Biophysics For The Life Sciences: Series Editor: Norma Allewell
Biophysics For The Life Sciences: Series Editor: Norma Allewell
Biophysics For The Life Sciences: Series Editor: Norma Allewell
1 Introduction ............................................................................................. 1
Linda O. Narhi
2 High-Throughput Biophysical Approaches to Therapeutic
Protein Development............................................................................... 7
Feng He, Vladimir I. Razinkov, C. Russell Middaugh,
and Gerald W. Becker
3 Techniques for Higher-Order Structure Determination ..................... 33
James Kranz, Fatma AlAzzam, Atul Saluja, Juraj Svitel,
and Wasfi Al-Azzam
4 Biophysical Techniques for Protein Size Distribution Analysis .......... 83
Ziping Wei and Alla Polozova
5 Qualification of Biophysical Methods for the Analysis
of Protein Therapeutics .......................................................................... 99
Yijia Jiang, Cynthia Li, and John Gabrielson
6 Application of Biophysics to the Early Developability
Assessment of Therapeutic Candidates and Its Application
to Enhance Developability Properties ................................................... 127
Hasige Sathish, Nicolas Angell, David Lowe, Ambarish Shah,
and Steven Bishop
7 Application of Biophysics in Formulation, Process,
and Product Characterization: Selected Case Studies ........................ 147
Satish K. Singh, Qin Zou, Min Huang, and Muralidhara Bilikallahalli
8 Biophysical Analysis in Support of Development of Protein
Pharmaceuticals ...................................................................................... 173
Sreedhara Alavattam, Barthelemy Demeule, Jun Liu, Sandeep Yadav,
Mary Cromwell, and Steven J. Shire
v
vi Contents
vii
viii Contributors
Linda O. Narhi
The last few decades have seen the evolution of protein therapeutics, from a few
drugs isolated from natural sources (such as insulin) to numerous engineered mol-
ecules that are designed to target specific diseases; these biotherapeutics comprise
an increasingly larger part of the commercial drug market (Walsh 2010; Pavlou and
Belsey 2005). While more difficult to make than the traditional small molecule
medications, proteins have the advantage of having a sustained half-life under phys-
iological conditions, can be targeted to specific sites, and do not have the same
concerns around metabolites generated during clearance in vivo, resulting in fewer
side effects. Initial protein therapeutics were hormones and growth factors that were
designed, using the new tools of molecular biology, to replace the native proteins
that were missing or nonfunctional in certain disease states. As the field and industry
matured, the focus shifted to designing molecules to act as antagonist or agonist
targeted to specific components (such as receptors or ligands) on a particular cell
type. Targeting to receptors commonly found on cancer cells has resulted in very
effective oncology treatments with fewer of the debilitating side effects of the tradi-
tional chemotherapy treatments. Inflammation is another area that has seen the
generation of effective biotherapeutics targeted to turn off the inflammatory response
of the effected part of the body in autoimmune diseases (for instance the joints in
rheumatoid arthritis). In order for proteins to be used as a drug, they must be stable,
retaining both activity and structure, during isolation and storage for up to 2 years,
often at high concentration. These are conditions that favor protein aggregation and
denaturation.
Proteins are complex macromolecules that are polymers of the 20 amino acids
connected through peptide bonds. This primary sequence is then folded into the
secondary, tertiary, and quaternary structure required for function. The secondary
structure consists of alpha helix, beta sheet, beta turns, or unstructured conforma-
tions which are defined by the configuration of the peptide backbone, while the
tertiary structure is defined as the overall global protein fold or three-dimensional
structure. Because of this higher order structure, a folded protein is stabilized by
interactions between amino acids that are located far away from each other on the
linear amino acid sequence. These can be hydrophobic, van der Waals or charge-
charge interactions, hydrogen bonds, salt bridges, disulfide bonds, etc.
Proteins can be simple single domain globular proteins like the growth factors
and cytokines that were the first biotech drugs or complex multi-domain molecules
such as the antibodies, the active form of which is actually made of several different
polypeptides held together by disulfide bonds between cysteines. Quaternary struc-
ture is the association of more than one polypeptide chain. This includes association
necessary for the formation of a functioning molecule, like the antibodies and
hemoglobin, association resulting in undesirable nonnative species such as hetero-
geneous partially unfolded protein aggregates that can be formed by stressing the
native monomer, and association into the well-defined but nonnative structures
characteristic of amyloids.
The conformation (overall global fold or three-dimensional structure) of a pro-
tein is important for maintaining biological activity and stability during long-term
storage. It can also impact the safety profile and biological consequences of bio-
therapeutics (Dobsen 2004; Wang et al. 2007). Proteins can be unfolded by many
different types of stress, including extremes of pH, high or low ionic strength and
other solution conditions, heat, and mixing of the air-liquid interface in solution. All
of these conditions can be encountered during the manufacturing, storage, and
administration of the protein therapeutic. The amino acid residues comprising the
molecule can also undergo chemical modifications which might or might not affect
other molecular properties.
Proteins have an inherent affinity to self-associate under the appropriate solution
conditions. This reaction in its simplest form is a concentration-dependent thermo-
dynamic equilibrium, but can also proceed as a complex, multistep reaction that
includes some irreversible steps. A very important irreversible reaction is the forma-
tion or shuffling of disulfide bonds between the monomeric subunits which can
occur under appropriate red/ox conditions. These self-associated species, or aggre-
gates, can range in size from dimer to species containing hundreds of monomers.
Protein aggregates themselves are comprised of a heterogeneous population contain-
ing molecules of different sizes, morphologies, and chemical modifications, formed
through different pathways and intermediates, all determined by the protein sequence
and the conditions to which the protein has been exposed (Narhi et al. 2010).
Even subtle changes in conformation can result in changes in the safety and effi-
cacy of the product, including decreased activity, inability to bind to receptors,
changes in the pharmacokinetics and pharmacodynamic profiles, and the potential
for immunogenicity. Biophysical methods are the principle techniques available to
determine if a potential protein product has the appropriate higher order structure
and can maintain the active conformation during manufacturing, storage, and delivery.
1 Introduction 3
Thus, biophysical techniques play a key role during the development of protein
therapeutics.
The development life cycle for pharmaceutical proteins begins with target iden-
tification and demonstration of the biological activity of an engineered protein. The
next steps are identification of lead therapeutic candidates with this desired activity
and selection of recombinant production cell lines. This is followed by process and
formulation development and characterization and finally clinical trials and then, if
successful, commercialization. The launch of a product changes the focus to prod-
uct consistency and lot release assays, exploration of different delivery devices and
therapeutic indications, comparability assessments, and support for product and
process failure investigations.
Attributes that are important to screen during cell line development include the
amount of monomer and aggregate of the target protein present in the cell culture
media prior to purification. This information will contribute to selection of the final
commercial cell line that delivers optimal yield and product quality. For selection of
the product candidate itself, in addition to biological activity, the ideal molecule will
need to be stable to process and storage conditions, including low pH for viral clear-
ance if it is a mammalian cell line-derived molecule or refolding conditions if it is
an Escherichia coli-derived protein in the form of inclusion bodies. The protein
therapeutic also needs to withstand storage in solution at 4–8°C for 2 years, often at
protein concentrations above 100 mg/mL (Mahler et al. 2009). Screening for this
type of stability usually involves stressing the material and then assessing the impact
of the stress on the integrity of the protein, with particular emphasis on protein
aggregation and irreversible unfolding of the native three-dimensional structures.
The requirements for the biophysical assays used to characterize therapeutic pro-
teins vary with stage of development. During the early stages of drug product devel-
opment, there is very little of the therapeutic protein available and little time to
perform the assessments; a relative ranking of the proteins or cell lines is often
sufficient to select the candidate to move forward into development. Therefore,
assays with minimal material requirements and high throughput are especially valu-
able, and categorizing candidates into pass or fail is an acceptable output.
Once the molecule to be developed as a therapeutic has been chosen, the focus
becomes developing the production process and formulation to be used for the com-
mercial product. At this stage material availability is no longer rate limiting, and
more rigorous techniques comparing the higher order structure of the actual mate-
rial obtained during the different processing steps can be used to ensure that the final
product was not irreversibly damaged by the conditions being used for its
manufacture.
During formulation development, the stability of the target protein is assessed in
different buffer compositions, pH, storage conditions, delivery devices, etc. These
studies typically involve the generation of many samples which need to be analyzed
in order to arrive at the optimal formulation conditions. Many of the principles that
apply during candidate selection apply here as well, and some of the same assays
can even be used. Actual long-term stability studies are also performed to demon-
strate that the protein drug in the formulation chosen has the desired shelf-life
4 L.O. Narhi
Chapters 6–10 present case studies organized to follow the product life cycle.
Chapter 6 demonstrates the use of higher order structure analysis during discovery
and candidate selection, Chap. 7 and 8 illustrates the use of biophysical techniques
during process and product development. Chapter 9 focuses on characterization of
protein aggregates, one of the key product attributes that have the potential to effect
product safety. Chapter 10 contains several case studies demonstrating how bio-
physical techniques are used to identify particles and sources of product and process
failure investigations. And finally Chap. 11 presents a regulatory perspective on
higher order structure analysis across the life cycle. It is our hope that this volume
will enhance the readers understanding and appreciation for the important role that
biophysics plays in successful therapeutic protein development.
References
Dobsen CM (2004) Principles of protein folding, misfolding, and aggregation. Semin Cell Dev
Biol 15:3–16
Mahler H-C, Friess W, Grauschopf U, Kiese S (2009) Protein aggregation: pathways, induction
factors and analysis. J Pharm Sci 98:2909–2934
Narhi LO, Jiang Y, Deshpande R, Kang S, Shultz J (2010) Approaches to control protein aggrega-
tion during bulk production. In: Wang W, Roberts CJ (eds) Aggregation of therapeutic proteins.
Wiley, Hoboken
Pavlou AK, Belsey MJ (2005) The therapeutic antibody market to 2008. Eur J Pharm Biopharm
59:389–396
Walsh G (2010) Biopharmaceutical benchmarks. Nat Biotechnol 28:917–924
Wang W, Singh S, Zeng DL, King K, Nema S (2007) Antibody structure, instability and formula-
tion. J Pharm Sci 96:1–26
Chapter 2
High-Throughput Biophysical Approaches
to Therapeutic Protein Development
2.1 Introduction
criteria to select protein candidates (described in Chap. 6). The selection philosophy
can be based on an established relationship between a particular biophysical
property and product quality, or simply follow the general assumption that better
biophysical characteristics can lead to more stable products if other methods cannot
differentiate the drug candidates. Typically, early stage development is conducted at
small scale employing in vivo and in vitro systems because protein availability is
limited. Thus, methods that consume less material but generate useful information
are the most attractive. Once a protein drug candidate is selected, development is
primarily focused on optimizing the parameters that effectively enable the manufac-
turing, packaging, storage, and delivery of the final product. Before these parame-
ters can be finalized for commercial processes, however, numerous analyses need to
be performed to define the space and limitations of physical conditions that best fit
the product. High-throughput biophysical tools often play an important role in these
efforts by providing a faster readout when the protein product is subjected to a range
of experimental conditions.
Two areas during protein therapeutic development frequently employing high-
throughput biophysical analyses are downstream and formulation development.
The goal of downstream development is to find the most suitable purification pro-
cess so that it can be properly scaled to deliver commercial product. Although the
final optimization usually takes place at scales similar to the commercial settings,
small-scale and high-throughput approaches are often utilized to predict protein
behavior. For example, multiwell plate-based chromatographic techniques com-
bined with high-throughput liquid handling instruments can provide a tremendous
amount of information on protein–resin interaction and therefore help guide purifi-
cation design with very affordable material input (Coffman et al. 2008). Although
the ability of a process to perform adequately at a large scale remains to be derived
empirically, small-scale development is critical because it provides opportunities to
test the potential manufacturing steps while varying physical parameters. In fact,
many regulatory authorities, when evaluating new drug product applications, require
such small-scale data that can demonstrate the robustness of the processes as well
as the edges of failure. Besides purification, the formulation development of protein
therapeutics also requires a large number of experimental trials to select suitable
conditions for final product presentation. The general aim is to find a formulation
that will best retain the physical, chemical, and biological properties of the product
while meeting the desired shelf life requirements. More recently, patient conve-
nience and comfort have become major factors that influence formulation develop-
ment. For instance, product formulations compatible with self-administration and
above-freezing storage are preferred when treating chronic diseases. To derive suit-
able formulations, large arrays of factors are typically screened during the develop-
ment phase for their ability to stabilize the drug product. Since it is well established
that physical instability may negatively impact protein therapeutics, biophysical
techniques offer relevant tools to monitor protein conformation in response to for-
mulation and storage conditions. An important area of focus during formulation
development is the evaluation of protein aggregation propensity. Protein aggregates
are thought to often impose detrimental effects on the therapeutic potency and side
2 High-Throughput Biophysical Approaches to Therapeutic Protein Development 9
effect profile of the drug which may even lead to significant clinical safety concerns
(Jiskoot et al. 2012; Rosenberg 2006). As a result, protein aggregation is fre-
quently used to differentiate formulation candidates. Over the last decade, it has
been well documented in the literature that high-throughput biophysical methods
are very capable of detecting the presence of protein aggregates over a wide size
range (Mach and Arvinte 2011). The feasibility and performance of a protein ther-
apeutic product can also be dependent on other key properties of the molecule. An
emerging example is protein viscosity. To reach the desired bioavailability, protein
drugs delivered via the subcutaneous route typically require high concentrations.
This often leads to high solution viscosity that may cause significant difficulties
during product manufacturing and administration of the protein (Shire et al. 2004).
Traditional analytical methods have limitations in material consumption and
throughput (Jezek et al. 2011), but newly developed biophysical techniques offer
significant advantages, including a reduction of sample volume as well as increases
in throughput capability (He et al. 2010a; Wagner et al. 2012).
In the following sections, the role of high-throughput biophysical analysis in the
development of protein therapeutics is discussed. Technical background for selected
biophysical methods is reviewed as well as their high-throughput utility. In addition,
a general introduction to empirical phase diagram (EPD) is presented as an example
of the application of high-throughput biophysical characterization and data inter-
pretation in the development of protein therapeutics.
Liquid chromatography (LC) is perhaps the most heavily used tool in the field of
biotechnology. LC methods are the main components for the purification and analy-
sis of protein therapeutics during their development cycle (Ahrer and Jungbauer
2006; Andrew and Titus 2001). Understandably, most LC technologies have adopted
the high-throughput scheme by automating and streamlining the sampling
mechanism. The principle of protein LC is based on the nature of protein–resin
interactions in a given liquid mobile phase. An initial major step towards high-
throughput LC technology is the implementation of high-performance liquid chro-
matography (HPLC) (Swadesh 2001). This technique uses a fast flow rate under
high enclosed pressure, instead of gravity force, to drive the mobile phase through
2 High-Throughput Biophysical Approaches to Therapeutic Protein Development 11
the column. This permits faster sampling throughput and improves the resolution of
sample partition. A recent advancement has been categorized as ultra-performance
liquid chromatography (UPLC) where smaller diameter particles and subsequently
higher pressures enable even faster separation times (Xiang et al. 2006). UPLC
technology has clearly shown its advantage in reversed phase (RP) chromatography
and has become the primary separation method for mass spectrometry analysis
(Stackhouse et al. 2011; Szapacs et al. 2010).
Another application of high-throughput LC technology is the development of
protein purification methods. The goal is to obtain the largest amount of the thera-
peutic protein in its active form while reaching the highest purity possible. The
range of such methods is currently extremely wide, and the specific purification
steps often need to be derived empirically. Primary methods include affinity, size-
exclusion, ion-exchange, and hydrophobic interaction chromatographies. Multiwell
plates or miniaturized columns are frequently used to screen a large number of
parameters including type of resin, protein loading and elution conditions, as well
as efficiency for the product of interest (Fahrner et al. 2001). The quantity of column
resin and protein material needed is often measured in microliters, and the experi-
mental steps are simply executed by gravity, centrifugal force, or a pump. This type
of high-throughput approach offers an opportunity to evaluate as many variables as
possible early in development with a minimum of time and protein. The outcome of
the screening can often help to select product-specific processes for further develop-
ment and may even allow companies to bypass intellectual property restricted com-
mon practices. The latest improvement in this field highlights the use of automated
liquid handling systems coupled to positive displacement liquid transfer technology
(Susanto et al. 2008; Wiendahl et al. 2008). Such a system has greater similarity to
large-scale chromatographic instruments used in manufacturing by generating a
pressurized liquid flow through the microcolumns. The results obtained using such
high-throughput techniques are believed to be more representative of a protein’s
behavior during large batch purification.
The path length of the vertical light absorbance is poorly controlled and highly
dependent on the amount of sample present in the well. In addition, proteins with
high extinction coefficients, such as monoclonal antibodies, can easily saturate the
light detector. New adjustable path length spectrophotometers have recently
become available, although their throughput is not as high and light scattering fre-
quently complicates the use of this method.
Compared to UV absorbance, high-throughput optical density (OD), or turbid-
ity, assessment is more widely applied to the protein therapeutic development.
Turbidity usually refers to the obscuration of a sample at wavelengths near the
visible light range where proteins in solution do not manifest significant absor-
bance. The most widely employed wavelength range for this purpose is 350–
400 nm, which avoids any specific absorbance arising from amino acid side chains
or common color pigments. Turbidity is proportional to the amount of light
blocked or scattered by the solution components. In a protein sample, aggregates
and precipitates are known to give rise to solution turbidity as measured by OD,
making it a quick method to assess the quality of protein samples with respect to
the presence of protein aggregates. High-throughput turbidity assessment is fre-
quently used during protein formulation development to evaluate a large number
of samples that are put on storage or under environmental stresses (Zhao et al.
2010; Capelle et al. 2009). Though quantitative determination of protein degrada-
tion is not usually possible with turbidity measurements, the information obtained
is sufficient to discriminate or rank order formulations. Turbidity assessment is
generally noninvasive and can be applied using a variety of spectroscopic instru-
mentation and sample cells, including microtiter plate readers and pharmaceuti-
cally relevant containers. The latter provide a unique opportunity to assess sample
quality of protein therapeutics in their actual storage units, and permit a real-time
monitoring of aggregation during the manufacturing and distribution of a protein
commercial product.
Another common tool for protein analysis is circular dichroism (CD). CD is a
technique that measures the difference in absorbance of left- and right-handed cir-
cularly polarized light. CD is generally divided into near-UV (250–350 nm) and
far-UV (190–250 nm) measurements (Li et al. 2011). Near-UV CD is sensitive to
the tertiary structure of proteins, due to the presence of optically active chromo-
phores including the aromatic amino acid side chains and disulfide bonds. Far-UV
CD, on the other hand, is used to study the secondary structure of proteins. Alpha
helix, beta sheet, turns, and disordered structure all display unique CD spectra.
Expanding the high-throughput capabilities of CD instruments is achieved either by
increasing the number of cuvettes that an instrument can employ or the use of autos-
amplers. In addition, it has recently become possible to obtain data from the near-
and far-UV regions in a single scan. CD is typically more time-consuming than
other light absorbance measurements and can be significantly affected by both light
scattering and absorption flattening phenomena. In order to minimize the interfer-
ence, high-concentration protein samples are often measured via the use of short
path length (µm range) cells (Harn et al. 2007).
2 High-Throughput Biophysical Approaches to Therapeutic Protein Development 13
The aromatic side chains in proteins serve as fluorophores which emit photons at
higher wavelengths once excited with UV range light. The excitation and emission
profile is highly dependent on the polarity of the side chain environment and, therefore,
is sensitive to the local tertiary structure of the protein (Lakowicz 2006). This fluores-
cent property is referred to as protein intrinsic fluorescence. Since most large proteins
have widely distributed aromatic amino acids, the overall intrinsic fluorescence pro-
vides a good measure of protein folding. Upon conformational changes, the intrinsic
fluorescence emission peak generally shifts in wavelength. If present, tryptophan (Trp)
emission dominates the fluorescence of proteins with tyrosine (Tyr) contributing indi-
rectly. This method is often applied to protein formulation studies to detect changes
in protein conformation as a result of stresses such as temperature, pH, and solute.
14 F. He et al.
Fig. 2.1 Representative differential scanning fluorometry spectra obtained with a therapeutic
protein on 96-well plate filled with different formulations for each well. Fluorescence intensity
in relative fluorescence units (RFU) is shown as a function of temperature. Protein concentration
is 1 g/L. The sample volume is 30 µL. Data were obtained using the Bio-Rad CFX96 RT-PCR
plate reader
Many companies that manufacture these instruments have realized the popularity of
this application and have included software options to obtain and analyze DSF data.
Another advantage of DSF is the wide range of protein concentration that can be
used. Depending on protein properties, transitions can be detected at as low as
0.05 mg/mL or as high as at 100 mg/mL. Another high-throughput method, the
ProteoStat™ assay from Enzo Life Sciences (Farmingdale, NY), provides an
improved thermal shift approach based on extrinsic fluorescence for assessment of
protein stability through monitoring protein aggregation, rather than protein
unfolding. These methods, which employ high-throughput technologies, further
expand the application of thermal analysis to modern pharmaceutics. While it is not
guaranteed that thermal stability correlates with a protein’s physical stability during
storage, better thermal stability usually indicates a greater energy requirement to
unfold the protein. When executed under similar experimental conditions, high-
throughput thermal analysis offers useful information that can be used to rank pro-
tein or formulation candidates. It seems safe to state that when all other properties
are comparable, the protein constructs or formulations which lead to better thermal
stability will always be more desirable.
18 F. He et al.
Fig. 2.2 High-throughput method for viscosity measurements based on dynamic light scattering
determination of the diffusion coefficient of polystyrene beads externally added into a protein
solution
ionic strength of the solution. Use of a DLS plate reader was a significant part of the
high-throughput analytical development (Sule et al. 2011).
In biopharmaceutical drug development DLS has been used not only for direct
detection and characterization of aggregation but also for the study of large colloidal
structures. Certain large colloid-like aggregates have been shown to inhibit enzymes
leading to false-positive HTS leads. These so-called promiscuous inhibitors were
detected and screened by DLS using a plate reader (Feng et al. 2005). Results from
such high-throughput assays for promiscuous inhibitory aggregates have been used
to develop new computational models of this phenomenon. A method for quantita-
tive characterization of macromolecular interactions using DLS has been introduced
in a temperature-controlled plate reader format (Hanlon et al. 2010). This technique
enabled determination of equilibrium dissociation constants and thermodynamic
parameters. The low volume of plate-based DLS reduced the sample amount to a
few microliters per experiment, with detection limits in the femtomolar range.
Biopharmaceutical products are often formulated at high concentrations to
maximize delivery dosage and efficiency, and solutions of some proteins become
very viscous at high concentrations (Yadav et al. 2010), creating significant prob-
lems for processes like purification, filtration, and injection through syringes.
Standard methods for viscosity measurements have low throughput and require
large quantities of protein. Thus, there is increasing demand for higher-throughput
viscosity screening. A DLS assay based on measurement of the diffusion coeffi-
cient of beads added directly to the protein solution is high throughput and run in
a multiwell plate format (He et al. 2010a). As shown in Fig. 2.2 the Stokes–Einstein
20 F. He et al.
equation can be used to calculate the viscosity of a protein solution using the
known radius of the added beads and the measured diffusion coefficient.
Furthermore, DLS measurements of diffusion coefficients as a function of protein
concentration can be used to derive the interaction parameter, kD, which has been
shown to correlate with protein properties such as viscosity and particulation pro-
pensity (Yadav et al. 2010; He et al. 2011). It is widely accepted that the second
virial coefficient, B22, obtained by SLS measurements contains information on pro-
tein–protein interaction (Printz et al. 2012). The kD parameter derived from DLS
measurements offers a simple way to compare samples under similar conditions
(discussed in Chap. 3).
particular factors like pH or ionic strength. This methodology has been applied to
the interpretation of CD and FTIR spectra obtained during the production of
antibodies (Greenfield 2006; Sellick et al. 2010). Another useful mathematical tool
is polynomial-based data fitting. This approach involves fitting an arbitrary poly-
nomial model to a limited data set and then using the same model to predict protein
behavior outside of the tested range. For example, discrete data at pH 5, 5.5, and 6
can be used to generate a polynomial model that best fit the experimental results.
The continuous mathematical model can then be used to predict results at pH 5.8,
and even at pH values outside of 5–6, if the assumption holds. The polynomial
method is especially effective when the source data set is large. Even more predic-
tive information can be obtained while considering multiple variables simultane-
ously (Sall et al. 2007).
selection for data analysis (entire spectra can also be used), averaging of multiple
data acquisitions, normalization, input matrix synthesis, singular value decompo-
sition, and finally color mapping of the most significant data using an RGB color
scheme. A detailed description of the method including the mathematics involved
is presented in Maddux et al. (2011). In general, far-UV circular dichroism (CD)
is used to monitor secondary structure although FTIR and Raman spectroscopies
can also be employed. Tertiary structure is most commonly analyzed with intrin-
sic fluorescence, near-UV CD, or high-resolution derivative absorption spectros-
copy. Dye binding using compounds such as 8-anilino naphthalene sulfonic acid
(ANS) is frequently used to probe the exposure of apolar regions in the protein.
Dissociation and association (including aggregation) are typically probed with
static and/or DLS (although see below). Overall thermal stability is often studied with
differential scanning calorimetry. The most common independent variables (i.e.,
forms of stress that have previously been employed) are temperature and pH although
a wide variety of other variables have been used as described below. Perhaps the
major limitation of the EPD approach has been the time and instrumentation neces-
sary to prepare an EPD. This has recently changed with the advent of equipment
capable of performing multiple different types of measurements simultaneously.
Originally an EPD typically required a fluorometer, CD spectropolarimeter, light
scattering system, and perhaps a DSC or FTIR spectrometer. Several newly
developed instruments now at least partially overcome this limitation. For example,
recent improvements in CD instruments now permit both near- and far-UV
spectra, near-UV absorption spectra, fluorescence, and SLS (turbidity or scatter-
ing at the fluorescence emission wavelength) to all be acquired simultaneously in
a four-position sample chamber under variable temperature conditions (Hu et al.
2011). This permits EPDs to be generated in less than a day. A similar “protein
machine” with a six-position sample chamber has also been recently described
(Maddux et al. 2012). Perhaps the simplest version of a system with rapid EPD
generation capability is a UV absorption spectrometer, typically of the diode array
variety, to permit sufficient resolution of derivative peaks (Kueltzo et al. 2003). At
high resolution (usually second), the derivative spectrum of a protein will usually
manifest distinct peaks for phenylalanine, tyrosine, and tryptophan (if present).
Since the residues are usually buried, Tyr is often interfacial, and Trp present in
highly variable environments, temperature, and pH-induced peaks shifts can fre-
quently provide a fairly detailed picture of a protein’s structural response to vari-
ous perturbations. Such data can easily be represented in the form of an EPD
(Kueltzo et al. 2003). Similarly, fluorescence microtiter plate-based fluorometers
have been developed which employ multiple fluorescence and light scattering
measurements as a function of temperature, as also described in the fluorescence
section. The latter is of especially high throughput, permitting the generation of
many EPDs in a single day. EPDs for four model proteins obtained from multiple
instruments, two CD-based spectrometers, and a high-throughput fluorometer are
shown in Fig. 2.3, where it can be seen that all produce similar EPDs although
small differences are apparent due to the various types of measurements used to
construct each EPD.
2 High-Throughput Biophysical Approaches to Therapeutic Protein Development 23
Fig. 2.3 Empirical phase diagrams (EPDs) of four model proteins (1) aldolase, (2) BSA,
(3) chymotrypsin, and (4) lysozyme constructed using data collected from various instruments:
(a) intrinsic fluorescence (FL) and static light scattering (SLS) data from a Photon Technology
International (PTI) fluorometer, and circular dichroism (CD) from an Applied Photophysics
Chirascan, (b) FL and SLS by an Avacta Optim 1000 microtiter plate fluorometer, (c) FL and CD
Applied Photophysics Chirascan, and (d) FL, SLS, CD, and UV absorbance by an Olis Protein
Machine
buffers, sugars, sugar alcohols, amino acids, polymers, detergents, and osmolytes is
used. In the initial screen relatively high concentrations of compounds are used with
their concentration dependence and use in combination later optimized. It is usually
wise to employ at least two methods: one sensitive to aggregation (e.g., light scat-
tering) and one to structural change (e.g., fluorescence, CD) for this purpose. DSC
is also commonly employed especially due to the recent availability of highly sensi-
tive high-throughput instruments. It is also possible to prepare EPDs in the presence
of selected stabilizers to permit a more detailed analysis/comparison of their effects
on a protein. The information thus obtained by a temperature/pH EPD thus provides
a basis for buffer and excipient selection at an early stage of pharmaceutical devel-
opment. Although not yet published, a new version of the EPD has been developed
in which the colors have actual physical measuring in contrast to the arbitrary
assignment of color in the original EPD.
High-throughput methods and EPDs can also be applied to a wide variety of differ-
ent situations, some of which will be briefly described here. Two commonly encoun-
tered forms of stress in the protein therapeutic area are freeze/thaw and shear. It is
often necessary to freeze and then thaw both during development and manufactur-
ing situations. An EPD can be created using the number of freeze/thaw cycles under
defined conditions as an independent variable accompanied by temperature and pH
stress. All three variables (temperature, pH, freeze/thaw cycles) can be combined
into a three-dimensional representation in which the EPD is presented as a colored
surface. Shear stress is also often encountered in the development, manufacture
(especially filling), and shipping of protein pharmaceuticals. To explore this poten-
tial degrading stress, the intensity of the shear can be varied by a mechanical pro-
cess such as stirring, shaking, or some other forms of agitation, and this is used as a
variable in EPD production.
Protein concentration is another important variable that has assumed increasing
importance with the use of high-concentration formulations. This variable can be
typically evaluated over the range of 0.05–300 g/L depending on the solubility of
the protein and the methods employed in the analysis. Proteins usually alter their
structure to little or no extent as a function of protein concentration, but aggregation
and surface adsorption are both highly concentration dependent. Thus, aggregation-
sensitive techniques such as light scattering are often of special importance in pro-
tein concentration-dependent studies. Another common variable of particular
importance is ionic strength. In the Debye–Huckel charge shielding regime (0–0.15
ionic strength), a number of intermediate concentrations should be evaluated to
probe electrostatic effects. At higher salt concentrations both preferential hydration
and binding effects usually dominate with salt concentration into the molar range
appropriately examined.
2 High-Throughput Biophysical Approaches to Therapeutic Protein Development 25
and sensitive way to see if structural identity has been obtained. As an example,
second-generation functional mutants of fibroblast growth factor one (FGF-1) have
been compared using EPDs and used to select molecules that are not dependent on
heparin for their activity (Alsenaidy et al. 2012). Such EPD-dependent comparisons
have even been performed with different rotavirus serotypes despite their individual
complexity (Esfandiary et al. 2010). Various mathematical (difference) methods
exist to facilitate such comparisons.
This approach as well as others allows a very complex data set to be presented in a
way that is readily understandable.
Although high-throughput biophysical methods become increasingly important
in the development of therapeutic proteins, there are still biophysical methods that
have not been adapted to a high-throughput format. Examples include circular
dichroism and mass spectrometry which are still in their infancy but which, no
doubt, will have high-throughput adaptations in the near future. It is clear that there
is a need for new and better high-throughput biophysical assays and that the needs
of the biopharmaceutical industry will play a role in their development.
References
Goldberg DS, Bishop SM, Shah AU, Sathish HA (2010) Formulation development of therapeutic
monoclonal antibodies using high-throughput fluorescence and static light scattering tech-
niques: role of conformational and colloidal stability. J Pharm Sci 100(4):1306–1315
Greenfield NJ (2006) Using circular dichroism spectra to estimate protein secondary structure. Nat
Protoc 1(6):2876–2890
Haidekker MA, Brady TP, Lichlyter D, Theodorakis EA (2005) Effects of solvent polarity and
solvent viscosity on the fluorescent properties of molecular rotors and related probes. Bioorg
Chem 33(6):415–425
Hanlon AD, Larkin MI, Reddick RM (2010) Free-solution, label-free protein-protein interactions
characterized by dynamic light scattering. Biophys J 98(2):297–304
Harn N, Allan C, Oliver C, Middaugh CR (2007) Highly concentrated monoclonal antibody solu-
tions: direct analysis of physical structure and thermal stability. J Pharm Sci 96(3):532–546
Hawe A, Sutter M, Jiskoot W (2008a) Extrinsic fluorescent dyes as tools for protein characteriza-
tion. Pharm Res 25(7):1487–1499
Hawe A, Friess W, Sutter M, Jiskoot W (2008b) Online fluorescent dye detection method for the
characterization of immunoglobulin G aggregation by size exclusion chromatography and
asymmetrical flow field flow fractionation. Anal Biochem 378(2):115–122
Hawe A, Filipe V, Jiskoot W (2010a) Fluorescent molecular rotors as dyes to characterize
polysorbate-containing IgG formulations. Pharm Res 27(2):314–326
Hawe A, Rispens T, Herron JN, Jiskoot W (2010b) Probing bis-ANS binding sites of different
affinity on aggregated IgG by steady-state fluorescence, time-resolved fluorescence and iso-
thermal titration calorimetry. J Pharm Sci 100(4):1294–1305
He F, Becker GW, Litowski JR, Narhi LO, Brems DN, Razinkov VI (2010a) High-throughput
dynamic light scattering method for measuring viscosity of concentrated protein solutions.
Anal Biochem 399(1):141–143
He F, Hogan S, Latypov RF, Narhi LO, Razinkov VI (2010b) High throughput thermostability
screening of monoclonal antibody formulations. J Pharm Sci 99(4):1707–1720
He F, Phan DH, Hogan S et al (2010c) Detection of IgG aggregation by a high throughput method
based on extrinsic fluorescence. J Pharm Sci 99(6):2598–2608
He F, Woods CE, Becker GW, Narhi LO, Razinkov VI (2011) High-throughput assessment of
thermal and colloidal stability parameters for monoclonal antibody formulations. J Pharm Sci
100(12):5126–5141
Hu L, Olsen C, Maddux N, Joshi SB, Volkin DB, Middaugh CR (2011) Investigation of protein
conformational stability employing a multimodal spectrometer. Anal Chem
83(24):9399–9405
Jezek J, Rides M, Derham B et al (2011) Viscosity of concentrated therapeutic protein composi-
tions. Adv Drug Deliv Rev 63(13):1107–1117
Jiskoot W, Randolph TW, Volkin DB et al (2012) Protein instability and immunogenicity: road-
blocks to clinical application of injectable protein delivery systems for sustained release.
J Pharm Sci 101(3):946–954
Karlsson R (2004) SPR for molecular interaction analysis: a review of emerging application areas.
J Mol Recognit 17(3):151–161
Khossravi M, Kao YH, Mrsny RJ, Sweeney TD (2002) Analysis methods of polysorbate 20: a new
method to assess the stability of polysorbate 20 and established methods that may overlook
degraded polysorbate 20. Pharm Res 19(5):634–639
Kueltzo LA, Ersoy B, Ralston JP, Middaugh CR (2003) Derivative absorbance spectroscopy and
protein phase diagrams as tools for comprehensive protein characterization: a bGCSF case
study. J Pharm Sci 92(9):1805–1820
Kung CE, Reed JK (1989) Fluorescent molecular rotors: a new class of probes for tubulin structure
and assembly. Biochemistry 28(16):6678–6686
Lakowicz JR (2006) Principles of fluorescence spectroscopy, 3rd edn. Springer, New York
Lee W, Fon W, Axelrod BW, Roukes ML (2009) High-sensitivity microfluidic calorimeters for
biological and chemical applications. Proc Natl Acad Sci USA 106(36):15225–15230
30 F. He et al.
Lerchner J, Wolf A, Wolf G et al (2006) A new micro-fluid chip calorimeter for biochemical
applications. Thermochim Acta 445:144–150
Lerchner J, Maskow T, Wolf G (2008) A new micro-fluid chip calorimeter for biochemical appli-
cations. Thermochim Acta 47:991–999
Li CH, Nguyen X, Narhi L et al (2011) Applications of circular dichroism (CD) for structural
analysis of proteins: qualification of near- and far-UV CD for protein higher order structural
analysis. J Pharm Sci 100(11):4642–4654
Mach H, Arvinte T (2011) Addressing new analytical challenges in protein formulation develop-
ment. Eur J Pharm Biopharm 78(2):196–207
Mach H, Middaugh CR (2011) Ultraviolet spectroscopy as a tool in therapeutic protein develop-
ment. J Pharm Sci 100(4):1214–1227
Mach H, Bhambhani A, Meyer BK et al (2011) The use of flow cytometry for the detection of
subvisible particles in therapeutic protein formulations. J Pharm Sci 100(5):1671–1678
Maddux NR, Joshi SB, Volkin DB, Ralston JP, Middaugh CR (2011) Multidimensional methods
for the formulation of biopharmaceuticals and vaccines. J Pharm Sci 100(10):4171–4197
Maddux NR, Rosen IT, Hu L, Olsen CM, Volkin DB, Middaugh CR (2012) An improved method-
ology for multidimensional high-throughput preformulation characterization of protein confor-
mational stability. J Pharm Sci 101(6):2017–2024
Malawski GA, Hillig RC, Monteclaro F et al (2006) Identifying protein construct variants with
increased crystallization propensity – a case study. Protein Sci 15(12):2718–2728
McEvoy M, Razinkov V, Wei Z, Casas-Finet JR, Tous GI, Schenerman MA (2011) Improved par-
ticle counting and size distribution determination of aggregated virus populations by asym-
metric flow field-flow fractionation and multiangle light scattering techniques. Biotechnol Prog
27(2):547–554
Otsuki S, Ishikawa M (2010) Wavelength-scanning surface plasmon resonance imaging for label-
free multiplexed protein microarray assay. Biosens Bioelectron 26(1):202–206
Owicki JC (2000) Fluorescence polarization and anisotropy in high throughput screening: perspec-
tives and primer. J Biomol Screen 5(5):297–306
Pantoliano MW, Petrella EC, Kwasnoski JD et al (2001) High-density miniaturized thermal shift
assays as a general strategy for drug discovery. J Biomol Screen 6:429–440
Peters WB, Frasca V, Brown RK (2009) Recent developments in isothermal titration calorimetry
label free screening. Comb Chem High Throughput Screen 12(8):772–790
Pierce MM, Raman CS, Nall BT (1999) Isothermal titration calorimetry of protein–protein interac-
tions. Methods 19(2):213–221
Printz M, Kalonia DS, Friess W (2012) Individual second virial coefficient determination of mono-
mer and oligomers in heat-stressed protein samples using size-exclusion chromatography-light
scattering. J Pharm Sci 101:363–372
Privalov GP, Privalov PL (2000) Problems and perspectives in microcalorimetry of biological mac-
romolecules. Methods Enzymol 323:31–62
Rathore AS, Devine R (2008) PDA workshop on “Quality by Design for Biopharmaceuticals:
Concepts and Implementation”, May 21–22, 2007, Bethesda, Maryland. PDA J Pharm Sci
Technol 62(5):380–390
Rathore AS, Winkle H (2009) Quality by design for biopharmaceuticals. Nat Biotechnol
27(1):26–34
Rich RL, Myszka DG (2000) Advances in surface plasmon resonance biosensor analysis. Curr
Opin Biotechnol 11(1):54–61
Rosenberg AS (2006) Effects of protein aggregates: an immunologic perspective. AAPS
J 8(3):E501–E507
Sall J, Creighton L, Lehman A (2007) JMP start statistics: a guide to statistics and data analysis
using JMP, 4th edn. SAS Press, Cary
Schäfer G, Schmidt H (2006) High-throughput spectroscopic viscosity measurement of nanocom-
posite sols with ETC-effect. J Sol–Gel Sci Technol 38(3):241–244
Schiffer CA, Dotsch V (1996) The role of protein-solvent interactions in protein unfolding. Curr
Opin Biotechnol 7(4):428–432
2 High-Throughput Biophysical Approaches to Therapeutic Protein Development 31
Schmidt R (2010) Dynamic light scattering for protein characterization. Encyclopedia of Analytical
Chemistry, 20–24
Sellick CA, Hansen R, Jarvis RM et al (2010) Rapid monitoring of recombinant antibody produc-
tion by mammalian cell cultures using Fourier transform infrared spectroscopy and chemomet-
rics. Biotechnol Bioeng 106(3):432–442
Shire SJ, Shahrokh Z, Liu J (2004) Challenges in the development of high protein concentration
formulations. J Pharm Sci 93(6):1390–1402
Siebert F, Hildebrandt P (2008) Vibrational spectroscopy in life science. Wiley-VCH, Weinheim
Stackhouse N, Miller AK, Gadgil HS (2011) A high-throughput UPLC method for the character-
ization of chemical modifications in monoclonal antibody molecules. J Pharm Sci
100(12):5115–5125
Sule SV, Sukumar M, Weiss WF, Marcelino-Cruz AM, Sample T, Tessier PM (2011) High-
throughput analysis of concentration-dependent antibody self-association. Biophys J
101(7):1749–1757
Susanto A, Knieps-Grünhagen E, von Lieres E, Hubbuch J (2008) High throughput screening for
the design and optimization of chromatographic processes: assessment of model parameter
determination from high throughput compatible data. Chem Eng Technol 31(12):1846–1855
Swadesh J (2001) HPLC: practical and industrial applications, 2nd edn. CRC Press, Boca Raton
Szapacs ME, Urbanski JJ, Kehler JR et al (2010) Absolute quantification of a therapeutic domain
antibody using ultra-performance liquid chromatography-mass spectrometry and immunoas-
say. Bioanalysis 2(9):1597–1608
Takeda H, Fukumoto A, Miura A, Goshima N, Nomura N (2006) High-throughput kinase assay
based on surface plasmon resonance suitable for native protein substrates. Anal Biochem
357(2):262–271
Tarazona MP, Saiz E (2003) Combination of SEC/MALS experimental procedures and theoretical
analysis for studying the solution properties of macromolecules. J Biochem Biophys Methods
56(1–3):95–116
Torres FE, Recht MI, Coyle JE, Bruce RH, Williams G (2010) Higher throughput calorimetry:
opportunities, approaches and challenges. Curr Opin Struct Biol 20(5):598–605
Verhaegen K, Baert K, Simaels J, Van Driessche W (2000) A high-throughput silicon microphysi-
ometer. Sens Actuators A Phys 82(1):186–190
Vermeir S, Nicolai BM, Verboven P et al (2007) Microplate differential calorimetric biosensor for
ascorbic acid analysis in food and pharmaceuticals. Anal Chem 79(16):6119–6127
Vincentelli R, Canaan S, Campanacci V et al (2004) High-throughput automated refolding screen-
ing of inclusion bodies. Protein Sci 13(10):2782–2792
Wagner M, Reiche K, Blume A, Garidel P (2012) Viscosity measurements of antibody solutions
by photon correlation spectroscopy: an indirect approach – limitations and applicability for
high-concentration liquid protein solutions. Pharm Dev Technol. doi:10.3109/10837450.2011.
649851
Wang L, Wang B, Lin Q (2008) Demonstration of MEMS-based differential scanning calorimetry
for determining thermodynamic properties of biomolecules. Sens Actuators B Chem
134:953–958
Wassaf D, Kuang G, Kopacz K et al (2006) High-throughput affinity ranking of antibodies using
surface plasmon resonance microarrays. Anal Biochem 351(2):241–253
Wiendahl M, Schulze Wierling P, Nielsen J et al (2008) High throughput screening for the design
and optimization of chromatographic processes – miniaturization, automation and paralleliza-
tion of breakthrough and elution studies. Chem Eng Technol 31(6):893–903
Xiang Y, Liu Y, Lee ML (2006) Ultrahigh pressure liquid chromatography using elevated tempera-
ture. J Chromatogr A 1104(1–2):198–202
Yadav S, Liu J, Shire SJ, Kalonia DS (2010) Specific interactions in high concentration antibody
solutions resulting in high viscosity. J Pharm Sci 99(3):1152–1168
Yau N (2011) Visualize this: the FlowingData guide to design, visualization, and statistics. Wiley,
Indianapolis
Zhao H, Graf O, Milovic N et al (2010) Formulation development of antibodies using robotic
system and high-throughput laboratory (HTL). J Pharm Sci 99(5):2279–2294
Chapter 3
Techniques for Higher-Order Structure
Determination
Proteins are the building blocks for major parts of living systems that are r esponsible
for all critical cellular functions. Their roles vary from catalyzing reactions
(enzymes) to facilitating movement (cytoskeletal and motor proteins) and acting as
messengers for signal transduction, to name a few. The ability of proteins to carry
out such broad structural and functional roles is due to their unique three-dimensional
structures, as well as thermodynamic characteristics that allow the functional form
of a protein to retain stability under many different environmental conditions, such
as changing pH or temperature. Understanding protein structure is a prerequisite for
the development of protein therapeutics, from the identification of potential targets
for therapeutic agents, being developed and administered as the therapeutic agent or
being used in more novel applications as carriers of therapeutic agents and sustained
released devices, etc. Control, optimization, and quantitation of protein structure are
critical to the success of protein pharmaceuticals.
Protein structure starts with the linear polymer composed of 20 different amino
acids covalently linked through peptide bonds; this is the primary structure or
sequence. The peptide bonds are also responsible for the formation of intraresidue
hydrogen bonding that gives rise to secondary structures. Each of the 20 different
amino acids has a specific side chain which determines the chemical and functional
properties of the residue; they are classified as nonpolar, polar, acidic, basic, or
aromatic.
The primary sequence can adopt stable configurations of local short-range sec-
ondary structures mediated by intraresidue peptide bond or “backbone” interac-
tions. Common examples of secondary structures are α-helices, β-sheets, turns, and
unstructured (random coil) segments. The structure of an α-helix is stabilized by
H-bonding between the nitrogen of the amide and the carbonyl carbons of peptide
bonds spaced four residues apart. β-sheets are composed of two or more different
linear sequences of at least 5–10 amino acids, stabilized by H-bonding between the
nitrogen of the amide of one amino acid and the carbonyl carbons of the amino acid
in the adjacent strand. β-sheets can contain strands which are far away from each
other in the linear amino acid sequence, but adjacent in the folded structure of the
protein. Parallel β-pleated sheets consist of strands that have the same (parallel)
orientation on the linear sequence or primary structure of the protein, while in the
more common antiparallel β-sheet, the amino acids at the strand–strand interface
are coming from opposite directions. Other secondary structures such as bends and
turns may exist in specific structures or conformations.
The final functional three-dimensional, or tertiary, structure of a protein is sta-
bilized by a large number of noncovalent interactions that are mediated almost
exclusively by side-chain interactions between amino acids that can be from very
different regions of the primary structure, but are brought into close proximity as
the protein folds. Tertiary structures result from further bonding between side
chains within the protein and with any water that may be present around the pro-
tein. The “hydrophobic effect,” which is a thermodynamic description of the driv-
ing forces stabilizing the tertiary or native structure of a protein, is due in large part
to the side-chain-specific interaction of H2O with a protein (Sharp et al. 1991;
Murphy and Freire 1992). The hydrophobic effect tends to facilitate the partition-
ing of polar amino acids on the outside of the folded protein, with nonpolar amino
acids clustering in the inside of the protein in its final stabilized structure. Examples
of tertiary contacts include hydrophobic (van der Waals) interactions, dipole- and
induced-dipole interactions, disulfide bonds, electrostatic and ionic interactions,
and hydrogen bonds.
Multidomain proteins are common, where different conserved functional
domains are present within a primary sequence and fold into separate domains often
connected only by the single strand of amino acids of the primary structure. These
domains can show cooperative behavior, with one domain influencing the folding of
other domains, or the folding of individual domains may be thermodynamically
independent of the rest of the protein. For example, different families of kinases
exist with a somewhat conserved catalytic domain but differ in terms of the auxiliary
3 Techniques for Higher-Order Structure Determination 35
domains that regulate kinase folding, function, and specificity. Likewise, many
proteins contain pro-domains that function only to facilitate proper folding, then
these are excised via proteolysis to produce the final folded active structure.
Quaternary structure refers to the association of folded domains from different
amino acid chains. It is in some ways an extension of tertiary interactions, arising
from ordered interactions between two or more polypeptides subunits. The bonds
in each monomer are the same as those found in the tertiary structure. The poly-
peptide subunits are attached either by peptide strand or noncovalent interactions.
Common quaternary structures include dimers (transcription factors), trimers
(some cytokines), and higher-ordered oligomers. Allostery is one additional
dynamic property that is common among multimeric proteins, where different
functional properties are observed for unique quaternary structures that are only
available to an assembly of monomers. One well-characterized example of allo-
stery is hemoglobin that gives rise to cooperative O2 binding as well as its differen-
tial affinity for O2 in the lungs and tissues based on the different conditions
(e.g., pH) of the surrounding tissues.
Structural analysis of protein therapeutics has been an important focus for the
biopharmaceutical industry, as described in the first chapter of this book. The goal
of successful protein therapeutic development is the creation of a therapeutic pro-
tein with native-like structure, good bioactivity, and stability that elicits a minimal
immune response. Moreover, protein conformational stability can correlate phe-
nomenologically with in vivo stability. Protein aggregation and chemical modifica-
tion are the two most common routes for protein degradation and loss, with
aggregation depending more on protein structure and stability (Putnam et al. 2010).
Aggregates are heterogeneous mixtures of native folded species, nonnative unfolded,
or partially unfolded species, self-associated into larger molecules or polymers.
Multi-monomeric aggregates may continue growing in regular shapes such as
fibrils, or as amorphous particles, and can range in size from oligomers ( nanometers)
to subvisible (microns) or visible (hundreds of microns) particulates (Stefani and
Dobson 2003).
Manufacturing processes are continually optimized during therapeutic devel-
opment to improve product yield and product quality, while changes in formula-
tion and delivery methods can also occur in response to information on the target
patient population. All of these changes could alter the physicochemical proper-
ties of the protein in use, including chemical modification, self-association, and
ultimately the tertiary structure and protein folding. Thus, there is a growing need
for the development of biophysical tools that are sensitive to subtle changes in
protein higher-order structure (HOS) during biopharmaceutical development.
This chapter provides a brief review of technologies that are commonly used and
available to characterize various higher-order structural properties, with an
intended application toward protein biotherapeutics. Applications of these tech-
niques throughout the product lifecycle are provided in many subsequent chapters
in this volume.
36 J. Kranz et al.
I
A = log 10 = ecl, (3.1)
I0
where A is the measured absorbance, I0 is the original intensity of the light and I is
the intensity of the light transmitted through the sample, l is the length of the sample
path length the light was shown through, c is the concentration of the chromo-
phores, and ε is the absorbance of a one molar solution of the chromophores, often
38 J. Kranz et al.
3.2.2 Fluorescence
properties of tryptophan residues (Bartlett and Radford 2009; Royer 1995; Royer
and Scarlata 2008). This is most clearly observed in systems with a single trypto-
phan (Epstein et al. 1971; Flanagan et al. 1992; Hynes and Fox 1991; Otto et al.
1994), but also in studies of proteins containing multiple tryptophan residues
(LeTilly and Royer 1993; Mann et al. 1993; Royer 1993). While the emission spec-
trum maximum of a buried tryptophan generally exhibits a red shift (lower energy,
higher frequency) upon protein unfolding, the peptide backbone and a number of
the amino acid side chains can alter tryptophan fluorescence spectra, generally
through excited-state quenching or electron transfer (Adams et al. 2002).
(I − I ⊥ )
A= . (3.2)
(I + 2 I ⊥ )
Fluorophores that are physically oriented such that their electronic dipole
moment is aligned in the plane of incident light are selectively excited. After rotational
40 J. Kranz et al.
A0 t
−1 = , (3.3)
A tc
where A0 is the limiting anisotropy, τ is the fluorescence lifetime, and τc is the rota-
tional correlation time of the macromolecule. For globular (approximately spheri-
cal) proteins, the rotational correlation time is a function of the apparent molecular
weight, M, of a protein by
hV hM
tc = = (n + h), (3.4)
RT RT
where η is the viscosity, n- is the specific volume of the protein, h is the relative
hydration, T is temperature (K), and R is the universal gas constant. For a typical
- ~ 0.73 mL/g and hydration levels are h ~ 0.23 g of H O per gram of pro-
protein, n 2
tein. This expression predicts that the correlation time of a hydrated protein is on
average ~30 % larger than would be predicted for a fully anhydrous protein (spheri-
cal approximation).
The apparent volume of a protein can be measured experimentally, solely from
measurements of protein anisotropy as a function of varying viscosity; changes in
viscosity are typically achieved by variable temperature combined with viscous
cosolvents, like glycerol. The Perrin equation is substituted with terms from Eq. 3.4,
giving
1 1 tRT
= + . (3.5)
A A0 t chV
The use of Eq. 3.4 to measure the apparent volume of a protein is one of the earli-
est applications of fluorescence in a biochemical system (Weber 1952a, b). As dis-
cussed in subsequent sections, the apparent volume of a protein can vary significantly
in response to changes in formulations, aggregation onset, and protein unfolding.
3 Techniques for Higher-Order Structure Determination 41
Quenching refers generically to any molecular process that decreases the intensity
of fluorescence. These include molecular rearrangements, energy transfer, excited-
state reactions, and quenching through collisions. Fluorescence quenching has been
widely studied as a source of information on biological systems. During collisional
quenching, the quenching agent must diffuse to the fluorophore during the lifetime
of the excited state. Upon contact, there is energy transfer via interactions of the
excited-state dipole and the dipole (or induced dipole) of the quencher, facilitating
return of the fluorophore to the ground state without photon emission.
The requirement of intermolecular contact between the quencher and fluoro-
phore has obvious applications to biochemical systems. Quenching measurements
can be indicative of the relative accessibility of fluorophores to quenchers. For exam-
ple, in systems with a few fluorophore groups, either buried in the protein interior of
soluble proteins or in lipid environments of transmembrane proteins, quenching
rates can be used to determine the permeability of proteins or membranes to the
extrinsic quencher. Likewise, quenching rates and lifetimes of fluorophore and
quencher groups can be useful in measurements of molecular diffusion. Details on
collisional quenching theory and applications are prevalent in the literature (Cantor
and Schimmel 1980; Lakowicz 2006).
calculated (dos Remedios and Moens 1995). This technique can also be used to
follow changes in the distance between parts of the same polypeptide with changes
in solution condition, interactions with other proteins, etc.
Most applications of FRET utilize extrinsic fluorophores, ones that involve
cofactors as in the process of photosynthesis or more commonly are covalently
attached to a protein via chemical means. These have found widespread applica-
tions in labeling of cellular proteins used in studies of cellular trafficking or in vivo
protein–protein interactions (Pollok and Heim 1999; Miyawaki 2011; Jares-Erijman
and Jovin 2003) and in studies of single molecules (Ferreon et al. 2011; Kubelka
et al. 2004). Within proteins there is intrinsic fluorescence energy transfer between
the tyrosine, the donor amino acid, and tryptophan, the acceptor molecule. These
effects are perhaps most observable in simplified, highly purified protein solutions
such as protein therapeutics or in studies of protein folding kinetics, where limits of
low signal intensity among Tyr and Trp FRET pairings are not so problematic.
In general, however, endogenous fluorophores are not heavily used for FRET-based
studies of proteins.
Almost 60 years ago, Weber demonstrated that the quantum yield of 1-anilino-
phthalene-8-sulfonic acid (1,8-ANS) increased significantly from a highly quenched
state in water (0.004) when bound to bovine serum albumin (0.75) (Daniel and
Weber 1966; Weber and Daniel 1966; Weber and Laurence 1954). The mechanism
of this effect involves a relief from collisional quenching from H2O molecules due
to direct binding to a hydrophobic drug binding site on albumin. More commonly,
the nonspecific binding of ANS to hydrophobic surfaces of proteins has been used
as a general indicator of partial or complete protein unfolding.
In contrast to intrinsic protein fluorescence, which arises from the naturally fluo-
rescent amino acids present in a given protein, extrinsic fluorescent dyes offer addi-
tional possibilities in the characterization of proteins. These can be covalently
attached to proteins (via amines as with lysine or the amino group of the N-terminus
or via free thiol groups on cysteine) or may interact with a protein via noncovalent-
binding interactions. While ANS and other extrinsic fluorescent dye probes are a
useful general probe of a change in the conformation of a protein (Hawe et al. 2008,
2011), they provide only a qualitative, nonlocalized means of investigating struc-
tural perturbations. As described in Chap. 2, these properties have been exploited
for high throughput assays, especially as screening assays, where qualitative rank-
ing of different molecules or conditions is the desired outcome. When more quanti-
tative assessment is needed, extrinsic fluorescence is generally complimented by
other techniques.
Extrinsic dyes work via excited-state reactions. These are molecular processes
which change the electronic or chemical structure of the excited state prior to relax-
ation via fluorescence. The best known example of an excited-state reaction is that
of phenol deprotonation, which occurs much more readily from its excited state due
3 Techniques for Higher-Order Structure Determination 43
to a shift of electrons from the phenolic hydroxyl onto the phenol ring, which
acidifies the hydroxyl group. Collisional quenching and solvent relaxation are other
well-characterized examples of excited-state reactions, and the ones most relevant
to extrinsic dyes.
Solvent relaxation is one aspect of the local environment of a polar fluorophore
that can significantly influence its emission spectrum. This is the origin of the
Stokes’ shift. Spectral shifts result from the interaction of the excited-state fluoro-
phore dipole with solvent dipoles, leading to another energy transfer pathway,
depicted in Fig. 3.2. This pathway for energy transfer is due in large part to the
kinetics of different phenomena and can result in large Stokes’ shifts of ANS and
similar dyes. Absorption lifetimes are affectively instantaneous (10−15 s) with respect
to atomic motions, the Franck–Condon principle. The excited-state lifetime of fluo-
rophores are typical 10−9 to 10−8 s. Solvent motions are generally on the order of
10−10 s, which allows sufficient time for solvent dipoles to align with the fluorophore
excited dipole and for non-radiative transfer to occur. Fluorescence still occurs, but
at a lower, relaxed energy level for the excited state. With the loss of the fluorophore
dipole upon fluorescence emission, the aligned solvent molecules are at a higher
energy state, which relaxes to the original ground state through non-radiative pro-
cesses. The end result is a fluorescence that is red shifted (lower energy) from the
original absorption frequency.
Steady-state fluorescence spectroscopy is the most common experimental tech-
nique used to follow the interactions of extrinsic dyes with proteins. Protein binding
44 J. Kranz et al.
The use of Fourier transform infrared (FTIR) spectroscopy to measure the second-
ary structure of protein molecules has dramatically expanded in the last few decades,
encouraged by the need to measure protein secondary structure in samples with
various physical forms without sample manipulation. This method can be used to
analyze molecules ranging from a few amino acids to large protein complexes, in
different physical forms including solutions, lyophilized or crystallized solids, gas,
and chemically modified or embedded in polymers (Perez et al. 2002; Carpenter
et al. 1998; Haris and Chapman 1995; Chalmers 2002; Griebenow and Klibanov
1995). Additionally, modern FTIR spectrophotometers require only small volumes
of solution for analysis, 10–100 μL, though at a relatively high concentration of
10 mg/mL and higher (Perez et al. 2002; Zuber et al. 1992). Compared to other
spectroscopic techniques, FTIR can produce high-quality spectra relatively easily
without the complications of background fluorescence, light scattering or problems
related to the size of the protein. However, the presence of water absorption, buffer
components, or other biological molecules produces problems of spectral overlap
that can only be successfully subtracted and resolved by mathematical approaches
(Susi et al. 1967; Venyaminov and Kalnin 1990a, b; Dong et al. 1990; Dousseau and
Pezolet 1990). Thus, FTIR spectroscopy can be used to study various biological
systems including proteins in pharmaceutical and non-pharmaceutical formulations
(Carpenter et al. 1998; Haris and Chapman 1995; Griebenow and Klibanov 1995;
Krimm and Bandekar 1986).
FTIR spectroscopy typically refers to the absorption of infrared light in the mid-
IR region (400−4,000 cm−1). The Michelson interferometer (Fig. 3.3) is the central
part of an FTIR spectrometer and is unique among other spectroscopic instrumen-
tation. Infrared radiation from the source is collected and collimated before it
strikes the beam splitter. The beam splitter ideally transmits one-half of the radia-
tion and reflects the other half. Both transmitted and reflected beams strike mirrors,
which reflect the two beams back to the beam splitter. Thus, one-half of the infra-
red radiation has first been reflected from the beam splitter to the moving mirror
(Mirror 2) and then back to the beam splitter prior to being shone through the
sample. The other half of the infrared radiation going to the sample first goes
through the beam splitter and then is reflected from the fixed mirror (Mirror 1)
back to the beam splitter. When these two optical paths are reunited, interference
occurs at the beam splitter because of the difference in optical paths caused by the
scanning of the moving mirror.
3 Techniques for Higher-Order Structure Determination 45
The interference signal measured by the detector as a function of the optical path
length difference is called the interferogram (Chalmers 2002; Griffiths and de
Haseth 1986; Smith 1996). In order to extract and present information as an absor-
bance or transmittance spectrum, the interferogram has to be Fourier transformed,
generating a single beam spectrum of IR absorption as a function of wavelength.
Instrumental and atmospheric contributions are superimposed in the primary IR
spectrum. To eliminate these contributions, a background spectrum obtained with-
out a sample must be recorded and subtracted. Finally, the sample spectrum must be
normalized, by dividing the sample spectrum signal, I, and by the background sig-
nal, I0. The result is a %-transmittance spectrum (as T = I/I0). For liquids, transmit-
tance is related to absorbance A as
I 1
A = log10 0 = log10 . (3.6)
I T
Physically, FTIR is a type of vibrational spectroscopy measuring the fluctuating
covalent bonds within a protein; these give rise to specific wavelengths for different
modes of vibration, the intensity relating to the number of bonded atoms in a com-
mon chemical environment. There are many modes of vibration a chemical bond can
exhibit in a given chemical structure which result in IR absorbance, even for simple
molecules such as CO2 (Fig. 3.4). Given the vast number of normal modes, the vibra-
tional FTIR spectrum is complex with many vibrational bands overlapping. However,
it is often possible to select a spectral region that provides specific structural informa-
tion. Table 3.1 shows spectral regions of common chemical structure vibrations. For
analysis the spectra are usually deconvoluted. Better visual resolution of the different
bands can be achieved by taking the second derivative, and the data is often presented
as the deconvoluted or second derivative of the raw spectrum.
46 J. Kranz et al.
Fig. 3.5 Schematic
representation of examples of
peptide bond vibrations from
amide I and amide II
Fig. 3.6 Amide I FTIR spectrum second derivative of lysozyme in 50 mM phosphate buffer at pH
7 and 25°C. Red line is showing the derivative curve fit. Dotted line is showing the curve fit of
individual peaks
Nearly all organic molecules and macromolecules are “optically active,” arising
from a lack of chemical symmetry. For proteins the symmetry is provided by the
three-dimensional structure of the folded protein. The CD spectrum arises from an
electronically symmetrical chromophore in an asymmetric environment. The chro-
mophores that absorb light in the UV range are the peptide backbone and the aro-
matic amino acids and disulfide bonds. The result is electromagnetic interactions
between neighboring chromophores that can be detected spectroscopically. There
are a number of ways in which optically active samples can alter properties of trans-
mitted light, including linearly polarized light, circularly polarized light, circular
birefringence, and circular dichroism (CD) spectroscopy. With CD spectroscopy,
one measures the differential absorption of either left-hand or right-hand compo-
nents of circularly polarized light, the output being a measure of ellipticity, θ, of a
sample as a function of different wavelengths. Because this is a difference spectrum,
the signal can be either positive or negative, with the blank at zero.
The contribution of different secondary structural elements to the total CD signal
is additive, but also wavelength-dependent (Table 3.3). Thus, CD can be used to
estimate the relative fraction of different secondary structural elements of a protein,
α-helix, β-sheet, and random coil, directly from the CD spectrum without any other
experimental input, as described extensively by Greenfield (Greenfield and Fasman
1969; Greenfield 1996, 2006a, b, c, 2007) as well as many others. However, the
structural environment around the aromatic residues can result in signals in the far-
UV CD spectrum that can complicate this analysis.
In the near-UV CD region the combination of aromatic and disulfide environ-
ments present in the protein structure results in a complex spectrum which is like a
finger print for that particular protein. Each of the aromatic amino acids has a char-
acteristic wavelength profile that corresponds to their absorbance spectra. Trp shows
a peak close to 290 nm with fine structure between 285 and 305 nm; Tyr has a peak
between 270 and 285 nm, with a shoulder at longer wavelengths often obscured by
the Trp band; Phe shows bands with fine structure between 250 and 265 nm.
Disulfide bonds also absorb in the near-UV region (weak broad absorption bands
from 250 to 280 nm), the changes in the dihedral angle of the disulfide bond will
result in a change in the signal in this region of the spectrum. The actual shape and
magnitude of the near-UV CD spectrum of a protein will depend on the protein
primary sequence, the number of each type of aromatic amino acids present, their
mobility, and the nature of their environment (H-bonding, polar groups, and polariz-
ability). While specific residue assignments cannot be made, changes in the protein
fold result in changes in the environment of the aromatic amino acids and disulfide
bonds and thus in the near-UV CD spectrum. As with fluorescence, the complexity
of the aromatic CD spectrum increases with increasing numbers of aromatic groups.
CD spectroscopy has been used extensively to characterize protein folding and
unfolding, both as a function of denaturants and temperature. It is a truly general
technique that is not limited to constraints of molecular size or chemical composition
and is a “label-free” approach that does not require secondary conjugation or chemi-
cal modification to obtain data on protein structural composition. Moreover, modern
improvements in CD instrumentation and data fitting have made it straightforward
to collect full spectra as a function of temperature, providing a means to quantitate
the presence of subtle fluctuations in protein structure that occurs at temperatures
below the primary protein unfolding transition.
Vibrational optical activity (VOA) is the field of spectroscopy associated with com-
bining infrared absorption or Raman scattering with the optical activity, manifested
as either optical rotation or CD. These include infrared vibrational circular dichro-
ism (VCD) and vibrational Raman optical activity (ROA). VCD is an extension of
electronic circular dichroism (ECD) described above, using an infrared spectrome-
ter (FTIR) to measure the difference in the IR intensity for left minus right circularly
polarized radiation, while ROA measures the difference in intensity of right minus
left circularly polarized Raman scattered radiation in several different instrument
configurations (Nafie 2011). VOA measurements are complementary to typical
spectroscopic methods such as CD and FTIR spectroscopy, but with additional sen-
sitivity in monitoring subtle changes in tertiary structure (Nafie 2011; Nafie and
Dukor 2007; Lakhani et al. 2009; Cao et al. 2008; Keiderling et al. 2006; Zhu et al.
2006; Barron 2006).
In a typical optical activity instrument, light emerging from a Michelson interfer-
ometer is polarization modulated using a photoelastic modulator (PEM), with light
oscillating between left and right circularly polarized states at the PEM frequency
3 Techniques for Higher-Order Structure Determination 51
(Fig. 3.7). The infrared light is modulated by the Michelson interferometer and then
enters the VCD portion of the setup. By passing through a linear polarizer and a
PEM, the Fourier-modulated IR beam is further modulated at the PEM frequency.
To obtain a VCD spectrum, the doubly modulated signal is first demodulated at the
PEM frequency by a lock-in amplifier, then Fourier transformed by a computerized
system (Keiderling et al. 1999).
A chiral molecule interacts differently with left and right circularly polarized light, in
contrast to classical IR spectroscopy in which vibrational excitation occurs with nonpo-
larized IR radiation and thus does not vary due to differences in chirality. In certain
applications of VCD, it is possible to infer relative absolute configuration between pairs
of structurally related molecules. Such correlations are strengthened due to the large
number of distinct vibrational bands in a typical mid-infrared and Raman scattering
spectrum. This high level of sensitivity to minor structural change gives optical activity
technologies, VCD and Raman optical activity, a high level of analytical power to eluci-
date both absolute molecular structure and conformation (Nafie and Dukor 2007).
Aromatic residues and polysaccharides are known to have overlap in the far-UV
CD region, but show good separation in a VCD spectrum (Shi et al. 2006; Dukor
and Keiderling 1991). The sensitivity of VCD to interactions of dipoles and the
combination of sign, pattern, and band shapes with frequency allow VCD to provide
local structural information and to distinguish between different secondary struc-
tures (Shi et al. 2006; Dukor and Keiderling 1991). Moreover, VCD, like the other
types of vibrational spectroscopy, can be used across different types of samples
including high concentration solutions (>50 mg/mL), solids, protein aggregates,
and foreign organic particles. VCD is useful for monitoring formation of amyloid-
like fibrils, with an order of magnitude increase in the VCD intensity upon fibril
formation, reporting on the long-range supramolecular chirality of protein fibrils
(Shi et al. 2006; Dukor and Keiderling 1991; Meyer et al. 2004; Baumruk and
Keiderling 1993; Yoder et al. 1997). Moreover, VCD spectroscopy should be trans-
parent to many commonly used formulation components that can interfere with CD
and FTIR spectroscopy measurements. Glycine is commonly used in protein
therapeutic drug product formulations, but the carboxyl group of free glycine in the
52 J. Kranz et al.
Like infrared (IR) spectroscopy, Raman spectroscopy is a method for probing struc-
tures and interactions of molecules by measuring the energies (or frequencies) of
molecular vibrations. The main advantage of Raman spectroscopy for applications
54 J. Kranz et al.
to proteins and their complexes is the fact that liquid water (both H2O and D2O) does
not confound the Raman effect. Also like FTIR samples in different physical forms
can be analyzed including solutions, suspensions, gels, precipitates, fibers, single
crystals, and amorphous solids, with no complicated sample preparation necessary.
Raman spectroscopy is nondestructive, with high specificity for certain structures
and applicability to large supermolecular assemblies (Nafie 1996).
IR and Raman differ fundamentally in the mechanism of interaction between
radiation and matter (Benevides et al. 2004). Raman scattering is inelastic light
scattering which occurs at wavelengths that are shifted from the incident light by
the energies of the molecular vibrations. Typical applications include structure
determination, multicomponent qualitative analysis, and quantitative analysis
(Nafie 1996; Dong et al. 1998). The vibrational spectrum of a molecule is com-
posed of bands representing active normal vibrations. The spectrum depends on
the masses of the atoms in the molecule, the strength of their chemical bonds, and
the atomic arrangement. Consequently, groups of atoms connected by certain types
of bonds have certain characteristic vibrations in the Raman spectra. Table 3.4
shows a set of common chemical structures with their frequency regions in Raman
spectra (Socrates 2004).
A Raman system typically consists of four major components (Fig. 3.10a): an
excitation source such as a laser, a sample illumination system, light collection
optics, a wavelength selector which can be either a filter or spectrophotometer, and
the detector which can be a photodiode array, a charge coupled device, or a photo-
multiplier tube. A sample is normally illuminated with a laser beam in the ultravio-
let (UV) visible (Vis) or near infrared (NIR) range (Dong et al. 1998). Other
state-of-the-art Raman spectrometer systems have been described for various bio-
logical applications such as UV resonance Raman [UVRR (Brennan et al. 1997;
Hashimoto et al. 1993; Asher et al. 1993)], Raman optical activity [ROA (Nafie
1996; Grauw et al. 1997; Barron et al. 1996)], and confocal Raman microscopy
(Goldstein et al. 1996). In all cases scattered light is collected with a lens and sent
through an interference filter or spectrophotometer to obtain the Raman spectrum of
the sample being analyzed. In the case of Raman microscopy, the microscope is
3 Techniques for Higher-Order Structure Determination 55
Fig. 3.10 Schematic of (a) basic Raman spectroscopy components and (b) energy level diagram
showing the states involved in Raman signal. The line thickness is roughly proportional to the
signal strength from the different transitions
used to visualize the sample and ensure that the spectrum is from a particular p article
or region of the sample.
Raman spectra are obtained when the light interacts with the molecule and dis-
torts (polarizes) the cloud of electrons round the nuclei to form a short-lived meta-
stable state from which photons are quickly reradiated (Fig. 3.10b). If only electron
cloud distortion is involved in scattering, the photons will be scattered with very
small frequency changes. This elastic scattering reaction is called Rayleigh scatter-
ing and is the dominant process following electron cloud distortion. However, if
light scattering induces relaxation through nuclear motions, then energy will be
transferred either from the incident photon to the molecule or from the molecule to
the scattered photon. In these cases the process is inelastic and the energy of the
scattered photon is different from that of the incident photon by one vibrational unit.
It is an inherently weak process in that only one in every 106–108 photons scattered
results in Raman scattering. However, it is very sensitive to subtle structural changes
and with the development of high-power modern lasers and microscopes this tech-
nique is now much more readily available and in use.
Changes in the molecular geometry or conformation that characterize many
molecular biological phenomena can produce relatively large changes in Raman
band positions, which are usually referred to as frequency shifts. Such shifts in the
Raman band frequencies are the basis for applications of the technique to analyze
protein secondary structures, monitor tertiary structure changes, determine side-
chain conformations, and detect intramolecular interactions. Because the molecular
geometry and force field are also sensitive to interactions between molecules,
Raman can be effectively applied to probe protein intermolecular interactions,
including the formation on specific complexes and large assemblies.
A typical protein spectrum contains virtually all of the fundamental vibrational
information on a protein or nucleic acid molecule, except for hydrogen stretch
modes which generate bands in the 2,400–3,600 cm−1 Raman interval (Barron et al.
2000). Raman intensities associated with protein secondary structure are typically
produced by the vibrations of the C=O and C–N chemical groups stretching in the
56 J. Kranz et al.
Table 3.5 Amide I and amide III Raman bands of representative polypeptides and proteins with
secondary structure assignments
Frequency (cm−1) Frequency (cm−1) Secondary structure
Molecule amide I amide III assignment
α-Poly-l-alanine 1,655 1,265–1,348 α-Helix
α-Poly-l-glutamate 1,652 1,290 α-Helix
α-Poly-l-lysine 1,645 1,295–1,311 α-Helix
β-Poly-l-alanine 1,669 1,226–1,243 β-Strand
β-Poly-l-glutamate 1,672 1,236 β-Strand
β-Poly-l-lysine 1,670 1,240 β-Strand
Poly-l-lysine, pH 4 1,665 1,243–1,248 Irregular
Poly-l-glutamate, pH 11 1,656 1,249 Irregular
cI (Lamda) repressor 1,675 1,245 Turns/irregular
(1–102)
Bacteriophage P22 subunit 1,655 1,235 β-Strand/turns
peptide bonds. Stretching vibrations of C–C, C–N, and C=O bonds also produce
intense Raman bands, especially if they involve the concerted symmetrical displace-
ments of side-chain skeletons. The Raman bands associated with the bending and
stretching modes of individual hydrogens in the protein substituent (e.g., C–H,
N–H, and O–H) are generally weak, but their collective spectral intensities, which
result from the large numbers of such groups in a protein, can be significant.
Generally, application of Raman spectroscopy to protein secondary structure
analysis focuses on changes in the amide I spectral region (1,640–1,680 cm−1),
which is primarily a carbonyl stretching mode, as well as the amide III (1,230–
1,310 cm−1), which combines both in-plane N–H bending and C–N stretching
motions. Table 3.5 lists the Raman frequencies assigned to amide I and amide III
bands of representative polypeptides and proteins of differing secondary structures.
Other conformation-sensitive bands have been identified in Raman and ultraviolet
resonance Raman (UVRR) spectra of peptide model compounds and proteins. For
example, amide II (1,500 and 1,600 cm−1), which involves substantial C–N stretch-
ing, is useful in studies of protein α-helical content (Wang et al. 1991). Polarized
Raman spectra of α-helical proteins also generate a useful secondary structure
marker near 1,340–1,345 cm−1, which can be exploited for α-helix orientation anal-
yses (Tsuboi et al. 2000). Additional amide-related modes near 1,390 cm−1 have
been proposed in UVRR spectra which have been assigned to the UVRR the over-
tone of amide V and appear to be sensitive to the local conformation of the peptide
linkage (Wang et al. 1991).
Certain Raman signatures are unique to protein tertiary structure, notably vibra-
tional frequencies of side-chain conformations. For example, the Raman band
between 2,500 and 2,600 cm−1 resulting from the cysteine sulfhydryl bond (S–H)
stretching vibration is a unique probe of local SH structure and dynamics (Raso
et al. 2001). In-plane vibrations of the rings of aromatic side chains such as trypto-
phan, tyrosine, and phenylalanine are also expected to produce Raman bands of
3 Techniques for Higher-Order Structure Determination 57
high intensity. In addition, vibrations that involve the displacement of heavy atoms
such as sulfur in C–S stretching modes of methionine and cysteine, S–S stretching
of cysteine, S–H stretching of cysteine, and Zn–S stretching in zinc metalloproteins
are expected to be relatively intense.
The indolyl moiety of tryptophan generates many prominent Raman bands, sev-
eral of which have been correlated with the local environment and geometry of the
tryptophan side chain in proteins (Miura et al. 1991; Siamwiza et al. 1975). For
example, normal mode W3 generates an intense and sharp Raman band in the
1,540–1,560 cm−1 interval. The tryptophan residue also generates a Fermi doublet
with components at 1,360 and 1,340 cm−1, with an intensity ratio (I1360/I1340) that
increases with increasing hydrophobicity of the indolyl ring environment and thus
serves as an indicator of local hydropathy. The normal mode W17 Raman band near
880 cm−1 is sensitive to indolyl N–H hydrogen-bond donation which exhibits a rela-
tively high value when the indolyl moiety is located in a highly hydrophobic envi-
ronment, such as the hydrophobic core of a globular protein. Normal mode W18,
which is an indole ring-breathing vibration, generates bands near 755 cm−1 in Raman
spectra of proteins with increased intensity when the hydrophobicity of the indolyl
ring environment decreases.
Raman spectroscopy becomes an effective technology for various applications in
protein biotechnology due to its capability to simultaneously measure the secondary
and tertiary structure of a protein sample and to be applied to various physical
forms. In the field of protein therapeutics, Raman spectroscopy has been used to
measure protein HOSs in solution, in lyophilized form, or embedded in biocompat-
ible polymer, throughout process and formulation development. Additionally,
Raman spectroscopy can be very useful investigating incidents during stability stud-
ies and during manufacturing.
et al. 2005). These techniques are more rapid, since they do not require backbone
resonance assignments necessary for NMR-based HX studies, but have diminished
resolution with respect to changes in stability on a per peptide, rather than a per resi-
due basis, providing a more heterogeneous view of protein fluctuations.
3.3 T
hermodynamic and Hydrodynamic Techniques:
Shape and Size
There are two general calorimetric approaches for investigating the thermodynam-
ics of proteins and protein–ligand (or excipient) interactions: titration calorimetry
(Velazquez-Campoy et al. 2004; Ladbury and Chowdhry 1996; Jelesarov and
Bosshard 1999; Weber and Salemme 2003) and scanning calorimetry (Jelesarov
and Bosshard 1999; Weber and Salemme 2003; Sanchez-Ruiz 2011; Spink 2008;
Krishinan and Brandts 1978; Brandts and Lin 1990). All calorimetric techniques are
“label-free,” that is, they do not require the presence of a particular fluorophore or
special labeling, but rather rely on direct measurement of thermodynamic changes
that arise during an experiment.
One of the most routine methods for candidate and formulation screening during
biopharmaceutical development is differential scanning calorimetry (DSC). In DSC
a reference cell (H2O or buffer) and a sample cell (containing the protein solution
60 J. Kranz et al.
being analyzed) are heated at a constant rate of change in temperature, and the
difference in the energy provided to the two cells, or heat capacity, in order to
achieve this rate is recorded. The output of DSC experiments is an evolution of heat
(enthalpy) corresponding to the net thermodynamic loss of tertiary structure and
rehydration of unfolded protein and a midpoint temperature, Tm, that represents the
point at which the sample is equally populated by folded and unfolded forms. The
thermodynamics of a reversible reaction can be obtained from these experiments,
while for irreversible reactions, relative differences in conformational stability or
stability to aggregation can be obtained from differences in Tm or in the temperature
at which unfolding begins. For multidomain proteins such as antibodies, it is often
possible to resolve the transition from each domain, providing information about
individual parts of the protein that is difficult to obtain by other techniques. The
application of DSC is somewhat limited due to the fact that it is a low-throughput
method that requires about 1 mg of protein.
DSC has applications in measuring ligand binding as well (Brandts and Lin
1990). Usually binding of a ligand increases the stability of the protein, increas-
ing the protein Tm in a manner that depends both on ligand affinity and on con-
centration. As described in Chap. 2, a higher-throughput approach that involves
heating a solution in the presence of an extrinsic fluorescent probe, which fluo-
resces only in the presence of unfolded protein, has been developed and is
broadly applied when material and time are limited. This technique is variously
referred to as differential scanning fluorimetry (Goldberg et al. 2010; He et al.
2009a, b; Li et al. 2011), ThermoFluor™ (Mezzasalma et al. 2007; Matulis et al.
2005), or a thermal shift assay (Kranz and Schalk-Hihi 2011; Cimmperman et al.
2008; Zubriene et al. 2009; Norvaisas et al. 2012; Zhang and Monsma 2010).
Each of these has been adapted to plate-based unfolding studies, which reduce
sample requirements by several orders of magnitude and allow for parallel
unfolding or ligand-binding studies. Thermal shift assays are quite useful in
excipient screens, allowing for rapid identification of stabilizing conditions for a
given protein. DSC can then be used as a sort of gold standard, providing support
for the results obtained with the higher-throughput methods and increasing
understanding of the underlying reactions.
Light scattering studies have been utilized extensively when structural informa-
tion is needed for protein molecules dissolved in solution (Van Holde et al. 2006).
For this purpose, both static light scattering (SLS) and dynamic light scattering
(DLS) studies (most commonly using light in the visible range) have been
employed; the former is primarily used for molecular weight and dimensional
analysis, and the latter is most useful for effective hydrodynamic size measure-
ments. Both these techniques and some key areas of application in protein devel-
opment are presented below.
3 Techniques for Higher-Order Structure Determination 61
SLS studies refer to the analysis of the intensity of the light scattered from a
molecule in solution. In simple terms, scattering can be defined as deflection of an
incident beam of light by a molecule. However, a number of phenomena take place
from the instant that light impacts the molecule to the moment it is dispersed. When
focused on a molecule, linearly polarized light, an electromagnetic wave moving in
a particular direction (e.g., x-axis) with its electric vector oscillating perpendicular
to this direction (e.g., z-axis) induces corresponding oscillation in the electrons of
the molecule. Such oscillations of the electron in the molecule result in generation
of oscillating dipoles and subsequent emission of some of the incident energy in the
form of light in a direction different from that of the incident light. When the size of
the scattering molecule is significantly smaller than the wavelength of the incident
light beam (about 50-fold), a phenomenon described as Rayleigh scattering occurs,
the ratio of the scattered light intensity to that of the incident light can be expressed
as either of the following two expressions depending on the angle of observation
with respect to either the z- or x-axis:
i 16p 4a 2 sin 2 j
= , (3.7)
I0 r2l 4
i 8p 4a 2
= 2 4 (1 + cos2 q ), (3.8)
I0 r l
where i is the intensity of the scattered light, i0 is intensity of the incident light, a is the
molecular polarizability, π is the angle between the direction of scattered light and the
direction of the oscillation of the electric vector (z-axis), r is the distance between the
point of observation and the position of the scattering particle, l is the wavelength of the
incident light, and q is the angle between the direction of the scattered light and that of
the incident beam (x-axis). It is apparent from this expression that (a) there will be no
scattering along the direction of the electric vector oscillation of the incident light, that
is, when π = 0, (b) the scattering intensity decreases with increasing distance from the
molecule and more significantly with increasing wavelength and (c) the scattering inten-
sity will be symmetrical in the forward and backward direction, that is, there is no angu-
lar dependence of the scattering light intensity. Equation 3.8, expressing the scattering
from a single molecule, can be expanded further to represent the scattering from a col-
lection of such molecules in a dilute solution to
2
dn
2p 2 n02
iq dc
= cM (1 + cos2 q ), (3.9)
I0 r l N
2 4
where iq is the cumulative scattered intensity from the collection of molecules
observed at an angle q from the direction of the incident light, n0 is the refractive
62 J. Kranz et al.
index of the solvent, dn/dc is the refractive index increment of the solution per unit
concentration of the added solute, N is Avogadro’s number, c is the molecule or
solute concentration, and M is the molecular weight of the solute. Equation 3.9 can
be further simplified to the more commonly used expression:
iq r2
Rq = = KcM , (3.10)
I 0 (1 + cos2 q )
where Rq is defined as the Rayleigh ratio. In Eq. 3.10, K is an instrumental con-
stant, since it is fixed for a given instrument and particular solute–solvent system,
defined as
2
dn
2p n
2 2
dc
0
K= . (3.11)
l N
4
Rq = KcM w , (3.12)
with Mw being the weight-averaged molecular weight of the solute since each asso-
ciated state contributes to the average molecular weight based on its total weight
fraction instead of its number fraction. Equation 3.7 is commonly employed in pro-
tein development studies to determine the molecular weight of protein in solutions
as a function of any given solution variable or over the shelf life of the product. It is
assumed here that the solution being studied is sufficiently dilute so that the inter-
molecular interaction between multiple protein molecules does not affect the overall
scattering behavior. In other words, the expression applies to ideal solutions.
However, this assumption is often invalid, and intermolecular interactions signifi-
cantly impact the scattering behavior of the molecules and solution.
In protein development, the most common application of static light scattering
has been its use for the determination of Mw of the protein in relatively dilute solu-
tions (Wyatt 1993). Since the formation of aggregates and other higher-order spe-
cies is considered a potential patient safety risk, early detection of propensity to
aggregate is critical to successful product development (Goldberg et al. 2011).
In addition to molecular weight measurements, static light scattering is also
employed to obtain dimensional information. For this to be possible, the size of the
molecule of interest has to be large enough to be treated as a point scatterer. In this
case, the scattering behavior for even a single molecule varies depending on the
location of the molecule where the incident light strikes. The scattered light thus
varies with the angle of measurement. With multiple light waves now capable of
striking the molecule simultaneously at different points, interference between the
3 Techniques for Higher-Order Structure Determination 63
light scattered from these multiple scattering locations needs to be accounted for.
The key attribute derived from the scattering data from larger molecules is the radius
of gyration (RG) defined as the root mean square average of the distance of the scat-
tering element from the center of mass of the molecule. For Rayleigh scattering to
be valid, l /RG should be in excess of about 50. Thus, for instruments equipped with
a 600 nm or higher laser, molecules with a radius of gyration of less than 12 nm can
be treated as point scatterers and will thus not exhibit any angular dependence of the
scattering intensity. Obviously in such cases, it will not be possible to determine the
radius of gyration of the particle and the primary application of light scattering will
be for the determination of molecular weight and the association state (e.g., mono-
mer, dimer). For globular proteins, the average hydrodynamic diameter (discussed
in the next section) has been found to be ~3 times the radius of gyration, and thus
when using a 600 nm laser, Rayleigh scattering will be valid for molecules with
radii as high as 36 nm (Mandel 1993). This is usually the case for most proteins in
development for therapeutic application. For example, for immunoglobulins, one of
the larger proteins frequently encountered in development with a mass of ~150 kD,
the hydrodynamic diameter is ~12 nm. In cases where information regarding the
dimensions of the molecular radius of gyration is required, using instruments with
incident light at less than 600 nm (ultraviolet light or X-ray) can help. But even this
approach has limitations since below 300 nm, protein molecules start absorbing,
and below about 160 nm, water itself starts absorbing. X-rays with wavelengths
below 10 nm, to which most materials are transparent, have been used at small
angles to overcome these limitations for analyzing dimensions of small and globular
protein molecules.
Static light scattering has been an excellent tool for characterizing the molecular
weight and the higher-order species in protein solutions. However, it has found lim-
ited application for effective size measurement in solution given the lack of angular
dependence for most of the molecules currently being developed. In instances where
hydrodynamic properties need to be measured, biophysical chemists have often
relied on the technique of DLS which is the focus of the next section.
Unlike static light scattering experiments, which capture the time-averaged scat-
tered light intensity of a solution sample, DLS is concerned with the fluctuations in
the intensity of the scattered light over relatively small time periods for a small
volume fraction of the total solution. On a molecular level, whereas static measure-
ments track the total scattered intensity from a collection of molecules, dynamic
studies track the fluctuations in the intensity of light scattered from a collection of
molecules as a function of time. This concept can be expanded to analysis of a small
volume element (or fraction) of the total solution wherein a large number of such
molecules are in constant motion. This motion is primarily due to Brownian motion,
but can also be due to intermolecular interactions. As a result of this constant motion
and the constantly changing location and orientation of these molecules in the finite
64 J. Kranz et al.
volume element, the light scattered from that volume element exhibits fluctuations
in intensity. This fluctuation in scattered intensity is used to calculate the average
size of the molecule in solution. Since water (or solvent) molecules closely associ-
ate with the molecule of interest and diffuse with it as a single entity, DLS measures
the size of the hydrated (or solvated) molecule.
A solution of larger molecules will undergo intensity fluctuations at a slower
rate, due to slower movements in and out of the portion of the volume being moni-
tored, compared to a solution of smaller-sized molecules. A DLS instrument tracks
these intensity fluctuations over nano- and microseconds and in one (and the most
commonly employed) variations of the technique correlates the intensity at time, t
with that at time t + Δt, t + 2Δt, and so on through an autocorrelation function repre-
sented as
〈i(t )•i(t + ∆t )〉
g ( 2 ) ( ∆t ) = , (3.13)
〈i 〉 2
where g(2) is the second-order autocorrelation function, i(t) is the scattered intensity
at time, t i(t + Δt) is the scattered intensity following an interval of Δt, and the
brackets (〈〉) indicate the averaged product of scattered intensity over time. The
symbol 〈i〉 represents the average values of scattered intensity. At short time inter-
vals (lower multiples of Δt), the correlation between scattered intensity values will
be high since the molecules would not have moved significantly from their initial
position and there is consistency in scattering states. Thus, higher and more stable
values of the autocorrelation function are obtained at shorter time intervals. As the
time delays become longer (at higher multiples of Δt), correlation between scattered
intensities decreases, since there is decreasingly little correlation between scattering
states, and more movement of the molecules in and out of the volume being moni-
tored. When a single component of the solution dominates the scattering from the
solution (monodisperse solutions), the autocorrelation function decays exponen-
tially such that
{ }
l n g ( 2 ) ( ∆t ) − 1 = ln a − 2q 2 D∆t , (3.15)
4pn q
q= sin , (3.16)
l 2
where q is the scattering vector, n represents the solution refractive index, q is the
scattering angle, l is the wavelength of incident light, a is an instrumental constant,
and D is the translational diffusion coefficient (referred to simply as the diffusion
coefficient) since it characterizes translational movement of a molecule from one
point to another rather than rotational movements about a given axis. Equation 3.15
can be fitted to the system-generated autocorrelation function data over time to cal-
culate the diffusion coefficient of a given component of the solution. It must be
3 Techniques for Higher-Order Structure Determination 65
noted that, similar to Rayleigh scattering, Eq. 3.15 can be used for the determination
of the diffusion coefficient only as long as the dimensions of the scattering mole-
cules are small compared to the wavelength of light. If this condition is not satisfied
and the molecular size approaches that of the wavelength of incident light, rota-
tional as well as internal movements of the molecules also contribute to the change
in the scattering pattern and intensity with time. In this case, the calculated autocor-
relation function does not accurately reflect the translational diffusion coefficient of
the molecule.
From the above discussion, it is apparent that DLS directly measures the diffu-
sion coefficient, not the size, of a given solute in solution. The effective or the
hydrodynamic size is subsequently derived through the Stokes–Einstein relation
assuming a spherical geometry for the molecule:
kT kT
Dm = = . (3.17)
f 3phdH
In Eq. 3.17, f is the frictional coefficient or a measure of drag on the diffusing
molecule, k is the Boltzmann constant, T is the absolute temperature, and η is the
solution viscosity. Therefore, the diffusion coefficient, sometimes called the self-
diffusion coefficient (Ds), of the solute, is inversely related to the hydrodynamic
diameter (dH) or the effective size of the hydrated solute molecule (Eberstein et al.
1994; Schümmer 1978). It is important to note that the Stokes–Einstein equation is
valid for spheres, and thus DLS measurements provide only an apparent hydrody-
namic size of the molecule assuming the particle under examination to diffuse in a
manner similar to a sphere of the same size. Consequently, globular proteins in DLS
measurements exhibit a smaller measured diameter (or larger Ds) compared to ran-
dom coils of the same molecular weight since random coils, due to more complex
and interfered motions and interactions with the solvent, diffuse more slowly in the
solution. Thus, DLS measurements incorporate the effect of both hydration and
shape of the molecule in measuring the size of the molecule. Any change in the
shape or hydration of the molecule that results in a change in the diffusion behavior
of molecules can be detected by DLS studies. Conformational changes in the pro-
tein molecules in solution, occurring as a function of solution environment, can thus
be readily detected via this technique.
The use of DLS is usually limited for particles in the submicron size (<1–2 μm).
Most common instruments have a practical upper limit of ~1 μm. Fortunately, most
protein monomers are significantly smaller, in low nanometer range, and thus can
be studied via DLS. Additionally, the formation of higher molecular weight or
aggregated species can also be detected since a considerable population of aggre-
gated species can exist at sizes below 1 μm.
The form of the Stokes–Einstein relation expressed in Eq. 3.17 assumes an ideal
solution behavior where intermolecular interactions, or in the case of protein solu-
tions protein–protein interactions, do not affect the movement of protein molecules
and thus do not impact the measured diffusion coefficient. However, this may not be
true in all cases. For ideal solutions, the diffusion coefficient depends on the
66 J. Kranz et al.
h ydrodynamic size of the protein molecule (in addition to viscosity and t emperature).
In other words, for systems in which protein–protein interactions are too weak to
influence the diffusion of the protein molecule, the diffusion coefficient is indepen-
dent of the protein concentration, and the measured dH is an absolute measure of the
effective hydrodynamic diameter of the protein. However, for solutions in which
protein–protein interactions affect translational movement of the protein molecules,
the diffusion coefficient is a function of the protein concentration as well. Thus,
DLS measurements should be carried out over a series of protein concentrations
instead of a single concentration in order to tease out the effect of molecular interac-
tions on molecular size measurements.
As discussed in several other chapters in this volume, DLS has become a potent
tool for use during candidate and formulation screening, where the ability to obtain
high throughput, relative qualitative results is perfectly suited for the requirements
of these stages of therapeutic development.
During an SV-AUC experiment, there are three forces acting on a single protein
molecule: (1) the gravitational force mω2r where m is the protein mass, ω2 is the
rotor angular velocity, and r is the distance from the center of rotation; (2) the buoy-
ancy force –mν(bar)ρω2 where ν (bar) is the protein partial specific volume and ρ is
the solvent density; and (3) the frictional force s(kT/D)ω2 where k is the Boltzmann
constant, T is the absolute temperature, and D is the diffusion constant. The sedi-
mentation coefficient s, measured in Svedberg units, is defined as s = v/ω2 where v is
the absolute velocity. The fundamental equation in AUC, the Svedberg equation, is
derived from the balance of the three forces acting on a single molecule:
3 Techniques for Higher-Order Structure Determination 67
s M (1 − ns )
= , (3.18)
D RT
where M is the molar mass, and R is the gas constant. The sedimentation and
diffusion of a protein in solution are measured in SV-AUC experiments as a series
of sedimentation profiles and can be described by the Lamm equation:
dc 1 d dc
= rD − sw 2 r 2 c , (3.19)
dt r dr dr
where χ is the concentration of the protein at a distance r from the center of rotation
at time t for a given angular velocity ω. Protein samples are usually not analyzed
under standard conditions, that is, aqueous solutions at 20°C; therefore, for the pur-
pose of comparing various samples, the s values obtained under different nonstan-
dard conditions can be converted to the standard conditions using the equation:
h h 1 − vr20,w
s20,w = s T,w s , (3.20)
h20,w hw 1 − vrT,w
where η20,w is the viscosity of water at 20°C, ηT,w is the viscosity of water at the
experimental temperature T, ηw is the viscosity of water at the experimental tem-
perature and ηs is the viscosity of the solvent at the same experimental temperature,
ρ20,w is the density of water, and ρT,s is the is the density of the solvent at the experi-
mental temperature. The theory behind protein sedimentation and AUC has been
described in detail in multiple books (van Holde et al. 2006; Machtle and Borger
2006; Scott and Schuck 2005; Philo 2005). Advanced AUC techniques and strate-
gies for experimentation and data analysis were described in several critical reviews
and tutorial articles (Philo 2006; Liu et al. 2006; Schuck et al. 2002).
In an SV-AUC experiment a protein sample is centrifuged, usually over several
hours, while the progress of centrifugation is monitored by a detection system, most
commonly UV absorbance. A typical result of SV-AUC analysis is presented in
Fig. 3.11. The primary information obtained from an AUC experiment is the sedi-
mentation coefficient. This can be converted (with certain presumptions about
shape) into other information better understood by general protein chemists work-
ing in formulation development, that is, molecular weight, MW distribution
(Fig. 3.11b), or hydrodynamic radius.
There are multiple approaches and also choices that are made during the perfor-
mance and subsequent data analysis of an AUC experiment that are important for
the interpretation of the data. The sample needs to be diluted to be in the linear
range of the detector (below optical density of 1.5 for UV absorbance) in buffer. As
many protein therapeutics are formulated at high concentration above 100 mg/mL,
the dilution can be significant. The first crucial decision is what buffer to use for
dilution; two approaches currently exist, to either use buffer without excipients that
68 J. Kranz et al.
a 5 b
5
4 4
3 3
c(s)
c(M)
2 2
1 1
magnified 20x magnified 20x
0 0
5 10 15 20 0 200 400 600 800 1000
Sedimentation coefficient (Svedberg) Molecular weight (kDa)
Fig. 3.11 SV-AUC analysis of an IgG analyzed in formulation buffer diluted to 0.5 mg/mL. Panel
a shows size distribution obtained using continuous c(s) distribution model, and panel b presents
c(M) distribution model
might interfere with the analysis or to use the original formulation buffer. Many
buffers used in protein formulation contain substantial concentrations of excipients,
mostly sugars, to enhance long-term stability. Dynamic gradients, caused by the
presence of the excipients, can have detrimental effects on the reliability of the data
obtained.
The sedimentation coefficient is the crucial parameter for HOS determination
and for detection of changes in HOS and/or changes in protein conformation and
shape. The value obtained from AUC experiments is in general dependent on pro-
tein concentration, as demonstrated in Fig. 3.12, and also solution conditions such
as density and viscosity. In order to eliminate the effect of protein concentration on
the sedimentation coefficient, it is necessary to extrapolate to zero protein concen-
tration. This is labor intensive because it requires at least several data points and
considerable experimental time. A more generally accepted approach is to compare
s values measured at the same protein concentration, for example, 0.5 mg/mL.
Similarly, the effect of density and viscosity on the sedimentation coefficients can
be compensated for by extrapolating the sedimentation coefficient s to standard
conditions at zero excipient concentration and 20°C (s20,w). The sedimentation coef-
ficient of an antibody monomer can be measured by AUC with great precision and
therefore changes in sedimentation coefficient values can be used to detect small
changes in protein conformation.
There are some situations where measurement of changes in s values has valu-
able application in formulation development: The conformational changes detected
by AUC can be also detected by other techniques, such as the spectrophotometric
techniques and ITC, described earlier in this chapter. However in some situations,
especially when the conformational changes are small, they may not be detected by
3 Techniques for Higher-Order Structure Determination 69
5 5
Process A Process B
4 4
3 3
c(s)
c(s)
2 2
Lot 3 Lot 3
1 1
Lot 2 Lot 2
Lot 1 Lot 1
0 0
2 4 6 8 10 2 4 6 8 10
Sedimentation coefficient (Svedberg) Sedimentation coefficient (Svedberg)
is remarkable considering that these were produced at two different locations at dif-
ferent points of time and analyzed by AUC in the original formulation buffer contain-
ing a considerable amount of excipients.
The SV-AUC method described in the previous section is currently extensively used
in protein development for aggregate characterization and for characterization and
for studies on the comparability of conformation. SE-AUC is an equilibrium method
and therefore does not provide any information on the molecular shape, unfolding,
3 Techniques for Higher-Order Structure Determination 71
Fig. 3.14 Sedimentation
1.5
equilibrium experiment of an
IgG antibody. The sample
was analyzed at three
0.0
3.3.4 Rheology
Rheology is the study of flow behavior of matter either in the liquid or the solid
state. However, for most practical purposes, it applies to liquids and semisolids and
other matter which undergo some degree of deformation or flow (viscous) and loss
of some or all of the applied energy as opposed to a material in which all of the
applied energy can be recovered (elastic, e.g., a perfect spring) (Schümmer 1978).
Viscosity (η) of a fluid is its resistance to flow, that is, its “internal friction” and is
the simplest expression of the rheological properties of a material. It can be defined
as the pressure [shear stress, τ] required to deform or to make a fluid flow at a unit
rate [shear rate, γ] in a direction perpendicular to the plane of applied pressure:
t
h= . (3.22)
g
Most of the fluids that we encounter in daily life exhibit viscosity which is a
function of the shear rate, that is, the viscosity will change as one varies the rate at
which the fluid is deformed. Such liquids are referred to as non-Newtonian. On the
other hand, for Newtonian fluids like water viscosity is not a function of the shear
rate but rather remains constant over a broad and practical range of shear rates
(Macosko 1994; Shaw 1992).
Rheological behavior of protein dispersions is intimately linked to the charac-
teristics of the protein itself as well as environmental factors. The characteristic
properties of the protein that are important in this aspect are its conformation,
shape, size, solubility, swellability, and charge which are influenced by environ-
mental conditions like temperature, concentration, pH, and ionic strength
(Hermansson 1972, 1979).
The viscosity of protein solutions is largely dependent on the (1) hydrodynamic
behavior of the protein molecules in solution and (2) protein–protein interactions
between adjacent molecules and their spatial distribution. Consequently, the study
of the behavior of protein solutions at infinite dilution of the protein when protein–
protein interactions are negligible can provide meaningful insight into the hydrody-
namic properties of the protein molecule which in turn can be related to the shape
and size or conformational changes in the protein. The problem of relating viscosity
of the colloidal dispersions with the fundamental nature of the dispersed particles
has been the subject of much experimental investigations and theoretical consider-
ations (Harding 1980, 1981; Lundqvist 1999). Such an effect is defined in terms of
viscosity functions such as relative (ηrel), specific (ηsp), and reduced or reduced spe-
cific viscosity (ηred):
h
hrel = , (3.23)
h0
3 Techniques for Higher-Order Structure Determination 73
h − h0
hsp = hrel − 1 = , (3.24)
h0
hred = hsp / c = (hrel − 1) / c, (3.25)
where η is the viscosity of the solution of concentration c and η0 is the viscosity of
the solvent. Intrinsic viscosity [η] is an intrinsic function of the dissolved/dispersed
macromolecule and is defined as the limiting value, at zero concentration, of relative
viscosity increment.
Two factors that contribute to this characteristic property of the dispersed particle
are particle shape and size/volume as summarized by the following relationship:
where, ν is the molecular shape factor known as the viscosity increment or universal
shape function (Kauzmann 1959; Yang 1961) and Vs is known as the swollen spe-
cific volume and is a measure of the solvent associated with the macromolecule. It
is defined as the volume of the macromolecule in solution per unit of anhydrous
mass of the macromolecule. Intrinsic viscosity is thus indirectly dependent on the
axial ratio of the molecule that governs its shape. It increases with increased axial
ratio and is therefore minimum for a sphere (Tanford and Buzzell 1956; Harding
1997). As a result of this dependence, intrinsic viscosity has been used as a com-
bined measure of the hydrodynamic volume and molecular shape.
Changes in intrinsic viscosity or in reduced viscosity at low concentration have
also been used as a measure of the extent of denaturation, which also reflects the
relationship with changing axial ratio (Suryaprakash and Prakash 2000; Ahmad and
Salahuddin 1976). Interestingly, for globular proteins unlike random coils, intrinsic
viscosity has been found to be independent of the molecular weight and depends
solely on the shape and size of the protein. In an excellent review, Harding (Harding
1997) has listed the values of intrinsic viscosity of numerous proteins and peptides.
A relatively simple method has been proposed in the literature for calculating the
axial ratio of proteins based on viscosity measurements and data fitting (Monkos
and Turczynski 1991). Rheological measurements have also been used to determine
the unfolding onset for protein in solution in the presence of an increasing concen-
tration of a denaturant like urea (Tu and Breedveld 2005).
A general application of intrinsic viscosity measurements has been to assess the
denaturation or conformational change of a protein molecule in solution. For a poly-
peptide poly-(his-ala-glu) studied in the pH range of 2.97–9.70, the intrinsic viscos-
ity varied from 8 mL/g at pH 2.97 to 33 mL/g at pH 4.99 to 55 mL/g at pH 9.70
indicating an increasing unfolding or conformational change with increase in solu-
tion pH. This is consistent with a general increase in the net negative charge on the
molecule due to the ionization of the glutamic acid residues and the deprotonation
of histidine residues with increasing solution pH. From the example it is apparent
that solution pH and ionic strength have a marked effect on the hydrodynamic
74 J. Kranz et al.
p roperties of the protein and therefore on the flow behavior of the system on the
whole. This effect is mediated through the effect on charge associated with the pro-
tein molecule, its folding and unfolding kinetics, and the solubility of the protein
that changes with the pH. In addition to the absolute shape and size effects, electro-
static charge can significantly influence the intrinsic viscosity of a protein molecule
in solution, especially if the molecule carries multiple charges in solution
(polyelectrolyte) like proteins. In theory three such contributions have been sum-
marized for a compact globular protein molecule: a “primary effect” due to the dif-
fuse double layer surrounding the protein molecule, a “secondary effect” due to
repulsion between the double layers of different molecules, and a “tertiary effect”
arising from the intermolecular repulsions that affect the shape of the molecule.
These three have been collectively referred to as “electroviscous” effects. Under
conditions of high net charge on the protein molecule, the electroviscous effect
contributes significantly to solution flow behavior. However, under conditions of
low molecular charge and/or high charge screening, as is the case with high ionic
strength solutions, electroviscous effects are not very prominent.
During the processing and delivery of protein therapeutics, there are multiple
steps that can be limited by viscosity, including filtration, chromatography, and
delivery through syringes. Thus, viscosity measurements can be important during
formulation and candidate screening and during process development.
3.4 Conclusions
Acknowledgements The authors would like to thank Dr. Lee Olszewski, Dr. Doug Nesta,
Dr. Aston Liu, and Dr. Joseph Rinella at GlaxoSmithKline Biopharm Development for supporting
this publication. We thank Dr. Rina K. Dukor from BioTools Inc. and Prof. Laurence A. Nafie from
Syracuse University for their insights and helpful discussion.
3 Techniques for Higher-Order Structure Determination 75
References
Adams PD, Chen Y, Ma K, Zagorski MG, Sonnichsen FD, McLaughlin ML, Barkley MD (2002)
Intramolecular quenching of tryptophan fluorescence by the peptide bond in cyclic hexapep-
tides. J Am Chem Soc 124:9278–9286
Ahmad F, Salahuddin A (1976) Reversible unfolding of the major fraction of ovalbumin by guani-
dine hydrochloride. Biochemistry 15(23):5168–5175
Al-Lazikani B, Lesk AM, Chothia C (1997) Standard conformations for the canonical structures
of immunoglobulins. J Mol Biol 273:927–948
Asher SA, Bormett RW, Chen XG, Lemmon DH, Cho N (1993) UV resonance Raman spectros-
copy using a new cw laser source: convenience and experimental simplicity. Appl Spectrosc
47:628–633
Bai Y, Sosnick TR, Mayne L, Englander SW (1995) Protein folding intermediates: native-state
hydrogen exchange. Science 269:192–197
Baldwin AJ, Kay LE (2009) NMR spectroscopy brings invisible protein states into focus. Nat
Chem Biol 5:808–814
Bandekar J (1992) Amide modes and protein conformation. Biochim Biophys Acta 1120:123–143
Barron LD (2006) Structure and behaviour of biomolecules from Raman optical activity. Curr
Opin Struct Biol 16:638–643
Barron LD, Hecht L, Bell AF, Wilson G (1996) Recent developments in Raman optical activity of
biopolymers. Appl Spectrosc 50:619–629
Barron LD, Hecht L, Blanch EW, Bell AF (2000) Solution structure and dynamics of biomolecules
from Raman optical activity. Prog Biophys Mol Biol 73:1–49
Bartlett AI, Radford SE (2009) An expanding arsenal of experimental methods yields an explosion
of insights into protein folding mechanisms. Nat Struct Mol Biol 16:582–588
Baumruk V, Keiderling TA (1993) Vibrational circular dichroism of proteins in H2O solution. J Am
Chem Soc 115:6939–6942
Benevides JM, Overman SA, Thomas GJ Jr (2004) Raman spectroscopy of proteins. Curr Protoc
Protein Sci, Chapter 17, Unit 17.18
Billeter M, Vendrell J, Wider G, Aviles FX, Coll M, Guasch A, Huber R, Wuthrich K (1992)
Comparison of the NMR solution structure with the X-ray crystal structure of the activation
domain from procarboxypeptidase B. J Biomol NMR 2:1–10
Billeter M, Wagner G, Wuthrich K (2008) Solution NMR structure determination of proteins revis-
ited. J Biomol NMR 42:155–158
Boehr DD, Dyson HJ, Wright PE (2006) An NMR perspective on enzyme dynamics. Chem Rev
106:3055–3079
Brandts JF, Lin LN (1990) Study of strong to ultratight protein interactions using differential scan-
ning calorimetry. Biochemistry 29:6927–6940
Brennan JF, Wang Y, Ramachandra RD, Feld MS (1997) Near-infrared Raman spectrometer sys-
tems for human studies. Appl Spectrosc 51:201–208
Byler DM, Susi H (1986) Examination of the secondary structure of proteins by deconvolved FTIR
spectra. Biopolymers 25:469–487
Cantor RC, Schimmel PR (1980) Biophysical Chemistry: techniques for the study of biological
structure and function, vol 2. W. H. Freeman, San Francisco
Cao XL, Dukor RK, Nafie LA (2008) Reduction of linear birefringence in vibrational circular
dichroism measurement: use of a rotating half-wave plate. Theor Chem Acc 119:69–79
Carpenter JF, Prestrelski SJ, Dong A (1998) Application of infrared spectroscopy to development
of stable lyophilized protein formulations. Eur J Pharm Biopharm 45:231–238
Chalmers J, Griffiths P (eds) (2002) Handbook of vibrational spectroscopy. Wiley, Chichester
Chance B (1953) The carbon monoxide compounds of the cytochrome oxidases. I. Difference
spectra. J Biol Chem 202:383–396
Chance B, Pappenheimer AM Jr (1954) Kinetic and spectrophotometric studies of cytochrome b5
in midgut homogenates of cecropia. J Biol Chem 209:931–943
76 J. Kranz et al.
Chill JH, Naider F (2011) A solution NMR view of protein dynamics in the biological membrane.
Curr Opin Struct Biol 21:627–633
Chothia C, Lesk AM, Tramontano A, Levitt M, Smith-Gill SJ, Air G, Sheriff S, Padlan EA, Davies
D, Tulip WR et al (1989) Conformations of immunoglobulin hypervariable regions. Nature
342:877–883
Cimmperman P, Baranauskiene L, Jachimoviciute S, Jachno J, Torresan J, Michailoviene V,
Matuliene J, Sereikaite J, Bumelis V, Matulis D (2008) A quantitative model of thermal stabi-
lization and destabilization of proteins by ligands. Biophys J 95:3222–3231
Daniel E, Weber G (1966) Cooperative effects in binding by bovine serum albumin. I. The binding
of 1-anilino-8-naphthalenesulfonate. Fluorimetric titrations. Biochemistry 5:1893–1900
D'Antonio J, Murphy BM, Manning MC, Al-Azzam WA (2012) Comparability of protein thera-
peutics: quantitative comparison of second-derivative amide I infrared spectra. J Pharm Sci
101:2025–2033
Davies DR, Cohen GH (1996) Interactions of protein antigens with antibodies. Proc Natl Acad Sci
USA 93:7–12
Dong A, Huang P, Caughey WS (1990) Protein secondary structures in water from second-
derivative amide I infrared spectra. Biochemistry 29:3303–3308
Dong J, Dinakarpandian D, Carey PR (1998) Extending the Raman analysis of biological samples
to the 100 micromolar concentration range. Appl Spectrosc 52:1117–1122
dos Remedios CG, Moens PD (1995) Fluorescence resonance energy transfer spectroscopy is a
reliable “ruler” for measuring structural changes in proteins. Dispelling the problem of the
unknown orientation factor. J Struct Biol 115:175–185
Dousseau F, Pezolet M (1990) Determination of the secondary structure content of proteins in
aqueous solutions from their amide I and amide II infrared bands. Comparison between classi-
cal and partial least-squares methods. Biochemistry 29:8771–8779
Dukor RK, Keiderling TA (1991) Reassessment of the random coil conformation: vibrational CD
study of proline oligopeptides and related polypeptides. Biopolymers 31:1747–1761
Eberstein W, Georgalis Y, Saenger W (1994) Molecular interactions in crystallizing lysozyme
solutions studied by photon correlation spectroscopy. J Cryst Growth 143:71–78
Eisenmesser EZ, Millet O, Labeikovsky W, Korzhnev DM, Wolf-Watz M, Bosco DA, Skalicky JJ,
Kay LE, Kern D (2005) Intrinsic dynamics of an enzyme underlies catalysis. Nature
438:117–121
Englander SW (2000) Protein folding intermediates and pathways studied by hydrogen exchange.
Annu Rev Biophys Biomol Struct 29:213–238
Englander SW (2006) Hydrogen exchange and mass spectrometry: a historical perspective. J Am
Soc Mass Spectrom 17:1481–1489
Englander SW, Mayne L, Bai Y, Sosnick TR (1997) Hydrogen exchange: the modern legacy of
Linderstrom-Lang. Protein Sci 6:1101–1109
Epstein HF, Schechter AN, Chen RF, Anfinsen CB (1971) Folding of staphylococcal nuclease:
kinetic studies of two processes in acid renaturation. J Mol Biol 60:499–508
Ferreon AC, Moran CR, Gambin Y, Deniz AA (2011) Single-molecule fluorescence studies of
intrinsically disordered proteins. Methods Enzymol 472:179–204
Flanagan JM, Kataoka M, Shortle D, Engelman DM (1992) Truncated staphylococcal nuclease is
compact but disordered. Proc Natl Acad Sci USA 89:748–752
Foldes-Papp Z, Demel U, Domej W, Tilz GP (2002) A new dimension for the development of
fluorescence-based assays in solution: from physical principles of FCS detection to biological
applications. Exp Biol Med (Maywood) 227:291–300
Fu K, Griebenow K, Hsieh L, Klibanov AM, Langer R (1999) FTIR characterization of the sec-
ondary structure of proteins encapsulated within PLGA microspheres. J Control Release
58:357–366
Goldberg DS, Bishop SM, Shah AU, Sathish HA (2010) Formulation development of therapeutic
monoclonal antibodies using high-throughput fluorescence and static light scattering tech-
niques: role of conformational and colloidal stability. J Pharm Sci
3 Techniques for Higher-Order Structure Determination 77
Goldberg DS, Bishop SM, Shah AU, Sathish HA (2011) Formulation development of therapeutic
monoclonal antibodies using high-throughput fluorescence and static light scattering tech-
niques: role of conformational and colloidal stability. J Pharm Sci 100(4):1306–1315
Goldstein SR, Kidder LH, Herne TM, Levin IW, Lewis EN (1996) The design and implementation
of a high-fidelity Raman imaging microscope. J Microsc 184:35–45
Grauw CJ, Otto C, Greve J (1997) Linescan Raman microspectrometry for biological applications.
Appl Spectrosc 51:1607–1612
Greenfield NJ (1996) Methods to estimate the conformation of proteins and polypeptides from
circular dichroism data. Anal Biochem 235:1–10
Greenfield NJ (2006a) Using circular dichroism spectra to estimate protein secondary structure.
Nat Protoc 1(6):2876–2890
Greenfield NJ (2006b) Determination of the folding of proteins as a function of denaturants, osmo-
lytes or ligands using circular dichroism. Nat Protoc 1:2733–2741
Greenfield NJ (2006c) Analysis of the kinetics of folding of proteins and peptides using circular
dichroism. Nat Protoc 1:2891–2899
Greenfield NJ (2007) Using circular dichroism collected as a function of temperature to determine
the thermodynamics of protein unfolding and binding interactions. Nat Protoc 1:2527–2535
Greenfield N, Fasman GD (1969) Computed circular dichroism spectra for the evaluation of pro-
tein conformation. Biochemistry 8:4108–4116
Griebenow K, Klibanov AM (1995) Lyophilization-induced reversible changes in the secondary
structure of proteins. Proc Natl Acad Sci USA 92:10969–10976
Griffiths PR, de Haseth JA (1986) Fourier transform infrared spectroscopy. Wiley Interscience,
New York
Hamuro Y, Weber PC, Griffin PR (2005) High-throughput analysis of protein structure by hydro-
gen/deuterium exchange mass spectrometry. Methods Biochem Anal 45:131–157
Harding SE (1980) The combination of the viscosity increment with the harmonic mean rotational
relaxation time for determining the conformation of biological macromolecules in solution.
Biochem J 189(2):359–361
Harding SE (1981) A compound hydrodynamic shape function derived from viscosity and molecu-
lar covolume measurements. Int J Biol Macromol 3(5):340–341
Harding SE (1997) The intrinsic viscosity of biological macromolecules. Progress in measure-
ment, interpretation and application to structure in dilute solution. Prog Biophys Mol Biol
68(2–3):207–262
Haris PI, Chapman D (1995) The conformational analysis of peptides using Fourier transform IR
spectroscopy. Biopolymers 37:251–263
Haris PI, Lee DC, Chapman D (1986) A Fourier transform infrared investigation of the structural
differences between ribonuclease A and ribonuclease S. Biochim Biophys Acta 874:255–265
Haris PI, Chapman D, Harrison RA, Smith KF, Perkins SJ (1990) Conformational transition
between native and reactive center cleaved forms of alpha 1-antitrypsin by Fourier transform
infrared spectroscopy and small-angle neutron scattering. Biochemistry 29:1377–1380
Hashimoto S, Ikeda T, Takeuuchi H, Harada I (1993) Utilization of a prism monochromator as a
sharp-cut bandpass filter in ultraviolet Raman spectroscopy. Appl Spectrosc 47:1283–1285
Hawe A, Sutter M, Jiskoot W (2008) Extrinsic fluorescent dyes as tools for protein characteriza-
tion. Pharm Res 25:1487–1499
Hawe A, Rispens T, Herron JN, Jiskoot W (2011) Probing bis-ANS binding sites of different affin-
ity on aggregated IgG by steady-state fluorescence, time-resolved fluorescence and isothermal
titration calorimetry. J Pharm Sci 100:1294–1305
He F, Hogan S, Latypov RF, Narhi LO, Razinkov VI (2009a) High throughput thermostability
screening of monoclonal antibody formulations. J Pharm Sci 99:1707–1720
He F, Phan DH, Hogan S, Bailey R, Becker GW, Narhi LO, Razinkov VI (2009b) Detection of IgG
aggregation by a high throughput method based on extrinsic fluorescence. J Pharm Sci
99:2598–2608
Hermansson AM (1972) Functional properties of proteins for foods. Lebensin-Wiss Technol 5:24
78 J. Kranz et al.
Hermansson AM (1979) Aspects of protein structure, rheology and texturization. Food Texture
Rheol [Proc Symp] 265–282
Herr AB, Ballister ER, Bjorkman PJ (2003) Insights into IgA-mediated immune responses from
the crystal structures of human FcalphaRI and its complex with IgA1-Fc. Nature
423:614–620
van Holde KE, Johnson WC, Ho (2006) Methods for the Separation and Characterization of
Macromolecules. In: van Holde KE, Johnson WC, Ho PS. Principles of Physical Biochemistry.
2nd ed. Upper Saddle River, NJ: Pearson Prentice Hall:213–275
Hynes TR, Fox RO (1991) The crystal structure of staphylococcal nuclease refined at 1.7 A resolu-
tion. Proteins 10:92–105
Jackson M, Mantsch HH (1995) The use and misuse of FTIR spectroscopy in the determination of
protein structure. Crit Rev Biochem Mol Biol 30:95–120
Jares-Erijman EA, Jovin TM (2003) FRET imaging. Nat Biotechnol 21:1387–1395
Jelesarov I, Bosshard HR (1999) Isothermal titration calorimetry and differential scanning calo-
rimetry as complementary tools to investigate the energetics of biomolecular recognition.
J Mol Recognit 12:3–18
Kamerzell TJ, Esfandiary R, Joshi SB, Middaugh CR, Volkin DB (2011) Protein-excipient interac-
tions: mechanisms and biophysical characterization applied to protein formulation develop-
ment. Adv Drug Deliv Rev 63:1118–1159
Kauzmann W (1959) Some factors in the interpretation of protein denaturation. Adv Protein Chem
14:1–63
Keiderling TA, Silva RA, Yoder G, Dukor RK (1999) Vibrational circular dichroism spectroscopy
of selected oligopeptide conformations. Bioorg Med Chem 7:133–141
Keiderling TA, Kubelka J, Hilario J (eds) (2006) Vibrational circular dichroism of biopolymers:
summary of methods and applications. CRC, Boca Raton
Kielec JM, Valentine KG, Wand AJ (2009) A method for solution NMR structural studies of large
integral membrane proteins: reverse micelle encapsulation. Biochim Biophys Acta
1798:150–160
Kranz JK, Schalk-Hihi C (2011) Protein thermal shifts to identify low molecular weight fragments.
Methods Enzymol 493:277–298
Krimm S, Bandekar J (1986) Vibrational spectroscopy and conformation of peptides, polypep-
tides, and proteins. Adv Protein Chem 38:181–364
Krishinan KS, Brandts JF (1978) Scanning calorimetry. Methods Enzymol 49:3–14
Kubelka J, Hofrichter J, Eaton WA (2004) The protein folding “speed limit”. Curr Opin Struct Biol
14:76–88
Kurouski D, Lombardi RA, Dukor RK, Lednev IK, Nafie LA (2010) Direct observation and pH
control of reversed supramolecular chirality in insulin fibrils by vibrational circular dichroism.
Chem Commun (Camb) 46:7154–7156
Kurouski D, Dukor RK, Lu X, Nafie LA, Lednev IK (2012) Spontaneous inter-conversion of insu-
lin fibril chirality. Chem Commun (Camb) 48:2837–2839
Ladbury JE, Chowdhry BZ (1996) Sensing the heat: the application of isothermal titration calorim-
etry to thermodynamic studies of biomolecular interactions. Chem Biol 3:791–801
Lakhani A, Malon P, Keiderling TA (2009) Comparison of vibrational circular dichroism instru-
ments: development of a new dispersive VCD. Appl Spectrosc 63:775–785
Lakowicz JR (2006) Principles of fluorescence spectroscopy, 3rd edn. Springer, New York
LeTilly V, Royer CA (1993) Fluorescence anisotropy assays implicate protein–protein interactions
in regulating trp repressor DNA binding. Biochemistry 32:7753–7758
Li CH, Li T (2009) Application of vibrational spectroscopy to the structural characterization of
monoclonal antibody and its aggregate. Curr Pharm Biotechnol 10:391–399
Li Y, Williams TD, Topp EM (2008) Effects of excipients on protein conformation in lyophilized
solids by hydrogen/deuterium exchange mass spectrometry. Pharm Res 25:259–267
Li Y, Mach H, Blue JT (2011) High throughput formulation screening for global aggregation
behaviors of three monoclonal antibodies. J Pharm Sci 100:2120–2135
3 Techniques for Higher-Order Structure Determination 79
Liu J, Andya JD, Shire SJ (2006) A critical review of analytical ultracentrifugation and field flow
fractionation methods for measuring protein aggregation. AAPS J 8:E580–E589
Loria JP, Berlow RB, Watt ED (2008) Characterization of enzyme motions by solution NMR
relaxation dispersion. Acc Chem Res 41:214–221
Lundqvist R (1999) Molecular weight studies on hydroxypropyl methyl cellulose. Part 2. Intrinsic
viscosity. Int J Polym Anal Charact 5(1):61–84
Ma S, Cao X, Mak M, Sadik A, Walkner C, Freedman TB, Lednev IK, Dukor RK, Nafie LA (2007)
Vibrational circular dichroism shows unusual sensitivity to protein fibril formation and devel-
opment in solution. J Am Chem Soc 129:12364–12365
Machtle W, Borger L (2006) Sedimentation velocity. In: Machtle W, Borger L (eds) Analytical
ultracentrifugation of polymers and nanoparticles. Heidelberg Springer:47–96
Macosko CW (1994) In: Macosko CW (ed) Rheology: principles, measurements and applications.
Wiley-VCH, New York, p 550
Mandel M (1993) Applications of dynamic light scattering to polyelectrolytes in solution. In:
Brown W (ed) Dynamic light scattering: the method and some applications. Oxford University
Press, New York, pp 319–371
Manley G, Loria JP (2012) NMR insights into protein allostery. Arch Biochem Biophys
519:223–231
Mann CJ, Royer CA, Matthews CR (1993) Tryptophan replacements in the trp aporepressor from
Escherichia coli: probing the equilibrium and kinetic folding models. Protein Sci 2:1853–1861
Markley JL, Ulrich EL, Westler WM, Volkman BF (2003) Macromolecular structure determina-
tion by NMR spectroscopy. Methods Biochem Anal 44:89–113
Matulis D, Kranz JK, Salemme FR, Todd MJ (2005) Thermodynamic stability of carbonic anhy-
drase: measurements of binding affinity and stoichiometry using ThermoFluor. Biochemistry
44:5258–5266
Meyer JD, Bai SJ, Rani M, Suryanarayanan R, Nayar R, Carpenter JF, Manning MC (2004) Infrared
spectroscopic studies of protein formulations containing glycine. J Pharm Sci 93:1359–1366
Mezzasalma TM, Kranz JK, Chan W, Struble GT, Schalk-Hihi C, Deckman IC, Springer BA, Todd
MJ (2007) Enhancing recombinant protein quality and yield by protein stability profiling.
J Biomol Screen 12:418–428
Mittermaier AK, Kay LE (2009) Observing biological dynamics at atomic resolution using NMR.
Trends Biochem Sci 34:601–611
Miura T, Takeuchi H, Harada I (1991) Raman spectroscopic characterization of tryptophan side
chains in lysozyme bound to inhibitors: role of the hydrophobic box in the enzymatic function.
Biochemistry 30:6074–6080
Miyawaki A (2011) Proteins on the move: insights gained from fluorescent protein technologies.
Nat Rev Mol Cell Biol 12:656–668
Miyazawa T, Shimanouchi T, Mizushima S (1956) Characteristic infrared bands of monosubsti-
tuted amides. J Chem Phys 24:408
Monkos K, Turczynski B (1991) Determination of the axial ratio of globular proteins in aqueous
solution using viscometric measurements. Int J Biol Macromol 13(6):341–344
Murphy KP, Freire E (1992) Thermodynamics of structural stability and cooperative folding
behavior in proteins. Adv Protein Chem 43:313–361
Nafie LA (1996) Vibrational optical activity. Appl Spectrosc 50:14A–26A
Nafie LA (2011) Vibrational optical activity: principles and applications. Wiley, Chichester
Nafie LA, Dukor RK (eds) (2007) Applications of vibrational optical activity in the pharmaceuti-
cal industry. Wiley, Chichester
Nashine VC, Hammes-Schiffer S, Benkovic SJ (2010) Coupled motions in enzyme catalysis. Curr
Opin Chem Biol 14:644–651
Norvaisas P, Petrauskas V, Matulis D (2012) Thermodynamics of cationic and anionic surfactant
interaction. J Phys Chem B 116:2138–2144
Otto MR, Lillo MP, Beechem JM (1994) Resolution of multiphasic reactions by the combination
of fluorescence total-intensity and anisotropy stopped-flow kinetic experiments. Biophys J
67:2511–2521
80 J. Kranz et al.
Perez C, Castellanos IJ, Costantino HR, Al-Azzam W, Griebenow K (2002) Recent trends in sta-
bilizing protein structure upon encapsulation and release from bioerodible polymers. J Pharm
Pharmacol 54:301–313
Perrin F (1926) J Phys Radium 1:390–401
Philo JS (2006) Is any measurement method optimal for all aggregate sizes and types? AAPS J
8:E564–E571
Philo JS (2005) Analytical ultracentrifugation. In: Borchardt RT, Middaugh CR, series eds.
Biotechnology: pharmaceutical aspects, Jiskoot W, Crommelin D, vol eds. Volume III: Methods
for structural analysis of protein pharmaceuticals. AAPS Press, Arlington, pp 379–412
Pollok BA, Heim R (1999) Using GFP in FRET-based applications. Trends Cell Biol 9:57–60
Putnam WS, Prabhu S, Zheng Y, Subramanyam M, Wang YM (2010) Pharmacokinetic, pharma-
codynamic and immunogenicity comparability assessment strategies for monoclonal antibod-
ies. Trends Biotechnol 28:509–516
Ramsey JD, Gill ML, Kamerzell TJ, Price ES, Joshi SB, Bishop SM, Oliver CN, Middaugh CR
(2009) Using empirical phase diagrams to understand the role of intramolecular dynamics in
immunoglobulin G stability. J Pharm Sci 98:2432–2447
Raso SW, Clark PL, Haase-Pettingell C, King J, Thomas GJ Jr (2001) Distinct cysteine sulfhydryl
environments detected by analysis of Raman S-hh markers of Cys–>Ser mutant proteins. J Mol
Biol 307:899–911
Rossmann MG (2001) Molecular replacement–historical background. Acta Crystallogr D Biol
Crystallogr 57:1360–1366
Royer CA (1993) Understanding fluorescence decay in proteins. Biophys J 65:9–10
Royer CA (1995) Fluorescence spectroscopy. Methods Mol Biol 40:65–89
Royer CA (2006) Probing protein folding and conformational transitions with fluorescence. Chem
Rev 106:1769–1784
Royer CA, Scarlata SF (2008) Fluorescence approaches to quantifying biomolecular interactions.
Methods Enzymol 450:79–106
Sanchez-Ruiz JM (2011) Probing free-energy surfaces with differential scanning calorimetry.
Annu Rev Phys Chem 62:231–255
Sapienza PJ, Lee AL (2010) Using NMR to study fast dynamics in proteins: methods and applica-
tions. Curr Opin Pharmacol 10:723–730
Schuck P, Perugini MA, Gonzales NR, Howlett GJ, Schubert D (2002) Size-distribution analysis
of proteins by analytical ultracentrifugation: strategies and application to model systems.
Biophys J 82:1096–1111
Schümmer P. Mechanics of Non-Newtonian Fluids. Von W. R. Schowalter. Pergamon Press,
Oxford–Frankfurt 1978. 1. Aufl., IX, 300 S., zahlr. Abb. u. Tab., geb., $ 35.00. Chemie
Ingenieur Technik. 1979;51(7):766–766
Schweitzer-Stenner R, Eker F, Huang Q, Griebenow K, Mosz P, Kozlowski P (2002) Structure
analysis of dipeptides in water by exploring and utilizing the structural sensitivity of amide III
by polarized visible Raman, FTIR spectroscopy and DFT based normal coordinate analysis.
J Phys Chem B 106:4294–4304
Scott DJ, Schuck P (2005) A brief introduction to the analytical ultracentrifugation of proteins for
beginners. In: Scott DJ, Harding SE, Rowe AJ (eds) Analytical ultracentrifugation techniques
and methods. RSC Publishing, Cambridge, pp 1–25
Sharp KA, Nicholls A, Friedman R, Honig B (1991) Extracting hydrophobic free energies from
experimental data: relationship to protein folding and theoretical models. Biochemistry
30:9686–9697
Shaw DJ (1992) Rheology. Colloidal and surface chemistry, 4th edn. Butterworth-Heinemann,
Boston
Shen Y, Lange O, Delaglio F, Rossi P, Aramini JM, Liu G, Eletsky A, Wu Y, Singarapu KK, Lemak
A, Ignatchenko A, Arrowsmith CH, Szyperski T, Montelione GT, Baker D, Bax A (2008)
Consistent blind protein structure generation from NMR chemical shift data. Proc Natl Acad
Sci USA 105:4685–4690
3 Techniques for Higher-Order Structure Determination 81
Shi Z, Chen K, Liu Z, Kallenbach NR (2006) Conformation of the backbone in unfolded proteins.
Chem Rev 106:1877–1897
Siamwiza MN, Lord RC, Chen MC, Takamatsu T, Harada I, Matsuura H, Shimanouchi T (1975)
Interpretation of the doublet at 850 and 830 cm−1 in the Raman spectra of tyrosyl residues in
proteins and certain model compounds. Biochemistry 14:4870–4876
Skinner AL, Laurence JS (2008) High-field solution NMR spectroscopy as a tool for assessing
protein interactions with small molecule ligands. J Pharm Sci 97:4670–4695
Skinner AL, Laurence JS (2010) Probing residue–specific interactions in the stabilization of pro-
teins using high-resolution NMR: a study of disulfide bond compensation. J Pharm Sci
99:2643–2654
Skinner JJ, Lim WK, Bedard S, Black BE, Englander SW (2012) Protein dynamics viewed by
hydrogen exchange. Protein Sci 21:996–1005
Smith BC (1996) Fundamentals of fourier transform infrared spectroscopy. CRC, Boca Raton
Socrates G (2004) Infrared and Raman characteristic group frequencies: tables and charts, 3rd edn.
Wiley, Chichester
Sondermann P, Oosthuizen V (2002) X-ray crystallographic studies of IgG-Fc gamma receptor
interactions. Biochem Soc Trans 30:481–486
Spink CH (2008) Differential scanning calorimetry. Methods Cell Biol 84:115–141
Stefani M, Dobson CM (2003) Protein aggregation and aggregate toxicity: new insights into pro-
tein folding, misfolding diseases and biological evolution. J Mol Med (Berl) 81:678–699
Stern KG (1937) Spectroscopy of catalase. J Gen Physiol 20:631–648
Stokes GG (1852) On the change of refrangibility of light. Philos Trans R Soc Lond
142:463–562
Surewicz WK, Mantsch HH (1988) New insight into protein secondary structure from resolution-
enhanced infrared spectra. Biochim Biophys Acta 952:115–130
Suryaprakash P, Prakash V (2000) Unfolding of multimeric proteins in presence of denaturants.
A case study of helianthinin from Helianthus annuus L. Nahrung 44(3):178–183
Susi H, Byler DM (1986) Resolution-enhanced Fourier transform infrared spectroscopy of
enzymes. Methods Enzymol 130:290–311
Susi H, Timasheff SN, Stevens L (1967) Infrared spectra and protein conformations in aqueous
solutions. I. The amide I band in H2O and D2O solutions. J Biol Chem 242:5460–5466
Tamm LK, Lai AL, Li Y (2007) Combined NMR and EPR spectroscopy to determine structures of
viral fusion domains in membranes. Biochim Biophys Acta 1768:3052–3060
Tanford C, Buzzell JG (1956) The viscosity of aqueous solutions of bovine serum albumin between
pH 4.3 and 10.5. J Phys Chem 60:225–231
Tsuboi M, Suzuki M, Overman SA, Thomas GJ Jr (2000) Intensity of the polarized Raman band
at 1340–1345 cm−1 as an indicator of protein alpha-helix orientation: application to Pf1 fila-
mentous virus. Biochemistry 39:2677–2684
Tsutsui Y, Wintrode PL (2007) Hydrogen/deuterium exchange-mass spectrometry: a powerful tool
for probing protein structure, dynamics and interactions. Curr Med Chem 14:2344–2358
Tu RS, Breedveld V (2005) Microrheological detection of protein unfolding. Phys Rev E Stat
Nonlin Soft Matter Phys 72(4 Pt 1):041914
Van Holde KE, Johnson WC, Ho PS (2006) Principles of physical biochemistry. Pearson/Prentice
Hall, Upper Saddle River
Velazquez-Campoy A, Ohtaka H, Nezami A, Muzammil S, Freire E (2004) Isothermal titration
calorimetry. Curr Protoc Cell Biol, Chapter 17, Unit 17.18
Venyaminov S, Kalnin NN (1990a) Quantitative IR spectrophotometry of peptide compounds in
water (H2O) solutions. I. Spectral parameters of amino acid residue absorption bands.
Biopolymers 30:1243–1257
Venyaminov S, Kalnin NN (1990b) Quantitative IR spectrophotometry of peptide compounds in
water (H2O) solutions. II. Amide absorption bands of polypeptides and fibrous proteins in
alpha-, beta-, and random coil conformations. Biopolymers 30:1259–1271
Wand AJ (2001) Dynamic activation of protein function: a view emerging from NMR spectros-
copy. Nat Struct Biol 8:926–931
82 J. Kranz et al.
Wang Y, Purrello R, Jordan T, Spiro TG (1991) UVRR spectroscopy of the peptide bond. J Am
Chem Soc 113:6359–6368
Weber G (1952a) Polarization of the fluorescence of macromolecules. II. Fluorescent conjugates
of ovalbumin and bovine serum albumin. Biochem J 51:155–167
Weber G (1952b) Polarization of the fluorescence of macromolecules. I. Theory and experimental
method. Biochem J 51:145–155
Weber G, Daniel E (1966) Cooperative effects in binding by bovine serum albumin. II. The binding
of 1-anilino-8-naphthalenesulfonate. Polarization of the ligand fluorescence and quenching of
the protein fluorescence. Biochemistry 5:1900–1907
Weber G, Laurence DJ (1954) Fluorescent indicators of adsorption in aqueous solution and on the
solid phase. Biochem J 56:xxxi
Weber PC, Salemme FR (2003) Applications of calorimetric methods to drug discovery and the
study of protein interactions. Curr Opin Struct Biol 13:115–121
Wormald MR, Petrescu AJ, Pao YL, Glithero A, Elliott T, Dwek RA (2002) Conformational stud-
ies of oligosaccharides and glycopeptides: complementarity of NMR, X-ray crystallography,
and molecular modelling. Chem Rev 102:371–386
Wyatt PJ (1993) Light scattering and the absolute characterization of macromolecules. Anal Chim
Acta 272(1):1–40
Yang JT (1961) The viscosity of macromolecules in relation to molecular conformations. Adv
Protein Chem 16:323–401
Yoder G, Polese A, Silva RAGD, Formaggio F, Crisma M, Broxterman QB, Kamphuis J, Toniolo
C, Keiderling TA (1997) Conformational characterization of terminally blocked L-(alpha Me)
Val homopeptides using vibrational and electronic circular dichroism. 3(10)-helical stabiliza-
tion by peptide–peptide interaction. J Am Chem Soc 119:10278–10285
Zhang R, Monsma F (2010) Fluorescence-based thermal shift assays. Curr Opin Drug Discov
Devel 13:389–402
Zhu F, Tranter GE, Isaacs NW, Hecht L, Barron LD (2006) Delineation of protein structure classes
from multivariate analysis of protein Raman optical activity data. J Mol Biol 363:19–26
Ziarek JJ, Peterson FC, Lytle BL, Volkman BF (2011) Binding site identification and structure
determination of protein-ligand complexes by NMR a semiautomated approach. Methods
Enzymol 493:241–275
Zuber G, Prestrelski SJ, Benedek K (1992) Application of Fourier transform infrared spectroscopy
to studies of aqueous protein solutions. Anal Biochem 207:150–156
Zubriene A, Matuliene J, Baranauskiene L, Jachno J, Torresan J, Michailoviene V, Cimmperman
P, Matulis D (2009) Measurement of nanomolar dissociation constants by titration calorimetry
and thermal shift assay – radicicol binding to Hsp90 and ethoxzolamide binding to CAII. Int J
Mol Sci 10:2662–2680
Chapter 4
Biophysical Techniques for Protein Size
Distribution Analysis
4.1 Introduction
Z. Wei
Analytical Development, Novavax, Rockville, MD 20850, USA
e-mail: [email protected]
A. Polozova (*)
Biopharmaceutical Development, Medimmune, 1 Medimmune Way,
Gaithersburg, MD 20878, USA
e-mail: [email protected]
Nanoparticle 0.02–1 μm Pros: unbiased size measurements in heteroge- Filipe et al. (2010)
tracking neous samples
analysis Cons: limited dynamic range (~107–109 particles/
mL)
Light obscuration Light 1–150 μm Pros: compendial test. Counts one particle at a Fujishita et al. (1995), Narhi
obscuration time, does not rely on distribution models et al. (2009)
Cons: sizing of protein particles is not reliable.
Does not work well with high protein
concentrations, turbid and viscous samples
85
(continued)
86
New technology Gas-phase 0.003–0.25 μm Pros: fast, direct measurement, high resolution Pease et al. (2008)
electropho- Cons: limited concentration range, electrospray
retic buffer different from formulation buffer
mobility
Resonant mass >0.1 μm Pros: measures single particle at a time, direct Burg et al. (2007)
measure- mass measurement, can distinguish different
ment types based on density
Cons: low sensitivity to particles with density
close to that of surrounding solution
Biophysical Techniques for Protein Size Distribution Analysis
87
88 Z. Wei and A. Polozova
Multiple static and dynamic light scattering techniques are widely used for protein
size heterogeneity analysis, including sizing and quantitation of aggregates (Follmer
et al. 2004; Philo 2006, 2009), distribution of glycosylation (Wen et al. 1996), and
reversible self-association (Wu et al. 1997; Attri and Minton 2005). In methods based
on static or classical light scattering, the molecular weight of molecules in solution is
derived from the light scattering theory equations. These equations link scattered
light intensity and concentration to the molecular weight. Multiangle light scattering
(MALS) applications are most commonly used for proteins in solution (Oliva et al.
2004). In a typical MALS measurement, light scattering intensity measured at three
or more angles is determined and used to calculate the molecular weight. For mole-
cules larger than 10 nm, the radius of gyration can be also calculated based on the
dependance of the light scattering intensity on the scattering angle.
Dynamic light scattering (DLS) relies on the analysis of an autocorrelation func-
tion derived from measurements of scattered light intensity fluctuations due to
Brownian motion of molecules in solution (Jossang et al. 1988). The translational
diffusion coefficient Dτ of protein molecules in the samples is derived from the
90 Z. Wei and A. Polozova
Light obscuration techniques (Fujishita et al. 1995; Narhi et al. 2009) can be applied
to the enumeration and size determination of large aggregates and particles (microm-
eter size). Measurements are based on blockage of laser light by individual particles
passing the detector in the flow cell. The change in light intensity reaching the
4 Biophysical Techniques for Protein Size Distribution Analysis 91
detector is converted to the particle size based on calibration with known particle
standards. Light obscuration is often used as a routine monitoring test for injectable
drugs to detect levels of subvisible particles (≥10 μm). Counting of a single particle
at a given time is an important advantage of this method, because it offers a direct
measurement of particle counts that does not require distribution models. However,
high-concentration protein solutions are not always compatible with this method.
For example, the presence of large amounts of background protein can reduce
contrast between the particles and the solution and interfere with measurements. In
addition, sizing of protein particles by light obscuration methods is often not reliable
because they have different optical properties and block different amounts of light
compared to commonly used polystyrene bead calibration standards. Introduction
of calibration standards similar in optical properties to large protein aggregates
would improve the sizing accuracy; however, such standards are still under
development and not yet commercially available.
Electrical sensing zone (ELS) analysis can be applied to large aggregates ranging in
size from 0.4 μm up to 1,200 μm (Narhi et al. 2009). Particle volume is a primary
parameter measured by this method. The sample is suspended in a weak electrolyte
solution. When large aggregates and particles pass through a small orifice with
applied voltage, they displace a volume of the electrolyte solution equivalent to their
volume. This results in a decrease in electric current, registered as a pulse. The size
information is derived from the magnitude of the pulse, which is directly proportional
to the particle volume. In addition to the size information, the ELS technique also
provides information on the particle counts for a given sample volume. The
requirements for an electrolyte solution and the use of multiple apertures to analyze
the entire aggregate and particle distribution may limit the applicability of this
technique.
4.2.7 Microscopy
Electron microscopy (EM) and atomic force microscopy (AFM) offer a high-
resolution capability for the characterization of conformational heterogeneity and
aggregates at the molecular level (Roux 1999; Thomson 2005). In electron
microscopy methods, the image is formed by a beam of electrons either passing
through (transmission EM) or reflecting off the specimen (scanning EM), while sets
of magnets serve as condenser and objective lenses. Because an electron beam is
used for imaging, EM offers very good spatial resolution. The theoretical resolution
limit is less than 10 Å; however, in practical applications with biological samples,
92 Z. Wei and A. Polozova
the resolution limit is ~1 nm, which still allows a detailed analysis of structure
(Roux 1999). The most promising high-resolution tool for detailed analysis of con-
formational heterogeneity and aggregate structure is cryo-transmission EM tomog-
raphy (Sandin et al. 2004; Banyay et al. 2004). A small droplet of aqueous protein
solution is flash-frozen to form a vitreous ice and directly analyzed in the TEM
without further sample manipulation or preparation. The three-dimensional images
reconstructed by computer software from a set of tilted-plane images provide
detailed information about aggregate order and structure. However, this technique is
still evolving, and, at this time, it is not routinely applied to the analysis of hetero-
geneity in protein solutions.
In AFM, the image is obtained by scanning the surface, usually in a raster pattern,
with very thin tip attached to a cantilever. The spatial resolution and sensitivity of
AFM depend on the mechanical properties of the cantilever and the sharpness of the
tip. Usually, 1 nm or better resolution can be achieved (Engel and Muller 2000).
AFM can provide valuable information on the structure of individual protein
molecules (Thomson 2005), as well as aggregates (Lee et al. 2011). However,
results should be interpreted with caution because this method might be prone to
sample preparation artifacts arising from adsorption on the support surfaces and
sample-tip interactions (Fechner et al. 2009).
Optical microscopy is one of the oldest techniques available for analysis of very
large aggregates, and it can be a useful screening tool for an initial assessment.
Resolution of optical microscopy is defined by the diffraction limit, which is
~0.5 μm for visual light. Large aggregates can be analyzed directly in solution;
however, some translucent particles might be difficult to see due to their low con-
trast with the surrounding solution. Alternatively, large aggregates can be captured
on a filter and detected under oblique light illumination (Barber et al. 1989),
although some thin and fragile particles might be lost during the sample preparation
process.
Over the past several years, flow microscopy has become an increasingly popular
tool for the analysis of large micron-size aggregates in protein solutions (Narhi et al.
2009; Sharma et al. 2007; Huang et al. 2009). This technique offers the combined
advantages of a particle counter and an optical microscope integrated into one
instrument (Sharma et al. 2007; Wuchner et al. 2010). The sample solution flows
through a narrow flow cell, typically 100–300 μm thick. Images magnified by a
10–20× microscope objective are captured by a high-speed camera operating at
8–10 frames per second and automatically analyzed by the image processing
software for presence of particles. Images of all detected particles in the sample are
collected in a database, and information on counts, particle size, transparency, and
morphology can be generated by the software. Large aggregates (particles) in the
2–300 μm range can be detected and analyzed. The ability to provide information
4 Biophysical Techniques for Protein Size Distribution Analysis 93
Mass spectrometry (MS) separates analytes based on their mass and charge (m/z
ratio). Electrospray ionization (ESI) and matrix-assisted laser desorption ionization
(MALDI) techniques are commonly used for the analysis of proteins. MS offers
high sensitivity and good mass resolution and has been a powerful tool used in a
large number of studies for determining protein molecular weights and investigating
protein complexes (Bich and Zenobi 2009). Although MS has been recently reported
to be a useful online detection tool for HPSEC analysis of protein aggregates
(Kukrer et al. 2010), there have been very limited applications of this technique to
study protein aggregation for several reasons (Wang et al. 2011). Firstly, the poten-
tial dissociation of non-covalently linked aggregates during ionization or transfer
into the gas phase could limit the applicability of MS-based techniques for size
distribution analysis. Secondly, most MS mass analyzers have upper-range limita-
tions in analysis of high molecular weight species. A time-of-flight (TOF) mass
analyzer provides a relatively wide m/z range and is often used for protein aggregate
analysis (Krutchinsky et al. 2000). Specialized detectors can be used to make the
m/z range of MALDI MS suitable for analysis of larger proteins and complexes
(Bich and Zenobi 2009; Bahr et al. 1996; Wenzel et al. 2005). Lastly, MS is usually
limited to a qualitative analysis of small oligomeric protein aggregates. However, a
recent study showed that MS was capable of probing the early stage of protein
aggregation in a semiquantitative fashion (Wang et al. 2011).
MS is a powerful tool to determine the structures, sequences, and sites of
N-linked and O-linked glycosylation (North et al. 2009). MS is widely used to
determine the molecular weights of protein fragments and is also capable of
identification of cleavage sites of fragments based on the accurate molecular mass
and known posttranslational modifications of fragments (Liu et al. 2008). Therefore,
MS provides more detailed structural information than other techniques for
glycoprotein size analysis and protein fragment analysis.
(Pease et al. 2008). Size distribution of protein species in the range of 3–250 nm can
be obtained, including distinct peaks for IgG monomers to pentamers. It provides a
fast and quantitative readout of the aggregate distribution. The limitations of the
method are that it works within a limited protein concentration range and the
electrospray step cannot be performed in buffer containing nonvolatile salts.
Additionally, the electrospray buffer may affect the aggregation state of proteins.
4.2.11 Discussion
Therapeutic protein products can have a very broad heterogeneous size distribution,
making size distribution analysis a challenging task. Because no single analytical
method is capable of covering the entire size range of all types of size variants, it is
critical to apply complementary tools to monitor and characterize the entire size dis-
tribution. Depending on the specific requirements for a product, different information,
such as molecular weight, distribution of different species, size, counts, or morphol-
ogy may be needed. Multiple biophysical methods presented in this chapter are capa-
ble of providing this information. Each method may provide a different type of
information depending on a measurement principle, and correlation between orthogo-
nal methods should be thoroughly assessed and understood.
As each analytical method has its own strengths and weaknesses, the choice of
analytical method(s) should be tailored to the sample type and to the intended
purpose. To ensure robust and meaningful size distribution measurements and to
avoid false conclusions, care must be taken in method selection and result
interpretation. Regardless of the method(s) chosen, the use of appropriate controls
is critical because sample preparation and handling may have a significant impact
on the results. Overall, a good knowledge of factors which may influence size
heterogeneity during sample handling and a thorough understanding of physical
principles of the chosen analytical techniques are keys for reliable measurements of
size distribution in protein solutions.
4 Biophysical Techniques for Protein Size Distribution Analysis 95
References
Attri AK, Minton AP (2005) Composition gradient static light scattering: a new technique for
rapid detection and quantitative characterization of reversible macromolecular hetero-
associations in solution. Anal Biochem 346:132–138
Bahr U, Rohling U, Lautz C, Strupat K, Schurenberg M, Hillenkamp F (1996) A charge detector
for time-of-flight mass analysis of high mass ions produced by matrix-assisted laser desorption/
ionization (MALDI). J Mass Spectrom Ion Process 153:9–21
Banyay M, Gilstring F, Hauzenberger E, Ofverstedt LG, Eriksson AB, Krupp JJ, Larsson O (2004)
Three-dimensional imaging of in situ specimens with low-dose electron tomography to analyze
protein conformation. Assay Drug Dev Technol 2:561–567
Barber TA, Lannis MD, Williams JG (1989) Method evaluation: automated microscopy as a com-
pendial test for particulates in parenteral solutions. J Parenter Sci Technol 43:27–47
Bich C, Zenobi R (2009) Mass spectrometry of large complexes. Curr Opin Struct Biol
19:632–639
Burg TP, Godin M, Knudsen SM, Shen W, Carlson G, Foster JS, Babcock K, Manalis SR (2007)
Weighing of biomolecules, single cells and single nanoparticles in fluid. Nature
446:1066–1069
Cao S, Pollastrini J, Jiang Y (2009a) Separation and characterization of protein aggregates and
particles by field flow fractionation. Curr Pharm Biotechnol 10:382–390
Cao S, Jiao N, Jiang Y, Mire-Sluis A, Narhi LO (2009b) Sub-visible particle quantitation in protein
therapeutics. Pharmeur Bio Sci Notes 2009:73–79
Cromwell ME, Hilario E, Jacobson F (2006a) Protein aggregation and bioprocessing. AAPS
J 8:E572–E579
Cromwell MEM, Felten C, Flores H, Liu J, Shire SJ (2006b) Self-association of therapeutic
proteins: implications for product development. In: Murphy RM, Tsai AM (eds) Misbehaving
proteins: protein (Mis)folding, aggregation, and stability. Springer, New York
Engel A, Muller DJ (2000) Observing single biomolecules at work with the atomic force micro-
scope. Nat Struct Biol 7:715–718
Fechner P, Boudier T, Mangenot S, Jaroslawski S, Sturgis JN, Scheuring S (2009) Structural infor-
mation, resolution, and noise in high-resolution atomic force microscopy topographs. Biophys
J 96:3822–3831
Filipe V, Hawe A, Jiskoot W (2010) Critical evaluation of Nanoparticle Tracking Analysis (NTA)
by NanoSight for the measurement of nanoparticles and protein aggregates. Pharm Res
27:796–810
Follmer C, Pereira FV, Da Silveira NP, Carlini CR (2004) Jack bean urease (EC 3.5.1.5) aggrega-
tion monitored by dynamic and static light scattering. Biophys Chem 111:79–87
Fujishita O, Sendo T, Hisazumi A, Otsubo K, Aoyama T, Oishi R (1995) The evaluation of sizing
accuracy of particle counters for parenteral drugs. PDA J Pharm Sci Technol 49:267–271
Gabrielson JP, Brader ML, Pekar AH, Mathis KB, Winter G, Carpenter JF, Randolph TW (2007a)
Quantitation of aggregate levels in a recombinant humanized monoclonal antibody formulation
by size-exclusion chromatography, asymmetrical flow field flow fractionation, and
sedimentation velocity. J Pharm Sci 96:268–279
Gabrielson JP, Randolph TW, Kendrick BS, Stoner MR (2007b) Sedimentation velocity analytical
ultracentrifugation and SEDFIT/c(s): limits of quantitation for a monoclonal antibody system.
Anal Biochem 361:24–30
Huang CT, Sharma D, Oma P, Krishnamurthy R (2009) Quantitation of protein particles in paren-
teral solutions using micro-flow imaging. J Pharm Sci 98:3058–3071
Jossang T, Feder J, Rosenqvist E (1988) Photon correlation spectroscopy of human IgG. J Protein
Chem 7:165–171
Krutchinsky AN, Ayed A, Donald LJ, Ens W, Duckworth HW, Standing KG (2000) Studies of
noncovalent complexes in an electrospray ionization/time-of-flight mass spectrometer. In:
Chapman J (ed) Mass spectrometry of proteins and peptides. Humana Press, Totowa
96 Z. Wei and A. Polozova
Kukrer B, Filipe V, Duijn EV, Kasper PT, Vreeken RJ, Heck AJR, Jiskoot W (2010) Mass spectro-
metric analysis of intact human monoclonal antibody aggregates fractionated by size-exclusion
chromatography. Pharm Res 27:2197–2204
Lee H, Kirchmeier M, Mach H (2011) Monoclonal antibody aggregation intermediates visualized
by atomic force microscopy. J Pharm Sci 100:416–423
Litzen A, Walter JK, Krischollek H, Wahlund KG (1993) Separation and quantitation of monoclo-
nal antibody aggregates by asymmetrical flow field-flow fractionation and comparison to gel
permeation chromatography. Anal Biochem 212:469–480
Liu J, Andya JD, Shire SJ (2006) A critical review of analytical ultracentrifugation and field flow
fractionation methods for measuring protein aggregation. AAPS J 8:E580–E589
Liu H, Gaza-Bulseco G, Lundell E (2008) Assessment of antibody fragmentation by reversed-
phase liquid chromatography and mass spectrometry. J Chromatogr B Analyt Technol Biomed
Life Sci 876:13–23
Mahler HC, Friess W, Grauschopf U, Kiese S (2009) Protein aggregation: pathways, induction
factors and analysis. J Pharm Sci 98:2909–2934
Manta B, Obal G, Ricciardi A, Pritsch O, Denicola A (2011) Tools to evaluate the conformation of
protein products. Biotechnol J 6:731–741
Moore JM, Patapoff TW, Cromwell ME (1999) Kinetics and thermodynamics of dimer formation
and dissociation for a recombinant humanized monoclonal antibody to vascular endothelial
growth factor. Biochemistry 38:13960–13967
Narhi LO, Jiang Y, Cao S, Benedek K, Shnek D (2009) A critical review of analytical methods for
subvisible and visible particles. Curr Pharm Biotechnol 10:373–381
North SJ, Hitchen PG, Haslam SM, Dell A (2009) Mass spectrometry in the analysis of N-linked
and O-linked glycans. Curr Opin Struct Biol 19:498–506
Oliva A, Llabres M, Farina JB (2004) Applications of multi-angle laser light-scattering detection
in the analysis of peptides and proteins. Curr Drug Discov Technol 1:229–242
Pease LF 3rd, Elliott JT, Tsai DH, Zachariah MR, Tarlov MJ (2008) Determination of protein
aggregation with differential mobility analysis: application to IgG antibody. Biotechnol Bioeng
101:1214–1222
Philo JS (2006) Is any measurement method optimal for all aggregate sizes and types? AAPS
J 8:E564–E571
Philo JS (2009) A critical review of methods for size characterization of non-particulate protein
aggregates. Curr Pharm Biotechnol 10:359–372
Rambaldi DC, Reschiglian P, Zattoni A (2011) Flow field-flow fractionation: recent trends in
protein analysis. Anal Bioanal Chem 399:1439–1447
Rivas G, Stafford W, Minton AP (1999) Characterization of heterologous protein-protein interac-
tions using analytical ultracentrifugation. Methods 19:194–212
Roux KH (1999) Immunoglobulin structure and function as revealed by electron microscopy. Int
Arch Allergy Immunol 120:85–99
Sandin S, Öfverstedt L-G, Wikström A-C, Wrange Ö, Skoglund U (2004) Structure and flexibility
of individual immunoglobulin G molecules in solution. Structure 12:409–415
Sharma DK, King D, Moore PD, Oma P, Thomas D (2007) Flow microscopy for particulate analy-
sis in parenteral and pharmaceutical fluids. Eur J Parenter Pharm Sci 12:97–101
Tatford OC, Gomme PT, Bertolini J (2004) Analytical techniques for the evaluation of liquid pro-
tein therapeutics. Biotechnol Appl Biochem 40:67–81
Thomson NH (2005) Imaging the substructure of antibodies with tapping-mode AFM in air: the
importance of a water layer on mica. J Microsc 217:193–199
Vendruscolo M (2007) Determination of conformationally heterogeneous states of proteins. Curr
Opin Struct Biol 17:15–20
Wang YJ, Shahrokh Z, Vemuri S, Eberlein G, Beylin I, Busch M (1996) Characterization, stability,
and formulations of basic fibroblast growth factor. Pharm Biotechnol 9:141–180
Wang G, Johnson AJ, Kaltashov IA (2011) Evaluation of electrospray ionization mass spectrometry
as a tool for characterization of small soluble protein aggregates. Anal Chem 84:1718–1724
4 Biophysical Techniques for Protein Size Distribution Analysis 97
5.1 Introduction
As described elsewhere (Jiang and Narhi 2006) and in other chapters of this book,
biophysical methods are routinely used by the biopharmaceutical industry to gain
information on protein higher-order structure and stability throughout the therapeu-
tic protein development life cycle. These methods include SV-AUC and SE-HPLC-LS
for size distribution analysis and molecular weight determination of species sepa-
rated by SE-HPLC; CD, FTIR, and DSC for secondary and tertiary structure and
overall conformation and thermal stability analysis; and light obscuration (LO)- and
micro-flow imaging (MFI)-based methods for subvisible particle analysis. The
results of these biophysical analyses have been included in regulatory filings as
required by ICH guidelines M4Q (R1), Q5E, and Q6B (ICH M4Q 2001; ICH Q5E
2005; ICH Q6B 1999). These results are included in the filing documents to demon-
strate that the therapeutic protein shows the predicted and expected properly folded
higher-order structure and retains the higher-order structure after manufacturing
process, site, and/or formulation changes. ICH guidelines M4Q (R1) (ICH M4Q
2001) on common technical document (CTD) for the registration of pharmaceuti-
cals for human use and Q6B (ICH Q5E 2005) on specifications state that for desired
Y. Jiang (*)
Process and Product Development, Amgen Inc., One Amgen Center Drive, 30E-1-C,
Thousand Oaks, CA 91320, USA
e-mail: [email protected]
C. Li
Process and Product Development, Amgen Inc., One Amgen Center Drive,
Thousand Oaks, CA 91320, USA
e-mail: [email protected]
J. Gabrielson
Analytical Sciences, Amgen Inc., 4000 Nelson Road, Longmont CO 80503, USA
e-mail: [email protected]
5.2 Qualification of CD
Circular dichroism (CD) spectroscopy is a technique that has been widely used in
structural biology for examining secondary structure, tertiary structure, conforma-
tional changes, and folding and binding interactions involving protein molecules. It
is also used by the biopharmaceutical industry as a characterization tool to study the
effect of manufacturing processes, formulation compositions and conditions, stor-
age conditions, and delivery systems on protein conformation (Johnson 1988,
1999a; Woody 1994, 1996; Manning and Woody 1989; Greenfield and Fasman
1969; Yang et al. 1986). This information is often included in regulatory filings fol-
lowing changes in any of these manufacturing conditions (ICH Q5E 2005; ICH
Q6B 1999). In recent years, a new extension of the conventional CD method, syn-
chrotron radiation circular dichroism (SRCD), has been shown to have a better sen-
sitivity for the higher-order structure of a protein by producing data to lower
wavelengths with higher information content and improved signal-to-noise levels
relative to those obtained using conventional benchtop instruments (Wallace 2005;
Lees and Wallace 2002; Miles et al. 2007; Wallace and Janes 2010). SRCD
spectroscopy can provide important static and dynamic structural information on
proteins in solution, including secondary structures of intact proteins and their
domains, protein stability, the differences between wild-type and mutant proteins, the
identification of natively disordered regions in proteins, and the dynamic p rocesses of
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 101
protein folding and membrane insertion and the kinetics of enzyme reactions (Miles
and Wallace 2006; Meersman et al. 2010; Drechsler et al. 2009). A new web-based
bioinformatics resource, the Protein Circular Dichroism Data Bank (PCDDB), has
also been created recently by Dr. Wallace’s laboratory which enables archiving,
access, and analyses of CD and SRCD spectra and supporting data (Powl et al.
2010; Whitmore and Wallace 2004; Wallace et al. 2006; Lees et al. 2006).
Unfortunately many of the formulation buffers used for therapeutic proteins interfere
with the SRCD, especially at the lower wavelengths required, and this technique can
usually not be applied to comparability evaluations or similar studies.
Even though CD spectroscopy has received considerable attention and been
widely used for biological molecules over the last 20 years by both industry and
academic scientists, the technique is mostly used in research and development labo-
ratories rather than quality control laboratories. This is mainly due to the complex-
ity of the analysis, the need for an experienced spectroscopist to run the studies, and
the qualitative nature of the spectra. The lack of defined best practices in making the
CD measurement and the variability in the data obtained from different laboratories
also make this an inappropriate technique for routine use in quality control labora-
tories. There have been a few publications describing standardization of best prac-
tices in making CD measurements (Van Stokkum et al. 1990; Kelly et al. 2005).
To understand the extent of comparability of CD data, the National Physical
Laboratory (NPL) has conducted an interlaboratory CD spectroscopy study in
recent years. Schiffmann et al. published a report titled “CD spectroscopy: an inter-
laboratory study” (Jones et al. 2004) and Ravi et al. (2010) published two papers
about the international comparability in spectroscopic measurements of protein
structure by CD (Schiffmann et al. 2004; Ravi et al. 2010). Both studies involved
academic and industry laboratories worldwide. The studies have highlighted some
of the current difficulties in obtaining comparable CD spectra from different instru-
ments and possible solutions to some of these problems, for instance, the iterative
spectral alignment method (ISAM). Both studies show that the uncertainties
involved in a CD measurement are rather complex, involving components relating
to calibration, path length, sample concentration, and instrumentation (Ravi et al.
2010; Whitmore and Wallace 2008; Miles et al. 2003, 2004, 2005). Therefore, a
careful calibration of the CD instrument and exact path lengths of the cuvettes
must be carried out to ensure equivalence of CD measurements when using dif-
ferent instruments and cuvettes. The calibration and standardization of CD
instruments has been described by Dr. Wallace et al. (Whitmore and Wallace 2008;
Miles et al. 2003). The authors suggested the procedure for proper calibration of a
CD spectrophotometer and proposed a method for standardization of protein spectra
obtained on any instrument. For example, calibrations should be done at more than
one wavelength in order to cover the wavelength range measured in a protein spec-
trum, and the linearity of the response should be demonstrated; the ratios of the
measured and expected ellipticity values at each wavelength should be used to
develop a model function for corrections.
For application of the method in the biopharmaceutical industry, one of the
important questions is the degree of comparability of the CD data between different
102 Y. Jiang et al.
Fig. 5.1 (a) Far-UV CD a
spectra of protein 1–4 6000
(spectral similarity compared
to that of protein 1 is 100% 4000
protein 1; 5.0% protein 2;
44.9% protein 3; 28.7% 2000
CD Ellipticity
protein 4). (b) Near-UV CD
spectra of protein 1–4 0
(spectral similarity compared
to that of protein 1 is 100% -2000 Protein-1
Protein-2
protein 1; 41.0% protein 2;
Protein-3
3.1% protein 3; 12.9% -4000
Protein-4
protein 4)
-6000
200 210 220 230 240
Wavelength (nm)
Tyr
b Phe Trp
80 S-S
40
CD Ellipticity
Protein-1
-40
Protein-2
Protein-3
-80 Protein-4
-120
250 300
Wavelength (nm)
similarity when comparing the spectra obtained with the same instrument and
cuvette independent of the analyst, location, and date; and (2) for far-UV CD, the
precision varies depending on the structural nature of the protein and the buffer
composition. For proteins with helical secondary structure in buffers with little
interference in the spectra, far-UV CD analysis shows low variability in spectral
similarity. For proteins with predominantly beta-sheet secondary structure and in
buffers where the excipients have absorbance or CD signals in this region, far-UV
CD analysis shows higher variability in spectral similarity. The sensitivity assess-
ment of CD spectroscopy is challenging because the unfolding of a protein by some
downstream processing conditions, for example, low pH or Gdn, is often reversible
or partially reversible when the stresses are removed. Thus a standard curve that
correlates spectral similarity to the percentage of denatured form of the protein can-
not be generated empirically. To address this issue, Li et al. used a computer-
simulated blending study of the spectra of the native protein and the unfolded form
104 Y. Jiang et al.
of the same protein to generate a standard curve that correlates the spectral s imilarity
of the measured spectrum to the native spectrum with the amount of unfolded form
of the protein present. To determine the sensitivity of CD spectroscopy, one would
have to combine the precision assessment of the technique under the required test
conditions (e.g., protein concentration and buffer) with the spectral similarity cor-
relation of the simulated blending study of denatured and native proteins. The
sensitivity of the method can be expressed as the percentage of denatured form of
the protein when the spectral similarity of the sample spectrum compared to the
native/control protein exceeds the precision of the method. Under different condi-
tions and with a different protein, the precision and sensitivity of the CD analysis
may also change.
The results from Li et al.’s qualification study indicate that CD spectroscopy can
be qualified for characterizing protein secondary and tertiary structures and for
comparability studies of biopharmaceuticals. With proper calibration of the instru-
ments and cuvettes, and standardized good operating practices, CD spectroscopy is
quite reproducible and able to detect changes induced in the secondary and tertiary
structure of a protein by lower pH and other possible denaturing reagents that are
present during therapeutic protein manufacturing processes. The sensitivity of the
technique to detect the changes in conformation that might be induced by the manu-
facturing process is dependent on the experimental variability and the nature and
extent of the structural changes in the particular protein being analyzed.
The results shown above provide the ground work on how the CD method should
be calibrated and qualified and demonstrate that the method is sufficiently precise to
monitor protein structure changes beyond the variability of the method.
Work has been done to quantify the percentages of different secondary structural
components for proteins in solution by FTIR (Vonhoffa et al. 2010; Susi and Byler
1987). There are two primary ways of doing this. The first uses curve fitting of the
second-derivative or self-deconvoluted spectra to obtain the relative amounts of dif-
ferent types of secondary structure based on the band areas (Susi and Byler 1986;
Byler and Susi 1986). The second involves peak fitting of the non-deconvolved and
baseline-corrected amide I bands and then obtaining the percentage of secondary
structures by correlating with the shape and intensity using an interval partial least
squares algorithm (Vedantham et al. 2000; Surewicz and Mantsch 1988). These
methods are helpful in determining the major secondary structure components and
estimating their percentages in a protein. However, these methods are not optimized
for assessing overall similarity of protein structure between samples from different
processes or formulations. The quantification of the secondary structure composi-
tion can vary depending on the methods and parameters used, and it involves con-
siderable mathematical manipulations including spectra manipulation, curve fitting,
area integration, and normalization. Therefore, they are not routinely carried out for
protein secondary structure assessment and qualification of FTIR by analysts in the
biopharmaceutical industry.
One other mathematical approach explored by Prestrelski et al. (1993) is to com-
pare the FTIR spectra quantitatively by using the correlation coefficient function.
Occasionally, the correlation coefficient value of the uncorrected spectra did not
agree with a visual assessment of the spectral similarity due to an offset in baselines
which led to an artificially low value. Conversely, if the spectra were baseline cor-
rected and peak positions were similar, but there were differences in relative peak
heights, the correlation value would be unreasonably high. To avoid this inconsis-
tency, Kendrick has developed a method to quantify the overlap of second-derivative
FTIR spectra to determine structural similarity between proteins (Kendrick et al.
1996). In this approach area-normalized second-derivative spectra were used and
compared. The authors found that quantifying the area of overlap between area-
normalized spectra provides a reliable, objective method to compare overall spectral
similarity and avoids the problems associated with calculation of the correlation
coefficient. However, due to normalization of the spectra, the area of overlap
approach cannot distinguish between true differences in the magnitude of FTIR
signals. Recently, D’antonio et al. 2012 showed a modified area of overlap method
that improved the differentiation power for quantitative comparison of FTIR spec-
tra. The authors also compared four different algorithms including the correlation
coefficient and area of overlap and summarized their strength and weakness for
studying comparability of protein therapeutics.
A new approach using Thermo Electron OMNIC software QC compare function
(Cover and Hart 1967; TQ Analyst Algorithm 2007–2010) has been applied to
directly compare FTIR spectra with minimal mathematical manipulation (Fig. 5.2),
which led to the identification of important performance characteristics for the qual-
ification of the FTIR method (Jiang et al. 2011). The QC compare function was
originally intended and used as a library searching tool for small-molecule identifi-
cation by FTIR (TQ Analyst Algorithm 2007–2010). It correlates the spectral
106 Y. Jiang et al.
2.0
1.0
0.0
Arbitrary unit
-1.0
-2.0
-3.0
-4.0
10-4
1720 1700 1680 1660 1640 1620 1600
Wavenumbers (cm-1)
Fig. 5.2 Second-derivative FTIR spectra of protein 1–5 (spectral similarity compared to that of
protein 1 is 100% protein 1 (pink); 46.3% protein 2 (purple); 36.3% protein 3 (green); 9.7% pro-
tein 4 (blue); 12.1% protein 5 (black))
information in the specified region of the sample and the reference spectra to
determine the similarity between the two. The results of the method are reported as
a value between 0% and 100%, which indicates how well the sample spectrum
matches the reference spectrum. A spectral similarity value of 100% indicates the
two spectra are identical.
The selection of an appropriate approach to numerically compare the spectra
should depend on the capability of the method, the ease of access and use of the
method, and the purpose of the study.
With the OMNIC QC compare algorithm, Jiang and colleagues evaluated the pre-
cision and sensitivity of FTIR for the analysis of protein secondary structure through
a multisite/instrument and multi-analyst study in an effort to qualify the method
(Jiang et al. 2011). Proteins containing different types of secondary structures such
as alpha-helical and beta-sheet were included to ensure that the precision assessment
would apply to secondary structure analysis of all proteins regardless of the specific
structural type. In addition inter-day repeatability and the effect of protein concentra-
tion differences on the precision of the method were also evaluated. The overall FTIR
spectral similarities of these analyses on the same proteins were compared. The sen-
sitivity of the method was evaluated by comparing the reference spectrum of a pro-
tein to that of the partially or fully unfolded protein after exposure to the denaturing
condition and also by blending studies where the spectrum of a reference was mixed
with that of a denatured protein. Standard curves were generated where the spectral
similarity was plotted as a function of the percentage of the unfolded protein.
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 107
Results by Jiang et al. (2011) demonstrate that the FTIR method for the analysis
and characterization of protein secondary structure is precise with the standard
deviations of the characteristic FTIR band frequencies <1.1 cm−1 and the spectral
similarity >93% for replicate measurements. The FTIR spectra remain the same
whether collected on the same day or during a three-day period and when the pro-
tein concentration deviates from the target concentration (≥30 mg/mL) by ≤10%
regardless of the maker of the instrument used or the structural family of the protein.
The authors also demonstrate that the method is sensitive and appropriate for assess-
ing protein secondary structure and changes in conformation resulting from stresses
that can be encountered during common biopharmaceutical manufacturing pro-
cesses. The sensitivity of the FTIR method is dependent on the extent of the struc-
tural changes induced and the magnitude of the resulting changes in the spectra.
The precision of the FTIR method is closely related to the signal-to-noise ratio
of the instrument, the sample concentration, the buffer components, and the spectral
region of interest. The sensitivity of the method to detect structural change depends
on the nature of the sample and relative spectral changes corresponding to the struc-
ture changes. Therefore, protein-specific qualification/verification of the method
may need to be performed to ensure the understanding of precision and sensitivity
of the method.
DSC has been used widely to study thermal dynamics of protein folding. The melt-
ing/thermal transition temperature(s) of proteins can be used to understand the energy
barrier for protein unfolding and thermal stability of proteins (Pyrpassopoulos et al.
2006; Johnson et al. 1995). DSC has also been used to study the binding interactions
of proteins with other molecules and the effect of mutations and the presence of
carbohydrates on protein thermal stability (Celej et al. 2006; Bruylants et al. 2005;
Johnston et al. 2011; Protasevich et al. 2010; Wen et al. 2008). The application of
DSC has been reviewed extensively in the area of characterization of macromole-
cules and their interactions during pharmaceutical product development (Chiu and
Prenner 2011; Bruylants et al. 2005). However, even though the method has been
used extensively to monitor and show changes in protein thermodynamic parame-
ters such as enthalpy and melting temperature (Tm), there are very limited published
data on the systematic qualification of the method to show that the DSC measure-
ments are precise, accurate, and sensitive for the analysis of protein conformation
and thermal stability. A protocol on measuring protein thermostability by DSC
was published by Makhatadze (Coligan et al. 2001). It provides a detailed proce-
dure for conducting DSC analysis including sample preparation and interpretation
of the results, as well as calibration of the instrument and maintenance of DSC
cells. Calibration of DSC was described by Gmelin and Sarge (Gmelin and Sarge
1995). They provided recommendations for instrument-independent calibration of
temperature, heat, heat flow rate of the scanning calorimeter, operation procedures,
108 Y. Jiang et al.
calibration substance, and data treatment algorithms which can be used for all
instruments.
In the early days of DSC analysis, various data analytical algorithms were devel-
oped and used to assess the effect of instrument configuration and scan rate on DSC
measurement results (Lopez and Freire 1987). During the use of DSC for purity
analysis of drug product, instrumental differences, sample size, heating rate, and
details of the calculation of such analysis were examined (Yoshii 1997). It was found
that changing the sample size or heating rate resulted in a difference in the effect of
purity value between two instruments. Collectively controlling the heating rate dur-
ing DSC measurement appears to be critical in obtaining reproducible results.
The precision of DSC measurements across different instruments and labs using
the same sample (polymeric material) and parameters has been assessed. The study
was organized by Empa (Swiss Federal Institute for materials testing and research)
(Schmid 2012; Affolter et al. 2001). It collected measured data on glass transition
(Tg) and melting point (Tm) and evaluated those using statistical methods to assess
the precision of DSC. The results show that variability of DSC measurements is
mainly caused by differences in the analyst, instrument, and calibration of the
instrument. For one-point temperature measurements such as Tg and Tm, good agree-
ments were observed with a standard deviation of repeatability at 0.3–1.0°C and
that of reproducibility at 1.0–2.1°C. The standard deviation increases significantly
with user-defined data evaluation.
A recent paper by Wen et al. (2011) explored the qualification of DSC for its
applications in thermal stability analysis of proteins. The authors assessed the preci-
sion and sensitivity of the DSC method through a multisite, instrument, and analyst
study using several proteins from different structural families including monoclonal
antibodies and cytokines and parameters including Tm and profile similarity. The
same experimental parameters such as heating/scan rate and similar protein concen-
trations were used for measurement on all instruments with the exception of one.
The results show that the Tm values obtained for the same protein by the same or
different instruments and/or analysts are quite reproducible, varying generally
<1.1°C with the same instrument and <1.5°C across different instruments.
The profile similarity values obtained using the OMNIC QC compare function
(discussed above in the CD and FTIR sections) for the same protein from the same
instrument are also high (> ~95%). The variability of inter-day DSC measurements
is very similar to that of the intra-day measurements (only slightly higher by
~0.1°C). The sensitivity of the DSC method for assessing protein thermal stability
and conformational changes was evaluated by several experiments including ana-
lyzing samples perturbed by pH 3 or 6 M guanidine HCl (Gdn). The results show
that DSC is able to detect changes in protein conformation caused by low pH and
denaturant to levels as low as 10% denatured protein present in the sample. The
authors then applied DSC to the analysis of pH stability and buffer screening of a
protein and candidate screening. They demonstrated that DSC is an appropriate
method for assessing protein thermal stability and conformational changes that may
result from manufacturing processes, formulation, and storage conditions and can
also be used to compare relative stability of candidate molecules.
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 109
the overall shape and charge of the size variant has been maintained when multiple
samples are measured under the same solution conditions. However, the accuracy
and precision of the measured sedimentation coefficient decreases as the concentra-
tion of the size variant decreases (Gabrielson et al. 2007b). In very pure samples
with low concentrations of fragmented and aggregated size variants (<1% of the
total protein mass), only the sedimentation coefficient of the most abundant species
(typically monomer) can be reliably determined.
In the biopharmaceutical industry, SV-AUC is primarily used as an orthogonal
method to SE-HPLC to quantitatively measure protein purity and secondarily used
to measure protein monomer sedimentation coefficients (Gabrielson et al. 2010).
Qualification of an SV-AUC method should reflect the fact that it will be used to
quantitatively measure these two attributes. Furthermore, two performance charac-
teristics of an SV-AUC method should be considered during qualification: (1) evalu-
ation of the method’s precision and (2) confirmation that the method is sufficiently
sensitive to provide quantitative results, thereby allowing meaningful conclusions to
be drawn.
Assessing the precision of SV-AUC, and determining which type(s) of precision
to evaluate, should be governed by how the method will be used. For example, if
SV-AUC analysis can typically be completed by one analyst using one instrument
for product characterization studies required during biotherapeutic development,
then it may not be necessary to study the reproducibility of the method. However, if
the number of samples requires multiple runs on different days, then understanding
the intermediate precision of the method is critical.
Experimental factors that impact the precision of SV-AUC measurements have
been reviewed elsewhere (Gabrielson et al. 2007b, 2010; Schuck et al. 2002;
Gabrielson and Arthur 2011; Arthur et al. 2009; Pekar and Sukumar 2007) and
include the instrument, rotor, centerpiece and other cell components, overall align-
ment of channels, temperature equilibration, analysis software and settings, and
sample characteristics. Even under the tightest experimental control currently
achievable, normal variation of these factors leads to variability in results of approx-
imately 0.4% for size variant quantification (Gabrielson and Arthur 2011; Arthur
et al. 2009) and about 0.02 S for sedimentation coefficient measurements (Gabrielson
and coworkers, personal observation, unpublished data). Much of the imprecision
arises from the specific centerpieces used (Arthur et al. 2009; Pekar and Sukumar
2007) and from the alignment of cells (Pekar and Sukumar 2007), factors that can
vary as much within a run as they do across runs or even across laboratories.
A robust qualification design for an SV-AUC method should therefore account for
normal variations from all relevant experimental factors, that is, those factors that
reflect how the method will be applied during development of a biotherapeutic.
Studies to evaluate SV-AUC method performance indicate that performance
parameters including linear range, limit of detection (LOD), limit of quantitation
(LOQ), and measurement precision are, in general, independent of the specific
protein product analyzed (Gabrielson et al. 2007b; Gabrielson and Arthur 2011;
Arthur et al. 2009). Variability in measured results is primarily due to experimental
factors, such as centerpiece differences, and depends to a lesser extent on sample
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 111
characteristics. Therefore, after sufficient data have been acquired to account for
relevant experimental factors, it is possible to qualify SV-AUC analysis of a new
protein product by simply verifying that the precision of sedimentation coefficient
measurements and the precision of size variant measurements are within a range
expected from previous experience with the method. Additional qualification exper-
iments may be warranted if the expected method performance is not achieved for a
particular product or sample type, for example, protein samples with reversible self-
association. In those cases, protein-specific intermediate precision should be more
rigorously defined.
The ability of SV-AUC to detect changes to the product, both to its size distribu-
tion profile and to the sedimentation coefficient of the monomer, is essential for
successful application of the method in biotherapeutic development. Therefore,
method qualification should include experiments demonstrating the suitability of
SV-AUC for detecting such changes. This may be accomplished by manipulating a
sample to perturb the protein structure and to produce higher levels of size variants.
A suitable amount of degradation can usually be obtained from thermal, UV, or pH
stresses. For example, subjecting a monoclonal antibody sample to low pH condi-
tions will often cause the sedimentation coefficient of the monomer to shift and
induce formation of high molecular weight size variants (aggregates), which can be
used to confirm the sensitivity of the method.
The sensitivity of the method can best be determined when changes to the mea-
sured output are correlated with changes deliberately made to the product. Changes
to the sedimentation coefficient of a protein and the amount of aggregation mea-
sured by SV-AUC for a different protein, both induced by exposure to low pH, are
shown in Fig. 5.3.
The data presented in Fig. 5.3 are instructive for several reasons. First, the
response of the method to a given stress condition may or may not be linear. In panel
A, the sedimentation coefficient responds nearly linearly to pH changes from 5.5 to
3.5 and then declines nonlinearly at more acidic pH, whereas in panel B the maxi-
mum aggregation levels are measured in a narrow pH range near 3.5; aggregation
levels decline markedly at higher and lower pH. Second, it is often convenient to
express the method sensitivity by normalizing the response, either by dividing by
the range of the independent variable (i.e., by calculating the slope if the response is
linear) or by dividing by an estimate of the variability of the method to express the
change in response as a number of standard deviations. For example, in panel A of
Fig. 5.3, a pH change from 5.5 to 3.5 results in a 0.2 Svedberg shift in the sedimen-
tation coefficient. The sensitivity of the method to pH could then be expressed as 0.1
Svedberg per pH unit (or five standard deviations per pH unit when normalized by
the method variability of 0.02 S). Finally, method sensitivity studies are most useful
when they evaluate a range wider than what the product will likely encounter during
manufacture, shipping, and long-term storage. This allows interpolation of the
response, rather than extrapolation to conditions which have not been studied.
Any number of stress conditions can be used to demonstrate the sensitivity of
SV-AUC. These may include thermal degradation, light/UV exposure, pH modifica-
tion, chemical denaturation, or primary structural modifications (e.g., oxidation).
112 Y. Jiang et al.
Fig. 5.3 Sedimentation
coefficients (a) and
aggregation levels (b)
measured by SV-AUC for
protein product samples
across the pH range indicated
Many considerations should inform the selection of product stress conditions with
which to determine the sensitivity of the method, including the likelihood that the
product will encounter a certain stress during manufacturing and the intended
change to the product. That is, a general stress like elevated temperature may be
desired to produce aggregation, whereas a more targeted stress like oxidation may
be useful to correlate the sedimentation rate of the protein with changes to its
primary structure. And finally, the intended purpose of the method should also be
considered when selecting product stress conditions. A different stress condition
may be desirable if the method will be used exclusively to characterize a drug sub-
stance manufacturing process compared to if the method will be used to evaluate
drug product comparability.
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 113
signal and noise estimates should be determined from the 90° detector. (For isotropic
scatters, such as proteins, the 90° detector has the highest signal-to-noise ratio.)
Each detector’s signal is taken to be the maximum signal (in volts) of the peak of
interest. The noise for each detector is estimated from a region without protein sig-
nal (i.e., a baseline region) using any one of several noise calculation approaches
such as the square root of the mean squared error.
The governing equations that relate the signals of each detector to the physical
attributes of the protein are provided in Eqs. 5.1, 5.2, and 5.3 (Wen et al. 1996):
2
dn
SLS = kLS Mc , (5.1)
dc
SUV = kUV elc, (5.2)
dn
SRI = kRI c, (5.3)
dc
where kLS (V mol L mL−2), kUV (V), and kRI (V g mg−1) are detector proportionality
constants and l (cm) is the path length of the UV cell, also a constant. The concen-
tration, c (mg mL−1), of the size variant can be determined from either Eq. 5.2 or
Eq. 5.3, depending on the choice of detector, when the extinction coefficient ε
(mL mg−1 cm−1) and the refractive index increment dn/dc (mL g−1) are provided. The
molecular weight of the size variant, M (g mol−1), is then determined from Eq. 5.1.
The error of the molecular weight measurement can be estimated by applying a
propagation of error analysis to Eq. 5.1. The generalized form of the error propaga-
tion formula (Taylor 1997) is shown in Eq. 5.4 for a response, f, that is a function of
independent variables, x, y, and z:
2 2 2
∂f ∂f ∂f
e f = ex + ey + ez ,
∂x ∂y ∂z (5.4)
where ef is the expected error in the calculated value of f and ex, ey, and ez are error
estimates for variables x, y, and z, respectively.
When LS and UV detectors are connected in series, the size variant concentration is
calculated from Eq. 5.2, which is then used to calculate the size variant molecular
weight from Eq. 5.1. It is assumed that some degree of uncertainty exists in the
protein extinction coefficient and refractive index increment, both of which are pro-
vided by the analyst. Propagation of error using Eq. 5.4 leads to the following
expression for the expected error in the molecular weight calculation (Eq. 5.5):
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 115
2 2 2 2
eM e e e e
= SLS + SUV + e + 4 dn / dc . (5.5)
M SLS SUV e dn / dc
As expected, the molecular weight accuracy depends on the error in the LS and
UV detector signals as well as uncertainty in the extinction coefficient and refrac-
tive index increment, which may not be explicitly known for peaks in the chromato-
gram. The error in each detector’s signal can be approximated by its noise (i.e.,
eSLS ≈ NLS) by assuming that inter-detector band broadening is either negligible or
accounted for by the data analysis software.
2 2 2
eM e e e
= SLS + SRI + dn / dc . (5.6)
M SLS SRI dn / dc
In this case, the error expression has only three terms: error in each detector’s
signal and uncertainty in the protein refractive index increment.
In the special case where all three detectors are connected in series, the refractive
index increment can be measured using the RI detection. This can be shown by
reexpressing Eq. 5.1 after solving for M and using Eqs. 5.2 and 5.3 to eliminate the
dn/dc term, as shown in Eq. 5.7:
SLS SUV
M=k , (5.7)
2
SRI e
where k is a collection of constants. Using all three detectors, an a priori estimate of
the refractive index increment is not required, but the extinction coefficient is required
to determine the concentration. Following similar error propagation analysis, an
expression for the error of the molecular weight calculation can be determined:
116 Y. Jiang et al.
2 2 2 2
eM e e e e
= SLS + SUV + 4 SRI + e . (5.8)
M SLS SUV SRI e
Because all three detectors are used, errors in each of their signals appear in the
overall error expression. Uncertainty in the protein extinction coefficient will also
impact the accuracy of the molecular weight determination.
Using the appropriate error expression based on the choice of detectors used, and
accounting for uncertainty in the protein extinction coefficient and/or refractive
index increment, allows an assessment of molecular weight accuracy to be per-
formed for any peak in any chromatogram. Thus, an evaluation of accuracy is not
made at a single point in time but is a dynamic assessment that varies with protein
load, size variant mass, and system noise. The accuracy also improves as analysts
apply more accurate estimates of extinction coefficients and refractive index
increments.
A graphical representation of the molecular weight error estimates from the three
detector arrangements is provided in Fig. 5.4. Fractional error in the molecular
weight (eM/M) is plotted as a function of the size variant concentration for different
levels of uncertainty in the extinction coefficient and/or refractive index increment.
At high protein concentrations (where S » N), the fractional errors in the detector
signal terms in Eqs. 5.5, 5.6, and 5.8 are negligible. Therefore, for cases 2 and 3 in
which LS and RI detection are used, the fractional error in molecular weight
approaches the fractional uncertainty in either dn/dn (LS-RI, Eq. 5.6, Fig. 5.4b) or ε
(LS-UV-RI, Eq. 5.8, Fig. 5.4c) at high protein concentrations. When UV detection
is used without RI detection, the limiting error in the molecular weight is higher
than when RI detection is used because uncertainty in both dn/dn and ε contribute
to the error estimate (LS-UV, Eq. 5.5, Fig. 5.4a). At low concentrations (where
S ~ N), the LS signal error term dominates the molecular weight error expression
(all panels of Fig. 5.4).
To show that the instrumentation is performing properly and the calculated molar
masses are meaningful, the measured masses by LS should be compared to a
standard of known mass such as bovine serum albumin. At least one such standard
should be analyzed within each sequence to confirm the system’s suitability. Then,
the molecular weights of size variants in the sample elution profile can be reported
with associated error estimates derived from Eqs. 5.5, 5.6, or 5.8.
An understanding of both the precision and accuracy of measured molecular
weights using LS detection should be considered when determining whether the
method will be appropriate for specific applications in the biopharmaceutical indus-
try. Addition of LS detection to an SE-HPLC method may be useful for product
profile characterization, contaminant identification, and product comparability
assessments. Nevertheless, the performance required of the method may be differ-
ent for each of these applications, which is why qualification of the method is criti-
cal to enable selection of appropriate uses of LS detection.
Fig. 5.4 Molecular weight fractional error estimates derived from error propagation analysis
(Eqs. 5.5, 5.6, and 5.8) and plotted across a range of protein concentrations. In each panel, different
curves represent different levels of uncertainty for ε and/or dn/dc. The curves were generated with
0, 2.5%, 5%, and 10% error for ε and dn/dc to encompass the expected range of uncertainty in
these parameters. (a) LS-UV detectors (Eq. 5.5). Fractional errors in ε/dn/dc from lowest curve to
highest: 0/0 (lowest curve); 0.025/0 (curve 2); 0.05/0 and 0/0.025 (indistinguishable, curve 3);
0.025/0.025 (curve 4); 0.05/0.025 (curve 5); 0/0.05, 0.1/0, and 0.025/0.05 (indistinguishable, curve
6); 0.05/0.05 and 0.1/0.025 (indistinguishable, curve 7); 0.1/0.05 (curve 8); 0/0.1, 0.025/0.1, and
0.05/0.1 (indistinguishable, curve 9); and ε = 0.1, dn/dc = 0.1 (highest curve). (b) LS-RI detectors
(Eq. 5.6). Fractional errors in dn/dc from lowest curve to highest: 0, 0.025, 0.05, and 0.1.
(c) LS-UV-RI detectors (Eq. 5.8). Fractional errors in ε from lowest curve to highest: 0, 0.025,
0.05, and 0.1. Data shown in all panels were generated using the following values: NUV = 0.00035
V; NLS = 0.0055 V; NRI = 0.00025 V; kUV = 10 V; kLS = 0.0015 V mol L mL−2; kRI = 20 V g mg−1;
M = 34,000 g mol−1; ε = 0.48 mL mg−1 cm−1; l = 1 cm; and dn/dn = 0.165 mL g−1
118 Y. Jiang et al.
range (i.e., <5 μm). The effect of the difference in refractive index between the
polystyrene calibration spheres and test particles on the apparent size of particles
determined by LO-based methods has been observed for other types of particles
such as lipids as well (Driscoll 2004). However, the demonstrated precision and
linearity, as well as the broadened size range of the modified LO method for protein
samples, enable the use of the method for monitoring effectiveness of filtration and
clearance of SbVPs during downstream processing steps, assessing the effect of
various process stresses on SbVP formation, and the contribution of primary con-
tainer components to the SbVP present (Cao et al. 2009, 2010).
The validation of the particle size method and instrument qualification have been
described by Rawle in the book on “Practical Approaches to Method Validation and
Essential Instrument Qualification”(Chan et al. 2011) and in the pharmacopeia
(United States Pharmacopeial Convention 2009; Council of Europe 2007) from dif-
ferent country/regions. The procedures for instrument calibration, performance
verification, and the use of particle counting standards are specified. Different
instrument vendors essentially follow the same calibration procedure for LO-based
instruments. This ensures the precision and accuracy of the measurements when
polystyrene particle counting standards are measured. The performance of the LO
method for protein therapeutics needs to be verified as shown by Cao et al. before it
is used for testing.
Even though LO-based methods have been demonstrated to be precise and are
widely used by the pharmaceutical industry to monitor SbVPs in their injectable
products, the information obtained from these methods is limited to the size and
concentration of the particles. In order to gain more understanding of the nature of
SbVP, MFI-based techniques were developed recently (Mahler et al. 2009; Huang
et al. 2009; Sharma et al. 2010a; Wuchner et al. 2010). Flow imaging techniques
allow the characterization of particle shape and morphology in addition to size and
concentration. It has been shown to be more sensitive than LO methods for measur-
ing subvisible proteinaceous particles in protein formulations.
Due to the potential challenges in detecting particles in opalescent mAb formula-
tions, the accuracy of MFI in determining size and particle counts in opalescent
solutions was investigated and compared to LO and membrane microscopy methods
by Sharma et al. (2010b). They used proteinaceous monoclonal antibody (mAb)
particles, generated either by chemical denaturation or agitation stress, polystyrene,
and glass particles as model systems for measurements in opalescent mAb solu-
tions. The authors demonstrate that the sizing and counting accuracies of MFI were
unaffected by the opalescence of the medium. Using glass particles as a model
system for proteinaceous particles, MFI was able to detect relatively low particle
concentrations (approximately 10 particle/mL) in opalescent solutions. MFI showed
excellent linearity (R(2) = 0.9969) for quantifying proteinaceous particles in opales-
cent solutions over a wide range of particle concentrations (approximately
20–160,000 particle/mL). The authors also show that the intensities of MFI particle
image are significantly different with respect to the transparency of proteinaceous
particles as a function of their size and mode of generation. The LO method signifi-
cantly underestimated proteinaceous particles, particularly those in the 2–10 μm
120 Y. Jiang et al.
size range. The less opaque proteinaceous particles, the more underestimated by the
LO method. Similar phenomenon was observed by Driscoll for lipid particles
(Driscoll 2004).
Polystyrene bead standards are generally used as a reference for calibrating and
validating particle analyzers. However, these standards, unlike protein particles, are
easily detected and do not challenge the sensitivity of optical instruments.
Instruments calibrated and verified only with beads can still exhibit significant dif-
ferences in measuring concentrations of protein particles. To ensure the accuracy of
the analysis of the size and concentration of suspended protein particles in protein
therapeutics, standards should resemble protein particles with respect to their trans-
parency and refractive index. Sharma et al. (2011) described the work on evaluating
a potential standard and the use of it to harmonize the performance of MFI instru-
ments. They have shown that the use of a suitable standard can significantly increase
measurement consistency when multiple instruments are used to characterize the
same protein particle suspension.
Similar to MFI, a micro-flow digital imaging (MDI) method was developed,
optimized, and qualified to detect and monitor protein particles in a high-
concentration IgG1 monoclonal antibody formulation by Wuchner et al. (2010).
The MDI method was qualified as a characterization assay with respect to accuracy
and precision of particle size and number using polystyrene standards (5–300 μm)
and protein particles (2 to >100 μm). The authors then used it to monitor the stabil-
ity profile of a 90 mg/mL IgG1 formulation stored at 2–8°C and −70°C for up to 18
months. The MDI assay results showed improved sensitivity in detecting subvisible
particles (≥5 μm) compared to the conventional LO-based method and a good over-
all correlation with the amount of visible protein particles (>70 μm). The MDI
results showed an accumulation of protein particles in certain size ranges and an
increase in the overall particle size distribution over time. The combined MDI
results and protein characterization studies have provided an enhanced understand-
ing of protein particulate formation during long-term storage of a high-concentration
IgG1 monoclonal antibody formulation.
In summary, the LO- and MFI-based methods can and should be qualified for
SbVP analysis in specific protein solutions. The methods offer sensitive and
reproducible measurement of particles in protein therapeutics solutions which
enables proper monitoring and control of SbVP levels during protein therapeutics
development, manufacturing, transportation, storage, and delivery.
5.8 Conclusion
The qualification of the biophysical methods including CD, FTIR, DSC, SV-AUC,
SE-HPLC-LS, and LO- and MFI-based methods has been reviewed. Consistent
instrument calibration and maintenance and proper sample handling are required to
ensure successful qualification of these methods for protein higher-order structure
and size distribution analysis. A numerical method was applied to compare the
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 121
overall spectra or profile similarity for the qualification of some of the methods such
as CD and FTIR. AUC and SE-HPLC-LS are first principle methods; therefore, the
qualification of these methods does not require the use of a standard. However, for
the proper qualification of the LO- and MFI-based methods, suitable particle
standards that closely resemble the optical properties and morphologies of protein
particles are still needed. For all of the biophysical methods, qualification is focused
on the precision of the methods and demonstration that the methods are suitable for their
intended applications. Successful qualification enables the understanding of the
method capability and reliable determination and control of product attributes.
Therefore, qualification of biophysical methods should be undertaken for protein
therapeutics to ensure the best possible knowledge of the product and its
manufacturing process.
References
Chan CC, Lam H, Zhang XM (2011) Practical approaches to method validation and essential
instrument qualification. Wiley, New York
Chiu MH, Prenner EJ (2011) Differential scanning calorimetry: an invaluable tool for a detailed
thermodynamic characterization of macromolecules and their interactions. J Pharm Bioallied
Sci 3(1):39–59
Coligan JE et al (2001) Measuring protein thermostability by differential scanning calorimetry.
In: Makhatadze GI (ed) Current protocols in protein science/editorial board. Wiley, New York
Council of Europe (2007) Particulate contamination: sub-visible particles, general chapter 2.9.19.
Council of Europe, Strasbourg
Cover TM, Hart PE (1967) Nearest neighbor pattern classification. IEEE Trans Inf Theory
IT-13(1):21–27
D’antonio J, Murphy BM, Manning MC, Al-Azzam WA (2012) Comparability of protein thera-
peutics: quantitative comparison of second-derivative amide I infrared spectra. J Pharm Sci.
doi:10.1002/jps.23133
Dam J et al (2004) Calculating sedimentation coefficient distributions by direct modeling of sedi-
mentation velocity concentration profiles. In: Methods in enzymology. Academic.
pp 185–212
Drechsler A, Miles AJ, Norton RC, Wallace BA, Separovic F (2009) Effect of lipid on the confor-
mation of the n-terminal region of equinatoxin II: a synchrotron radiation CD study. Eur
Biophys J 39:121–127
Driscoll DF (2004) Examination of selection of light-scattering and light-obscuration acceptance
criteria for lipid injectable emulsions. Pharm Forum 30:2244–2253
Fu K, Griebenow K, Hsieh L, Klibanov AM, Langer R (1999) FTIR characterization of the sec-
ondary structure of proteins encapsulated within PLGA microspheres. J Controlled Release
58:357–366
Gabrielson JP, Arthur KK (2011) Measuring low levels of protein aggregation by sedimentation
velocity. Methods 54(1):83–91
Gabrielson JP et al (2007a) Quantitation of aggregate levels in a recombinant humanized monoclo-
nal antibody formulation by size-exclusion chromatography, asymmetrical flow field flow frac-
tionation, and sedimentation velocity. J Pharm Sci 96(2):268–279
Gabrielson JP et al (2007b) Sedimentation velocity analytical ultracentrifugation and SEDFIT/
c(s): limits of quantitation for a monoclonal antibody system. Anal Biochem 361:24–30
Gabrielson JP, Arthur KK, Stoner MR et al (2010) Precision of protein aggregation measurements
by sedimentation velocity analytical ultracentrifugation in biopharmaceutical applications.
Anal Biochem 396(2):231–241
Gmelin E, Sarge St.M (1995) Calibration of differential scanning calorimeters. IUPAC, Pure Appl
Chem 67:1789–1800
Greenfield NJ (1996) Methods to estimate conformation of proteins and polypeptides from CD
data. Anal Biochem 235(1):1–10
Greenfield NJ, Fasman GD (1969) Computed circular dichroism spectra for the evaluation of pro-
tein conformation. Biochemistry 8:4108–4116
Gross PC, Zeppezauer M (2010) Infrared spectroscopy for biopharmaceutical protein analysis.
J Pharm Biomed Anal 53:29–36
Haines-Nutt RF, Munton TJ (1984) Particle size measurement in intravenous fluids. J Pharm
Pharmacol 36(8):534–536
Haris PI, Chapman D (1995) The conformational analysis of peptides using Fourier transform IR
spectroscopy. Biopolym (Pept Sci) 37:251–263
Huang CT, Sharma D, Oma P, Krishnamurthy R (2009) Quantitation of protein particles in paren-
teral solutions using micro-flow imaging. J Pharm Sci 98(9):3058–3071
ICH M4Q (2001) The common technical document for the registration of pharmaceuticals for
human use: quality. US Department of Health and Services, FDA, CDER
ICH Q2(R1) (2005) Validation of analytical procedures: test and methodology. In: ICH Harmonised
Tripartite guideline in international conference on harmonisation of technical requirements for
registration of pharmaceuticals for human use
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 123
Lopez MO, Freire E (1987) Dynamic analysis of differential scanning calorimetry data. Biophys
Chem 27(1):87–96
Mahler HC, Friess W, Grauschopf U, Kies S (2009) Protein aggregation: pathways, induction fac-
tors and analysis. J Pharm Sci 98:2909–2934
Manning MC, Woody RW (1989) Theoretical study of the contribution of aromatic side chains to
the circular dichroism of basic bovine pancreatic trypsin inhibitor. Biochemistry
28:8609–8613
Meersman F, Atilgan C, Miles AJ, Bader R, Shang W, Matagne A, Wallace BA, Koch MHJ (2010)
Consistent picture of the reversible thermal unfolding of hen egg-white lysozyme from experi-
ment and molecular dynamics. Biophysical J 99:2255–2263
Miles AJ, Wallace BA (2006) Synchrotron radiation circular dichroism spectroscopy of proteins
and applications in structural and functional genomics. Chem Soc Reviews 35:39–51
Miles AJ, Wien F, Lees JG, Rodger A, Janes RW, Wallace BA (2003) Calibration and standardisa-
tion of synchrotron radiation circular dichroism (SRCD) amplitudes and conventional circular
dichroism (CD) spectrophotometers. Spectroscopy 17:653–661
Miles AJ, Wien F, Wallace BA (2004) Redetermination of the extinction coefficient of camphor-b-
sulfonic acid, A calibration standard for circular dichroism spectroscopy. Anal Biochem
335:338–339
Miles AJ, Wien F, Lees JG, Wallace BA (2005) Calibration and standardisation of synchrotron
radiation and conventional circular dichroism spectrometers. Part 2: factors affecting magni-
tude and wavelength. Spectroscopy 19:43–51
Miles AJ, Hoffman SV, Tao Y, Janes RW, Wallace BA (2007) Synchrotron radiation circular
dichroism (SRCD) spectroscopy: new beamlines and new applications in biology. Spectroscopy
21:245–255
Narhi LO, Jiang Y, Cao S, Benedek K, Shnek D (2009) A critical review of analytical methods for
subvisible and visible particles. Curr Pharm Biotechnol 10(4):373–381
Pekar A, Sukumar M (2007) Quantitation of aggregates in therapeutic proteins using sedimenta-
tion velocity analytical ultracentrifugation: practical considerations that affect precision and
accuracy. Anal Biochem 367:225–237
Perczel A, Fasman GD (1993) Effect of spectral window size on circular dichroism spectra decon-
volution of proteins. Biophys Chem 48:19–29
Perczel A, Hollosi M, Tusnady G, Fasman GD (1991) Convex constraint analysis: a natural decon-
volution of circular dichroism curves of proteins. Protein Eng 4:669–679
Philo JS (2006) Is any measurement method optimal for all aggregate sizes and types? AAPS J
8(3):564–571
Powl AM, O'Reilly AO, Miles AM, Wallace BA (2010) Synchrotron radiation circular dichroism
spectroscopy-defined structure of the C-terminal domain of NaChBac and its role in channel
assembly. Proc Natl Acad Sci 107:14064–14069
Prestrelski S, Tedeschi N, Arakawa T, Carpenter J (1993) Biophys J 65(2):661–671
Protasevich I, Yang Z, Wang C, Atwell S, Zhao X, Emtage S, Wetmore D, Hunt JF, Brouillette CG
(2010) Thermal unfolding studies show the disease causing F508del mutation in CFTR ther-
modynamically destabilizes nucleotide-binding domain 1. Protein Sci 19(10):1917–1931
Pyrpassopoulos S, Vlassi M, Tsortos A, Papanikolau Y, Petratos K, Vorgias CE, Nounesis G
(2006) Equilibrium heat-induced denaturation of chitinase 40 from Streptomyces thermoviola-
ceus. Proteins 64(2):513–523
Ravi J, Rakowska PD, Garfagnini T et al (2010) International comparability in spectroscopic
measurements of protein structure by circular dichroism: CCQM-P59.1, Metrologia 47(6):
631–641
Ravi J, Schiffmann D, Tantra R et al (2010) International comparability in spectroscopic measure-
ments of protein structure by circular dichroism: CCQM P59, Metrologia 47:1A
Schiffmann D, Butterfield DM, Yardley RE, Knight A, Windsor SA, Jones C (2004) Val-CiD
Appendix A: CD spectroscopy: an inter-laboratory study. DQL-AS 009 ISSN: 1744–0602
Schmid M (2012) Precision of DSC measurements: results of interlaboratory tests. http://www.
eurostar-science.org/MAP1/Schmid.pdf. Accessed 7 Feb 2012
5 Qualification of Biophysical Methods for the Analysis of Protein Therapeutics 125
Wen J, Arthur K, Chemmalil L, Muzammil S, Gabrielson JP, Jiang Y (2011) Applications of dif-
ferential scanning calorimetry for thermal stability analysis of proteins: qualification of DSC.
J Pharm Sci. doi:10.1002/jps.22820
Whitmore L, Wallace BA (2004) DICHROWEB, An online server for protein secondary structure
analyses from circular dichroism spectroscopic data. Nucleic Acids Res 32:668–673
Whitmore L, Wallace BA (2008) Protein secondary structure analyses from circular dichroism
spectroscopy: methods and reference databases. Biopolymers 89:392–400
Woody RW (1994) Contributions of tryptophan side chains to the far-ultraviolet circular dichroism
of proteins. Eur Biophys J 23:253–262
Woody RW (1996) Theory of circular dichroism of proteins, circular dichroism and the conforma-
tional analysis of biomolecules. In: Fasman GD (ed) Circular dichroism and the conforma-
tional analysis of biomolecules. Plenum, New York
Wuchner K, Büchler J, Spycher R, Dalmonte P, Volkin DB (2010) Development of a microflow
digital imaging assay to characterize protein particulates during storage of a high concentration
IgG1 monoclonal antibody formulation. J Pharm Sci 99(8):3343–3361
Wyatt PJ (1993) Light scattering and the absolute characterization of macromolecules. Anal Chim
Acta 272:1–40
Yang JT, Wu CSC, Martinez HM (1986) Calculation of protein conformation from circular dichro-
ism. Methods Enzymol 130:208–269
Yoshii K (1997) Application of differential scanning calorimetry to the estimation of drug purity:
various problems and their solutions in purity analysis. Chem Pharm Bull 45(2):338–343
Zölls S, Tantipolphan R, Wiggenhorn M, Winter G, Jiskoot W, Friess W, Hawe A (2011) Particles
in therapeutic protein formulations, Part 1: overview of analytical methods. J Pharm Sci.
doi:10.1002/jps.23001
Chapter 6
Application of Biophysics to the Early
Developability Assessment of Therapeutic
Candidates and Its Application to Enhance
Developability Properties
For the past few decades, both the cost and the complexity of therapeutic protein
development have soared to levels that are becoming unsustainable (Adams and
Brantner 2010). Additionally, there is a significant increase in the attrition rate of
programs that can be attributed to, for example, complex disease pathways, an
increased safety bar, and a greater number of highly complex molecules in develop-
ment (Kola and Landis 2004). According to a recent analysis, the success rate for IND
to market is approximately 25–30% for biotech companies (DiMasi and Grabowski
2007). Both the spiraling cost associated with drug development and the demands
from end users for low-cost prescription drugs are forcing the industry to look for
ways to increase efficiencies. One approach is cycle time reduction, a strategy that
translates to a reduced and more efficient development timeline so that a drug can
move faster to phase I and faster to market. This chapter focuses on current and future
challenges in developing therapeutic biological drugs and how the use of novel bio-
physical tools to screen for potential developability and manufacturability risks can
improve product quality, streamline development, increase manufacturing efficiency,
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 127
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2_6,
© Springer Science+Business Media New York 2013
128 H. Sathish et al.
and increase end-user convenience. Apart from selecting low-risk molecules, early
developability screening can also aid in identification of platform compatible leads,
which can significantly reduce both development cost and time.
More novel target classes, such as G-protein-coupled receptors (GPCRs), ion chan-
nels, and intracellular targets may not be evolutionarily ideally suited for antibody
therapy. Therefore, these targets may require the development of therapeutic mole-
cules that are inherently more novel than the prototypical IgG, such as alternative
protein fold scaffolds or engineered receptor ligands. Recent advances in protein
expression and production techniques, as well as improved in vitro libraries and
selection strategies, have started to expand the coverage of these targets that are
implicated in a range of different diseases (Lundstrom 2005; Ackerman and
Clapham 1997; Naylor and Beech 2009). However, the types of therapeutic mole-
cules (discussed in Sect. 6.2.2) that need to be engineered to bind these novel targets
are far less well characterized in terms of large-scale development and manufacture
than prototypical mAbs and are trending towards being highly complex, resulting in
a greater attrition rate during discovery and development (Ma and Zemmel 2002).
The field of antibody engineering has matured rapidly over the past decade so that
properties such as affinity, potency, specificity, pharmacokinetics, and effector
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 129
function are now readily tunable (Igawa et al. 2011). Also, the pharmaceutical
industry is increasingly focusing on engineering molecules with novel mechanisms
of action (Ma and Zemmel 2002).
For molecules with an established mechanism of action, the focus is on enhanced
safety, higher affinity to target, and increased half-life. The first generation of mAbs
to reach the market typically exhibited dissociation constants (Kds) in the micromo-
lar range, in line with the in vivo ceiling of the B-cell response (Foote and Eisen
1995; Batista and Neuberger 1998). In vitro affinity maturation methods and tech-
nologies [reviewed in Igawa et al. (2011)] have allowed researchers to go far beyond
these limits, reaching femtomolar Kds in many cases (Boder et al. 2000; Finch et al.
2011). Increasing the affinity of an antibody to a given antigen is achieved by target-
ing the variable domains, as these are regions of the molecule that have evolved to
interact directly with the antigen, particularly the complementarity determining
regions (CDRs). Affinity maturation typically results in several mutations in the
CDR loops. As the CDR loops are surface exposed, engineering changes may result
in changes to biophysical behavior, such as an increase in aggregation propensity
and loss of stability (Chennamsetty et al. 2009a; Pepinsky et al. 2010).
Over the last decade, there has been an increased effort in glycoengineering and
the engineering of Fc variants in order to modulate effector functions [reviewed in
Kubota et al. (2009)] and improve pharmacokinetic properties through modified
interaction with the neonatal Fc receptor (Presta 2008; Stewart et al. 2011;
Dall’Acqua et al. 2006a, b). Glycosylation in IgG molecules is essential for struc-
tural integrity, solubility, and stability (Krapp et al. 2003; Mimura et al. 2000;
Kayser et al. 2011). Any modulation of glycosylation can, in turn, have significant
impact on solubility, manufacturability, and long-term storage. Likewise, engi-
neering the Fc region of the IgG molecule to modulate effector function or phar-
macokinetics may have unintended effects on the biophysical properties of the
molecule, such as stability and aggregation propensity. The unintended effects of
glycoengineering and the engineering of Fc variants may not become apparent
until larger-scale manufacture and formulation to support clinical studies.
The search for molecules with novel mechanisms of action has resulted in an
increased development of antibody-based alternative formats, such as multivalent
antibodies (Carter 2011). Recently, a new set of protein therapeutics based on alter-
native protein scaffolds to immunoglobulin domains has entered clinical develop-
ment (Gebauer and Skerra 2009). Each of these novel therapeutics is likely to
present their own challenges for development, which will doubtless become appar-
ent over the coming decade.
Chemistry, manufacturing, and controls (CMC) design spaces are another area of
challenge in the commercialization of protein therapeutics. For molecules with a
precedented mechanism of action, enhanced product performance can be achieved
130 H. Sathish et al.
As explained in Sect. 6.2.2, there is greater focus at the discovery stage on obtaining
attributes such as higher affinity, enhanced safety, and increased half-life, which
translates to highly complex, often novel molecules. From the development per-
spective, there is an increased demand to enhance product performance and patient
comfort by use of a high-concentration liquid formulation in specialty devices, with
an additional burden of decreasing the overall development timeline. Taken together,
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 131
the objective of quick and cheap screening. Also key is better predictive strategies,
which should have the ability to predict the behavior of the protein during manufac-
turing and storage.
A typical early screen for a mAb might initially focus on a property, such as speci-
ficity or strength (affinity) of binding. Subsequent secondary screens and assays
typically include specificity against related molecules to the target and binding to
species homologues (to support preclinical safety studies). In this way a “screening
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 133
6.6.1 Aggregation
The propensity for a molecule to aggregate is one of the single most influential and
challenging factors for formulation development. Candidates that are inherently
resistant towards aggregation are considered superior. Nonnative aggregation is a
multistep process and can be explained by the Lumry–Eyring framework. Each
state has a different molecular configuration and energy, and each reaction step
proceeds through an energy barrier (ΔG*, difference in energy or activation free
energy). The step that has highest ΔG* is the rate-limiting step (Chi et al. 2003a).
The first step in this process, which is a reversible process, gives rise to an
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 135
with thermal ramping (reviewed in Chap. 2). These techniques can provide through-
put on the order of 96–384 samples per hour and typically only require 10s of
micrograms. Aggregation and particle formation can also be directly monitored by
techniques such as dynamic light scattering (DLS), HIAC light obscuration, micro-
flow imaging (MFI), and electrical charge-based methods (e.g., Coulter counter). Of
these methods, DLS offers the greatest flexibility for early-stage use based on
throughput (plate-based DLS) and material required. However, recent advances
have resulted in improvements in particle detection including scanning ion occlu-
sion spectroscopy (qNano), chip-based mass measurements (Archimedes), and
nanoparticle tracking analysis (NanoSight). Each of these methods offers the pos-
sibility of increased sensitivity and throughput with a decrease in sample consump-
tion, resulting in more screening possibilities.
As shown in Table 6.1, proteins have multiple liabilities that are prone to chemical
degradation during manufacture and storage. The rate of degradation depends on
both the solution condition and the structure of the molecule (Sinha et al. 2009;
Vlasak and Ionescu 2011). Typically some of the most common liabilities like
deamidation, nonenzymatic hydrolysis, and oxidation are highlighted very early and
easily by in silico primary sequence liability analysis. Theoretically, molecules with
the fewest chemical liabilities will allow one to develop a wider process and formu-
lation design space. For molecules with identified liabilities, there will be an addi-
tional burden of determining the effect of the liabilities on product quality and
designing appropriate process and formulation controls. Unlike physical stability
studies, the slow rates of reaction associated with accessing chemical stability results
in time-consuming forced degradation studies. For example, evaluation of the effect
of deamidation on function and stability requires forced degradation studies at ele-
vated temperature and pH, followed by liquid chromatography–mass spectrometry
(LCMS) and subsequent selected ion monitoring (SIM). Similarly fragmentation
propensity can be determined either using size-based separation techniques or side-
chain chemistry-based analytical methods like reversed-phase high-performance
liquid chromatography (RPHPLC). Even though HT analytics are available to fol-
low fragmentation, the slow rate of reaction of forced degradation studies that can
accurately predict nonenzymatic fragmentation under storage conditions typically
requires time-consuming incubation studies in the range of 4–6 weeks.
Other important developability criteria that are often neglected during early screen-
ing are the solution properties of a molecule at high concentration. Solution properties
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 137
such as solubility, phase separation, and viscosity can often become the rate-limiting
step during processing, filtration, drug delivery, etc., as solution properties of a mol-
ecule are governed by various attractive and repulsive nonspecific intermolecular
interactions [reviewed in Saluja and Kalonia (2008) and Tadros (2011)]. Under con-
centrated solution conditions, the intermolecular distance becomes less than the
molecule’s diameter. This reduction in intermolecular distance will not only increase
the collision frequency but will also increase the duration of such an encounter
(Scherer et al. 2010; Minton 2005). Thus at high concentrations, the nature of, and
relative contributions from, various nonspecific interactions is dictated by both solu-
tion conditions and the surface properties of a molecule, including net charge,
charge distribution, and surface hydrophobicity. From a development perspective,
the consequences of such unfavorable interactions are a change in the protein’s
solution behavior potentially resulting in issues such as poor solubility, phase sepa-
ration, and/or increased viscosity (Saluja and Kalonia 2008; Chari et al. 2009).
Typically, such unfavorable interactions can be controlled during development by
modulating solution conditions like ionic strength and/or pH to develop appropriate
formulation controls (Salinas et al. 2010; Yadav et al. 2012; Liu et al. 2005), but in
some cases, depending on strength of intermolecular interactions, it may not be
possible to achieve this control. Quite often the solution conditions might have an
unintended opposite effect on various solution properties and overall stability
(Salinas et al. 2010).
The rheological properties of a molecule under a given buffer system have typi-
cally been ascertained using a cone- and plate-type instrument to determine viscos-
ity. The dimensions of the cone and plate instruments have been dramatically
reduced over the years, resulting in a correlating decrease in the volume of material
required. However, the time associated with these measurements has not fallen to
the same degree. New techniques such as viscosity–DLS (see He et al., Chap. 2, this
volume) and chip-based measurements provide an attractive alternative to the tradi-
tional cone and plate approach. Alternatively, solution properties can also be
predicted by accurate measurements of net protein–protein interactions through col-
loidal stability (Yadav et al. 2012; George and Wilson 1994; Mehta et al. 2012). The
osmotic second virial coefficient (B22), which is a thermodynamic parameter, is often
used to directly quantify overall protein–protein interactions (Chi et al. 2003a, b;
Valente et al. 2005; Payne et al. 2006). A positive B22 value indicates an overall
dominance of repulsive interactions, where as a negative B22 value indicates an over-
all attractive intermolecular interactions. The B22 of a colloidally active system can
be accurately measured by various techniques like DLS (Yadav et al. 2011), static
light scattering (SLS), analytical ultracentrifugation (AUC), osmometry, and self-
interaction chromatography (SIC) (Le Brun et al. 2010).
One of the main reasons for limited screening of these undesirable solution prop-
erties during early development is the limited availability of HT techniques. Most of
the available techniques mentioned above are low throughput in nature and also
require large amounts of material. But recently significant progress has been made
in this area, and several HT methods are available to either directly measure some
of the solution properties (reviewed in Chap. 2) or accurately predict solution
138 H. Sathish et al.
properties (Johnson et al. 2009; Bajaj et al. 2007; Sule et al. 2011; Tessier et al.
2008). Availability of such HT techniques allows one to screen and select candidate
molecules with desired surface properties or interaction potential, which is a mole-
cule that has a B22 value close to 0 under the desired solution conditions. Applying
an approach of selecting or engineering a molecule with desired surface properties,
or interaction potential, will enable the formulation scientist to focus mostly on
aggregation or other chemical degradations, thus widening the formulation develop-
ment design space.
One consequence of such an approach is that there is a potential for conflict
between the different parameters that are being simultaneously screened. For exam-
ple, if antibody affinity is found to correlate with an unfavorable biophysical char-
acteristic, such as aggregation propensity, careful consideration would need to be
given as to which characteristic would be most economic to engineer/mitigate
against the aggregation propensity. Examples of candidate therapeutic antibodies
that have been extensively reengineered to improve biophysical properties and as
such improve developability have begun to be described (Pepinsky et al. 2010).
As stated in Sect. 6.2, for molecules with a precedented mechanism of action, the
focus is on enhanced safety, higher affinity to target, and increased half-life. The
efficacy and potency of a molecule can be fine-tuned by efficient mediation of effec-
tor functions (Anderson et al. 1997; Weng and Levy 2003). This case study demon-
strates how engineering of Fc variants to modulate these activities can have an
unintentional effect on a molecule’s developability and how this unintentional effect
can be screened for and detected early by applying appropriate biophysical tools. In
this case a triple mutant S239D/A330L/I332E (resulting in mAbA-3M) was intro-
duced into the CH2 domain of a human mAb (mAbA). Compared to its wild type,
these mutations resulted in an approximately 10–100 fold increase in mAbA-3M
binding to its receptor, but at a price of reduced stability (Fig. 6.1) (Oganesyan et al.
2008; Anandakumar et al. 2008). Subsequent developability-accelerated stability
studies indicated a significant increase in the aggregation rate at 40 °C (Fig. 6.2).
Based on various biophysical studies like DSC and denaturant unfolding, it was
speculated that the reduced conformational stability was due to minor structural
perturbations in the region around the CH2 domain. By solving the structure of the
mutant, Oganesyan et al. determined that the mutations resulted in opening of the
cleft in the CH2 domain of the Fc fragment (Oganesyan et al. 2008). They concluded
that the enhanced interactions with human Fcγ(gamma)RIIIA could be due to
enhanced Fc openness as well as the introduction of hydrophobic and electrostatic
interactions at the corresponding interface (Oganesyan et al. 2008). Even though
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 139
Case Study 2 examines three protein candidates to ascertain if any have potential
liabilities with respect to manufacturing or formulation. Three IgG2 monoclonal
140 H. Sathish et al.
Fig. 6.2 Comparison of stability of mAbA and its 3M mutant obtained under accelerated condi-
tions. Stability of both proteins was monitored by incubating the protein at 50 mg/mL at 40 °C and
monitoring the changes in purity and aggregation by HPSEC
antibody (mAb) candidates against the same target, molecule H, molecule J, and
molecule K, were provided for assessment. Using biophysical and accelerated incu-
bation studies, these molecules were screened to see if they could adhere to a for-
mulation and processing platform.
DSC was used to monitor the thermal stability of the candidates in the formulation
at 1 mg/mL. At pH 5.2, all the candidates exhibited an apparent Fab Tm between
77 °C and 80 °C. The observation of relatively higher Fab Tm value indicated lower
stability/developability risk. This was further confirmed by an incubation study per-
formed at 70 mg/mL mAb concentration in an acetate pH 5.2 buffer. Samples were
incubated in the buffer and stored at 4 °C, 25 °C, or 37 °C for 2–4 weeks. For sam-
ples stored at 37 °C, all candidates exhibited similar increases in HMW (high
molecular weight) species (between 0.3% and 0.4% increase from t = 0) over a
period of 4 weeks. As expected, no significant change in the SE-HPLC % main peak
(MP) was observed for any of the three candidates upon storage at 4 °C for 4 weeks
or at 25 °C for 2 weeks. To simulate the pH jump upon physiological administra-
tion, candidates were subjected to pH change from pH 5.2 to ~7.0 and held for 24 h
at 37 °C. An increase in % HMW species was noted for all the candidates after 24 h
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 141
To test the compatibility with the possible solution conditions encountered during
purification, the candidates were exposed to representative buffers at pH values
between 3.0 and 7.4. Potential changes in thermal stability, higher-order structure,
and resulting degradation were monitored by applying various biophysical tools and
SE-HPLC analysis. At pH 7.4 as expected the DSC thermogram showed three endo-
therms corresponding to the unfolding of the CH2, Fab, and CH3 domains. However,
at pH 3.0, only a single endotherm was observed for all candidates indicating either
increased inter-domain interactions (cooperative unfolding) or unfolding of some of
the domains (even at temperatures below the starting point of the DSC analysis) at
low pH. The conformation of molecule K appeared to be the most stable, as indi-
cated by the higher Tm of the Fab domain at pH 3.5 and 5.0, followed by molecule
J and molecule H. When reversed into PBS by dialysis, no significant irreversible
change was observed for any of the three candidates, and the DSC thermograms
were superimposable.
The tertiary structure of the candidates was studied by near-UV CD. The spectra
of the three candidates exhibited similar features, with varying intensities of mean
residue ellipticity (MRE). The spectra for all candidates at pH 3.5 and 5.0 were simi-
lar suggesting that no conformational changes occurred at pH 3.5. At pH 3.0, the
near-UV CD spectrum of all candidates showed a marked loss in MRE that indicated
the protein molecules had lost a significant amount of tertiary structure at this pH.
Most of the loss in the near-UV CD signal occurred in the region between 250 nm
and 280 nm suggesting a marked change in the disulfide environment at pH 3.0. The
near-UV CD spectra of the reversibility samples in PBS overlay very well with the
protein in pH 5 reversed into PBS at pH 7.2 indicating that there were no irreversible
changes in the tertiary structure of any of the candidates induced by exposure to
lower pH. The effect of low pH on the protein conformation and its reversibility
were further evaluated by ANS-binding studies, which revealed an increase in pro-
tein surface hydrophobicity with a decrease in pH. Upon reversing the solution con-
dition to PBS, all candidates exhibited full reversibility in surface hydrophobicity.
Implications of low pH-induced structural perturbation on candidate stability
were evaluated by incubating the candidates at different pH. None of the candidates
exhibited a decrease in % main peak by SE-HPLC after 6 h of exposure to actual
solution pH values (pH 3.7, 5.0 and 7.4). Whereas, when exposed to pH 3.0, all
candidates exhibited aggregation and a loss in % main peak determined by
SE-HPLC. Molecule K exhibited the most aggregation (84%) followed by molecule
H (78%) and molecule J (77%). The results obtained by SEC were further supported
by DLS studies, which showed the presence of HMW species at pH 3.0 for all three
candidates. At pH 3.5 and 5.0, the size distribution was superimposable and the
hydrodynamic diameter (dH) was close to that of a monomer (approximately
142 H. Sathish et al.
10–11 nm), suggesting a more monomer-like species at higher solution pH. The
polydispersity index (PdI, an indicator of size heterogeneity in the sample with
lower values <0.1 suggesting a homogeneous population) also showed a decreasing
trend from pH 3.0 to 5.0. For samples reversed into PBS, a decrease in size distribu-
tion was not observed in the pH 3.0 samples indicating that some irreversible
changes were produced. However, full reversibility was observed for samples
exposed to pH values of 3.5 and 5 and then dialyzed into PBS, consistent with
observations made from the other biophysical techniques tested during this study.
This result was expected since the samples at pH 3.5 and 5.0 exhibited a homoge-
neous size distribution.
Biophysical and stability data obtained during early developability screening
indicated all three candidates were low-risk molecules from a developability/
manufacturability perspective and adhered to the formulation and processing
platforms.
6.8 Conclusion
Acknowledgements Authors wish to thank Nancy Craighead at MedImmune for critical review
of this chapter.
References
Ackerman MJ, Clapham DE (1997) Ion channels–basic science and clinical disease. N Engl J Med
336(22):1575–1586
Adams CP, Brantner VV (2010) Spending on new drug development1. Health Econ
19(2):130–141
Anandakumar R, Harn NR, Sathish HA, LeachWL, Oliver CN, Bishop SM (2008) American
Association of Pharmaceutical Scientist, National Biotech Conference, 2008
Anderson DR, Grillo-Lopez A, Varns C, Chambers KS, Hanna N (1997) Targeted anti-cancer
therapy using rituximab, a chimaeric anti-CD20 antibody (IDEC-C2B8) in the treatment of
non-Hodgkin’s B-cell lymphoma. Biochem Soc Trans 25(2):705–708
Bajaj H, Sharma VK, Kalonia DS (2007) A high-throughput method for detection of protein self-
association and second virial coefficient using size-exclusion chromatography through simul-
taneous measurement of concentration and scattered light intensity. Pharm Res
24(11):2071–2083
Batista FD, Neuberger MS (1998) Affinity dependence of the B cell response to antigen: a thresh-
old, a ceiling, and the importance of off-rate. Immunity 8(6):751–759
Bee JS, Stevenson JL, Mehta B, Svitel J, Pollastrini J, Platz R, Freund E, Carpenter JF, Randolph
TW (2009) Response of a concentrated monoclonal antibody formulation to high shear.
Biotechnol Bioeng 103(5):936–943
Bekard IB, Asimakis P, Bertolini J, Dunstan DE (2011) The effects of shear flow on protein struc-
ture and function. Biopolymers 95(11):733–745
Biddlecombe JG, Craig AV, Zhang H, Uddin S, Mulot S, Fish BC, Bracewell DG (2007)
Determining antibody stability: creation of solid–liquid interfacial effects within a high shear
environment. Biotechnol Prog 23(5):1218–1222
Boder ET, Midelfort KS, Wittrup KD (2000) Directed evolution of antibody fragments with mon-
ovalent femtomolar antigen-binding affinity. Proc Natl Acad Sci USA 97(20):10701–10705
Carter PJ (2011) Introduction to current and future protein therapeutics: a protein engineering
perspective. Exp Cell Res 317(9):1261–1269
Cellmer T, Bratko D, Prausnitz JM, Blanch HW (2007) Protein aggregation in silico. Trends
Biotechnol 25(6):254–261
Chan AC, Carter PJ (2010) Therapeutic antibodies for autoimmunity and inflammation. Nat Rev
Immunol 10(5):301–316
Chari R, Jerath K, Badkar AV, Kalonia DS (2009) Long- and short-range electrostatic interactions
affect the rheology of highly concentrated antibody solutions. Pharm Res 26(12):2607–2618
Chennamsetty N, Voynov V, Kayser V, Helk B, Trout BL (2009a) Design of therapeutic proteins
with enhanced stability. Proc Natl Acad Sci USA 106(29):11937–11942
Chennamsetty N, Helk B, Voynov V, Kayser V, Trout BL (2009b) Aggregation-prone motifs in
human immunoglobulin G. J Mol Biol 391(2):404–413
Chi EY, Krishnan S, Randolph TW, Carpenter JF (2003a) Physical stability of proteins in aqueous
solution: mechanism and driving forces in nonnative protein aggregation. Pharm Res
20(9):1325–1336
Chi EY, Krishnan S, Kendrick BS, Chang BS, Carpenter JF, Randolph TW (2003b) Roles of
conformational stability and colloidal stability in the aggregation of recombinant human granu-
locyte colony-stimulating factor. Protein Sci 12(5):903–913
Dall’Acqua WF, Cook KE, Damschroder MM, Woods RM, Wu H (2006a) Modulation of the
effector functions of a human IgG1 through engineering of its hinge region. J Immunol
177(2):1129–1138
144 H. Sathish et al.
Dall’Acqua WF, Kiener PA, Wu H (2006b) Properties of human IgG1s engineered for enhanced
binding to the neonatal Fc receptor (FcRn). J Biol Chem 281(33):23514–23524
DiMasi JA, Grabowski HG (2007) The cost of biopharmaceutical R&D: is biotech different?
Manag Decis Econ 28(4–5):469–479
Famm K, Hansen L, Christ D, Winter G (2008) Thermodynamically stable aggregation-resistant
antibody domains through directed evolution. J Mol Biol 376(4):926–931
Finch DK, Sleeman MA, Moisan J, Ferraro F, Botterell S, Campbell J, Cochrane D, Cruwys S,
England E, Lane S, Rendall E, Sinha M, Walker C, Rees G, Bowen MA, Schneider A, Liang
M, Faggioni R, Fung M, Mallinder PR, Wilkinson T, Kolbeck R, Vaughan T, Lowe DC (2011)
Whole-molecule antibody engineering: generation of a high-affinity anti-IL-6 antibody with
extended pharmacokinetics. J Mol Biol 411(4):791–807
Foote J, Eisen HN (1995) Kinetic and affinity limits on antibodies produced during immune
responses. Proc Natl Acad Sci USA 92(5):1254–1256
Fowler SB, Poon S, Muff R, Chiti F, Dobson CM, Zurdo J (2005) Rational design of aggregation-
resistant bioactive peptides: reengineering human calcitonin. Proc Natl Acad Sci USA
102(29):10105–10110
Gebauer M, Skerra A (2009) Engineered protein scaffolds as next-generation antibody therapeu-
tics. Curr Opin Chem Biol 13(3):245–255
George A, Wilson WW (1994) Predicting protein crystallization from a dilute solution property.
Acta Crystallogr D Biol Crystallogr 50(Pt 4):361–365
Goldberg DS, Bishop SM, Shah AU, Sathish HA (2011) Formulation development of therapeutic
monoclonal antibodies using high-throughput fluorescence and static light scattering tech-
niques: role of conformational and colloidal stability. J Pharm Sci 100:1306–1315
Hageman MJ (2006) Solubility, solubilization and dissolution in drug delivery during lead optimi-
zation. Springer, New York
He F, Becker GW, Litowski JR, Narhi LO, Brems DN, Razinkov VI (2010) High-throughput
dynamic light scattering method for measuring viscosity of concentrated protein solutions.
Anal Biochem 399(1):141–143
Igawa T, Tsunoda H, Kuramochi T, Sampei Z, Ishii S, Hattori K (2011) Engineering the variable
region of therapeutic IgG antibodies. MAbs 3(3):243–252
Jespers L, Schon O, Famm K, Winter G (2004) Aggregation-resistant domain antibodies selected
on phage by heat denaturation. Nat Biotechnol 22(9):1161–1165
Johnson DH, Parupudi A, Wilson WW, DeLucas LJ (2009) High-throughput self-interaction chro-
matography: applications in protein formulation prediction. Pharm Res 26(2):296–305
Karlsson M, Ekeroth J, Elwing H, Carlsson U (2005) Reduction of irreversible protein adsorption
on solid surfaces by protein engineering for increased stability. J Biol Chem
280(27):25558–25564
Kayser V, Chennamsetty N, Voynov V, Forrer K, Helk B, Trout BL (2011) Glycosylation influ-
ences on the aggregation propensity of therapeutic monoclonal antibodies. Biotechnol
J 6(1):38–44
Kola I, Landis J (2004) Can the pharmaceutical industry reduce attrition rates? Nat Rev Drug
Discov 3(8):711–715
Krapp S, Mimura Y, Jefferis R, Huber R, Sondermann P (2003) Structural analysis of human
IgG-Fc glycoforms reveals a correlation between glycosylation and structural integrity. J Mol
Biol 325(5):979–989
Kubota T, Niwa R, Satoh M, Akinaga S, Shitara K, Hanai N (2009) Engineered therapeutic anti-
bodies with improved effector functions. Cancer Sci 100(9):1566–1572
Lauer TM, Agrawal NJ, Chennamsetty N, Egodage K, Helk B, Trout BL (2012) Developability
index: a rapid in silico tool for the screening of antibody aggregation propensity. J Pharm Sci
101(1):102–115
Le Brun V, Friess W, Bassarab S, Muhlau S, Garidel P (2010) A critical evaluation of self-interaction
chromatography as a predictive tool for the assessment of protein-protein interactions in pro-
tein formulation development: a case study of a therapeutic monoclonal antibody. Eur J Pharm
Biopharm 75(1):16–25
6 Application of Biophysics to the Early Developability Assessment of Therapeutic… 145
Liu J, Nguyen MD, Andya JD, Shire SJ (2005) Reversible self-association increases the viscosity
of a concentrated monoclonal antibody in aqueous solution. J Pharm Sci 94(9):1928–1940
Lundstrom K (2005) Structural genomics of GPCRs. Trends Biotechnol 23(2):103–108
Ma P, Zemmel R (2002) Value of novelty? Nat Rev Drug Discov 1(8):571–572
Mehta CM, White ET, Litster JD (2012) Correlation of second virial coefficient with solubility for
proteins in salt solutions. Biotechnol Prog 28(1):163–170
Mimura Y, Church S, Ghirlando R, Ashton PR, Dong S, Goodall M, Lund J, Jefferis R (2000) The
influence of glycosylation on the thermal stability and effector function expression of human
IgG1-Fc: properties of a series of truncated glycoforms. Mol Immunol 37(12–13):697–706
Minton AP (2005) Influence of macromolecular crowding upon the stability and state of associa-
tion of proteins: predictions and observations. J Pharm Sci 94(8):1668–1675
Naylor J, Beech DJ (2009) Extracellular ion channel inhibitor antibodies. Open Drug Discov
J 1:36–42
Nelson AL, Dhimolea E, Reichert JM (2010) Development trends for human monoclonal antibody
therapeutics. Nat Rev Drug Discov 9(10):767–774
Norde W (1986) Adsorption of proteins from solution at the solid–liquid interface. Adv Colloid
Interface Sci 25(4):267–340
Oganesyan V, Damschroder MM, Leach W, Wu H, Dall’Acqua WF (2008) Structural characteriza-
tion of a mutated, ADCC-enhanced human Fc fragment. Mol Immunol 45(7):1872–1882
Payne RW, Nayar R, Tarantino R, Del Terzo S, Moschera J, Di J, Heilman D, Bray B, Manning
MC, Henry CS (2006) Second virial coefficient determination of a therapeutic peptide by self-
interaction chromatography. Biopolymers 84(5):527–533
Pepinsky RB, Silvian L, Berkowitz SA, Farrington G, Lugovskoy A, Walus L, Eldredge J, Capili
A, Mi S, Graff C, Garber E (2010) Improving the solubility of anti-LINGO-1 monoclonal
antibody Li33 by isotype switching and targeted mutagenesis. Protein Sci 19(5):954–966
Prassler J, Thiel S, Pracht C, Polzer A, Peters S, Bauer M, Norenberg S, Stark Y, Kolln J, Popp A,
Urlinger S, Enzelberger M (2011) HuCAL PLATINUM, a synthetic Fab library optimized for
sequence diversity and superior performance in mammalian expression systems. J Mol Biol
413(1):261–278
Presta LG (2008) Molecular engineering and design of therapeutic antibodies. Curr Opin Immunol
20(4):460–470
Remmele RL Jr, Nightlinger NS, Srinivasan S, Gombotz WR (1998) Interleukin-1 receptor (IL-
1R) liquid formulation development using differential scanning calorimetry. Pharm Res
15(2):200–208
Salinas BA, Sathish HA, Bishop SM, Harn N, Carpenter JF, Randolph TW (2010) Understanding
and modulating opalescence and viscosity in a monoclonal antibody formulation. J Pharm Sci
99(1):82–93
Saluja A, Kalonia DS (2008) Nature and consequences of protein-protein interactions in high
protein concentration solutions. Int J Pharm 358(1–2):1–15
Scherer TM, Liu J, Shire SJ, Minton AP (2010) Intermolecular interactions of IgG1 monoclonal
antibodies at high concentrations characterized by light scattering. J Phys Chem B
114(40):12948–12957
Sethuraman A, Belfort G (2005) Protein structural perturbation and aggregation on homogeneous
surfaces. Biophys J 88(2):1322–1333
Sinha S, Zhang L, Duan S, Williams TD, Vlasak J, Ionescu R, Topp EM (2009) Effect of protein
structure on deamidation rate in the Fc fragment of an IgG1 monoclonal antibody. Protein Sci
18(8):1573–1584
Stewart R, Thom G, Levens M, Guler-Gane G, Holgate R, Rudd PM, Webster C, Jermutus L, Lund
J (2011) A variant human IgG1-Fc mediates improved ADCC. Protein Eng Des Sel
24(9):671–678
Sule SV, Sukumar M, Weiss WF, Marcelino-Cruz AM, Sample T, Tessier PM (2011) High-
throughput analysis of concentration-dependent antibody self-association. Biophys
J 101(7):1749–1757
146 H. Sathish et al.
Tadros T (2011) Interparticle interactions in concentrated suspensions and their bulk (rheological)
properties. Adv Colloid Interface Sci 168(1–2):263–277
Tessier PM, Jinkoji J, Cheng YC, Prentice JL, Lenhoff AM (2008) Self-interaction nanoparticle
spectroscopy: a nanoparticle-based protein interaction assay. J Am Chem Soc 130(10):
3106–3112
Thomas CR, Geer D (2011) Effects of shear on proteins in solution. Biotechnol Lett
33(3):443–456
Valente JJ, Payne RW, Manning MC, Wilson WW, Henry CS (2005) Colloidal behavior of proteins:
effects of the second virial coefficient on solubility, crystallization and aggregation of proteins
in aqueous solution. Curr Pharm Biotechnol 6(6):427–436
Vlasak J, Ionescu R (2011) Fragmentation of monoclonal antibodies. MAbs 3(3):253–263
Wendorf JR, Radke CJ, Blanch HW (2004) Reduced protein adsorption at solid interfaces by sugar
excipients. Biotechnol Bioeng 87(5):565–573
Weng WK, Levy R (2003) Two immunoglobulin G fragment C receptor polymorphisms indepen-
dently predict response to rituximab in patients with follicular lymphoma. J Clin Oncol
21(21):3940–3947
Yadav S, Scherer TM, Shire SJ, Kalonia DS (2011) Use of dynamic light scattering to determine
second virial coefficient in a semidilute concentration regime. Anal Biochem 411(2):292–296
Yadav S, Shire SJ, Kalonia DS (2012) Viscosity behavior of high-concentration monoclonal anti-
body solutions: correlation with interaction parameter and electroviscous effects. J Pharm Sci
101(3):998–1011
Chapter 7
Application of Biophysics in Formulation,
Process, and Product Characterization:
Selected Case Studies
7.1 Introduction
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 147
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2_7,
© Springer Science+Business Media New York 2013
148 S.K. Singh et al.
manufactured before and after such changes, comparability studies are performed
(ICH Q5E 2005). In order to show that a product is comparable before and after
the changes, they are required to have a high degree of similarity in their physico-
chemical and biological properties. Biophysical characterization is an important
component of comparability studies to demonstrate similarity in higher-order
structure of the product. For this purpose, low-resolution structural information is
obtained through spectroscopic techniques such as far- and near-UV circular
dichroism (CD) spectroscopy, fluorescence spectroscopy, and Fourier-transform
infrared (FTIR) spectroscopy. These methods provide a signature for the second-
ary and tertiary structure of the molecule. Hydrodynamic analyses such as light-
scattering spectroscopy, AUC, and AFFFF can provide information about
quaternary structure. The case study below on an IgG2 antibody is presented
wherein product characterization by various biophysical methods was used to
explain the observed differences before and after a process-scale and formulation
change and to establish criteria for comparability.
7.2.1 Methods
Far- and near-UV CD spectra were collected on two drug substance lots with pilot-
scale formulation (Lots 1 and 2) and eight other lots with full-scale formulation
(Lots 3–10). For heat denaturation experiments, CD spectra were taken every 5°C
from 25 to 75°C. Three separate denaturation experiments were done for the assess-
ment of experimental uncertainty. All spectra were corrected with matched buffer
blanks. The spectral similarity was assessed by Similarity Match analysis in TQ
Analyst (Thermo Scientific). The DSC thermogram was generated at a scan rate of
1°C/min, and the molar excess heat capacity was calculated and deconvoluted to a
non-two-state model. Tryptophan fluorescence spectra were collected using an exci-
tation wavelength of 290 nm. All spectra were corrected with matched buffer blank.
For the Analytical ultracentrifugation sedimentation velocity (AUC-SV) experi-
ments, the samples were diluted to 0.5 mg/mL with the matched buffer and spun at
45,000 rpm. The absorbance data at 280 nm was collected and analyzed using Sedfit
(Schuck 2000). For details on the methods used, including the underlying theory,
please see Chap. 3 in this volume.
7.2.2 Results
Simple visual examination of both the far- and near-UV CD spectra in Fig. 7.1
shows differences between the pilot-scale and full-scale materials. Differences in
far-UV CD spectra (Fig. 7.1a) are mainly at 215–218 nm, suggesting alteration in
β (beta)-sheet content. The near-UV CD spectra (Fig. 7.1a) differ largely at around
290 nm, mainly attributable to the change in chiral environment around tryptophan
7 Application of Biophysics in Formulation, Process, and Product Characterization… 149
Fig. 7.1 Far-UV (a) and near-UV (b) CD spectra of pilot-scale and full-scale drug substance lots
in clinical and commercial formulations, respectively. Lots 1 and 2 are pilot-scale lots; Lots 3–10
are full-scale lots. Lots 10 and 10* are the reference material in full-scale/commercial and pilot-
scale/clinical formulation, respectively (see Table 7.1)
residues (Strickland 1974; Kelly and Price 2000). Intrinsic tryptophan fluorescence
spectroscopy analysis also shows difference between pilot-scale and full-scale
materials (Fig. 7.2), which is consistent with the difference seen in the near-UV
CD spectra. However, the change in fluorescence is not easily interpreted. The
150 S.K. Singh et al.
Fig. 7.2 Tryptophan emission fluorescence spectra of all different lots excited at 295 nm.
See Fig. 7.1 or Table 7.1 for lot information
Fig. 7.3 AUC-SV profiles of a pilot-scale/clinical formulation lot (Lot 2) and a full-scale/
commercial formulation lot (Lot 10)
for comparison. This application provides a single number for spectral comparison
and supplements the qualitative visual assessment. The use of this commercial
algorithm also eliminates tedious and error-prone spreadsheet calculations pre-
sented previously (Zou and Luo 2010). The calculated match values for the ten
different lots versus the reference standard (Lot 10) are listed in Table 7.1; the dif-
ference between pilot-scale and full-scale lots is in agreement with the visual
inspection. To establish a possible cut-off value for the spectral difference between
“similar” and “dissimilar,” comparison among a series of spectra generated by heat
denaturation was conducted (Fig. 7.4a, b). The spectra at the temperatures lower
than the onset denaturation temperature (60°C) as determined by DSC (Fig. 7.5)
are compared pair-wise using the similarity analysis. The resulting similarity
match values are subject to the t-test at 95% confidence level to deduce the lower
limit of the similarity match value used as the spectral similarity cut-off values
(87.27% and 96.50% for far- and near-UV CD spectra, respectively). As shown in
Table 7.1, only one lot (Lot 6) in the full-scale material is not similar to the refer-
ence standard in the far-UV CD analysis. However, there is no difference in its
near-UV CD spectrum. Careful examination of the far-UV CD spectrum of this lot
reveals a high level of noise in this data, and therefore, the spectrum does not have
as good a signal to noise ratio, confounding the comparison to the other samples.
However, the similarity analysis is sensitive enough to pick up small differences in
152 S.K. Singh et al.
Fig. 7.4 Far- (a) and near-UV (b) CD spectra of the IgG2 antibody at different temperatures
the spectra and should serve as a good tool for comparability studies. Other type of
data analyses such as data deconvolution for far-UV CD spectra can provide the
estimate of relative abundance of different secondary structures (Sreerama and
Woody 2000). However, it is judged to be not sufficiently sensitive to be used in
comparability studies (Zou and Luo 2010).
7 Application of Biophysics in Formulation, Process, and Product Characterization… 153
Fig. 7.5 The DSC thermogram of the reference standard (Lot 10). The traces are the fits from the
non-two-state model are shown
Table 7.1 Quantitative similarity analysis for pilot-scale and full-scale drug substance lots in
clinical and commercial formulations, respectively
Similarity match Similarity match
Drug substance value for far UV value for near UV
Lot # /formulation CD (%) CD (%)
10 (reference Full-scale/commercial 100 100
standard)
1 Pilot-scale/clinical 64.16 86.12
2 Pilot-scale/clinical 61.51 86.31
3 Full-scale/commercial 93.06 99.47
4 Full-scale/commercial 90.80 99.60
5 Full-scale/commercial 90.06 99.57
6 Full-scale/commercial 85.98a 99.60
7 Full-scale/commercial 94.78 99.41
8 Full-scale/commercial 89.71 99.45
9 Full-scale/commercial 92.52 99.58
10* (reference Pilot-scale/clinical 58.88 87.30
standard in clinical
formulation)
The values in bold are below the limits obtained from the heat denaturation study (87.27% and
96.50% for far- and near-UV CD, respectively)
a
Difference attributable to poor quality of spectrum collected
154 S.K. Singh et al.
7.2.3 Conclusions
quantify aggregates, but the upper limit of size range is limited to approximately
50 nm. Thus, this technique can assess oligomers but misses the larger aggregates
in the submicron and micron range (den Engelsman et al. 2011). Limitations of SEC
in terms of accuracy and specificity, when evaluating aggregates, have been dis-
cussed elsewhere (Philo 2009; Carpenter et al. 2010). Orthogonal methods such as
AUC and AFFFF are therefore recommended to supplement the SEC results.
The second case study presented in this chapter involves an Fc-fusion protein
which showed significant aggregation during storage at accelerated and stressed
conditions as measured by SEC. Application of AUC in two modes (sedimentation
velocity and gravitational sweep) provided some interesting insights into the self-
association behavior and impact of salt in the mobile phase for SEC.
7.3.1 Methods
The SEC analysis was performed using a TSKgel G3000SW column (Tosoh
Bioscience) with a mobile phase buffer of 200 mM sodium phosphate, 450 mM
NaCl, and pH 6.5. The elution profile was monitored at 214 nm. A sample volume
of 20 μL at 0.5 mg/mL was injected. AUC sedimentation velocity was run at
20,000 rpm and absorbance data at 280 nm was collected. The data were analyzed
using Sedfit (Schuck 2000). For the concentration-dependent study, samples at three
different concentrations were analyzed at 50,000 rpm, and the data were fitted using
Sedphat for reversible association (Schuck 2003). Gravitational sweep AUC
156 S.K. Singh et al.
Fig. 7.7 AUC-SV sedimentation coefficient distribution for the stability samples of an Fc-fusion
protein
(AUC-gs) was performed between 3,000 rpm and 50,000 rpm, and the absorbance
data at 280 nm was analyzed using SedAnal (Stafford and Braswell 2004). For more
detailed discussion of the techniques, please see Chap. 3.
7.3.2 Results
dissociate at high salt concentrations in the mobile phase was considered. The dis-
crepancy prompted further development of a more suitable SEC method. To further
characterize the aggregates profile, gravitational sweep AUC (AUC-gs) with speeds
between 3,000 and 45,000 rpm was employed. This technique expands the dynamic
range and allows detection of even larger aggregates if present, although quantita-
tion is difficult (Stafford and Braswell 2004). The resultant profile shown in Fig. 7.8
demonstrates that some very large aggregates are present in the stressed sample.
Comparison of the size distribution from the single speed at 20,000 rpm to that from
the variable speeds showed some interesting similarities as well as differences. Both
methods captured the large aggregate at the apparent sedimentation coefficient of
55 S (78 S in standard condition, Fig. 7.8), but the gravitational sweep method
detected even larger aggregates at 1,800 S. Due to the changing speed during the
sedimentation, it is not possible to determine the true size and mass of aggregates in
Fig. 7.8. However, if treated as compact spheres, the minimal size and molar masses
of the aggregates in an AUC-gs run can be estimated, and this data is provided in
Table 7.4. The results show that a broad size range of aggregates are formed in this
biotherapeutic after being subject to stress conditions.
In addition to the broad distribution of aggregates, the AUC profile at 20,000 rpm
shows two overlapping main peaks (Fig. 7.7). This suggests the existence of a
dynamic equilibrium during the sedimentation process. To further understand the
potential for reversible self-association, a concentration-dependent AUC-SV analy-
sis was applied to the sample at 5°C. Figure 7.9 shows the size distributions at three
different concentrations. The concentration-dependent shift in sedimentation coef-
ficient of the second peak suggests that there is rapid association/dissociation equi-
librium at the timescale of the sedimentation experiment (Schuck 2003). Global
data fitting to the Gilbert–Jenkins theory implemented in Sedphat gives a dimer
dissociation constant of 40 μM and the free energy of the association as −5.9
kcal/mol, suggesting a moderate strength protein–protein interaction (Horton and
Lewis 1992). It is likely that this association is dependent on salt concentration so
that this equilibrium does not occur in the SEC mobile phase, and the SEC analysis
therefore does not detect the presence of protein self-association. It may be possible
that the non-Fc part of the fusion protein has pH- and salt-dependent self-association
behavior too.
158
Table 7.3 AUC-SV results of the stability samples for the Fc-fusion protein
Other higher
Monomer Dimer Large aggregates order aggregates
Sample
treatment Mol. wt. Sed. coeff. Mol. wt. Sed. Coeff. Mol. wt. Sed. coeff.
(1 month) (kDa) (S20oC,w) Fraction % (kDa) (S20oC,w) Fraction % (kDa) (S20oC,w) Fraction % Fraction %
5 °C 85 4.1 21.5 222 7.7 78.0 1.2a
25 °C 86 4.7 28.5 202 8.0 68.7 3.0a
35 °C 7 78.2 33.9 27.8b
a
species that are larger than the “dimer” peak
b
species that are larger than the 7 MDa peak
S.K. Singh et al.
7 Application of Biophysics in Formulation, Process, and Product Characterization… 159
Fig. 7.8 Gravitational sweep AUC analysis of the 35°C 1-month stability sample of an Fc-fusion
protein showing the wide range of aggregate sizes present
Table 7.4 Size and molar mass of the hypothetical compact spheres for the
species in the gravitational sweep AUC (also see Fig. 7.8)
Minimum molecular weight Minimum hydrodynamic radius
Species (MDa) for compact spheres (nm) for compact spheres
1800 S 565.6 54.7
245 S 28.4 20.2
55 S 3.0 9.6
15 S 0.4 5
4.5 S 0.1 2.7
7.3.3 Conclusions
Orthogonal characterization based upon AUC in different modes, allowed the nature
of aggregation behavior of the Fc-fusion protein to be comprehensively elucidated
in contrast to the SEC analysis. The observed discrepancy prompted an improve-
ment in the SEC method, while detailed analysis of the AUC runs showed the exis-
tence of a reversible self-association behavior in the system at the timescale of the
AUC experiment.
160 S.K. Singh et al.
Fig. 7.9 Concentration-dependent AUC-SV analysis of the 5°C 1-month sample showing the
concentration-dependent shift in position of the second peak
7.4.1 Methods
7.4.2 Results
Fig. 7.10 Dynamic light-scattering size distribution of the particles in drug substance stability
samples. Both backscattering (173°) and forward-scattering (13°) results from unfiltered and fil-
tered samples are shown
162 S.K. Singh et al.
Fig. 7.11 Real-time monitoring of particle formation in the filtered samples by DLS at 37°C
Fig. 7.12 A schematic diagram illustrating the mechanism for visible particle formation in the
IgG4 antibody
7.4.3 Conclusions
A real-time DLS study was able to help elucidate the mechanism of visible particle
formation and explain the lack of detection of aggregation by SEC. Once aggregates
reach a certain size range, they rapidly coalesce to form larger visible particles, thus
effectively eliminating all but a small amount of aggregates, detectable by SEC,
from the solution. No aggregates in intermediate size ranges are detectable, and
therefore, no changes are seen in the SEC results, despite the appearance of visible
particles.
7.5.1 Methods
Nile red extrinsic dye-based fluorescence was carried out according to the procedure
described by (Demeule et al. 2007), with some modifications. Nile red was dissolved
in DMSO to obtain a 10 mM stock solution. Three microliter of the dye stock solution
was added to 2,997 μL of protein at 20 mg/mL for the fluorescence experiments.
Thermal incubation was conducted by preheating the cuvette to the target temperature
and then adding 3 mL of dye + protein solution into the cuvette at time zero.
An IgG2 mAb was prepared at 20 mg/mL in 20 mM histidine buffer pH 5.5, without
any other excipients or stabilizers. Samples were purposely stressed to induce aggrega-
tion including thermal treatment at 40°C (7 weeks) or 50°C (10 days), or 20 cycles of
freeze/thaw stress. The level of aggregation was measured by SEC, and the samples
were also studied by CD, intrinsic fluorescence, and bis-ANS fluorescence, apart from
the Nile red fluorescence described above. An aggregate sample was also created by
mixing aliquots of 40°C 7 weeks and 5°C control at 1:1 (v/v) ratio.
7.5.2 Results
SEC analysis of the test solutions showed fairly low and similar levels of aggre-
gates (labeled as high molecular mass species, HMMS) (see Table 7.5).
Characterization of these samples by circular dichroism (Fig. 7.13), intrinsic
fluorescence (not shown), bis-ANS fluorescence (Fig. 7.14), and Nile red fluo-
rescence (Fig. 7.15) showed little differences between them. DLS analysis did
not pick up any significant particles beyond the main peak at around 9 nm (data
not shown). However, when samples containing the Nile red dye were incubated
at 63°C (just below the onset of CH2 domain-melting peak by DSC), and moni-
tored for peak fluorescence, an interesting trend was seen. Once at temperature,
the peak Nile red fluorescence intensity of the samples increased in a linear
fashion that was distinct for the different stresses (Fig. 7.16). The freeze/thaw
cycled sample and the control sample were very similar in the evolution of the
peak fluorescence while the thermally stressed samples followed a distinctly
different pattern with a more rapid increase in the peak intensity. The results of
Fig. 7.13 Circular dichroism spectra of stressed samples of IgG2 mAb. Sample treatments are as
follows: 5°C control control unstressed sample, FT20 20 cycles of freeze/thaw stress, 40°C 7 weeks
thermal stress at 40°C for 7 weeks, 50°C 10 days thermal stress at 50°C for 10 days
Fig. 7.14 Extrinsic (bis-ANS) fluorescence spectra of stressed samples of IgG2 mAb. Sample treat-
ments are as follows: 5°C control control unstressed sample, FT20 20 cycles of freeze/thaw stress,
40°C 7 weeks thermal stress at 40°C for 7 weeks, 50°C 10 days thermal stress at 50°C for 10 days
166 S.K. Singh et al.
Fig. 7.15 Extrinsic (Nile red) fluorescence spectra of stressed samples of IgG2 mAb. Sample treat-
ments are as follows: 5°C control control unstressed sample, FT20 20 cycles of freeze/thaw stress,
40°C 7 weeks thermal stress at 40°C for 7 weeks, 50°C 10 days thermal stress at 50°C for 10 days
Fig. 7.16 Time course of Nile red peak fluorescence for stressed samples of IgG2 mAb under incuba-
tion at 63°C. Sample treatments are as follows: 5°C control control unstressed sample, FT20 20 cycles
of freeze/thaw stress, 40°C 7 weeks thermal stress at 40°C for 7 weeks, 50°C 10 days thermal stress at
50°C for 10 days, 40°C 7 weeks 5°C control 1:1 is a 1:1 by volume mixture of the thermally stressed
and unstressed control samples. Linear regression plots are shown (see Table 7.6)
7 Application of Biophysics in Formulation, Process, and Product Characterization… 167
Table 7.6 Linear regression analysis for time course of Nile red peak fluorescence from stressed
samples of IgG2 mAb shown in Fig. 7.16
Sample treatments HMMS (%) Slope (h−1) Intercept R2
5°C control 0.9 0.0662 22.062 0.9465
Freeze/thaw cycles FT20 2.0 0.0633 20.377 0.9824
Heat treated 40°C 7 weeks + 5°C 1.0 0.0795 51.973 0.9963
control 1 + 1 v/v
Heat treated 40°C 7 weeks 1.1 0.0897 61.477 0.9797
Heat treated 50°C 10 days 1.9 0.1231 48.408 0.9983
a linear model fit are summarized in Table 7.6. The fit parameters do not corre-
late with the measured HMMS values, but they form two distinct groups. It is
clear from the fluorescence data in Fig. 7.15 that the aggregate (levels) or the
applied stress per se cannot be distinguished by their ability to bind the hydro-
phobic dye Nile red. However, when incubated at 63°C, it is likely that mole-
cules that have small structural perturbations and misfolds as a consequence of
the stress history of the samples, begin to unfold in the cuvette. As this occurs,
the Nile red binding and fluorescence response increases. The data in Fig. 7.16
therefore suggests that molecules that underwent freeze/thaw stress, either
aggregate or refold back to the native state when thawed. Therefore, few par-
tially unfolded or misfolded species are created by this stress, and the response
with incubation (in Fig. 7.16) is similar to that of the control sample. However,
when the mAb is thermally stressed, a variety of such species are created which
are not detectable by SEC since they are not aggregates, and are not distinguish-
able by conventional spectroscopic techniques since they are either too few and/
or too similar to the native molecule.But the structural susceptibility of these
molecules is amplified by incubating the sample at 63°C, where the presence of
the Nile red dye is able to capture this phenomenon, which would otherwise be
too small energetically to be detected by DSC (data not shown).
7.5.3 Conclusions
Detection and control of aggregation is a key aspect for the successful develop-
ment of a biotherapeutic product. Exposed hydrophobic patches either by design
or through some degree of misfolding are generally understood to act as seeds for
the nucleation step of aggregation. Such regions may be present in only a small
fraction of the population and would therefore be difficult to detect. An ability to
detect small amounts of such precursors would therefore be useful as a predictor
of long-term behavior of a particular batch of the molecule. Ability to detect small
differences in the exposed hydrophobic regions can also be of value in selecting
among candidate molecules, as an indicator of their aggregation propensity.
Following upon the above concept of exposed hydrophobic regions, it was decided
to use bis-ANS binding as a probe. The proposed hypothesis is that bis-ANS would
bind more efficiently (tighter and faster) to such hydrophobic regions.
7.6.1 Methods
7.6.2 Results
Three monoclonal antibodies were evaluated for their interaction with the hydro-
phobic dye bis-ANS. The titration curves are fit to the model as shown in Table 7.7.
7 Application of Biophysics in Formulation, Process, and Product Characterization… 169
Table 7.7 Thermodynamic parameters of bis-ANS binding to different IgG mAbs and HSA
derived from ITC data at 25°C fit to a two-class binding-site model
Protein pI N1 K1 ΔG1 ΔH1 TΔS1 N2 K2 ΔG2 ΔH2 TΔS2
IgG2 Mol A 8.6 13 11 −9.6 0.18 −9.77 22 0.12 −6.93 −0.20 6.26
IgG2 Mol B 8.0 10 20 −9.9 0.11 −9.98 18 0.16 −7.16 −1.38 5.72
IgG2 Mol C 8.3 3 10 −9.6 0.10 −9.66 17 0.19 −7.19 −1.67 5.54
HSA 4.9 11 82 −10.8 0.29 −11.0 9 0.46 −7.72 −1.91 5.81
The pI of each molecule is also listed. Note the opposite signs of ΔH and –TΔS for the two
classes of binding sites for all proteins
N = ± 1 (SD); K = 1 × 106 M (5% SD); ΔG = kcal/mol (5% SD); ΔH = kcal/mol (5% SD);
−TΔS = kcal/mol (5% SD)
HSA was also included in the study as a reference. The results show that bis-ANS
binds to the molecules with varying stoichiometry and affinity. The data was ana-
lyzed using a two-class binding-site model. Analysis shows that the first class of
binding sites are specific and entropically driven (KD1 = 0.0–0.2 µM; ΔH positive;
TΔS negative; KD1 is inverse of K1, the association constant). This is likely a result
of the interaction of the naphthalene ring structure of bis-ANS molecules with
hydrophobic surfaces/pockets of the protein. The second class of binding sites are
nonspecific, low-affinity sites (KD2 = 10–40 µM; ΔH negative; TΔS positive; KD2 is
inverse of K2, the association constant), probably a result of hydrophilic/ionic
interactions through the sulfonic acid groups. For reference, although the HSA–
bis-ANS N1 value is similar to the mAbs (despite being a smaller protein), the
equilibrium constant (KD1) is significantly larger, indicating a stronger affinity of
bis-ANS for the more hydrophobic HSA protein. Based on the affinity and stoi-
chiometry of the first binding site (N1), the aggregation propensity driven by
hydrophobicity is rank-ordered as HSA > IgG2 MolA > IgG2 MolB > IgG2 MolC.
For aggregation to occur as a consequence of the presence of hydrophobic
patches, the patches must be accessible. It was hypothesized that the patches that are
accessible would be the first to which bis-ANS would bind, and thus the molecule
with more of these accessible patches would have a faster initial binding of bis-
ANS. To distinguish between the molecules on the basis of the (extent of their)
highly accessible hydrophobic patches, binding kinetics was studied by the stopped-
flow technique. Stopped-flow measurements on the four test molecules are shown in
Fig. 7.17 with the model fit data in Table 7.8. A biphasic-binding model was used
for the data analysis. Hydrophobic HSA has the highest k1, indicating the fastest
binding rate, and also correlates with the high aggregate level (% HMMS) measured
by SEC. Among the mAbs, the low general level of aggregation did not correlate
with the kinetics. However, based on k1, the rank order of aggregation propensity is
as follows HSA > IgG2 MolA > IgG2 MolC > IgG2 MolB.
The rank ordering based on stopped-flow kinetics was found to correlate with the
long-term stability behavior of the mAbs.
170 S.K. Singh et al.
Fig. 7.17 Stopped-flow binding kinetics for mAbs and HSA with bis-ANS. Raw data is compared
at 20s time scale.
Table 7.8 Bis-ANS binding kinetic parameters derived from a biphasic kinetic model fit to
stopped-flow data
HMMS
Protein (%) (SEC) k1 (s−1) k2 (s−1) A1 (RFES)a A2 (RFES)a R2
IgG2-Mol A 0.3 2.44 ± 0.02 0.048 ± 0.003 3.34 ± 0.23 15.15 ± 0.48 0.99
IgG2-Mol B 1.0 1.41 ± 0.24 0.122 ± 0.010 2.18 ± 0.11 4.49 ± 0.10 0.94
IgG2-Mol C 0.7 1.78 ± 0.09 0.093 ± 0.002 1.71 ± 0.34 40.7 ± 0.18 0.99
HSA 17.3 11.82 ± 0.35 1.32 ± 0.05 3.16 ± 0.54 7.01 ± 0.27 0.98
a
Relative fluorescence emission signal
7 Application of Biophysics in Formulation, Process, and Product Characterization… 171
7.6.3 Conclusions
7.7 Conclusions
References
Carpenter JF, Randolph TW, Jiskoot W et al (2009) Overlooking subvisible particles in therapeutic
protein products: gaps that may compromise product quality. J Pharm Sci 98:1202–1205
Carpenter JF, Randolph TW, Jiskoot W, Crommelin DJA, Middaugh CR, Winter G (2010) Potential
inaccurate quantitation and sizing of protein aggregates by size exclusion chromatography:
essential need to use orthogonal methods to assure the quality of therapeutic protein products.
J Pharm Sci 99:2200–2208
Cromwell MEM, Hilario E, Jacobsen F (2006) Protein aggregation and bioprocessing. AAPS J
8:E572–E579
Dam J, Schuck P (2000) Calculating sedimentation coefficient distributions by direct modeling of
sedimentation velocity concentration profiles. In: Johnson ML, Brand L (eds) Methods in
enzymology, vol 384. Part E: numerical computer methods. Elsevier Academic, New York,
pp 185–212
Demeule B, Gurny R, Arvinte T (2007) Detection and characterization of protein aggregates by
fluorescence microscopy. Int J Pharm 329:37–45
den Engelsman J, Garidel P, Smulders R, Koll H, Smith B, Bassarab S, Seidl A, Hainzl O, Jiskoot
W (2011) Strategies for the assessment of protein aggregates in pharmaceutical biotech product
development. Pharm Res 28:920–933
Horton N, Lewis M (1992) Calculation of the free energy of association for protein complexes.
Protein Sci 1:169–181
ICH Q5E (2005) Comparability of biotechnological/biological products. Available from
www.ICH.org
Kelly SM, Price NC (2000) The use of circular dichroism in the investigation of protein structure
and function. Curr Protein Pept Sci 1:349–384
Kozlowski S, Swann P (2006) Current and future issues in the manufacturing and development of
monoclonal antibodies. Adv Drug Del Rev 58:707–722
172 S.K. Singh et al.
Philo JS (2009) A critical review of methods for size characterization of non-particulate protein
aggregates. Curr Pharm Biotechnol 10:359–372
Roberts CJ (2006) Non-native protein aggregation: Pathways, kinetics, and shelf-life prediction.
In: Murphy RM, Tsai AM (eds) Misbehaving proteins: protein misfolding, aggregation, and
stability. Springer, New York, pp 17–46
Sackett DL, Wolff J (1987) Nile red as a polarity-sensitive fluorescent probe for hydrophobic pro-
tein surfaces. Anal Biochem 167:228–234
Schuck P (2000) Size distribution analysis of macromolecules by sedimentation velocity ultracen-
trifugation and Lamm equation modeling. Biophys J 78:1606–1619
Schuck P (2003) On the analysis of protein self-association by sedimentation velocity analytical
ultracentrifugation. Anal Biochem 320:104–124
Singh SK (2011) Impact of product-related factors on immunogenicity of biotherapeutics. J Pharm
Sci 100:354–387
Sreerama N, Woody RW (2000) Estimation of protein secondary structure from circular dichroism
spectra: comparison of CONTIN, SELCON, and CDSSTR methods with an expanded refer-
ence set. Anal Biochem 287:252–260
Stafford WF, Braswell EH (2004) Sedimentation velocity, multi-speed method for analyzing poly-
disperse solutions. Biophys Chem 108:273–279
Strickland HE (1974) Aromatic contributions to circular dichroism spectra of proteins. CRC Crit
Rev Biochem 2:113–175
Weiss WFI, Young TM, Roberts CJ (2009) Principles, approaches, and challenges for predicting
protein aggregation rates and shelf life. J Pharm Sci 98:1246–1277
Zou Q, Luo Y (2010) Biophysical characterization for product comparability. BioPharm Int Suppl
12–17
Chapter 8
Biophysical Analysis in Support of Development
of Protein Pharmaceuticals
8.1 Introduction
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 173
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2_8,
© Springer Science+Business Media New York 2013
174 S. Alavattam et al.
Gibson et al. (2011) recently have reported the use of HTS methods for protein solu-
bility using an IgG1 mAb. The authors were able to modify a PEG-induced precipi-
tation method in a 96 well format and used ultra violet-visible (UV–Vis) spectroscopy
to screen various buffer compositions and pH that help to maintain mAb solubility.
Relative solubility profiles of both chimeric and human IgG1 mAbs were deter-
mined using this HTS method. He et al. (2010a) reported the use of extrinsic fluo-
rescence in a 96-well format to detect and quantify IgG aggregation. It can be
envisioned that several other analytical techniques, such as turbidity, DLS, and vari-
ous chromatographic methods, can be utilized in the future in an HTS format to help
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 177
Fig. 8.2 Bar plots comparing kD (filled square) and mAb solution viscosity (open square) in
(a) 20 mM His-OAc, pH 5.5, (b) 30 mM His-Cl, pH 6.0, (c) 200 mM Arg-Cl, pH 5.0, and
(d) 200 mM Arg-Succ, pH 5.5. Scatter plots display correlation between kd and viscosity for the
corresponding mAbs in (e) 20 mM His-OAc, pH 5.5, (f) 30 mM His-Cl, pH 6.0, (g) 200 mM
Arg-Cl, pH 5.0, and (h) 200 mM Arg-Succ, pH 5.5. Viscosity was measured at 175 mg/mL by cone
and plate rheometry (published previously in Connolly et al. 2012)
178 S. Alavattam et al.
mAbs provided a linear fit and showed the relation as kd = 1.06 A2M − 8.9, where M
is the molecular weight of the mAb. Overall, this methodology using DLS could be
used to analyze A2 using low protein quantities and is amenable to HTS as demon-
strated in this work.
Biophysical measurements that are labor-intensive such as AUC are not amenable to a
high-throughput methodology but can provide important information to guide further
development of a protein therapeutic. As an example, Lu et al. (2008) used sedimenta-
tion velocity AUC to demonstrate its usefulness in monitoring long-term stability and
molecular integrity of antibodies. Specifically, their data was used to support the notion
that a single point mutant in the hinge region of an IgG4 (S241 to P241) led to an
increase in stability of the molecule against freeze–thaw-induced aggregation.
AUC has also been used to characterize the complexes formed in vitro between
an anti IgE–IgG1 antibody and IgE (Liu et al. 1995). Monoclonal antibodies that
bind to free-circulating IgE can be used for the treatment of IgE-induced allergic
asthma, as they prevent the loading of IgE on mast cells or basophils, which can
result in release of inflammatory molecules such as leukotrienes and histamine after
exposure to an allergen (Fig. 8.3). Theoretically since there are two sites on each IgE
where anti IgE can bind and since each anti IgE molecule is bivalent, the complexes
could become very large (Fig. 8.4). Sedimentation velocity (Fig. 8.5) and equilib-
rium measurements were used to determine weight-average molecular weight and
size distribution (Fig. 8.6) to clearly show that this did not happen and that the com-
plexes that formed were of limiting size. The complexes formed are dependent on
the molar ratio of IgE:anti IgE as shown in a schematic diagram (Fig. 8.7). The
formation of these complexes dictates the pharmacokinetics of the drug therapy
since IgE has a clearance time of ~6 h whereas an IgG1, due to the binding to FcRN
neonatal receptor, has a typical half-life of ~2 weeks in serum. The complexes that
form between IgE and anti IgE take on the long half-life typically seen for an IgG1
(Fox et al. 1996). The amount of free IgE in plasma should be related to the clear-
ance rate of anti IgE:IgE complexes, free IgE, unbound anti IgE, and relative bind-
ing affinities of high-affinity receptor for IgE and that of IgE with anti IgE (Fig. 8.8).
Thus, the dose required for effective lowering of free IgE in plasma will be related
to stability of the complexes when interacting with high-affinity receptor. Since it is
possible to detect the formed IgE:anti IgE complexes using sedimentation velocity
AUC, it is possible to perform competitive binding experiments using AUC. AUC
experiments were performed with preformed anti IgE:IgE complexes at a molar
ratio of 6:1 where there is an excess of anti IgE and a predominance of a trimeric
species consisting of two anti IgE molecules bound to one IgE (Fig. 8.7). A soluble
180 S. Alavattam et al.
Fig. 8.3 Mechanism of action of an anti IgE mAb in the treatment of IgE allergic-mediated d isease.
The anti IgE mAb can inhibit IgE synthesis and binds free-circulating IgE preventing the IgE from
interacting with the FcεRIα high-affinity receptors on the surface of mast cells or basophils
Fig. 8.4 The theoretical interaction of IgE (dark grey) with an anti IgE mAb (light grey) via bind-
ing of the two high-affinity Fc receptor sites on IgE
form of the high-affinity receptor, sFcεRIα, was then added at several molar ratios,
and the results (Fig. 8.9a) clearly showed that the receptor has a greater affinity for
IgE than anti IgE necessitating excess dosing of anti IgE to effectively lower free-
circulating IgE. In particular, at a molar ratio of IgE:anti IgE of 1:6, all the IgE is
incorporated into a trimeric complex at ~13.3 S, and upon addition of soluble recep-
tor at 0.1:1 of complex, there is a reduction of this 13.3 S peak as well as a slight
increase in the baseline between 8 and 9 S and an increase of the unbound anti IgE
peak at ~7 S. As more soluble receptor is added, there are additional increases in the
anti IgE peak and the appearance of a peak that is likely the receptor:IgE complex,
which was characterized as a dimer (Fig. 8.7) in earlier work using sedimentation
velocity and equilibrium AUC as well as static light scattering (SLS) (Liu et al.
1997). These data were generated using the differential sedimentation method of
Stafford (Stafford 1992), which does not take into account diffusion resulting in a
broadening of the peak and is the likely reason that at lower concentrations it is
difficult to detect a single peak representative of the receptor:IgE dimer. As an
extension of this technique, a competition binding AUC experiment was also done
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 181
Fig. 8.5 Differential sedimentation coefficient distribution of IgE (solid line) and anti IgE (dotted
line) monomers at 0.64 mg/mL (a); IgE and anti IgE complexes at various molar ratios (b and c)
in PBS at 10°C. The molar ratios of IgE:anti IgE were as follows: (b) 1:1 (solid line), 1:3 (dash-
dotted line), 1:6 (dashed line), and 1:10 (dotted line): (c) 1:1 (solid line), 3:l (dash-dotted line), 6:1
(dashed line), and 10:1 (dotted line). The sedimentation coefficients have been corrected to the
standard condition of water at 20°C. No faster moving species was observed in early scanning
(previously published in Liu et al. 1995)
using a genetically engineered version of an anti IgE, referred to as anti IgEZ, which
was designed to have a greater affinity for IgE. The competition binding analysis
shows that anti IgEZ does indeed have higher affinity since even at a ratio of soluble
receptor to IgE at 1:1 there is very little disruption of the complex (Fig. 8.9b).
These examples are typical of how biophysics can be used to study the behavior
of biotherapeutics in vitro and help in determining adequate dosing. However, rather
little information on chemical and physical stability is available once the drug is
administered to humans, since characterization under physiological conditions
182 S. Alavattam et al.
Fig. 8.6 Sedimentation equilibrium analysis of IgE and anti IgE mAb complex formation in PBS
at 10°C. The weight-average molecular weights of complexes at different molar ratios were
obtained by analyzing the data from three different rotor speeds (5,000, 7,000, and 10,000 rpm) as
a single ideal species simultaneously. The error bars correspond to a 95% confidence interval
(adapted from Liu et al. 1995)
requires specialized tools. Recent advances in hardware have resulted in the capa-
bility to use fluorescence optics in AUC and have been specifically used to detect an
anti IgE molecule binding to IgE in a complex matrix such as human serum
(Demeule et al. 2009a). Two main differences were noticed for the anti IgE binding
to IgE in serum vs. PBS. The absence of the 21 S peak and the presence of an 8.7 S
peak in serum for the anti IgE complex with IgE, instead of a 7.3 S peak in PBS,
were noticeable. The absence of the 21 S peak and the different profile observed in
serum compared to PBS underlines the importance to characterize the molecules
under physiologically relevant conditions. Additionally, the absence of the 7.3 S
peak that is replaced by a 8.7 S peak further confirms the distinct behavior in serum
compared to PBS. The authors hypothesize that affinity of an anti IgE molecule
towards IgE was higher in serum compared to that in PBS and that the largest anti
IgE:IgE complex observed in serum was smaller than expected. AUC equipped with
fluorescence optics was clearly demonstrated to be useful to characterize biophar-
maceuticals under physiological conditions. Direct characterization in serum may
now allow for better drug candidate selection and should be carefully considered
during molecular assessment in early research stages.
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 183
Fig. 8.7 Schematic diagram of complex formation by IgE and anti IgE and IgE and soluble
high-affinity Fc receptor, sFcεRIα (adapted from Liu et al. 1997)
Fig. 8.8 Overall scheme for clearance of IgE (purple), anti IgE (red), and formation and clearance
of complexes at excess anti IgE vs. binding of IgE with high-affinity receptor, FcεRIα (grey), on
mast cells and basophils
Fig. 8.9 Differential sedimentation coefficient distribution of anti IgE, FcεRIα, and IgE:anti IgE
and IgE:FcεRIα complexes (a) and anti IgEZ, FcεRIα, and IgE:anti IgEZ and IgE:FcεRIα com-
plexes (b): assessment of competition of binding of soluble high-affinity receptor, FcεRIα, with
either preformed IgE:anti IgE or IgE:anti IgEZ complexes using AUC sedimentation velocity
the overall conformational changes are large enough to elicit a significant change in
the determined values. SEC when used with an on-line light scattering detector to
determine weight-average molecular weight can also detect mis-folded conformers
as long as the shape change is large enough to allow for separation from the prop-
erly folded protein monomer (Philo 2006).
Spectrophotometric techniques such as circular dichroism (CD) have often been
used to assess the folded state of a protein. Secondary structures (including α helices
and β sheets) can be identified in the far-UV region, 190–240 nm, of the CD spec-
trum. These ellipticity changes stem from the distinct chiral positioning of the
amide chromophores within different secondary structures (Johnson 1990). In addi-
tion, the local environment of aromatic chromophores such as tryptophan, tyrosine,
and phenylalanine gives rise to a CD signal in the near-UV region, 240–340 nm, that
can be used to infer tertiary structure changes (Johnson 1990). Thus, CD can be
used to determine if the structure of a recombinant DNA-derived protein is that
expected of “natural sourced” proteins or whether the structure is altered when
exposed to stress conditions. As an example, the far-UV spectrum of a mAb refer-
ence material from Genentech showed a minimum at 218 nm indicative of the
β-sheet character consistent with IgG1 antibodies (Brahms and Brahms 1980).
Comparison of both near- and far-UV spectra between reference material and the
stress panel shows no differences in structure by CD, as shown in Fig. 8.10, indicat-
ing that any subtle structural changes due to the applied stress (pH, oxidation, etc.)
may not be picked up in this technique. However, it has been shown recently by Li
et al. (2011) that CD can be used to monitor conformational changes in proteins
during manufacturing process conditions. Li et al. (2011) describe the effect of pH
or denaturants during purification and show a quantitative method to compare CD
spectra using the OMNIC QC compare algorithm. Near-UV CD spectrum of several
protein candidates was assessed either in sodium citrate buffer (pH 3.0) or in PBS
(pH 7.4). Careful evaluation of the near-UV CD spectrum of these candidates
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 185
Fig. 8.10 Far- and near-UV CD spectra of a monoclonal antibody after exposure to different stress
conditions
i ndicated that fewer changes were induced by low pH on candidate 1 over candidate
2, and hence the former may be more amenable to manufacturing processes at low
pH. Far-UV CD spectrum of the same two candidates indicated that both proteins
underwent the loss of the native β structure when incubated at pH 3.0; however, the
secondary structure of candidate 1 was found to be relatively more stable than can-
didate 2.
Although CD is often used on its own to probe conformation of proteins, it can
be a very valuable tool when used in conjunction with other biophysical techniques.
The self-association properties and conformation of recombinant DNA-derived
human relaxin, a pregnancy hormone, were studied by sedimentation equilibrium
analytical ultracentrifugation and CD (Shire et al. 1991). The sedimentation equilib-
rium AUC data were consistent with a monomer–dimer self-association model with
an association constant of ~6 × 105 M−1. An approximate five fold increase in weight
fraction of human relaxin monomer elicited by dilution of the protein resulted in no
change in the far-UV CD spectrum at 220 nm.
In contrast, after the same increase in weight fraction of monomer, the near-UV
circular dichroism spectra for human relaxin showed a significant decrease in the
intensity of the CD bands near 277 and 284 nm. Although human relaxin has two
tryptophan residues, the near-UV CD spectra exhibit only a broad shoulder near
295 nm rather than the strong CD bands often found for tryptophan. Moreover, there
is little change in this broad band after dilution of human relaxin to concentrations
that resulted in a five fold increase in the monomer weight fraction (Fig. 8.11).
186 S. Alavattam et al.
Fig. 8.11 Near-UV circular dichroism of human relaxin at 0.5 mg/mL (solid line) and 20 μg/mL
(dotted line). Relaxin at 0.5 mg/mL was thermostated at 20 °C in a 1-cm cell, whereas relaxin at
20 μg/mL was in an unthermostated 10-cm cylindrical cuvette. The temperature in the sample
compartment was ~27°C during the data collection process. The CD data were collected at 0.25-
nm intervals at a spectral bandwidth of 0.5 nm and are the result of an average of three scans using
an average time for each single data point collection of 5 s for the 0.5 mg/mL samples and the
result of an average of ten scans using an average time for each single data point collection of 10 s
for the 20 μg/mL sample. The weight fraction of human relaxin monomer estimated from the
determined association constant by sedimentation equilibrium AUC of 100 (g/L)−1 is 0.13 at
0.5 mg/mL and 0.50 at 20 μg/mL (adapted from Shire et al. 1991)
These data suggest that dissociation of the human relaxin dimer to monomer is not
accompanied by large overall changes in secondary structure or alteration in the
average tryptophan environment, whereas there is a significant change in the tyro-
sine environment. This conclusion was affirmed by the X-ray crystal structure of
human relaxin, which crystallized as a dimer with the lone tyrosine from each
monomer at the dimer interface (Eigenbrot et al. 1991). Thus, the solution studies
were in good agreement with the crystal studies, suggesting that the determined
crystal structure is very similar to the structure of the protein in solution.
Another spectroscopic technique widely used to assess conformational changes
in proteins is Fourier transform infrared (FTIR) spectroscopy, which can be used to
probe secondary structural elements including α helices and β sheets in solution as
well as in solid-state dosage forms. These structural features appear in FTIR spectra
as broad, characteristic absorption bands in the regions 1,700–1,620 cm–1 (Amide I)
and 1,600–1,500 cm–1 (Amide II), among others (Byler and Susi 1986; Dong et al.
1990; Jackson and Mantsch 1995). These absorption bands are caused by a combi-
nation of bending and stretching vibrations of bonds along the peptide backbone.
Since the technique can be used to assess conformation of proteins in solution as
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 187
Fig. 8.12 Inverted second derivative of the amide I IR spectra of rhDNase I: (a) in aqueous
solution (EGTA treated, no incubation), (b) lyophilized-untreated power (solid line) and lyophi-
lized EGTA-treated power (dotted line) (no incubation for either sample), (c) deamidated form
(aqueous protein incubated in 1 mM CaCl2 for 120 days at 40°C), and (d) aggregated form (aque-
ous EGTA-treated protein incubated in the absence of exogenous calcium ions for 120 days at
40°C) (published previously in Chen et al. 1999)
well as in the solid state, it allows for assessments of excipients used to stabilize the
protein during drying. Studies using omalizumab clearly showed that the native
conformation of a monoclonal antibody could be preserved when a lyoprotectant
such as sucrose was added to the formulation. Freeze drying in the absence of lyo-
protectant resulted in formation of covalent aggregates, which were linked by disul-
fide bonds as shown by nonreducing and reducing SDS polyacrylamide gel
electrophoresis (SDS PAGE) (Andya et al. 2003). FTIR has also been used to inves-
tigate the conformational stability of Pulmozyme in the aqueous and solid states
(lyophilized). Pulmozyme is a recombinant DNA-derived DNase used for the treat-
ment of cystic fibrosis and requires calcium ions for stability and activity (Chen
et al. 1999). Exogenous calcium can be removed by treatment with EGTA leaving
one tightly bound calcium ion per rhDNase molecule. Analysis of the FTIR spectra
in the amide III region in either the aqueous or lyophilized state demonstrated that
removal of exogenous Ca2+ by EGTA treatment had little effect on the secondary
structure (Fig. 8.12, Table 8.1). This result for the aqueous state was confirmed
using CD that also showed that there was no large overall change in the secondary
or tertiary structure upon the removal of calcium. The primary degradation route for
rhDNase in solution is deamidation. For the EGTA-treated protein, there was also
severe covalent aggregation, e.g., formation of intermolecular disulfides facilitated
188 S. Alavattam et al.
Table 8.1 FTIR analyses of various rHDNase I samplesa (published previously in Chen et al.
1999)
Secondary structure (%)
Sample α helix β sheet Otherb
Aqueous solution, with Ca 2+,c,d
21 ± 2 23 ± 3 56 ± 6
Aqueous solution, EGTA treated 20 ± 2 26 ± 2 54 ± 3
Lyophilized, with Ca2+,d 13 ± 2 41 ± 3 46 ± 2
Lyophilized, EGTA treated 14 ± 2 45 ± 3 41 ± 3
Deamidated forme 21 ± 1 25 ± 2 54 ± 2
Aggregated formf 10 ± 1 44 ± 4 46 ± 3
a
The secondary structure of rhDNase was calculated by Gaussian curve fitting the original amide
III spectra
b
Other secondary structure includes random coil and turns and extended chains
c
The aqueous solution contained 1 mM calcium chloride
d
Data from Saluja and Kalonia (2004)
e
Aqueous protein incubated in 1 mM Cacl2 for 120 days at 40°C
f
Aqueous EGTA-treated protein incubated in the absence of exogenous calcium ions for 120 days
at 40°C
8.3.4 A
ssessing Impact of Chemical Changes
on Physical Stability
While chemical changes are examined using various analytical methods, they typi-
cally do not monitor structural changes that may accompany the chemical change.
Liu et al. (2008) have recently reported the structural changes that accompany Met
oxidation in the CH2 domain of an Escherichia coli expressed Fc protein. The
authors report that methionine oxidation led to subtle changes in the protein confor-
mation and used various biophysical techniques such as CD, DSC, and NMR to
confirm their results. Using far-UV CD, it was reported that methionine oxidation
led to an ellipticity decrease of about 10% at 218 nm, indicating a small but detect-
able change in the secondary structure of the protein. Near-UV CD spectra showed
a positive increase in ellipticity indicating that Met oxidation also leads to changes
in the tertiary structure of the protein as well. The authors also used 2D, 1H–15N
HSQC NMR experiments to corroborate their findings on the structural impact of
Met oxidation. The resonances of many of amino acids in close proximity to the
oxidized Met are affected post H2O2 treatment indicating that upon oxidation, the
protein undergoes a dramatic conformational change. Using transgenic mice with
human FcRn, Wang et al. (2011) have recently reported that oxidized Met residues
(Met 252 and Met 428) have significantly shorter serum half-life.
190 S. Alavattam et al.
Bertolotti-Ciarlet et al. (2009) have used surface plasmon resonance and cell
binding assays to study the impact of methionine oxidation on the binding of two
humanized IgG1 antibodies to Fcγ receptors and to the neonatal Fc receptor (FcRn).
SPR analysis showed an increase in kd values that was approximately proportional
to level of Met oxidation, reaching a several-fold higher value in highly oxidized
species. The authors conclude that while Met oxidation did not result in substantial
changes to Fcγ binding (except FcγRIIa), binding to FcRn was significantly affected.
Wang et al. (2011) further evaluated the impact of Met oxidation on serum half lives
of two humanized IgG1 mAbs in transgenic mice with human FcRn. Results
obtained from these studies corroborate the fact that Met oxidation leads to signifi-
cant reduction of half-life in serum after Met oxidation in the Fc region of the mAb
and correlates directly to the level of Met oxidation and binding constants as mea-
sured by SPR. The above studies indicate that techniques that help understand and
prevent methionine oxidation during mAb production would be a significant addi-
tion to the arsenal used in characterizing proteins.
8.3.5 B
iophysical Techniques in High-Concentration
Formulation Development
[h]c
h = h0 exp (8.1)
1 − k [h]c
v
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 191
Fig. 8.13 Viscosity of mAbs1–3 as a function of protein concentration. mAb1, mAb2, and mAb3
aqueous samples have a composition of 266 mM sucrose, 16 mM histidine, and 0.03% polysorbate
20 at pH 6.0. mAb2 reconstituted lyophilized samples are at either 100 mg/mL MAb2 in 240 mM
trehalose, 40 mM histidine, and 0.04% polysorbate 20 or 125 mg/mL in 300 mM trehalose, 50 mM
histidine, and 0.05% polysorbate 20, pH 6. The solid curves are the result of a nonlinear regression
of all the data to the modified Mooney equation using a solvent viscosity of 1.1 mPas and intrinsic
viscosity of 6.3 cm3/g (adapted from Liu et al. 1995)
where h0 is the solution viscosity, η is the solvent viscosity, [η] is the intrinsic vis-
cosity of the protein, k is a “crowding factor,” and ν is the Simha parameter related
to the ellipsoid of revolution used to model the protein. The viscosity–concentration
profile for mAb1, on the other hand, cannot be described by the hard-sphere model
(Fig. 8.13, Liu et al. 2005) suggesting that there are other interactions in addition to
those involving excluded volume that mitigate the viscosity behavior of this particu-
lar mAb. It was hypothesized by Liu et al. that net-attractive interactions that result
in reversible concentration-dependent self-association increase the viscosity of a
concentrated monoclonal antibody in aqueous solution. At low concentrations the
three monoclonal antibodies are essentially monomeric in solution as determined by
SEC analysis. Thus, in order to determine if, in fact, mAb1 undergoes self-associa-
tion at high concentration, the analysis needs to be done at high concentration. This
is a challenging task since most analyses are done at low concentrations, typically
at 1 mg/mL. AUC sedimentation equilibrium has been used to investigate protein
self-association at high concentration by using thin plastic gaskets as centerpieces
(Minton and Lewis 1981), but this is not an easy method due to deformations of the
gasket material when torquing the AUC cells. The thin plastic gaskets that result in
a narrow cell pathlength are necessary in order to deal with the high refractive index
192 S. Alavattam et al.
that results in a deviation of the light path away from the detecting photomultiplier
in the centrifuge (Gonzalez et al. 2003). An alternative technique previously
described by Minton is preparative AUC (Minton 1989). In this method a prepara-
tive centrifuge is used at low speeds to generate sedimentation equilibrium concen-
tration gradients. After equilibrium is attained the centrifuge tubes are loaded onto
a low-volume fraction collector (Brandel®, Brandel Inc, Gaithersburg, MD), and
5 μL fractions are dispensed into a 96-well plate. All the wells are then diluted to
bring the absorbance reading down to within the dynamic range of the UV plate
reader (Fig. 8.14). The resulting absorbance as a function of radial position is then
used to determine the apparent weight-average molecular weight, Mw, app, at each
radial position using the following equation:
_
M w,app (1 − v r )(r 2 − r0 2 )
C (r ) = C0 exp (8.2)
2 RT
where c(r) is the protein_ concentration at radial position r, c0 is the initial loading
protein concentration, v is the partial specific volume, ρ is the buffer density, ω is
the angular velocity, and r0 is the reference radial position. At these high concentra-
tions there is a huge amount of non-ideality. Since mAb2 and mAb3 viscosity can
be accounted for using the Mooney equation, it is assumed that mAb2 and 3 exist
mainly as monomeric molecules in solution, and the non-ideality correction for the
charge and excluded volume effects can be obtained in the absence of added NaCl
for an IgG1 monomer with 150 kDa molecular weight. These corrections are
obtained from the relationship between apparent molecular weight, Ma, at weight/
volume concentration c and actual molecular weight, M (Chatelier and Minton
1987):
d ln g
M a = M 1 + c (8.3)
dc
where γ is the activity coefficient of the monoclonal antibody. Assuming that mAb2
and 3 are essentially monomeric in solution leads to a multiplicative correction fac-
tor as a function of c when M is set equal to 150 kDa. The resulting corrected molec-
ular weight for mAb1 (Fig. 8.14) supports the hypothesis that this monoclonal
antibody undergoes a concentration-dependent reversible self-association. Another
technique, SLS, at high concentrations has essentially corroborated the AUC analy-
sis (Scherer et al. 2010). A rigorous analysis of the SLS data shows that mAb1 self-
associates much more than mAb2. Thus, while literature has focused on
protein–protein interactions in dilute solutions, this work has emphasized the need
to develop newer techniques to understand intra-protein and inter-protein interac-
tions in highly concentrated protein solutions, especially because self-association of
proteins under such conditions appears to be crucial to appreciate the underlying
solution viscosity (Yadav et al. 2010). Kanai et al. proposed that the observed self-
association of mAb1, at pH 6.0, originates from multiple Fab–Fab interactions.
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 193
More recently, Yadav et al. (2011) showed that the presence of specific attractive
interactions at pH 6.0 leads to the self-association and high viscosity of mAb-1 at
high concentrations. They also demonstrate that exposed charged residues, espe-
cially histidyl residues, in the CDR of mAb1 are critical in determining the self-
associating and highly viscous behavior observed at high concentrations. The
authors used various biophysical techniques including CD, sedimentation equilib-
rium, high-frequency ultrasonic rheology, DLS, zeta potential, and net charge
194 S. Alavattam et al.
8.3.6 B
iophysical Techniques to Support Clinical
In-Use Studies
Pharmacists are responsible for setting a “beyond use” date based on USP 797,
wherein the beyond use date for the compounded sterile preparation (CSP) is
defined as the time by which the compounded preparation must be used to avoid
risks for product degradation, contamination, etc. Physical and chemical stability of
the CSP can be difficult to maintain over extended storage, especially since the
formulation components are diluted within the IV bag contents. Recent published
reports have suggested the use of extended time, beyond that recommended by the
manufacturer, for the storage and administration of CSP. These recommendations
were based on inadequate analytical testing of the CSP. Alavattam et al. 2012 dem-
onstrate that setting of the beyond use date should be carefully assessed using the
appropriate biophysical and analytical methods, given the fact that important excipi-
ents that help stabilize the protein get diluted upon dilution in saline IV bags. This
study shows that as the ratio of mAb to polysorbate 20 increased, the levels of aggre-
gates also increased with agitation as detected by the SEC assay. This suggests that
there is less surfactant to compete with the protein for the air–liquid interface in the
IV bag, thereby allowing mAb to aggregate at the interface. This study also suggests
that while the ratio of mAb to polysorbate 20 is an important contributing factor to
form soluble aggregates, headspace also played a key role. No clear trending was
observed in the subvisible particle analysis using the light obscuration (HIAC–
Royco) method. When performing particle analysis, there may be variability in the
particulate measurements attributed to sample handling and/or instrument accuracy,
and care must be taken to ensure the proper procedures are followed (Cao et al.
2010). However, it was evident that mAb1 and mAb2 with 0% polysorbate had a
larger number of subvisible particles at the ≥10 μm size in bags that contained head-
space, suggesting at least 0.0001% polysorbate may be necessary to protect the
protein at the air–water interface from particulate formation. It appeared that sub-
visible particles increased upon agitation in IV bags with headspace in mAb1 solu-
tions that did not contain polysorbate 20, whereas the subvisible particles were not
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 195
significantly different for mAb2 under similar solution c onditions. When mAb3 was
diluted tenfold into PO IV bags and subjected to agitation, a significant increase in
soluble aggregates was observed (13.5% over 1 h). A larger increase in soluble
aggregates was observed for mAb3 than for mAb1 and mAb2 over similar time
scales despite higher polysorbate 20 levels in the IV bags illustrating the high sen-
sitivity of mAb3 to agitation. Visible precipitation of the protein was observed on
further agitation of mAb3 in the IV bag, and no additional testing was performed on
samples that were agitated for longer periods of time. A clear correlation between
subvisible particles and soluble aggregates could not be found in this study.
Kumru et al. (2012) recently reported the physical stability of an IgG4 monoclo-
nal antibody upon dilution into intravenous (i.v.) bags containing 0.9% saline.
Soluble aggregates and subvisible particles were characterized using a variety of
analytical (including SEC) and biophysical methods such as light obscuration,
nanoparticle tracking analysis (NTA), microflow-digital imaging (MFI), and turbid-
ity measurements. Characterization studies with FTIR microscopy and extrinsic
fluorescence spectroscopy demonstrated that isolated particles contained native-like
secondary structure with partially altered tertiary structure, compared with heat-
denatured and non-stressed controls. Transmission electron microscopy (TEM) and
MFI analysis showed particles had an amorphous morphology of varying sizes.
Demeule et al. (2009b) recently reported a case study on characterizing trastu-
zumab samples diluted in either saline or dextrose solutions using asymmetrical
flow field-flow fractionation (FFF), fluorescence spectroscopy, fluorescence micros-
copy, and TEM. When trastuzumab samples were analyzed by FFF using a standard
separation method, no difference could be seen between trastuzumab diluted in
sodium chloride and trastuzumab diluted in dextrose. However, during FFF mea-
surements made with appropriate changes in the protocol, trastuzumab aggregates
were detected in 5% dextrose. This suggests that trastuzumab aggregates are over-
looked if the FFF analysis of trastuzumab diluted in dextrose solution is performed
using 0.9% NaCl. The authors were also able to confirm aggregates using fluores-
cence microscopy and TEM.
All the above studies emphasize the need to perform orthogonal biophysical tests
to clearly understand different mechanisms for protein degradation in IV bags and
support the proper use of protein therapeutics in clinical trials.
8.3.7 B
iophysical Techniques in Antibody Drug Conjugate
Formulation Development
has been one of the more extensively used techniques to confirm SEC
analysis(Berkowitz 2006), as described in Chap. 5.
Fig. 8.16 Concordance plot of AUC sedimentation velocity and SEC analysis of a monoclonal anti-
body. AUC, analytical ultracentrifugation; HMWS, high-molecular weight species; SEC, size exclu-
sion high-performance liquid chromatography. The error bars represent two standard deviations from
n = 3 determination. All other data points denote a single determination. Circles denote samples that
have HMWS levels below the LOQ of the AUC technique (Gabrielson and Arthur 2011)
8.4.2 F
low Field-Flow Fractionation
as an Orthogonal Technique
Another orthogonal technique that has been used to characterize protein aggregates
is FFF (Liu et al. 2006; Rambaldi et al. 2011). This technique is a flow-based sepa-
ration method, whereby separation is achieved by applying an externally generated
field that is perpendicular to laminar flow within a buffer filled open channel.
Although several different external fields have been used (Giddings 2000), charac-
terization of protein aggregates has been done mainly by application of solution
flow as the external perpendicular field and is termed flow FFF. This method,
which covers a wide range of molecular sizes from 0.001 to 50 μm in size, provides
assessments of size and quantity of protein aggregates without the use of standards
or solid-state matrices. Among all the flow-based FFF methods, asymmetrical flow
FFF (AF4), where only the bottom wall of the channel has a semipermeable mem-
brane, is the one most commonly used for therapeutic proteins (Fraunhofer and
200 S. Alavattam et al.
Table 8.3 Analysis of a mixture of three proteins (described in the text) using AF4, SEC, and
AUC
% Species
% % Species % Species AUC
MW Species AF4 (n = 5, Accuracya SEC (n = 5, Accuracya (n = 6, Accuracya
Samples (kDa) (actual) mean ± σ) AF4 (%) mean ± σ) SEC (%) mean ± σ) AUC (%)
Protein I 150 32.99 31.97 ± 0.13 1.02 32.80 ± 0.03 0.19 31.1 ± 0.00 1.9
Protein II 100 33.88 35.53 ± 0.08 −1.65 34.31 ± 0.05 −0.43 33.0 ± 0.00 0.9
Protein III 50 33.13 32.50 ± 0.05 0.63 32.87 ± 0.04 0.26 35.9 ± 0.00 −2.8
a
Accuracy was determined by subtracting the experimental values from the actual values
Winter 2004). The applied cross flow drives macromolecules towards the
membrane, and differences in diffusion coefficients create a concentration distribu-
tion in the laminar flow, resulting in different elution times. These elution times
allow for determination of the translational diffusion coefficient, which can be
used to compute the molecular weight assuming spherical structures for all spe-
cies. Although there is no solid-state support for potential interaction of protein,
there have been issues with protein interaction with the channel membrane. Liu
et al. (2012) have investigated this problem and show that with correct choice of
membrane and solvent conditions, it is possible to use this technology to confirm
results by SEC. In particular, they compared the analysis of three mAb prepara-
tions, protein I (a full-length monoclonal antibody), protein II (a single-armed anti-
body, i.e., with only one Fab), and protein III (a Fab fragment) using SEC, AUC,
and AF4. The results show very good agreement between all three techniques
(Table 8.3) demonstrating that AF4 is a viable technique that can be used for con-
firmatory studies.
This chapter has covered a variety of biophysical techniques, which have been used
in the development of protein biopharmaceuticals, including some perspective of
formulation development and long-term stability. It has not meant to be an all-
encompassing review but rather to demonstrate in what areas of development bio-
physics has been of use to aid in the production of a stable drug product and to
compliment the other chapters in this book. We have not discussed the use of differ-
ent biophysical methods to characterize protein particulates since that was covered
in Chap. 4. We also have highlighted some of the ways we have used biophysical
analysis at Genentech and hope this will be useful for researchers who wish to apply
such technologies.
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 201
References
Gabrielson JP, Arthur KK (2011) Measuring low levels of protein aggregation by sedimentation
velocity. Methods 54(1):83–91
Gibson TJ et al (2011) Application of a high-throughput screening procedure with PEG-induced
precipitation to compare relative protein solubility during formulation development with IgG1
Monoclonal antibodies. J Pharm Sci 100(3):1009–1021
Giddings JC (2000) The field-flow fractionation family: underlying principles. In: Schimpf ME,
Caldwell KD, Giddings JC (eds) Field-flow fractionation handbook. Wiley, New York
Gonzalez JM, Rivas G, Minton AP (2003) Effect of large refractive index gradients on the perfor-
mance of absorption optics in the Beckman XL-A/I analytical ultracentrifuge: an experimental
study. Anal Biochem 313(1):133–136
He F et al (2010a) Detection of IgG aggregation by a high throughput method based on extrinsic
fluorescence. J Pharm Sci 99(6):2598–2608
He F et al (2010b) High-throughput dynamic light scattering method for measuring viscosity of
concentrated protein solutions. Anal Biochem 399(1):141–143
He F et al (2011) Screening of monoclonal antibody formulations based on high-throughput ther-
mostability and viscosity measurements: design of experiment and statistical analysis. J Pharm
Sci 100(4):1330–1340
Hollander I, Kunz A, Hamann PR (2008) Selection of reaction additives used in the preparation of
monomeric antibody-calicheamicin conjugates. Bioconjug Chem 19(1):358–361
Houde D et al (2009) Characterization of IgG1 conformation and conformational dynamics by
hydrogen/deuterium exchange mass spectrometry. Anal Chem 81(14):2644–2651
Hvidt A, Linderstrom-Lang K (1954) Exchange of hydrogen atoms in insulin with deuterium
atoms in aqueous solutions. Biochim Biophys Acta 14(4):574–575
Jackson M, Mantsch HH (1995) The use and misuse of FTIR spectroscopy in the determination of
protein structure. Crit Rev Biochem Mol Biol 30(2):95–120
Ji JA et al (2009) Methionine, tryptophan, and histidine oxidation in a model protein, PTH: mecha-
nisms and stabilization. J Pharm Sci 98(12):4485–4500
Johnson WC Jr (1990) Protein secondary structure and circular dichroism: a practical guide.
Proteins 7(3):205–214
Kamerzell TJ et al (2011) Protein-excipient interactions: mechanisms and biophysical character-
ization applied to protein formulation development. Adv Drug Deliv Rev 63(13):1118–1159
Kumru OS et al (2012) Compatibility, physical stability, and characterization of an IgG4 monoclo-
nal antibody after dilution into different intravenous administration bags. J Pharm Sci
101(2):3636–3650
Lauer TM et al (2012) Developability index: a rapid in silico tool for the screening of antibody
aggregation propensity. J Pharm Sci 101(1):102–115
Lehermayr C et al (2011) Assessment of net charge and protein-protein interactions of different
monoclonal antibodies. J Pharm Sci 100(7):2551–2562
Li CH et al (2011) Applications of circular dichroism (CD) for structural analysis of proteins:
qualification of near- and far-UV CD for protein higher order structural analysis. J Pharm Sci
100(11):4642–4654
Liu J et al (1995) Characterization of complex formation by humanized anti-IgE monoclonal anti-
body and monoclonal human IgE. Biochemistry 34(33):10474–10482
Liu J, Ruppel J, Shire SJ (1997) Interaction of human IgE with soluble forms of IgE high affinity
receptors. Pharm Res 14(10):1388–1393
Liu J et al (2005) Reversible self-association increases the viscosity of a concentrated monoclonal
antibody in aqueous solution. J Pharm Sci 94(9):1928–1940
Liu J, Andya JD, Shire SJ (2006) A critical review of analytical ultracentrifugation and field flow
fractionation methods for measuring protein aggregation. AAPS J 8(3):E580–E589
Liu D et al (2008) Structure and stability changes of human IgG1 Fc as a consequence of methio-
nine oxidation. Biochemistry 47(18):5088–5100
Liu J et al (2012) Assessing and improving asymmetric flow field-flow fractionation of therapeutic
proteins. In: Williams KR, Caldwell KD (eds) Field flow fractionation in biopolymer analysis.
Springer, New York
8 Biophysical Analysis in Support of Development of Protein Pharmaceuticals 203
Lu Y et al (2008) The effect of a point mutation on the stability of IgG4 as monitored by analytical
ultracentrifugation. J Pharm Sci 97(2):960–969
Manta B et al (2011) Tools to evaluate the conformation of protein products. Biotechnol
J 6(6):731–741
Minton AP (1989) Analytical centrifugation with preparative ultracentrifuges. Anal Biochem
176(2):209–216
Minton AP, Lewis MS (1981) Self-association in highly concentrated solutions of myoglobin: a
novel analysis of sedimentation equilibrium of highly nonideal solutions. Biophys Chem
14(4):317–324
Patel AR, Kerwin BA, Kanapuram SR (2009) Viscoelastic characterization of high concentration
antibody formulations using quartz crystal microbalance with dissipation monitoring. J Pharm
Sci 98(9):3108–3116
Philo JS (2006) Is any measurement method optimal for all aggregate sizes and types? AAPS J
8(3):E564–E571
Rambaldi DC, Reschiglian P, Zattoni A (2011) Flow field-flow fractionation: recent trends in pro-
tein analysis. Anal Bioanal Chem 399(4):1439–1447
RNCOS (2012) Global-protein-therapeutics-market-forecast-to-2015. http://www.pharmaceutical-
market-research.info/research/PMAAAWUF-Global-Protein-Therapeutics-Market-
Forecast-to-2015.shtml
Robinson NE, Robinson AB (2004) Prediction of primary structure deamidation rates of asparagi-
nyl and glutaminyl peptides through steric and catalytic effects. J Pept Res 63(5):437–448
Ross PD, Minton AP (1977) Hard quasispherical model for the viscosity of hemoglobin solutions.
Biochem Biophys Res Commun 76(4):971–976
Saluja A, Kalonia DS (2004) Measurement of fluid viscosity at microliter volumes using quartz
impedance analysis. AAPS PharmSciTech 5(3):e47
Saluja A, Kalonia DS (2005) Application of ultrasonic shear rheometer to characterize rheological
properties of high protein concentration solutions at microliter volume. J Pharm Sci
94(6):1161–1168
Saluja A et al (2007) Ultrasonic storage modulus as a novel parameter for analyzing protein-
protein interactions in high protein concentration solutions: correlation with static and dynamic
light scattering measurements. Biophys J 92(1):234–244
Saluja A et al (2010) Diffusion and sedimentation interaction parameters for measuring the second
virial coefficient and their utility as predictors of protein aggregation. Biophys J 99(8):2657–2665
Samra HS, He F (2012) Advancements in high throughput biophysical technologies: applications
for characterization and screening during early formulation development of monoclonal anti-
bodies. Mol Pharm 9(4):696–707
Scherer TM et al (2010) Intermolecular interactions of IgG1 monoclonal antibodies at high con-
centrations characterized by light scattering. J Phys Chem B 114:12948
Schuck P (2000) Size-distribution analysis of macromolecules by sedimentation velocity ultracen-
trifugation and Lamm equation modeling. Biophys J 78(3):1606–1619
Shire SJ, Holladay LA, Rinderknecht E (1991) Self-association of human and porcine relaxin as
assessed by analytical ultracentrifugation and circular dichroism. Biochemistry
30(31):7703–7711
Shire SJ, Shahrokh Z, Liu J (2004) Challenges in the development of high protein concentration
formulations. J Pharm Sci 93(6):1390–1402
Stafford WF 3rd (1992) Boundary analysis in sedimentation transport experiments: a procedure
for obtaining sedimentation coefficient distributions using the time derivative of the concentra-
tion profile. Anal Biochem 203(2):295–301
Stephan JP, Kozak KR, Wong WLT (2011) Challenges in developing bioanalytical assays for char-
acterization of antibody-drug conjugates. Bioanalysis 3(6):677–700
Tartaglia GG et al (2005) Prediction of aggregation rate and aggregation-prone segments in poly-
peptide sequences. Protein Sci 14(10):2723–2734
Tartaglia GG et al (2008) Prediction of aggregation-prone regions in structured proteins. J Mol
Biol 380(2):425–436
204 S. Alavattam et al.
9.1 Introduction
T. Wang • S.B. Joshi • O.S. Kumru • S. Telikepalli • C.R. Middaugh • D.B. Volkin (*)
Department of Pharmaceutical Chemistry, Macromolecule and Vaccine Stabilization Center,
University of Kansas, 2030 Becker Drive, Lawrence, KS 66047, USA
e-mail: [email protected]; [email protected]; [email protected]; [email protected];
[email protected]; [email protected]
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 205
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2_9,
© Springer Science+Business Media New York 2013
206 T. Wang et al.
also increased. This situation leads to improved product quality with lower and
more consistent impurity levels. Similarly, potentially more stable protein formula-
tions may be developed in the future as our ability to monitor and control the forma-
tion of protein aggregates and particles improves with the introduction of more
sensitive biophysical methods (Carpenter et al. 2010b; Mire-Sluis et al. 2011).
The case studies in this chapter are presented as a series of illustrative examples
based on the size of the protein aggregates being examined. This approach permits
better comparisons across laboratories as investigators study different types of
aggregates from a wide range of proteins using a variety of different physicochemi-
cal methods (Narhi et al. 2011). Case studies examining the use of biophysical
methods to better characterize soluble protein aggregates in the size range of
1–100 nm are first presented followed by examples utilizing different biophysical
techniques to better characterize submicron-sized protein particles (0.1–1 µm). The
last section then presents examples of monitoring and evaluating larger subvisible
(1–100 µm) and visible (>100 µm) particles in protein formulations. For each sec-
tion, a brief overview of the methods used in the case studies are presented followed
by descriptions of recent literature examples that have employed a combination of
biophysical methods to better characterize the nature and composition of protein
aggregates and particles formed from different environmental stresses.
A variety of analytical approaches have been used to examine the size, amount, and
nature of soluble protein aggregates under both nondenaturing native and denatur-
ing conditions. For example, size-exclusion HPLC (SE-HPLC), sedimentation
velocity analytical ultracentrifugation (SV-AUC), dynamic light scattering (DLS),
and field-flow fractionation (FFF) techniques are commonly performed using non-
denaturing solution conditions to determine the extent and size of soluble protein
aggregates. Under denaturing conditions, that is, in the presence of additives such as
sodium dodecyl sulfate (SDS) or urea, protein aggregates are typically detected by
electrophoresis (SDS-PAGE, capillary SDS) or chromatography (SE-HPLC with
SDS or urea in the mobile phase, d-SEC). These electrophoretic and chromato-
graphic approaches are generally used to better characterize the size, amount, and
nature of soluble protein aggregates (e.g., presence or absence of native and nonna-
tive covalent, disulfide cross-links). More recently, mass spectrometry has been
employed to determine the size and nature of soluble protein aggregates. A brief
overview of the key analytical techniques used in the specific case studies described
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 207
in this chapter is provided below, followed by a few illustrative examples from the
literature which have used combinations of these analytical techniques to evaluate
the size, amount, and nature of different types of soluble protein aggregates in
the size range of 1–100 nm.
Due to its ease of use and high sample throughput, size-exclusion high-performance
liquid chromatography (SE-HPLC) is the most commonly utilized technique for
sizing and quantifying soluble protein aggregates in the molecular weight range of
5–1,000 kDa (Mahler et al. 2009; Philo 2009; Arakawa et al. 2010). Larger protein
aggregates cannot be accurately evaluated for their size since they migrate in the
void volume of the column or detected at all if they cannot pass through the column
matrix. Typically, larger aggregates need to be removed by filtration or centrifuga-
tion prior to injection onto the SE-HPLC column. One disadvantage of SE-HPLC
for use in sizing soluble protein aggregates is that the molecular weight (MW)
calibration standards typically used with this method are globular in nature. In con-
trast, protein aggregates are not necessarily spherical resulting in inaccurate MW
estimates. In addition, during sample preparation or during the chromatography
process, protein aggregates may reversibly associate or dissociate and will therefore
either not be detected or not represent the equilibrium state of interest (Carpenter
et al. 2009; Mahler et al. 2009). Nonetheless, SEC protein stability studies are abun-
dant and are a well-accepted approach (Oliva et al. 2001; Gabrielson et al. 2007;
Kiese et al. 2008, 2010; Tyagi et al. 2009; Van Buren et al. 2009; Bond et al. 2010;
Nayak et al. 2011). Since an often major limitation of SE-HPLC is the potential for
aggregates to interact with the column matrix, orthogonal methods such as analyti-
cal ultracentrifugation, FFF, or DLS, which monitor the hydrodynamic properties of
proteins free in solution, are now being more frequently used to verify SE-HPLC
results (Carpenter et al. 2010a; Bond et al. 2010).
An analytical approach often used to confirm SE-HPLC results is SV-AUC (Liu and
Shire 1999; Gabrielson and Arthur 2011). Briefly, an optical centrifugation cell is
loaded in a rotor that contains two chambers, one sample and one reference, with
detection of protein concentration in the cells by UV absorbance. The sample is
centrifuged at high g-force forming a moving boundary that migrates to the outside
of the rotor over time. Lower molecular weight species take longer to sediment than
their higher molecular weight (HMW) counterparts and this behavior is monitored
by absorbance at 280 nm over time. Software data analysis packages (e.g., SEDFIT)
are then used to calculate the percentage of each protein species based on their rela-
tive sedimentation time (Dam and Schuck 2004; Schuck 2000, 2004). Compared to
SE-HPLC, SV-AUC requires no dilution of the protein into a mobile phase and is
208 T. Wang et al.
Various light scattering techniques have been employed to monitor protein aggrega-
tion in solution. DLS not only determines the size of protein aggregates in the 1 nm
to 1 µm size range but also can provide information about the polydispersity of the
sample (Kiese et al. 2008, 2010; Mahler et al. 2005). A DLS instrument records the
time-dependent fluctuations in the amount of scattered light when particles are
moving under Brownian motion. From the changes in these intensity fluctuations, it
is possible through autocorrelation analysis to determine a diffusion coefficient of
the particles and, through the Stokes–Einstein equation, their hydrodynamic radius
(Lomakin et al. 1999). Sizes can be estimated by an averaging technique known as
cumulant analysis or from various forms of deconvolution analysis that to a limited
extent resolve individual populations that differ in size by a factor of two or more.
As examples of the utility of DLS for monitoring protein aggregation, Mahler et al.
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 209
(2005) and Kiese et al. (2008) have monitored the aggregation behavior of IgG1
monoclonal antibodies during stirring and agitation under a variety of solution con-
ditions. Static light scattering (SLS) can monitor protein aggregation by measuring
changes in the average intensity of scattered light and if performed at different
angles and concentrations can determine changes in a protein molecule’s radius of
gyration (Murphy 1997). The radius of gyration of species in protein solutions can
also be determined by small angle neutron scattering (SANS), for example, in the
detection of microaggregates in different insulin preparations (Heldt et al. 2011).
There are now numerous examples using SE-HPLC in conjunction with UV absor-
bance, refractive index, and multiangle light scattering (MALS) detectors to obtain
absolute molecular weight of soluble, aggregated protein species (Bond et al. 2010;
Wen et al. 1996; Ye 2006). In addition, UV optical density spectroscopy is routinely
used to monitor the presence of aggregation through increases in optical density at
320–350 nm (Mach and Middaugh 2011). This approach has been widely used to
assess the ability of pharmaceutical excipients to minimize protein aggregate for-
mation in a high-throughput format (Bhambhani et al. 2012).
The size, extent, and nature of soluble protein aggregates can also be evaluated
under denaturing conditions. Electrophoresis is a standard way to study protein
aggregation, in the molecular weight range of 5–500 kDa (Mahler et al. 2009), using
an electric field to separate macromolecules based on their charge and weight. The
most common electrophoretic technique used to separate and quantify protein
aggregates of different molecular weight is SDS polyacrylamide gel electrophoresis
(SDS-PAGE). SDS-PAGE not only allows size estimation but can also be used to
estimate purity levels and evaluate for the presence of covalent, reducible (e.g.,
disulfide-linked) aggregates (Mahler et al. 2009). Noncovalent protein aggregates,
however, are generally not detected since they are usually disassociated during sam-
ple preparation due to the presence of SDS. Capillary electrophoresis (CE) offers an
attractive alternative to conventional gel electrophoresis due to its robustness, speed,
enhanced resolution, and reproducibility as well as its automation and quantitative
features. CE techniques are now routinely used in the biopharmaceutical industry in
the quality control environments for purity testing as well as assessing stability and
integrity of protein drugs. Although a number of CE modes are available, the most
common technique used is capillary electrophoresis SDS (CE-SDS). A detailed
description of these techniques can be found elsewhere (Kraly et al. 2006; Rustandi
et al. 2008; Kilar 2003). In addition to electrophoresis, SE-HPLC can be run with
SDS, urea, or guanidine hydrochloride in the mobile phase (denaturing SE-HPLC or
d-SEC) to separate disrupted protein aggregates under both reduced and nonreduced
conditions (Michels et al. 2007). Often, dissociation of aggregates occurs prior to
unfolding of the monomeric species as the concentration of the unfolding agent is
increased permitting further analysis. CE-SDS is more commonly used in quality
control environments to quantitatively determine protein aggregate size and purity
210 T. Wang et al.
intact proteins are directly analyzed in the mass spectrometer without prior
enzymatic digestion. This approach gives a molecular mass of the intact protein,
their major posttranslational modifications and can potentially detect the pres-
ence of protein aggregates.
Not only have both of these two approaches been used to examine protein aggre-
gates, but they have also been coupled to the analysis of kinetic exchange rates of
hydrogen for deuterium in the protein amide backbone (H/D exchange-MS), to gain
insight into the role of protein flexibility and dynamics in protein aggregation path-
ways. The rate of H/D exchange in the amide backbone hydrogens in a protein dif-
fers in flexible or disordered regions (lacking stable hydrogen bonding) compared to
more tightly folded regions (protected from H/D exchange) (Konermann et al. 2011).
By quenching the H/D exchange reaction at different time points and enzymatically
digesting the protein under quenched conditions, the rate of H/D exchange in differ-
ent regions of the protein can be determined by LC–MS analysis of each peptide over
time (Marcsisin and Engen 2010). Furthermore, regions protected by aggregation
can be identified. Recent advances in chromatographic separations such as UPLC
technology, combined with software advances in the analysis of H/D exchange rates
in hundreds of peptides, has permitted the characterization of the conformational
dynamics of IgG monoclonal antibodies by H/D exchange-MS (Houde et al. 2009).
Surface plasmon resonance (SPR) is a method based on energy transfer from light
(photons) to the electrons on a metal surface (Ahrer et al. 2003). As polarized light
interacts with the surface of a biosensor chip coated with gold, an evanescent wave
is created that is highly sensitive to the changes in the dielectric constant of the adja-
cent medium. This energy transfer can create an angle of minimum reflectance (the
SPR angle) under certain conditions. The change in SPR angle is proportional to the
mass of bound material at the chip surface (Thillaivinayagalingam et al. 2010). By
analyzing real-time changes of the SPR angle (SPR sensorgrams) as fluid passes
over the surface of the biosensor, both the extent of analyte binding to the surface as
well as binding rate kinetics (kon, koff) and binding affinities (KD) can be obtained.
SPR has been extensively applied to the quantitative analysis of protein-ligand and
protein-protein interactions and facilitated by the availability of automated equip-
ment. One requirement for the application of SPR to biological molecules is the
successful immobilization of sufficient amounts of ligand to the surface of the sen-
sor chip. Chemical coupling methods (e.g., through amines, thiols) covalently bind
the targeted ligand onto the biosensor surface. Since a lack of control over the ori-
entation of protein ligands might affect subsequence analyte binding, indirect cou-
pling methods use a second ligand (e.g., an antibody or avidin) in which the target
ligand binds to the immobilized protein creating a more homogenously oriented
bound ligand (Wear et al. 2005).
212 T. Wang et al.
This section will cover the three case studies illustrating a variety of analytical
approaches to characterize different types of soluble protein aggregates, using
methods under native-like (SE-HPLC and AUC) and denaturing conditions
(SE-HPLC and CE with SDS). The first study uses a novel application of SE-HPLC,
along with orthogonal testing by SV-AUC, to improve the sensitivity of aggregate
detection for a variety of different types of aggregates of an IgG1 monoclonal anti-
body generated from different environmental stresses. The second study examines
similar analytical approaches to characterize a high molecular weight aggregate of
a recombinant therapeutic glycoprotein. The third utilizes analytical methods under
denaturing conditions (SE-HPLC and capillary electrophoresis with SDS) to deter-
mine the level of reducible and nonreducible protein aggregates in stressed samples
of a monoclonal antibody.
In the first case study, an IgG1 monoclonal antibody was subjected to different
environmental conditions such as elevated temperature, decreased pH, and light
exposure to generate different aggregates. Bond et al. were able to accurately, and
with improved sensitivity compared to light scattering or MALS, quantify the
amount of aggregate in force-degraded IgG1 monoclonal antibody solutions using
a dual-wavelength approach (DW-SE-HPLC) with very good recovery (Bond et al.
2010). This optimized SE-HPLC technique monitored aggregate, monomer, and
fragment peaks at two ultraviolet wavelengths (214 and 280 nm). It was determined
that the limit of quantitation for this new method was 0.04%, which was much lower
than the 0.2% found for the previously used single wavelength SE-HPLC method.
In addition, the new approach extended the dynamic range ~five to ten fold. As
shown in Fig. 9.1, different environmental stresses resulted in formation of different
types of oligomeric species and soluble protein aggregates (dimer vs. multimer as
well as covalent cross-links vs. noncovalent interactions). After assay optimization
to minimize interactions with the column matrix, the DW-SE HPLC results corre-
lated well with SV-AUC data in terms of the ability to quantify aggregate levels in
solutions (Fig. 9.1). Interestingly, a high-concentration formulation of the IgG1
monoclonal antibody, when stressed at 40°C for 3 months, showed a low-level
formation of a different type of oligomeric species, a noncovalently associated
dimer (Bond et al. 2010).
Several investigations have now appeared using SV-AUC to confirm the results
obtained with SE-HPLC, specifically to verify a lack of effect of SE-HPLC sample
handling and mobile phase conditions on aggregate formation and dissolution
(Berkowitz 2006; Hughes et al. 2009). In one interesting example with a complex
protein, Hughes and colleagues quantitatively investigated the ability of SE-HPLC
and SV-AUC to detect known amounts of protein aggregates of recombinant human
acid alpha-glucosidase (rhGAA), an enzyme with seven N-lined glycosylation
sites and an apparent molecular weight of 110 kDa. Overall, the authors found a
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 213
Fig. 9.1 Analytical characterization of IgG1 monoclonal antibody solutions containing soluble
aggregates measured by orthogonal methods, SE-HPLC (a, b, c) and SV-AUC (d, e, f). Protein
aggregates of varying size and composition were formed by exposure to different environmental
stresses: (a, d) control, (b, e) heating under acidic pH conditions, and (c, f) light exposure. Adapted
from Bond et al. (2010)
good correlation between the two methods, since the amount of the soluble
aggregates measured was not statistically different (Hughes et al. 2009) (Table 9.1).
It should be pointed out, however, that other studies in the literature have found
significantly higher percentages of soluble protein aggregates when measured by
SV-AUC compared to SE-HPLC. This suggests that the SE-HPLC approach may
require additional method development to prevent aggregate adsorption to the col-
umn matrix with certain proteins. For example, in work by Gabrielson et al. with a
214 T. Wang et al.
Table 9.1 Quantitative analysis of aggregate levels as determined by SV-AUC and SE-HPLC for
forced degraded samples of recombinant human acid alpha-glucosidase (rhGAA)
rhGAA forced aggregate (%)
0 0.25 0.5 1.25 2.5 4 5 10 20
% Aggregation (AUC), 0.4 0.9 0.6 1.3 2.5 3.1 4.2 8.5 18.1
N = 6a
SD 0.3 0.7 0.4 0.5 1.1 0.3 0.4 0.7 1.3
RSD (%) 67.0 84.0 69.4 35.7 44.0 8.8 10.4 8.1 7.0
% Aggregation (SEC), N = 9 0.2 0.3 0.5 1.0 1.9 3.0b 4.1 8.4 17.5
SD 0.02 0.03 0.02 0.04 0.03 0.04 0.04 0.04 0.42
RSD (%) 9.2 10.8 3.7 3.7 1.6 1.2 1.0 0.5 2.4
Published in Hughes et al. (2009)
a
Samples were analyzed in a single AUC run using six different cells
b
These SEC data were obtained on six samples
Fig. 9.2 Analytical characterization of monomers, dimers, and high molecular weight (HMW)
aggregates of a monoclonal antibody as measured by SE-HPLC (native-like conditions) and d-SEC
and CE-SDS (denaturing conditions). (a) SE-HPLC, (b) d-SEC, and (c) CE-SDS with indicated
levels of percent nonreducible aggregates. Published in Michels et al. (2007)
Despite recent advances in the use of biophysical methods to better characterize the
higher-order structure and conformational stability of proteins, especially when
combining multiple techniques with advanced data analysis approaches (Maddux
et al. 2011), the determination of a protein’s biological activity remains the “gold
standard” in terms of monitoring subtle conformational changes. Two case studies
are presented in this section, first demonstrating correlations of certain physical
properties of protein aggregates with biological activity (Remmele et al. 2006) and
second showing how biophysical binding assays using biosensors (SPR) can be
employed to assess the Fc receptor binding activities of different aggregates and
complexes of two different monoclonal antibodies (Luo et al. 2009).
In the case of an IgG1 monoclonal antibody (epratuzumab), samples were shown
to contain both monomers (~150 kDa) and dimers (~300 kDa). The dimers con-
tained ~70% covalent cross-links (as measured by CE-SDS) and consisted of three
different assemblies including Fab-Fab, Fc-Fc, and Fab-Fc complexes. The biologi-
cal activity of the dimers, in terms of relative potency in a cell-based bioassay, was
shown to be twice that of the monomer (i.e., equal activity on a weight basis indicat-
ing that the dimer has twice the number of binding sites as the monomer) (Remmele
et al. 2006). A more recent investigation examined the binding affinities of different
mAb complexes and aggregates using an SPR biosensor (Luo et al. 2009). Similar
to the previous study, monomeric and aggregated forms of the mAbs had similar
antigen-binding properties in the Fab region. It was demonstrated, however, that the
monomers and aggregates showed differences in their Fc region’s ability to bind
different Fc-gamma receptors. As shown in Fig. 9.3a, b, the monomeric form of
216 T. Wang et al.
Fig. 9.3 The Fc receptor binding of the immune complexes and aggregates of mAb1 (a and b) and
mAb2 (c and d) as measured by SPR binding assays. Two different Fc receptors were evaluated:
FcγRIIA (a and c) and FcγRIIIB (b and d). The SPR sensorgrams for the binding of mAb mono-
mer, dimer/HMW, and their immune complexes (containing 1:1 molar ratio of monomeric anti-
body and multivalent antigen) are overlaid. Published in Luo et al. (2009)
mAb1 manifested different binding properties to both the Fc-gamma RIIA and
Fc-gamma RIIIB receptors compared to the dimeric aggregate and a multivalent
immune complex formed between mAb1 and its antigen. Similar results are shown
in Fig. 9.3c, d for monomeric mAb2, high molecular weight aggregates of mAb2,
and a multivalent immune complex formed between the mAb2 and its antigen. The
dimers and multimers of these mAbs showed higher in vitro binding affinities to
Fc-gamma receptors as well (Luo et al. 2009).
Although the use of mass spectrometry to better characterize soluble protein aggre-
gates formed during storage has been more limited in its applications to date, MS
has been pursued for many years to better characterize well-ordered amyloid fibril
protein aggregates. For example, nanoflow ESI combined with Q-TOF mass
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 217
spectrometer has been utilized to study the self-assembly, aggregation, and amyloid
fibril formation of insulin under different conditions (Nettleton et al. 2000). Another
example is the use of ESI-quadrupole MS to quantify β(beta)-amyloid monomers in
the presence of potential therapeutic agents and to identify and rank order the com-
pounds which inhibit the aggregation of β(beta)-amyloid monomers (Cheng and
van Breemen 2005). The initial applications of MS technology to characterize pro-
tein aggregates under pharmaceutical conditions have combined chromatographic
separations with subsequent MS analysis. For example, Van Buren et al. character-
ized dimers of an IgG after isolation and enzymatic digestions (Van Buren et al.
2009). A combination of SE-HPLC, sample dialysis, and ESI-TOF MS was recently
used to identify and characterize different IgG aggregates (Kukrer et al. 2010). In
addition to complex sample handling with limited throughput, these MS approaches
have certain limitations including potential changes in the aggregate profile during
sample dialysis and a bias of MS results toward lower MW species when monitor-
ing a sample containing a mixture of oligomers (Kaltashov et al. 2012).
More sophisticated MS approaches are now being developed to examine and
characterize protein aggregates without these experimental limitations as described
in the following three case studies: (1) use of EDI-MS to directly monitor in real-
time heat-induced protein deformation and aggregation (Wang et al. 2011), (2) use
of ion mobility MS to characterize intact protein assemblies in the gas phase, and
(3) the use of H/D exchange-MS to characterize different proteins (Kheterpal and
Wetzel 2006; Zhang et al. 2011, 2012).
As temperature increases, loss of a protein’s conformational integrity results in
formation of high-charge-density ions and oligomers that can be easily identified by
higher m/z ratios. A new design of a temperature-controlled ESI source not only
permits direct monitoring of protein unfolding and aggregation but also ensures the
efficiency of the heating process and prevents cooling of the protein solution during
the sample introduction to the ESI interface (Wang et al. 2011). With this new
approach, both reversible and irreversible protein unfolding events were captured
by ESI-MS for cytochrome-C and glucocerebrosidase (GCase), respectively.
The ability of ESI mass spectra to directly monitor the formation of dimers, trimers,
tetramers, and pentamers during heat treatment of GCase is shown in Fig. 9.4.
Ion mobility spectrometry (IMS) is an analytical technique used to separate and
identify ionized molecules in the gas phase based on their mobility in a carrier buf-
fer gas. When coupled with mass spectrometry, ion mobility mass spectrometry
(IM-MS) enables the separation and detection of isomers and conformers. More
recently, the technology (also referred to as native MS) has been successfully used
to characterize larger macromolecules, such as proteins and their biological com-
plexes, in the gas phase (Uetrecht et al. 2010). Although Native MS has been
employed to examine the composition and stability of numerous large biological
complexes, confirmation of results with other methods (examining the same protein
complex in the solution state) is typically performed (van Duijn 2010). Kaddis et al.
used ESI-IMS to study the general size dimensions of proteins as protein molecules
were physically transitioned from the solution to the gas phase (Kaddis et al. 2007).
Traveling voltage-wave ion mobility mass spectrometry (TWIMS) is a recent
218 T. Wang et al.
advancement in the technology coupling an ion mobility unit to Q-TOF MS. The
application of TWIMS to the characterization of hepatitis B viral capsids (MW of
~3,000 and ~4,000 kDa) as well as GroEL and its complexes with smaller proteins
(MW of 801 and 857 kDa, respectively) has been recently reviewed (Uetrecht et al.
2010).
Hydrogen–deuterium exchange linked to MS (H/D exchange-MS) analysis has
been used to probe the secondary structure of amyloid fibrils of insulin (Kheterpal
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 219
and Wetzel 2006). This work combined an ESI-MS system with a bottom-up experi-
mental approach (peptide mapping). First, protonated fibrils were collected by
centrifugation, washed, resuspended, and incubated in deuterated buffer to promote
exchange of labile protons with deuterium. Second, the exchange reaction was
quenched and the fibrils were disaggregated. Finally, the monomers were subjected
to proteolysis and analyzed by MS. The results showed that both the N- and
C-terminal segments (1–19 and 35–40) are readily exchangeable (not involved in
the core structure), whereas the fragment containing residues 20–34 was highly
protected from exchange. H/D exchange-MS has also been used to the study of the
effect of protein flexibility and sequences on protein aggregate formation with bio-
pharmaceutical products including interferon-gamma (Tobler and Fernandez 2002)
and, more recently, with the model enzyme lactate dehydrogenase (LDH) (Zhang
et al. 2011) as well as a monoclonal antibody (Zhang et al. 2012). For smaller pro-
teins such as interferon-gamma and LDH, the rate of H/D exchange can be moni-
tored in the intact protein, as well as at the peptide level (after digestion with pepsin),
to identify sites of increased and decreased flexibility. In contrast, due to their large
size, mAbs are only analyzed after pepsin digestion and H/D exchange reactions
must be carefully examined across hundreds of peptides. The H/D mass spectra of
native and freeze-thaw (F/T)-induced aggregates of LDH are shown in Fig. 9.5. It
can clearly be seen that F/T-induced aggregates consist primarily of unfolded pro-
tein. Peptide level analysis identified several short sequences (13–31, 109–117, and
133–143) with moderate protection from H/D exchange indicating that these
sequences may be the regions of intermolecular interactions within the LDH aggre-
gates induced by freeze-thaw (Zhang et al. 2011). In contrast, in the case of a mono-
clonal antibody, F/T aggregates had similar H/D protection maps compared to
soluble protein, indicating a primarily native-like conformation (Zhang et al. 2012).
Fig. 9.5 Hydrogen–deuterium exchange mass spectrometry (H/D-MS) spectra of native lactate
dehydrogenase (LDH) and freeze/thaw aggregated LDH obtained after 10 s, 10 and 180 min of
deuterium labeling. The dotted spectra in the top and bottom panels are unlabeled and fully labeled
controls, respectively. Published in Zhang et al. (2011)
(QCMs). A brief overview of the key analytical techniques used in the case studies
presented in this chapter is provided below, followed by a few illustrative examples
from the literature which have employed useful combinations of these techniques.
generated due to the piezoelectric nature of the quartz material. When the mechanical
oscillations are close to the fundamental frequency of the crystal, a resonant oscil-
lation is achieved. Changes in the resonant frequency are related to the mass accu-
mulated on the crystal (Sauerbrey 1959). Commercial systems are designed to
reliably measure mass changes up to ~100 µg (microgram) with a sensitivity of
~1 ng cm−2 (O’Sullivan and Guilbault 1999). Although QCM has been used suc-
cessfully in other fields (e.g., vacuum deposition, thin film deposition control),
application to biological samples has been limited due to its inherent nonspecificity.
Recently, chemical modification of the electrode surface enabled the real-time study
of bovine insulin aggregation to form amyloid fibrils (Knowles et al. 2011).
Immersion of resonators into aqueous solutions, however, significantly reduces
their resolution due to the damping effect of solution viscosity (Burg et al. 2007).
A recently developed nanomechanical resonator, Archimedes, contains a solution
channel inside a hollow resonator that is surrounded by vacuum to overcome this
damping effect. As a result, Archimedes has improved mass resolution (orders of
magnitude) over a commercial QCM for aqueous measurements (Burg et al. 2007).
In a comparative study by Filipe et al. (2010), the ability of NTA and DLS to size
submicron particulates using polystyrene bead standards of various sizes was initially
examined. Although both techniques can accurately size monodisperse standards,
NTA offers the advantage of also being able to count the number of particles in the
sample. In addition, it was possible to visualize the particles in the samples using
NTA. When particle standards of different sizes were mixed together, NTA can more
easily distinguish their relative amounts compared to DLS. In this study, DLS detected
the largest particle standard in a mixture, while NTA detected each of the standards
and thereby did not skew the reported size distribution. The formation of submicron-
sized protein aggregates was then monitored with heat treated IgG and metal-oxidized
insulin. NTA was again better at analyzing sample polydispersity than DLS. As shown
in Fig. 9.6, NTA can be used to monitor particle counts and size distribution in real
time during the formation of IgG submicron-sized particulates during heating at 50°C.
Recently, Filipe et al. used NTA in conjunction with fluorescence single-particle
tracking analysis to obtain the size of fluorescently labeled monoclonal IgG and HSA
when subjected to heat stress (Filipe et al. 2011).
Nanomechanical resonators, including QCMs, have achieved extraordinary sen-
sitivity in the detection of mass accumulation on surfaces in a vacuum. The detec-
tion of most bimolecular samples, however, requires a solution environment. As
described above, a recently developed nanomechanical resonator, Archimedes,
includes a solution channel inside a hollow resonator that permits solution measure-
ments (Burg et al. 2007). This device is able to measure the mass of objects that are
not attached to the resonator surface, such as floating polystyrene beads, agglutinating
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 223
microspheres, and bacteria (Burg et al. 2007; Rumi Chunara et al. 2007; Michel
Godin et al. 2007). The instrument vendor’s website (http://www.affinitybio.com/
applications/protein_formulations.php) provides examples (from currently unpub-
lished data) of the detection of submicron-sized IgG aggregates. Both size and num-
ber distributions of protein particles in solution can be displayed in the form of a
histogram.
Fig. 9.6 Monitoring of IgG aggregation and submicron particle formation at 50°C in the NTA
(NanoSight) instrument sample chamber. The size distribution (middle panels) with the corre-
sponding NTA video frame (left panels) and 3D graph (size vs. intensity vs. concentration, right
panels) are shown. Published in Filipe et al. (2010)
and negative stain TEM to study the assembly of monoclonal antibody into highly
ordered structures in the presence of multivalent carboxylate ions (Esue et al. 2009).
As shown in Fig. 9.8, TEM images demonstrate smaller bundles that were composed
of straight, single filaments with an average diameter of 4 nm. In addition, the pres-
ence of larger bundles with diameters as large as 200 nm, containing midsized
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 225
Fig. 9.7 Atomic force microscopy (AFM) visualization of submicron-size protein particles from
a monoclonal antibody solution: (a) mAb aggregate (in water), (b) the same mAb aggregate in
different viewing area. Published in Lee et al. (2011)
Current analytical methods utilized to count and size subvisible and visible particu-
lates in the size range of 1–100 µm and over 100 µm, respectively, include coulter
counter, light obscuration (LO), microflow digital imaging (MFI), light microscopy,
and visual assessments. Techniques employed to better characterize the morphology
and composition of subvisible and visible protein particulates include zeta potential
measurement, flow cytometry, fluorescence, and FTIR microscopy as well as SEM-
EDX. A brief discussion of each technique will be presented below followed by
relevant case studies that have incorporated these techniques to size, count, and
characterize subvisible and visible protein particles.
226 T. Wang et al.
Fig. 9.8 Transmission electron microscopy (TEM) visualization of submicron and subvisible
protein particles from a monoclonal antibody solution containing citrate. Analysis of micrographs
showed large bundles containing smaller filament bundles. Individual straight filaments were of
varying lengths and had an average diameter of 4 nm. Scale bars are 1 µm and 100 nm. Published
in Esue et al. (2009)
The Coulter method counts subvisible particles in the size range of ~0.5–50 µm
(micrometer). Unlike LO and MFI, the Coulter method does not rely on the optical
characteristics of the particles, but rather on the perturbation of an electric field.
Single particles flow through a pore and electrical pulses are generated, which pro-
duce a change in conductance. This change is directly proportional to the particle
volume. To measure a change in conductance, the Coulter method requires a con-
ductive buffer to permit an electric field to form. In some cases, this can be accom-
plished by adding a small amount of highly concentrated NaCl (or other electrolyte)
to the buffer (Barnard et al. 2012). Since only single particles can transverse the
pore, the sample must be diluted. Both dilution and the addition of salt during sam-
ple preparation are of concern when monitoring protein particulates since these
steps may either disassociate aggregates or induce their formation. Until recently,
Coulter counter measurements required large sample volumes (10 ml), but newer
instruments have significantly decreased sample volumes to as little as ~0.1 ml
(Barnard et al. 2012; Rhyner 2011; Roberts et al. 2012). These reduced sample vol-
umes are far more practical for use with protein containing samples.
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 227
The principle of light obscuration (LO) is simple: the sample is passed through a
chamber containing a light source. The amount of light that is blocked is directly
proportional to the number and size of the particles. This technique is widely used
as the “gold standard” for measuring subvisible particles in pharmaceutical applica-
tions since both US and EU pharmacopeia methods have been described. The con-
centration and size of particles ranging from 2 to 100 µm (micrometer) can be
determined. Advantages of this technique include speed and the ability to cover the
entire subvisible size range in a single measurement. Disadvantages of LO include
a lack of morphological information and potential artifacts arising from air bubbles.
Therefore, samples are typically degassed prior to analysis (Huang et al. 2009;
Chrai et al. 1987). In addition, LO has been shown to undercount the number of
proteinaceous subvisible particles as described in more detail below.
During analysis by light microscopy, particles in solution are drawn through a grid-
lined filter, dried, and either observed by a standard light microscope or subsequently
stained with fluorescent or nonfluorescent dyes to enhance resolution (Li et al. 2007;
Demeule et al. 2007a, 2009). Particles can then be counted manually or by automated
methods. Advantages of staining protein aggregates include enhanced detection of
smaller particles and the selective binding of certain dyes to proteinaceous particles
(Li et al. 2007; Rosenberg et al. 2009). Artifacts resulting from sample preparation,
handling, and/or filtering are again the major pitfalls in using microscopy. In addi-
tion, particles <10 µm can be difficult to visualize and care should be taken to ensure
amorphous or fibril shaped particles do not pass through the filter.
228 T. Wang et al.
thousands of individual particles per second are measured using several detectors
(forward scatter, side scatter and fluorescent detectors). Light scattering and/or fluo-
rescence emission is unique to each cell/particle. Thus, a combination of the two
can be used to distinguish different particles or cells in a heterogeneous sample. The
capability of flow cytometer to characterize individual particles makes it an attrac-
tive technique for use in the study of subvisible particles/aggregates in protein for-
mulations. This technique has been widely used in the field of cellular and molecular
biology, but its potential for use in the study of subvisible particles in protein formu-
lations has only recently been recognized (Ludwig et al. 2011; Mach et al. 2011).
This type of microscopy can identify cells, cellular constituents, and particulate
matter with a high degree of specificity and is rapidly expanding in use for pharma-
ceutical applications. The technique is used to study samples that can fluoresce
either by themselves or after treating with fluorescent probes/dyes (Lichtman and
Conchello 2005). Fluorescence microscopy can be very useful for the detection of
subtle changes in the aggregation state of the proteins (David Bernard Williams
2009; Demeule et al. 2007a). Some of the dyes used to selectively monitor
intermolecular β(beta)-sheet interactions often present among proteins that form
β(beta)-amyloid structures are Nile red, Congo red, and Thioflavin T (Hawe et al.
2008a; Khurana et al. 2001; Naiki et al. 1989; Kim et al. 2003). Covalent labeling of
proteins with donor and acceptor fluorophores allows the use of total internal reflec-
tion fluorescence microscopy to image aggregated proteins (Hillger et al. 2007).
shape, and size of each material present can be visualized through these chemical
images. These images capture the spectral features unique to each material present
and provide chemical information about the multicomponent system. Creating a
map, however, is a time-consuming process especially if both high spatial resolution
and high signal to noise ratio as required (Prati et al. 2010).
Energy-dispersive X-ray spectrometer (EDX or EDS), which utilizes the X-ray gen-
erated when the electron beam bombards the sample to collect information on
chemical composition of a sample, can be connected to both TEM and SEM (Joseph
Goldstein et al. 2003; David Bernard Williams 2009).
Although MFI is a relatively new technique, there are already many interesting case
studies employing the technology to count and size subvisible particles in protein for-
mulations. Examples include measuring subvisible-sized protein particles induced by
contact with ceramic filling pumps or syringes (Nayak et al. 2011; Majumdar et al.
2011) as well as from exposure to different environmental stresses (Joubert et al. 2011;
Luo et al. 2011; Fradkin et al. 2011; Barnard et al. 2010). Several studies also include
comparisons of MFI or Coulter methods to light obscuration in terms of counting both
protein particles and polystyrene bead standards (Huang et al. 2009; Wuchner et al.
2010; Sharma et al. 2010a, b). The major difference in MFI and LO results for counting
subvisible protein particles is probably due to their translucent nature which results in
undercounting by LO. A recent study by Demule et al. (2010) directly compared count-
ing of protein particles obtained from LO, MFI, and Coulter instruments. One set of
results is shown in Fig. 9.9. These data demonstrate that the smallest amount of protein
subvisible particles detected is obtained with LO, followed by MFI and the Coulter
method. The results were dependent on the dilution of the sample prior to analysis
(Fig. 9.9) as well as the particle size cutoff used during the analysis (Demeule et al.
2010). The authors’ explanation for obtaining higher subvisible counts with the Coulter
method is its insensitivity to differences in the refractive index between the protein par-
ticles and the solvent. On the other hand, a study by Singh and colleagues showed that
the MFI and the Coulter methods generated comparable data except with highly concen-
trated protein samples, although the authors acknowledge that sample handling or dif-
ferences in the measurement methods can potentially affect the final results (Singh et al.
2010). Taken together, these studies again highlight the importance of using orthogonal
techniques to count and size protein particulates in the subvisible range to determine
which methods yield the most reproducible data for an individual protein formulation.
A detailed study by Wuchner and colleagues investigated the effect of long-term
storage of a high-concentration mAb formulation on formation of subvisible and visible
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 231
Fig. 9.9 Comparing the performance of the Coulter counter, flow microscopy (MFI), and light
obscuration (LO) for measuring subvisible particles (>3.5 µm) in protein solutions. Results are
reported as average ± standard deviation. Counting was performed at several protein concentrations
with the reported particle numbers corrected for the dilution factor and represent the particle con-
tent of the stock (150 mg/ml) solution. The particle counts of the formulation buffer are included.
Published in Demeule et al. (2010)
protein particles (Wuchner et al. 2010). Direct comparison of LO and MFI revealed five
to tenfold higher subvisible protein particle counts in the MFI measurements, consis-
tent with the results described above with other mAbs. This study proceeded to charac-
terize the count and size of subvisible and visible particles during long-term storage as
measured by MFI. The results found that a shift occurred in the particle size distribution
over time, with larger subvisible particles accumulating during long-term storage
(Fig. 9.10). This result was not observed by LO measurements. The accumulation of
visible particles was also monitored by a combination of visible assessments, inverted
microscopy, and MFI measurements (for protein particles larger than 100 µm; see
Fig. 9.10). The three techniques showed good overall correlation in terms of the count
of visible particle formed during long-term storage (Wuchner et al. 2010).
Fig. 9.10 Differential particle levels (subvisible- and visible-sized categories) in a 90 mg/ml IgG1
monoclonal antibody solution during storage for 18 months at 2–8°C as measured by MFI. Data
were collected from two to four preparations and presented as the average of differential particle
number per particle size range (ECD) with error bars of ±1 SD. Results from a frozen retain stored
for 12 months at −70°C were used as a surrogate for time 0. Published in Wuchner et al. (2010)
publications have shown the use of Nile red staining and fluorescence microscopy
for the detection and characterization of protein aggregates and particulates
(Demeule et al. 2007a, 2009; Sutter et al. 2007). Caution, however, needs to be
taken while using fluorescent dyes for protein characterization since these dyes may
also induce formation of aggregates and particulates because of their tendency to
perturb protein structure (Engelhard and Evans 1995).
234 T. Wang et al.
Fig. 9.11 Bright-field image (left) of unstained antibody A aggregates (in a 10 mM phosphate
buffer pH 7 solution containing 0.8 mg/ml protein) and a fluorescence image (right) of the same
antibody aggregates stained with Nile red. Both methods showed aggregates similar in size and
shape. Published in Demeule et al. (2007a, b)
Fig. 9.12 Representative ATR FTIR spectrum of a protein particle (b), overlaid with IgG1 refer-
ence spectrum (a), and polydimethylsiloxane reference spectrum (c). Protein stability sample con-
tained 90 mg/mL IgG1 stored for 16 months at 2–8°C. Published in Wuchner et al. (2010)
in the FTIR spectrum of protein particles at a 16-month stability time point were
comparable to those obtained for a reference IgG1 preparation (Fig. 9.12). Additional
bands characteristic of silicone were detected for most of the particles examined.
The protein particulates were further characterized by SEM-EDX to study their
composition (Wuchner et al. 2010) as shown in Table 9.2. In addition to silicon,
some of the protein particles contained aluminum and fluorine. These compositional
data from FTIR microscopy and SEM-EDX analysis indicated heterogeneity across
the particles examined, suggesting protein particle formation is a heterogeneous
process (Table 9.2).
This chapter has described recent case studies on the use of different analytical tech-
niques to better characterize the nature of protein aggregates and particles. The focus
was on recent work from different investigators who are exploring both new tech-
niques as well as combinations of analytical approaches, to better understand the
number, size range, and composition of protein aggregates and particles in protein
formulations. It is well recognized that no one analytical measurement can detect all
types and sizes of protein aggregates (Philo 2006). Some of the newer biophysical
approaches used to monitor protein aggregate and particle formation have come
from the field of protein aggregation diseases (Aguzzi and O’Connor 2010).
236 T. Wang et al.
Table 9.2 Analytical characterization of protein particles isolated from long-term stability
samples of a 90 mg/ml IgG1 monoclonal antibody solution formulation stored at 2–8°C
Age of
Sample material Inverted microscopy FTIR SEM-EDX
Batch 1 11 mo Some large and many small Protein (6/6), C,N,O (4/4), Si
translucent particles of typical silicone (6/6) (4/4)
proteinaceous morphology
Batch 2 14 mo Some large and many small Protein (5/5), C,N,O (3/3), Si
translucent particles of typical silicone (5/5) (3/3), Al
proteinaceous morphology (1/3), F (1/3)a
Batch 3 11 mo Some large and many small Protein (7/7), C,N,O (6/6), Si
translucent particles of typical silicone (4/7) (4/6), Al (4/6)
proteinaceous morphology
Batch 4 11 mo Some large and many small Protein (10/10), C,N,O (4/4), Si
translucent particles of typical silicone (3/4)
proteinaceous morphology (10/10)
Published in Wuchner et al. (2010)
mo, months at 2–8°C
Inverted microscopy: large particles through microscope approximately >50 µm, small particles
approximately <20 µm
FTIR microscopy: between 5 and 10 particles per sample were analyzed. The frequencies of main
components (spectral matches) are indicated in parentheses (number of particles with component
per total number or particles analyzed)
SEM-EDX: between 3 and 6 particles per sample were analyzed. The frequencies of the detected
elements are indicated in parentheses (number of particles with element per total number of par-
ticles analyzed)
a
Fluorine was detected at trace level
As new case studies are published demonstrating a further enhanced ability to detect
protein aggregates and particles under varying conditions, our mechanistic under-
standing of the interrelationships of protein aggregation pathways (Philo and
Arakawa 2009; Wang et al. 2010) as well as specific and nonspecific protein-protein
interactions leading to protein self-association (Saluja and Kalonia 2008) will con-
tinue to improve. The interrelationships between chemical and physical stabilities
are also emerging as an important new direction since some aggregates may contain
an altered nature (e.g., mixed disulfide formation and oxidation) as well as the pres-
ence of nonprotein nucleating species such as metals (Carpenter et al. 2009; Narhi
et al. 2011; Joubert et al. 2011; Luo et al. 2011). Similarly, although not the focus of
this chapter, our mechanistic understanding of the interrelationships of protein
aggregate composition and size to immunogenic potential is increasing (Carpenter
et al. 2010b; Mire-Sluis et al. 2011; Rosenberg 2006; Maas et al. 2007). Based on
this enhanced understanding, more rational approaches to the selection and optimi-
zation of stabilizing excipients to minimize protein aggregation and particle forma-
tion should result (Kamerzell et al. 2011). The combination of utilizing better
analytical methods along with an enhanced understanding of protein aggregation
pathways should ultimately lead to important improvements in the impurity levels
and long-term stability of protein formulations.
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 237
References
Harding SA, Din JN, Sarma J et al (2007) Flow cytometric analysis of circulating platelet-
monocyte aggregates in whole blood: methodological considerations. Thromb Haemost
98(2):451–456
Harper JD, Lieber CM, Lansbury PT Jr (1997) Atomic force microscopic imaging of seeded fibril
formation and fibril branching by the Alzheimer’s disease amyloid-beta protein. Chem Biol
4(12):951–959
Hawe A, Friess W, Sutter M, Jiskoot W (2008a) Online fluorescent dye detection method for the
characterization of immunoglobulin G aggregation by size exclusion chromatography and
asymmetrical flow field flow fractionation. Anal Biochem 378(2):115–122
Hawe A, Sutter M, Jiskoot W (2008b) Extrinsic fluorescent dyes as tools for protein characteriza-
tion. Pharm Res 25(7):1487–1499
Hawe A, Hulse WL, Jiskoot W, Forbes RT (2011) Taylor dispersion analysis compared to dynamic
light scattering for the size analysis of therapeutic peptides and proteins and their aggregates.
Pharm Res 28(9):2302–2310
He F, Phan DH, Hogan S et al (2010) Detection of IgG aggregation by a high throughput method
based on extrinsic fluorescence. J Pharm Sci 99(6):2598–2608
Heldt CL, Sorci M, Posada D, Hirsa A, Belfort G (2011) Detection and reduction of microaggre-
gates in insulin preparations. Biotechnol Bioeng 108(1):237–241
Hillger F, Nettels D, Dorsch S, Schuler B (2007) Detection and analysis of protein aggregation
with confocal single molecule fluorescence spectroscopy. J Fluoresc 17(6):759–765
Houde D, Arndt J, Domeier W, Berkowitz S, Engen JR (2009) Characterization of IgG1 conforma-
tion and conformational dynamics by hydrogen/deuterium exchange mass spectrometry. Anal
Chem 81(14):5966
Huang C-T, Sharma D, Oma P, Krishnamurthy R (2009) Quantitation of protein particles in paren-
teral solutions using micro-flow imaging. J Pharm Sci 98(9):3058–3071
Hughes H, Morgan C, Brunyak E et al (2009) A multi-tiered analytical approach for the analysis
and quantitation of high-molecular-weight aggregates in a recombinant therapeutic glycopro-
tein. AAPS J 11(2):335–341
Hunt G, Moorhouse KG, Chen AB (1996) Capillary isoelectric focusing and sodium dodecyl
sulfate-capillary gel electrophoresis of recombinant humanized monoclonal antibody HER2.
J Chromatogr A 744(1–2):295–301
Jorio H, Tran R, Meghrous J, Bourget L, Kamen A (2006) Analysis of baculovirus aggregates
using flow cytometry. J Virol Methods 134(1–2):8–14
Joseph Goldstein DN, Joy D, Lyman C, Echlin P, Lifshin E, Sawyer L, Michael J (2003) Scanning
electron microscopy and X-ray microanalysis, 3rd edn. Plenum, New York
Joubert MK, Luo Q, Nashed-Samuel Y, Wypych J, Narhi LO (2011) Classification and character-
ization of therapeutic antibody aggregates. J Biol Chem 286(28):25118–25133
Kaddis CS, Lomeli SH, Yin S et al (2007) Sizing large proteins and protein complexes by electrospray
ionization mass spectrometry and ion mobility. J Am Soc Mass Spectrom 18(7):1206–1216
Kaltashov IA, Bobst CE, Abzalimov RR, Wang G, Baykal B, Wang S (2012) Advances and chal-
lenges in analytical characterization of biotechnology products: mass spectrometry-based
approaches to study properties and behavior of protein therapeutics. Biotechnol Adv
30(1):210–222
Kamerzell TJ, Esfandiary R, Joshi SB, Middaugh CR, Volkin DB (2011) Protein-excipient interac-
tions: mechanisms and biophysical characterization applied to protein formulation develop-
ment. Adv Drug Deliv Rev 63(13):1118–1159
Kheterpal I, Wetzel R (2006) Hydrogen/deuterium exchange mass spectrometry—a window into
amyloid structure. Acc Chem Res 39(9):584–593
Khurana R, Uversky VN, Nielsen L, Fink AL (2001) Is Congo red an amyloid-specific dye? J Biol
Chem 276(25):22715–22721
Kiese S, Papppenberger A, Friess W, Mahler HC (2008) Shaken, not stirred: mechanical stress
testing of an IgG1 antibody. J Pharm Sci 97(10):4347–4366
Kiese S, Pappenberger A, Friess W, Mahler H-C (2010) Equilibrium studies of protein aggregates
and homogeneous nucleation in protein formulation. J Pharm Sci 99(2):632–644
240 T. Wang et al.
Madsen R, Cherris R, Shabushnig JG, Hunt DG (2009) Visible particulates in injections—a history
and a proposal to revise USP general chapter injections (1). Pharmacop Forum 35:1383–1387
Mahler HC, Muller R, Friess W, Delille A, Matheus S (2005) Induction and analysis of aggregates
in a liquid IgG1-antibody formulation. Eur J Pharm Biopharm 59(3):407–417
Mahler H-C, Friess W, Grauschopf U, Kiese S (2009) Protein aggregation: pathways, induction
factors and analysis. J Pharm Sci 98(9):2909–2934
Majumdar S, Ford BM, Mar KD, Sullivan VJ, Ulrich RG, D’souza AJM (2011) Evaluation of the
effect of syringe surfaces on protein formulations. J Pharm Sci 100(7):2563–2573
Marcsisin SR, Engen JR (2010) Hydrogen exchange mass spectrometry: what is it and what can it
tell us? Anal Bioanal Chem 397(3):967–972
Markovich RJ, Taylor AK, Rosen J (1997) Drug migration from the adhesive matrix to the polymer
film laminate facestock in a transdermal nitroglycerin system. J Pharm Biomed Anal
16(4):651–660
Michel Godin AKB, Burg TP, Babcock K, Manalis SR (2007) Measuring the mass, density, and
size of particles and cells using a suspended microchannel resonator. Appl Phys Lett 91:3
Michels DA, Brady LJ, Guo A, Balland A (2007) Fluorescent derivatization method of proteins for
characterization by capillary electrophoresis-sodium dodecyl sulfate with laser-induced fluo-
rescence detection. Anal Chem 79(15):5963–5971
Middaugh CR, Joshi SB (2011) Spectroscopy of vaccines. In: Singh M, Srivastava IK (eds)
Development of vaccines, from discovery to clinical testing. Wiley, Hoboken, pp 263–292
Mire-Sluis A, Cherney B, Madsen R, Polozova A, Rosenberg A, Smith H, Arora T, Narhi L (2011)
Analysis and immunogenic potential of aggregates and particles. BioProcess Int 9(10):38–43
Murphy RM (1997) Static and dynamic light scattering of biological macromolecules: what can
we learn? Curr Opin Biotechnol 8(1):25–30
Naiki H, Higuchi K, Hosokawa M, Takeda T (1989) Fluorometric determination of amyloid fibrils
in vitro using the fluorescent dye, thioflavin T1. Anal Biochem 177(2):244–249
Narhi LO, Schmit J, Bechtold-Peters K, Sharma D (2011) Classification of protein aggregates.
J Pharm Sci 101(2):493–498
Nayak A, Colandene J, Bradford V, Perkins M (2011) Characterization of subvisible particle for-
mation during the filling pump operation of a monoclonal antibody solution. J Pharm Sci
100(10):4198–4204
Nettleton EJ, Tito P, Sunde M, Bouchard M, Dobson CM, Robinson CV (2000) Characterization
of the oligomeric states of insulin in self-assembly and amyloid fibril formation by mass spec-
trometry. Biophys J 79(2):1053–1065
O’Sullivan CK, Guilbault GG (1999) Commercial quartz crystal microbalances – theory and appli-
cations. Biosens Bioelectron 14(8–9):663–670
Oliva A, Llabres M, Farina JB (2001) Comparative study of protein molecular weights by size-
exclusion chromatography and laser-light scattering. J Pharm Biomed Anal 25:833–841
Pease LF 3rd, Elliott JT, Tsai DH, Zachariah MR, Tarlov MJ (2008) Determination of protein
aggregation with differential mobility analysis: application to IgG antibody. Biotechnol Bioeng
101(6):1214–1222
Philo JS (2006) Is any measurement method optimal for all aggregate sizes and types? AAPS
J 8(3):E564–E571
Philo JS (2009) A critical review of methods for size characterization of non-particulate protein
aggregates. Curr Pharm Biotechnol 10(4):359–372
Philo JS, Arakawa T (2009) Mechanisms of protein aggregation. Curr Pharm Biotechnol
10(4):348–351
Prati S, Joseph E, Sciutto G, Mazzeo R (2010) New advances in the application of FTIR micros-
copy and spectroscopy for the characterization of artistic materials. Acc Chem Res
43(6):792–801
Qi P, Volkin DB, Zhao H et al (2009) Characterization of the photodegradation of a human IgG1
monoclonal antibody formulated as a high-concentration liquid dosage form. J Pharm Sci
98(9):3117–3130
242 T. Wang et al.
Remmele RL Jr, Callahan WJ, Krishnan S et al (2006) Active dimer of Epratuzumab provides
insight into the complex nature of an antibody aggregate. J Pharm Sci 95(1):126–145
Reschiglian P, Zattoni A, Roda B, Michelini E, Roda A (2005) Field-flow fractionation and bio-
technology. Trends Biotechnol 23(9):475–483
Rhyner MN (2011) The Coulter principle for analysis of subvisible particles in protein formula-
tions. AAPS J 13(1):54–58
Roberts GS, Yu S, Zeng Q et al (2012) Tunable pores for measuring concentrations of synthetic
and biological nanoparticle dispersions. Biosens Bioelectron 31(1):17–25
Roda B, Zattoni A, Reschiglian P et al (2009) Field-flow fractionation in bioanalysis: a review of
recent trends. Anal Chim Acta 635(2):132–143
Rosenberg AS (2006) Effects of protein aggregates: an immunologic perspective. AAPS
J 8(3):E501–E507
Rosenberg E, Hepbildikler S, Kuhne W, Winter G (2009) Ultrafiltration concentration of monoclo-
nal antibody solutions: development of an optimized method minimizing aggregation. J Membr
Sci 342(1–2):50–59
Ross S, Morrison ID (2002) Colloidal dispersions. Wiley Interscience, New York
Rumi Chunara MG, Knudsen SM, Manalis SR (2007) Mass-based readout for agglutination
assays. Appl Phys Lett 91:3
Rustandi RR, Washabaugh MW, Wang Y (2008) Applications of CE SDS gel in development of
biopharmaceutical antibody-based products. Electrophoresis 29(17):3612–3620
Salas-Solano O, Tomlinson B, Du S, Parker M, Strahan A, Ma S (2006) Optimization and valida-
tion of a quantitative capillary electrophoresis sodium dodecyl sulfate method for quality con-
trol and stability monitoring of monoclonal antibodies. Anal Chem 78(18):6583–6594
Salman A, Erukhimovitch V, Talyshinsky M, Huleihil M, Huleihel M (2002) FTIR spectroscopic
method for detection of cells infected with herpes viruses. Biopolymers 67(6):406–412
Saluja A, Kalonia DS (2008) Nature and consequences of protein-protein interactions in high
protein concentration solutions. Int J Pharm 358(1–2):1–15
Sauerbrey G (1959) Use of quartz vibration for weighing thin films on a microbalance. J Phys
155:206–212
Schmitt C, Bovay C, Rouvet M, Shojaei-Rami S, Kolodziejczyk E (2007) Whey protein soluble
aggregates from heating with NaCl: physicochemical, interfacial, and foaming properties.
Langmuir 23(8):4155–4166
Schuck P (2000) Size-distribution analysis of macromolecules by sedimentation velocity
ultracentrifugation and lamm equation modeling. Biophys J 78(3):1606–1619
Schuck P (2004) A model for sedimentation in inhomogeneous media. II. Compressibility of aque-
ous and organic solvents. Biophys Chem 108(1–3):201–214
Sejersen M, Salomonsena T, Ipsen R, Clark R, Rolin C, Engelsen S (2007) Zeta potential of pectin-
stabilised casein aggregates in acidified milk drinks. Int Dairy J 17:302–307
Shapiro HM (1995) Practical flow cytometry. Wiley-Liss, New york
Sharma DK, Oma P, Pollo MJ, Sukumar M (2010a) Quantification and characterization of subvis-
ible proteinaceous particles in opalescent mAb formulations using micro-flow imaging.
J Pharm Sci 99(6):2628–2642
Sharma DK, King D, Oma P, Merchant C (2010b) Micro-flow imaging: flow microscopy applied
to sub-visible particulate analysis in protein formulations. AAPS J 12(3):455–464
Singh SK, Afonina N, Awwad M et al (2010) An industry perspective on the monitoring of subvis-
ible particles as a quality attribute for protein therapeutics. J Pharm Sci 99(8):3302–3321
Strehl R, Rombach-Riegraf V, Diez M et al (2011) Discrimination between silicone oil droplets
and protein aggregates in biopharmaceuticals: a novel multiparametric image filter for sub-
visible particles in microflow imaging analysis. Pharm Res 27:1–9
Sutter M, Oliveira S, Sanders NN et al (2007) Sensitive spectroscopic detection of large and dena-
tured protein aggregates in solution by use of the fluorescent dye Nile red. J Fluoresc
17(2):181–192
9 Case Studies Applying Biophysical Techniques to Better Characterize Protein… 243
10.1 Introduction
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 245
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2_10,
© Springer Science+Business Media New York 2013
246 Z.-Q. Wen et al.
and the container per se. They include both visible (>100 μm) protein aggregates
growing from single protein molecule either through self-association or partial
denaturation and then aggregation (Narhi 2012) and/or particles generated from the
materials of the primary container such as glass (Iacocca and Allgeier 2007), and
silicone oil (Wen et al. 2009), which is used as a lubricant in prefilled syringes
(PFSs). Extraneous particulate matter or foreign contaminants are material residues
carried over into the primary container that were generated during the manufactur-
ing process of the primary container by the vendor, as well as unexpected foreign
particles from manufacturing environment occasionally fall into the container dur-
ing filling process. For instance, tungsten oxide (Wen et al. 2007) and polymer resi-
due left inside of PFSs (Liu et al. 2010) were from a tungsten pin and nylon pin used
for PFS manufacturing. The identification of these microparticles plays a critical
role in determining the root cause of the nonconformance, including the exact
original source of the particle and the mechanism of particulate generation. This in
turn enables process engineers and formulation scientists to take preventive and cor-
rective action and to help quality and toxicological scientists to assess the product
and safety impact, ultimately determining the disposition or rejection of that par-
ticular lot.
There are many analytical techniques that are employed to determine the physi-
cal size and chemical identities of the microparticles. The most frequently employed
are optical microscopy to determine the morphology and physical size, vibrational
microspectroscopies including Fourier transform infrared (FTIR) and Raman for
the determination of the chemical identity (Humecki 1995; Li et al. 2009; Wen
2007; Wen et al. 2008), and scanning electron microscope (SEM) with energy dis-
persive spectroscopy (EDS) for elemental composition (Goldstein et al. 1992). In
addition, FTIR and Raman microscopy can provide higher-order structural informa-
tion of protein the microparticle may contain (Wen 2007; Wen et al. 2008). Moreover,
information on whether the protein particles are denatured and/or aggregated can be
assessed by the derivative vibrational spectra as well. Using SEM/EDS analysis,
one can further provide a mapping of the distribution of individual elements
throughout the microparticle to determine if the specific element is evenly distrib-
uted in the microparticle or merely heterogeneous among other components in the
particle (Liu et al. 2010).
In this chapter, we will give a brief overview of microparticle analysis and thera-
peutic protein manufacturing NC investigation as currently performed in our labora-
tory and throughout the industry. The general approach to microparticle analysis
and NC incident investigation will be outlined, and the principle of the analytical
procedures will be presented. The goal of this chapter is to give insight into mic-
roparticle NC investigation, the challenges it presents, and how it fits into the overall
application of biophysical and analytical methods to biopharmaceutical develop-
ment and manufacturing. All of the cases presented in the chapter are from previous
investigations that involve protein therapeutics contained in primary containers
including prefilled glass syringes and glass vials. These incidents cover a broad
range of materials that could be encountered during the fill and finishing processes
during manufacturing of protein therapeutics.
10 Investigation of Nonconformance During Protein Therapeutic Manufacturing 247
Figure 10.1 shows the flow chart of the scheme of microparticle analysis. The
majority of the microparticle samples are obtained from manufacturing sites and
quality control labs that conducted visual inspection of products contained in pre-
filled glass syringe or glass vials. The primary containers that have microparticles
are normally sent to the forensic lab for analysis at a controlled temperature of
4–8°C. At the beginning of each investigation, a picture of this primary container is
taken with a digital camera or a stereomicroscope to record the particles in the
original state. It is recommended to record images of the entire container as well as
the featured area that has the particles to make sure that their integrity was not
compromised during transportation before the container is opened. In particular,
for glass vials, a picture of the vial cap should be recorded before opening it. This
is important to avoid potential controversy regarding if the vial had been punctured
or not. This is particularly important for vials containing commercial drug product
and returned as part of a customer complaint from patients or clinical centers. If the
vial has been punctured by a needle, the source of the microparticles may not be
caused by the manufacturing process.
The majority of the microparticles can be isolated through filtration using filter
appropriate for the planned analysis. A gold-coated membrane filter (Rapid-ID,
Berlin, Germany) facilitates FTIR-microscopic analysis of the particle directly on
the filter without the need to pick up the particle and transfer it to a potassium bro-
mide (KBr) disc. A polycarbonate membrane filter (Millipore Inc. Billerica) is bet-
ter for SEM/EDS analysis, as it has no strong gold signal to interfere with other
element analysis on the EDS spectrum. If the microparticle is larger than 100 µm, it
may be retrieved directly with an appropriate probe (tungsten probe or plastic spat-
ula) under a stereomicroscope in a laminar flow hood in a clean room (Class 100).
The microparticles that are retained on the filters are then examined with a stereo-
microscope to record their size and morphology, followed by FTIR-microscopic
analysis. If they can be identified definitively by FTIR at this step (especially if they
are organic in nature), further SEM/EDS analysis may not be required. However, for
metallic and inorganic materials, SEM/EDS analysis is a must as these materials
may not exhibit appropriately detailed FTIR or Raman spectra. Even for organic or
polymeric materials, SEM/EDS analysis can provide additional information regard-
ing the elemental composition of the materials to substantiate the vibrational spec-
troscopic analysis. In some cases the microparticles inside the glass container may
be identified using in situ Raman microscopy (Cao et al. 2009, 2010). If the samples
are amenable to in situ Raman analysis, this will be the ideal situation as Raman
microscopy can provide identification without the need to open the container to
isolate the microparticle and is a nondestructive technique.
For complex microparticles that are composed of multiple components, the above-
mentioned techniques may not be sufficient to provide a definitive determination,
and additional chromatographic and spectroscopic techniques such as inductively
coupled plasma mass spectrometry (ICP-MS), gas chromatography (GC), high-
performance liquid chromatography (HPLC/MS), and nuclear magnetic resonance
(NMR) spectroscopy may be required. This may eventually lead to a full-scale com-
prehensive trace material analysis, which is beyond the scope of this chapter.
and Raman analysis. Three major optical microscopes are commonly employed as
the first step for microparticle analysis. A Carl Zeiss Stemi 2000C stereomicroscope
can be employed to take pictures of microparticles that have been isolated onto fil-
ters. An Axioskop 2 MAT polarized microscope (Carl Zeiss Inc, Peabody) is
employed to examine microparticles that may have birefringence such as liquid crys-
tal polymer fiber. The polarized microscope can help to identify some materials that
have a distinct birefringence such as cotton cellulose fibers, based on their optical
properties. For in situ examination of microparticles contained inside the prefilled
glass syringe and large-size glass vials, a long working distance and high-resolution
optical microscope is required. These requirements are fulfilled with a Keyence
high-magnification optical microscope VHX-600 or similar instrument (Keyence,
Itasca), which can take high-magnification-quality images of very tiny particles in
situ. The three microscopes are selectively employed according to the specific needs
of investigations and the type of containers and samples being analyzed.
FTIR and Raman microscopy are two vibrational microspectroscopic techniques that
provide complementary information on the molecular structure of materials (Humecki
1995; Li et al. 2009; Wen 2007; Wen et al. 2008). They are the result of a marriage
between vibrational spectroscopy and optical microscopy. The combination of a
microscope and a vibrational spectrometer offers the power to focus easily on a spe-
cific microparticle using the microscopic unit and then to record the vibrational spec-
trum of this selected particle with the spectrometer. They are the primary techniques
for determining the identities of microparticles composed of polymeric or organic
compounds at the molecular level. Some inorganic materials can also be examined by
FTIR and Raman microscopy if they exhibit strong vibrational spectra.
The principles of FTIR and Raman spectroscopy are detailed in Chap. 3. FTIR-
microscopy is generally preferred as the analytical tool for microparticle analysis for
particles that have been isolated on a filter as the FTIR-microscope has been devel-
oped into a very sophisticated tool that is relatively easy to operate. Moreover, a large
FTIR database of reference spectra with more than a quarter of a million compounds
is now available from Bio-Rad Know-it-all (Shermann and Brodbelt 2002) that can
be used to aid in particle identification. For microparticles composed of a single
component, the FTIR spectrum of the microparticle is compared with the reference
FTIR spectral library; if a perfect match is found to a reference spectrum from the
library, the identity of the microparticle can be established. In cases where there is no
perfect match, the chemical and structural information from a set of best matching
reference spectra could still provide information on the category or structural family
of the material, and to narrow down the possibilities and the direction the investiga-
tion should take. Further analysis relies on the experienced vibrational spectroscopist
to interpret the FTIR spectra based on the structural chemistry using the empirical
group frequency analysis approach, to determine the most likely composition based
250 Z.-Q. Wen et al.
on the combination of functional groups and any knowledge relevant to the particles
(Li et al. 2009; Smith 1999). In many cases, an identification of the type or category
of a material can be very helpful in finding the root cause of the microparticles. This
type of information can also be used to assess the potential chemical toxicity of the
particles and for making decisions on product quality impact.
Raman microscopy is employed as a complementary technique to FTIR-
microscopy. However, it has the advantage over FTIR-microscopy in that it can
perform in situ analysis without the particle isolation procedure (Goldstein et al.
1992; Cao et al. 2009) for microparticles contained in glass vials and prefilled glass
syringes. This is because the visible laser beam used in Raman can penetrate through
the container without any interference from the glass or the aqueous solution. The
glass (or plastic) containers strongly absorb infrared radiation and thus prevent
FTIR measurements in situ. Another unique feature of Raman microscopy is that
molecules that have rich electron density, such as conjugated aromatic compounds,
show strong Raman bands (Chaps. 3, 13, and 16). Inorganic crystal material can
exhibit strong Raman scattering in the low-frequency region as well due to the
vibrational modes of the crystal lattice. The weakness of Raman microscopy is rela-
tively low instrument sensitivity compared to FTIR-microscopy, resulting in longer
acquisition time of a spectrum. Samples with a strong fluorescent background can
also interfere with the ability to obtain a Raman spectrum. Currently the reference
database for Raman spectra is also much smaller (15,000 Raman spectra) than the
FTIR spectral database (a quarter of a million). Fortunately, the Raman database of
organic and polymeric materials has been expanding rapidly in recent years with the
continued improvement of the instrumentation, the increased ease of use, and the
accumulation of Raman spectra of various materials in labs around the world.
Raman microscopy will play an even greater role for microparticle analysis in the
pharmaceutical industry in the near future.
SEM coupled with the energy dispersive X-ray spectroscopy (EDS) is a chemical
microanalysis technique that utilizes X-rays emitted from the sample upon
bombardment with an electron beam to characterize the elemental composition of
the materials (Goldstein et al. 1992). The SEM function provides a high-resolution
micrograph of the microparticle and the EDS provides a spectrum of the elementary
composition. When the microparticle is irradiated by the electron beam, electrons
are ejected from the atoms of the sample. The resulting electron vacancy is filled by
electrons from higher orbit shells, and the characteristic X-rays of the element are
emitted. The EDS spectrum consists of the characteristic X-rays of the elements,
which is used for qualitative determinations of the elements in the sample (Goldstein
et al. 1992). SEM/EDS is particularly useful for identification of microparticles
from metallic and inorganic materials that do not produce molecular vibrational
10 Investigation of Nonconformance During Protein Therapeutic Manufacturing 251
Fig. 10.2 The micrograph of microparticles (top panel) and the FTIR spectrum (up trace) of a
microparticle in a pre-filled syringe and the reference FTIR spectrum of a protein (lower trace)
Fig. 10.3 The SEM micrographs and the EDS spectra of a microparticle and a tungsten chip found
in a pre-filled syringe
10 Investigation of Nonconformance During Protein Therapeutic Manufacturing 253
high temperatures in the range of 1000°C. The tungsten pin was used to create the
hole of the glass barrel where the needle would later be inserted as the glass was
heated at temperatures approaching the glass melting point. Oxidization of the tung-
sten pin surface can generate tungsten oxide, since the glass syringe barrel was
made in an open air environment at this high temperature. These tungsten-rich com-
pounds can deposit (most likely via sublimation) on the interior of the glass syringe,
primarily near the tip of the barrel (Liu et al. 2010). During the manufacturing pro-
cess of the PFSs, the barrels were rinsed with jets of 80°C water for injection prior
to final packaging and sterilization; however, the tungsten deposits may not be suf-
ficiently soluble to be completely removed by the washing process (Liu et al. 2010).
The possibility of the presence of tungsten residues in the syringe was confirmed by
the analysis of white deposits seen in the funnel area of empty syringe barrels. The
root cause of tungsten element in the microparticles was thus attributed to the tung-
sten pins used at high temperature during the manufacturing process of the glass
barrel (Wen et al. 2007). The residual tungsten species remained in the barrel and
were carried over to the product during the fill and finishing process, at which point
they came in contact with the protein therapeutic. This resulted in particle formation
on stability which was then found by visual inspection.
To understand the role of tungsten compounds in protein particulation, subse-
quent experiments spiking tungsten oxide compounds and used tungsten pin extracts
into a protein solution were conducted (Jiang et al. 2009). These studies demon-
strated that some proteins under specific solution conditions indeed precipitated into
microparticles with the addition of tungsten compounds at a level above 1 ppm of
used tungsten pin extractables. The induced microparticles contained both proteins
and tungsten as confirmed by SEM/EDS (Jiang et al. 2009). Moreover, the tungsten
was also distributed evenly throughout these microparticles as revealed by the EDS
Cameo mapping of tungsten (see Fig. 10.4) (Liu et al. 2010). Biophysical character-
ization of the protein–tungsten aggregates in solution using Raman, FTIR, and DLS
confirmed that the protein particles induced by tungsten in this case were reversible.
They were associated with the formation of tungsten oxide colloids, which require
specific solution conditions and a stoichiometry ratio (1:1) between protein and
tungsten (Jiang et al. 2009). To prevent microparticle formation in protein drug
products induced by tungsten residues, a specification of below 0.5 ppm tungsten
was established for the lots of PFSs. The manufacturer of the PFSs also changed
their manufacturing processes to minimize the tungsten residues in PFSs.
In this case, visible particles were initially observed in a few frozen placebo vials of
the protein therapeutic. The particles in the placebo vials were visible to the naked
eye, appeared translucent to white, and amorphous. However, no similar particles
were seen in the protein therapeutic vials which had the exact same formulation as
the placebo. The particles were first isolated on a membrane filter and examined
with optical microscopy followed by FTIR-microscopic analysis.
254 Z.-Q. Wen et al.
Fig. 10.4 The cameo mapping of Tungsten element in a microparticle and its EDS spectrum,
which confirms that the tungsten element is distributed evenly on the entire microparticle, which is
a mixture of protein and tungsten element
Figure 10.5 shows the results of the optical microscopic and FTIR-microscopic
analyses. The top panel of Fig. 10.5 shows the micrograph of the microparticles in
the size range of 20–250 µm. The bottom panel of Fig. 10.5 is the FTIR spectra of
the particles together with a reference FTIR spectrum of polysorbate-20 and a refer-
ence FTIR spectrum of silica gel from the Bio-Rad FTIR spectral library (Shermann
and Brodbelt 2002). The FTIR spectrum of the particle showed bands at 1,736 cm−1
(C═O stretching), 2,926 and 2,853 cm−1 (CH2 stretching), 1,460 cm−1 (CH2 bend-
ing), and 1,070 cm−1 (CO stretching), respectively. These FTIR features of the par-
ticles are very similar to the FTIR reference spectrum of polysorbate-20 except that
the strongest band is centered at 1,070 cm−1 in the particle spectrum, while polysor-
bate-20 shows two strong bands at 1,100 and 2,800–2,950 cm−1 region. Since poly-
sorbate-20 was part of the placebo formulation, it was initially suspected to be the
major component and the source of the particles. However, careful examination of
the FTIR spectrum of the particle found that there are a few bands of medium
intensity at 958 and 802 cm−1 that could not be attributed to polysorbate-20.
Moreover, the CH2 stretching bands at 2,925 and 2,853 cm−1 had much weaker signals
than those of polysorbate-20, which has many CH2 groups from its polyoxyethylene
10 Investigation of Nonconformance During Protein Therapeutic Manufacturing 255
Fig. 10.5 The optical micrograph of the microparticles (top) and FTIR spectra of a microparticle,
the reference FTIR spectra of Polysorbate-20 and of a silica gel
chain and the long fatty acid ester moiety. This suggested that the particulates may
have contained polysorbate-20 but may have other components which contribute to
the FTIR bands that cannot be assigned to surfactant (Fig. 10.6).
Subsequent SEM/EDS analysis of the particles revealed an extra element Si in
addition to the two major expected organic elements of C and O. This indicated
that the particles included silicon-containing material. Trace aluminum was also
found in some of the particles (data not shown here) by SEM/EDS. Moreover, a
Cameo mapping of the particle with Si element showed that the silicon was distrib-
uted evenly throughout the particle (see Fig. 10.5). The vial was made of borosili-
cate glass, and silicon can be dissolved from the glass vial in aqueous solution
(Perera and Doremus 1991). Quantitative comparative analysis of the silicon and
other glass elements (boron, aluminum, sodium) by ICP-MS revealed that the pla-
cebo glass vial containing visible particles had 30 times silicon in solution than a
control placebo plastic vial (below the detection limit of 150 ppb) (Ricci et al. 2011).
These analyses strongly suggested that the visible particles were composed of silica
256 Z.-Q. Wen et al.
Fig. 10.6 Secondary electron image of a representative particle on gold coated filter (Left) the sili-
con element map displays Si in blue (Right). Note that Si is evenly distributed throughout the
particle and it matches the particle shape well shown on the left
and polysorbate-20 and that the glass container itself was the source of the silica
particles. The root cause analysis then focused on why the placebo vials had the
visible particles, but the product vials did not, and what caused the placebo vial to
generate visible particulates that contain silicon.
Reviewing the manufacturing history of the protein product and placebo vials
revealed some subtle differences in the post-fill handling and storage temperatures
between the product and placebo vials. To maximize the stability of drug product, it
was frozen at −30°C immediately after fill. In contrast, the placebo was considered
to be robust and therefore their storage temperature was not deemed critical. For
practical and logistical considerations, the placebo vials were stored at 2–8°C after
fill until they were needed for clinical labeling and packaging. Since the demand
varied over the course of the clinical studies, the time of storage at 2–8°C varied
across the placebo lots by months. On the other hand, the product vials were frozen
and stored at −30°C both after fill and packaging and at the clinical trial sites.
Moreover, it was realized that the product and placebo were formulated at neutral
pH 7, a pH that can accelerate glass dissolution (Perera and Doremus 1991). Reports
in the literature indicated that silica, the dominant component of the glass vials, is
less stable in neutral pH than under relatively acidic conditions (Perera and Doremus
1991). The combination of neutral pH and storage of the placebo vials at 2–8°C for
a period apparently increased the dissolution of glass into solution in the form of
silicic acids, which then polymerized into silica gel (Hsu-Chiang et al. 2005). The
polysorbate-20 molecules were merely associated to the silica gel particulates, but
not the cause. Scanning of the inner surface of the placebo glass vials by SEM
revealed that there were defects in the glass surface of the liquid contact area, but
not on the nonliquid contact area above the fill line (data not shown here). This fur-
ther reinforces the hypothesis that the silica gel particles were indeed formed from
the glass vial through glass dissolution. Since the protein drug product was frozen
at −30°C immediately after fill and was stored at −30°C, the silica dissolution was
greatly minimized at lower temperature for the product container (Ricci et al. 2011).
Therefore, the storage condition of the placebo was changed to match that of the
protein therapeutic product, reducing the storage time of the placebo at 2–8°C to
less than 48 h.
10 Investigation of Nonconformance During Protein Therapeutic Manufacturing 257
Fig. 10.7 Optical micrograph, FTIR spectra of the bubble-like particle (top trace FTIR is the
bubble like microparticle and the bottom trace FTIR is the reference spectrum of silicone oil
Fig. 10.8 The two SEM/EDS spectra are obtained from the bubble like particles. The top panel of
the EDS spectrum was obtained on the protein component of the bubble-like particle. The bottom
panel of the EDS spectrum was obtained on the bubble-like micro-particle that containing silicone
oil and protein
the Si–C stretching bands at 865 and 800 cm−1. The FTIR spectra were consistent
with a protein–silicone oil mixture.
The SEM/EDS analyses, shown in Fig. 10.8, supported the FTIR results as the
EDS spectra of the two particles differed only in the presence of Si, which is attrib-
uted to the presence of silicone oil in the particle shown in Fig. 10.7b. Other than Si,
both particles were found to contain C, N, O, and S. This is consistent with the FTIR
results as a proteinaceous material containing the basic elements of sulfur, carbon,
oxygen, and nitrogen, although some C and O signals could also be contributed by
the underlying polycarbonate membrane filter. The EDS spectra also showed P, Na,
and Cl, which can be attributed to the formulation buffer. The SEM micrographs
demonstrated morphological similarities of both “dried bubble” particles. The
source of the silicone oil was the PFS, as silicone oil is coated in a very thin layer of
a few hundred nanometers on the glass barrel surface as a lubricant (Wen et al.
2009). The generation of these bubble-like particles was attributed to the silicone
oil/protein/buffer interaction during the transportation (vibration) of the products
via dewetting processes as silicone oil is hydrophobic, but is in contact with aqueous
solution. Tiny silicone oil droplets on the glass surface can migrate together to form
relatively large visible bubble-like particles associated with protein molecules.
10.4 Conclusions
NC investigations and root cause analyses are critical to protein therapeutic devel-
opment and commercialization. Physical particulate contaminants are one of the
types of NC that can occur during the fill and finishing process of biopharmaceutical
10 Investigation of Nonconformance During Protein Therapeutic Manufacturing 259
products. When this occurs, the physical–chemical nature of the particulate matter
must be identified as one of the first steps in determining the origin and the root
cause to allow remediation. This requires the use of the appropriate tools from the
biophysical toolbox as described above.
It is worthwhile to point out that many physical–chemical contaminants found in
the final biopharmaceutical products were originally generated during the process-
ing of the drug product container and were associated with the container manufac-
turing equipment, being carried over into the final commercial drug product. Other
times, it is the glass container per se that interacted with the formulation compo-
nents of the protein products to generate particulate issues such as glass delamina-
tion. To quickly resolve NC investigations, detailed analysis of materials present,
and determination of their root cause and source are mandatory. Moreover, close
collaboration among the raw material vendor, primary container manufacturer, and
the protein pharmaceutical manufacturer is required for resolution of the NCs and
to take appropriate preventive measures.
References
11.1 Introduction
Proteins are complex molecules that affect gene expression, cellular structure,
metabolism, reproduction, inter-/intracellular signaling, immune responses, and virtu-
ally every aspect of life. Function of proteins depends on the orientation of the func-
tional groups of their building blocks. For example, the presence and orientation of
the three critical amino acids (Ser, His, and Asp) in the catalytic triad of serine prote-
ases, such as coagulation factors, determines their ability to perform their biologic
role, i.e., cleave a peptide bond and, in the case of the coagulation cascade, activate
the substrate. Furthermore, nearby regions (secondary and tertiary structures) are
Disclaimer The findings and conclusions in this chapter have not been formally disseminated by
the Food and Drug Administration and should not be construed to represent any agency determina-
tion or policy.
E.B. Struble • C. Kimchi-Sarfaty • Z.E. Sauna • E. Marszal (*)
Division of Hematology, Office of Blood Research and Review,
Center for Biologics Evaluation and Research (CBER), Food and Drug
Administration (FDA), 29 Lincoln Drive, Bethesda, MD 20892, USA
e-mail: [email protected]; [email protected];
[email protected]; [email protected]
J.F. Cipollo
Division of Bacterial, Parasitic and Allergenic Products,
Office of Vaccine Research and Review, CBER, FDA, 29 Lincoln Drive,
Bethesda, MD 20892, USA
e-mail: [email protected]
J.A. Ragheb
Division of Therapeutic Proteins, Office of Biotechnology Products, Center for Drug Evaluation
and Research, FDA, 29B Lincoln Drive, Bethesda, MD 20892, USA
e-mail: [email protected]
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 261
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2_11,
© Springer Science+Business Media New York 2013
262 E.B. Struble et al.
essential for substrate specificity and, in the case of coagulation factors, enhancement
of the biologic activity by cofactors and attenuation of the effect by inhibitors.
Unlike the bonds that determine the structures of small molecules, the interactions
driving the complex structure of proteins—hydrogen bonds, electrostatic interac-
tions, and van der Waals forces—are noncovalent in nature. As such, the energy that
maintains the stability of a protein’s three-dimensional structure is inherently low,
in many cases not much greater than the thermal energy at room temperature (Pfeil
and Privalov 1976). Thus, increases in temperature result in protein unfolding
(Griko et al. 1988). Other environmental conditions such as pH (Anderson et al.
1990), solvents (Grothe et al. 2009), salts, small molecule additives, impurities
(Bolen and Baskakov 2001; Maity et al. 2009; Rosgen 2007; England and Haran
2011), and even other proteins (Hartl and Hayer-Hartl 2009) can influence the three-
dimensional structure and result in complete or partial unfolding of the protein.
Destabilized proteins often form aggregates, which may vary from dimers with a
size in the low nanometer range to visible agglomerates of thousands of protein
molecules reaching a size of several millimeters. Aggregated proteins lose physio-
logic activity (Meager et al. 2011; Carpenter et al. 2010a) and may acquire new,
toxic attributes. Examples include diseases associated with protein aggregation such
as systemic amyloidoses (Obici et al. 2005), neurological disorders (Aguzzi and
O’Connor 2010), and, as recently shown, cancer (Xu et al. 2011). Interestingly, it has
been postulated that in some diseases toxicity is associated with oligomeric interme-
diates rather than the large aggregates (Haass and Selkoe 2007; Klein et al. 2001).
The formation of protein aggregates is a major concern in the manufacture and
clinical use of protein therapeutics. Protein aggregates have been associated with
enhanced immune responses both in animal studies and clinical trials (Dresser
1962; Braun et al. 1997; Moore and Leppert 1980). Aggregates may stimulate the
innate immune system to promote an adaptive immune response by enhancing anti-
gen presentation and the release of cytokines, leading to immunogenicity. Cytokine
release may also directly induce more immediate infusion-associated reactions.
The effects of the resulting antidrug antibodies (ADA) can range from a small
alteration in the pharmacokinetic properties of the therapeutic protein to a decrease
or complete loss of efficacy due to immune-mediated clearance or neutralization of
the biologic activity of the drug. When both the ADA and the therapeutic protein are
present at sufficiently high levels, such as may occur with therapeutic monoclonal
antibodies, toxicity can occur due to deposition of circulating immune complexes.
In the most severe cases, the immune reaction to the offending therapeutic protein
(e.g., erythropoietin) can cause a breakdown of immune tolerance, resulting in neu-
tralization of the patient’s endogenous counterpart to the therapeutic.
Given the consequences to both activity and toxicity, conformational changes
and protein aggregation affect product quality and can affect product safety. Product
safety and efficacy are determined during clinical trials and must not decline over
the shelf life of a given lot or the market life of a product. It is, thus, essential that
the structural and functional features of the product be characterized to an accept-
able extent during product development and that each batch of the product has simi-
lar characteristics at release and throughout its shelf life. In part, this is ensured by
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 263
product release testing, but because this testing may not be capable of detecting
subtle changes that can occur in the product due to variations in manufacturing
processes, product consistency over time must be ensured by tight control of the
manufacturing process. In addition, any major manufacturing change should be
supported by detailed product characterization and comparability studies, including
a comparison of product stability profiles.
Factors that affect protein structure, aggregation, and stability are numerous,
and, while we acknowledge that not all of them can be controlled during the manu-
facture of biotherapeutic products, the principles outlined in this chapter are meant
to be science-based good manufacturing considerations to be applied when and if
appropriate and feasible for a given specific product.
It cannot be overemphasized that assuring the quality of biologic products is a
complex process that can be achieved only through tight control of process parame-
ters and in-depth characterization beginning with product development and continu-
ing through the lifecycle of the product, including routine testing of every batch and
long-term monitoring of product stability. Recent technical developments allow for
the manufacture of biosynthetic protein therapeutics under tightly defined conditions.
This has increased our potential to limit impurities and control biosynthetic protein
characteristics to a greater extent than is possible for naturally derived products.
When coupled with extensive characterization of protein structure, these advance-
ments better ensure protein comparability after a manufacturing change.
Since Anfinsen’s Nobel Prize winning work (Anfinsen 1973), it is well established
that the primary sequence of the polypeptide chain determines a protein’s three-
dimensional structure. Since then, the discovery of the protein basis of “inborn
errors of metabolism” (Garrod 1908; McCarthy 2011; Benoit et al. 2010; Elia and
Albanese 2010; Percy and Rumi 2009; Bentham and Bhattacharya 2008; Scriver
2007; Pasinelli and Brown 2006; Harrison et al. 2008) and the experience acquired
in the development and clinical use of engineered proteins (Alexander et al. 2007;
Meeker et al. 1996; Stepanenko et al. 2008) have underscored that even seemingly
small changes in protein primary sequence can affect structure and stability. In the
case of therapeutic proteins made from recombinant sources, this understanding
serves to reinforce the importance of the amino acid sequence, its implications for
aggregation propensity, and, as a result, product efficacy and safety.
However, the amino acid sequence is not the only determinant of the protein
three-dimensional structure. Indeed, folding kinetics and the structure of proteins
264 E.B. Struble et al.
Protein folding and its glycosylation can be affected by conditions encountered during
the upstream manufacturing process, including both the intra- and extracellular
milieu. Chaperones such as the Hsp70s play an essential role in correct folding and
secretion of proteins (Kramer et al. 2009). During protein synthesis and
posttranslational modifications, binding to chaperones protects the polypeptide
chain from collapse and incorrect folding that could occur in the absence of the yet
to be translated, downstream sections or from aggregation in the highly crowded cell
environment. Although chaperones are found across all organisms, there are mecha-
nistic differences as well as differences in expression levels amongst commonly
used recombinant protein production systems and cell lines (Hartl and Hayer-Hartl
2009). To add to this complexity, the availability of chaperones depends on other
factors such as temperature, nutrient state (Neuhofer et al. 1999), and the expression
level of the recombinant protein. Chaperone availability in turn can affect folding,
aggregation, and yield of the product. These factors should be carefully analyzed not
only when the process is being developed but also when changes in media formula-
tion and other process parameters are being considered or implemented.
Many cell culture process variables can affect glycosylation. Manufacturing
scale, fermentation type, and fermentation conditions all must be considered.
Bioreactor pH, manganese ions, dissolved oxygen, ammonia concentration, and
temperature have all been shown to affect glycosylation patterns (Hossler et al.
2009). The three major modes of production—batch, fed-batch, and perfusion—all
have different waste product accumulation and nutrient depletion profiles, which
can have profound effects on glycosylation profiles. Accumulation of ammonia,
principally through glutamine and asparagine metabolism, raises media pH, which
can diminish the activity of the late Golgi glycosyltransferases (Thorens and Vassalli
1986). Manganese is required for the activity of some glycosyltransferases, includ-
ing those involved in late carbohydrate modifications such as the addition of
N-acetylneuraminic acid and galactose. Nucleotide sugar precursors and dolichol,
which are substrates for glycosylation biosynthesis, can become depleted, thus
affecting glycosylation efficiency (Castro et al. 1995; Jenkins et al. 1994;
Kochanowski et al. 2008). The mode of production dictates the necessary level of
control over these factors, and therefore, its choice must be weighed with care.
Shear stress has also been shown to affect late Golgi processing (Senger and Karim
2003) and should be taken into account, especially when the scale of production is
changed (Hossler et al. 2009). All of these factors should be considered early in the
development of the manufacturing process as differences in cellular stresses due to
culture scale and production mode can have profound effects on glycosylation pro-
files and thus the conformation, efficacy, and safety of the protein product.
266 E.B. Struble et al.
Protein therapeutic products are derived from highly complex starting materials—human
plasma or blood, transgenic animal milk, bacterial, plant, insect, yeast, and mam-
malian cells or tissues. Multistep purification protocols are often used that may
include precipitation and filtration steps, chromatography steps, and freeze/thawing.
In addition, viral inactivation operations involving filtration, elevated temperatures,
low pH exposure, and/or the use of solvent/detergents are frequently employed.
Virtually every step of the manufacturing process may have an effect on protein
structure (Vazquez-Rey and Lang 2011; Cromwell et al. 2006). Therapeutic pro-
teins are often purified from a mixture containing cell and/or tissue components
based on different surface characteristics that affect solubility and interactions with
solid media. Conditions selected for precipitation operations and chromatography
steps involve pH changes, addition of a nonpolar solvent, or a change in ionic
strength. These changes may alter the distribution of charges on the protein surface
and the properties of the solvent, affecting the interactions of the protein with both
itself and the solvent and thus potentially impact the integrity of the protein
(Zimmerman 2006; Lewis and Nail 1997; Shukla et al. 2007). Temperature also has
an effect on the strength of these interactions, with a rise in temperature strengthen-
ing hydrophobic and reducing ionic interactions. However, ultimately, an increase
in temperature results in higher kinetic energy and will eventually lead to partial
unfolding. Unfolded protein species are highly susceptible to formation of aggre-
gates which are more stable in common formulation media than the native form of
the protein. Additionally, low temperatures may induce cold denaturation, and inter-
action with ice during lyophilization, an operation that involves freezing and dehy-
dration, may also destabilize protein structure (Rathore and Rajan 2008; Tang and
Pikal 2004). Mechanical stress, which affects therapeutics when the protein solu-
tions are pumped, mixed, filtered, or filled, may lead to structural distortions and
aggregation of the protein product (Rathore and Rajan 2008). In addition, chemical
modifications (e.g., oxidation and deamidation) occurring during the purification
process or upon storage may lead to conformational changes, activity reduction, and
aggregation (Luo et al. 2011; Patel et al. 2011). Extractables and leachables are
other potential sources of chemical modifications that may impact protein stability
(Huang et al. 2011). Lastly, lyophilization, used in the manufacture of many biolog-
ics, can result in local concentration and pH changes that may induce conforma-
tional changes in proteins leading to aggregation.
Protein stability upon storage and transportation greatly depends on the protein’s
physical state, i.e., on whether the protein is dissolved or lyophilized and on the
composition of the product formulation. The formulation should be carefully
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 267
optimized not only to ensure long shelf life but also to enable successful completion
of the last steps of manufacturing (Jorgensen et al. 2009; Shire 2009). Other impor-
tant factors include temperature (and potential temperature excursions), presence of
interfaces such as a liquid–air interface, and interactions with the container-closure
system, including leachables, extraneous particles such as glass, metal, and silicone
oil; the latter of which is commonly used as a lubricant in syringes and on stoppers
(Bee et al. 2011). Product agitation during transport and handling may also induce
aggregation of the protein product (Kiese et al. 2008; Thirumangalathu et al. 2009).
Thus, in addition to thorough characterization of the final container product, careful
analysis of the effect of transportation parameters on the protein conformation and
aggregation state of the product is important.
Protein therapeutics, in contrast to small molecule drugs, can never be fully charac-
terized. Their structure is complex, small structural changes are difficult to recog-
nize, and contaminating factors that co-purify with the protein product and which
exist in minute amounts may not be well defined. The quality of biological materials
is verified by final lot release testing of the product and by ensuring the consistency
of manufacture. The latter is achieved by establishing in-process parameters and
product limits that have been demonstrated to deliver a defined quality product.
Extended product characterization is performed for licensure and at the time of any
major manufacturing change. As product quality must be maintained throughout
shelf life, quality attributes are evaluated during stability studies.
268 E.B. Struble et al.
The characterization, utility, and limitations of these and other techniques have been
discussed in numerous reports (see, e.g., Philo 2009; Huang et al. 2009; Fraunhofer
and Winter 2004; Filipe et al. 2010 and Chap. 3 and 4 of this volume). Although the
techniques have evolved in sophistication and sensitivity, analyzing therapeutic pro-
teins or their aggregates in biological media remains a major challenge.
There is an interest in industry to use high throughput screens to assess higher-
order structure. Among the established methods of measuring protein stability, ther-
mal and chemical denaturation monitored by intrinsic fluorescence were the earliest
to be automated (Stites et al. 1995) and are constantly being improved upon (Edgell
et al. 2003; Aucamp et al. 2005). This technology has been reduced to the nanoliter
scale (Gaudet et al. 2010).
A technique that does not depend on the intrinsic fluorescence of a protein, which
can vary widely depending on the local environment and number of fluoroactive
residues, is differential scanning fluorimetry (DSF). This method, based on moni-
toring fluorescence of a specific dye that has increased affinity for the hydrophobic
regions of protein surfaces, has been found to be applicable to a broad range of
proteins (Ericsson et al. 2006, Chap. 2 of this volume), and good correlations have
been found between melting temperature values determined from CD thermal dena-
turation and DSF (Lavinder et al. 2009).
One goal of adopting high throughput assays for determining protein stability is
to sort out the stable variants in a library. To reduce the search space, considerable
effort has been expended in developing first-principles computational methods to
predict the stability of proteins. These programs have relied, to a large extent, on
knowledge-based potentials based on distances between residue pairs or backbone
torsion angles (Gilis and Rooman 2000), statistical potentials (Parthiban et al.
2006), empirical potentials that describe the physical interactions that contribute to
protein stability (Yin et al. 2007), and different types of machine learning tools
(Capriotti et al. 2005; Cheng et al. 2006; Masso and Vaisman 2008). While most
programs require knowledge of the structure of the target protein, some methods
can predict stability changes in proteins of unknown structure with reasonable accu-
racy (Capriotti et al. 2005). While these algorithms are unlikely to produce a com-
prehensive predictive model in the foreseeable future, they can be used in conjunction
with experimental methods.
The above method enumeration is by no means exhaustive. A recent industry
effort to review and evaluate analytical methods available for protein aggregates
detection and characterization was reported upon by den Engelsman et al. (den
Engelsman et al. 2011). Furthermore, there are a large number of publications on
biophysical methods that can be used as a guide to select an appropriate set of
methods. There are a few methods which, although at present are not used widely
by industry, show great potential for evaluating therapeutic biologics and should
be explored further. For example, H/D exchange in conjunction with MALDI
mass spectrometry, another robust and well-characterized method, has recently
been adapted to measure protein stabilities in living cells (Ghaemmaghami and
Oas 2001). It was also recently employed to detect conformational changes asso-
ciated with posttranslational modifications of a monoclonal antibody preparation.
270 E.B. Struble et al.
1
FDA guidance documents contain recommendations that reflect FDA’s current thinking on given
issues and are intended to assist the industry in carrying out its obligations under statutes and regu-
lations. They do not create or confer any rights for or on any person and do not operate to bind FDA
or the public. They are publically available on the FDA website. http://www.fda.gov/Drugs/
GuidanceComplianceRegulatoryInformation/Guidances/default.htm.
2
ICH is an organization that brings together regulators and pharmaceutical industry representatives
from Europe, Japan, and the USA. The ICH mission is to facilitate development of safe, effective,
and high-quality drugs in the most resource-efficient manner by harmonization of guidelines and
requirements for product registration.
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 271
For approval of a new pharmaceutical product, the ICH M4Q(R1) guidance for
industry ICH M4Q(R1) recommends the submission of details on primary, second-
ary, and higher-order structure for “the desired product and product-related sub-
stances.” In addition to information on biological activity of the product, the
guidance recommends the submission of information on the molecular mass, details
of “posttranslational forms (e.g., glycoforms),” and a schematic amino acid sequence
showing glycosylation sites or other posttranslational modifications.
Part B of ICH Q5 guidance discusses the need for characterization of the nucleic
acid coding sequence or the transcription products, as appropriate. This guidance,
published in 1995, although it does not specifically address the effect of the sequence
on protein folding, points out that “the genetic sequence of recombinant proteins
produced in living cells can undergo mutations that could alter the properties of the
protein with potential adverse consequences to patients.” As we know today, even
an alteration in nucleotide sequence that has no effect on the amino acid sequence
may affect the rate of translation and folding of a protein.
ICH Q6B guidance focuses on product physicochemical characterization, which
generally includes determination of the composition, physical properties, and pri-
mary structure of the product. The guidance provides examples of product attributes
that can be considered for structural characterization, such as amino acid composi-
tion, terminal sequencing, peptide mapping, determination of the primary sequence,
location of disulfide bonds, and characterization of the carbohydrate content and
structure. Physicochemical characterization includes determination of the molecular
weight, isoform pattern, extinction coefficient, electrophoretic pattern, liquid chro-
matography patterns, and spectroscopic profiles. The guidance emphasizes that opti-
mal physicochemical characterization varies from product to product and that new
technologies are still being developed and should be applied when appropriate.
Demonstration of product purity presents a challenge, and the results may depend
on the method used (ICH Q6B). The relative purity of the product is often deter-
mined in terms of specific activity expressed in units of biological activity present
in a mass unit of the product. Specific activity determined in such a way is also
highly method dependent. Thus, the purity is assessed by using a combination of
analytical techniques. However, small conformational changes in the protein struc-
ture may be difficult to identify, and guidance on how to efficiently select a method(s)
that identifies such changes in an individual protein is lacking.
ICH Q6B also provides guidance on the types of impurities that may be found in
a product. Impurities may be either process related (i.e., derived from the manufac-
turing process) or product related, e.g., degradation products that may be inactive and
affect the safety profile of the product. Protein aggregates are included in the latter
category. Impurities can have a known structure(s), may be partially characterized, or
remain unidentified. When adequate quantities of impurities are present or can be
272 E.B. Struble et al.
generated by stressing the protein, the guidance advises that they be characterized to
the greatest possible extent and, where feasible, their biological activity evaluated.
Product characterization sets a basis for establishing specifications, which repre-
sent a set of test methods and acceptance criteria that are used in routine product
testing and are chosen not to characterize the product but rather to confirm its
quality. In contrast to methods used for product characterization that need to be
“scientifically sound and provide results that are reliable” (ICH Q5E) (i.e., quali-
fied), methods used in determining specifications and stability should be validated
(ICH Q6B; ICH Q5C; ICH Q2 (R1)). This limits the selection of the methods that
can be considered for routine product testing because some of the biophysical meth-
ods are difficult to validate.
The structural and functional properties of a protein product can change with
time. Thus, it is essential to establish criteria for an acceptable product stability pat-
tern and to determine the product shelf life. The stability of a biological product is
evaluated following ICH Q1A and ICH Q5C guidances. It is recommended that
attributes at risk for change during storage and likely to influence quality, safety,
and/or efficacy be monitored. The manufacturer should develop specifications that
provide assurance that changes in the purity and potency of the product are detected.
ICH Q1A states that “the testing should cover, as appropriate, the physical, chemi-
cal, biological, and microbiological attributes” and that “validated stability-
indicating analytical procedures should be applied.”
The ICH Q6B guidance notes that the inherent degree of structural heterogeneity
in proteins due to the biosynthetic processes used by living organisms or resulting
from manufacture and/or storage necessitates that the pattern of heterogeneity be
defined and lot-to-lot consistency and comparability between lots used in preclini-
cal and clinical studies be demonstrated. The guidance states that “if a consistent
pattern of product heterogeneity is demonstrated, an evaluation of the activity, effi-
cacy, and safety (including immunogenicity) of individual forms may not be neces-
sary.” When process changes and degradation products result in heterogeneity
patterns which differ from those observed in the material used during preclinical
and clinical development, the guidance recommends evaluation of the significance
of such alterations. The ICH Q6B guidance also notes that under certain circum-
stances, physicochemical tests may replace a biological assay to measure the bio-
logical activity. However, such instances are limited to cases where sufficient
physicochemical information about the drug exists, including when the higher-order
structure can be thoroughly established by qualified methods, a relevant physico-
chemical correlation with biologic activity is demonstrated, and a well-established
manufacturing history exists.
Changes to the manufacturing process may have an effect on product composition
and its structural properties. A determination of comparability of the product pre- and
post-manufacturing change can be based on a combination of analytical testing and,
in some cases, nonclinical and clinical data when physicochemical or biological
assays are not considered adequate to confirm that there has been no adverse effect
on the product. ICH Q5E guidance states that “generally, quality data on the pre- and
post-change product are generated, and a comparison is performed that integrates and
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 273
evaluates all data collected, e.g., routine batch analyses, in-process control, process
validation/evaluation data, characterisation and stability, if appropriate. The compari-
son of the results to the predefined criteria should allow an objective assessment of
whether or not the pre- and post-change product are comparable.”
Since manufacturing process changes may have an impact on protein structure,
ICH Q5E advises manufacturers to “attempt to determine that higher-order struc-
ture (secondary, tertiary, and quaternary structure) is maintained.” If the appropriate
higher-order structural information cannot be obtained, the guidance indicates that
a relevant biological activity assay with appropriate precision and accuracy could
indicate that changes in conformational structure have not occurred. However, it
should be noted that the mass of protein aggregates in the product can be so small
that their presence has no meaningful impact on the functional test results. The
small mass of protein aggregates, especially in the subvisible size range, and their
heterogeneity present a significant analytical challenge.
ICH Q5E advises further that to address the full spectrum of physicochemical
properties, more than one analytical technique may be needed to evaluate the same
quality attribute, e.g., the secondary and tertiary structures and presence of impuri-
ties. In such cases, each method should be based on different physicochemical or
biological principles to maximize the chance that differences in the product caused
by a change in the manufacturing process are detected.
Proteins may be sensitive to changes in buffer composition, processing and hold-
ing conditions, and the use of organic solvents. Even slight modifications to manu-
facturing, storage, and handling may affect the stability of the product. For example,
the presence of trace amounts of metal ions might activate proteases or catalyze
chemical modifications of the product that may lead to protein aggregation.
Therefore, ICH Q5E recommends initiation of real-time/real temperature stability
studies on the product potentially affected by the change. Stability studies per-
formed under accelerated conditions, such as increased temperature, light intensity,
agitation, and freeze–thaw cycles, may be particularly helpful in identifying subtle
changes that are not detectable by routine product characterization, such as may
occur following a manufacturing change.
et al. 2010b). These limitations include sample dilution, which has the potential to
shift the equilibrium for soluble aggregate formation, and exclusion/adsorption of
aggregates on the SE-HPLC matrix. Also, extinction coefficients may be conforma-
tion dependent, and using the same extinction coefficient for different protein
species may lead to erroneous results. Solutions proposed include the use of orthog-
onal methods, determination of mass recovery, confirmation of method suitability
by using stressed protein samples, and development of new media. For HPLC and
other methods used for protein aggregates characterization, development of protein
standards would be a useful and important advance toward standardizing various
characterization methods.
Until recently, protein aggregates with sizes between 100 nm and 10 μm were
neither quantified nor characterized (Carpenter et al. 2009), in part due to the lack of
relevant analytical techniques. Recently, there is increasing interest in characteriza-
tion of protein aggregates in this size range. It has been postulated that aggregates of
this size may be of particular importance due to risk of an immune response stimu-
lated by their presence. Fortunately, recent advances in the analytical field, including
the development of new methods (e.g., resonant mass measurement-based tech-
niques, nanoparticle tracking analysis) and expanding the dynamic range of existing
methods (flow microscopy), may eventually allow this gap in our understanding of
the impact of subvisible particles on product quality and safety to be bridged.
Standards for the amount of particulate matter in the final container of injectable
therapeutics are recommended by the pharmacopeias (ICH Q6B). United States
Pharmacopeial Convention (USP)3 requirements for particulate matter in parenteral
products are described in general chapters “Injections” (USP <1>) and “Particulate
Matter in Injections” (USP <788>). According to the USP, therapeutic protein solu-
tions need to be “essentially free from visible particles.” Each container of the prod-
uct should be visually inspected for the presence of observable particulate and other
foreign matter (“visible particulates”). Every container showing evidence of visible
particulates should be rejected. However, this standard was set for “particulate mat-
ter” defined as “extraneous mobile undissolved particles, other than gas bubbles,
unintentionally present in the solutions.” As such, it cannot be directly applied to all
biological products because protein aggregates are often unavoidable in protein ther-
apeutics, some of which may contain protein aggregates even in the visible size range.
The meaning of the term “essentially free” has been under discussion for an
extended period of time, and attempts to implement some numerical standards have
been made (Madsen et al. 2009). Recently, USP has proposed a new general chapter
“Visible particulates in injections” (USP <790>), which includes a specification for
determining whether a product meets the criterion of being “essentially free” of par-
ticles. The chapter also states that “some products, such as those derived from pro-
teins, may contain intrinsic particles of agglomerates” and in such cases recommends
meeting the requirements of individual monographs.
3
USP is a standard setting organization that can aid in ensuring the quality and safety of drugs,
dietary supplements, and food ingredients.
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 275
Particles with sizes ≥10 μm have been quantified by using the USP chapter
<788> methods, light obscuration, and/or microscopic test. The methods and the
acceptance criteria were developed for extraneous foreign particles and should not
be applied to biologics without critical analysis. The microscopy method, which is
applicable to non-proteinaceous products, should not be used for characterizing
protein aggregates because they are fragile and translucent, may not be visible on
the membrane used for filtration, or may break down and be filtered out.
The limitations of the light obscuration method include the need for large sample
volume, which given biological product final container volumes may render the
method very costly. The details of sample preparation are also problematic for bio-
logical materials; protein aggregates may dissociate or form under the sample
degassing conditions (sonication) recommended in chapter <788>. Also, the method
cannot distinguish between protein aggregates and extraneous, non-proteinaceous
matter (e.g., silicone oil). More objective and data-driven ways to determine the
acceptance criteria for particulates in biologic products are being developed based
on process performance together with the product safety profile.
The exceptions that are not subject to the requirements of chapter <788> include
parenteral products for which the labeling specifies the use of a final filter prior to
administration, provided scientific data exist to justify the exemption, such as in-use
qualification of the filter. However, we note that the presence of protein aggregates
represents not only a potential safety problem but is also a product quality issue.
In-line filters may become clogged by protein aggregates, exposing patients to the
inconvenience and potential safety risk of having filters changed during infusion of
a large volume biologics; thus, the quantity of the aggregates should be controlled
even if the product is filtered.
Some of the above limitations were considered during the development of a new
USP general chapter “Subvisible particulate matter in therapeutic protein injec-
tions” (USP <787>), which addresses the presence of subvisible particles in bio-
therapeutics and has been published in the Pharmacopeial Forum for comments.
Also of note are two recent improvements in methodology. Due to technical
advancements, the light obscuration method can be validated to monitor particle
size down to 2 μm. In addition, new and promising flow imaging methods may help
characterize particulate matter content in the size range that earlier was mainly
assessed by light obscuration.
Quantification and characterization of protein aggregates is not an easy task. The
complexity is related to differences in protein propensity to aggregate, heterogene-
ity of aggregation, and unique aggregate-associated potential risk. As discussed
above, even with the existing gaps, many analytical methods are available for use,
although identifying the relevant test(s) for a specific product is not trivial given the
complexities of the protein therapeutic products. It is extremely encouraging that
members of the biologics community, which includes industry, academic, and gov-
ernment scientists and regulators, are working together to develop a better under-
standing of protein aggregates in protein therapeutics and to devise effective
methods to identify and mitigate the associated risks (Marszal and Fowler 2012).
276 E.B. Struble et al.
11.6 Conclusion
References
Aguzzi A, O’Connor T (2010) Protein aggregation diseases: pathogenicity and therapeutic per-
spectives. Nat Rev Drug Discov 9:237–248
Alexander PA, He Y, Chen Y, Orban J, Bryan PN (2007) The design and characterization of two
proteins with 88 % sequence identity but different structure and function. Proc Natl Acad Sci
USA 104:11963–11968
Anderson DE, Becktel WJ, Dahlquist FW (1990) pH-induced denaturation of proteins: a single
salt bridge contributes 3–5 Kcal/Mol to the free energy of folding of T4 lysozyme. Biochemistry
29:2403–2408
Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181:223–230
Appa RS, Theill C, Hansen L, Moss J, Behrens C, Nicolaisen EM, Klausen NK, Christensen MS
(2010) Investigating clearance mechanisms for recombinant activated factor VII in a perfused
liver model. Thromb Haemost 104:243–51
Aucamp JP, Cosme AM, Lye GJ, Dalby PA (2005) High-throughput measurement of protein sta-
bility in microtiter plates. Biotechnol Bioeng 89:599–607
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 277
Moore WV, Leppert P (1980) Role of aggregated human growth hormone (hGH) in development
of antibodies to hGH. J Clin Endocrinol Metab 51:691–697
Neuhofer W, Muller E, Grunbein R, Thurau K, Beck FX (1999) Influence of NaCl, urea, potassium
and pH on HSP72 expression in MDCK cells. Pflugers Arch 439:195–200
Obici L, Perfetti V, Palladini G, Moratti R, Merlini G (2005) Clinical aspects of systemic amyloid
diseases. Biochim Biophys Acta 1753:11–22
Parthiban V, Gromiha MM, Schomburg D (2006) CUPSAT: prediction of protein stability upon
point mutations. Nucleic Acids Res 34:W239–W242
Pasinelli P, Brown RH (2006) Molecular biology of amyotrophic lateral sclerosis: insights from
genetics. Nat Rev Neurosci 7:710–723
Patel J, Kothari R, Tunga R, Ritter NM, Tunga BS (2011) Stability considerations for biopharma-
ceuticals, Part 1. Overview of protein and peptide degradation pathways. BioProcess Int
9:20–31
Percy MJ, Rumi E (2009) Genetic origins and clinical phenotype of familial and acquired erythro-
cytosis and thrombocytosis. Am J Hematol 84:46–54
Pfeil W, Privalov PL (1976) Thermodynamic investigations of proteins. III. Thermodynamic
description of lysozyme. Biophys Chem 4:41–50
Philo JS (2009) A critical review of methods for size characterization of non-particulate protein
aggregates. Curr Pharm Biotechnol 10:359–372
Rademacher TW, Parekh RB, Dwek RA (1988) Glycobiology. Annu Rev Biochem 57:785–838
Rathore N, Rajan RS (2008) Current perspectives on stability of protein drug products during
formulation, fill and finish operations. Biotechnol Prog 24:504–514
Rosgen J (2007) Molecular basis of osmolyte effects on protein and metabolites. Methods Enzymol
428:459–486
Scriver CR (2007) The PAH gene, phenylketonuria, and a paradigm shift. Hum Mutat
28:831–845
Senger RS, Karim MN (2003) Effect of shear stress on intrinsic CHO culture state and glycosyl-
ation of recombinant tissue-type plasminogen activator protein. Biotechnol Prog 19:1199–209
Sharp PM, Cowe E, Higgins DG, Shields DC, Wolfe KH, Wright F (1988) Codon usage patterns
in Escherichia coli, Bacillus subtilis, Saccharomyces cerevisiae, Schizosaccharomyces pombe,
Drosophila melanogaster and Homo sapiens; a review of the considerable within-species diver-
sity. Nucleic Acids Res 16:8207–8211
Shire SJ (2009) Formulation and manufacturability of biologics. Curr Opin Biotechnol
20:708–714
Shukla AA, Gupta P, Han X (2007) Protein aggregation kinetics during Protein A chromatography:
case study for an Fc fusion protein. J Chromatogr A 1171:22–28
Sola RJ, Rodriguez-Martinez JA, Griebenow K (2007) Modulation of protein biophysical proper-
ties by chemical glycosylation: biochemical insights and biomedical implications. Cell Mol
Life Sci 64:2133–2152
Stepanenko OV, Verkhusha VV, Shavlovsky MM, Kuznetsova IM, Uversky VN, Turoverov KK
(2008) Understanding the role of Arg96 in structure and stability of green fluorescent protein.
Proteins 73:539–551
Stites WE, Byrne MP, Aviv J, Kaplan M, Curtis PM (1995) Instrumentation for automated deter-
mination of protein stability. Anal Biochem 227:112–122
Struble EB, Ladner JE, Brabazon DM, Marino JP (2008) New crystal structures of ColE1 Rom and
variants resulting from mutation of a surface exposed residue: implications for RNA-
recognition. Proteins 72:761–768
Tang X, Pikal MJ (2004) Design of freeze-drying processes for pharmaceuticals: practical advice.
Pharm Res 21:191–200
Thirumangalathu R, Krishnan S, Ricci MS, Brems DN, Randolph TW, Carpenter JF (2009)
Silicone oil- and agitation-induced aggregation of a monoclonal antibody in aqueous solution.
J Pharm Sci 98:3167–3181
Thorens B, Vassalli P (1986) Chloroquine and ammonium chloride prevent terminal glycosylation
of immunoglobulins in plasma cells without affecting secretion. Nature 321:618–620
11 Higher-Order Structure and Protein Aggregate Characterization of Protein… 281
Uchida E, Morimoto K, Kawasaki N, Izaki Y, Abdu Said A, Hayakawa T (1997) Effect of active
oxygen radicals on protein and carbohydrate moieties of recombinant human erythropoietin.
Free Radic Res 27:311–323
Vazquez-Rey M, Lang DA (2011) Aggregates in monoclonal antibody manufacturing processes.
Biotechnol Bioeng 108:1494–1508
Wright A, Morrison SL (1998) Effect of C2-associated carbohydrate structure on Ig effector func-
tion: studies with chimeric mouse-human IgG1 antibodies in glycosylation mutants of Chinese
hamster ovary cells. J Immunol 160:3393–3402
Xu J, Reumers J, Couceiro JR, De SF, Gallardo R, Rudyak S, Cornelis A, Rozenski J, Zwolinska
A, Marine JC et al (2011) Gain of function of mutant P53 by coaggregation with multiple
tumor suppressors. Nat Chem Biol 7:285–295
Yin S, Ding F, Dokholyan NV (2007) Modeling backbone flexibility improves protein stability
estimation. Structure 15:1567–1576
Zimmerman TP (2006) Yield improvement for manufacture of α1-proteinase inhibitor. Vox Sang
91:309–315
Index
A application, 155
Absorbance spectroscopy, 36–38 Fc-fusion protein, 158
ADA. See Antidrug antibodies (ADA) SEC, 156
ADCs. See Antibody drug Antibody drug conjugates (ADCs)
conjugates (ADCs) chemical and biophysical techniques,
AFM. See Atomic force microscope (AFM) 195–196
Aggregation DSC analysis, 196–197
CDR loops, 129 in vivo assays and DAR, 196
conformational and colloidal stability, 135 SEC, 196, 197
in silico protein, 131 small molecule cytotoxic drugs, 197
Lumry–Eyring framework, 134–135 T-DM1, 195
soluble protein, 206–219 Antidrug antibodies (ADA), 262
thermal stability, 135–136 Atomic force microscope (AFM)
Aggregation propensity cantilever, 221
circular dichroism, 164–165 measurements, 221
extrinsic fluorescence, 164, 165 morphology, submicron protein
high molecular mass species, 164 particulates, 223, 225
linear regression analysis, 167 AUC. See Analytical ultracentrifugation
Nile red, 163 (AUC)
stressed samples, 166
thermal incubation, 164
Analytical ultracentrifugation (AUC) B
advantages and limitations, 89 Biopharmaceuticals development, protein
orthogonal technique to SEC HOS. See Higher-order structure
concordance plot, 198, 199 (HOS)
HMWS and LMWS, 198 Biophysical analysis, protein
mAbs, 198 pharmaceuticals
SE-AUC, 70–71 AUC, 198–199
sedimentation velocity and sedimentation characterization (see Biophysical
equilibrium, 88–89 characterization)
SV-AUC, 66–70 description, 174–175
use, 88–89 FFF, 199–200
Analytical ultracentrifugation- orthogonal techniques and SEC, 197
gravity-sweep, 148 protein stability development, 173–174
Analytical ultracentrifugation-sedimentation screening assessments (see Screening
velocity (AUC-SV) assessments, biophysical analysis)
L.O. Narhi (ed.), Biophysics for Therapeutic Protein Development, Biophysics 283
for the Life Sciences 4, DOI 10.1007/978-1-4614-4316-2,
© Springer Science+Business Media New York 2013
284 Index