rantanen b,⁎
a
b
Research Center Pharmaceutical Engineering GmbH (RCPE), Graz, Austria
Department of Pharmacy, Faculty of Health and Medical Sciences, University of Copenhagen, Copenhagen DK-2100, Denmark
a r t i c l e
i n f o
Available online 11 April 2015
Keywords:
Preformulation
Solid-state
Drug development
Formulation
Drug delivery
Process analytical technologies
Quality control
Dissolution testing
a b s t r a c t
Almost 100 years after the discovery of the Raman scattering phenomenon, related analytical techniques
have emerged as important tools in biomedical sciences. Raman spectroscopy and microscopy are frontier,
non-invasive analytical techniques amenable for diverse biomedical areas, ranging from molecular-based drug
discovery, design of innovative drug delivery systems and quality control of finished products. This review
presents concise accounts of various conventional and emerging Raman instrumentations including associated
hyphenated tools of pharmaceutical interest. Moreover, relevant application cases of Raman spectroscopy in
early and late phase pharmaceutical development, process analysis and micro-structural analysis of drug delivery
systems are introduced. Finally, potential areas of future advancement and application of Raman spectroscopic
techniques are discussed.
© 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
Contents
1.
2.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Enabling development of Raman techniques . . . . . . . . . . . . . . .
2.1.
Raman sampling configurations . . . . . . . . . . . . . . . . . .
2.2.
Fibre optics Raman probes, stand-off and handheld Raman spectrometer
2.3.
Simultaneous and hyphenated techniques . . . . . . . . . . . . .
2.4.
Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . .
3.
Applications of Raman spectroscopy in pharmaceutical product design . . .
3.1.
Early drug development . . . . . . . . . . . . . . . . . . . . .
3.2.
Characterization of drug delivery systems . . . . . . . . . . . . .
3.3.
Process measurements . . . . . . . . . . . . . . . . . . . . . .
3.4.
Product performance testing and quality control for the final product .
4.
Perspectives and future directions . . . . . . . . . . . . . . . . . . . .
Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3
4
4
5
6
7
7
7
8
10
12
13
14
14
15
1. Introduction
☆ This review is part of the Advanced Drug Delivery Reviews theme issue on
“Pharmaceutical applications of Raman spectroscopy — from diagnosis to therapeutics”.
⁎ Corresponding author at: Department of Pharmacy, Faculty of Health and Medical
Sciences, University of Copenhagen, Universitetsparken 2, 2100 Copenhagen, Denmark.
Tel.: +45 35 33 65 85; fax: +45 35 33 60 30.
E-mail address:
[email protected] (J. Rantanen).
1
These authors contributed equally to this work.
2
Author's present address: F. Hoffmann-La Roche Ltd., Pharma Technical Development
Actives, Grenzacherstrasse 124, CH-4070 Basel, Switzerland.
Since 1928, when Raman and Krishnan discovered the new type of
secondary radiation later termed “Raman scattering”, there has been a
constant flow of development in the instrumentation using this principle.
With the advancements in photonics and optoelectronics, as well as a
rapidly expanding range of applications, Raman spectrometers,
microscopes and allied analytical tools have continuously evolved
over the decades [1]. Several efforts in spectrometric hardware have
invested in tackling the intervention of the stronger elastic Rayleigh
http://dx.doi.org/10.1016/j.addr.2015.04.003
0169-409X/© 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
4
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
signal to the weaker Raman (inelastic) scattering signal. These efforts
have resulted in the highly efficient Rayleigh line filters [2]. Historically,
the developments of Raman tools were also augmented by the discoveries
of various phenomena other than the basic Raman effect, such as
resonance Raman (RR), coherent anti-Stokes Raman scattering
(CARS) and surface-enhanced Raman scattering (SERS) [1]. Raman
spectroscopy has proven to be a useful tool in a vast number of fields,
including pharmaceuticals, in vivo biomedical studies, process control,
environmental sciences, semiconductors, catalysts, glasses, pigments,
archaeology and forensic sciences. Raman analytical techniques have
been increasingly implemented at different stages of drug discovery and
development. This includes chemical identification, molecular biology
research and diagnostics, preformulation, solid form screening,
bioanalysis, formulation analytics in late phase drug development,
process analytics, quality control, raw material qualification and
counterfeit identification.
The analytical versatility of Raman molecular spectroscopic
techniques allows for investigating a wide diversity of transparent,
translucent, opaque, and coloured samples including solids, semi-solids,
suspensions, and solutions. This facilitates non-invasive analysis of
synthetic reaction mixtures, biological specimens, and intermediates to
various finished dosage forms (such as powders, tablet/capsule, creams,
heterogeneous suspensions, and syrups) with minimal or no sample
preparations [3]. Spontaneous instrumental evolutions of Raman-based
techniques have progressed from traditional laboratory spectrometers
to a wide range of miniaturized and tunable lasers, optical filters, spectrographs, interferometers, charge-coupled device (CCD), microprocessor
control and Raman probes. These have improved the mobile applications
to be non-contact, process-friendly and allow for remote analysis through
glass containers, well plates, and even aqueous samples [4]. Today,
Raman-based tools are cheaper, smaller, smarter and faster, and
the analysis of “real world samples” in-line from production lines
or inside the packaged containers has gone from a concept idea to a
well-established practice for different pharmaceutical products [5].
This has led to the recognition of Raman spectroscopy by regulatory
authorities for innovative analysis.
Spectral analysis of pharmaceuticals using Raman-based techniques
presents some additional benefits over mid- or near-infrared (IR)
spectroscopy. Because Raman spectroscopy is a scattering technique,
there is no need for a reference light path (as needed for IR/NIR);
therefore, it is amenable to fibre optics and allows for remote sampling.
Higher lateral spatial and depth resolution is attainable by (confocal)
Raman microscopy than by IR microscopy. In many ways, it is possible
to characterize samples better than with Fourier-transform IR (FT-IR)
spectroscopy. However, any possible fluorescence from the sample
should be taken into consideration. A single scan of a typical Raman
measurement can collect spectral data in the range of 4000–40 cm−1.
In addition to the fingerprint region between 4000 and 400 cm−1, the
low frequency or far IR region (400–40 cm−1) of Raman spectra covers
some of the important vibration modes that are relevant for the
identification of different solid-state forms [6]. Water being a weak
Raman scatterer makes it possible to successfully analyze aqueous
samples by Raman spectroscopy. In many cases, sharper spectroscopic
contrast between API and excipients offer superior quantitative analysis
capability.
Raman spectrometers comprise a laser light source, focusing optics,
and spectrograph(s) consisting of a dispersing element and a detector
[2,7]. The diverse Raman laser sources deliver in the range of UV to NIR
region, the most ubiquitous being the visible light laser [8]. As Raman
efficiency is inversely related to the fourth power of the wavelength,
the selection of a suitable source requires the accounts of sample nature
and the spectrometer geometry. Additionally, samples with possible
fluorescence interference require a longer excitation wavelength.
Dispersive Raman spectrometers utilize a holographic grating monochromator as the spectrograph, whereas FT-Raman instruments typically
utilize the Michelson interferometers. A Raman detector is generally a
photon collecting photomultiplier tube/CCD or a germanium detector
(FT-Raman). Raman microscopes are equipped with a conventional
optical microscope as a sampling device and thus can perform localized sample analysis, enabling hyperspectral chemical imaging by
wide field imaging or line/point mapping [1]. Furthermore, confocal
Raman microscopy (CRM) can axially discriminate signals originating from selective depth within the sample using the confocal hole
[9]. The fundamental working principles and components of Raman
instrumentations, from conventional to advanced instrumentations,
are available in many standard books and excellent review papers [1,
4,7,8,10]. We have excluded the basic quantum process associated
with various types of Raman scattering phenomena, but these are
also covered in many books and papers. Because the present review
aims to provide a broad perspective on the application of various
Raman tools in drug development, only very brief accounts are
reported [11].
2. Enabling development of Raman techniques
2.1. Raman sampling configurations
Different existing Raman sampling geometries are schematically
illustrated in Fig. 1. A traditional backscattering Raman analysis acquires
data from a small spot of an analyte. Therefore, the resulting spectral
output may fail to entirely represent the static and heterogeneous sample.
Sample rotation during spectral acquisition and the temporal averaging of
the acquired data, the spatial averaging of the data acquired by scanning
different regions of sample, and simultaneous wide-angle illumination
(WAI) are configurations available to overcome the issues related to
sub-sampling [12–14]. Most analysis of pharmaceutical solid samples
utilizes WAI, which involves the illumination of a large volume of samples
using wider laser beam and selective collection of Raman scattering
within the covered area. A dispersive Raman probe constituting
multiple optical fibres, known as PhAT (Pharmaceutical Area Testing),
measures a significantly larger sample volume [12,15]. Furthermore, the
configuration enabling the collection of signals from locations laterally
offset away (hundred micrometres to centimetres in some cases)
from the illuminated area is called spatially offset Raman spectroscopy
(SORS) [16]. Typical SORS geometry irradiates the laser at the centre of
the ring and Raman collection from the circumference, the radius
being the spatial offset [17]. The avoidance of surface interference
with this technique can facilitate the depth profiling of the sample.
Spectral acquisition representing a bulk sample can be performed
using transmission Raman spectroscopy (TRS), wherein the incident
and the collection beam path are separated to the extreme on opposite sides of sample [18]. TRS potentially avoids the sub-sampling
problem of heterogeneous samples and yields (semi) averaged
spectral data of the bulk composition for turbid or opaque materials
[19–23].
Surface-enhanced Raman scattering (SERS) and tip-enhanced
Raman scattering (TERS) are extensively researched and implemented
surface-sensitive Raman techniques [24–27]. In brief, a SERS signal
enhances via dipolar localized surface plasmon resonance originating
from the interaction of visible light and a nanoscale rough surface
noble-metal substrate coated with analyte [28]. The interaction of
sample with the intense electric field generated from substrateincident light interaction increases the magnitude of induced dipole of
the sample, thus enhancing the Raman scattering signal [24]. TERS is a
high resolution SERS variant combined with atomic force microscopy
(AFM) for the localization of incident light at sufficient proximity to
the analyte [29–31]. The surface plasmon excited at a sharp-metal tip
brought onto the sample surface initiates the SERS effect upon laser
irradiation [32]. In addition to surface sensitive Raman techniques, signal
enhancing Raman nonlinear techniques such as coherent anti-Stoke
Raman scattering (CARS) are increasingly used for the spectroscopic
imaging of pharmaceuticals [33,34]. CARS technique involves a coherent
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
5
Fig. 1. Different Raman spectroscopic configurations (Modified from [1,12,134–136]).
interaction between two high energy lasers (incident and pump laser)
collinearly irradiated on the sample producing strong Raman scattering.
The CARS signal inherently lacks the fluorescence interference. Several
recent reviews cover the details of these techniques [35–37].
2.2. Fibre optics Raman probes, stand-off and handheld Raman spectrometer
Fibre optics-based remote Raman systems are extensively used in
drug discovery and development. The construction essentially consists
6
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
of the sampling probe that is remote from the dispersive Raman
spectrometer and is connected via optical fibres acting as transmitters
for incident and collected light via total internal reflection [38]. The
separation can be as much as 100 m and can thus benefit the remote
analysis of samples [1]. The common fibre probe Raman setup consists
of a single excitation fibre surrounded coaxially by a bundle of separate
multiple collection fibres [39] (Fig. 1). Additionally, a double fibre probe
with separate illuminating and collecting fibres as well as a single fibre
probe for both the incident and collection beams is also common [1].
Both the distance and angle of incident beam fibre and collection
beam fibre are very important. Overall, fibre optic arrangement
provides extreme flexibility for monitoring samples that are at harsh
or physically inaccessible locations, e.g., varying process environments.
The fibre material should be devoid of Raman interference and resistant
to solarisation, especially using UV radiation. Interested readers are
referred to an excellent review published recently by Latka et al. on
fibre optic Raman probes [40].
Various factors, including the collection efficiency, S/N ratio, flexibility,
fibre-cladding, probe diameter, working distance, and confocality, are
important for fabricating fibre-based Raman sensors [40]. The requirements can be quite different for integrating a Raman probe as a PAT tool
in a particular location of a manufacturing line versus using it as an
in vivo diagnostic tool or with plate readers for solid-form screening.
The probe head is important in the case of immersion probes in solids,
semi-solids or solutions. A bevelled fibre tip enclosed in an air window
is reported to minimize the influence of the refractive index on the
depth resolution [41]. The resistant window of the Raman transparent
material housing collection lens is operable in contact mode with a
chemically reactive or high pressure environment [42]. Optimal focal
length immersion probes provide better results for the dispersive
media [43]. Wide varieties of TRS probes, SORS probes and ready-made
or customizable Raman probes are commercially available for in situ
and real time analysis of facile transformations in solid, liquid and melt
at non-ambient pressure; humidity; shear; temperature conditions
during flow; levitated drying; mixing; coating; spraying; extrusion;
and compression [42,44–46]. Furthermore, a recently devised Raman
fibre optic microstructure light guiding hollow-core and a miniaturized
sample holder for flowing gas or liquid analyte showed promising
quantitative capability for picomolar drug concentration in physiological
medium and for the mixture of gases in human breath [47,48]. Various
fibre-based Raman systems include but are not limited to the WAI probe
(PhAT), polarization Raman probes and imaging fibre probes, fibre coupled
confocal Raman, video probe, and a chemical/high pressure/temperature
resistant immersion probe [40]. Fibre optics-based Raman techniques are
well-recognized to be versatile and elegant in situ/in process monitoring
PAT tools for various pharmaceutical unit operations and processes
[49–51]. Moreover, micro-Raman sensors have been extensively developed in the medical and diagnostic analysis field [52–54], in vivo analysis
[55], and skin and cosmetic research and applications [56].
The other variant of the fibre-based Raman technique is a stand-off
spectrometer configuration that generally physically separates the
instrument from the sample [4]. This enables remote analysis using
telescopic signal collecting system(s) with the stand-off distance up to
500 m [57]. The incident laser source and collecting telescopic lens are
assembled together to make a probe that illuminates the sample and
collects the returning signal. This assembly is connected to the dispersing
and detecting spectrograph via optical fibre. Both dispersive backscattering and SORS systems are available in stand-off setup [58]. Stand-off tools
are widely used for remote analysis of packaged pharmaceuticals, as well
as toxic and explosive materials [59–61].
Substantial progression in miniaturization of Raman instruments
can be observed by the significantly diminished size and weight of
spectrometers over the last decade [62,63]. A recent report by Crocombe
includes extensive details of commercial portable handheld Raman
devices [64]. Major areas include the expeditious and cost-effective
testing of packaged samples for quality control, incoming material
inspections, at-line analysis of biopharmaceuticals, and counterfeit
analysis [65–69]. Portable Raman instruments consist of a miniaturized
laser source, a dispersive component, and detectors that are connected
via micro-fibre optics with the Raman probe. Furthermore, the incorporation of wireless technologies such as Bluetooth, WiFi and GPS into new
technologies has enabled remote operation of these handheld Raman
devices. Major commercial handheld devices work with a laser source
at 785 nm, whereas newer systems operate with a non-fluorescent source
(1030 or 1064 nm) [64]. A recently reported portable FT-Raman analyzer
and SERS system with a laser source less likely to cause eye damage
(1550 nm) shows versatility for mobile outdoor application [70,71].
Hand held SORS are potentially implementable for rapid, non-invasive
detection of drugs and explosives [72,73]. Although handheld Raman
instruments deploy an inbuilt (partial) spectral library and chemometric
data analysis, reliable calibration and relative intensity correction are
important for robust material identification. Various baseline correction
methodologies for handheld Raman instruments have been reported [74].
2.3. Simultaneous and hyphenated techniques
Cutting-edge research is moving towards the development of
analytics that deploy multiple measuring principles while (quasi-)
simultaneously analyzing the same sample. The complementary
information on various kinetic and thermodynamic events generated
using such integrated tools can provide an excellent opportunity for
understanding and more deeply characterizing complex materials. One
of the major challenges of integration will be avoiding the temporal lag
of response based on sensing principles and problems related to data
handling [75]. Thanks to the recent increasing ubiquity and portability
of the probe-based and compact Raman spectrometers, the spectrometers
are now more amenable to integration with many other analytical tools,
thus enabling in situ spectroscopy.
Attempts to integrate Raman systems with thermal analysis tools are
increasing. Raman was interfaced with a commercial thermal analysis
system to study dehydration and phase conversion of hydrate [76,77]. A
commercial DSC-Raman is equipped with fibre Raman mounted onto a
power compensation DSC furnace. A custom Raman probe-interfaced
heat flux DSC is also reported for polymer curing analysis [78]. A recent
study reports that a Raman system was integrated with ultrafast DSC,
enabling the facile characterization of subtle intermediate structural
changes of a metastable solid form [79].
Water–solid interactions are an extremely complex and important
aspect in pharmaceutical systems. The chemical speciation of molecular
changes occurring during moisture sorption or desorption can improve
the characterization of complex product microstructures [80,81]. The
relatively poorer Raman activity of water led to the coupling of a Raman
fibre probe with a dynamic vapour sorption (DVS) system [82,83] to
study molecular modes of moisture-API and/or excipient interactions and
crystallization in solid samples [84]. Previously, a controlled humidity cell
was combined with Raman microscope or a PhAT probe was implemented
in DVS to minimize gravimetric influence on spectral data [85,86].
Coupling Raman systems with microscopy can enable topochemical
profiling of the microscopic surface observed by optical microscopy,
mesoscopic features visualized via electron microscopy or highly localized
nanoscopic surface properties revealed by AFM [87]. Near-field scanning
optical microscopy (NSOM) integrated with a Raman system has been
developed as a high resolution system [88]. There are commercially
available modular multi-microscopic configurations combining AFM,
NSOM, confocal Raman and TERS [89]. These nanophotonic approaches
have yet to be applied to nanoscale pharmaceutical analysis [90–92].
The Raman integrated SEM (RISE) microscope, a recent promising
innovation, can enable nanoscopic molecular distribution analysis and
depth profiling by generating the chemical and topographical information
of the same sample area [93,94].
A tool interfacing the Raman principle with circular dichroism
spectroscopy measures Raman optical activity (ROA) and is considered
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
complementary to vibrational circular dichroism [95,96]. ROA exploits the
interaction of an optical activity tensor with the circular polarizability of a
chiral molecule and can thus provide structural information on chirality,
enantiopurity and so on [97,98]. Typical ROA samples are liquids or
solutions of peptide and protein, polymers, nucleic acid, carbohydrates,
oils, and biological molecules [99–105].
Simultaneous Raman-coupled X-ray and neutron techniques are
some of the recent exciting developments in this field [106–108].
The concerted measurement of Raman scattering with small and/or
wide-angle X-ray or neutron scattering, absorption or diffraction
yields powerful multi-scale structural data [109]. Such integrations
can be employed for the in situ monitoring of heterogeneous catalytic
reactions with extremely high (spatio-) temporal resolution [110–113].
Most of the reported studies utilize the installation of fibre-based optics
with remote Raman in the synchrotron beamline using a specially
customized sample holder [114–117]. A recent study reveals simultaneous measurement using in-house small- and wide-angle X-ray scattering with a high energy laboratory source coupled to a Raman probe for
complex polymers [118]. A Raman beam scanner is often used for
matching the X-ray or neutron beam and Raman laser spot. Multichannel
in situ assembly, including an X-ray Raman coupled with a mass
spectrometer, is also reported [119,120]. The correlation data obtained
from Raman vibrational changes and diffusional, crystallographic or
supra-molecular surface structures can elucidate structures of complex
API, excipient or delivery systems [121–123].
One ongoing area of development is towards multimodal microscopy/
spectroscopy, including Raman with other multi-techniques in fibre
optics systems [124–126]. An interest in multimodal vibrational spectroscopy has emerged with the aim to integrate Raman systems with NIR
systems [127]. The fusion of the separate NIR and Raman images acquired
from the same area of a formulation was reported to provide additional
value over individual techniques [128,129]. In contrast, attempts of
Raman-NIR integration increased due to identical measurement
speed and the possibility of single sampling configuration, indicating
the possibility of heterospectral Raman/NIR 2D correlation analysis
[130,131]. Additionally, Raman integrated with optical coherence
tomography (OCT) has a potential application to pharmaceutical
products [132]. Sequential or simultaneous data acquisition from the
same sample area using a combined Raman-OCT probe can provide
OCT-guided Raman chemical analysis [133].
Raman systems are sometimes bridged with different separation
techniques to act as a detector. Separated elution of multiple components
of an analyte using chromatography or electrophoresis can be detected
and possibly quantified by the hyphenated Raman spectrometer
[137–142]. Interested readers are referred to the relevant book for
the details on the technical aspects of using infrared or Raman as a
detector for chromatographic techniques [143].
2.4. Data analysis
Large datasets can be generated using Raman spectroscopy, which
creates a challenge in terms of extracting useful information from
these complex data. Univariate analysis is considered the easiest, most
prevalent and most robust data analysis approach and, in many cases,
can provide sufficient information and reliable predictability [144]. The
complex and vast amount of data generated using Raman spectroscopy
often demands multivariate data analysis approaches [145]. Common
multivariate methods include principal component analysis (PCA), partial
least square regression (PLS), classical least squares (CLS), multivariate
curve resolution (MCR) and partial least square discriminant analysis
(PLS-DA). Various preprocessing and variable selection methods prior to
multivariate modelling, as well as validation of the model after data
analysis, become crucial for deriving reliable quantitative information
[144–150]. Chemometric treatments and multivariate approaches
have been extensively published with regard to micro-Raman data
processing [146,151,152].
7
3. Applications of Raman spectroscopy in pharmaceutical product
design
3.1. Early drug development
Preformulation studies are an essential component of the drug development process and are performed to understand the physicochemical
properties of candidate drugs to support formulation development for
preclinical and clinical studies. Biopharmaceutics is the study of how
the physicochemical properties of the candidate drugs, the formulation/
delivery system and the route of administration affect the rate and extent
of drug absorption. Overall, it is important to understand the pharmacokinetic profile (absorption, distribution, metabolism, excretion) of the
candidate drugs during early development. Raman spectroscopy has
been used in this research area. Determination of 5-fluorouracil in saliva
using SERS is reported by Farquharson et al. [153]. Moreover, it is
important to interrogate the interaction potential of the candidate
drug compounds with the cytochrome P450 3A4 (CYP 3A4) enzyme
to predict first-pass metabolism and drug–drug interactions. Mak et al.
demonstrated the application of Resonance Raman spectroscopy for
understanding binding interactions of APIs with CYP3A4 [154].
An important part of preformulation is solid-form screening aiming
to identify the optimal solid form of the candidate drug compound as
early as possible. This can help avoid crisis during later stages of drug
development [155–157]. The lattice vibrations for different solid forms
results in unique and characteristic Raman spectra for each of these
forms. Therefore, Raman spectroscopy has long been used in pharmaceutical industry for this task. The structural backbone of a given solid
form can be examined by Raman spectroscopy because the spectral
features attributable to the lattice vibrational mode (shifts typically
near 200 cm−1) are discernible with the use of modern notch filters.
These filters allow the data collection of spectral data within 50 cm−1
of the excitation laser (which is not possible with FTIR). In this way,
the low frequency (LF) Raman region provides richer information
on the solid-state of materials. Therefore, the ability of rapid and
non-invasive solid form analysis via Raman spectroscopy has been
integrated into different high-throughput and low quantity solid
form screening platforms of APIs including microfluidic polymorph
and cocrystal screening [158–166].
Raman spectroscopy has been extensively used for understanding
phase transformations such as polymorphic changes, anhydrate −
hydrate transitions and amorphization [45,167–178], as well as for
identifying the mechanisms of co-crystal formation [179,180]. Taylor
et al. studied the influence of particle size on the dehydration behaviour
of trehalose dihydrate of two different particle sizes using in situ FT
Raman spectroscopy [177]. Although smaller particles (b45 μm) progressively underwent amorphization during dehydration, the larger particles
(N 425 μm) led to a crystalline anhydrate form. DSC-Raman facilitates the
visualization and understanding of molecular manifestations underlying
transient thermal transitions, such as polymorphic transformation of
crystalline pharmaceuticals [181–184]. As shown in Fig. 2, solid–solid
transition between enantiotropic polymorphs of sulfathiazole can be
clearly distinguished with the simultaneously recorded DSC-Raman data
[184]. Transformation of form III to form I observed in the DSC
thermogram is confirmed by the corresponding change in the integrated
Raman peak areas of respective polymorphs.
Raman spectroscopy was utilized as a real-time analytical technique
to monitor the anhydrate-to-hydrate transformation of a wide diversity
of APIs in an aqueous environment by Wikström et al. [178]. The transformation was found to be dependent on API chemistry and surface
properties and to be strongly affected by extraneous factors such as
degree of shear force and the presence of seeds. A recent publication
reports the use of integrated Raman-neutron scattering to study API
polymorphism [185]. In a recent study, Raman-coupled SEM was
shown to identify solution-mediated polymorphic transformation
of sulphathiazole at the dissolving crystal surface [186].
8
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
Fig. 2. DSC thermogram of sulfathiazole form III showing transition to form I and the
subsequent melting event of form I. Inset shows the simultaneously measured Raman
peak areas of corresponding polymorphs.
Reprinted (adapted) with permission from [184] Copyright (2009) John Wiley & Sons Inc.
mechanisms governing the reactions of different quercetin structures
and thus enabled the understanding of structure–instability correlations.
Furthermore, SERS analysis in Ag-doped sol–gel-filled capillaries was carried out to identify a nerve agent and its hydrolytic degradation products
in aqueous solution [192]. The kinetics of disproportionation of a crystalline new drug candidate–citrate salt in both buffered and unbuffered
aqueous slurries was measured using online Raman spectroscopy [193].
Apart from small molecules, determination of protein–complex
structures using XRD alone can be ambiguous. Wen et al. demonstrated
that Raman spectroscopy offers certain advantages over XRD for studies
of detailed local environments and intermolecular interactions of side
chains of proteins, both in solution and in crystals [194]. The authors
used visible Raman, Raman optical activity and UV Raman Resonance
spectroscopy to determine the conformation and local environments of
a recombinant human interleukin-1 receptor antagonist. Hédoux et al.
investigated the thermal denaturation behaviour of the lysozyme in aqueous solution with and without trehalose using Raman spectroscopy [195].
The spectral analysis of low frequency and amide region elucidated the
dynamics of protein–solvent and protein–trehalose interactions in the
native state and during the denaturation process. Trehalose addition
was shown to retard protein dynamics and thus reduce aggregation.
3.2. Characterization of drug delivery systems
Understanding solid state transformations as a function of humidity
and temperature is quite important for investigating the storage stability
of the selected solid form of the API. DVS is a frequently used method for
understanding the complex transformation pathways among various
anhydrate–hydrate forms as a function of humidity and temperature
[187]. However, the gravimetric data in DVS need to be further correlated
with the detailed structural changes. The utmost utility of the DVS-Raman
technique is in monitoring moisture- or solvent sorption/desorptioninduced solid-state transformation (e.g., anhydrate/hydrate, crystallization) of pharmaceuticals [188]. Gift et al. explored the application of the
DVS-Raman technique to gain a mechanistic understanding of watersolid interactions [84]. The authors studied various phenomena such as
stoichiometric and non-stoichiometric hydrate formation, deliquescence,
amorphous-crystalline transformations and capillary condensation.
Furthermore, the authors demonstrated that quantitative information
on transformation kinetics can be easily derived from the collected
Raman spectra at various humidities. Feth et al. derived useful information using a DVS-Raman tool [86]. The authors identified phase transitions
among non-stoichiometric hydrate forms using a combination of Raman
spectroscopy and DVS (Fig. 3a to d). In another study on the same
compound, Feth et al. used hot stage Raman spectroscopy to understand
the behaviour of form-1 (2.25 eq. hydrate) as a function of temperature
and correlated the phenomena using DSC (Fig. 3e and f) [189]. The
authors observed that form-1 transforms to form-2 (1 eq. water) above
60 °C followed by melting above 90 °C.
One of the important tasks during preformulation is understanding
the degradation profile of the drug in the presence of various excipients
under different stress conditions. Although HPLC and LC–MS analysis is
needed to obtain a complete picture of the degradation profile of the API
under various stress conditions, Raman spectroscopy can give useful
hints of possible degradation in the API. Savolainen et al. prepared
amorphous indomethacin samples using different preparation techniques
from α and γ forms of indomethacin as starting material. The authors
observed that Raman spectroscopy was the most sensitive technique
among spectroscopic techniques for distinguishing both the molecular
differences of amorphous samples prepared by different methods as
well as differential degradation behaviour of differently prepared
amorphous samples [190]. Sensitive Raman methods such as SERS
have also been used to study the autoxidation, molecular fragmentation,
dimerization and polymerization of quercetin flavonoids under alkaline
conditions in an aqueous solution and on an Ag nanoparticle [191].
SERS spectra acquired at different pH unravelled the molecular
Raman spectroscopy can be used for microstructural characterization
of drug delivery systems, as well as to understand drug–excipient
interactions in the formulation. Raman chemical imaging has been
utilized to determine the size distribution of API microparticles and to
determine the API distribution homogeneity in a composite formulated
tablet [196,197]. This type of information can be extremely useful
for the development of innovative drug products with increasingly
complex and highly engineered structures [198–201]. Scoutaris et al.
used confocal Raman microscopy (CRM) to study the distribution of
printed drug delivery systems of felodipine–PVP mixtures [198]. Chen
et al. used Raman spectroscopy to characterize nanosized PEGylatedreduced graphene oxide [202]. The latter was developed as an effective
carrier for a photo-thermally controllable drug delivery system with the
intention of increasing the effectiveness of tumour therapy. Hasa et al.
used Raman spectroscopy to elucidate drug–polymer interactions in
co-ground tablets prepared using cross-linked PVP or AcDiSol® [203].
One of the strategies for developing a modified release formulation is
the use of multi-layered tablets. Choi et al. developed geometric
three-layered tablets with various mechanical properties to achieve a
sustained release once-a-day formulation for a model drug, tamsulosin
HCl dihydrate [204]. The authors successfully used Raman imaging to understand the drug migration inside the dosage forms during dissolution,
which were in turn used to obtain a mechanistic insight into the drug release profile from the complex matrix system.
Amorphous solid dispersions are one of the effective solubilizing
formulations available for poorly water soluble APIs. Amorphous APIs
and carriers are widely characterized using Raman spectroscopy, even
in the presence of moisture [205]. The nature of water molecules in
hygroscopic amorphous polymer matrices such as PVP and PVP–VA
was discerned using FT-Raman spectroscopic analysis as a function of
moisture content [206]. Raman spectroscopy and imaging offer excellent
opportunities for understanding drug–polymer interactions, miscibility
and spatial phase distribution in solid dispersion formulations [151,167,
207,208]. Qian et al. performed Raman microscopic profiling on two
PVP–VA-based solid dispersions of a drug candidate, BMS-A, with the
same average composition (40% drug) but prepared by hot melt extrusion
(HME) using screw speeds of 50 RPM (HME 1) and 225 RPM (HME 2).
Despite the detection of a single identical glass transition temperature
for both extrudates, the drug underwent partial crystallization from
HME 1 after two months storage at ambient condition, whereas HME 2
remained amorphous. As shown in Fig. 4, the Raman chemical image of
HME 1 clearly demonstrates the evidence of poorer homogeneity with
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
9
Fig. 3. Phase transitions of a non-stoichiometric hydrate (form-1) (a) water sorption gravimetric curve: sorption isotherm (b) Raman peak position vs relative humidity plot during desorption
and sorption cycles, (c) distribution plots of the three hydrate forms during a water-sorption experiment determined by multivariate curve resolution (MCR) analysis of the Raman spectra,
(d) phase diagram of hydrate phases as a function of humidity (Form 1 — 2.25 eq. water, Form 2 — 1 eq. water, Form 3 — 1.5 eq. water), (e) variable temperature Raman hot-stage experiment,
(f) MDSC thermogram for form-1 in an aluminium pin-holed pan.
(a) to (d) reprinted (adapted) with permission from [86] Copyright (2010) John Wiley & Sons Inc, (e) and (f) reprinted (adapted) with permission from [189] Copyright (2010) Elsevier.
Fig. 4. Raman chemical imaging of HME 1 (50 RPM) (a) and HME 2 (225 RPM) (b) of BMS A–PVPVA (40% drug). The histogram inset shows the population distribution of different local
drug concentrations. In the colour index displayed on the right, blue and red represent the lowest and highest drug concentration, respectively.
Reprinted (adapted) with permission from [209] Copyright (2010) Elsevier.
10
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
local drug concentration up to 50% in some regions as observed in the
histogram inset. In contrast, fair compositional homogeneity was
discernible for HME 2 with narrow distribution of drug concentration
centred at 40%. This is in agreement with the superior physical stability
of the latter dispersion [209].
Furthermore, Raman spectroscopy can also provide useful information
about residual crystallinity in solid dispersion formulations and can be
used to study the crystallization kinetics of amorphous API from solid
dispersion formulations prepared by different techniques. The limit
of detection of the Raman method for the trace crystallinity in solid
dispersion is often comparable to that of XRD [210]. Time resolved
Raman spectroscopy has been used to study the kinetics of crystallization
of several amorphous APIs from the pure state as well as from molecular
dispersions at elevated temperature and/or humidity [176,211]. Owing to
the superior sensitivity of the phonon peaks of nanocrystallites (~30 Å)
generated during early stage crystallization of amorphous pharmaceuticals, there is growing interest in using LF Raman susceptibility to monitor
crystallization from amorphous pharmaceuticals [6]. Additionally, LF
Raman spectroscopy can be used to monitor structural evolution during
amorphization by cryo-grinding and pressure. Hedoux et al. obtained
distinct LF Raman spectra of amorphous indomethacin generated by
pressurizing and quench cooling/cryo-grinding the drug, and the authors
suggested a clear distinction between their local structures, pointing to
polyamorphism [167].
Raman spectroscopy can also provide useful information for the
fabrication of implants. Esrafilzadeh et al. developed multifunctional
conducting polymer-based hybrid fibres to achieve electrically controlled
release of ciprofloxacin. The authors used Raman spectroscopy to study
the electropolymerisation mechanism of the fibres [212]. The integrity
and microstructure of drug eluting stents are difficult to characterize
using conventional analytics because drug formulation is present as a
thin coat over the wire. CRM was used to analyze the chemical morphology of rapamycin/PLGA coatings on stents, from the surface through the
bulk of the coatings [213]. Biggs et al. published excellent results on the
comprehensive elucidation of the sirolimus-eluting coronary stent
(CYPHER™) during drug elution using CRM correlated with AFM [214].
Drug-rich regions revealed by surface and sub-surface Raman chemical
images were linked with AFM topography of the coating, which indicated
a link between drug release behaviour and device microstructure and
pore networks. Raman spectroscopic and imaging techniques have
been used for the characterization of different topical drug delivery
systems with in vitro, ex vivo and in vivo studies on the cutaneous penetration of active drugs from dermatological products [54,215,216].
Lipid-based drug delivery systems have gained interest in recent
years for enhancing the oral bioavailability of BCS class II and IV drugs
by means of increasing solubility and permeability [217–220]. Raman
spectroscopy and Raman imaging techniques have been successfully
employed for the characterization of lipid-based drug delivery systems
[201,221–224]. Breitkreitz et al. used Raman image spectroscopy for the
homogeneity evaluation of semi-solid self-emulsifying drug delivery
systems of the model compound, atorvastatin calcium [222]. Schaefer
et al. used CRM to probe the temperature-controlled release of encapsulated molecules from phospholipid vesicles in liposome formulations
[223]. Stillhart et al. determined drug precipitation kinetics during
in vitro lipolysis of a lipid-based drug delivery system using Raman
spectroscopy as a real-time monitoring tool [221]. The same group
recently extended the use of in-line Raman spectroscopy to monitor
in vitro drug dispersion, lipolysis and supersaturation of lipid-based
formulation using in vitro lipolysis integrated with a continuous
absorptive environment [225].
The characterization of spatial drug distribution on the carrier
surface is challenging but highly relevant to aerodynamics and API
lung deposition from dry powder for inhalation (DPI). Raman spectroscopy is also used in the analysis of respirable dosage forms [226] and to
gain insight into the controlled release behaviour of inhalable polymeric
microspheres for treatment of pulmonary hypertension [227]. High
performance LF Raman spectroscopy was shown to exhibit an excellent
ability to distinguish the solid-state (amorphous and crystalline) from
small mass fractions of API (b 3% w/w) present in multicomponent
respirable powders [226]. Theophilus et al. utilized Raman spectroscopy
to analyze the drug (co)deposited following the aerodynamic impaction
tests of a metred dose inhaler containing a salmeterol/fluticasone propionate combination, as well as the two drugs tested separately. The result
implied that the drug combination in a single device significantly increased the co-deposition over using separate inhalers [228]. Raman
chemical imaging using coupled Raman and optical imaging was
shown to be a promising method for characterizing particle size and distribution, together with chemical identity for a representative corticosteroid in aqueous nasal spray suspension [229]. CRM is also used as a
fast and convenient method for the analysis of content uniformity, homogeneity and the polymorphic form of a drug substance
distributed within a spray-dried lipid-based inhalable powder [230].
Recently, Fussell et al. showed the capability of CARS imaging to identify
drug particle clusters and their size on the carrier particle surface of a
DPI formulation [231]. As shown in Fig. 5, CARS images, when combined
with SEM images, provide information on the behaviour of morphological
changes and drug distribution onto the carrier surface as a function of
mixing time.
Raman spectroscopy can also be informative for the characterization
of protein and peptide drug delivery systems [232–234]. A Raman
spectroscopy-based method for in situ monitoring of secondary
structural composition of proteins during frozen and thawed storage
was developed by Roessl et al. [232]. Raman optical activity (ROA)
bears high potential for application towards protein characterization in
the native state as well as in a formulated environment [235]. Raman
chemical imaging was used to study the distribution of orally delivered
parathyroid hormone within sucrose spheres beneath the entericcoating layer and to visualize the enteric coating integrity [236].
3.3. Process measurements
The concept of process analytical technology (PAT) is becoming
increasingly important in the pharmaceutical industry. Raman spectroscopy is being considered a feasible PAT tool as it can provide molecular
level insight into chemical and physical phenomena taking place during
pharmaceutically relevant unit operations, which can be useful for
obtaining real-time information from these processes. Various analytical
applications of Raman spectroscopy and relevant pitfalls have been
compiled in literature reviews [49,237]. These cover several processes
related to the manufacturing of solid dosage forms such as synthesis,
crystallization, mixing, granulation, tableting and coating. Here, we have
compiled the recent advancements and selected examples from various
processes.
Synthesis and fermentation processes are often the source of smallmolecular and bio-macromolecular APIs. Raman spectroscopy has been
used to monitor the fermentation process of antibiotics [238,239].
Raman spectroscopy has also been used to monitor mammalian cellbased bioprocesses [240,241]. The quality of the cell culture media is a
crucial factor affecting bioprocess performance and the quality of the
final product in biopharmaceutical manufacturing. Calvet et al. demonstrated the application of SERS to the simple and fast analysis of cell
culture media degradation [242]. Crystallization is a crucial unit operation
used for the separation and purification of solid products. Batch crystallization is ideally performed in an aqueous environment. Therefore, Raman
spectroscopy is a very useful tool for process control and monitoring the
desired polymorphic form during crystallization [147,149,237,243,244].
Pataki et al. demonstrated the application of fibre optic-coupled Raman
spectroscopy for real-time monitoring of the desired polymorph of
carvedilol during the cooling crystallization process. The authors successfully tested the Raman signal feedback method for the production of
the desired form II of carvedilol [245].
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
11
Fig. 5. Correlative CARS–SEM image of salmeterol xinafoate mixed with α-lactose monohydrate particle intended for DPI (0.4 wt.% drug loading) after mixing times of 0.5 min (left) and
600 min (right).
Reprinted with permission from [231] Copyright (2014) John Wiley & Sons Inc.
Tablets and capsules are still considered the preferred dosage forms.
The manufacture of tablets involves several unit operations such as
milling, powder blending, granulation, drying, compaction and coating.
It is crucial to control the quality of the intermediate products generated
during various unit operations to manufacture the final quality product.
Raman spectroscopy has been found to be a very useful PAT tool for inprocess monitoring to test for the desired intermediate product after
each unit operation [246]. De Beer et al. applied Raman spectroscopy
as a PAT tool for the end-point control of a powder blending process
[247]. The authors demonstrated that Raman spectroscopy enabled inline and real-time monitoring of the blend homogeneity and also led
to a better understanding of the process. The effect of particle size,
shape, density and/or light absorption capability on the Raman
sampling depth during in situ monitoring of API-excipient blending by
wide area illumination (WAI) was recently studied by Allan et al.
[248]. Poutiainen et al. developed a mechanistic understanding of the
evolution of granule structure and drug content homogeneity during
fluidized bed granulation using X-ray microtomography and confocal
Raman spectroscopy [249]. Wikström et al. studied the effect of various
process parameters on the rate and extent of theophylline anhydrate to
monohydrate conversion during high-shear wet granulation using inline Raman spectroscopy [250]. A solvent-mediated transformation
model was shown to estimate the time scale of the transformation
onset for this compound. The higher sampling volume and improved
robustness against sampling distance fluctuation was obtained for
monitoring phase transformation using large spot Raman optics
compared to the standard immersion Raman probes. Christensen et al.
studied the influence of excipients on the process-induced transformation of piroxicam anhydrate to monohydrate during wet granulation
using Raman spectroscopy [148]. Raman spectroscopy is also a valuable
tool for monitoring the coating process [251–253]. Knop et al. recently
published a literature review on PAT-tools for process control in
pharmaceutical film coating applications. The authors suggested that
in-line Raman spectroscopy is a promising PAT tool for monitoring the
coating process of solid dosage forms, irrespective of the coating
apparatus or the core type (tablets or pellets) [252]. Wirges et al.
successfully implemented Raman spectroscopy on a production scale
to monitor the coating process. The authors also validated the Raman
spectroscopic method at a production scale according to the European
Medicine Agency (EMA) guidelines with respect to the special
requirements of the applied in-line model development strategy [251].
Hot melt extrusion (HME) has gained significant attention in the
pharmaceutical industry in the last five years for the rapid production
of solid dispersions [254]. Raman spectroscopy has also been used as a
PAT tool for inline and real-time monitoring of the concentration of
APIs in the extrudate or of the solid state of the extrudates, as well as
to provide information about drug–polymer interactions during the
hot melt extrusion process [254,255]. Saerens et al. demonstrated an
application of Raman spectroscopy for in-line monitoring of the solid
state of the hot melt extrudates of Celecoxib–Eudragit E PO formulations
prepared with varying drug loads (30%, 40%, 50%), temperatures (130 °C,
140 °C, 150 °C), and screw configurations. A Raman probe was inserted
into the die head to monitor the solid state of the extrudates before
they were forced through the die. The temperature at the Raman
measurement point was kept constant at 110 °C. As shown in Fig. 6a,
the in-line Raman measurements can easily distinguish glassy solid solutions (red), crystalline solid dispersions (blue) and partially crystalline
solid dispersions (green). Furthermore, Raman spectra also provided
useful information on the process history of the samples. As shown in
Fig. 6c and d, a large variation in the average die pressure was noticed
between different experiments. The authors suggested that this variation
was caused by two factors: (i) higher processing temperature, which may
reduce the viscosity of the melt and lower the die pressure; and (ii) higher
drug load, which leads to a plasticization effect. Higher pressure at the
measurement point resulted in a decrease in Raman intensity [256]. In
addition to solid-state analysis, transmission Raman spectroscopy (TRS)
and CRM was used to investigate the process-induced degradation of
spironolactone during the preparation of solid dispersion with Eudragit
E by HME [257]. The correlation of TRS results of the ratio of drug to
degradation product, to that obtained by Raman chemical imaging and
to the results of chromatographic analysis indicated the application of
TRS as a rapid and non-invasive quality control technique.
Recent advancements in the field of bio-macromolecules have
opened a new avenue in the development of protein drug delivery
systems. Freeze drying /lyophilization is often used in the manufacturing of protein drug delivery systems. Physical stability of proteins
plays a crucial role in the quality of final protein formulations
[258–260]. Raman spectroscopy can be applied as a non-invasive technique for the in-line monitoring of crystallization and polymorphic transformation of APIs and/or excipients during freeze-drying [261]. De Beer
et al. proposed a strategy for the implementation of a PAT system in
freeze-drying processes using in-line Raman spectroscopy and mannitol solutions as a model system for the freeze-drying process [262]. Waard et al.
developed a novel process, “controlled crystallization during freezedrying”, using in-line Raman spectroscopy during the freeze-drying process
to understand the crystallization behaviour of solvent and solute components during various phases of the freeze drying process [263]. Hedoux
et al. used micro-Raman spectroscopy for in situ monitoring of proteins
during lyophilization. The authors investigated the influence of the lyophilization parameters on the degree of protein denaturation [260].
12
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
Fig. 6. In-line Raman measurements of Celecoxib — Eudragit E PO formulations (varying drug loads, 30%, 40%, 50%; temperatures,130 °C, 140 °C, 150 °C; and screw configurations, FF, FB,
2 F); (a) PCA scores plot (red cluster, glassy solid solutions; green cluster, partially crystalline solid dispersions; blue cluster, crystalline solid dispersions); (b) PCA loadings plot for PC1 scores in
(a); (c) PCA scores plot from in-line collected raw Raman spectra at 110 °C (different score values reflect process history); (d) Raman spectral intensity with mean melt temperatures and
varying die pressures (varying intensities is due to differences in viscosity at different processing temperatures (130 °C, 140 °C, 150 °C)).
Reprinted (adapted) with permission from [256] Copyright (2014) Elsevier.
3.4. Product performance testing and quality control for the final product
Quality control of finished pharmaceutical products is often
considered to be a challenging task due to the complexity of the dosage
form and the presence of multiple excipients in the final dosage form
[264]. Mostly, APIs give very sharp and significant Raman spectra,
whereas excipients do not give very intense Raman signals [265].
This makes Raman spectroscopy a very useful tool for performance
testing of the final product and to correlate the results of performance
testing with the changes in crystal structure or morphology. Additionally,
Raman spectra are almost entirely free from interference of water bands.
This allows monitoring of the solid form changes during dissolution
testing. However, the detection of API solid-states in the presence of
solvent media is limited by the Raman signal strength and spectral bandwidths, which are significantly broader in solutions.
Drug content assays and impurity profiles are two critical quality
attributes of a finished dosage form. A low resolution Raman device is
useful for rapid and on-line content uniformity determination in solid
oral dosage forms such as tablets [266]. Owing to the multicomponent
nature of finished dosage forms, various multivariate data treatment
and chemometric methods are useful for the rapid detection and
quantification of drug and excipients [267]. Therefore, the accuracy
of content uniformity analysis and quantification also depends on
the selection of a sufficiently predictive multivariate model for a
particular dosage form [145]. Raman spectroscopic content analysis
and the quantification of powder blend formulations and heterogeneous
samples require due consideration of the influence of variations in
optical path length on the Raman signal due to particle size and shape
distribution and the implementation of suitable corrections and chemometric techniques [268,269]. TRS is found to be well suited for content
analysis for formulations such as bulk tablets (even coated) with minimal error and for accurate compositional analysis of drug suspension
in turbid media [13,270–272]. TRS mode was also recently implemented
for the quality control of cocrystal tablets [273] and the analysis of solid
dispersions [274]. Successful quantitative application of TRS was demonstrated for multi-component pharmaceutical capsules in intact forms
owing to the absence of interferences from capsule shell signal and
insensitivity to the capsule fill weight and size [275,276]. A recent publication reported the use of Raman for the quality control of drug products
that are carried by astronauts into space [277]. The space-induced degradation of acetaminophen was monitored using Raman spectroscopy by
measuring the API and its toxic degradation product (aminophenol).
Another recent study reported the quantification of aminophenol at
concentrations of 0.025% to 0.2% in an acetaminophen tablet using a
SERS chemical imaging approach [278]. In addition to drug content and
impurity profiles, detection of drug counterfeiting has been a serious
challenge for pharmaceutical dosage forms [279,280]. Raman spectroscopy, especially handheld devices and portable units, enables
analysis through a container such as a blister pack or coloured vial
without sample preparation, thus allowing rapid and non-invasive
material identification [281,282]. Spatially offset Raman spectroscopy
(SORS) has been applied for rapid and non-invasive identification of
pharmaceutical raw materials as well as for the characterization of
genuine and fake tablets through blister packs [283,284]. Furthermore,
Raman microscopy and two-dimensional correlation spectroscopy
have been employed for the analysis of counterfeit pharmaceutical
tablets enclosed within the packaging [285–287]. However, the
identification and quantification of falsified medicines in complex,
multi-component dosage forms using low resolution Raman spectroscopic data necessitates strong and reliable chemometric data analysis
models [286,288].
One of the strengths of solid-state Raman spectroscopy is the ability
to produce a unique signature of vibrational bands representing the 3-D
arrangement of a particular polymorphic form of a crystalline solid.
Therefore, subtle alteration in the solid-state structure, including
conformational state changes due to polymorphic modifications,
can provide quantitatively specific Raman signals owing to the
diverse phonon vibration of crystals. Likewise, a hydrate can often
be readily quantified in the presence of an anhydrate in the mixture
using Raman spectra. Bulk amorphous form presents a broader spectra
or loss of different peaks present in its crystalline counterpart. Extensive
studies are available on the quantification of solid forms using Raman
spectroscopy, microscopy and chemical image analysis in static, multicomponent mixtures or during various pharmaceutical processes as
aforementioned, often accompanied by suitable statistical or chemometric treatment of data and multivariate calibration model development
[147,171]. The readers are referred to a review by Strachan et al. on the
application of Raman spectroscopy for the quantification of pharmaceutical solid forms [289]. Proper matching of the particle size distribution of
calibration mixtures with that of the actual mixture is necessary to
avoid the influence of particle size on Raman intensity, thus avoiding
quantitative prediction errors [290–292]. Appropriate sample preparation
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
and homogeneity of the calibration sample is necessary to develop an
accurate calibration model [293]. Sampling volume and data processing
are important factors relating to the productivity and quantification
power of Raman-based polymorphic and hydrate/anhydrate analysis
methods [144]. Raman spectroscopy has been used to quantify binary or
ternary polymorphic and/or pseudopolymorphic forms of APIs in powder
mixtures, crystallinity in amorphous/crystalline mixtures, polymorphic
forms and amorphous mixtures of excipient, solid forms of APIs, and
excipients in tablet form [150,294–299]. Various studies report the superiority of the TRS method in terms of speed of analysis, sampling statistics,
peak sharpness and low wavenumber region (solid-state lattice modes)
for the facile quantification of different polymorphs of APIs in binary
mixtures as well as in the presence of excipients in tablets (in blister
packages) or capsules [19,271,300–303]. In situ Raman spectroscopy
was used to monitor and quantify intermediate to final solid forms
generated during the dehydration of hydrate forms of piroxicam and
carbamazepine, and the quantitative models built were exploited by
solid state quantification during fluid bed drying [304,305].
Drug release from solid dosage forms such as tablets and capsules is
a complex process that is accompanied by concomitantly occurring
physicochemical, solution-mediated solid-state transformations at the
surface and the bulk of solids in contact with the aqueous media.
Raman spectroscopy can enable the understanding of the solid-state
manifestation via dissolution by analyzing formulations before dissolution and the residual solids remaining during dissolution. Moreover, the
lack of water signal interference and inherent capability of data collection
in aqueous systems have led to the increasing popularity of fibre optic
Raman probes or Raman microscopy as powerful tools for in situ monitoring of various solution-mediated phase transformations, which includes
monitoring during drug dissolution and the release testing of powder or
compacts in various in vitro dissolution apparatuses and flow-through
dissolution devices, even in biorelevant media [306–308]. Recent studies
also showed the utilization of Raman spectroscopy orthogonally coupled
with UV imaging for dissolution monitoring [309,310]. Interested readers
are referred to a review on real-time Raman analysis for studying
solution-mediated phase transformations by Greco and Bogner [311].
The choice between Raman spectroscopy versus microscopy depends
on the type of investigation and dosage form (compact versus powder).
For example, Raman microscopy is suitable for the system undergoing
relatively slow transformations monitored in one single crystal or on a
small portion of a compact [312]. Raman spectroscopy is better for
quantitative purposes in fast conversions with a large sample volume.
Various solution mediated phase transformations accompanying drug
dissolution can be monitored by Raman spectroscopy/microscopy,
including anhydrate to hydrate transformation [309], crystallization kinetics of amorphous pharmaceuticals [313,314], disproportionation of
salt forms of API [310,315,316] and polymorphic transformations [317].
Strong hydrogen bonding interactions between carbamazepine and
sodium taurocholate, a component of fasted state simulated intestinal
fluid (FaSSIF), were inferred from in situ Raman spectroscopic data collected during dissolution of carbamazepine in FaSSIF [307]. This is related
to the inhibition of anhydrate to dihydrate transformation in FaSSIF owing
to the formation of needle anhydrate. This implies that in situ Raman
spectroscopy can be a powerful tool for monitoring solid formdissolution media interaction. In addition to conventional Raman
systems, Hyperspectral CARS microscopy has been increasingly
used to monitor in vitro drug dissolution from different solid oral
dosage forms and to relate dissolution with drug distribution in
formulation [318–322]. CARS was used for the real-time spatial
visualization of the change in the solid-state of theophylline from
anhydrate to the less soluble monohydrate during dissolution from
the lipid-based oral dosage form [323]. Haaser et al. employed
Raman microscopy to chemically image the component distribution and
structural changes during dissolution testing of sustained release
extrudate formulations of theophylline anhydrate in a tripalmitin matrix
with polyethylene glycol [324].
13
In situ Raman spectroscopy can also provide deeper insight into the
dissolution behaviour of amorphous APIs or dispersions. Alonzo et al.
investigated the crystallization behaviour of felodipine and indomethacin
from the slurry of amorphous forms exposed to the dissolution medium
at 25 °C and 37 °C in the absence and presence of externally added
polymers such as PVP, HPMC or HPMC-AS using a fibre-optic Raman
system equipped with an immersion probe [325]. As observed in Fig. 7a,
the conversion of amorphous indomethacin to α-indomethacin at 37 °C
could be evidenced with the disappearance of a peak unique to the amorphous form (at 753 cm−1) and the appearance of the characteristic peak
of α-indomethacin (at 1648 cm−1). A univariate quantification model
was applied based on the peak attributable at 1648 cm−1 and thus for
correlating crystallization kinetics with the dissolution profile (Fig. 7b).
Amorphous felodipine crystallized rapidly at 37 °C and slowly at 25 °C,
whereas indomethacin crystallized much slower at both temperatures,
generating a certain extent of supersaturation. Upon addition of polymer
in the solution phase, the supersaturation was maintained due to the
drastic reduction in overall crystallization rates for both the APIs.
Likewise, crystallization of APIs from piroxicam–PVP amorphous
solid dispersion prepared by cryo-milling from the slurry in simulated
gastric fluid was monitored using a Raman spectrometer coupled with
two coaxial fibre optic probes [326]. Induction of crystallization into
piroxicam monohydrate occurred within 1 min for amorphous solid
dispersion with PVP K25 and within 3 min for PVP K90 with no further
spectral change appearing after 15 min. Ferreira Molina et al. recently
published study on the drug release mechanism from cisplatin-loaded
polymeric xerogel using a setup that combines Raman spectroscopy
with synchrotron X-ray absorption spectroscopy and UV–Vis
spectroscopy [327].
4. Perspectives and future directions
Raman spectroscopy and associated analytics have had a long
development and many applications. Several specialized innovations
are continuing to evolve within pharmaceuticals as a central focus.
With relatively cost-effective fabrication, extensive varieties of Raman
instruments are available today. This includes highly sophisticated
Raman-integrated assemblies, from high performance instruments
applicable to high end research to low resolution portable instruments
for quality control, screening and material qualifications. This bifurcated
trend of technological development will continue to support a wide
diversity of multi-stage pharmaceutical research.
Dedicated Raman spectroscopic installation at the synchrotron
beamline holds novel application potential for multi-methodological
monitoring of complex phase transitions in pharmaceutical product
design and engineering. Raman integrated optical levitator cell design
can be used to understand the genesis of unique microstructure and
physical states of APIs and excipients in formulation while drying a
single droplet levitated under varying temperatures and pressure environments [46]. Thanks to the flexibility of Raman devices, the incentives
of interfacing various analytical tools with Raman have led to a booming
area of pharmaceutical nanotechnology. The field of material engineering
will also benefit from these developments. Simpler adaptation of sophisticated Raman experiments such as SERS, TERS and CARS will appeal to
untrained professionals [328]. Even the fields of diagnostics, theranostics,
and medical and in vivo research will benefit from multi-microscopic
platforms that include Raman microscopy as an integral part. For
example, a recently fabricated disposable Raman probe head for
use with magnetic resonance imaging (MRI) has shown potential as a
new avenue of in vivo diagnostic application [329,330]. Furthermore,
multimodal image-guided Raman endoscopy techniques hold promising
future potential for the real-time, in-vivo diagnosis of oesophageal
cancer [331]. Raman spectroscopy can also be considered to be a future
diagnostic tool for non-invasive glucose monitoring through skin [332].
Potential pharmaceutical research applications can demonstrate the
14
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
Fig. 7. (a) Raman spectra of indomethacin slurries showing, from top to bottom, amorphous indomethacin after exposure to buffer (37 °C) for 120 min, 90 min and 20 min, unexposed
amorphous indomethacin and α-indomethacin. (b) Normalized peak intensity (■) of α indomethacin at 1648 cm−1 as a function of time and dissolution of amorphous indomethacin
at 37 °C as a function of time (□).
Reprinted (adapted) with permission from [325] Copyright (2010) Springer Link.
commercial economic viability of transferring customary laboratorydeveloped technologies into the real world.
Innovative fibre optics-based Raman probes with operational
flexibility will further incentivize the field of pharmaceutical PAT applications. The anticipated future megatrend of continuous pharmaceutical
manufacturing within the Quality by Design (QbD) framework may
embrace the tailored Raman probe hardware that is adaptable to a wide
variety of unit operations [257,333,334]. Process monitoring for
blue-sky mini-manufacturing technologies that are envisioned as
personalized/individualized drug delivery techniques, such as medicine
printing, dripping, film casting, microfluidics, and coacervation, reckon
the use of versatile Raman sensors as an intelligent PAT system in coming
years [335]. In addition, advancements in the biomedical and diagnostic
application of Raman techniques will further facilitate personalized drug
delivery. One such example is the use of carbon nanotubes in advanced
diagnostic tools [336,337]. Portable Raman devices with improved practicalities such as unconstrained size and shape as well as simple optics are
mountable in several inaccessible parts of production lines. Another possible important application in the pharmaceutical industry is the coupling
of Raman to detect (in situ) extrinsic or intrinsic mobile micro-particles
that are optically trapped, known as Raman tweezers spectroscopy, in
products or the manufacturing process [338]. Furthermore, the simplicity
and compactness of handheld and portable Raman devices will further
increase their application for anti-counterfeiting and quality control
purposes [339]. Eventually, a juvenile field for future application
will develop around small, handheld Raman instruments that can
enable the tracking of drug products from factory to pharmacy
shelves or provide quality control at outpatient clinics and hospitals.
As an example, a recent proof-of-concept study showed that the
accuracy of quantitative analysis of hospital compounded ganciclovir
by Raman spectroscopy is comparable to that by chromatographic
analysis [340].
In addition to hardware, novel data handling and chemometric
methods, Raman libraries and algorithms will further support the
emerging instrumental development and their pharmaceutical application. Extremely large volumes of data collected during PAT monitoring
or complex Raman chemical imaging require data analysis models to
achieve efficient trending and meaningful interpretation. Moreover,
multimodal and integrated spectroscopy generates data that are
interesting for heterospectral 2D correlation treatment and extract
maximum information. Multi-dimensional correlation Raman spectroscopy studies are dominant in chemical characterizations using timeresolved measurements or as the function of various types of perturbation
such as temperature and composition [341–343]. This implies that there
are enormous opportunities in applying 2D correlation Raman spectroscopy with different external perturbations and correlation formalism for
pharmaceutical analysis, which has not been very common [344–346].
For example, such an experiment could be very useful for probing
heteromolecular interactions in complex formulations, such as amorphous solid dispersions.
Abbreviations
AFM
API
CARS
CCD
CD
CLS
CRM
DPI
DVS
DSC
FT-IR
HME
LF
MCR
NIR
NMR
NSOM
OCT
PAT
PCA
PhAT
PLS
PLS-DA
PMT
ROA
RRS
SAXS
SEM
SESORS
SORS
SPM
SRS
STM
TERS
TRS
UV
WAI
XRD
Atomic Force Microscopy
Active Pharmaceutical Ingredient
Coherent Anti-stokes Raman Scattering
Charge-Coupled Device
Circular Dichroism
Classical Least Squares
Confocal Raman Spectroscopy
Dry Powder for Inhalation
Dynamic Vapour Sorption
Differential Scanning Calorimetry
Fourier-Transform Infrared
Hot Melt Extrusion
Low Frequency
Multivariate Curve Resolution
Near-IR
Nuclear Magnetic Resonance
Near-field Scanning Optical Microscopy
Optical Coherent Tomography
Process Analytical Technology
Principal Component Analysis
Pharmaceutical Area Testing
Partial Least Square Regression
Partial Least Square Discriminant Analysis
Photo-multiplier Tube
Raman Optical Activity
Resonance Raman spectroscopy
Small-Angle X-ray Scattering
Scanning Electron Microscopy
Surface-Enhanced SORS
Spatially Offset Raman Spectroscopy
Scanning Probe Microscopy
Stimulated Raman scattering
Scanning Tunnelling Microscopy
Tip-Enhanced Raman Scattering
Transmission Raman spectroscopy
Ultraviolet
Wide Angle Illumination
X-ray Diffraction
Acknowledgements
Acknowledgement Ioan Notingher is an EPSRC Fellow (EP/
L025620/1) and acknowledges the support from the National
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
Institute for Health Research (NIHR) under its Invention for Innovation (i4i) Programme (grant numbers II-AR-0209-10012).The views
expressed are those of the author(s) and not necessarily those of the
NHS, the NIHR or the Department of Health.
• http://www.controldevelopment.com/products/probes/productprobes-raman-probe.php
• http://www.augustus-pharmaceuticals.com/alt.php?content=about
• http://www.azonano.com/article.aspx?ArticleID=1538
• http://pharmacy.ku.dk/research/section-of-pharmaceutical-tech
nology/
• http://wallpoper.com/wallpaper/pills-tablets-406870
References
[1] Y.-S. Li, J.S. Church, Raman spectroscopy in the analysis of food and pharmaceutical
nanomaterials, J. Food Drug Anal. 22 (2014) 29–48.
[2] A. Mahadevan-Jansen, Raman spectroscopy: from benchtop to bedside, in: T. Vo-Dinh
(Ed.)Biomedical Photonics Handbook 2003, pp. 759–786.
[3] S.E.J. Bell, Quantitative analysis of solid dosage formulations by Raman spectroscopy,
in: S. Šašic (Ed.), Pharmaceutical Applications of Raman Spectroscopy, John Wiley &
Sons Inc. 2007, pp. 29–64.
[4] X. Zhu, T. Xu, Q. Lin, Y. Duan, Technical development of Raman spectroscopy: from
instrumental to advanced combined technologies, Appl. Spectrosc. Rev. 49 (2013)
64–82.
[5] S. Garrigues, M. de la Guardia, Non-invasive analysis of solid samples, TrAC, Trends
Anal. Chem. 43 (2013) 161–173.
[6] A. Hédoux, L. Paccou, Y. Guinet, J.-F. Willart, M. Descamps, Using the low-frequency
Raman spectroscopy to analyze the crystallization of amorphous indomethacin,
Eur. J. Pharm. Sci. 38 (2009) 156–164.
[7] L.A. Nafie, Theory of Raman scattering, in: I.R. Lewis, H. Edwards (Eds.),Handbook of
Raman Spectroscopy: From the Research Laboratory to the Process Line 2001,
pp. 1–10.
[8] E. Smith, G. Dent, Modern Raman Spectroscopy: A Practical Approach, John Wiley
& Sons Ltd., West Sussex, 2005. 224.
[9] T. Haefele, K. Paulus, Confocal Raman microscopy in pharmaceutical development,
in: T. Dieing, O. Hollricher, J. Toporski (Eds.), Confocal Raman Microscopy, Springer,
Berlin Heidelberg 2011, pp. 165–202.
[10] F. LaPlant, Lasers, spectrographs, and detectors, in: P. Matousek, M.D. Morris (Eds.),
Emerging Raman Applications and Techniques in Biomedical and Pharmaceutical
Fields, Springer, Berlin Heidelberg 2010, pp. 1–24.
[11] B.D. Patel, P.J. Mehta, An overview: application of Raman spectroscopy in pharmaceutical field, Curr. Pharm. Anal. 6 (2010) 131–141.
[12] K. Shin, H. Chung, Wide area coverage Raman spectroscopy for reliable quantitative
analysis and its applications, Analyst 138 (2013) 3335–3346.
[13] J. Johansson, A. Sparén, O. Svensson, S. Folestad, M. Claybourn, Quantitative transmission Raman spectroscopy of pharmaceutical tablets and capsules, Appl. Spectrosc. 61
(2007) 1211–1218.
[14] J. Johansson, S. Pettersson, S. Folestad, Characterization of different laser irradiation
methods for quantitative Raman tablet assessment, J. Pharm. Biomed. Anal. 39
(2005) 510–516.
[15] M. Kim, H. Chung, Y. Woo, M. Kemper, New reliable Raman collection system using
the wide area illumination (WAI) scheme combined with the synchronous intensity
correction standard for the analysis of pharmaceutical tablets, Anal. Chim. Acta 579
(2006) 209–216.
[16] P. Matousek, I.P. Clark, E.R.C. Draper, M.D. Morris, A.E. Goodship, N. Everall, M. Towrie,
W.F. Finney, A.W. Parker, Subsurface probing in diffusely scattering media using
spatially offset raman spectroscopy, Appl. Spectrosc. 59 (2005) 393–400.
[17] N.A. Macleod, M.D. Morris, P. Matousek, Characterization of deep layers of tissue and
powders: spatially offset Raman and transmission Raman spectroscopy, in: P.
Matousek, M.D. Morris (Eds.), Emerging Raman Applications and Techniques in
Biomedical and Pharmaceutical Fields, Springer, Berlin Heidelberg 2010, pp. 47–69.
[18] P. Matousek, A.W. Parker, Bulk Raman analysis of pharmaceutical tablets, Appl.
Spectrosc. 60 (2006) 1353–1357.
[19] P. Matousek, A.W. Parker, Non-invasive probing of pharmaceutical capsules using
transmission Raman spectroscopy, J. Raman Spectrosc. 38 (2007) 563–567.
[20] R. Baker, P. Matousek, K.L. Ronayne, A.W. Parker, K. Rogers, N. Stone, Depth profiling of
calcifications in breast tissue using picosecond Kerr-gated Raman spectroscopy,
Analyst 132 (2007) 48–53.
[21] K. Buckley, P. Matousek, Recent advances in the application of transmission Raman
spectroscopy to pharmaceutical analysis, J. Pharm. Biomed. Anal. 55 (2011)
645–652.
[22] N. Everall, I. Priestnall, P. Dallin, J. Andrews, I. Lewis, K. Davis, H. Owen, M.W.
George, Measurement of spatial resolution and sensitivity in transmission and
backscattering Raman spectroscopy of opaque samples: impact on pharmaceutical
quality control and Raman tomography, Appl. Spectrosc. 64 (2010) 476–484.
[23] T. Deckert-Gaudig, M. Richter, D. Knebel, T. Jähnke, T. Jankowski, E. Stock, V.
Deckert, A modified transmission tip-enhanced Raman scattering (TERS) setup
provides access to opaque samples, Appl. Spectrosc. 68 (2014) 916–919.
15
[24] M.D. Sonntag, E.A. Pozzi, N. Jiang, M.C. Hersam, R.P. Van Duyne, Recent advances in
tip-enhanced Raman spectroscopy, J. Phys. Chem. Lett. (2014) 3125–3130.
[25] V. Deckert, Tip-enhanced Raman spectroscopy, J. Raman Spectrosc. 40 (2009)
1336–1337.
[26] E. Bailo, V. Deckert, Tip-enhanced Raman scattering, Chem. Soc. Rev. 37 (2008)
921–930.
[27] S. Schlücker, Surface-enhanced Raman spectroscopy: concepts and chemical applications, Angew. Chem. Int. Ed. 53 (2014) 4756–4795.
[28] A.J. McQuillan, The discovery of surface-enhanced Raman scattering, Notes Rec. R.
Soc. 63 (2009) 105–109.
[29] J. Wessel, Surface-enhanced optical microscopy, J. Opt. Soc. Am. B 2 (1985)
1538–1541.
[30] B. Pettinger, G. Picardi, R. Schuster, G. Ertl, Surface-enhanced and STM-tipenhanced Raman spectroscopy at metal surfaces, Single Mol. 3 (2002) 285–294.
[31] K.F. Domke, D. Zhang, B. Pettinger, Toward Raman fingerprints of single dye molecules
at atomically smooth Au(111), J. Am. Chem. Soc. 128 (2006) 14721–14727.
[32] J. Stadler, T. Schmid, R. Zenobi, Developments in and practical guidelines for tipenhanced Raman spectroscopy, Nanoscale 4 (2012) 1856–1870.
[33] M.D. Duncan, J. Reintjes, T.J. Manuccia, Imaging biological compounds using the
coherent anti-Stokes Raman scattering microscope, Opt. Eng. 24 (1985) 242352.
[34] L. Opilik, T. Schmid, R. Zenobi, Modern Raman imaging: vibrational spectroscopy on
the micrometer and nanometer scales, Annu. Rev. Anal. Chem. 6 (2013) 379–398.
[35] W. Min, C.W. Freudiger, S. Lu, X.S. Xie, Coherent nonlinear optical imaging: beyond
fluorescence microscopy, Annu. Rev. Phys. Chem. 62 (2011) 507–530.
[36] C.L. Evans, X.S. Xie, Coherent anti-Stokes Raman scattering microscopy: chemical
imaging for biology and medicine, Annu. Rev. Anal. Chem. 1 (2008) 883–909.
[37] S. Ruschin, Y. Ben‐Aryeh, Coherent trapping of a single excitation in a multimolecular
system, J. Chem. Phys. 71 (1979) 476–480.
[38] R.L. McCreery, M. Fleischmann, P. Hendra, Fiber optic probe for remote Raman
spectrometry, Anal. Chem. 55 (1983) 146–148.
[39] S.D. Schwab, R.L. McCreery, Versatile, efficient Raman sampling with fiber optics,
Anal. Chem. 56 (1984) 2199–2204.
[40] I. Latka, S. Dochow, C. Krafft, B. Dietzek, J. Popp, Fiber optic probes for linear and
nonlinear Raman applications — current trends and future development, Laser
Photonics Rev. 7 (2013) 698–731.
[41] N.J. Everall, Modeling and measuring the effect of refraction on the depth resolution of
confocal Raman microscopy, Appl. Spectrosc. 54 (2000) 773–782.
[42] A.S. Pereira, C.A. Perottoni, J.A.H. da Jornada, Raman spectroscopy as a probe for in
situ studies of pressure-induced amorphization: some illustrative examples, J.
Raman Spectrosc. 34 (2003) 578–586.
[43] H.-J. van Manen, R. Bloemenkamp, O.F. van den Brink, Focal length determination of
Raman immersion ball probes in diverse media, Appl. Spectrosc. 63 (2009) 378–380.
[44] J.B. Nanubolu, J.C. Burley, Investigating the recrystallization behavior of amorphous
paracetamol by variable temperature Raman studies and surface Raman mapping,
Mol. Pharmaceutics 9 (2012) 1544–1558.
[45] S.-Y. Lin, W.-T. Cheng, The use of hot-stage microscopy and thermal micro-Raman
spectroscopy in the study of phase transformation of metoclopramide HCl
monohydrate, J. Raman Spectrosc. 43 (2012) 1166–1170.
[46] S. Baer, C. Esen, A. Ostendorf, Phase equilibrium measurements of acoustically
levitated squalane–CO2 mixtures by Raman spectroscopy, J. Raman Spectrosc. 45
(2014) 680–685.
[47] T. Frosch, D. Yan, J. Popp, Ultrasensitive fiber enhanced UV resonance Raman sensing
of drugs, Anal. Chem. 85 (2013) 6264–6271.
[48] S. Hanf, R. Keiner, D. Yan, J. Popp, T. Frosch, Fiber-enhanced Raman multigas
spectroscopy: a versatile tool for environmental gas sensing and breath analysis,
Anal. Chem. 86 (2014) 5278–5285.
[49] K.L. Davis, M.S. Kemper, I.R. Lewis, Raman spectroscopy for monitoring real-time
processes in the pharmaceutical industry, Pharmaceutical Applications of Raman
spectroscopy, John Wiley & Sons, Inc. 2007, pp. 117–162.
[50] B. Smith-Goettler, On-line PAT applications of spectroscopy in the pharmaceutical
industry, Process Analytical Technology, John Wiley & Sons, Ltd., West Sussex
2010, pp. 439–461.
[51] G. Zhou, R. Guenard, Z. Ge, Infrared and Raman spectroscopy for process development,
Handbook of Vibrational Spectroscopy, John Wiley & Sons, Ltd., West Sussex, 2006.
[52] H. Sato, H. Shinzawa, Y. Komachi, Fiber-optic Raman probes for biomedical and
pharmaceutical applications, in: P. Matousek, M.D. Morris (Eds.), Emerging
Raman Applications and Techniques in Biomedical and Pharmaceutical Fields,
Springer, Berlin Heidelberg 2010, pp. 25–45.
[53] P. Matousek, N. Stone, Recent advances in the development of Raman spectroscopy
for deep non-invasive medical diagnosis, J. Biophotonics 6 (2013) 7–19.
[54] N.S. Eikje, Y. Ozaki, K. Aizawa, S. Arase, Fiber optic near-infrared Raman spectroscopy
for clinical noninvasive determination of water content in diseased skin and
assessment of cutaneous edema, J. Biomed. Opt. 10 (2005) 014013.
[55] P.D.A. Pudney, E.Y.M. Bonnist, P.J. Caspers, J.-P. Gorce, C. Marriot, G.J. Puppels, S.
Singleton, M.J.G. van der Wolf, A new in vivo Raman probe for enhanced applicability
to the body, Appl. Spectrosc. 66 (2012) 882–891.
[56] J.C.C. Day, N. Stone, A subcutaneous Raman needle probe, Appl. Spectrosc. 67
(2013) 349–354.
[57] A.J. Hobro, B. Lendl, Stand-off Raman spectroscopy, TrAC, Trends Anal. Chem. 28
(2009) 1235–1242.
[58] C. Eliasson, P. Matousek, Noninvasive authentication of pharmaceutical products
through packaging using spatially offset raman spectroscopy, Anal. Chem. 79 (2007)
1696–1701.
[59] W.J. Olds, E. Jaatinen, P. Fredericks, B. Cletus, H. Panayiotou, E.L. Izake, Spatially offset
Raman spectroscopy (SORS) for the analysis and detection of packaged pharmaceuticals and concealed drugs, Forensic Sci. Int. 212 (2011) 69–77.
16
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
[60] E.L. Izake, S. Sundarajoo, W. Olds, B. Cletus, E. Jaatinen, P.M. Fredericks, Standoff Raman
spectrometry for the non-invasive detection of explosives precursors in highly
fluorescing packaging, Talanta 103 (2013) 20–27.
[61] I. Malka, A. Petrushansky, S. Rosenwaks, I. Bar, Detection of explosives and latent
fingerprint residues utilizing laser pointer-based Raman spectroscopy, Appl.
Phys. B 113 (2013) 511–518.
[62] N.A. O'Brien, C.A. Hulse, D.M. Friedrich, F.J. Van Milligen, M.K. von Gunten, F. Pfeifer,
H.W. Siesler, Miniature Near-infrared (NIR) Spectrometer Engine for Handheld
Applications, 2012. (pp. 837404-837404-837408).
[63] P. Colomban, The on-site/remote Raman analysis with mobile instruments: a
review of drawbacks and success in cultural heritage studies and other associated
fields, J. Raman Spectrosc. 43 (2012) 1529–1535.
[64] R.A. Crocombe, Handheld spectrometers: the state of the art, Proc. SPIE, 8726 2013,
pp. 87260R–872614R.
[65] S. Assi, Raw material identification using dual laser handheld Raman spectroscopy,
Eur. Pharm. Rev. 18 (2013) 25–31.
[66] M. Hajjou, P. Lukulay, Potential use of handheld Raman devices as tools for screening
medicines for quality, BioPharma Asia (2014) 14–21.
[67] M. Hajjou, Y. Qin, S. Bradby, D. Bempong, P. Lukulay, Assessment of the performance of
a handheld Raman device for potential use as a screening tool in evaluating medicines
quality, J. Pharm. Biomed. Anal. 74 (2013) 47–55.
[68] B.R. Wood, P. Heraud, S. Stojkovic, D. Morrison, J. Beardall, D. McNaughton, A portable
Raman acoustic levitation spectroscopic system for the identification and environmental monitoring of algal cells, Anal. Chem. 77 (2005) 4955–4961.
[69] J. Zheng, S. Pang, T.P. Labuza, L. He, Evaluation of surface-enhanced Raman scattering
detection using a handheld and a bench-top Raman spectrometer: a comparative
study, Talanta 129 (2014) 79–85.
[70] C. Brouillette, H. Huang, W. Smith, S. Farquharson, Raman spectroscopy using
1550 nm (retina-safe) laser excitation, Appl. Spectrosc. 65 (2011) 561–563.
[71] H. Huang, C. Shende, A. Sengupta, F. Inscore, C. Brouillette, W. Smith, S. Farquharson,
Surface-enhanced Raman spectra of melamine and other chemicals using a
1550 nm (retina-safe) laser, J. Raman Spectrosc. 43 (2012) 701–705.
[72] E. Lee, Raman spectral imaging on pharmaceutical products, Infrared and Raman
spectroscopic imaging, Wiley-VCH Verlag GmbH & Co. KGaA 2009, pp. 377–402.
[73] B. Cletus, W. Olds, P.M. Fredericks, E. Jaatinen, E.L. Izake, Real-time detection of
concealed chemical hazards under ambient light conditions using Raman spectroscopy, J. Forensic Sci. 58 (2013) 1008–1014.
[74] R.A. Grothe, D.A. Wright, in: U.S.P.a.T. Office (Ed.), Methods of Automated Spectral
Peak Detection and Quantification Without User Input, Thermo Finnigan LLC, 2011.
[75] J. Aaltonen, C.J. Strachan, K. Pöllänen, J. Yliruusi, J. Rantanen, Hyphenated spectroscopy
as a polymorph screening tool, J. Pharm. Biomed. Anal. 44 (2007) 477–483.
[76] A. Kern, W.J. Collins, R.T. Cambron, N.L. Redman-Furey, Use of a TG/DTA/Raman
system to monitor dehydration and phase conversions, J. ASTM Int. 2 (2005) 1–10.
[77] W.J. Collins, C. DuBois, R.T. Cambron, N.L. Redman-Furey, A. Bigalow Kern, Development and evaluation of a TG/DTA/Raman System, J. ASTM Int. 2 (2005) 1–10.
[78] Simultaneous DSC–Raman analysis of the curing of an epoxy thermoset using
RamanRxn1™, AZO Materials, 2014.
[79] L. Wei, J. Jiang, M. Shan, W. Chen, Y. Deng, G. Xue, D. Zhou, Integration of ultrafast
scanning calorimetry with micro-Raman spectroscopy for investigation of metastable
materials, Rev. Sci. Instrum. 85 (2014) 074901.
[80] B.H. Stuart, A Fourier transform Raman study of water sorption by Nylon 6, Polym.
Bull. 33 (1994) 681–686.
[81] B.H. Stuart, A Fourier transform Raman study of water sorption by Kevlar-49,
Polym. Bull. 35 (1995) 727–733.
[82] D. Burnett, D. Pearse, N. Malde, Combining Raman Spectroscopy With Gravimetric
Vapour Sorption Analysis for Pharmaceutical Materials Surface Measurement Systems
Ltd, Application Note, 2012.
[83] N. Sandler, M. Savolainen, A. Saupe, C.J. Strachan, T. Rades, Applications of Raman
spectroscopy in aqueous environments, Pharm. Technol. Eur. 19 (2007) 1–8.
[84] A.D. Gift, L.S. Taylor, Hyphenation of Raman spectroscopy with gravimetric analysis
to interrogate water–solid interactions in pharmaceutical systems, J. Pharm.
Biomed. Anal. 43 (2007) 14–23.
[85] K.L.A. Chan, O.S. Fleming, S.G. Kazarian, D. Vassou, G.D. Chryssikos, V. Gionis, Polymorphism and devitrification of nifedipine under controlled humidity: a combined FTRaman, IR and Raman microscopic investigation J. Raman Spectrosc. 35 (2004)
353–359.
[86] M.P. Feth, J. Jurascheck, M. Spitzenberg, J. Dillenz, G. Bertele, H. Stark, New technology
for the investigation of water vapor sorption–-induced crystallographic form transformations of chemical compounds: a water vapor sorption gravimetry–dispersive
Raman spectroscopy coupling, J. Pharm. Sci. 100 (2011) 1080–1092.
[87] K. Williams, R. Bennett, A. Brooker, R. Bormett, T. Prusnick, New methods in
Raman spectroscopy-combining other microscopes, Microsc. Microanal. 9 (2003)
1094–1095.
[88] D.A. Schmidt, I. Kopf, E. Bründermann, A matter of scale: from far-field microscopy
to near-field nanoscopy, Laser Photonics Rev. 6 (2012) 296–332.
[89] P. Dorozhkin, E. Kuznetsov, A. Schokin, S. Timofeev, V. Bykov, AFM + Raman
Microscopy + SNOM + tip-enhanced Raman: instrumentation and applications,
Microsc. Today (2010) 28–32.
[90] E.d.B. Santos, F.A. Sigoli, I.O. Mazali, Study of structure of the TiO2–MoO3 bilayer
films by Raman spectroscopy, Mater. Res. Bull. 60 (2014) 242–246.
[91] B.R. Wood, M. Asghari-Khiavi, E. Bailo, D. McNaughton, V. Deckert, Detection of
nano-oxidation sites on the surface of hemoglobin crystals using tip-enhanced
Raman scattering, Nano Lett. 12 (2012) 1555–1560.
[92] K.A. Antonio, Z.D. Schultz, Advances in biomedical Raman microscopy, Anal. Chem.
86 (2013) 30–46.
[93] Products: Optik & Photonik 2/2014, Opt. Photon. 9 (2014) 20–26.
[94] G. Wille, X. Bourrat, N. Maubec, A. Lahfid, Raman-in-SEM, a multimodal and multiscale
analytical tool: performance for materials and expertise, Micron 67 (2014) 50–64.
[95] M. Baranska, K. Chruszcz-Lipska, Raman optical activity: a powerful technique
to investigate essential oil components, Nat. Prod. Commun. 5 (2010)
1417–1420.
[96] A.J. Thorvaldsen, B. Gao, K. Ruud, M. Fedorovsky, G. Zuber, W. Hug, Efficient calculation
of ROA tensors with analytical gradients and fragmentation, Chirality 24 (2012)
1018–1030.
[97] N. Daugey, T. Brotin, N. Vanthuyne, D. Cavagnat, T. Buffeteau, Raman optical activity of
enantiopure cryptophanes, J. Phys. Chem. B 118 (2014) 5211–5217.
[98] K. Chruszcz-Lipska, E.W. Blanch, Raman optical activity of biological samples, in: M.
Baranska (Ed.), Optical Spectroscopy and Computational Methods in Biology and
Medicine, Springer, Netherlands 2014, pp. 61–81.
[99] C. Merten, L.D. Barron, L. Hecht, C. Johannessen, Determination of the helical screw
sense and side-group chirality of a synthetic chiral polymer from Raman optical
activity, Angew. Chem. Int. Ed. 50 (2011) 9973–9976.
[100] E. Blanch, L. Barron, Raman optical activity of biological molecules, in: P. Matousek,
M.D. Morris (Eds.), Emerging Raman Applications and Techniques in Biomedical
and Pharmaceutical Fields, Springer, Berlin Heidelberg 2010, pp. 153–177.
[101] K. Chruszcz-Lipska, A. Jaworska, E. Szczurek, M. Baranska, (−)-R-Mevalonolactone
studied by ROA and SERS spectroscopy, Chirality 26 (2014) 453–461.
[102] K. Chruszcz-Lipska, E.W. Blanch, In situ analysis of chiral components of pichtae
essential oil by means of ROA spectroscopy: experimental and theoretical Raman
and ROA spectra of bornyl acetate, J. Raman Spectrosc. 43 (2012) 286–293.
[103] L.D. Barron, F. Zhu, L. Hecht, G.E. Tranter, N.W. Isaacs, Raman optical activity: an
incisive probe of molecular chirality and biomolecular structure, J. Mol. Struct.
834–836 (2007) 7–16.
[104] L. Ashton, B. Czarnik-Matusewicz, E.W. Blanch, Application of two-dimensional
correlation analysis to Raman optical activity, J. Mol. Struct. 799 (2006) 61–71.
[105] L. Ashton, L.D. Barron, L. Hecht, J. Hyde, E.W. Blanch, Two-dimensional Raman and
Raman optical activity correlation analysis of the [small alpha]-helix-to-disordered
transition in poly(L-glutamic acid), Analyst 132 (2007) 468–479.
[106] M.A. Adams, S.F. Parker, F. Fernandez-Alonso, D.J. Cutler, C. Hodges, A. King,
Simultaneous neutron scattering and Raman scattering, Appl. Spectrosc. 63
(2009) 727–732.
[107] W. van Beek, A. Urakawa, M. Milanesio, XRD–Raman and modulation excitation
spectroscopy, In-situ Characterization of Heterogeneous Catalysts, John Wiley &
Sons, Inc. 2013, pp. 411–439.
[108] M.A. Newton, W. van Beek, Combining synchrotron-based X-ray techniques with
vibrational spectroscopies for the in situ study of heterogeneous catalysts: a view
from a bridge, Chem. Soc. Rev. 39 (2010) 4845–4863.
[109] G.K. Bryant, H.F. Gleeson, A.J. Ryan, J.P.A. Fairclough, D. Bogg, J.G.P. Goossens, W.
Bras, Raman spectroscopy combined with small angle X-ray scattering and wide
angle X-ray scattering as a tool for the study of phase transitions in polymers,
Rev. Sci. Instrum. 69 (1998) 2114–2117.
[110] J. Xu, L. Ouyang, W. Mao, X.-J. Yang, X.-C. Xu, J.-J. Su, T.-Z. Zhuang, H. Li, Y.-F. Han,
Operando and kinetic study of low-temperature, lean-burn methane combustion
over a Pd/γ-Al2O3 catalyst, ACS Catal. 2 (2012) 261–269.
[111] A.M. Beale, A.M.J. van der Eerden, K. Kervinen, M.A. Newton, B.M. Weckhuysen,
Adding a third dimension to operando spectroscopy: a combined UV–Vis, Raman
and XAFS setup to study heterogeneous catalysts under working conditions,
Chem. Commun. 3015–3017 (2005).
[112] W.V. Beek, F. Carniato, S. Kumar, G. Croce, E. Boccaleri, M. Milanesio, Studying
modifications and reactions in materials by simultaneous Raman and X-ray powder
diffraction at non-ambient conditions: methods and applications, Phase Transit. 82
(2009) 293–302.
[113] U. Bentrup, J. Radnik, U. Armbruster, A. Martin, J. Leiterer, F. Emmerling, A.
Brückner, Linking simultaneous in situ WAXS/SAXS/Raman with Raman/ATR/
UV–vis spectroscopy: comprehensive insight into the synthesis of molybdate
catalyst precursors, Top. Catal. 52 (2009) 1350–1359.
[114] C. Kongmark, V. Martis, A. Rubbens, C. Pirovano, A. Lofberg, G. Sankar, E. BordesRichard, R.-N. Vannier, W. Van Beek, Elucidating the genesis of Bi2MoO6 catalyst
by combination of synchrotron radiation experiments and Raman scattering,
Chem. Commun. (2009) 4850–4852.
[115] S. Kumar, F. Carniato, A. Arrais, G. Croce, E. Boccaleri, L. Palin, W. van Beek, M.
Milanesio, Investigating surface vs bulk kinetics in the formation of a molecular
complex via solid-state reaction by simultaneous Raman/X-ray powder diffraction,
Cryst. Growth Des. 9 (2009) 3396–3404.
[116] W. van Beek, O.V. Safonova, G. Wiker, H. Emerich, SNBL, a dedicated beamline for combined in situ X-ray diffraction, X-ray absorption and Raman scattering experiments,
Phase Transit. 84 (2011) 726–732.
[117] E. Boccaleri, F. Carniato, G. Croce, D. Viterbo, W. van Beek, H. Emerich, M. Milanesio, In
situ simultaneous Raman/high-resolution X-ray powder diffraction study of transformations occurring in materials at non-ambient conditions, J. Appl. Crystallogr. 40
(2007) 684–693.
[118] R. Hirose, T. Yoshioka, H. Yamamoto, K.R. Reddy, D. Tahara, K. Hamada, K. Tashiro,
In-house simultaneous collection of small-angle X-ray scattering, wide-angle X-ray
diffraction and Raman scattering data from polymeric materials, J. Appl. Crystallogr.
47 (2014) 922–930.
[119] D.S. Wragg, R.E. Johnsen, M. Balasundaram, P. Norby, H. Fjellvåg, A. Grønvold, T.
Fuglerud, J. Hafizovic, Ø.B. Vistad, D. Akporiaye, SAPO-34 methanol-to-olefin catalysts
under working conditions: a combined in situ powder X-ray diffraction, mass
spectrometry and Raman study, J. Catal. 268 (2009) 290–296.
[120] D.S. Wragg, A. Grønvold, A. Voronov, P. Norby, H. Fjellvåg, Combined XRD
and Raman studies of coke types found in SAPO-34 after methanol and propene
conversion, Microporous Mesoporous Mater. 173 (2013) 166–174.
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
[121] S. Magazù, F. Migliardo, S.F. Parker, Vibrational properties of bioprotectant mixtures of
trehalose and glycerol, J. Phys. Chem. B 115 (2011) 11004–11009.
[122] G. Bellavia, L. Paccou, Y. Guinet, A. Hédoux, How does glycerol enhance the
bioprotective properties of trehalose? Insight from protein–solvent dynamics, J.
Phys. Chem. B 118 (2014) 8928–8934.
[123] F. Migliardo, S. Magazù, M.A. Gonzalez, C. Mondelli, Study of the Boson peak and
fragility of bioprotectant glass-forming mixtures by neutron scattering, Adv.
Mater. Sci. Eng. 2013 (2013) 6.
[124] T. Meyer, M. Chemnitz, M. Baumgartl, T. Gottschall, T. Pascher, C. Matthäus, B.F.M.
Romeike, B.R. Brehm, J. Limpert, A. Tünnermann, M. Schmitt, B. Dietzek, J. Popp,
Expanding multimodal microscopy by high spectral resolution coherent anti-Stokes
Raman scattering imaging for clinical disease diagnostics, Anal. Chem. 85 (2013)
6703–6715.
[125] N. Vogler, A. Medyukhina, I. Latka, S. Kemper, M. Böhm, B. Dietzek, J. Popp, Towards
multimodal nonlinear optical tomography — experimental methodology, Laser
Phys. Lett. 8 (2011) 617–624.
[126] O.R. Šćepanović, Z. Volynskaya, C.-R. Kong, L.H. Galindo, R.R. Dasari, M.S. Feld, A multimodal spectroscopy system for real-time disease diagnosis, Rev. Sci. Instrum. 80 (2009).
[127] W. Wang, A Multimodal Spectrometer for Raman Scattering and Near-infrared
Absorption Measurement, University of Manitoba, 2012.
[128] F.C. Clarke, M.J. Jamieson, D.A. Clark, S.V. Hammond, R.D. Jee, A.C. Moffat, Chemical
image fusion. The synergy of FT-NIR and Raman mapping microscopy to enable a
more complete visualization of pharmaceutical formulations, Anal. Chem. 73
(2001) 2213–2220.
[129] M.B. Lopes, J.M. Bioucas-Dias, M.A. Figueiredo, J.C. Wolff, N. Mistry, J. Warrack,
Comparison of near infrared and Raman hyperspectral unmixing performances for
chemical identification of pharmaceutical tablets, Hyperspectral Image and Signal
Processing: Evolution in Remote Sensing (WHISPERS), 2011 3rd Workshop on
2011, pp. 1–4.
[130] N. Djeu, in: U.S.P.a.T. Office (Ed.), Combined Fiber-optic Absorption and Emission
Measurement Apparatus, University of South Florida, 2010.
[131] W. Wang, J. Paliwal, A multimodal spectrometer for Raman scattering and nearinfrared absorption measurement, Vib. Spectrosc. 74 (2014) 13–19.
[132] C.A. Patil, J. Kalkman, D.J. Faber, J.S. Nyman, T.G. van Leeuwen, A. Mahadevan-Jansen,
Integrated system for combined Raman spectroscopy–spectral domain optical
coherence tomography, J. Biomed. Opt. 16 (2011) (011007-1–011007-10).
[133] K.M. Khan, H. Krishna, S.K. Majumder, K.D. Rao, P.K. Gupta, Depth-sensitive Raman
spectroscopy combined with optical coherence tomography for layered tissue
analysis, J. Biophotonics 7 (2014) 77–85.
[134] D. Pestov, X. Wang, G.O. Ariunbold, R.K. Murawski, V.A. Sautenkov, A. Dogariu, A.V.
Sokolov, M.O. Scully, Single-shot detection of bacterial endospores via coherent
Raman spectroscopy, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 422–427.
[135] S.J. Fraser, K.C. Gordon, Raman spectroscopy in the study of pharmaceuticals: the
problems and solutions to sub-sampling and data analysis, Eur. Pharm. Rev. 19
(2014) 3–8.
[136] N. Hayazawa, Y. Saito, M. Motohashi, M. Iyoki, S. Kawata, Nanoscale characterization of
localized strain in crystals by tip-enhanced Raman spectroscope in reflection mode,
Proc. SPIE, 6324 2006, pp. 63240L–63248L.
[137] T. Yamamoto, Y. Murakami, J. Motoyanagi, T. Fukushima, S. Maruyama, M. Kato,
An analytical system for single nanomaterials: Combination of capillary
electrophoresis with Raman spectroscopy or with scanning probe microscopy
for individual single-walled carbon nanotube analysis, Anal. Chem. 81 (2009)
7336–7341.
[138] A. Lucotti, M. Tommasini, M. Casella, A. Morganti, F. Gramatica, G. Zerbi, TLC–surface
enhanced Raman scattering of apomorphine in human plasma, Vib. Spectrosc. 62
(2012) 286–291.
[139] H.F. Boelens, R.J. Dijkstra, P.H. Eilers, F. Fitzpatrick, J.A. Westerhuis, New background
correction method for liquid chromatography with diode array detection, infrared
spectroscopic detection and Raman spectroscopic detection, J. Chromatogr. A 1057
(2004) 21–30.
[140] S.D. Cooper, M.M. Robson, D.N. Batchelder, K.D. Bartle, Development of a universal
Raman detector for microchromatography, Chromatographia 44 (1997) 257–262.
[141] A. Ruddick, D.N. Batchelder, K.D. Bartle, A.C. Gilby, G.D. Pitt, Development of a Raman
detector for capillary electrophoresis, Appl. Spectrosc. 54 (2000) 1857–1863.
[142] P.C. Chen, C.C. Joyner, S.T. Patrick, R.M. Royster, L.L. Ingham, Gas chromatography–
multiplex coherent Raman spectroscopy, Anal. Chem. 75 (2003) 3066–3072.
[143] J.M. Peter, Application of infrared and Raman spectroscopy for detection in liquid
chromatographic separations, Hyphenated and Alternative Methods of Detection
in Chromatography, CRC Press 2011, pp. 99–144.
[144] J. Rantanen, H. Wikström, F.E. Rhea, L.S. Taylor, Improved understanding of factors
contributing to quantification of anhydrate/hydrate powder mixtures, Appl.
Spectrosc. 59 (2005) 942–951.
[145] M. Fransson, J. Johansson, A. Sparén, O. Svensson, Comparison of multivariate
methods for quantitative determination with transmission Raman spectroscopy
in pharmaceutical formulations, J. Chemom. 24 (2010) 674–680.
[146] J.X. Wu, F. Van den Berg, J. Rantanen, T. Rades, M. Yang, Current advances and future
trends in characterizing poorly water-soluble drugs using spectroscopic, imaging
and data analytical techniques, Curr. Pharm. Des. 20 (2014) 436–453.
[147] E. Simone, A.N. Saleemi, Z.K. Nagy, Application of quantitative Raman spectroscopy for
the monitoring of polymorphic transformation in crystallization processes using a
good calibration practice procedure, Chem. Eng. Res. Des. 92 (2014) 594–611.
[148] N.P.A. Christensen, C. Cornett, J. Rantanen, Role of excipients on solid-state properties
of piroxicam during processing, J. Pharm. Sci. 101 (2012) 1202–1211.
[149] J. Cornel, C. Lindenberg, M. Mazzotti, Quantitative application of in situ ATR-FTIR and
Raman spectroscopy in crystallization processes, Ind. Eng. Chem. Res. 47 (2008)
4870–4882.
17
[150] A. Heinz, M. Savolainen, T. Rades, C.J. Strachan, Quantifying ternary mixtures of
different solid-state forms of indomethacin by Raman and near-infrared spectroscopy,
Eur. J. Pharm. Sci. 32 (2007) 182–192.
[151] M.C. Breitkreitz, R.J. Poppi, Trends in Raman chemical imaging, Biomed. Spectrosc.
Imaging 1 (2012) 159–183.
[152] K.C. Gordon, C.M. McGoverin, Raman mapping of pharmaceuticals, Int. J. Pharm.
417 (2011) 151–162.
[153] S. Farquharson, C. Shende, F.E. Inscore, P. Maksymiuk, A. Gift, Analysis of 5-fluorouracil
in saliva using surface-enhanced Raman spectroscopy, J. Raman Spectrosc. 36 (2005)
208–212.
[154] P.J. Mak, I.G. Denisov, Y.V. Grinkova, S.G. Sligar, J.R. Kincaid, Defining CYP3A4 structural
responses to substrate binding. Raman spectroscopic studies of a nanodiscincorporated mammalian cytochrome P450, J. Am. Chem. Soc. 133 (2011) 1357–1366.
[155] A.Y. Lee, D. Erdemir, A.S. Myerson, Crystal polymorphism in chemical process development, Annu. Rev. Chem. Biomol. Eng. 2 (2011) 259–280.
[156] S.R. Chemburkar, J. Bauer, K. Deming, H. Spiwek, K. Patel, J. Morris, R. Henry, S.
Spanton, W. Dziki, W. Porter, J. Quick, P. Bauer, J. Donaubauer, B.A. Narayanan, M.
Soldani, D. Riley, K. McFarland, Dealing with the impact of ritonavir polymorphs on
the late stages of bulk drug process development, Org. Process Res. Dev. 4 (2000)
413–417.
[157] M. Meyer, A. Straughn, E. Jarvi, G. Wood, F. Pelsor, V. Shah, The bioinequivalence of
carbamazepine tablets with a history of clinical failures, Pharm. Res. 9 (1992)
1612–1616.
[158] L.Y. Pfund, A.J. Matzger, Towards exhaustive and automated high-throughput
screening for crystalline polymorphs, ACS Comb. Sci. 16 (2014) 309–313.
[159] P. Láng, V. Kiss, R. Ambrus, G. Farkas, P. Szabó-Révész, Z. Aigner, E. Várkonyi, Polymorph screening of an active material, J. Pharm. Biomed. Anal. 84 (2013) 177–183.
[160] T. Kojima, S. Tsutsumi, K. Yamamoto, Y. Ikeda, T. Moriwaki, High-throughput
cocrystal slurry screening by use of in situ Raman microscopy and multi-well
plate, Int. J. Pharm. 399 (2010) 52–59.
[161] S. Al-Dulaimi, A. Aina, J. Burley, Rapid polymorph screening on milligram quantities of
pharmaceutical material using phonon-mode Raman spectroscopy, CrystEngComm
12 (2010) 1038–1040.
[162] J. Aaltonen, M. Allesø, S. Mirza, V. Koradia, K.C. Gordon, J. Rantanen, Solid form
screening — a review, Eur. J. Pharm. Biopharm. 71 (2009) 23–37.
[163] M. Allesø, S. Velaga, A. Alhalaweh, C. Cornett, M.A. Rasmussen, F.v.d. Berg, H.L.d.
Diego, J. Rantanen, Near-infrared spectroscopy for cocrystal screening. A comparative
study with Raman spectroscopy, Anal. Chem. 80 (2008) 7755–7764.
[164] M.L. Peterson, S.L. Morissette, C. McNulty, A. Goldsweig, P. Shaw, M. LeQuesne, J.
Monagle, N. Encina, J. Marchionna, A. Johnson, J. Gonzalez-Zugasti, A.V. Lemmo,
S.J. Ellis, M.J. Cima, Ö. Almarsson, Iterative high-throughput polymorphism studies
on acetaminophen and an experimentally derived structure for form III, J. Am.
Chem. Soc. 124 (2002) 10958–10959.
[165] S. Goyal, M.R. Thorson, G.G.Z. Zhang, Y. Gong, P.J.A. Kenis, Microfluidic approach to
cocrystal screening of pharmaceutical parent compounds, Cryst. Growth Des. 12
(2012) 6023–6034.
[166] M.R. Thorson, S. Goyal, Y. Gong, G.G.Z. Zhang, P.J.A. Kenis, Microfluidic approach to
polymorph screening through antisolvent crystallization, CrystEngComm 14
(2012) 2404–2412.
[167] A. Hédoux, Y. Guinet, M. Descamps, The contribution of Raman spectroscopy to the
analysis of phase transformations in pharmaceutical compounds, Int. J. Pharm. 417
(2011) 17–31.
[168] A. Hédoux, Y. Guinet, L. Paccou, F. Danède, P. Derollez, Polymorphic transformation
of anhydrous caffeine upon grinding and hydrostatic pressurizing analyzed by
low-frequency raman spectroscopy, J. Pharm. Sci. 102 (2013) 162–170.
[169] A.M. Amado, M.M. Nolasco, P.J.A. Ribeiro-Claro, Probing pseudopolymorphic transitions in pharmaceutical solids using Raman spectroscopy: hydration and dehydration
of theophylline, J. Pharm. Sci. 96 (2007) 1366–1379.
[170] M. Sardo, A.M. Amado, P.J.A. Ribeiro-Claro, Pseudopolymorphic transitions of
niclosamide monitored by Raman spectroscopy, J. Raman Spectrosc. 39 (2008)
1915–1924.
[171] N. Chieng, T. Rades, J. Aaltonen, An overview of recent studies on the analysis of
pharmaceutical polymorphs, J. Pharm. Biomed. Anal. 55 (2011) 618–644.
[172] S.X.M. Boerrigter, C.J.M. van den Hoogenhof, H. Meekes, P. Bennema, E. Vlieg,
P.J.C.M. van Hoof, In situ observation of epitaxial polymorphic nucleation of the
model steroid methyl analogue 17 norethindrone, J. Phys. Chem. B 106 (2002)
4725–4731.
[173] F. Tian, J.A. Zeitler, C.J. Strachan, D.J. Saville, K.C. Gordon, T. Rades, Characterizing
the conversion kinetics of carbamazepine polymorphs to the dihydrate in aqueous
suspension using Raman spectroscopy, J. Pharm. Biomed. Anal. 40 (2006) 271–280.
[174] S. Piqueras, L. Duponchel, R. Tauler, A. de Juan, Monitoring polymorphic transformations by using in situ Raman hyperspectral imaging and image multiset analysis,
Anal. Chim. Acta 819 (2014) 15–25.
[175] Y. Hu, H. Wikström, S.R. Byrn, L.S. Taylor, Estimation of the transition
temperature for an enantiotropic polymorphic system from the transformation
kinetics monitored using Raman spectroscopy, J. Pharm. Biomed. Anal. 45 (2007)
546–551.
[176] T. Gnutzmann, R. Kahlau, S. Scheifler, F. Friedrichs, E.A. Rossler, K. Rademann, F.
Emmerling, Crystal growth rates and molecular dynamics of nifedipine,
CrystEngComm 15 (2013) 4062–4069.
[177] L. Taylor, A. Williams, P. York, Particle size dependent molecular rearrangements
during the dehydration of trehalose dihydrate-in situ FT-Raman spectroscopy,
Pharm. Res. 15 (1998) 1207–1214.
[178] H. Wikström, J. Rantanen, A.D. Gift, L.S. Taylor, Toward an understanding of the factors
influencing anhydrate-to-hydrate transformation kinetics in aqueous environments,
Cryst. Growth Des. 8 (2008) 2684–2693.
18
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
[179] K.-S. Lee, K.-J. Kim, J. Ulrich, In situ monitoring of cocrystallization of salicylic acid4,4′-dipyridyl in solution using Raman spectroscopy, Cryst. Growth Des. 14 (2014)
2893–2899.
[180] N. Rodriguez-Hornedo, S.J. Nehm, K.F. Seefeldt, Y. Pagan-Torres, C.J. Falkiewicz,
Reaction crystallization of pharmaceutical molecular complexes, Mol. Pharmaceutics
3 (2006) 362–367.
[181] J. Huang, M. Dali, Evaluation of integrated Raman-DSC technology in early pharmaceutical development: characterization of polymorphic systems, J. Pharm. Biomed.
Anal. 86 (2013) 92–99.
[182] H.R.H. Ali, H.G.M. Edwards, M.D. Hargreaves, T. Munshi, I.J. Scowen, R.J. Telford,
Vibrational spectroscopic characterisation of salmeterol xinafoate polymorphs
and a preliminary investigation of their transformation using simultaneous in
situ portable Raman spectroscopy and differential scanning calorimetry, Anal.
Chim. Acta 620 (2008) 103–112.
[183] J.F. Kauffman, L.M. Batykefer, D.D. Tuschel, Raman detected differential scanning
calorimetry of polymorphic transformations in acetaminophen, J. Pharm. Biomed.
Anal. 48 (2008) 1310–1315.
[184] H.R.H. Ali, H.G.M. Edwards, I.J. Scowen, Insight into thermally induced solid-state
polymorphic transformation of sulfathiazole using simultaneous in situ Raman
spectroscopy and differential scanning calorimetry, J. Raman Spectrosc. 40 (2009)
887–892.
[185] M.P.M. Marques, R. Valero, S.F. Parker, J. Tomkinson, L.A.E. Batista de Carvalho,
Polymorphism in cisplatin anticancer drug, J. Phys. Chem. B 117 (2013) 6421–6429.
[186] Á. Munroe, D.M. Croker, Å.C. Rasmuson, B.K. Hodnett, Solution-mediated polymorphic
transformation of FV sulphathiazole, Cryst. Growth Des. 14 (2014) 3466–3471.
[187] D. Raijada, A. Bond, F. Larsen, C. Cornett, H. Qu, J. Rantanen, Exploring the solid-form
landscape of pharmaceutical hydrates: transformation pathways of the sodium
naproxen anhydrate–hydrate system, Pharm. Res. 30 (2013) 280–289.
[188] L. Helmdach, M.P. Feth, J. Ulrich, Online analytical investigations on solvent-,
temperature- and water vapour-induced phase transformations of citric acid,
Cryst. Res. Technol. 47 (2012) 967–984.
[189] M.P. Feth, N. Nagel, B. Baumgartner, M. Brockelmann, D. Rigal, B. Otto, M.
Spitzenberg, M. Schulz, B. Becker, F. Fischer, C. Petzoldt, Challenges in the development of hydrate phases as active pharmaceutical ingredients—an example, Eur. J.
Pharm. Sci. 42 (2011) 116–129.
[190] M. Savolainen, A. Heinz, C. Strachan, K.C. Gordon, J. Yliruusi, T. Rades, N. Sandler,
Screening for differences in the amorphous state of indomethacin using multivariate visualization, Eur. J. Pharm. Sci. 30 (2007) 113–123.
[191] Z. Jurasekova, C. Domingo, J.V. Garcia-Ramos, S. Sanchez-Cortes, Effect of pH on the
chemical modification of quercetin and structurally related flavonoids characterized
by optical (UV–visible and Raman) spectroscopy, Phys. Chem. Chem. Phys. 16
(2014) 12802–12811.
[192] S. Farquharson, A. Gift, P. Maksymiuk, F. Inscore, Surface-enhanced Raman spectra
of VX and its hydrolysis products, Appl. Spectrosc. 59 (2005) 654–660.
[193] P.J. Skrdla, D. Zhang, Disproportionation of a crystalline citrate salt of a developmental
pharmaceutical compound: characterization of the kinetics using pH monitoring and
online Raman spectroscopy plus quantitation of the crystalline free base form in
binary physical mixtures using FT-Raman, XRPD and DSC, J. Pharm. Biomed. Anal.
90 (2014) 186–191.
[194] Z.-Q. Wen, X. Cao, A. Vance, Conformation and side chains environments of recombinant human interleukin-1 receptor antagonist (rh-IL-1ra) probed by raman, raman
optical activity, and UV–resonance raman spectroscopy, J. Pharm. Sci. 97 (2008)
2228–2241.
[195] A. Hédoux, L. Paccou, Y. Guinet, Relationship between β-relaxation and structural
stability of lysozyme: microscopic insight on thermostabilization mechanism by
trehalose from Raman spectroscopy experiments, J. Chem. Phys. 140 (2014) 225102.
[196] P.-Y. Sacré, P. Lebrun, P.-F. Chavez, C.D. Bleye, L. Netchacovitch, E. Rozet, R.
Klinkenberg, B. Streel, P. Hubert, E. Ziemons, A new criterion to assess distributional
homogeneity in hyperspectral images of solid pharmaceutical dosage forms, Anal.
Chim. Acta 818 (2014) 7–14.
[197] A. Kuriyama, Y. Ozaki, Assessment of active pharmaceutical ingredient particle size
in tablets by raman chemical imaging validated using polystyrene microsphere
size standards, AAPS PharmSciTech 15 (2014) 375–387.
[198] N. Scoutaris, M.R. Alexander, P.R. Gellert, C.J. Roberts, Inkjet printing as a novel
medicine formulation technique, J. Control. Release 156 (2011) 179–185.
[199] M.-A. Yessine, J.-C. Leroux, Membrane-destabilizing polyanions: interaction with
lipid bilayers and endosomal escape of biomacromolecules, Adv. Drug Deliv. Rev. 56
(2004) 999–1021.
[200] T.J. Kamerzell, R. Esfandiary, S.B. Joshi, C.R. Middaugh, D.B. Volkin, Protein–excipient
interactions: mechanisms and biophysical characterization applied to protein
formulation development, Adv. Drug Deliv. Rev. 63 (2011) 1118–1159.
[201] S. Wartewig, R.H.H. Neubert, Pharmaceutical applications of Mid-IR and Raman
spectroscopy, Adv. Drug Deliv. Rev. 57 (2005) 1144–1170.
[202] J. Chen, H. Liu, C. Zhao, G. Qin, G. Xi, T. Li, X. Wang, T. Chen, One-step reduction
and PEGylation of graphene oxide for photothermally controlled drug delivery,
Biomaterials 35 (2014) 4986–4995.
[203] D. Hasa, B. Perissutti, M.R. Chierotti, R. Gobetto, I. Grabnar, A. Bonifacio, S.
Dall'Acqua, S. Invernizzi, D. Voinovich, Mechanochemically induced disordered
structures of vincamine: the different mediation of two cross-linked polymers,
Int. J. Pharm. 436 (2012) 41–57.
[204] D.H. Choi, K.H. Kim, J.S. Park, S.H. Jeong, K. Park, Evaluation of drug delivery profiles
in geometric three-layered tablets with various mechanical properties, in vitro–
in vivo drug release, and Raman imaging, J. Control. Release 172 (2013) 763–772.
[205] P. Tong, L. Taylor, G. Zografi, Influence of alkali metal counterions on the glass transition temperature of amorphous indomethacin salts, Pharm. Res. 19 (2002)
649–654.
[206] L.S. Taylor, F.W. Langkilde, G. Zografi, Fourier transform Raman spectroscopic study
of the interaction of water vapor with amorphous polymers, J. Pharm. Sci. 90
(2001) 888–901.
[207] L. Taylor, G. Zografi, Spectroscopic characterization of interactions between PVP
and indomethacin in amorphous molecular dispersions, Pharm. Res. 14 (1997)
1691–1698.
[208] J. Breitenbach, W. Schrof, J. Neumann, Confocal Raman-spectroscopy: analytical
approach to solid dispersions and mapping of drugs, Pharm. Res. 16 (1999)
1109–1113.
[209] F. Qian, J. Huang, Q. Zhu, R. Haddadin, J. Gawel, R. Garmise, M. Hussain, Is a distinctive
single Tg a reliable indicator for the homogeneity of amorphous solid dispersion? Int. J.
Pharm. 395 (2010) 232–235.
[210] R. Palermo, S. Short, C. Anderson, H. Tian, J. Drennen III, Determination of figures of
merit for near-infrared, raman and powder X-ray diffraction by net analyte signal
analysis for a compacted amorphous dispersion with spiked crystallinity, J. Pharm.
Innov. 7 (2012) 56–68.
[211] W. Sinclair, M. Leane, G. Clarke, A. Dennis, M. Tobyn, P. Timmins, Physical stability
and recrystallization kinetics of amorphous ibipinabant drug product by Fourier
transform Raman spectroscopy, J. Pharm. Sci. 100 (2011) 4687–4699.
[212] D. Esrafilzadeh, J.M. Razal, S.E. Moulton, E.M. Stewart, G.G. Wallace, Multifunctional
conducting fibres with electrically controlled release of ciprofloxacin, J. Control.
Release 169 (2013) 313–320.
[213] A. Belu, C. Mahoney, K. Wormuth, Chemical imaging of drug eluting coatings:
combining surface analysis and confocal Raman microscopy, J. Control. Release
126 (2008) 111–121.
[214] K.B. Biggs, K.M. Balss, C.A. Maryanoff, Pore networks and polymer rearrangement
on a drug-eluting stent as revealed by correlated confocal Raman and atomic
force microscopy, Langmuir 28 (2012) 8238–8243.
[215] H. Garvie-Cook, K. Frederiksen, K. Petersson, R.H. Guy, S. Gordeev, Characterization
of Topical Film-Forming Systems Using Atomic Force Microscopy and Raman
Microspectroscopy, Mol. Pharmaceutics 12 (2015) 751–757.
[216] Q. Zhang, K.L. Andrew Chan, G. Zhang, T. Gillece, L. Senak, D.J. Moore, R. Mendelsohn,
C.R. Flach, Raman microspectroscopic and dynamic vapor sorption characterization of
hydration in collagen and dermal tissue, Biopolymers 95 (2011) 607–615.
[217] A. Müllertz, A. Ogbonna, S. Ren, T. Rades, New perspectives on lipid and surfactant
based drug delivery systems for oral delivery of poorly soluble drugs, J. Pharm.
Pharmacol. 62 (2010) 1622–1636.
[218] A. Saupe, K.C. Gordon, T. Rades, Structural investigations on nanoemulsions, solid lipid
nanoparticles and nanostructured lipid carriers by cryo-field emission scanning
electron microscopy and Raman spectroscopy, Int. J. Pharm. 314 (2006) 56–62.
[219] N. Thomas, R. Holm, T. Rades, A. Mullertz, Characterising lipid lipolysis and
its implication in lipid-based formulation development, AAPS J. 14 (2012) 860–871.
[220] N. Thomas, R. Holm, M. Garmer, J.J. Karlsson, A. Mullertz, T. Rades, Supersaturated selfnanoemulsifying drug delivery systems (Super-SNEDDS) enhance the bioavailability
of the poorly water-soluble drug simvastatin in dogs, AAPS J. 15 (2013) 219–227.
[221] C. Stillhart, G. Imanidis, M. Kuentz, Insights into drug precipitation kinetics during
in vitro digestion of a lipid-based drug delivery system using in-line Raman spectroscopy and mathematical modeling, Pharm. Res. 30 (2013) 3114–3130.
[222] M.C. Breitkreitz, G.P. Sabin, G. Polla, R.J. Poppi, Characterization of semi-solid SelfEmulsifying Drug Delivery Systems (SEDDS) of atorvastatin calcium by Raman
image spectroscopy and chemometrics, J. Pharm. Biomed. Anal. 73 (2013) 3–12.
[223] J.J. Schaefer, C. Ma, J.M. Harris, Confocal Raman microscopy probing of temperaturecontrolled release from individual, optically-trapped phospholipid vesicles, Anal.
Chem. 84 (2012) 9505–9512.
[224] P. Li, S.R. Hynes, T.F. Haefele, M. Pudipeddi, A.E. Royce, A.T.M. Serajuddin, Development of clinical dosage forms for a poorly water-soluble drug II: formulation and
characterization of a novel solid microemulsion preconcentrate system for oral
delivery of a poorly water-soluble drug, J. Pharm. Sci. 98 (2009) 1750–1764.
[225] C. Stillhart, G. Imanidis, B. Griffin, M. Kuentz, Biopharmaceutical modeling of drug
supersaturation during lipid-based formulation digestion considering an absorption
sink, Pharm. Res. 31 (2014) 1–19.
[226] H. Wang, M.A. Boraey, L. Williams, D. Lechuga-Ballesteros, R. Vehring, Lowfrequency shift dispersive Raman spectroscopy for the analysis of respirable dosage
forms, Int. J. Pharm. 469 (2014) 197–205.
[227] A. Saigal, W.K. Ng, R.B.H. Tan, S.Y. Chan, Development of controlled release
inhalable polymeric microspheres for treatment of pulmonary hypertension, Int.
J. Pharm. 450 (2013) 114–122.
[228] A. Theophilus, A. Moore, D. Prime, S. Rossomanno, B. Whitcher, H. Chrystyn, Codeposition of salmeterol and fluticasone propionate by a combination inhaler,
Int. J. Pharm. 313 (2006) 14–22.
[229] W. Doub, W. Adams, J. Spencer, L. Buhse, M. Nelson, P. Treado, Raman chemical
imaging for ingredient-specific particle size characterization of aqueous suspension
nasal spray formulations: a progress report, Pharm. Res. 24 (2007) 934–945.
[230] C. Schoenherr, T. Haefele, K. Paulus, G. Francese, Confocal Raman microscopy to
probe content uniformity of a lipid based powder for inhalation: a quality by design
approach, Eur. J. Pharm. Sci. 38 (2009) 47–54.
[231] A.L. Fussell, F. Grasmeijer, H.W. Frijlink, A.H. de Boer, H.L. Offerhaus, CARS microscopy
as a tool for studying the distribution of micronised drugs in adhesive mixtures for
inhalation, J. Raman Spectrosc. 45 (2014) 495–500.
[232] U. Roessl, S. Leitgeb, S. Pieters, T. De Beer, B. Nidetzky, In situ protein
secondary structure determination in ice: Raman spectroscopy-based process
analytical tool for frozen storage of biopharmaceuticals, J. Pharm. Sci. 103
(2014) 2287–2295.
[233] H. Grohganz, D. Gildemyn, E. Skibsted, J.M. Flink, J. Rantanen, Rapid solid-state
analysis of freeze-dried protein formulations using NIR and Raman spectroscopies,
J. Pharm. Sci. 100 (2011) 2871–2875.
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
[234] Z.-Q. Wen, Raman spectroscopy of protein pharmaceuticals, J. Pharm. Sci. 96
(2007) 2861–2878.
[235] F.J. Ramirez, B. Nieto-Ortega, J. Casado, J.T. Lopez Navarrete, The first chiral Raman
spectrum report of a protein: a perspective of 20 years, Chem. Commun. 49 (2013)
8893–8895.
[236] W.P. Kelley, S. Chen, P.D. Floyd, P. Hu, S.G. Kapsi, A.S. Kord, M. Sun, F.G. Vogt,
Analytical characterization of an orally-delivered peptide pharmaceutical product,
Anal. Chem. 84 (2012) 4357–4372.
[237] J. Rantanen, Process analytical applications of Raman spectroscopy, J. Pharm.
Pharmacol. 59 (2007) 171–177.
[238] Y.T. Li, L.L. Qu, D.W. Li, Q.X. Song, F. Fathi, Y.T. Long, Rapid and sensitive in-situ
detection of polar antibiotics in water using a disposable Ag-graphene sensor based
on electrophoretic preconcentration and surface-enhanced Raman spectroscopy,
Biosens. Bioelectron. 43 (2013) 94–100.
[239] S.J. Clarke, R.E. Littleford, W.E. Smith, R. Goodacre, Rapid monitoring of antibiotics
using Raman and surface enhanced Raman spectroscopy, Analyst 130 (2005)
1019–1026.
[240] N.R. Abu-Absi, B.M. Kenty, M.E. Cuellar, M.C. Borys, S. Sakhamuri, D.J. Strachan, M.C.
Hausladen, Z.J. Li, Real time monitoring of multiple parameters in mammalian cell
culture bioreactors using an in-line Raman spectroscopy probe, Biotechnol. Bioeng.
108 (2011) 1215–1221.
[241] B. Li, B.H. Ray, K.J. Leister, A.G. Ryder, Performance monitoring of a mammalian cell
based bioprocess using Raman spectroscopy, Anal. Chim. Acta 796 (2013) 84–91.
[242] A. Calvet, A.G. Ryder, Monitoring cell culture media degradation using surface
enhanced Raman scattering (SERS) spectroscopy, Anal. Chim. Acta 840 (2014)
58–67.
[243] Z.K. Nagy, G. Fevotte, H. Kramer, L.L. Simon, Recent advances in the monitoring,
modelling and control of crystallization systems, Chem. Eng. Res. Des. 91 (2013)
1903–1922.
[244] J. Cornel, C. Lindenberg, J. Schöll, M. Mazzotti, Raman spectroscopy, Industrial
Crystallization Process Monitoring and Control, Wiley-VCH Verlag GmbH & Co.
KGaA 2012, pp. 93–103.
[245] H. Pataki, I. Csontos, Z.K. Nagy, B. Vajna, M. Molnar, L. Katona, G. Marosi, Implementation of Raman signal feedback to perform controlled crystallization of carvedilol, Org.
Process Res. Dev. 17 (2012) 493–499.
[246] T. De Beer, A. Burggraeve, M. Fonteyne, L. Saerens, J.P. Remon, C. Vervaet, Near
infrared and Raman spectroscopy for the in-process monitoring of pharmaceutical
production processes, Int. J. Pharm. 417 (2011) 32–47.
[247] T.R.M. De Beer, C. Bodson, B. Dejaegher, B. Walczak, P. Vercruysse, A. Burggraeve, A.
Lemos, L. Delattre, Y.V. Heyden, J.P. Remon, C. Vervaet, W.R.G. Baeyens, Raman
spectroscopy as a process analytical technology (PAT) tool for the in-line monitoring
and understanding of a powder blending process, J. Pharm. Biomed. Anal. 48 (2008)
772–779.
[248] P. Allan, L.J. Bellamy, A. Nordon, D. Littlejohn, J. Andrews, P. Dallin, In situ monitoring of
powder blending by non-invasive Raman spectrometry with wide area illumination, J.
Pharm. Biomed. Anal. 76 (2013) 28–35.
[249] S. Poutiainen, J. Pajander, A. Savolainen, J. Ketolainen, K. Järvinen, Evolution of
granule structure and drug content during fluidized bed granulation by X-ray
microtomography and confocal Raman spectroscopy, J. Pharm. Sci. 100 (2011)
5254–5269.
[250] H. Wikström, P.J. Marsac, L.S. Taylor, In-line monitoring of hydrate formation during
wet granulation using Raman spectroscopy, J. Pharm. Sci. 94 (2005) 209–219.
[251] M. Wirges, A. Funke, P. Serno, K. Knop, P. Kleinebudde, Development and in-line
validation of a Process Analytical Technology to facilitate the scale up of coating
processes, J. Pharm. Biomed. Anal. 78–79 (2013) 57–64.
[252] K. Knop, P. Kleinebudde, PAT-tools for process control in pharmaceutical film coating
applications, Int. J. Pharm. 457 (2013) 527–536.
[253] A. Bogomolov, M. Engler, M. Melichar, A. Wigmore, In-line analysis of a fluid bed
pellet coating process using a combination of near infrared and Raman spectroscopy,
J. Chemom. 24 (2010) 544–557.
[254] M. Wilson, M.A. Williams, D.S. Jones, G.P. Andrews, Hot-melt extrusion technology
and pharmaceutical application, Ther. Deliv. 3 (2012) 787–797.
[255] L. Saerens, L. Dierickx, B. Lenain, C. Vervaet, J.P. Remon, T.D. Beer, Raman spectroscopy
for the in-line polymer–drug quantification and solid state characterization during a
pharmaceutical hot-melt extrusion process, Eur. J. Pharm. Biopharm. 77 (2011)
158–163.
[256] L. Saerens, D. Ghanam, C. Raemdonck, K. Francois, J. Manz, R. Krüger, S. Krüger, C.
Vervaet, J.P. Remon, T. De Beer, In-line solid state prediction during pharmaceutical
hot-melt extrusion in a 12 mm twin screw extruder using Raman spectroscopy,
Eur. J. Pharm. Biopharm. 87 (2014) 606–615.
[257] T. Vigh, G. Drávavölgyi, P.L. Sóti, H. Pataki, T. Igricz, I. Wagner, B. Vajna, J. Madarász, G.
Marosi, Z.K. Nagy, Predicting final product properties of melt extruded solid dispersions from process parameters using Raman spectrometry, J. Pharm. Biomed. Anal.
98 (2014) 166–177.
[258] H. Trnka, J. Rantanen, H. Grohganz, Well-plate freeze-drying: a high throughput
platform for screening of physical properties of freeze-dried formulations,
Pharm. Dev. Technol. (2014) 1–9.
[259] A. Hedoux, L. Paccou, S. Achir, Y. Guinet, Mechanism of protein stabilization by
trehalose during freeze-drying analyzed by in situ micro-raman spectroscopy, J.
Pharm. Sci. 102 (2013) 2484–2494.
[260] A. Hedoux, L. Paccou, S. Achir, Y. Guinet, In situ monitoring of proteins during
lyophilization using micro-Raman spectroscopy: a description of structural changes
induced by dehydration, J. Pharm. Sci. 101 (2012) 2316–2326.
[261] S. Romero-Torres, H. Wikström, E.R. Grant, L.S. Taylor, Monitoring of mannitol
phase behavior during freeze-drying using non-invasive Raman spectroscopy,
PDA J. Pharm. Sci. Technol. 61 (2007) 131–145.
19
[262] T.R.M. De Beer, M. Allesø, F. Goethals, A. Coppens, Y. Vander Heyden, H. Lopez De
Diego, J. Rantanen, F. Verpoort, C. Vervaet, J.P. Remon, W.R.G. Baeyens, Implementation of a process analytical technology system in a freeze-drying process using
Raman spectroscopy for in-line process monitoring, Anal. Chem. 79 (2007)
7992–8003.
[263] H. de Waard, T. De Beer, W.J. Hinrichs, C. Vervaet, J.-P. Remon, H. Frijlink, Controlled
crystallization of the lipophilic drug fenofibrate during freeze-drying: elucidation of
the mechanism by in-line Raman spectroscopy, AAPS J. 12 (2010) 569–575.
[264] L.S. Taylor, F.W. Langkilde, Evaluation of solid-state forms present in tablets by
Raman spectroscopy, J. Pharm. Sci. 89 (2000) 1342–1353.
[265] M. de Veij, P. Vandenabeele, T. De Beer, J.P. Remon, L. Moens, Reference database
of Raman spectra of pharmaceutical excipients, J. Raman Spectrosc. 40 (2009)
297–307.
[266] H. Wikström, S. Romero-Torres, S. Wongweragiat, J.A. Stuart Williams, E.R. Grant, L.S.
Taylor, On-line content uniformity determination of tablets using low-resolution
Raman spectroscopy, Appl. Spectrosc. 60 (2006) 672–681.
[267] D. Cebeci Maltaş, K. Kwok, P. Wang, L.S. Taylor, D. Ben-Amotz, Rapid classification
of pharmaceutical ingredients with Raman spectroscopy using compressive detection strategy with PLS–DA multivariate filters, J. Pharm. Biomed. Anal. 80 (2013)
63–68.
[268] Z.-P. Chen, L.-M. Li, J.-W. Jin, A. Nordon, D. Littlejohn, J. Yang, J. Zhang, R.-Q. Yu,
Quantitative analysis of powder mixtures by Raman spectrometry: the influence
of particle size and its correction, Anal. Chem. 84 (2012) 4088–4094.
[269] J. Zhang, S.-Z. Liu, J. Yang, M. Song, J. Song, H.-L. Du, Z.-P. Chen, Quantitative
spectroscopic analysis of heterogeneous systems: chemometric methods for the
correction of multiplicative light scattering effects, Rev. Anal. Chem. 32 (2013)
113–125.
[270] C. Eliasson, N.A. Macleod, L.C. Jayes, F.C. Clarke, S.V. Hammond, M.R. Smith, P.
Matousek, Non-invasive quantitative assessment of the content of pharmaceutical
capsules using transmission Raman spectroscopy, J. Pharm. Biomed. Anal. 47
(2008) 221–229.
[271] C.M. McGoverin, M.D. Hargreaves, P. Matousek, K.C. Gordon, Pharmaceutical
polymorphs quantified with transmission Raman spectroscopy, J. Raman Spectrosc.
43 (2012) 280–285.
[272] K. Shin, P.K. Duy, S. Park, Y.-A. Woo, H. Chung, Development of a particle-settling
tolerant transmission Raman scheme for analysis of suspension samples, Analyst
139 (2014) 2813–2822.
[273] J.C. Burley, A. Alkhalil, M. Bloomfield, P. Matousek, Transmission Raman spectroscopy
for quality control in model cocrystal tablets, Analyst 137 (2012) 3052–3057.
[274] T. Vigh, T. Horváthová, A. Balogh, P.L. Sóti, G. Drávavölgyi, Z.K. Nagy, G. Marosi,
Polymer-free and polyvinylpirrolidone-based electrospun solid dosage forms for
drug dissolution enhancement, Eur. J. Pharm. Sci. 49 (2013) 595–602.
[275] M.D. Hargreaves, N.A. Macleod, M.R. Smith, D. Andrews, S.V. Hammond, P.
Matousek, Characterisation of transmission Raman spectroscopy for rapid quantitative analysis of intact multi-component pharmaceutical capsules, J. Pharm.
Biomed. Anal. 54 (2011) 463–468.
[276] Y. Lee, J. Kim, S. Lee, Y.-A. Woo, H. Chung, Simple transmission Raman measurements
using a single multivariate model for analysis of pharmaceutical samples contained in
capsules of different colors, Talanta 89 (2012) 109–116.
[277] C. Shende, W. Smith, C. Brouillette, S. Farquharson, Drug Stability Analyzer for Long
Duration Spaceflights, Proc. SPIE, 9112, 2014 91120R-91125R.
[278] C. De Bleye, P.Y. Sacré, E. Dumont, L. Netchacovitch, P.F. Chavez, G. Piel, P.
Lebrun, P. Hubert, E. Ziemons, Development of a quantitative approach using
surface-enhanced Raman chemical imaging: first step for the determination
of an impurity in a pharmaceutical model, J. Pharm. Biomed. Anal. 90 (2014)
111–118.
[279] K. Dégardin, Y. Roggo, P. Margot, Understanding and fighting the medicine counterfeit
market, J. Pharm. Biomed. Anal. 87 (2014) 167–175.
[280] S. Kovacs, S.E. Hawes, S.N. Maley, E. Mosites, L. Wong, A. Stergachis, Technologies
for detecting falsified and substandard drugs in low and middle-income countries,
PLoS One 9 (2014) e90601.
[281] R.L. McCreery, A.J. Horn, J. Spencer, E. Jefferson, Noninvasive identification of materials
inside USP vials with raman spectroscopy and a raman spectral library, J. Pharm. Sci.
87 (1998) 1–8.
[282] C. Ricci, L. Nyadong, F. Yang, F.M. Fernandez, C.D. Brown, P.N. Newton, S.G.
Kazarian, Assessment of hand-held Raman instrumentation for in situ screening
for potentially counterfeit artesunate antimalarial tablets by FT-Raman spectroscopy
and direct ionization mass spectrometry, Anal. Chim. Acta 623 (2008) 178–186.
[283] C. Ricci, C. Eliasson, N. Macleod, P. Newton, P. Matousek, S. Kazarian, Characterization
of genuine and fake artesunate anti-malarial tablets using Fourier transform infrared
imaging and spatially offset Raman spectroscopy through blister packs, Anal. Bioanal.
Chem. 389 (2007) 1525–1532.
[284] M. Bloomfield, D. Andrews, P. Loeffen, C. Tombling, T. York, P. Matousek, Non-invasive
identification of incoming raw pharmaceutical materials using Spatially Offset Raman
Spectroscopy, J. Pharm. Biomed. Anal. 76 (2013) 65–69.
[285] K. Kwok, L.S. Taylor, Analysis of the packaging enclosing a counterfeit pharmaceutical
tablet using Raman microscopy and two-dimensional correlation spectroscopy, Vib.
Spectrosc. 61 (2012) 176–182.
[286] K. Kwok, L.S. Taylor, Raman spectroscopy for the analysis of counterfeit tablets,
Infrared and Raman Spectroscopy in Forensic Science, John Wiley & Sons, Ltd.
2012, pp. 561–572.
[287] K. Kwok, L.S. Taylor, Analysis of counterfeit Cialis® tablets using Raman microscopy
and multivariate curve resolution, J. Pharm. Biomed. Anal. 66 (2012) 126–135.
[288] Q. Gao, Y. Liu, H. Li, H. Chen, Y. Chai, F. Lu, Comparison of several chemometric
methods of libraries and classifiers for the analysis of expired drugs based on
Raman spectra, J. Pharm. Biomed. Anal. 94 (2014) 58–64.
20
A. Paudel et al. / Advanced Drug Delivery Reviews 89 (2015) 3–20
[289] C.J. Strachan, T. Rades, K.C. Gordon, J. Rantanen, Raman spectroscopy for quantitative
analysis of pharmaceutical solids, J. Pharm. Pharmacol. 59 (2007) 179–192.
[290] Y. Hu, H. Wikström, S.R. Byrn, L.S. Taylor, Analysis of the effect of particle size on
polymorphic quantitation by Raman spectroscopy, Appl. Spectrosc. 60 (2006)
977–984.
[291] N. Townshend, A. Nordon, D. Littlejohn, J. Andrews, P. Dallin, Effect of particle properties of powders on the generation and transmission of Raman scattering, Anal. Chem.
84 (2012) 4665–4670.
[292] N. Townshend, A. Nordon, D. Littlejohn, M. Myrick, J. Andrews, P. Dallin, Comparison of
the determination of a low-concentration active ingredient in pharmaceutical tablets
by backscatter and transmission Raman spectrometry, Anal. Chem. 84 (2012)
4671–4676.
[293] N. Chieng, S. Rehder, D. Saville, T. Rades, J. Aaltonen, Quantitative solid-state analysis of
three solid forms of ranitidine hydrochloride in ternary mixtures using Raman
spectroscopy and X-ray powder diffraction, J. Pharm. Biomed. Anal. 49 (2009) 18–25.
[294] M.E. Auer, U.J. Griesser, J. Sawatzki, Qualitative and quantitative study of polymorphic
forms in drug formulations by near infrared FT-Raman spectroscopy, J. Mol. Struct.
661–662 (2003) 307–317.
[295] K. Kachrimanis, D.E. Braun, U.J. Griesser, Quantitative analysis of paracetamol
polymorphs in powder mixtures by FT-Raman spectroscopy and PLS regression,
J. Pharm. Biomed. Anal. 43 (2007) 407–412.
[296] D.E. Braun, S.G. Maas, N. Zencirci, C. Langes, N.A. Urbanetz, U.J. Griesser, Simultaneous
quantitative analysis of ternary mixtures of d-mannitol polymorphs by FT-Raman
spectroscopy and multivariate calibration models, Int. J. Pharm. 385 (2010) 29–36.
[297] S.A. Schönbichler, L.K.H. Bittner, A.K.H. Weiss, U.J. Griesser, J.D. Pallua, C.W. Huck,
Comparison of NIR chemical imaging with conventional NIR, Raman and ATR-IR
spectroscopy for quantification of furosemide crystal polymorphs in ternary powder
mixtures, Eur. J. Pharm. Biopharm. 84 (2013) 616–625.
[298] L.S. Taylor, G. Zografi, The quantitative analysis of crystallinity using FT-Raman
spectroscopy, Pharm. Res. 15 (1998) 755–761.
[299] M. Farias, R. Carneiro, Simultaneous quantification of three polymorphic forms of
carbamazepine in the presence of excipients using Raman spectroscopy, Molecules
19 (2014) 14128–14138.
[300] M.C. Hennigan, A.G. Ryder, Quantitative polymorph contaminant analysis in tablets
using Raman and near infra-red spectroscopies, J. Pharm. Biomed. Anal. 72 (2013)
163–171.
[301] L.B. Lyndgaard, F. van den Berg, A. de Juan, Quantification of paracetamol through
tablet blister packages by Raman spectroscopy and multivariate curve resolutionalternating least squares, Chemom. Intell. Lab. Syst. 125 (2013) 58–66.
[302] J.C. Burley, A. Aina, P. Matousek, C. Brignell, Quantification of pharmaceuticals
via transmission Raman spectroscopy: data sub-selection, Analyst 139 (2014) 74–78.
[303] D.M. Croker, M.C. Hennigan, A. Maher, Y. Hu, A.G. Ryder, B.K. Hodnett, A comparative
study of the use of powder X-ray diffraction, Raman and near infrared spectroscopy
for quantification of binary polymorphic mixtures of piracetam, J. Pharm. Biomed.
Anal. 63 (2012) 80–86.
[304] K. Kogermann, J. Aaltonen, C.J. Strachan, K. Pöllänen, P. Veski, J. Heinämäki, J.
Yliruusi, J. Rantanen, Qualitative in situ analysis of multiple solid-state forms
using spectroscopy and partial least squares discriminant modeling, J. Pharm. Sci.
96 (2007) 1802–1820.
[305] K. Kogermann, J. Aaltonen, C.J. Strachan, K. Pöllänen, J. Heinämäki, J. Yliruusi, J.
Rantanen, Establishing quantitative in-line analysis of multiple solid-state transformations during dehydration, J. Pharm. Sci. 97 (2008) 4983–4999.
[306] J. Aaltonen, P. Heinänen, L. Peltonen, H. Kortejärvi, V.P. Tanninen, L. Christiansen, J.
Hirvonen, J. Yliruusi, J. Rantanen, In situ measurement of solvent-mediated phase
transformations during dissolution testing, J. Pharm. Sci. 95 (2006) 2730–2737.
[307] P. Lehto, J. Aaltonen, M. Tenho, J. Rantanen, J. Hirvonen, V.P. Tanninen, L. Peltonen,
Solvent-mediated solid phase transformations of carbamazepine: effects of simulated
intestinal fluid and fasted state simulated intestinal fluid, J. Pharm. Sci. 98 (2009)
985–996.
[308] K. Greco, T.L. Bergman, R. Bogner, Design and characterization of a laminar flowthrough dissolution apparatus: comparison of hydrodynamic conditions to those
of common dissolution techniques, Pharm. Dev. Technol. 16 (2011) 75–87.
[309] J. Østergaard, J.X. Wu, K. Naelapää, J.P. Boetker, H. Jensen, J. Rantanen, Simultaneous UV
imaging and Raman spectroscopy for the measurement of solvent-mediated phase
transformations during dissolution testing, J. Pharm. Sci. 103 (2014) 1149–1156.
[310] J.P. Boetker, M. Savolainen, V. Koradia, F. Tian, T. Rades, A. Müllertz, C. Cornett, J.
Rantanen, J. Østergaard, Insights into the early dissolution events of amlodipine using
UV imaging and Raman spectroscopy, Mol. Pharmaceutics 8 (2011) 1372–1380.
[311] K. Greco, R. Bogner, Solution-mediated phase transformation: significance during
dissolution and implications for bioavailability, J. Pharm. Sci. 101 (2012) 2996–3018.
[312] F. Tian, H. Qu, M. Louhi-Kultanen, J. Rantanen, Mechanistic insight into the evaporative
crystallization of two polymorphs of nitrofurantoin monohydrate, J. Cryst. Growth 311
(2009) 2580–2589.
[313] K. Greco, R. Bogner, Crystallization of amorphous indomethacin during dissolution:
effect of processing and annealing, Mol. Pharmaceutics 7 (2010) 1406–1418.
[314] M. Savolainen, K. Kogermann, A. Heinz, J. Aaltonen, L. Peltonen, C. Strachan, J. Yliruusi,
Better understanding of dissolution behaviour of amorphous drugs by in situ solidstate analysis using Raman spectroscopy, Eur. J. Pharm. Biopharm. 71 (2009) 71–79.
[315] K. Greco, D.P. McNamara, R. Bogner, Solution-mediated phase transformation of
salts during dissolution: investigation using haloperidol as a model drug, J.
Pharm. Sci. 100 (2011) 2755–2768.
[316] K. Greco, R. Bogner, Solution-mediated phase transformation of haloperidol mesylate
in the presence of sodium lauryl sulfate, AAPS PharmSciTech 12 (2011) 909–916.
[317] F. Wang, J.A. Wachter, F.J. Antosz, K.A. Berglund, An investigation of solventmediated polymorphic transformation of progesterone using in situ Raman spectroscopy, Org. Process Res. Dev. 4 (2000) 391–395.
[318] M. Jurna, M. Windbergs, C.J. Strachan, L. Hartsuiker, C. Otto, P. Kleinebudde, J.L.
Herek, H.L. Offerhaus, Coherent anti-Stokes Raman scattering microscopy to monitor
drug dissolution in different oral pharmaceutical tablets, J. Innov. Opt. Health Sci. 02
(2009) 37–43.
[319] A.L. Fussell, P. Kleinebudde, J. Herek, C.J. Strachan, H.L. Offerhaus, Coherent antiStokes Raman scattering (CARS) microscopy visualizes pharmaceutical tablets
during dissolution, J. Vis. Exp. 89 (2014) 1–6.
[320] A. Fussell, E. Garbacik, H. Offerhaus, P. Kleinebudde, C. Strachan, In situ dissolution
analysis using coherent anti-Stokes Raman scattering (CARS) and hyperspectral
CARS microscopy, Eur. J. Pharm. Biopharm. 85 (2013) 1141–1147.
[321] M. Windbergs, M. Haaser, C.M. McGoverin, K.C. Gordon, P. Kleinebudde, C.J.
Strachan, Investigating the relationship between drug distribution in solid lipid
matrices and dissolution behaviour using Raman spectroscopy and mapping, J.
Pharm. Sci. 99 (2010) 1464–1475.
[322] A.L. Fussell, P.T. Mah, H. Offerhaus, S.-M. Niemi, J. Salonen, H.A. Santos, C. Strachan,
Coherent anti-Stokes Raman scattering microscopy driving the future of loaded
mesoporous silica imaging, Acta Biomater. 10 (2014) 4870–4877.
[323] M. Windbergs, M. Jurna, H.L. Offerhaus, J.L. Herek, P. Kleinebudde, C.J. Strachan,
Chemical imaging of oral solid dosage forms and changes upon dissolution using
coherent anti-Stokes Raman scattering microscopy, Anal. Chem. 81 (2009)
2085–2091.
[324] M. Haaser, M. Windbergs, C.M. McGoverin, P. Kleinebudde, T. Rades, K.C. Gordon,
C.J. Strachan, Analysis of matrix dosage forms during dissolution testing using
Raman microscopy, J. Pharm. Sci. 100 (2011) 4452–4459.
[325] D. Alonzo, G.Z. Zhang, D. Zhou, Y. Gao, L. Taylor, Understanding the behavior of
amorphous pharmaceutical systems during dissolution, Pharm. Res. 27 (2010)
608–618.
[326] K. Kogermann, A. Penkina, K. Predbannikova, K. Jeeger, P. Veski, J. Rantanen, K.
Naelapää, Dissolution testing of amorphous solid dispersions, Int. J. Pharm. 444
(2013) 40–46.
[327] E.F. Molina, C.V. Santilli, S.H. Pulcinelli, S. Blanchandin, F. Baudelet, V. Briois, Multispectroscopic monitoring of cisplatin-derived species delivery from ureasil
polyether hybrid matrix, Phase Transit. 84 (2011) 687–699.
[328] L. Polavarapu, A.L. Porta, S.M. Novikov, M. Coronado-Puchau, L.M. Liz-Marzán, Pen-onpaper approach toward the design of universal surface enhanced Raman scattering
substrates, Small 10 (2014) 3065–3071.
[329] P.C. Ashok, B.B. Praveen, M. Rube, B. Cox, A. Melzer, K. Dholakia, Development of a
Fiber Based Raman Probe Compatible With Interventional Magnetic Resonance
Imaging, 2014. (pp. 89351J-89351J-89356).
[330] K.-Y. Ju, S. Lee, J. Pyo, J. Choo, J.-K. Lee, Bio-inspired development of a dual-mode
nanoprobe for MRI and Raman imaging, Small 11 (2015) 84–89.
[331] M.S. Bergholt, W. Zheng, K. Lin, K.Y. Ho, M. Teh, K.G. Yeoh, J.B. So, Z. Huang, In vivo
diagnosis of esophageal cancer using image-guided Raman endoscopy and biomolecular modeling, Technol. Cancer Res. Treat. 10 (2011) 103–112.
[332] J. Lipson, J. Bernhardt, U. Block, W.R. Freeman, R. Hofmeister, M. Hristakeva, T.
Lenosky, R. McNamara, D. Petrasek, D. Veltkamp, S. Waydo, Requirements for
calibration in noninvasive glucose monitoring by Raman spectroscopy, J. Diabetes
Sci. Technol. 3 (2009) 233–241.
[333] C. Cahyadi, W. Sasithornwetchakun, A. Topark-Ngarm, L. Chan, P. Heng, An evaluation of content uniformity of tablets prepared by quasi-continuous drug coating
using near infrared and Raman spectroscopy, J. Near Infrared Spectrosc. 22
(2014) 239–248.
[334] R. Singh, A. Sahay, F. Muzzio, M. Ierapetritou, R. Ramachandran, A systematic
framework for onsite design and implementation of a control system in a continuous
tablet manufacturing process, Comput. Chem. Eng. 66 (2014) 186–200.
[335] J. Zhang, Y. Ying, B. Pielecha-Safira, E. Bilgili, R. Ramachandran, R. Romañach, R.N.
Davé, Z. Iqbal, Raman spectroscopy for in-line and off-line quantification of poorly
soluble drugs in strip films, Int. J. Pharm. 475 (2014) 428–437.
[336] S. Kruss, A.J. Hilmer, J. Zhang, N.F. Reuel, B. Mu, M.S. Strano, Carbon nanotubes as
optical biomedical sensors, Adv. Drug Deliv. Rev. 65 (2013) 1933–1950.
[337] M.S. Dresselhaus, A. Jorio, M. Hofmann, G. Dresselhaus, R. Saito, Perspectives on carbon
nanotubes and graphene Raman spectroscopy, Nano Lett. 10 (2010) 751–758.
[338] Z. Pilát, J. Ježek, J. Kaňka, P. Zemánek, Raman Tweezers in Microfluidic Systems for
Analysis and Sorting of Living Cells, Proc. SPIE, 8947, 2014 89471M-89479M.
[339] H. Wu, F.G. Haibach, E. Bergles, J. Qian, C. Zhang, W. Yang, Miniaturized handheld
hyperspectral imager, Proc. SPIE, 9101, 2014 91010W-91016W.
[340] A. Amin, P. Bourget, F. Vidal, F. Ader, Routine application of Raman spectroscopy in the
quality control of hospital compounded ganciclovir, Int. J. Pharm. 474 (2014) 193–201.
[341] I. Noda, Generalized two-dimensional correlation method applicable to infrared,
Raman, and other types of spectroscopy, Appl. Spectrosc. 47 (1993) 1329–1336.
[342] H. Shinzawa, K. Hashimoto, H. Sato, W. Kanematsu, I. Noda, Multiple-perturbation
two-dimensional (2D) correlation analysis for spectroscopic imaging data, J. Mol.
Struct. 1069 (2014) 176–182.
[343] Q. Bi, J. Chen, X. Li, J.-j. Shi, X. Wang, J. Zhang, D. Gao, Y. Zhai, Y. Zhao, S. Weng, Y. Xu, I.
Noda, J. Wu, Investigation on the dipole–dipole interactions between tetramethylurea
and acetonitrile by two-dimensional asynchronous spectroscopy, J. Mol. Struct. 1069
(2014) 264–271.
[344] H.M. Kim, H.S. Park, Y. Cho, S.M. Jin, K.T. Lee, Y.M. Jung, Y.D. Suh, Noninvasive
deep Raman detection with 2D correlation analysis, J. Mol. Struct. 1069 (2014)
223–228.
[345] I. Noda, Frontiers of two-dimensional correlation spectroscopy. Part 2. Perturbation
methods, fields of applications, and types of analytical probes, J. Mol. Struct. 1069
(2014) 23–49.
[346] V.A. Shashilov, I.K. Lednev, Two-dimensional correlation Raman spectroscopy for
characterizing protein structure and dynamics, J. Raman Spectrosc. 40 (2009)
1749–1758.