Advanced Spectroscopy in Biopharmaceutical Manufacturing
eBook
Published: October 27, 2025
Credit: Thermo
The biopharmaceutical industry demands rapid, accurate and regulatory-compliant testing to ensure product quality and process efficiency. However, traditional analytical methods often lack the speed or sensitivity needed for complex biologics. As a result, innovative spectroscopic techniques are transforming every stage of bioprocessing, from material verification to real-time release testing.
This eBook highlights how vibrational and UV-Vis spectroscopy enable precise, non-destructive analysis across upstream, downstream and final product workflows.
Download this eBook to discover:
- How MIR, NIR, Raman and UV-Vis spectroscopy improve analytical precision and efficiency
- New applications for process monitoring and critical quality attribute assessment
- Real-world examples demonstrating faster, data-driven biopharmaceutical development
BioPharmaceutical
approach with spectroscopy
Compendium
Summary
Heavily-regulated biopharmaceutical manufacturers are increasing their use of the
molecular spectroscopy techniques mid-infrared (MIR), near infrared (NIR), and
Raman UV-Vis spectroscopy because of these techniques’ rapid, accurate analysis
capabilities and their complementary nature. MIR spectroscopy is the analytical tool
of choice for material verification in small molecule manufacturing due to its simplicity
of implementation and its reliability and specificity. Recently, Raman spectroscopy has
gained popularity in large molecule manufacturing, since it has increased sensitivity
because of the resonance enhancement caused by the large size of molecules, as
well as sensitivity to polymorphism. For certain applications like positive raw material
verification of protein purification resins, the use of photoacoustic spectroscopy with
Fourier-transform infrared (FTIR) offers unique selectivity and sensitivity. Vibrational
spectroscopy plays a major role for analysis in upstream, downstream, and fill-finish
processes. To support upstream processes, MIR, NIR, and Raman spectroscopy
can be utilized for multi-attribute raw material testing. Further, in downstream
processing, critical quality attributes (CQAs) like glycosylation, aggregation, and
degradation can be determined at line or inline using Raman spectroscopy. Recently,
NIR spectroscopy was demonstrated to have a wide variety of potential applications
to improve speed and efficiency in different downstream unit operations, including
capture chromatography, protein PEGylation reactions, and tangential flow ultrafiltration.
Real-time release testing (RTRT) in biopharmaceutical manufacturing has increasing
importance. In this regard, macro-Raman measurements through primary packaging
offer faster alternative tests for CQAs like pH, osmolality, and potency (strength), along
with the positive identification of the drug product.
In addition, identification can be accomplished with macro- and micro-Raman
spectroscopy in lyophilized biopharmaceutical analysis of cake morphology,
siliconization, distribution of the drugs along with foreign particulate.
Currently, CQAs such as moisture and potency are commonly determined using
destructive and time-intensive techniques like Karl Fischer titration. Since NIR
spectroscopy is very sensitive to moisture, employing NIR for such analysis allows
accurate and repeatable analysis of lyophilized cake through glass vials
or containers in less than a minute.
Another technique, UV-Vis spectroscopy is used for quantitative analysis across the
biopharmaceutical industry. Applications range from quantifying small volumes of
genomic material to assessing protein aggregation in large volume samples. Accurate
quantitation is a critical step within R&D and QA/QC.
Protein secondary structure elucidation
using FTIR spectroscopy 4
Protein concentration prediction in cell cultures 8
Protein aggregation identified through
UV-Visible absorption spectroscopy 11
Quantify protein and peptide preparations
at 205 nm 16
Spectrophotometric Analysis of Ibuprofen
According to USP & EP Monographs 19
Observation of gold nanoshell plasmon
resonance shifts after bioconjugation 23
Color analysis for pharmaceutical products
using UV-Visible absorption techniques 26
The NanoDrop Eight Spectrophotometer
detects contaminating nucleic acids in
mammalian DNA and RNA preparations 32
Enabling real-time release of final products
in manufacturing of biologics 35
Table of Contents
Protein secondary structure elucidation using
FTIR spectroscopy
Application note
Keywords
FTIR, ATR, protein structure
elucidation, Biocell calcium fluoride
cell, ConcentrateIR2 ATR, transmission
Author
Suja Sukumaran
Thermo Fisher Scientific, USA
Abstract
Fourier-transform infrared (FTIR) spectroscopy is one of the most versatile analytical
tools used across various disciplines. In this study, the Thermo Scientific™ Nicolet™
iS10 and Nicolet iS50 FTIR Spectrometers, equipped with attenuated total reflection
(ATR) FTIR and transmission FTIR, were used for the determination of protein
secondary structures. Structure calculations based on a protein database as well as
spectral deconvolution are discussed. The analyses were quick and easy.
Introduction
Protein secondary structure describes the repetitive conformations of proteins and
peptides. There are two major forms of secondary structure, the α-helix and β-sheet,
so named for the patterns of hydrogen bonds between amine hydrogen and carbonyl
oxygen atoms that create the peptide backbone of a protein.1 Understanding protein
secondary structure is important to gain insight into protein conformation and
stability. For example, temperature dependent analysis of the secondary structure is
critical in determining storage conditions for maintaining active therapeutic proteins.2
Protein secondary structure is also crucial in understanding the structure–function
relationship and enzyme kinetics of various proteins.3
FTIR has long been established as a powerful analytical technique to investigate
protein secondary structure and local conformational changes.1, 4 A typical protein
infrared (IR) spectrum often contains nine amide bands, with vibrational contributions
from both protein backbone and amino acid side chains. Among which, of particular
pertinence to protein secondary structure are amide I and amide II bands. The
absorptions associated with C=O stretching are denoted as amide I, whereas those
associated with N–H bending are amide II. Since both C=O and N–H bonds are
involved in the hydrogen bonding between different moieties of secondary structure,
the positions of both amide I and amide II bands are sensitive to the secondary
structure composition of a protein,3, 4 although the amide II band is widely viewed
as a less useful predictor for quantifying the secondary structure of proteins.
4
The shifts in the amide I band are often small compared
to the intrinsic width of the band, resulting in one broad peak
instead of a series of resolved peaks for each type of the
secondary structure. Mathematical procedures such as Fourier
self-deconvolution and second derivatives can be used
to resolve the overlapping bands for the quantitative analysis
of protein secondary structure.3 Table 1 shows the
secondary structure band assignments for proteins in water.
Note that all assignments are depicted as a range, as the
exact position of each peak varies from protein to protein due
to the differences in hydrogen bonding interactions and the
environment of the proteins.
With a range of sampling techniques, including transmission,
ATR, and infrared reflection absorption spectroscopy (IRRAS),
FTIR is particularly advantageous in terms of its versatility and
general applicability compared to other analytical techniques
for protein secondary structure analysis. Protein sample forms
suitable for FTIR analysis include lyophilized powders, water
solution, and colloids, to name a few. We report herein two
examples of protein secondary structure determination using
transmission FTIR and ATR, respectively. Both methods are
fast, consume a minute amount of sample, and require minimal
sample preparation.
Experiment
All proteins were procured from Sigma-Aldrich (MO, USA)
and used as received. For the transmission studies, a BioCell™
Calcium Fluoride Cell (Biotools, Jupiter, FL) was used, and
all measurements were carried out at ambient temperature. A
10 μL protein solution was placed at the center of the window,
and the protein solution was sandwiched between the two
CaF2 windows, and placed in the holder. The concentration
of protein tested was between 6 and 12 mg/mL. A 6 μm
path length was created by sandwiching the two CaF2
windows. CaF2 windows are suited for water-based sample
analysis. As water has a significant absorption peak at
1,645 cm-1 region, a small path length of 6 μm can effectively
avoid saturated water peaks.
A purged Nicolet iS10 FTIR Spectrometer, equipped with
a DTGS detector, was used for transmission analysis. The scan
parameters used were 256 scans with a resolution of 4 cm-1.
The Thermo Scientific Smart OMNI-Transmission™ Accessory
allows for a quick purge of the chamber, eliminating the need
for water vapor subtraction in most analyses. Secondary
structure analysis of the buffer-subtracted spectra was carried
out using the built-in feature of the PROTA-3S™ FT-IR Protein
Structure Analysis Software. Secondary structure calculation
in PROTA-3S software is based on a database of 47 secondary
structures (for more information visit www.btools.com).
For ATR analysis, a ConcentratIR2™ Multiple Reflection ATR
Accessory (Harrick Scientific Products, Inc. Pleasantville,
NY) with diamond crystal was used in a Nicolet iS50 FTIR
spectrometer equipped with a mercuric cadmium telluride
(MCT) detector. The diamond ATR has ten internal reflections
with a nominal angle of incidence of 45 degrees. A 10 μL
protein solution in phosphate buffer was dried on the surface
of the ATR crystal under a stream of nitrogen. Scan parameters
used were 256 scans and a resolution of 4 cm-1. Secondary
structure determination was carried out using the peak resolve
feature of the OMNIC™ Software.
Results and discussion
Transmission-FTIR with Bio Cell
Figure 1 shows the overlay of three FTIR spectra: phosphate
buffer, cytochrome C at 6 mg/mL and 12 mg/mL in
phosphate buffer, respectively. At first glance, the spectra
are predominantly water bands. The three spectra show little
difference, even at a high protein concentration of 12 mg/mL.
Table 1. Secondary structure band assignments for protein in water.2
Figure 1. Transmission-FTIR spectra for cytochrome C in phosphate buffer (cytc_12) at 12 mg/mL and
6 mg/mL (cytc_6), and phosphate buffer blank.
Secondary structure Band assignment in water
α-Helix 1,648–1,657 cm-1
β-Sheet
(high-frequency component)
1,623–1,641 cm-1
1,674–1,695 cm-1
Random 1,642–1,657 cm-1
Coils 1,662–1,686 cm-1
Absorbance
Wavenumbers (cm-1)
3.0
2.8
2.6
2.4
2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
3,500 3,000 2,500 2,000 1,500
5
Next, the buffer spectrum was subtracted from the raw protein
spectra using the PROTA-3S software, and the results are
shown in Figures 2A (cytochrome C) and 2B (concanavalin).
The amide I and II peaks are clearly discernible in both
spectra. The amide I peak position for cytochrome C spectra
is 1,654 cm-1, suggesting an α-helix dominant secondary
structure. For concanavalin A, the amide I peak centers
at 1,633 cm-1, and there is also a noticeable shoulder peak at
1,690 cm-1 (red circle), indicative of the β-sheet component
and its associated high-frequency component.2
Table 2 summarizes the secondary structure prediction using
the PROTA-3S software. The cytochrome C has 45% α-helix
and 5% β-sheet, whereas concanavalin A has 42% β-sheet
and 4% α-helix. Differences in secondary structure
composition between X-ray and FTIR data are likely due to
the physicochemical state of the protein samples such as
crystalline versus solution, temperature, pH, buffer conditions,
etc. Furthermore, different prediction algorithms could have
slightly varying outputs.7 Notwithstanding the differences in
analytical technique, sample state, and prediction algorithms,
the secondary structure elucidation by FTIR using PROTA-3S
software is largely in line with that from X-ray. Transmission-
FTIR measurements combined with PROTA-3S software offer
a facile and fast means to analyze the secondary structure of
proteins in solution2, 3 with minimal sample prep.
ATR-FTIR with ConcentratIR2 Accessory
When the quantity and concentration of protein are limited,
FTIR measurements with the ConcentratIR2 Multiple
Reflection ATR offer a better alternative than transmission-
FTIR spectroscopy. The unique design of this ATR accessory
allows for the direct measurement of protein powders, gels,
solutions as well as proteins dried on the ATR surface.
When concentrating proteins on the crystal surface, caution
should be exercised in buffer selection since buffer will also
concentrate on the surface of the crystal.
Only those buffers with minimum or no peaks in the amide I
and II region should be selected. Figure 3 shows the ATR-FTIR
spectra of BSA in phosphate buffer, dried on the crystal from a
1 mg/mL solution. In addition to the amide I and II bands, there
are spectral features of the side chain, such as 1,515 cm-1 from
tyrosine and 1,498 cm-1 from aspartic acid. Side chain peaks
are critical for the elucidation of protonation and de-protonation
states of various amino acids.2
α-Helix (%) β-Sheet (%) Random (%)
Protein FTIR X-ray FTIR X-ray FTIR X-ray
Cytochrome C 45 41 5 0 50 59
Concanavalin A 4 0 42 48 54 52
Figure 2. FTIR spectra of (a) cytochrome C and (b) concanavalin A after the buffer spectrum was subtracted using PROTA-3S software.
Table 2. Comparison of secondary structure calculation from FTIR
(PROTA-3S) and X-ray data.
Figure 3. Amide I and II for 1 mg/mL BSA analyzed using ConcentratIR2 ATR on the Nicolet iS50 FTIR
Spectrometer equipped with an MCT detector.
Absorbance
Wavenumbers (cm-1)
Wavenumbers (cm-1)
0.045
0.040
0.036
0.030
0.025
0.020
0.015
0.010
1,800
1,750 1,700 1,650 1,600 1,550 1,500 1,450 1,400
Wavenumbers (cm-1)
1
0.8
0.6
0.4
0.2
0
-0.2
1,750 1,700 1,650 1,600 1,550 1,500 1,450
1,700 1,600 1,500 1,400 1,300 1,200 1,100 1,000
Absorbance
1
0.8
0.6
0.4
0.2
0
-0.2
Absorbance
6
Peak deconvolution of the amide I peak (Figure 4) of BSA
was carried out using the OMNIC software. It is important
to note that second derivative analysis is often performed
prior to deconvolution to clearly identify the peaks
required for peak fitting.2 In the current study, the second
derivative peaks obtained (results not shown) are well
correlated to the secondary structure peak assignments
in Table 1. In order to obtain a good peak shape for
peak fitting, a baseline correction on the amide I region
was also performed. Baseline correction also effectively
excluded the contributions from the amide II region. The
deconvolution of amide I resulted in 5 peaks, and the area
under each peak was then evaluated against the total area.
Amide I peak deconvolution shows a secondary structure
composition of 47% α-helix, 3% β-sheet, 24% coils, and
26% random, which is to published FTIR5 and X-ray data.
Conclusion
In this note, we have demonstrated two examples of protein
secondary structure elucidation using FTIR spectroscopy.
Transmission-FTIR measurements combined with
PROTA-3S software provides a facile means to analyze
secondary structure of proteins in solution with minimal sample
preparation. When the quantity and concentration
of protein are limited, ATR-FTIR offers a better alternative
by drying the proteins in ATR crystals directly. The data were
collected using an older model, the Nicolet iS10 Spectrometer.
An improved model, the Nicolet iS20 Spectrometer, offers
superior speed and performance over this predecessor model.
References
1. Elliott, A., Ambrose, E. J. Structure of synthetic polypeptides,
Nature (1950) 165, 921-922.
2. Jackson, M., Mantsch, H.H. The use and misuse of FTIR spectroscopy
in the determination of protein structure, Crit. Rev. Biochem. Mol. Biol.
(1995) 30, 95-120.
3. Barth, A. Infrared spectroscopy of proteins, Biochim. Biophys.
Acta (2007) 1767, 1073-1101.
4. Byler, D.M., Susi, H. Examination of the secondary structure of proteins
by deconvolved FTIR spectra, Biopolymers (1986) 25, 469-487.
5. Surewicz, W.K., Mantsch, H.H. New insight into protein secondary
structure from resolution-enhanced infrared spectra, Biochim. Biophys.
Acta (1988) 952, 115-130.
6. Sukumaran, S., Hauser, K., Maier, E., Benz, R., Mantele, W. Tracking the
unfolding and refolding pathways of outer membrane protein porin from
Paracoccus denitrificans, Biochemistry (2006) 45, 3972–3980.
7. Klose, D., Janes R.W. 2Struc – the protein secondary structure analysis
server, Biophysical Journal (2010) 98, 454-455.
Learn more at thermofisher.com/brighteroutcomes
Figure 4: Peak deconvolution of amide I peak of BSA using Peak Resolve function of OMNIC software.
Absorbance
Wavenumbers (cm-1)
1.00
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
1,720 1,700 1,680 1,660 1,640 1,620 1,600
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. BioCell and PROTA-3S are trademarks of BioTools, Inc. ConcentratIR2
is a trademark of Harrick Scientific Products, Inc. All other trademarks are the property of Thermo Fisher Scientific and its subsidiaries
unless otherwise specified. AN52985_E 4/22
Protein concentration prediction in cell cultures
The next stage in NIR bioprocess analysis
Application note
Author
Todd Strother, PhD, Thermo Fisher
Scientific, Madison, WI, USA
Introduction
Biologically produced materials are an increasingly important aspect in many
industrial processes including those related to pharmaceuticals, food, diagnostics,
and fuels. Most of these biologicals are produced in fermentors and bioreactors
in which specialized cell cultures grow and manufacture the molecule of interest.
Many different types of cells are used in culturing and producing biopharmaceutical
products including genetically engineered bacterial and yeast cells. However
a majority of the products are proteins cultured from mammalian systems such
as Chinese hamster ovary (CHO), green monkey (VERO), or human embryonic kidney
(HEK) cell lines. Many of these products are large complex proteins, hormones
or polysaccharides that are impossible or difficult to manufacture in large quantities
any other way. A recent survey of the US Food and Drug Administration noted that
there are over 350 biologicals approved for various uses, including vaccines and
diagnostic and therapeutically important antibodies.
Bioprocesses that produce the desired materials by nature rely on complex biological
systems to synthesize their useful products. While typical chemical manufacturing
processes have relatively little variability, the inherent complexity of biological systems
makes a great deal of variability from batch to batch inevitable. As a consequence
of the complexity and variability of the processes, it has been estimated that 30%
of the production batches need to be reprocessed for quality reasons, which results
in a tenfold loss in profit. Industries that rely on these complex biological systems
benefit greatly from closely monitoring the growth of their cell cultures and production
of the target molecule. Process analytical technology (PAT) initiatives in bioprocesses
improve the overall product quality, reducing waste by accounting for this
inherent variability.
Monitoring and controlling cell culture conditions greatly reduces this variability and
results in improved target protein production. Fourier transform near-infrared (FT-NIR)
spectroscopy has proven to be a useful technology for monitoring and controlling
manufacturing processes including more specific bioprocess applications. It is also
part of PAT initiatives across many industries including bioprocessing. Previous work
performed on cell cultures using NIR spectroscopy has usually focused on monitoring
and controlling nutrients, waste products, cell densities and other parameters related
to the health of the cell culture. While these parameters are useful for determining the
relative health of the cell culture, the more important parameter of interest is the actual
production and concentration of the target molecule. Very few NIR studies have
determined and measured protein concentrations in actual cell culture conditions.
This application note demonstrates the feasibility of using the Thermo Scientific™
Antaris™ MX FT-NIR Process Analyzer (Figure 1) to predict protein concentrations at
biologically relevant concentrations in dynamic cell cultures.
Figure 1. Antaris MX FT-NIR Process
Analyzer used for collecting thespectroscopic
information from the cell cultures.
8
NIR spectroscopy uses light between 10,000 and 4,000
cm-1 to determine the identity and quantity of a variety of
materials. Most molecules of interest absorb light in this region
through combination or overtone vibrations. The advantage
of performing spectral analysis on these absorption bands is
that the light is able to penetrate more deeply into the material
under analysis and does not require dilution or manipulation
of the sample. Therefore NIR analyzers can be coupled directly
into a process stream or tank where spectral analysis can
be performed without human intervention. FT-NIR has been
implemented in many different industrial, pharmaceutical
and other process settings for many years and has proven to
be extremely valuable in collecting real-time analytical data
automatically. When used in process environments, the Antaris
MX FT-NIR Process Analyzer is easily coupled to process
control computers where it is an integral part of maintaining
optimal manufacturing conditions. Because of these
advantages and the need to control the inherently variable
biological systems found in cell culture technologies, NIR is an
excellent choice for analyzing different components
in bioreactors including proteins.
Methods
Chinese hamster ovary (CHO) cell cultures were grown at
optimal conditions until the cell concentrations reached
approximately one million per millemeter, representing a typical
cell density for a young and growing culture. Samples of the
cell culture were tested on a Nova BioProfile® analyzer to
determine concentrations of glucose, glutamine, lactate, and
ammonia. The concentrations of these materials changed
throughout the experiment and accounted for some variability
that might be encountered across multiple cultures. The
concentrations were variously and singularly altered by
spiking the samples with nutrients or waste products or
diluting the samples with unaltered cell culture. Each of those
four components was altered so that two or three different
concentrations were represented for each. Table 1 lists the
concentration ranges for the various nutrient, waste, and
protein components of the tested samples. This methodology
also has the effect of removing covariance between the
different components and protein present.
Ultrapure bovine albumin protein was added to the solutions
to represent target protein synthesized by the cells. Genetically
modified cell cultures are designed to produce the target
protein in large quantities almost exclusively to all other cellular
proteins. As a result, the protein concentrations in the cell
culture media will often approach and exceed 5.0 g/L and
consist almost entirely of the target molecule. Albumin protein
is an excellent mimic for recombinant proteins because it
is available in extremely pure form and contains NIR active
groups essentially identical to a typical target protein from a
cell culture. In this case, purity is extremely important because
any extraneous material present will also have a NIR signal and
would lead to confounding results. The albumin protein material
was carefully weighed and added to the cell cultures
in concentrations ranging from 0.16 to 5.0 g/L. Over 35 different
solutions were produced that had a range of nutrient and
waste as well as protein concentrations. These varied solutions
resulted in 54 spectra that were used to build the chemometric
method and 20 spectra that were used to validate that method.
The cell culture samples were scanned with an Antaris MX
FT-NIR Process Analyzer in the range between 10,000 and
4,000 cm-1. The analyzer was coupled to a transflectance
probe with an adjustable path length. The gap distance was
set to 1.25 mm for a total path length of 2.5 mm. Sixteen scans
were averaged per spectrum and were collected using eight
wavenumber resolution with a gain of 0.1. Sample time took
approximately 15 seconds. Two spectra were collected per
sample. Figure 2 shows images of the probe before insertion
into a cell culture sample and during spectral collection.
The sample spectra were loaded into the Thermo Scientific
TQ Analyst™ Pro Edition Software for chemometric analysis
using a partial least squares (PLS) method with a constant
pathlength. The spectra were analyzed in the first derivative
using a Norris smoothing filter. Two regions were used for the
analysis: 8,910 to 5,340 cm-1 and 4,830 to 4,340 cm-1 These
two regions collected information across a wide range of
data points while avoiding the totally attenuating water peak
centered around 5,100 cm-1. Figure 3 shows representative
raw spectra and the first derivative spectra of the samples.
Table 1. Concentration ranges of various components.
The solutions represent over 35 different protein concentrations that
also vary in concentrations of nutrient and waste components.
Figure 2. Transflectance probe used for data collection. Left panel shows
the design of the probe with the adjustable pathlength. Right panel shows
probe inserted into cell culture during data collection.
Component Range (g/L)
Protein 0.16–5.00
Glucose 7.98–8.12
Glutamine 0.28–0.58
Lactate 0.45–0.90
Ammonia 0.05–2.39
9
Learn more at thermofisher.com/brighteroutcomes
Results
PLS analysis of the protein concentrations in the various cell
culture samples revealed excellent predictive capabilities within
the range of materials tested. The 54 spectra used to develop
the PLS method are shown on a calibration plot (Figure 4) that
compares the calculated protein concentrations versus the
actual concentrations.
The calibration plot can be used to determine how well the
method predicts the actual protein concentrations in the
samples. The plot developed by the chemometric method
resulted in a correlation coefficient of 0.977. Root mean square
error of calibration (RMSEC) was 0.33 g/L and the Root mean
square error of prediction (RMSEP) calculated from the 20
validation samples was 0.31 g/L. Additionally, the Root mean
square error of cross validation (RMSECV) was 0.51 g/L. These
errors indicate that the protein concentration in the cell culture
samples can be predicted to 0.5 g/L or less. Approximately 1⁄3
of this error was attributed to the balance used to weigh the
protein material.
Conclusions
Measuring protein concentrations in living dynamic cell cultures
was successfully performed with the Antaris MX FT-NIR
Process Analyzer. Protein concentration is a critical parameter
in determining the success and quality of a cell culture in
manufacturing a viable end product. This NIR technique
successfully demonstrates the ability to measure and monitor
protein concentrations in real time at relevant concentrations.
The developed method shows excellent correlation with actual
protein concentrations between 0.16 and 5.0 g/L and with
errors of less than 0.5 g/L.
This application demonstrates the continued capability of the
Antaris MX FT-NIR Process Analyzer to be successfully used
in bioprocess environments where it can safely, accurately and
automatically monitor and control cell cultures. While previous
NIR studies have monitored cell culture conditions to promote
optimal protein production, few have actually monitored and
predicted protein concentrations. This feasibility study shows
the power of the Antaris MX FT-NIR Process Analyzer to
correctly predict target protein concentrations in a live and
dynamic cell culture.
Figure 3. Representative raw spectra showing the variability present
in the cell culture samples. Regions of analysis avoided the attenuated
water peak at 5,100 cm-1. Inset shows the first derivative spectra used
for the PLS chemometric method.
Figure 4. Calibration plot comparing the calculated protein concentrations
to the actual concentrations from the PLS method. Root mean square
errors are approximately 0.5 g/L or less. Blue circles (o) represent spectra
used to create the method, purple crosses (+) are spectra used to validate
the method.
Absorbance
Wavenumbers (cm-1)
10,000
-1
Calculated
Actual
9,000 8,000 7,000 6,000 5,000
5
Correlation coefficient: 0.977
RMSEC: 0.33 g/L
RMSEP: 0.31 g/L
RMSECV: 0.51 g/L
5
-1
1.4
1.3
1.2
1.1
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN51791 4/22 M
Protein aggregation identified through
UV-Visible absorption spectroscopy
Introduction
Misfolded or denatured proteins can associate in solution,1
forming insoluble aggregates (Figure 1). This process is often
irreversible, effectively removing useful proteins from solution
and making the detection of aggregates critical for further
downstream use of protein solutions. This is particularly
important when studying unstable or abnormal proteins, which
are more likely to form aggregates.2,3
The formation of protein aggregates in the body has also
been linked to several diseases, including Alzheimer’s and
Parkinson’s disease.1,4,5 In the pharmaceutical industry,
protein therapeutics, such as insulin,6 have been developed
to effectively treat a variety of diseases but have been difficult
to synthesize.7 The presence of aggregates in these products
can lead to lower product yields and can reduce the efficacy
of the final therapeutic.5, 8 For example, protein therapeutics
that undergo aggregation have been linked to lowered immune
responses and, in some cases, can even induce
allergic reactions.8
In the food industry, protein composition can have a large
impact on the palatability of the final product. Protein
aggregates can significantly change a food’s organoleptic
properties (e.g., taste, smell, etc.), as well as the digestibility of
the material.5
Size-exclusion chromatography has previously been used
to identify the presence of aggregates in a sample.9 This
characterization method is time-consuming, however, and
sample retrieval can be difficult. An alternative method for
the detection of protein aggregates uses UV-visible (UVVis)
absorption spectroscopy, a technique that measures a
sample’s light absorption. Aggregates in solution are known
to scatter incoming light, resulting in an apparent absorption
artifact across the entire spectrum.5, 10 This scattering artifact
does not represent the true absorption of the sample and
instead indicates that the solution contains aggregates large
enough to scatter the incoming light.
In this application note, UV-Vis absorption spectroscopy was
used to identify the presence of protein aggregates in aqueous
bovine gamma globulin (BGG) samples. Aggregation was
induced in these samples using heat or the addition of NaCl.
An integrating sphere was further used to measure the scatterfree
spectra of the samples. Scatter-correction methods were
used to determine the concentration of free, non-aggregated
BGG in solution.
Figure 1. Visualization of protein aggregation induced by heat or changes
in ionic strength.
Application note
11
Scattering appears as a raised baseline at longer wavelengths
but also influences the apparent absorption across the entire
spectrum and is highly dependent on the wavelength of
the incident light. This influence can be estimated using the
following equation:
Ascatter = log (I0 /Ino scatter) + Aoffset = log (I0 /I0-(f/λ4))) + Aoffset (1)
In the equation above, Ascatter is the scattering artifact/apparent
absorption due to scattering, I0 is the intensity of the light
before it interacts with the sample, Ino scatter is the intensity of the
light that reaches the detector (not scattered by the solution),
f is an arbitrary scaling factor, λ is wavelength in nanometers,
and Aoffset is an offset. This equation uses Beer’s law,
A = log (I0 /I) (2)
and the relationship between the wavelength of light and the
intensity of the scattered light, which is defined by the Rayleigh
equation,12
Iscatter ∝ (1/λ4) (3)
to determine an estimated intensity of the scattered light (Iscatter).
Assuming I0 is 1 and the intensity of the scattered light is less
than 1, Equation 2 includes only two parameters that must be
fit to determine the scattering contribution. The relationship
between scattering intensity and wavelength indicates that
there is a larger effect in the UV region (Figure 3a), where there
are prominent absorption features for proteins. This effect must
therefore be carefully corrected.
Figure 3b shows the data corrected using two different
methods. The first, referred to as “baseline correction,” involves
taking the average of the absorption reported in the spectral
region in which the sample should not absorb. The calculated
average is then subtracted from each point in the spectrum, as
described by:
Acorrected,λ=Ameasured,λ-Aaverage,(330-350 nm)) (4)
Experimental
Absorption spectra were collected using a Thermo Scientific™
Evolution™ One Plus UV-Vis Spectrophotometer. Samples
were held in a 10 mm quartz cuvette, and measurements were
collected between 220 and 400 nm. A stock 1.1 mg/mL BGG
solution was made by diluting standard Thermo Scientific
Pierce™ BGG Standard (2.0 mg/mL, Lot Number MH162604)
with phosphate buffer (PBS, 1×) to achieve the appropriate
concentration. A 5.3 M NaCl solution in phosphate buffer was
made by dissolving 1.5 g NaCl (Fisher Scientific) in 6.0 mL of
phosphate buffer. BGG samples were prepared as described
in Table 1.
BGG samples were heated using a single-cell Peltier accessory
at 75˚C for 30 or 60 minutes. Sample measurements were
collected using a Thermo Scientific™ Evolution™ ISA-220
Integrating Sphere Accessory in transmission geometry.
The collected data was reported using the Kubelka-Munk
transformation. An 8˚ wedge was used for optimized light
collection. After integrating sphere measurements were
completed, Sample 4 (Table 1) was filtered using a syringe filter.
The absorption spectrum of the filtrate was then measured
using the Evolution One Plus Spectrophotometer, without the
Evolution ISA-220 Accessory.
Results
The absorption spectrum of BGG (not aggregated), depicted in
Figure 2a (blue curve), is in agreement with literature values.11
Upon addition of NaCl, the entire spectrum appears to have
a higher absorbance, an artifact resulting from the presence
of larger particulates. Increased ionic strength of a protein
solution (due to high salt concentration) has been shown to
induce protein aggregation;4 this scattering signal can therefore
be attributed to the presence of small BGG aggregates.
Scattering is observed regardless of the visual (clear, nonturbid)
appearance of the solution (Figure 2c). This indicates
that, while it is difficult to confirm through visual observation
alone, aggregate scattering can be measured using UV-Vis
absorption, and the technique can be used as a test for
protein aggregation.
BGG sample Volume
of 1.1 mg/
mL BGG
(mL)
Volume
of PBS
(mL)
Volume
of 5.3 M
NaCl (mL)
Temperature
(°C)
NaCl
concentration
(M)
1 25.0 0.00 1.0 1.0 0.0
2 25.0 2.65 1.0 0.0 1.0
3 75.0 (60 min
incubation)
0.00 1.0 1.0 0.0
4 75.0 (30 min
incubation)
0.00 1.0 1.0 0.0
Table 1. BGG solution preparation.
Figure 2. Absorption spectra of 0.55 mg/mL BGG in PBS with (red) and
without (blue) 2.65 M NaCL. Images of a solution of BGG with (b) and
without (c) 2.65 M NaCl.
0.0
0.2
0.4
0.6
Wavelength (nm)
Absorbance
260 280 300 320 340
0 M NaCl
2.65 M NaCl
a. b.
c.
12
The concentration of free, non-aggregated BGG in the sample
was found to be 0.54 mg/mL using Beer’s law:
A = clε (6)
In the equation above, A is the measured absorbance, c is the
concentration, l is the path length (1 cm), and ε is the extinction
coefficient of the protein. Therefore, the concentration of
proteins that contribute to aggregation in this sample is
0.01 mg/mL.
For samples with a relatively low scatter contribution, the
mathematical scatter-correction method works well. However,
for samples that are visibly cloudy/turbid, this correction is
not ideal, as only a small portion of the light is allowed to
interact with the detector. To study a sample that is turbid,
a 0.55 mg/mL BGG sample was held at 75˚C for 60 minutes
using a single-cell Peltier accessory for the Evolution One Plus
Spectrophotometer, producing a cloudy solution (Figure 4b).
The resulting absorption spectrum is depicted in Figure 4a. The
scattering artifact present indicates that ~30% of the light is
transmitted through the sample at 310 nm, where BGG begins
to absorb, and even less is transmitted at shorter wavelengths.
This suggests there is a high concentration of aggregates
present in this heated sample.
As mentioned previously, the small amount of light reaching
the detector makes it difficult to mathematically correct for
scattering. Instead, an integrating sphere can be used—this
accessory allows for the collection of scattered light diffusely
reflected off the inner walls of the sphere. As the diffuse light
reflects many times, it can be uniformly collected, removing the
scattering artifact. To correct for the scatter shown in Figure
4a, a spectrum for the aggregated BGG sample (Table 1,
Sample 3) was collected using an Evolution ISA-220 Accessory.
Through the instrument software, the signal was reported using
Kubelka-Munk units, F(R), which is proportional to both the
absorption coefficient, k, and scattering coefficient, s, of
the material:
F(R)=k/s (7)
In this equation, Ameasuredλ is the absorption spectrum collected,
Aaverage, (330–350 nm) is the average of the absorption measured
between 330 and 350 nm, and Acorrectedλ is the corrected
absorption spectrum. The resulting spectrum is shown in
Figure 3b (green curve); the maximum absorption from the
band is still higher than that of the untreated BGG sample. This
does not match the expected result, as formation of aggregates
should remove free BGG from solution, leading to a lower
concentration and lower absorbance in the region of interest.
Consequently, the “baseline correction” does not properly
account for the scattering artifact present in the
collected spectrum.
The second method, called “scatter corrected”, fits the long
wavelength baseline to Equation 1, where f and A0 are fit
such that the resulting function matches the long wavelength
signal well. The scattering function described in Figure 3a was
fit using f = 6.1 x 108 and A0 = 0.006. The resulting scatter
function was then subtracted from the absorption spectrum, as
shown in the following equation,
A(corrected,λ = Ameasured,λ – Ascatter,λ (5)
where Ascatterλ is the calculated scatter estimate. This correction
results in the yellow spectrum in Figure 3b. Unlike the baseline
corrected spectrum (green curve, Figure 3b), the maximum
absorption of the scatter-corrected spectrum is below the
absorption maximum of the spectrum for untreated BGG,
as expected.
Figure 3. a) Estimated scattering calculated using Equation 1. b)
Absorption spectra of BGG with and without NaCL. Baseline-corrected
data is shown in green, calculated using Equation 4. Scatter-corrected
data is shown in yellow, calculated using Equation 5.
Figure 4. Absorption spectrum of 0.55 mg/mL BGG following a 60-minute
incubation at 75˚C.
0.0
0.5
1.0
1.5
2.0
Wavelength (nm)
Absorbance
260 280 300 320 340
0.55 mg/mL BGG:
60 min heated
a. b.
0
40
20
60
80
100x10-3
Wavelength (nm)
Absorbance
260 280 300 320 340
Estimated Scatter
0.0
0.2
0.4
0.6
Wavelength (nm)
Absorbance
260 280 300 320 340
0 M NaCl
2.65 M NaCl
Scatter Corrected
2.65 M NaCl
2.65 M NaC
Baseline Corrected
13
Using Equation 11, the concentration of non-aggregated
materials in the BGG sample was found to be 0.20 mg/mL,
implying 0.35 mg/mL of BGG contributed to the formation of
aggregates in this sample. To verify this equation, the BGG
sample containing aggregates was filtered using a syringe filter
and the absorption spectrum of the filtrate was collected using
a traditional cell holder. Using Beer’s law, the concentration
of the BGG filtrate was found to be 0.20 mg/mL, matching
the calculated concentration determined using the integrating
sphere. This further implies that BGG aggregates in solution do
not absorb an appreciable amount of light in the spectral region
of interest for this sample.
Figure 5 demonstrates the Kubelka-Munk spectrum of the
BGG solution shown in Figure 4; the scattering signal is largely
removed from the spectrum.
F(R) is not equivalent to absorbance, indicating Beer’s law
cannot be used to determine concentration from the collected
results. However, as F(R) is proportional to the absorption
coefficient, it is also proportional to the absorbance, A, and the
concentration, c, of the free proteins in solution:
F(R) ∝ A∝ c. (8)
To determine the concentration of aggregated and nonaggregated
proteins in solution using the Kubelka-Munk
formula, the fully non-aggregated sample (control) was
measured using the integrating sphere. The resulting Kubelka-
Munk spectrum collected is shown in Figure 6a (gray curve). A
second BGG sample heated to 75˚C for 30 minutes (Table 1,
Sample 4), which also resulted in a large scattering artifact, was
analyzed using the Evolution ISA-220 Accessory as well.
If the collected F(R) of the sample at a given wavelength is
assumed to be equivalent to the concentration of the proteins
in solution multiplied by some constant, b, that is shared
between all BGG samples, then we can construct a series
of equations:
Fcontrol (R) = ccontrolb (9)
Fsample (R)=csampleb (10)
csample = ccontrol * Fsample(R)/Fcontrol(R) (11)
The equations above can be used to relate the concentration of
non-aggregated BGG in the sample that was incubated at 75˚C
(csample) to the concentration of the non-aggregated BGG control
(ccontrol), the Kubelka-Munk signal of the sample (Fsample(R)), and
the control (Fcontrol(R)). For more complex samples, constructing
a standard curve with multiple control samples of differing
concentration would be a more effective analysis tool.
Figure 6. a) Kubelka-Munk spectra of 0.55 mg/mL BGG after a 30-minute
incubation at 75˚C (blue) and 0.55 mg/mL non-aggregated BGG (gray).
b) Absorption spectra of filtered 0.55 mg/mL BGG after a 30-minute
incubation at 75˚C (brown) and 0.55 mg/mL non-aggregated BGG (orange).
The incubated BGG sample was filtered using a Millipore Millex-GV
PVDFA filter.
Figure 5. Kubelka-Munk spectrum of 0.55 mg/mL BGG after a 60-minute
incubation at 75˚C.
Wavelength (nm)
0.0
0.2
0.4
0.6
1.0
0.8
F(R)
240 260 280 300 320 340
0.0
0.6
0.4
0.2
0.8
1.0
1.2
Wavelength (nm)
Absorbance
250 260 270 280 290 300 310
0.55 mg/mL BGG: Not Aggregated
0.55 mg/mL BGG: Aggregated
F(R)
0.0
0.5
1.0
1.5
2.0
Wavelength (nm)
250 260 270 290 300 310
0.55 mg/mL BGG: Not Aggregated
0.55 mg/mL BGG: Aggregated
14
Learn more at thermofisher.com/evolution
Conclusion
Protein aggregates in solution can quickly be detected using
the Evolution One Plus UV-Visible Spectrophotometer. For
samples with a low concentration of aggregate present, the
resulting scattering artifact can be corrected by estimating
the scattering contribution and subtracting that estimate from
the measured spectrum. For highly scattering solutions, the
Evolution ISA-220 Integrating Sphere Accessory works well
in removing the scattering artifact from the spectrum. The
concentration of free proteins in solution can then be solved for
the corresponding spectrum of a known standard or a series of
known standards.
References
1. Weids, A.J., Ibstedt, S., Tamás, M.J.; Grant, C.M., Sci. Rep. 2016, 6, 24554.
2. Chen, Z.; Huang, C.; Chennamsetty, N; Xu, X.; Li, Z.J., J. Chromatogr. A, 2016,
1460, 110 – 122.
3. Jubete, Y.; Maurizi, M.R., Gottesman, S., J. Biol. Chem., 1996, 271, 48,
30798-30803.
4. Da Vela, S.; Roosen-Runge, F.; Skoda, M.W.A.; Jacobs, R.M.J.; Seydel, T.;
Frielinghaus, H.; Sztucki, M.; Schweins, R.; Zhang, F.; Schreiber, F., J. Phys. Chem.
B 2017, 121, 23, 5759 – 5769.
5. Pignataro, M.F.; Herrera, M.G.; Dodero, V.I.; Molecules 2020, 25, 20, 4854.
6. Johnson, I.S., Science, 1983, 219, 4585, 632 – 637.
7. Mitragotri, S.; Burke, P.A.; Langer, R., Nat. Rev. Drug Discov., 2014, 13, 9,
655 – 672.
8. Lundahl, M.L.E.; Fogli, S.; Colavita, P.E.; Scanlan, E.M., RSC Chem. Biol., 2021, 2,
1004-1020.
9. Hawe, A.; Friess, W.; Sutter, M.; Jiskoot, W., Anal. Biochem. 2008, 38, 115 – 122.
10. Hall, D.; Zhao, R.; Dehlsen, I.; Bloomfield, N.; Williams, S.R.; Arisaka, F.; Goto, Y.;
Carver, J.A., Anal. Biochem., 2016, 489, 78 – 94.
11. Smith, E.L.; Coy, N., J. Biol. Chem., 1946, 164, 1, 367 – 370.
12. Yao, G.; Li, K.A.; Tong, S.Y., Anal. Chim. Acta, 1999, 398, 319 – 327.
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN53585_E 06/22M
Quantify protein and peptide preparations at 205 nm
NanoDrop One Spectrophotometer
Application note
Abstract
Life scientists can quantify peptide and protein samples on the Thermo Scientific™
NanoDrop™ One/OneC Microvolume UV-Vis Spectrophotometers using the
A205 preprogrammed direct absorbance application. The new A205 application
offers a choice of methods for peptides that contain Tryptophan and Tyrosine
residues in their sequence as well as peptides that completely lack aromatic amino
acids. The A205 application offers enhanced sensitivity for peptide quantification
in seconds from only 2 μL of sample.
Introduction
Researchers have always needed ways to quickly quantify various biomolecules
(e.g., protein and nucleic acid preparations) as a routine part of their workflows. This
information helps them make informed decisions before proceeding with downstream
experiments. There are many protein quantification methods to choose from
including gravimetric approaches, colorimetric assays, direct spectrophotometric
UV measurements (such as A280), and amino acid analysis. All of these methods
have their strengths and weaknesses. Direct spectrophotometric microvolume UV
measurements are a popular choice for researchers because they are simple to
perform, require no reagents or standards, and consume very little sample. The
NanoDrop One Spectrophotometer has preprogrammed applications (Figure 1) for
direct quantification of proteins using absorbance measurements at 280 nm and
205 nm. This application note specifically describes how to use the Protein A205
application to quantify protein samples.
A protein’s peptide backbone absorbs light in the deep UV region (190 nm-220 nm),
and this absorbance can be used for protein sample quantitation. The A205 protein
quantitation method has several advantages over the direct A280 protein method
such as lower protein-to-protein variability (because A205 extinction coefficients
are not based on amino acid composition) and higher sensitivity (because of the
high molar absorptivity proteins have at 205 nm). However, technical limitations
made it difficult to obtain these measurements in the past. Spectrometers’ stray
light performance, deep UV linearity, and protein buffers containing UV-absorbing
components have all added to the challenge of obtaining A205 data. The NanoDrop
One patented sample-rentention technology and low stray light performance have
simplified quantification of small amounts of protein by A205 methods.
In this application note, we discuss the three A205 measurement options included
in the NanoDrop One Protein A205 application and present performance data for
each option.
Figure 1. NanoDrop One Proteins Home screen
showing available preprogrammed applications
for protein quantitation.
16
A205 extinction coefficients for peptide and
protein measurements
The NanoDrop One Protein A205 application allows customers
to choose from three different options (Figure 2). The selected
option will automatically determine the extinction coefficient
that will be used to calculate the protein concentration based
on the sample absorbance at 205 nm.
• ε₂₀₅=31 method
• Scopes method²
• Other = custom method ε₂₀₅1mg/mL
Previous studies showed that most protein solutions at
1 mg/mL have extinction coefficients (ε₂₀₅1mg/mL) ranging
from 30 to 35². The ε₂₀₅ of 31 mL mg-1cm-1 is an extinction
coefficient often used for peptides lacking tryptophan and
tyrosine residues¹. The Scopes method gives a more accurate
ε₂₀₅, especially for proteins containing a significant amount
of tryptophan (Trp) and tyrosine (Tyr) residues. The increased
accuracy of this method takes into account the significant
absorbance at 205 nm contributed by the aromatic side chains
of Trp and Tyr. This method uses an A280/A205 ratio in its
equation to correct for Trp and Tyr side-chain absorbance³.
Recently, Anthis and Clore proposed the use of a
sequence-specific ε₂₀₅ calculation (e.g., custom/Other method),
which is suitable for a wide range of proteins and peptides¹.
This method is appropriate for pure preparations of proteins
or peptides whose amino acid sequences are known.
A205 performance on the NanoDrop One
Preparations of polymyxin, a cationic detergent antibiotic with
a peptide backbone, but no Trp or Tyr residues, were made in
0.01% Brij® 35 buffer and measured on the NanoDrop One and
the Thermo Scientific™ Evolution™ UV-Vis Spectrophotometers.
To ensure the validity of the measurements taken with the
Evolution Spectrophotometer instrument, the polymyxin
preparations were diluted in 0.01% Brij buffer to ensure that
the measurements taken were within the linear range of the
detector. For measurements on the NanoDrop One instrument
2 μL of sample were pipetted directly on the sample pedestal,
while a 10 mm quartz cuvette was used for measurements on
the Evolution Spectrophotometer. The polymyxin concentration
data obtained on both instruments (Table 1) were plotted
(Figure 3). Regression line shows that protein concentration
results from the NanoDrop One instrument are in good
agreement to the results obtained on a traditional high end
UV-Vis spectrophotometer using a cuvette.
Figure 2. NanoDrop One Protein A205 methods selection screen.
Table 1. Various preparations of Polymyxin were measured on the
NanoDrop One and Evolution Spectrophotometers. Five(5) replicates of
each solution were measured on the NanoDrop One instrument using the
205=31 application. Solutions with absorbance over 1.0A were diluted and
measured in triplicate on the Evolution Spectrophotometer instrument.
Figure 3. Polymyxin concentrations calculated with the Evolution
Spectrophotometer and NanoDrop One instruments were plotted.
Regression line shows that protein concentration measurements
on the NanoDrop One instrument are in good agreement to those
obtained on a traditional high end UV-Vis spectrophotometer.
Target
[Conc]
mg/mL
NanoDrop One Evolution
A205
Std.
Dev.
[Conc]
mg/mL
[Conc]
mg/mL
0 -0.01 0.04 -0.18 -0.02
5 0.11 0.01 3.60 5.05
10 0.27 0.01 8.84 10.53
15 0.44 0.02 14.08 17.09
50 1.68 0.01 54.14 55.32
100 3.39 0.01 109.44 108.48
200 6.64 0.03 214.16 222.50
0
50
100
150
200
250
NanoDrop One [Conc] μg/mL
A205 Performance using Polymyxin
Evolution Spectrophotometer
[Conc] μg/mL
0 50 100 150 200 250
R² = 0.99923
17
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2023 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN52774_EN 0423 M
Schedule a free trial at thermofisher.com/ndfreetrial
To assess the effect that the extinction coefficients used at
205 nm (i.e., Scopes and ε₂₀₅=31 methods) would have on the
result, we prepared dilutions of three different proteins with
varied amounts of aromatic residues: bovine serum albumin
(BSA, 3 Trp and 21 Tyr residues), lysozyme (6 Trp and 3 Tyr
residues) and polymyxin (no Trp, no Tyr). These preparations
were measured on the NanoDrop One instrument using the
ε₂₀₅=31 and Scopes methods (Table 2).
Conclusion
To assess NanoDrop One Spectrophotometer performance
at A205, we compared polymyxin concentration results
obtained with the NanoDrop One and the Evolution
benchtop Spectrophotometers, which have excellent stray
light performance. Table 1 shows that the NanoDrop One
instrument provided very consistent results between replicate
measurements at 205 nm with standard deviations below
0.04A. In addition, the results obtained with both instruments
were comparable (Figure 3). Comparison between the A205
methods (Scopes and ε₂₀₅=31 methods) offered in the
NanoDrop One A205 application shows that the number of
tryptophan and tyrosine residues has a large effect on the
calculated concentration (Table 2). This is because tryptophan
is the largest contributor to A280 absorbance, and the Scopes
method uses the A280/A205 ratio to correct for aromatic
side-chain absorbance at A205.
Our results show that A205 quantification using the ε₂₀₅=31
method gives comparable results when proteins have only a
few tryptophan residues.
One limitation of the A205 method is that many of protein
buffers commonly used have absorbance at 205 nm. Before
using this technique, we recommend checking the protein
buffer for any contribution to the absorbance at 205 nm.
References
1. Anthis, NJ and Clore, GM 2013. Sequence-specific determination of protein and
peptide concentrations by absorbance at 205 nm. Protein Science 22:851-858.
2. Goldfarb, AR, Saidel, LJ, Mosovich E 1951. The ultraviolet absorption spectra of
proteins. Journal of Biological Chemistry 193(1):397-404.
3. Scopes, RK 1974. Measurement of protein by spectrophotometry a 205 nm.
Analytical Biochemistry 59:277-282.
Protein
Preparation
# of Trp of Tyr
Trp Tyr A205 STDV
[Concentration]
e₂₀₅=31 (μg/mL)
[Concentration]
Scopes Method (μg/mL)
BSA 1 3 21 3.960 0.013 127.73 131.80
BSA 2 3 21 37.271 0.218 1202.30 1261.71
BSA 3 3 21 70.044 0.239 2259.48 2387.91
BSA 4 3 21 129.170 1.458 4166.77 4345.20
BSA 5 3 21 271.027 0.851 8742.81 9198.13
Lysozyme 1 6 3 29.069 0.169 937.71 795.95
Lysozyme 2 6 3 53.651 0.545 1730.68 1459.05
Lysozyme 3 6 3 102.713 0.668 3313.32 2814.79
Polymyxin 1 0 0 0.112 0.015 3.60 3.12
Polymyxin 2 0 0 0.274 0.014 8.84 10.12
Polymyxin 3 0 0 0.437 0.021 14.08 16.03
Polymyxin 4 0 0 1.678 0.014 54.14 60.99
Polymyxin 5 0 0 3.393 0.014 109.44 125.16
Polymyxin 6 0 0 6.639 0.034 214.16 244.87
Table 2. Comparison of different A205 methods for various protein and peptide preparations on the NanoDrop One Spectrophotometer.
Spectrophotometric Analysis of Ibuprofen
According to USP and EP Monographs
Performing pharmaceutical identification tests with
an Evolution UV-Visible Spectrophotometer
Introduction
Monographs outlined by the United States Pharmacopeia
(USP) and European Pharmacopoeia (EP) contain tests,
procedures, and acceptance criteria that help to ensure drug
ingredients and drug products conform to the published
requirements for strength, quality, and purity. These
monographs contain detailed instructions utilizing a variety
of analytical instrumentation for performing identification
tests, purity tests, and tests to limit the amount of undesirable
impurities. Although the general requirements governing the
performance of an analytical instrument used in a monograph
will be outlined in its own general chapter, additional
instrument requirements needed to perform a test may be
indicated in the individual monographs.
This document will highlight the utilization of UV-Visible
spectrophotometers to perform essential monograph tests.
As spectrophotometric tests are featured in hundreds of
monographs, UV-Visible spectrophotometers are essential
analytical instrumentation for every pharmaceutical quality
control laboratory. In this work, a Thermo Scientific™
Evolution™ Pro using Thermo Scientific™ Insight™ Pro Software
will be used to perform the USP and EP spectrophotometric
identification tests for ibuprofen highlighted in their
respective monographs.
Ibuprofen is an active pharmaceutical ingredient that
is used as a medication for treating pain, fever, and
inflammation. The Ibuprofen monographs published by
USP and EP contain a variety of identification tests utilizing
infrared absorption spectroscopy, ultraviolet visible
spectroscopy, melting point analysis, and chromatography
to help confirm the quality of ibuprofen samples.
Application note
Figure 1. Chemical Structure of Ibuprofen, C₁₃H₁₈O₂.
19
Experimental
EP Identification Test
The spectrophotometric ibuprofen identification test according
to the EP ibuprofen monograph requires measuring a test
sample and comparing the ratios of its absorbance values to
confirm they are within an acceptable range.
A 500 μg/mL ibuprofen solution was prepared by dissolving
50 mg in a solution of 0.1 M sodium hydroxide in a 100 mL
volumetric flask using the sodium hydroxide solution to fill to the
total volume. The sodium hydroxide solution was used as the
blank. The monograph required a spectrum of the ibuprofen
solution to be obtained with a wavelength range of 240 nm –
300 nm, a bandwidth of 1 nm, and a scan speed of less than
or equal to 50 nm/min. The Insight Pro software experimental
parameters used to obtain the spectrum are shown in Figure
2. The programed ratio equations for evaluating the results are
shown in Figure 3.
The spectrum of the ibuprofen test sample is shown in Figure 4.
Visual inspection of the spectrum shows absorption maxima at
264 nm and 272 nm and a shoulder at 258 nm which align with
the monograph requirement for the identification of ibuprofen.
Along with this spectrum, Insight Pro Software automatically
calculates the ratio values required to verify the identity of
ibuprofen using the programed equations from Figure 3. The
absorbance ratio of A₂₆₄/A₂₅₈ was 1.25 which was within
the 1.20 – 1.30 requirement highlighted in the ibuprofen
monograph. The Absorbance ratio of A₂₇₂/A₂₅₈ gave a result
of 1.00 which was also within the monograph requirement
of 1.00 – 1.10. These absorbance ratios along with the
visual inspection of the spectrum confirm the identity of
ibuprofen according to the Ultraviolet and visual absorption
spectrophotometry test in the EP ibuprofen monograph.
Figure 2. Experimental parameters of EP ibuprofen test. Figure 3. Ratio calculations for EP ibuprofen test.
Figure 4. EP ibuprofen test sample spectrum.
0.0
0.2
0.4
0.6
0.8
1.0
Wavelength (nm)
Absorbance
240 250 260 270 280 290 300
20
USP Identification Test
In the USP spectrophotometric identification test for ibuprofen,
a reference standard sample and a test sample are prepared
using identical procedures and the spectra are compared
to confirm the test sample exhibits absorption maxima and
minima only at the same wavelengths as those of the reference
sample. Additionally, the molar absorptivity at two different
wavelengths are calculated and compared between reference
standard and sample to ensure they do not differ within a
certain percentage.
A 250 μg/mL solution of an ibuprofen standard was prepared
by dissolving 25 mg in a solution of 0.1M sodium hydroxide in a
100 mL volumetric flask using the sodium hydroxide solution to
fill to the total volume. A test ibuprofen solution was prepared in
the same way as the standard. The sodium hydroxide solution
was used as the blank in the experiment.
Spectral measurements of the reference standard and sample
were obtained using a similar procedure as the EP Identification
Test in Figure 2 but scanning from 200 nm – 400 nm. The
spectral data should be similar to the results shown in Figure 4
with maxima and minima at the same wavelengths.
The absorbance values at 264 nm and 273 nm of both the
standard and test solution were measured according to the
USP guidelines using the Fixed method on the Evolution Pro
Spectrophotometer with the following parameters as shown in
Figure 5.
The absorptivity values at each wavelength for both the
standard and test sample were calculated using the Beer-
Lambert Law where ε is the molar absorptivity in L/mol•cm
units, A is the absorbance value, l is the pathlength in cm
and c is the concentration in mol/L:
The molar absorptivity is calculated using the absorbance
values and the weight of the samples. An example is shown
below where the molar absorptivity of the standard sample
at 264 nm was calculated as shown below:
The measured weight, absorbance at each wavelength, and
calculated absorptivity at each wavelength for the reference
standard and test sample are shown in Table 1.
The percent difference between the absorptivities of the
sample and standard at both 264 nm and 273 nm
were calculated using the following formula where the
Average ε is the average of the Test ε and Standard ε at
each respective wavelength:
Using this formula, we obtain a percent difference between
the Standard and Sample of 1.9% for the 264 nm absorptivity
and 1.4% for the 273 nm absorptivity. The USP test requires
the difference between the respective absorptivities at 264
nm and 273 nm to be less than or equal to 3.0%. Since the
test sample has a percent difference of less than 3.0% at both
wavelengths, it meets the identification requirement in the
ibuprofen USP monograph.
Figure 5. Experimental parameters of USP ibuprofen test.
Table 1. USP Measurement Data.
Weight
(mg) A₂₆₄ A₂₇₃ ε₂₆₄ ε₂₇₃
Standard 25.8 0.4732 0.3905 385.83 318.35
Sample 24.9 0.4644 0.3849 378.61 313.85
A
l x c
ε =
Sample ε – Standard ε
Average ε
Percent Difference = x 100
0.4732
1 cm x 0.00122649 mol/L
ε =
ε = 385.83 L/mol
21
Ordering information
Product Part number
Evolution One UV-Vis Spectrophotometer 840-341400
Evolution One Plus UV-Vis Spectrophotometer 840-341500
Evolution Pro UV-Vis Spectrophotometer 840-340200
Conclusion
Spectrophotometers are utilized in hundreds of
pharmaceutical monographs which make them essential
instrumentation for confirming the identity of drug ingredients
and drug products. The Thermo Scientific Evolution Series
Spectrophotometers are ideal for performing these tests due
to their versatility, ease of use, and superior performance.
In this document both the USP and EP identification tests
for each respective ibuprofen monograph was completed
with an Evolution Pro Spectrophotometer. The identity of
an ibuprofen test sample was confirmed according to the
USP requirements when compared to a standard ibuprofen
sample. The identity of an ibuprofen test sample was also
confirmed according to the EP requirements through visual
inspection and by comparing absorbance ratios.
References
1. United States Pharmacopeia and National Formulary (USP 43-NF 38),
Monographs, Ibuprofen
2. European Pharmacopeia (EP 9.6), Monographs, Ibuprofen
Learn more at thermofisher.com/uv-vis
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN53349_E 07/22M
Observation of gold nanoshell plasmon
resonance shifts after bioconjugation
Using the NanoDrop One Microvolume UV-Vis Spectrophotometer
Application note
Authors
Kejian Li1, Megan N. Dang1, Alexis B.
Duffy1 and Emily S. Day1,2,3
1 University of Delaware, Dept. of Biomedical
Engineering, Newark, DE, USA
2 University of Delaware, Department
of Materials Science and Engineering,
Newark, DE, USA
3 Helen F. Graham Cancer Center and
Research Institute, Newark, DE, USA
RNA interference (RNAi)-based therapy has shown great
potential in improving the study and treatment of diseases
whose genetic underpinnings are known. However,
challenges such as susceptibility to nuclease degradation,
low cellular uptake, or rapid clearance from circulation
impede the successful preclinical and clinical application
of RNAi therapeutics.1 To overcome these limitations, small
interfering RNAs (siRNAs) or microRNAs (miRNAs) can be
conjugated to nanoparticles (NPs), such as nanoshells (NS),
to improve their stability, cellular uptake, and blood
circulation time, thus resulting in increased effectiveness.2, 3, 4
Prior to using RNA-NP conjugates in therapeutic applications,
it is critical to confirm successful RNA conjugation to the NP.
One common method to confirm molecule loading onto goldbased
NPs involves evaluating the surface plasmon resonance
(SPR) spectra of the NPs before and after functionalization;
successful RNA attachment will typically cause a slight
red-shift in the peak SPR wavelength. Traditionally, UV-Vis
spectrophotometers are used to analyze the optical properties
of gold-based NPs. For example, the peak absorbance can be
utilized to determine NP concentration via Beer’s Law and to
evaluate changes due to any surface modification. However,
conventional cuvette-based UV-Vis spectrophotometers
have limited linear range due to the use of a standard fixed
pathlength (10 mm) cuvette, and they often require relatively
large sample volumes (ranging from 0.5 mL to 3 mL). This is
not ideal for conserving precious samples such as NPs coated
with expensive RNA molecules. Furthermore, the need to
dilute samples to fit the operating range of the instrument is
time-consuming and increases the likelihood for inaccurate
measurements. Alternative measurement techniques that
require less volume and allow analysis of concentrated samples
without dilution would be ideal.
Recent work has shown that the Thermo Scientific™
NanoDrop™ One Microvolume UV-Vis Spectrophotometer
can be used to accurately measure highly concentrated NP
samples without dilution, owing to its surface tension system
and auto-ranging pathlength technique.5, 6 For example, 150 nm
diameter NS can be measured at concentrations up to 100 pM
with high reproducibility.5 In this application note, the use of the
Nanodrop One instrument to observe shifts in the SPR of NS
after conjugation to thiol-modified siRNA duplexes and
23
methoxy-poly(ethylene glycol)-thiol (mPEG-SH; a passivating
agent) was investigated. The results indicate that the Nanodrop
One instrument can serve as a microvolume alternative to
traditional cuvette-based spectrophotometers for qualitatively
confirming RNA and PEG loading on gold-based NPs via
plasmon resonance shifts.
Experimental procedures
NS were synthesized by published protocols via the Oldenburg
method.7 First, 3-5 nm diameter gold colloid was made by
the Duff method8 from hydrogen tetrachloroaurate (III) hydrate
(HAuCl4) (VWR), tetrakis(hydroxymethyl)phosphonium chloride
(VWR), and 1 N sodium hydroxide (Fisher Scientific). The
gold colloid was then combined with 120 nm diameter silica
spheres functionalized with 3-aminopropyltriethoxysilane
(Nanocomposix) and 1 M sodium chloride (NaCl) and
rocked for 3-4 days at room temperature to create “seed”
nanoparticles. The seed was purified twice via centrifugation
at 3000 rpm for 30 minutes each and resuspended in Milli-Q®
water (Sigma) to an optical density at 530 nm (OD530nm)
of 0.1, as determined using a cuvette-based UV-Vis
spectrophotometer. The diluted seed was mixed with additional
HAuCl4 diluted in potassium chloride followed by addition
of a small volume of 37% formaldehyde (VWR). The mixed
solution was rapidly agitated to form complete gold shells and
purified twice via centrifugation at 500 g for 15 minutes each.
Additionally, NS were treated with 0.1% diethyl pyrocarbonate
(DEPC) (Sigma) for 3 days rocking at 37°C to render the NS
RNase-free. All materials described were purchased or treated
with DEPC to be RNase-free prior to use.
siRNA oligonucleotides were purchased as single strands
from IDT DNA, with sequences listed in Table 1. Thiolated
sense strands were mixed with complementary non-thiolated
antisense strands in equimolar amounts, boiled at 95° C for
5 min in a thermomixer, and then slowly cooled to 37° C over
1 hour to facilitate siRNA duplexing. RNase-free NS were
diluted to OD800nm = 1.5 in Milli-Q water (as measured on
a cuvette-based spectrophotometer). Next, 10% Tween-20
and 5 M NaCl were added to final concentrations of 0.2%
and 12 mM, respectively, and the NS incubated for 5 min
at room temperature. Then, siRNA duplexes were added to
a final concentration of 200 nM, and the solution was bath
sonicated and rocked at 4° C for 3 hours. NaCl was then
added incrementally to a final concentration of 400 mM prior to
rocking overnight at 4° C. The following day, 5 kDa mPEG-SH
was diluted in Milli-Q water to 1 mM and added to NS to a final
concentration of 10 μM. After rocking for 4 hours at 4° C, the NS
solution was purified via centrifugation at 500 g for 5 minutes
3 times, resuspended in RNase-Free 1X phosphate buffered
saline (PBS) with 100 X less volume of the starting NS, and
stored at 4° C until use.
For conventional spectrophotometry, bare NS and siRNA-NS
(diluted 100-fold in water) were placed in 1-cm pathlength
disposable cuvettes and analyzed on a reference UV-Vis
spectrophotometer from 1,100 nm to 400 nm. The NS
concentrations were calculated from Beer’s Law using
the peak extinction (OD at ~800 nm) as determined by the
spectrophotometer and the theoretical extinction coefficient
of NS with 120 nm diameter silica cores and 15 nm thick
gold shells. This revealed the initial bare NS and siRNA-NS
had a concentration of 6.9 pM and 150 pM, respectively.
To prepare samples for measurement with the NanoDrop
One Spectrophotometer, the bare NS were concentrated by
centrifugation at 500 g for 15 minutes, followed by removal of
the supernatant and dilution in water to 100 pM. The siRNANS
were directly diluted in water to 100 pM. The 100 pM bare
NS and siRNA-NS solutions were measured on a NanoDrop
One Spectrophotometer from 850 nm to 190 nm by pipetting
2 μL aliquots directly onto the sample pedestal. Between
measurements, the NanoDrop One instrument sample pedestal
was cleaned using a lint-free lab wipe. The auto pathlength
option was turned on in the NanoDrop One Spectrophotometer
software for each measurement.
Name Sequence
siRNA sense GCU GAU AUU GAC GGG CAG UAU /
iSpPC//iSpPC//3ThioMC3-D/
siRNA antisense AUA CUG CCC GUC AAU AUC AGC
Table 1: siRNA sense and antisense RNA sequences used in this work,
denoted 5’ to 3’. iSpPC is a photo-cleavable 10-atom spacer molecule,
while 3ThioMC3-D is a thiol modification that facilitates attachment to
gold NS.
Results
The absorption spectra of 150 nm NS, before (bare NS) and
after (siRNA-NS) conjugation to thiolated siRNA and mPEGSH
at concentrations of 100 pM are shown in Figure 1. These
spectra reveal the bare NS and siRNA-NS have a peak
plasmon resonance at ~795 nm and ~804 nm, respectively,
which is consistent with the spectra obtained using a
reference spectrophotometer. The slightly red-shifted peak
post functionalization, which maintains the overall shape and
intensity of the spectra, provides evidence of successful siRNA
and mPEG-SH conjugation. This was corroborated by dynamic
light scattering and zeta potential measurements, as well as by
siRNA loading quantification via OliGreen assay.2, 4, 9 Notably, the
spectra produced by the NanoDrop One Spectrophotometer
were highly accurate and reproducible. Very little sample
volume (2 μL) was used in the measurement, and no dilution
was required for the analysis of highly concentrated samples
(100 pM).
24
Conclusions
This study demonstrates that the NanoDrop One
Spectrophotometer can be used as a simple and reliable
method to evaluate the surface modification of NS. The
NanoDrop One Spectrophotometer can produce highly reliable
results due to its built-in Thermo Scientific™ Acclaro™ Sample
Intelligence Technology, which identifies common contaminants
or other anomalies that may impact measurement accuracy.
Additionally, the NanoDrop One Spectrophotometer allows
the users to measure highly concentrated samples in 1–2 μL
without dilution and produce full spectral data in seconds
compared to a traditional cuvette-based spectrophotometer.
These advantages save valuable time and money and help
determine the quality and quantity of the sample before
use in downstream applications. The ease of operation and
small sample size requirement make the NanoDrop One
Spectrophotometer an ideal and valuable instrument to
characterize the properties of surface-modified NPs.
Figure 1: (A) UV-Vis spectra for Bare NS and siRNA-NS at concentrations of 100 pM, as measured on the NanoDrop One Spectrophotometer. n=3. (B)
Zoom in of the UV-Vis spectra peak for Bare NS and siRNA-NS (750 nm to 850 nm).
References
1. Wang, T.; Shigdar, S.; Shamaileh, H. A.; Gantier, M. P.; Yin, W.; Xiang, D.; Wang, L.;
Zhou, S. F.; Hou, Y.; Wang, P.; et al. Challenges and Opportunities for siRNA-Based
Cancer Treatment. Cancer Lett. 2017, 387, 77−83.
2. Riley, R. S.; Dang, M. N.; Billingsley, M. M.; Abraham, B.; Gundlach, L.; Day, E. S.
Evaluating the Mechanisms of Light-Triggered SiRNA Release from Nanoshells for
Temporal Control Over Gene Regulation. Nano Lett. 2018, 18, 3565−3570.
3. Artiga, Á.; Serrano-Sevilla, I.; De Matteis, L.; Mitchell, S. G.; De La Fuente, J. M.
Current status and future perspectives of gold nanoparticle vectors for siRNA
delivery. J. Mater. Chem. B 2019, 7, 876-896.
4. Dang, M. N.; Gomez Casas, C; Day, E. S. Photoresponsive miR-34a/Nanoshell
Conjugates Enable Light-Triggered Gene Regulation to Impair the Function of
Triple-Negative Breast Cancer Cells. Nano Lett. 2021, 21(1), 68-76.
5. Li, K.; Kapadia, C. H.; Dang, M. N.; Day, E. S. Quantification of
gold nanoshells using the NanoDrop One Microvolume UV-Vis
Spectrophotometer. https://www.thermofisher.com/document-connect/
document-connect.html?url=https%3A%2F%2Fassets.thermofisher.
com%2FTFS-Assets%2FMSD%2FApplication-Notes%2Fquantification-goldnanoshells-
nanodrop-one-uv-vis-spectrophotometer-an53464.pdf
6. Kapadia, C. H.; Melamed, J. R.; Day, E. S. Quantification of gold nanoparticles
using the NanoDrop One Microvolume UV-Vis Spectrophotometer. http://assets.
thermofisher.com/TFS-Assets/MSD/Application-Notes/AN53100-quantificationgold-
nanoparticle.pdf
7. Oldenburg, S. J.; Averitt, R. D.; Westcott, S. L.; Halas, N. J. Nanoengineering of
optical resonances. Chem. Phys. Lett. 1998, 288 (2−4), 243−247.
8. Duff, D. G.; Baiker, A.; Edwards, P. P. A new hydrosol of gold clusters. 1. formation
and particle size variation. Langmuir. 1993, 9, 2301−2309.
9. Melamed, J. R.; Riley, R. S.; Valcourt, D. M.; Billingsley, M. M.; Kreuzberger, N. L.;
Day, E. S. Chapter 1: Quantification of siRNA duplexes Bound to Gold Nanoparticle
Surfaces. In Biomedical Nanotechnology Humana Press: New York, NY, USA, 2017.
Learn more at thermofisher.com/nanodrop
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN53609_E 07/22M
Color analysis for pharmaceutical products using
UV-Visible absorption techniques
Introduction
The collection of reflected light by our eyes leads to the
perception of an object’s color, specifically light in the visible
range of the electromagnetic spectrum (~400 nm – 700 nm).
As our eyes are sensitive to variations in color and brightness,1
small changes in the color of an object can be easily observed.
In pharmaceutical manufacturing, the color of a drug product
is important to analyze for QA/QC purposes. Not only is it
necessary to minimize batch-to-batch variations for aesthetic
purposes, but changes to the color of a product can have
implications for the quality of the products. Specifically, variations
from the anticipated color could indicate impurities are present
in the product or that the material has degraded.2–4 This is
particularly important for materials which are easily decomposed,
including light, moisture, and oxygen/air-sensitive substances.5
Figure 1: Diagram of how the color of an object is perceived.
Qualitatively, a comparison of the color of a finished drug product
with an accepted standard can be used to ensure the material’s
color matches. However, inherently this methodology will introduce
Application note
person-to-person variations.6 Additionally, environmental effects,
such as the light source or the presence of shadows, can influence
the perceived color. As the color of a material comes from the
reflected visible light, spectroscopic measurements of a material in
the visible spectral range can be used to provide a more rigorous
and quantitative method for assessing color. Consequently, a
UV-Visible spectrophotometer can be used to measure either the
percent of light transmitted (%T) or reflected (%R) across the
visible spectrum for this purpose. As either of these measurement
geometries can be used, this analysis can be applied to both liquid
and solid products.
The American Society for Testing and Materials (ASTM),7 as well
as USP <1061>,8 have detailed descriptions of the mathematics
that can be used to assign the sample’s color a coordinate in a
graphical representation of color, also referred to as a color space.
The tristimulus values, calculated through the equations 1 – 3,
are the basis of most other color spaces developed by the
Comission Internationale de l’Eclairage (CIE).9 These formulas
include the measured reflectance (R(λ)), the spectral power of
an illuminant (S(λ)), a color matching function (x(λ),y(λ),z(λ)), and
the normalization factor (k).
(1)
(2)
(3)
26
As described previously, the color of an object is highly
dependent on environmental factors, such as light source
and the field of view of the object. For example, the intensity
of the light across the visible spectrum can be very different
for various light sources and can lead to differences in how
the color is observed. In the tristimulus equations, this factor
is taken into account through the inclusion of the spectral
power of the illuminant, S(λ). A standardized intensity spectrum
describing the spectral illuminant power as a function of
wavelength was developed to describe a typical intensity
spectrum for common illuminants (e.g., room lights, daylight),
and is included in equations 1 – 3. Additionally, the observer
angle, which defines the field of view of the material, can also
alter the perceived color and is also accounted for in tristimulus
equations through the color-matching functions.
The tristimulus values can condense the measured visible
spectrum of a sample down to a single coordinate, however,
the coordinate space is not uniform.9 The lack of uniformity
can lead to issues gauging the difference between the
color of a sample and the color of a reference standard. In
pharmaceutical applications, specifically in QA/QC functions,
the ability to compare the sample to an accepted standard, as
well as establish acceptance criteria, is critical. Consequently,
a uniform color space must be used instead. CIE developed
a set of mathematical functions which convert the calculated
tristimulus coordinates into a uniform, cylindrical (CIE L*a*b*)
or spherical (CIE L*C*h*) coordinate system (Figure 2), which is
built on opposing color theory.
Figure 2: CIE L*a*b* and CIE L*C*h* coordinates
Coordinates for the more commonly used CIE L*a*b* color space
are generated through the following mathematical functions,7, 8
where X, Y, and Z are the calculated tristimulus values and Xn,
Yn
, and Zn are the tristimulus values of a perfectly reflecting
white diffuser. Here L* describes how light (100) or dark (0)
the materials are, a* represents how red (positive) or green
(negative) the sample is, and b* demonstrates how yellow
(positive) or blue (negative). As this transformation results in a
more uniform color space, a better representation of the color
difference (ΔE*) between the sample and a standard can be
developed. The color difference formula (eq 7) describes how a
color difference is mathematically determined,
where L*sam, a*sam, and b*sam represent the CIE L*a*b* values for
the sample and L*std, a*std, and b*std represent the CIE L*a*b*
values for the standard.8 As a rule of thumb, two colors are
considered to be indistinguishable from one another by eye if
the color difference between the two substances is less than 3.
The CIE L*C*h* color space uses the same coordinate system
as the CIE Lab system, except it reports the chroma (Cab*) and
hue (hab*) of the substance in place of a* and b*. Chroma is
calculated through equation 8,
and describes how colorful a substance is wherein a small
Cab* represents a more pale or muted color, while a large Cab*
describes a substance with a very vibrant color. Hue describes
the color of the object and is calculated through equation 9.
Color analysis can be a quick and useful tool for assessing the
overall quality of a given product prior to further downstream
processing. Through UV-Visible absorption spectroscopy,
the analysis can be made more rigorous, allowing for a more
accurate measurement of color. Herein, we describe how
color analysis can be applied to both solid and liquid samples
using the Thermo Scientific™ Evolution™ Spectrophotometers
and Thermo Scientific™ Insight™ Pro Software. Furthermore,
descriptions of the USP requirements for color analysis of samples
are explained in relation to the instrumental analysis method.
(4)
(5)
(6)
(7)
(8)
(9)
Thermo Scientific Evolution Spectrophotometers
27
Experimental
Materials
USP color-matching solutions were prepared based on
descriptions in USP’s chapter <631>,10 which includes methods
to analyze and report the color of solution phase samples.
Briefly, three stock solutions were generated:
• 0.27 M CoCl2 • 6H2O (red solution)
• 0.17 M FeCl3 • 5H2O (yellow solution)
• 0.23 M CuSO4 • 5H2O (blue solution)
These solutions were mixed in different proportions to prepare
the color-matching solutions A – T as defined in USP <631>
(see Table 1).10
Table 1: Proportions of stock color solutions used to prepare color
matching solutions A – T based on USP <631>.10
Color
Matching
Solution
Volume
CoCl2 •
6H2O (mL)
Volume
FeCl3 •
5H2O (mL)
Volume
CuSO4 •
5H2O (mL)
Volume
H2O (mL)
A 0.1 0.4 0.1 4.4
B 0.3 0.9 0.3 3.5
C 0.1 0.6 0.1 4.2
D 0.3 0.6 0.4 3.7
E 0.4 1.2 0.3 3.1
F 0. 1.2 0.0 3.5
G 0.5 1.2 0.2 3.1
H 0.2 1.5 0.0 3.3
I 0.4 2.2 0.1 2.3
J 0.4 3.5 0.1 1.0
K 0.5 4.5 0.0 0.0
L 0.8 3.8 0.1 0.3
M 0.1 2.0 0.1 2.8
N 0.0 4.9 0.1 0.0
O 0.1 4.8 0.1 0.0
P 0.2 0.4 0.1 4.3
Q 0.2 0.3 0.1 4.4
R 0.3 0.4 0.2 4.1
S 0.2 0.1 0.0 4.7
T 0.5 0.5 0.4 3.6
For comparison against a more realistic example, two different
cough syrups were analyzed. One sample was labeled
“Daytime” and the other “Night-time.” Additionally, a set of four
antacid tablets of different colors were analyzed herein. The
tablets were crushed into powders using a mortar and pestle.
Instrument parameters
UV-Visible measurements described herein were collected
using an Evolution One Plus Spectrophotometer. For all
samples, spectral measurements spanning 280 nm and
780 nm were collected using a 1.0 nm spectral bandwidth and
2 nm data interval.
The USP color-matching solutions were measured in transmission
geometry and reported as % Transmission (%T), and the cough
syrup samples were reported in absorption units. For both sample
sets, deionized water was used to establish a 100% transmission
baseline as the blank solution. All USP matching solutions were
measured using a plastic 10 mm cuvette, while the cough syrup
samples were measured in a 10 mm and 1 mm quartz cuvette.
The antacid samples were measured in reflection geometry
using an integrating sphere accessory (ISA-220) with a powder
cell holder. A white Spectrlon© disk was used to establish a
100% reflection baseline as the blank. The resulting data was
reported as % Reflectance (%R).
Color analysis parameters
For all samples described herein, the CIE L*a*b* color values were
calculated using Insight Pro Software. The D65 illuminant with a
10˚ observer angle was chosen to reflect the color of all samples.
Color difference measurements were also performed through this
software feature. All calculations performed correspond to the
descriptions outlined in USP <1061>8 and ASTM-E308.7
Results and discussion
Analysis of liquid samples—color matching solutions
According to USP <631>, color-matching solutions are to
be used as a comparison point against the produced liquid
product to ensure the product matches the expected color.
As many liquid-based pharmaceutical products are yellow in
hue, the USP monograph includes a procedure for making a
set of standard solutions of varying yellow (Figure 3d).10 EP has
a different procedure outlined for making color standards and
includes a wider range of colors, including brown, green and
blue, among others.11
As shown in Figure 3d, some samples appear by eye to be similar
and almost indistinguishable in color. However, as the purpose
of these standards is to serve as different matching solutions,
the variations in the color may be slight and difficult to compare
without instrumental methods like UV-Visible color analysis. To
demonstrate this concept, the percent transmittance of each
matching solution was collected and are shown in Figures 3a – 3c.
a b c d
Figure 3: Absorption spectra of USP color matching solutions (a) A – G, (b) H – N, (c) O – T. (d) An image of the USP color matching solutions.
From these spectra, it is clear there are small differences in the
transmittance, and consequently absorption, of each matching
solution; however, color difference calculations were needed to
rigorously compare the colors. As described previously, the CIE
L*a*b* values were calculated using the Insight Pro Software.
A select set of color-matching standards were chosen for
comparison and are included in Table 2 as these standards
(Soln. A and B, Soln. J and K, and Soln. Q and R) appear similar
enough to each other in color that they are difficult to tell apart.
Table 2: CIE Lab and color difference values for select USP color matching
solutions (A, B, J, K, Q, R). Color difference calculations were carried out
for samples which appear similar by eye.
Solution L* a* b* ΔE*
A 87.5 0.5 28.5
9.7
B 83.3 2.4 37.0
J 69.1 12.0 80.0
12.5
K 73.9 12.5 91.5
Q 85.1 2.6 28.3
5.2
R 88.1 2.5 24.0
The color difference values calculated between matching
solutions A and B, J and K, and Q and R are relatively low;
however, a numerical limit is required to put these difference
values into context. In the pharmaceutical industry, different
formulations may require different methods of comparison
against a color-matching standard. For example, one product
may need to have no discernable color (achromatic), while
another must meet a minimum color value. Consequently,
USP has developed a set of criteria which can be used to set
acceptable limits for the calculated color difference from a
standard (Table 3).
There are four main test limits which can be used depending
on the color expectations for the analyzed product. Each test
defines a limit to an acceptable color difference between the
material and a given standard. For a sample which should have
no color, the first test in Table 3 (colorless/achromatic) defines
the necessary color difference limit as ΔE* < 1, where the colormatching
standard is purified water.
For samples where the sample has an expected color, there
are a few different options for analysis. If the color must
match a given standard color exactly, the second test in
Table 3 (Indiscernible from Standard) is required. Here, the
color difference between the product and the color matching
standard is used and must be less than 3. As mentioned
previously, this defines the color difference that is discernable
by the human eye.10 The last two analyses define maximum
and minimum color limits. Here, a sample can either be more
or less colorful than a given standard. USP defines Δhab*, the
difference in hue between the sample and matching standard
chosen must be less than 15. When setting the maximum or
minimum color limit, instead of comparing the color difference
against a number, two different analyses are required: one
where the color of the standard is compared to the color of
pure water (ΔEstd*) and one where the color of the product is
compared against pure water (ΔE*).
As the color difference values shown in Table 2 are intended
to determine how similar the color of the two solutions are to
one another, this analysis would follow the “Indiscernible from
Standard” test. The passing criteria would require a calculated
color difference of less than 3. For each set of standards, the
color difference exceeds this limit, indicating they fail this test
and are distinguishable from one another. This result highlights
how small differences in color can be analyzed through the
instrumental method, where it is difficult to perceive visually.
Analysis of liquid samples—cough syrup
The color-matching standards are ideal solutions with
optimized component concentrations to produce a measurable
spectrum in a standard 10 mm cuvette. Real samples may not
be manufactured to produce UV-Visible absorption spectra that
can be easily measured under these conditions. For example,
Figure 4a includes the absorption spectra of a “Daytime” and
“Night-time” cough syrup measured in a 10 mm cuvette. By
eye, the “Daytime” syrup appears orange while the “Night-time”
syrup appears red/purple.
As shown, both samples absorb greatly at wavelengths shorter
than 550 nm (A > 3). In UV-Visible absorption measurements,
it is good practice not to use highly absorptive samples for
calculations or quantification, as very little light is allowed to
pass through the sample and be detected by the system. For
example, an absorption of 3 indicates 99.9% of the incident
light is absorbed by the sample, leaving 0.1% of the light
collected by the detector. Consequently, the absorption spectra
in Figure 4a are not ideal for color analysis and result in the
values described in Table 4.
Figure 4: Absorption spectra of "Daytime" and "Night-time" cough syrup
collected using a (a) 10 mm and (b) 1 mm quartz cuvette. (c) An image of the
"Daytime" and "Night-time" cought syrup).
Daytime Night-time
a
b
c
Table 3: Passing criteria for color difference tests from USP <631>.10 For the
maximum and minimum color difference measurements, ΔEstd* refers to the
color difference between a matching standard and purified water while ΔE*
refers to the color difference of the sample against purified water.
Test Color
Standard
Passing
Criteria
1 Colorless
(Achromatic)
Purified Water ΔE* < 1
2 Indiscernible
from Standard
Color Matching
Solution
ΔE* < 3
3 Maximum
Color
Purified Water ΔE* < ΔEstd*
4 Minimum Color Purified Water ΔE* > ΔEstd*
Table 4: CIE L*a*b* values for "Daytime" and "Night-time" cough syrup samples. Spectra were measured using a 10 mm and 1 mm path length.
L* a* b*
Sample 10 mm cuvette 1 mm cuvette 10 mm cuvette 1 mm cuvette 10 mm cuvette 1 mm cuvette
Daytime 67.5 79.7 62.0 40.3 116.2 86.0
Night-time 40.2 62.3 68.5 72.2 69.2 27.6
To avoid issues for highly absorptive samples, instead a short
pathlength cuvette can be used as absorption is directly
proportional to pathlength according to Beer’s law (eq. 10),
where A is the collected absorbance, c is the concentration of
the analyte, l is the path length, and ε is the molar absorptivity
of the analyte. Changing the path length also circumvents the
need to dilute the sample, avoiding some waste of the material.
Herein, both cough syrup samples were measured using a
1 mm cuvette, resulting in the absorption spectra in Figure 4b.
Compared to the spectra shown in Figure 4c, the spectra
collected show much more clearly the absorption features
present in the sample. Included in Table 4 are the resulting
color values based on the spectra collected with a shorter path
length. These reported values are very different from the values
calculated using the spectra collected with a longer path length.
It is important to note that changing the path length not only
changed the perceived lightness/darkness of the sample (L*), but
also how red/green (a*) and how blue/yellow (b*) the samples
appear. This observation further illustrates the importance of
measuring highly absorptive samples in a shorter path length
to avoid significant deviations in the calculated color values. As
good practice, quantification should only be performed when the
highest peak absorption in the spectral region of interest is 1 A
or lower. Given the calculated color values will be sensitive to the
chosen path length, it is important any standard used for color
difference calculations be measured using the same path length.
Analysis of solid samples
USP <631> specifically refers to color analysis procedures
for liquids; however, color analysis can be performed using
solid samples as well, according to USP <1061>.8,11 For
pharmaceutical analysis, the color of a solid drug product
can also have implications on the quality of the material,3–6 as
described previously; however, it can also be used to indicate
the dosage of a given product as well as comply with a
company’s branding or marketing needs.6 For solid materials,
measurements in reflection geometry are appropriate as it is
difficult to pass light through a solid material without scattering
effects. As described in equations 1 – 3, the tristimulus values,
and therefore the CIE L*a*b* values, can be calculated using
reflectance data, allowing for color analysis of solid samples.
a
b
Figure 5 – (a) An image of the four antacid tablets measured. (b)
Reflectance spectra of four antacid tablets (blue—Tablet A, dark
green—Tablet B, brown—Tablet C, and light green—
Tablet D) and a white
reflectance standard (Spectralon).
Figure 5b includes the percent reflectance spectra (%R) of
four antacid tablets (Fig. 3a) of varying colors. By eye, Tablets
A – D appear white, yellow, orange, and red, respectively. The
calculated CIE L*a*b* values for each sample are included in
Table 5, along with the color values for a white Spectralon®
reference material (99% reflectance). Color difference
calculations were then performed to determine how different
each antacid tablet was from the white reference material.
Tablets B – D resulted in very high color differences (between
23 and 27) with respect to the reference standard, as
anticipated as these samples are visually very different from
the white standard. Tablet A, which appears white by eye, is
closer in color to the reference, with a color difference of 8.7
compared with the color difference of the other three tablets,
however as the calculated color difference is greater than 3, it
is distinguishable from the reference standard and would fail a
color matching test.
(10)
Table 5: Calculated CIE L*a*b* color values and color difference values for
antacid tablets. Color Difference Calculations were carried out using the
color values for the Spectralon® reference as the standard.
Sample L* a* b* ΔE*
Spectralon®
Reference
100.0 0.0 0.0 —
Tablet A 92.8 0.3 3.4 7.92
Tablet B 92.8 -5.8 21.7 23.6
Tablet C 88.1 13.7 17.0 24.9
Tablet D 82.5 19.3 8.7 27.5
30
Learn more at thermofisher.com/evolution
Conclusion
Color analysis can be an effective and quick method for QA/QC
in pharmaceutical manufacturing. As shown in the experiments
described herein, color analysis can be performed using the
Evolution UV-Visible Spectrophotometers to carefully determine a
material’s color without person-to-person variations, allowing for
a quantitative analysis of a produced pharmaceutical. Additionally,
these measurements demonstrate the ability to analyze both liquid
and solid samples following USP color analysis procedures.
References
1. Ng, S.E., Tay, Y.B.; Ho, T.Y.K.; Ankit; Mathews, N., Inorganic Electrochromic
Transistors as Environmentally Adaptable Photodetectors, Nano Energy, 2022, 97,
107142.
2. Zhou, L.; Vogt, F. G.; Overstreet, P. -A.; Dougherty, J. T.; Clawson, J. S.; Kord, A. S.,
A Systematic Method Development Strategy for Quantitative Color Measurement in
Drug Substances, Starting Materials, and Synthetic Intermediates, J. Pharm. Innov.,
2011, 6, 217 – 231.
3. Yamazaki, N.; Taya, K.; Shimokawa, K.-I., Ishii, F., The Most Appropriate Storage
Method in Unit-Dose Package and Correlation between Color Change and
Decomposition Rate of Aspirin Tablets, Int. J. Pharm., 2010, 396, 105 – 110.
4. Oram, P. D.; Strine, J., Color Measurement of a Solid Active Pharmaceutical
Ingredient as an Aid to Identifying Key Process Parameters, J. Pharm. Biomed.
Anal., 2006, 40, 1021 – 1024.
5. Berberich, J., Dee, K.-H., Hayauchi, Y., Pörtner, C., A New Method to Determine
Discoloration Kinetics of Uncoated White Tablets Occurring During Stability
Testing – An Application of Instrumental Color Measurements in the Development
Pharmaceutics, Int. J. Pharm., 2002, 234, 55 – 66.
6. Hetrick, E. M.; Vannoy, J.; Montgomery, L. L.; Pack, B. W., Integrating Tristimulus
Colorimetry into Pharmaceutical Development for Color Selection and Physical
Appearance Control: A Quality-by-Design Approach, J. Pharm. Sci., 2013, 102,
2608 – 2621.
7. ASTM International. Standard Practice for Computing the Color of Objects by Using
the CIE System; ASTM E308-08; West Conshohocken, PA.
8. United States Pharmacopeia and National Formulary. <1061> Color – Instrumental
Measurement. In: USP–NF. Rockville, MD: USP
9. Subert, J.; Cizmarik, J., Application of Instrumental Colour Measurements in
Development and Quality Control of Drugs and Pharmaceutical Excipients,
Pharmazie, 2008, 63, 331 – 336.
10. United States Pharmacopeia and National Formulary. <631> Color and Achromicity.
In: USP–NF. Rockville, MD: USP.
11. European Pharmacopoeia. 2.2.2. Degree of Coloration of Liquids. In: European
Pharmacopoeia. Strasbourg, France: European Pharmacopoeia.
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN56364_E 11/22M
The NanoDrop Eight Spectrophotometer
detects contaminating nucleic acids in
mammalian DNA and RNA preparations
Introduction
Understanding nucleic acid sample quality and quantity is integral
for many life science applications, reducing the occurrence of
costly delays caused by troubleshooting downstream experimental
failures. The Thermo Scientific™ NanoDrop™ Eight Microvolume
UV-Vis Spectrophotometer measures eight samples at a time
and provides you the ability to measure the concentration of
biomolecules for high-throughput assays using a 1–2 μL sample
size without the need for dilutions. With a measurement time of
less than 20 seconds, you can easily insert the NanoDrop Eight
Spectrophotometer into your high-throughput workflows.
The Thermo Scientific Acclaro™ Sample Intelligence Technology
integrated within the NanoDrop Eight Spectrophotometer’s
Technical notes
software utilizes chemometrics to detect RNA in dsDNA sample
preparations and dsDNA in RNA preparations to then calculate a
corrected dsDNA or RNA concentration, respectively. Historically,
the A260/A280 purity ratio has been utilized to assess nucleic
acid sample purity; however, nucleic acid contaminants at
low concentrations, such as RNA contamination in dsDNA
samples, have a negligible effect on the purity ratio, and the
contaminant identity is not easily determined by a change in the
A260/A280 purity ratio or by visualizing the UV-Vis spectrum.
Acclaro Technology’s contaminant analysis capability eliminates
the need for purity ratio assumptions and reports the contaminant
present, contaminant absorbance, and a corrected sample
concentration (Figure 1).
Figure 1: Acclaro Technology’s contaminant analysis screen outlining
the original concentration, corrected dsDNA concentration, and the
absorbance contribution of RNA contamination. The original spectrum is
shown in green, the corrected spectrum in pink, and the contaminating
RNA spectrum in orange.
32
Materials and methods
Total RNA and genomic DNA from mouse tissue (BioChain
Institute Inc., R1334035-50 and D1334999-G01) and RNA and
genomic DNA from the MCF-7 cell line (BioChain Institute Inc.,
R1255830-50 and D1255830) were dialyzed and diluted in
tris-EDTA buffer (TE pH 8.0, Fisher Scientific, BP2473500) and
made into various DNA/RNA mixtures according to percentage
of absorbance contribution. Triplicates of each mixture were
measured on the NanoDrop Eight Spectrophotometer using fresh
1.0 μL aliquots per replicate for the dsDNA and RNA applications.
The NanoDrop Eight Spectrophotometer’s Acclaro Technologycorrected
results from the mouse and MCF-7 DNA/RNA mixtures
were compared with the theoretical concentration and the
original, uncorrected concentration in Figures 2 and 3 using the
dsDNA and RNA applications, respectively. Acclaro Technology
calculated an original, uncorrected concentration and a corrected
concentration based on a modified Beer’s Law equation and the
absorbance contribution at 260 nm.
Figure 2: Comparison of the concentration
reported by the Acclaro Technology for
different sample compositions of DNA and
RNA based on percentage of absorbance
contribution. DNA and RNA from either the
MCF-7 cell line or mouse tissue were mixed
according to absorbance percentage and were
measured using the dsDNA application. The
mean original concentration (blue bars), the
theoretical concentration (orange bars), and the
mean Acclaro Technology software-corrected
concentration (gray bars) were reported by the
NanoDrop Eight Spectrophotometer’s software.
Error bars represent the standard deviation.
Figure 3: Comparison of the concentration
reported by the Acclaro Technology for different
sample compositions of DNA and RNA based
on percentage of absorbance contribution.
DNA and RNA from either the MCF-7 cell line
or mouse tissue were mixed according to
absorbance percentage and were measured
using the RNA application. The mean original
concentration (blue bars), the theoretical
concentration (orange bars), and the mean
Acclaro Technology software-corrected
concentration (gray bars) were reported by the
NanoDrop Eight Spectrophotometer’s software.
Error bars represent the standard deviation.
Results
In Figures 2 and 3, the Acclaro Technology’s software-corrected
mean concentration from the NanoDrop Eight Spectrophotometer
was graphed against the original, uncorrected concentration
and the theoretical concentration for the mouse and MCF-7
DNA/RNA mixtures with standard deviation shown as error bars.
Since nucleic acids absorb at 260 nm, the original, uncorrected
concentration is inflated compared to the Acclaro Technology’s
software-corrected concentration when DNA and RNA are both
contributing to absorbance.
With the inclusion of the Acclaro Technology in the NanoDrop
Eight Spectrophotometer’s software, the corrected nucleic acid
concentration was calculated after correcting for the contaminant
absorbance contribution. This feature allows for simultaneous
nucleic acid purity and quantity assessments. All the Acclaro
Technology’s software-corrected concentrations fall within ±20%
of the theoretical concentration, with most samples within ±10%.
33
Learn more at thermofisher.com/nanodropeight
Conclusion
Contaminating nucleic acids in dsDNA or RNA preparations
can cause costly delays in applications such as qPCR, where
exact quantitation is crucial for a successful experiment.
Since RNA and dsDNA both absorb at 260 nm, the true
nucleic acid concentration will be overestimated with a
copurified contaminant present. This overestimation can lead
to experimental failures and require extensive troubleshooting.
The ease with which the Acclaro Technology corrects for
contaminating nucleic acids will save time, effort, and associated
costs by improving sample purity and quantity assessments.
The function of the Acclaro Technology makes the nucleic
acid purity assessment clear and simple. With each
measurement of a nucleic acid sample, the NanoDrop Eight
Spectrophotometer takes quality assessment a step further
by outlining the contaminant identification, absorbance
contribution, and the corrected sample concentration. The
results from the experiments above indicate the NanoDrop Eight
Spectrophotometer, which includes the Acclaro Technology in
its software, can be implemented into many molecular biology
workflows to obtain an accurate and advanced nucleic acid
evaluation for downstream success.
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2021 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. TN53470 1121 M
Enabling real-time release of final products in
manufacturing of biologics
Application note
Keywords
DXR3 SmartRaman, spectrometer,
biopharmaceutical, GMP,
real-time release testing, QbD, RTRT,
manufacturing, multi-attribute testing
Authors
Shaileshkumar Karavadra,
David James, and Arnaud Di Bitetto,
Hemel Hempstead, UK
Introduction
Biopharmaceuticals (or biologics) are manufactured using biological-expression
systems (such as mammalian, bacterial, and insect cells) and have spawned
a large and growing biopharmaceutical industry (BioPharmaceuticals). The
structural and chemical complexity of biologics, combined with the intricacy
of cell-based manufacturing, imposes a huge analytical burden to correctly
characterize and quantify both processes (upstream) and products (downstream).
In small-molecule manufacturing, advances in analytical and computational
methods have been extensively exploited to generate process analytical
technologies (PAT) that are now used for routine process control, leading to more
efficient processes and safer medicines.
Raman spectroscopy is a vibrational spectroscopy technique with several useful
properties (non-destructive, non-contact, high molecular-specificity,
and robustness) that make it particularly suited for PAT applications in which
molecular information (composition and variance) is required.
Typical good manufacturing practice (GMP) operations involve performing an
extensive set of tests according to approved specifications before the material
is released to the market or for further processing. Recent ICH guidelines
(ICH Q8, Q9, Q10, and Q11), however, suggest an alternative real-time release
strategy to provide assurance of product quality prior to release. Real-time
release testing uses the principles of the pharmaceutical Quality by Design
(QbD) to optimize release and stability testing. A combination of manufacturing
process understanding, process control, and product knowledge can be used
to demonstrate that the material was made according to GMP.
The exact approach to real-time release testing (RTRT) will vary depending on
the process requirements. The RTRT strategy may be based on control of process
parameters, monitoring of product attributes, or on a combination of both at
appropriate steps throughout the process. Critically, the RTRT strategy should
be based on a firm understanding of the process and the relationship between
process parameters, in-process material attributes, and product attributes.
Quality, cost, and speed are the major drivers for implementing in-line monitoring,
at-line monitoring, and real-time release.
35
Here, we review some of the most important applications
of Raman spectroscopy to the manufacturing and analysis
of biopharmaceuticals. This article covers two aspects of the
biopharmaceutical-manufacturing process: identity/variance
testing of raw materials and cell culture media; and multiattribute
product testing of a biologic drug product or final
product testing of a biologic drug product.
Raw material characterization
Acceptance of raw materials today is often predicated on
small-scale functional testing and/or limited analytical methods,
which may not be representative of at-scale performance.
This leads, in some cases, to fluctuating process outputs and,
in extreme cases, not meeting predefined release criteria.
Furthermore, many clinical products are developed using
a small number of batches resulting in a narrow range of raw
material variation and thus a limited process understanding.
Especially in upstream cell culture, the unforeseen variability
of various components of the cell culture media can impact
a product’s micro-heterogeneity and its critical quality
attributes (CQA).
Multi-attribute tests for high-risk raw materials may include
identity test, quantitative test for the concentration of key
ingredients in a raw material, batch-to-batch variability test,
and degradation tests.
One high-risk raw material encountered in biologics
manufacturing is cell culture media. Identification of cell
culture media samples by traditional liquid chromatography
(LC) methods, such as amino acid or vitamin analysis, has
high costs and requires significant analytical expertise and
laboratory space. Raman spectroscopy offers many potential
benefits, such as low cost, portability, and potentially limited
skill required to operate the instruments.
Buffers are another set of critical raw materials used in
downstream manufacturing. Osmolality is a measure of
concentration and is considered a critical quality attribute and
critical process parameter in bioprocessing. The yield and
quality of a biologic are highly dependent on the optimization
of the downstream process. Identity testing along with
osmolality of buffers can be carried out using a multi-attribute
method based on principal component analysis and partial list
squares. Rapid testing of buffers through single-use flexi bags
can be carried out using the fiber optics probe of the Thermo
Scientific™ DXR3 SmartRaman Spectrometer at the point
of use with no need for sample preparation.
Final product identity testing
Final product identification of biologics pre- and post-shipment
is another regulatory requirement. Product testing for identity
through different kinds of primary packaging (glass vials,
syringes, glass bottles) poses a significant analytical challenge
in the manufacturing of biologics. Fill finish sites may not have
the necessary analytical expertise to carry out the tests and
may have to send the samples to the parent site or external lab
for testing, incurring time and money.
Moreover, biologics or small molecule drug products would
also have to undergo retesting upon importation either from
a third country in the EU member state or the USA when drug
products have been sent to the USA from other countries.
A full list of tests is typically carried out, including final product
identity testing. For biopharma manufacturers, this involves
either sending the samples back to the parent site for analysis
or employing third-party labs in the country of import. This
increases significant costs and delays in the delivery of highly
needed drug products.
End product identity testing/final product identity testing
of biologics after fill-finish or pre-shipping to the fill-finish line
is carried out by a variety of analytical techniques depending
on the molecule/registration dossier.
For example, the verification test for biologic proteins is peptide
mapping—a long-established workflow for protein identification
using LC/mass spectrography (MS). This complex separation
technique requires protein extraction and clean-up, enzyme
digestion, one or more stages of liquid chromatography, and
two phases of mass spectrometry before the final spectrum is
matched against protein databases. Although it is a standard
methodology, peptide mapping necessitates an analytical
lab with qualified technical resources, entails extensive time
for preparation, and introduces significant costs in solvents,
columns, and analytical equipment.
The DXR3 SmartRaman Spectrometer, with its high sensitivity
and resolution, allows characterization of the drug product
by evaluating the fingerprint region of the molecule. Therefore,
the DXR3 SmartRaman Spectrometer’s unique capability with
sampling flexibility ensures repeatable measurements, and
subsequent analysis allows rapid method development and
deployment.
We ran a feasibility study for multinational drug manufacture
whereby the primary goal was to set up a rapid multiattribute
end product test to differentiate 15 different types
of drug products and determine the concentration of the two
preservatives in the drug products.
36
Band frequency (cm-1) Region Vibrational mode Protein structure assignments
870–1,150 Backbone,
skeletal stretch
Cα-C, Cα-Cβ, Cα-N Secondary structure elements: α-helix,
β-sheets, less-ordered structure
1,200–1,340 Amide III N-H in-plane, Cα-N stretch Hydrogen bonding, secondary structure
1,400–1,480 Side chain
deformations
CH2 and CH3 deformations Local environments, intermolecular
interactions of side chains
1,510–1,580 Amide II N-H deformations and C-N
stretch (observed in UVRR and
not conventional Raman spectra)
Local environments, intermolecular
interactions of side chains
1,630–1,700 Amide I C=0 stretch N-H in-plane bending Secondary structure elements:
α-helix, β-sheet, less-ordered structure
For this feasibility test we were given 15 different types
of biologic drug products that varied in concentration from
0.5 mg/mL to 6 mg/mL. Concentration of two preservatives
A and B ranged from 0.85 mg/mL to 5.0 mg/mL and
0.42 mg/mL to 3.91 mg/mL respectively.
These commercial drug products were supplied in their native
glass vials varying in size and volume. A picture of such glass
vials is shown below (Figure 1).
Reversed-phase high-performance liquid chromatography
(HPLC) is currently used for the final product identity test and
quantitative measurement of two preservatives in the final
drug product.
DXR3 SmartRaman Spectrometer with universal sampling plate
and 180-degree sampling module was used to acquire spectra
of 15 drug products. To acquire each spectrum, a 532 nm laser
with 40 mW power and 1 minute of scanning time was used.
Ten spectra were acquired for each sample to accommodate
the variability of glass vials and scattering effects.
DXR3 SmartRaman Spectrotometer offers excellent selectivity,
repeatability, and full wavelength range to characterize
biologics based on the characteristic band assignment (Table 1
and Figure 2).
Figure 1. Typical native glass vials.
Table 1. Characteristic Raman band assignment.
Figure 2. DXR3 SmartRaman spectrum showing characteristic bands of a biologic drug product.
Raman intensity
Raman shift (cm-1)
3,800
3,600
3,400
3,200
3,000
2,800
2,600
2,400
2,200
2,000
1,800
1,600
1,400
1,200
1,000
800
600
400
200
0
400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200 2,400 2,600 2,800 3,000
37
Figure 3 shows the spectra of a sample containing a drug
product against its placebo. It is imperative to establish that
technique chosen for a feasibility study. In this case, Raman
spectroscopy is sensitive enough to detect the differences
between the drug product and its placebo. DXR3 SmartRaman
Spectrometer offers high sensitivity to determine the significant
differences between placebo and actual drug products.
Figure 4 is showing spectra of different classes of drug
products. These spectra were utilized to build the discriminant
analysis method on the Thermo Scientific™ TQ Analyst™
Software. TQ Analyst Software is a validated qualitative and
quantitative method building software offering full compliance
for pharmaceutical applications.
The discriminant analysis classification technique can
be used to determine the class or classes of known materials
that are most similar to an unknown material by computing
the unknown’s distance from each class center in Mahalanobis
distance units. The discriminant analysis technique is typically
used to screen incoming materials or final products to
determine if they are compound/molecule a, b, or c.
Discriminant analysis methods typically specify at least two
classes of known materials, but the method also works with
only one class. Multiple standards may be used to describe
each class (at least one class must contain two or more
standards). Multiple regions of the spectrum may be used for
the analysis.
Figure 3. Raman spectra of drug product and its placebo and variance spectrum.
Figure 4. Raman spectra of different classes of drug products.
Raman intensity Raman intensity
Raman shift (cm-1)
Raman shift (cm-1)
8,500
8,000
7,500
7,000
6,500
6,000
5,500
5,000
4,500
4,000
3,500
3,000
2,500
2,000
1,500
1,000
500
1,600 1,400 1,200 1,000 800 600
3,000
2,800
2,600
2,400
2,200
2,000
1,800
1,600
1,400
1,200
1,000
800
600
500 1,000 1,500
3,400
3,200
3,000
2,800
2,600
2,400
2,200
2,000
1,800
1,600
1,400
1,200
1,000
400 600 800 1,000 1,200 1,400
Raman intensity
Raman shift (cm-1)
38
What does discriminant analysis do?
A discriminant analysis method applies the spectral information
in the specified region or regions of an unknown sample
spectrum to a stored calibration model to determine which class
of standards is most similar to the unknown.
When the method is used to analyze an unknown sample or a
class, the software performs a principal component analysis on
the spectra of the standards and uses those results to determine
score values for the unknown sample spectrum. The score plots
are used to produce Mahalanobis distance values, which in turn
are used to rank the classes.
The result of a discriminant analysis is the name of the class or
classes that are most similar to the spectrum of the unknown
sample. The Mahalanobis distance between the unknown sample
and each reported class can also be reported. The closer each
distance value is to zero, the better is the match.
After cross-validation, principal component scores plot
revealed the class differentiation and the report indicated that
all the classes of the different products were correctly identified
with no mismatches to indicate false positives.
Quantitative analysis of biologics
for preservative A and preservative B
As part of this feasibility study, our client also wanted to
determine if the DXR3 SmartRaman Spectrometer test
could be utilized to replace the HPLC test for measuring the
concentration of two preservatives in their drug products. The
level of preservative A was 0.85 mg/mL to 3.07 mg/mL and that
of preservative B was 0.32 mg/mL to 2.57 mg/mL.
Pure samples of preservatives A and B were acquired
as references, and to ascertain their presence in the final
drug formulation.
Figure 5. Analysis of preservative A and preservative B.
Actual class Mismatch Calculated class Calculated distance Next class Next distance
Product D Product D 0.5809 C 4.5556
Product A Product A 1.9869 I 12.9617
Product B Product B 1.3796 E 25.1324
Product C Product C 0.5417 D 3.8568
Product D Product D 0.8466 M 9.0495
Product I Product I 1.7709 A 13.9064
Product M Product M 0.5284 S 3.3881
Product O Product O 0.2244 X 17.3044
Product R Product R 0.5419 C 4.4691
Product T Product T 0.5944 X 2.3213
Product X Product X 0.79 T 3.1646
Product S Product S 1.1837 M 3.0829
Product N Product N 1.0954 U 15.1798
Product U Product U 0.1603 T 9.1738
Product S Product S 1.8544 N 22.1624
39
Samples of varying concentrations as per table 1 were
acquired using the same parameters as of spectra acquired for
identity test through 3 mL vial. Figure 6 is showing the spectra
of the drug product with the two preservatives.
Four standards with the reference values were supplied
in 3 mL and 10 mL vials and a validation sample to test the
model for 3 mL and 10 mL vials.
Four spectra per standard were acquired and used to build the
chemometric method. The final drug product samples were
scanned with a DXR3 SmartRaman Spectrometer to acquire
spectra in the range of 3500 to 50 cm-1 and captured with
a single exposure of the CCD, avoiding stitching artifacts. The
sample time took approximately 1 minute. Three spectra were
collected per sample. The sample spectra were loaded into TQ
Analyst Software for chemometric analysis using a partial least
squares (PLS) method.
Preservative A
(mg/mL)
Preservative B
(mg/mL)
Standard 1
3 mL and 10 mL
0.85 0.42
Standard 2
3 mL and 10 mL
1.27 1.12
Standard 3
3 mL and 10 mL
1.57 1.75
Standard 4
3 mL and 10 mL
3.07 2.57
Validation – 3 mL 1.57 1.75
PLS results for 3 mL Cartridge
Preservative A
(mg/mL)
Preservative B
(mg/mL)
Validation sample:
3 mL
1.58
actual 1.57
1.71
actual 1.75
Real Sample in
solution: 3 mL
1.56
actual 1.55
1.69
actual 1.77
Real sample in
suspension: 3 mL
0.72
actual 0.69
1.23
actual 1.58
Table 2. Calibration and validation sample.
Table 3. Validation result for 3 mL sample.
Figure 6. Spectrum in blue is from pure preservative A and spectrum in red is from pure preservative B.
Figure 7. Spectra showing varying concentration of preservatives in final drug product.
Raman intensity
Raman shift (cm-1)
Raman intensity
Raman shift (cm-1)
1,200
1,100
1,000
900
800
700
600
500
400
300
200
100
0
3,500 3,000 2,500 2,000 1,500 1,000 500
1,800 1,600 1,400 1,200 1,000 800 600 400
30,000
25,000
20,000
15,000
10,000
5,000
0
40
Results
PLS analysis of the final drug product samples revealed
excellent predictive capabilities within the range of materials
tested. The spectra used to develop the PLS method for 3 mL
cartridge are shown on calibration plots (Figure 8 and Figure 9)
that compare the calculated preservative concentrations
versus the actual concentrations. The calibration plot can be
used to determine how well the method predicts the actual
preservative concentrations in the samples. The plot developed
by the chemometric method resulted in a correlation coefficient
of 0.998 for preservative A. Root mean square error of
calibration (RMSEC) was 0.0425 mg/mL, and the Root mean
square error of prediction (RMSEP) calculated was 0.0372
for preservative A. The additional method for preservative B
resulted in in a correlation coefficient of 0.999. The RMSEC
was 0.0316 mg/mL, and the calculated RMSEP was 0.0496.
The method was able to accurately predict the 3 mL validation
sample and a real sample in solution (Table 3). The prediction
can be improved when suspensions are allowed to settle and
liquid phase is analyzed.
When 10 mL vial calibration samples were added to the above
PLS method, method performance remained the same and
was able to accurately predict the validation samples (Table 4).
Conclusions
A multi-attribute test to establish Final product identification
and predicting concentrations of preservatives was done with
the DXR3 SmartRaman Spectrometer by developing
a discriminant analysis method and partial least square
method. The final drug product identification test is part
of release testing and current methods used are timeconsuming
and laborious. This Raman technique successfully
demonstrates the ability to measure and monitor preservative
concentrations either in the lab environment or at the line.
The method developed shows excellent correlation with actual
preservative concentrations with errors comparable to the
reference analysis method. This application demonstrates the
continued capability of the DXR3 Raman Spectrometer
to be successfully used in bioprocess environments for
implementing multi-attribute final product testing of biologics.
Apart from the examples shown here, DXR3 SmartRaman
Spectrometer can be used to implement at-line control
strategies to monitor protein concentration, excipients
concentration, and critical quality attributes like osmolality and
pH. Many such examples are cited in the literature for Raman
applications in biopharma manufacturing.
Learn more at thermofisher.com/brighteroutcomes
Figure 8. PLS model for preservative A —3 mL cartridge.
Figure 9. PLS model for preservative B —3 mL cartridge.
Table 4. Validation results for 3 mL 10 mL vials.
References:
1. European Medicines Agency. Guideline on Real Time Release Testing.
https://www.ema.europa.eu/en/documents/scientific-guideline/guideline-realtime-
release-testing-formerly-guideline-parametric-release-revision-1_en.pdf
(accessed June 15, 2021).
2. Buckley K, Ryder AG. (2017). Applications of Raman spectroscopy
in biopharmaceutical manufacturing: a short review. Applied Spectroscopy,
71(6): 1085-1116. https://aran.library.nuigalway.ie/handle/10379/7177
(accessed June 15, 2021).
PLS 3 mL cart and 10 mL vials
Preservative A
(mg/mL)
Preservative B
(mg/mL)
Validation sample:
3 mL
1.58
actual 1,57
1.71
actual 1,75
Real sample in
solution: 3 mL
1.56
actual 1.55
1.65
actual 1.77
Real sample in
suspension: 3 mL
0.80
actual 0.69
1.21
actual 1.58
Real sample in
suspension: 10 mL
0.73
actual 0.68
1.32
actual 1.57
Calculated
Actual
3.2
0.7
0.7 3.2
Calculated
Actual
2.7
0.3
0.3 2.7
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2022 Thermo Fisher Scientific Inc. All rights reserved. All trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN53435 0522/M
For research use only. Not for use in diagnostic procedures. For current certifications, visit thermofisher.com/certifications
© 2023 Thermo Fisher Scientific Inc. All rights reserved. BioCell and PROTA-3S are trademarks of BioTools, Inc. ConcentratIR2 is a trademark of Harrick Scientific
Products, Inc. All other trademarks are the property of Thermo Fisher Scientific and its subsidiaries unless otherwise specified. CS56393_E_08/23M
Notes
Learn more at thermofisher.com
Brought to you by
Download the eBook for FREE Now!
Information you provide will be shared with the sponsors for this content. Technology Networks or its sponsors may contact you to offer you content or products based on your interest in this topic. You may opt-out at any time.
Experiencing issues viewing the form? Click here to access an alternate version