Sie sind auf Seite 1von 9

International Journal of Pharmaceutics 436 (2012) 282–290

Contents lists available at SciVerse ScienceDirect

International Journal of Pharmaceutics


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

Long-term stability of diluted solutions of the monoclonal antibody rituximab


Muriel Paul ∗ , Victoire Vieillard, Emmanuel Jaccoulet, Alain Astier
APHP, GH Henri Mondor, Département de Pharmacie, Unité Pharmaceutique de Recherche en Essais Cliniques, 51 avenue du Maréchal de Lattre de Tassigny, 94010 Créteil, France

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

Article history: As it is an important challenge for pharmacists to access the stability of monoclonal antibodies because
Received 23 February 2012 of the widespread of centralized preparation units, we conducted a study to evaluate the physicochem-
Received in revised form 29 June 2012 ical and biological stability of diluted rituximab at 1 mg/mL over six months at 4 ◦ C. We also conducted
Accepted 30 June 2012
the study at 40 ◦ C to demonstrate that all methods employed were stability indicating. Various protein
Available online 10 July 2012
characterization methods were used to determine changes in physicochemical properties of rituximab,
including size-exclusion chromatography, dynamic light scattering, turbidimetry, cation-exchange chro-
Keywords:
matography, second-derivative ultraviolet and infrared spectroscopy, and peptide mapping. Cell culture
Rituximab
Physicochemical stability
was used to assess biological stability. We demonstrated that diluted rituximab stored at 4 ◦ C in polyole-
Biological stability fine bags remained stable for at least six months. No physical or chemical instability was observed, and
Diluted solution the biological activity was fully maintained. Size exclusion chromatography did not show polymerization
Monoclonal antibody or fragmentation. No difference was noticed in the hydrodynamic diameter of RTX. No additional peak or
decrease in the areas under curve was found by cation exchange chromatography. The thermal aggrega-
tion curves and their derived thermodynamic parameters were unchanged. The primary, secondary and
tertiary structures of the protein were not modified. Finally, the direct cytotoxic effect of rituximab was
fully maintained.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction Many of these stress factors are routinely encountered during


the preparation, purification, shipping or storage of protein prod-
Monoclonal antibodies (mAbs) are a class of biopharmaceuti- ucts (Jiskoot et al., 1990). Therefore, careful control of these factors
cal products mostly used as therapeutic agents. They represent helps preserve the stability of the protein, which is a real challenge
the fastest-growing segment of the biopharmaceutical market for pharmaceutical manufacturers. Additionally, the preparation
(Nicolaides et al., 2006). Rituximab (RTX), the first IgG1 mAb suc- must be stabilized against deleterious environmental factors to
cessfully used in onco-hematology, is a chimeric mAb indicated in preserve its pharmacological properties and to reduce potential
Non-Hodgkin’s Lymphoma (Cheung et al., 2007), rheumatoid pol- immunogenicity (Schellekens, 2002; Hermeling et al., 2004; Hwang
yarthritis (Edwards et al., 2004) and chronic lymphoid leukemia. It and Foote, 2005; Sharma, 2007a,b).
is directed against CD20, which is present in all the B-cell lines, and RTX is formulated as a 10 mg/mL solution containing sodium cit-
induces a depletion of CD20-positive cells (Pedersen et al., 2002). rate dihydrate as buffer (pH 6.5) (Wang et al., 2007) and surfactant
Because of its proteic nature, RTX may go through a variety of polysorbate 80. Their thermodynamic and physicochemical prop-
chemical and physical degradation processes, as recently reviewed erties favor the equilibrium of attractive and repulsive strengths
by Manning (Manning et al., 2010). Physical instability mostly con- and thus avoid aggregation (Chi et al., 2003). The use of citrate
cerns low energy bonds (such as hydrogen bonds), and includes buffer slows the deamidation rates by altering the conformation
adsorption on to surfaces, unfolding and aggregation (Wang, 1999; of the protein (Zheng and Janis, 2006).
Chi et al., 2003; Wang et al., 2007). Chemical instability, such as However, according to the manufacturer’s instructions, RTX
asparagine deamidation (the most common chemical degradation must be diluted into a solution of 0.9% sodium chloride and admin-
process), aspartic acid isomerization, oxidation or disulfide bond istered within 24 h. Therefore, in practice, the stability of diluted
shuffling, concerns covalent bond modifications (Wakankar and RTX would be only 24 h. There is surprisingly little information on
Borchardt, 2006; Chumsae et al., 2007). mAbs’ storage stability, except for some therapeutic immunoglob-
ulins that exhibit a long shelf life (over one year) in liquid state.
Moreover, Bakri reported in 2006 that bevacizumab was stable for
a period of six months at 4 ◦ C (Bakri et al., 2006). Thus, it can be
∗ Corresponding author at: Département de Pharmacie, GH H. Mondor, 51 avenue hypothesized that this recommended short-term stability is mainly
du Maréchal de Lattre de Tassigny, 94010 Créteil, France. Tel.: +33 149812760. focused on avoiding bacterial contamination, for handling in wards
E-mail address: muriel.paul@hmn.aphp.fr (M. Paul). (Bardin et al., 2011).

0378-5173/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ijpharm.2012.06.063
M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290 283

It is an important challenge for pharmacists to assess the stabil- (Sigma® centrifuge; Osterode, Germany) before SEC, CEX and UV
ity of biotechnology-derived anticancer drugs such as antibodies spectrophotometric analysis. All buffers were filtered at 0.22 ␮m
because of the widespread, ready-to-use, centralized preparation and all dilution steps were conducted under laminar airflow.
of these drugs in pharmacy departments. Thus, the objective of this
study was to assess the physical, chemical and biological stability 2.5. Turbidity determination (Mahler et al., 2005)
of diluted RTX (1 mg/mL) over six months under an optimal storage
temperature (Bardin et al., 2011). The turbidimetric method is often used to estimate the amount
As mAbs are more susceptible to degradation at 40 ◦ C than 4 ◦ C of visible protein aggregates by measuring the optical density of
(Liu et al., 2006), we also conducted experiments at 40 ◦ C to deter- samples based on light scattering in the near-UV or visible regions
mine the different instability mechanisms and to demonstrate that (Thirumangalathu et al., 2006). Aggregation of RTX was monitored
the methods used were good indicators of protein stability. by measuring absorbance at 350 nm (Seefeldt et al., 2005) using a
Various protein characterization methods were used to deter- UV–VIS spectrophotometer (Cary 50 Probe, Varian, Inc., Palo Alto,
mine changes in physicochemical properties of RTX, including USA). Here, the turbidance (T ) was measured as absorbance at
size-exclusion chromatography (SEC), dynamic light scatter- 350 nm, where none of the known intrinsic chromophores in the
ing (DLS) and turbidimetry, cation-exchange chromatography protein formulation absorb (Wang, 2005).
(CEX), second-derivative ultraviolet and Fourier-transform infrared The samples were diluted to 0.4 mg/mL in buffer (20 mM
spectroscopy (FT-IR), and peptide mapping after trypsin and KH2 PO4 , pH 6.0). The mean value of each experimental sample
endopeptidase digestion. Cell culture was used to assess biological absorbance was calculated and compared to the mean value of the
stability. reference sample.

2. Materials and methods 2.6. Tertiary structure analysis (Balestrieri et al., 1980)

2.1. Reagents The tertiary structure of RTX was examined by second deriva-
tive UV spectroscopy. This method allows for an evaluation of the
Disodium hydrogen phosphate, sodium azide, potassium dihy- impact of dilution and storage on the aromatic amino acids of
drogen phosphate and trifluoroacetic acid (TFA) were purchased the mAb. Indeed, the position of second derivative UV absorbance
from Merck (Darmstadt, Germany). Ammonium bicarbonate, ace- peaks is sensitive to the polarity of the microenvironment of the
tonitrile, dithiothreitol (DTT), guanidine hydrochloride (GnHCl), aromatic amino acids and therefore can be used to provide a global
iodoacetic acid, 2,3-bis-(2-methoxy-4-nitro-5-sulfophenyl)-2H- picture of the tertiary structure of a protein. A Cary 50 Probe
tetrazolium-5-carboxanilide, sodium salt (XTT), and phenazine UV–visible spectrophotometer (Varian) was used. The spectra were
methyl sulfate (PMS) were obtained from Sigma–Aldrich (St Louis, collected in the near-UV region from 320 to 250 nm in a 1 cm
MO, USA). Formic acid was purchased from Prolabo (Fontenay- path length quartz cuvette. Five major peaks were studied (252-
sous-Bois, France). Isotonic saline was provided by Fresenius Kabi Phe, 258-Phe, 275-Tyr, 284-Thy/Trp and 292-Trp), as previously
(Louviers, France). All reagents were of analytical grade. reported for IgG2, and also two minima at 287 and 295 nm (Servillo
et al., 1982; Kueltzo et al., 2007). The spectra of each sample (D30,
2.2. Materials D90 and D180) were compared to the spectra of reference sample.

RTX solutions were obtained from the commercially avail- 2.7. Determination of molecular hydrodynamic diameters
able product Mabthera® (Roche, Switzerland), formulated in
7.35 mg/mL sodium citrate dihydrate, 9 mg/mL sodium chloride, Dynamic light scattering (DLS) is a sensitive method to evaluate
sodium hydroxide and hydrochloric acid to obtain a pH of 6.5, and the size of aggregate populations containing aggregates of 1 nm
polysorbate 80 (PS80) as a stabilizing agent. The formulation was to 6 ␮m. The analysis is based on the measurement of fluctuating
presented in glass vials as a concentrated solution of 10 mg/mL scattered light intensity due to Brownian particles motion, which
ready to dilute in 0.9% sodium chloride solution, according to the also defines diffusion.
manufacturer’s instructions. A Malvern Zetasizer Nano ZS (Worcestershire, UK) laser light
scattering system was used to obtain DLS measurements with a
2.3. Methods 633 nm laser source. The NanoZS® software was used for data
acquisition and analysis. A total of 400 ␮l of each sample at
For this study, RTX diluted bags were prepared from Mabthera® 1 mg/mL was added in a 1 cm path length quartz microcuvette
(Roche Laboratories) by dilution with 0.9% sodium chloride in rinsed before use with 0.22 ␮m filtered water. All the operations
Freeflex® polyolefin bags (Freeflex® - Fresenius Kabi, France). The were performed under laminar airflow. The size and hydrodynamic
final concentration was 1 mg/mL. Two different storage temper- diameter distributions of the detected aggregates were obtained by
atures were tested: 4 ◦ C [2–8 ◦ C] and 40 ◦ C in an ICH climatic the measurement and integration of the light scattering intensity.
incubator (60% of relative humidity Climacell® ; Bioblock Fisher Sci-
entific, Illkirch, France). For each temperature, three batches were 2.8. Thermal denaturation curves (Wang, 1999; Harn et al., 2007)
prepared, and all RTX diluted bags were protected against light with
opaque plastic bags. With increases in temperature or simply with time, the sec-
ondary, tertiary and quaternary structures of a protein may change
2.4. Samples preparation and lead to protein unfolding and/or aggregation, which are crit-
ical factors of physical instability. These events can be monitored
Samples were withdrawn from diluted RTX bags under sterile because dramatic changes in protein size during denaturation can
conditions in a laminar flow hood to prevent potential bacterial or be easily identified with DLS techniques. We selected a heating
particulate contaminations at days D0, D14, D30, D90 and D180. rate of 1 ◦ C/min to draw the heating curves, allowing the system
A reference sample was obtained from the immediate dilution to be a quasi-thermodynamic equilibrium showing good repeti-
of fresh RTX vials at 1 mg/mL, and was freshly prepared before tion of the results. So, the automated measurements were collected
each step. The samples were centrifuged at 5000 rpm for 5 min using a 1 ◦ C incremental temperature ramp and a three-minute
284 M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290

equilibrium time at each temperature measurement from 25 ◦ C to temperature was set at 25 ◦ C. The parameters studied were the AUC,
90 ◦ C. The temperature of aggregation (Tagg ), at which the aggre- the retention time and the molecular weight.
gated fraction is 0.5 during thermal transition, can be determined The molecular weight of the peaks was calculated using a stan-
using the nonlinear least squares method to fit the diameter depen- dard solution including several molecular weight markers ranging
dence on the temperature. The thermodynamics parameters were from 1.2 to 150 kDa (hydroxocobalamin, cytochrome C, albumin
calculated as previously by Farruggia et al. (1997). and cetuximab), and using a calibration curve fitted as a third-
degree polynomial function.
2.9. Secondary structure analysis (Dong et al., 1997)
2.10.3. Peptide mapping (Dick et al., 2009)
The secondary structure of RTX was monitored by second All samples were concentrated at 16 mg/mL using Amicon®
derivative FT-IR spectroscopy in the conformation-sensitive amide Ultra-10 centrifugal filter units (Millipore) by centrifugation at
I band region (1600–1700 cm−1 ), which is due almost entirely to 4000 × g for 20 min. One milliliter of the concentrated sample was
C O stretch vibrations of peptide linkages (Kong and Yu, 2007). withdrawn for the next steps. These steps are identical to those
This method has been extensively used in protein denatura- described par our team in a previous article (Lahlou et al., 2009) and
tion/aggregation studies (Wang, 1999). The samples (D0, D30, D60 consisting in denaturation and trypsin and endopeptidase diges-
and D180 at 4 ◦ C) were concentrated to 16 mg/mL using 10 kDa tion.
Centricon concentration devices (Millipore, Molsheim, France). IR
spectra were recorded with a 2000 FT-IR system (Perkin-Elmer, 2.11. Direct cytotoxicity assay (Yang et al., 2003)
les Ulis, France). For each spectrum, a 256-scan interferogram was
collected in single-beam mode, with a 4 cm−1 resolution. The sam- To detect potential changes in RTX biological activity after
ple cell consisted of CaF2 windows separated by a Teflon spacer of long-term storage, RTX direct cytotoxic (intrinsic toxicity) on
15 ␮m. The spectrum of reconstituted buffer (identical composition CD20-expressing cells was determined through concentration-
to the one used in the commercialized drug) was obtained under dependent cytotoxicity curves on a human lymphoma cell line.
identical conditions. For the analysis, the spectra were initially con- RAJI lymphoma cells (ATCC number: CCL-86) were cultured in
verted to absorbance followed by subtraction of the corresponding dye-free RPMI 1640 medium supplemented with 10% fetal bovine
buffer spectrum. Second-derivative spectra were calculated with serum and 1% glutamine at 37 ◦ C in a humidified CO2 incubator (5%
the Spectrum® software (Perkin-Elmer) to calculate peak areas CO2 ). After dilution at 0.5 × 105 cells/mL, 100 ␮L aliquot of the cell
and determine the percentage area under each peak. All spectra suspension were transferred to each well of a 96-well plate and
obtained were the mean of three separate scans. The absorbance cultured with various concentrations of RTX (0, 4.5, 9, 22.5, 45, 90,
wavelengths for the different secondary structures were those 135, 270 and 450 ␮g/mL) for three days. Twelve wells were seeded
described by different authors (Dong et al., 1997; Matheus et al., for each concentration tested. To assess cell viability, 50 ␮L of a
2006; Kong and Yu, 2007). solution containing 50 ␮g of XTT and 0.38 ␮g of PMS were added.
The cells were incubated at 37 ◦ C for 4 h. The absorbance of each
2.10. Chromatographic analysis (Lahlou et al., 2009) well at 450 nm was measured using an automated plate reader.
The results were expressed as the percentage of inhibition in com-
All chromatographic analyses were realized with a Dionex parison to control (without RTX). The IC50 and IC25 were calculated
Station consisting of a Dionex Ultimate 300 chromatographic sys- using the Michaelis–Menten model. The AUCs of the cytotoxicity
tem coupled with a Chromeleon® chromatography workstation curves were also calculated, using the trapezoidal method.
(Dionex Corporation, Sunnyvale, CA, USA). The system is equipped
with a gradient pump with degas option and gradient mixer, UV 2.12. Statistical analysis
detector and autosampler.
The results are presented as mean ± standard deviation (SD)
2.10.1. Cation exchange chromatography (CEX) obtained from three separate samples and for each storage con-
The different ionic species of RTX were separated on a Dionex dition and concentration. All data were analyzed with the Kyplot
Propac® WCX-10 cation exchange column (4 mm × 250 mm) with a software (V2. Koichi Yoshioka). Analysis of variance (two-way
WCX-10 guard column (4 mm × 50 mm, Dionex, Voisins le Breton- ANOVA) or non-parametric Mann–Whitney tests, if required, were
neux, France). Buffer A was prepared with 50 mM KH2 PO4 , 0.05% used to compare data. Statistical significance was defined as
NaN3 pH 6.0, and buffer B was prepared with 50 mM KH2 PO4 , 1 M p < 0.05.
sodium chloride and 0.05% NaN3 (w/v), both in deionized water.
The flow rate was 1 mL/min, and the injected volume was 100 ␮L. 3. Results
The elution gradient started with 100% of A to achieve 50% of A and
B after 20 min, followed by an equilibrium phase of 5 min to obtain 3.1. Turbidity
100% of buffer A. The elution was monitored at 280 nm. The mean
value of the area under the curve (AUC) and the retention time (rt) No variation in optical density was observed for six months
were compared to those of the reference sample. at 4 ◦ C (mean: 0.0418 ± 0.0016). At 40 ◦ C, the rate of aggregation
increased with the temperature, and a linear increase in optical
2.10.2. Size-exclusion chromatography (SEC) density (r2 = 0.95) was observed. After six months at this temper-
SEC analysis is a classical method used in protein studies to ature, an increase in optical density was found, but remained low
detect potential fragmentation and aggregation (small size aggre- (0.0614 ± 0.0001).
gates), corresponding to physical and chemical instabilities. A
7.8 mm × 300 nm TSK-GEL G4000 SWXL column (Tosoh Biosciences 3.2. Tertiary structure analysis
Corporation, Montgomeryville, PA) with a guard column was used,
and the elution was monitored at 280 nm. A total of 100 ␮L of each No modification or shift of spectrum was observed during the
centrifuged sample (1 mg/mL) were eluted isocratically with a pH six-month storage at 4 ◦ C for the peaks corresponding to pheny-
7.0 buffer [0.1 M sodium sulfate, 0.1 M sodium hydrogen phosphate lalanine (260 nm) and tyrosine/tryptophan (285 nm). However, as
and 0.05% sodium azide (w/v)], at a flow rate of 0.6 mL/min. The showed in Fig. 1, there was a significant structural change after
M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290 285

Table 1
Mean hydrodynamic (nm) diameter of diluted RTX (1 mg/mL) after different storage conditions.

Storage (days) Main peak 2nd Peak 3th Peak PdIa


◦ ◦
4 C, 25 C and up to 6 months 11.18 ± 0.24 − − 0.07 ± 0.030
40 ◦ C 1 month
40 ◦ C 3 months 11.6 ± 0.32 90.04 ± 11.6 600.7 ± 35.7 0.38 ± 0.038
40 ◦ C 6 months 11.32 ± 0.55 72.6 ± 19.5 527.1 ± 148 0.49 ± 0.039
a
Polydispersity index.

1,2

0,8

%fu
0,6
Tm = 71.7 ± 0.1 Tm = 72.5 ± 0.3 °C

0,4

0,2

0
Fig. 1. Second derivative UV spectra of samples obtained immediately after recon- 40 50 60 70 80 90
stitution (red line) or after storage at 40 ◦ C during 1 month (blue line) and 6 months Temperature °C
(black line).The concentration of samples was 0.4 mg/ml. The largest peak position
shifts were observed in the minimum at 299 nm and in 3 maxima at 301, 296 nm and Fig. 2. Thermal denaturation curves of Rtx (1 mg/ml). The evolution of aggregated
269 nm. One nm shift was observed for one minimum and 3 maxima as indicated fractions of Rtx in function of temperature was followed by the determination of
by narrows. (For interpretation of the references to color in this figure legend, the mean diameter using dynamic light scattering. No difference was observed between
reader is referred to the web version of the article.) curves at D0 and D180 at 4 ◦ C. The Tm was not different after 6 months of storage
at 4 ◦ C. After 6 months at 40 ◦ C, the Tm value was lesser than this obtained after
six months at 40 ◦ C. Weak red shifts (1 nm) were observed in the reconstitution. Each curve represents the mean ± SD of 3 separate experiments. To
maximum at 299 nm and in 3 minima at 301, 296 nm and 269 nm, (䊉), T6 months (). At 40 ◦ C: T6 months (). In all cases, a typical sigmoidal plot was
obtained (r2 = 0.99).
respectively, suggesting an alteration of the tertiary structure, and
demonstrating that the method is a good indicator of changes in
stability. in all cases (r2 = 0.99). The Tagg values determined from this
curve at D0 and D180 at 4 ◦ C were not different (72.5 ± 0.3 ◦ C
3.3. Determination of hydrodynamic diameters vs. 72.7 ± 0.7 ◦ C) (Table 2). Nevertheless, at 40 ◦ C, a decrease was
observed (71.7 ± 0.1 ◦ C vs. 72.5 ± 0.3 ◦ C; p > 0.001 vs. 4 ◦ C). The H
The results are reported in Table 1. The initial mean hydrody- observed were always endothermic, in agreement with published
namic diameter of RTX was 11.18 ± 0.24 nm, as expected for the data for a thermal denaturation process. Storage at 40 ◦ C induced
monomeric form of a mAb of about 150 kDa (Matheus et al., 2006). a decrease in enthalpy and entropy parameters of the unfolding
There was no significant change in this value in any of the storage process as compared to storage at 4 ◦ C (Table 2). Moreover, this
conditions up to 180 days at 4 ◦ C or one month at 40 ◦ C. Moreover, decrease was also observed with the mT parameter.
a low polydispersity index (PdI < 0.1) was found. However, the size
distribution by intensity showed two additional peaks at approxi- 3.5. Secondary structure analysis
mately 100 nm and 600 nm for the storage under stress conditions
after 90 days. Moreover, the cumulative distribution showed poly- No modification of the FT-IR spectra was observed after six
dispersity (PdI = 0.49). months at 4 ◦ C. The spectra of D0 and D180 samples were super-
imposable, and no shift was noted (considering both near and
3.4. Thermal denaturation far IR). The similarity coefficient was close to one (0.968 ± 0.016
vs 0.969 ±0.005 at D0 and after 6 months at 4 ◦ C, respectively).
Fig. 2 presents the evolution of aggregated fractions of RTX In Table 3, the percentages of the different secondary structures
at different temperatures. A typical sigmoidal plot was obtained are presented, calculated from the amide I bands. The initial

Table 2
Tm values and thermodynamic parameters of RTX calculated for different storage conditions.

Different storage conditions Tm (◦ C) mT (cal mol−1 K) S (cal mol−1 K−1 ) H (kcal mol−1 )

D0 72.5 ± 0.3 1613 ± 42.7 1433.0 ± 78.6 495.0 ± 26.9


D6 months 4 ◦ C 72.7 ± 0.7 1910 ± 17.4 1262.0 ± 91.3 425.0 ± 41.7
D3 months 40 ◦ C 71.9 ± 0.4 456 ± 13.8 1195.0 ± 35.3 415.0 ± 12.5
D6 months 40 ◦ C 71.7 ± 0.1 266 ± 14.5 1124.0 ± 48.6 387.5 ± 36.8
286 M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290

Table 3
FT-IR deconvoluted amide I band frequencies, assignments and percentage for Rituximab (16 mg/mL) stored in different temperatures in Nacl 0.9%.

RTX solution Conditions Frequencies (cm−1 ) Assignments % spectra band assignment 2nd derivative

Just after dilution ␤ Structures 49.1 ± 2.1


1614, 1621, 1628, Intermolecular ␤-sheet 8.9 ± 1.8
1644, 1689 Intramolecular ␤-sheet 24.6 ± 3.1
1633 ␤-sheet and extended strand 15.6 ± 1.7
1651 Random 20.3 ± 1.8
1659, 1668, 1674, 1681 ␤- turn 30.6 ± 2.2
1656 ␣-helix 0

After 3 months at 4 ◦ C ␤ Structures 51.7 ± 2.8


1614, 1621, 1628 Intermolecular ␤-sheet 7.8 ± 0.7
1644, 1689 Intramolecular ␤-sheet 27.7 ± 4.6
1633 ␤-sheet and extended strand 16.2 ± 0.7
1651 Random 20.9 ± 2.7
1659, 1668, 1674, 1681 ␤- turn 27.4 ± 1.9
1656 ␣-helix 0

After 6 months at 4 ◦ C ␤ Structures 51.1 ± 1.7


1614, 1621, 1628 Intermolecular ␤-sheet 8.0 ± 0.4
1644, 1689 Intramolecular ␤-sheet 29.0 ± 4.4
1633 ␤-sheet and extended strand 14.1 ± 2.0
1651 Random 18.5 ± 0.4
1659, 1668, 1674, 1681 ␤- turn 30.4 ± 7.3
1656 ␣-helix 0

After 6 months at 40 ◦ C ␤ Structures 47.6 ± 2.3


1614, 1621, 1628 Intermolecular ␤-sheet 8.2 ± 5.1
1644, 1689 Intramolecular ␤-sheet 22.9 ± 2.5
1633 ␤-sheet and extended strand 16.5 ± 1.4
1651 Random 24.2 ± 1.8
1659, 1668, 1674, 1681 ␤- turn 28.2 ± 0.6
1656 ␣-helix 0

percent of ␤-sheet was approximately 50%. However, although rt = 1.4 ± 0.05 min was found in all chromatograms corresponding
there was no change in native ␤-sheet content in temperature- to the buffer. In the optimal storage conditions (4 ◦ C), no signifi-
induced aggregates (47.6% ± 2.3 vs. 49.1% ± 2.1; NS), minor cant change in chromatogram profiles was detected, even after six
changes were observed for other ␤-structures: random structure months. There was no significant difference between the AUC of
(24.2% ± 1.8 vs. 20.3% ± 1.8); and ␤-sheet and extended strand the reference sample and the other samples stored at 4 ◦ C up for six
(15.6% ± 1.7 vs. 16.5% ± 1.4; NS), which were non-significant, prob- months, while a decrease in AUC was observed at 40 ◦ C from three
ably because of the low number of samples tested (n = 3). However, months and reaching 45% at six months. Two new peaks were found
far IR shifts were observed as storage duration and tempera- just before the main peak after six months at 40 ◦ C. The retention
ture increased (Fig. 3), explaining the similarity coefficient found times were 2.9 min and 3.3 min, respectively (Fig. 4).
(0.7624 vs. 0.9695). These results indicate that storage of diluted
RTX solutions at 40 ◦ C, but not at 4 ◦ C, induces modifications in RTX
secondary structure. 3.7. Size-exclusion chromatography (SEC)

Only one peak for the RTX reference sample was found. The
3.6. Cation exchange chromatography (CEX) retention time of the reference sample was 19.02 ± 0.01 min,
and the mean AUC was 79.47 ± 3.93 mAU min. The molec-
A major peak with a retention time (rt) = 4.27 ± 0.02 min was ular weight found was 147.5 ± 1.0 kDa, as expected (Vermeer
observed for the reference sample. A minor second peak with a and Norde, 2000a). A second peak, at 22.7 min, was observed

Fig. 3. Second-derivative IR spectra of RTX in the 1700–1600 cm−1 region (amide I bands) obtained immediately after reconstitution (blue line) or after storage at 40 ◦ C
during 6 months (black line). After 6 months of storage, modifications are observed. Changes were observed for ␤-structures (random structure and ␤-sheet and extended
strand). Far IR shifts were also observed. The samples were concentrated to 16 mg/mL using 10 kDa Centricon concentration devices (Millipore).
M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290 287

Fig. 4. Cation-exchange chromatographic profile of RTX (1 mg/ml) using a WCX-10


column at pH 6.0 eluted with a linear salt gradient from 0 to 50% of 1M NaCl and
detection at 280 nm. Decrease in AUC was observed at 40 ◦ C from three months and
reaching 45% at six months. Changes of shape of main peak were also observed. Two
new peaks were found just before the main peak after six months at 40 ◦ C (inset). Fig. 5. Elution profile of RTX (1 mg/ml) chromatography using a TSK-GEL G4000
SWXL column with detection at 280 nm. Samples were centrifuged before injection.
Stressed sample of diluted RTx (1 mg/ml) stored at 40 ◦ C for 6 months exhibiting an
additional peak at about 22 KDa (light chain) as result of structural breakage.
for all conditions corresponding to the buffer. Regardless of
the conditions of storage, the retention time and AUC did not
change up to three months. No significant change in chro- molecular weight of this fragment was estimated to be
matogram profiles was observed for RTX samples stored for 22.1 ± 1.2 kDa. Moreover, after six months at 40 ◦ C, a decrease in
six months at 4 ◦ C. Nevertheless, after six months at 40 ◦ C, a AUC was found (−12.8%), while the decrease only reached 5% after
weak fragmentation was found (Fig. 5) corresponding to an addi- three months at 40 ◦ C. Finally, no additional peak of higher molec-
tional peak eluting after the original form (rt = 20.8 min). The ular weight was detected.

Fig. 6. Peptide map of RTX (1 mg/ml). Samples were digested by trypsin and the generated peptides were analyzed by HPLC on a C18 column in gradient mode at 280 nm. A
difference appeared between samples stored at 4 ◦ C (black line) and at 40 ◦ C (blue line), indicating a chemical degradation of the mAb primary structure. (For interpretation
of the references to color in this figure legend, the reader is referred to the web version of the article.)
288 M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290

Table 4
Cytotoxicity of RTX stored in different conditions on RAJI cells.

D0 4 ◦C 40 ◦ C

D30 D180 D30 D180

Number of experiments 10 2 5 3 2
IC50 (␮g/mL) 137.2 ± 28.18 134.7 ± 21.6 125.8 ± 36.5 331.7 ± 32.2b 573.9a ± 181.1b
IC25 (␮g/mL) 41.9 ± 8.3 45.6 ± 60.56 35.8 ± 11.9 83.5 ± 8.8b 119 ± 33.9b
R2 0.994 ± 0.035 0.986 ± 0.014 0.993 ± 0.031 0.992 ± 0058b 0.996 ± 0.002b
AUC 25,334 ± 2316 25,785 ± 1284 25,077 ± 2451 17,910 ± 329b 14,301 ± 433.2b
a
By extrapolation.
b
p < 0.01 vs D0.

80 (1 mg/mL in saline) stored at 4 ◦ C in polyolefine bags remained sta-


ble for at least six months. No physical or chemical instability was
found, and the biological activity was fully maintained. No visible
70
aggregation was found by turbidimetry. SEC did not show poly-
merization or fragmentation, and retention time and AUC were
60 identical for the control and the solutions stored for six months at
4 ◦ C. No significant difference was found considering the hydrody-
% of inhibition

50 namic diameter of RTX, in all dilutions and storage conditions, up to


three months. No additional peak or decrease in the AUC was found
by CEX. The thermal aggregation curves and their derived ther-
40 modynamic parameters were unchanged, suggesting the absence
of intra- or intermolecular interactions, and a conserved struc-
30 tural integrity without appreciable destabilization. The primary,
secondary and tertiary structures of the protein were not modi-
fied as demonstrated by peptide mapping, and second derivative
20 IR and UV spectroscopy. Finally, the direct cytotoxic effect of RTX
was fully maintained.
10 These results are not surprising, considering the presence of a
significant number of disulfide bonds, intimate domain–domain
interactions and the high content in ␤-sheets (Wang et al., 2007).
0 This gives mAbs a good stability and resistance to moderate ther-
0 100 200 300 400 500
mal stress, as compared to other proteins and revealed by Tm or Tagg
RTX concentration (µg/ml) values higher than in other proteins (range of 70–80 ◦ C vs. range of
40–65 ◦ C for globular protein) (Wang, 1999). Therefore, it is surpris-
Fig. 7. RTX direct cytotoxic on CD20-expressing cells was determined through ing that the manufacturer recommends the use of RTX within 24 h
concentration-dependent cytotoxicity curves (concentration of Rtx: 0, 4.5, 9, 22.5,
after dilution. This recommendation could be only based on bacte-
45, 90, 135, 270 and 450 ␮g/mL) on a human lymphoma cell line (RAJI lymphoma
cells (ATCC number: CCL-86)). Twelve wells were seeded for each concentration riological criteria, because it is not supported by physicochemical
tested. The results were expressed as the percentage of inhibition in comparison and biological studies of instability.
to control (without RTX). The cytotoxic effects of RTX were not different after six Several previous studies demonstrated the good stability of
months of storage at 4 ◦ C.
several mAbs. Liu et al. showed that a high temperature was
necessary to observe clear physicochemical instability (six months
3.8. Peptide mapping at 40 ◦ C) (Liu et al., 2006). Usami et al. found that storage at 37 ◦ C for
14 days did not alter the tested mAb (Usami et al., 1996). Cordoba
No change in chromatographic profile was observed for samples showed that four humanized IgG1 were stable for one month at 5 ◦ C
stored at 4 ◦ C for six months. At D90, a difference appeared between (Cordoba et al., 2005). Finally, Bakri indicated a six-month stabil-
samples stored at 4 ◦ C and at 40 ◦ C (Fig. 6), indicating a chemical ity at 4 ◦ C for diluted bevacizumab (Bakri et al., 2006). However,
degradation of the mAb primary structure. this study was only based on immunoassay methods measur-
ing bevacizumab concentration, and very recently, it has been
3.9. Direct cytotoxicity assay suggested that these analytical methods are insufficient to prove
the stability of these complex molecules (Sreedhara et al., 2011).
A shown in Table 4 and Fig. 7, the cytotoxic effects Indeed, the combination of several orthogonal methods, each hav-
of RTX were not different after six months of storage at ing its own characteristic measuring principle, for example, by size,
4 ◦ C. Indeed, the IC50 and AUC were not different (137.2 ± quantification or structure, should be performed to study the sta-
28.18 ␮g/mL vs. 125.8 ± 36.5 ␮g/mL and 25,334 ± 2316 vs. bility of biological products (Mahler et al., 2009). Therefore, Pabari
25,077 ± 2451 ␮g/mL/min, respectively). In contrast, storage demonstrated that trastuzumab in solution was resistant to vari-
at 40 ◦ C induced a significant decrease in RTX cytotoxic proper- ous processing stresses (sonication, freeze–thawing, lyophilization
ties (IC50 × 4.2) associated with an important decrease in AUC and spray drying) using different orthogonal methods (physico-
(45%), thus demonstrating a severe alteration of the mAb biologic chemical and biological studies) (Pabari et al., 2011). The use of
activity. such orthogonal methods is also suggested in the current European
Medicines Agency (EMA) draft guideline on “Production and quality
4. Discussion control of monoclonal antibodies and related substances” (Mahler
et al., 2009) and also by the recent consensus conference on sta-
The analysis of diluted RTX stability using several different bility studies for anticancer drugs (Bardin et al., 2011). According
analytical and biological methods demonstrated that diluted RTX to these recommendations, all methods employed in this study,
M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290 289

widely used for protein analysis, are appropriate to determine a structural changes, as shown previously (Wan et al., 1999; Harris
beyond-use date. et al., 2001).
To obtain information on the most susceptible degradation The modifications of the tertiary structure, induced by an
pathways and to confirm that the analytical methods were good increase in temperature, is probably due to changes in the degree of
indicators of stability, we exposed RTX under an elevated temper- protonation/deprotonation or ionization of aromatic amino acids,
ature stress condition. We chosen a temperature of 40 ◦ C because such as tyrosine, tryptophan or phenylalanine. In accordance with
it seemed to be a better threshold for studying stability than higher this hypothesis, we observed a 1 nm red shift for three minima
temperatures (Duddu and Dal Monte, 1997). and one maximum, probably due to an electronic differentiation
Physical instability, which is usually the first response to stress process, as previously described by Balestrieri et al. (1980).
of high molecular weight molecules such as mAbs, was observed Considering the secondary structure, we found that the initial
only in stress conditions and after three or six months. Optical den- percentage of ␤-sheets is lower (50% vs 65%–70%) and the per-
sity, which is a good a marker of visible aggregation, remained low centage of random coil is higher (20% vs 5%) than those previously
(0.0614 ± 0.0001 vs. 0.0418 ± 0.0016) even after 40 ◦ C for 6 months, published for IgG by different authors (Dong et al., 1997; Matheus
confirming that aggregation is not the major degradation pathway et al., 2006). This difference could be due to the presence of buffers
for IgG mAbs after thermal stress (Liu et al., 2006), as opposed and surfactant agents in the commercial solution of rituxumab.
to what is observed under mechanical stress (Lahlou et al., 2009; Indeed, the addition of phosphate buffer induces an increase of ran-
Sreedhara et al., 2011). The results of the DLS analysis, which is dom coil (10%) (Vermeer et al., 1998). The adjunction of surfactant
very sensitive in detecting submicron aggregates (also described allows obtaining 50% of ␤ − sheet structure with 20% of random coil
as “subvisible” or “soluble” aggregates), revealed two subpopu- structure (Vermeer and Norde, 2000a) that explains our results.
lations with a 9-fold and 60-fold higher hydrodynamic diameter All RTX modifications observed were due to an increase in tem-
(90.0 ± 11.6 nm and 600.7 ± 35.7 nm), as compared to the reference perature, and it is well known that increased temperature decreases
sample (11.18 ± 0.24 nm) when diluted RTX was stored for 90 days protein stability in thermodynamic terms by favoring unfolded
at 40 ◦ C. These results suggest the presence of subvisible aggregates states and by increasing disorders in the system (Speed et al., 1997;
that were not detected by turbidimetry, in accordance with pre- Wang, 1999). Protein stability results from an equilibrium between
vious studies of protein aggregation under thermal (Vermeer and destabilizing and stabilizing forces. Destabilizing forces are mainly
Norde, 2000b; Hawe et al., 2009) and mechanical stress (Mahler due to a large increase in unfolding entropy (dissipative forces, loss
et al., 2005; Lahlou et al., 2009). SEC also revealed that there was of conformational entropy of native state), and stabilizing forces
no subvisible aggregation after six months at 4 ◦ C. The absence of (intramolecular interactions) are mainly due to non-covalent inter-
aggregation after six-month storage at 4 ◦ C is very important, con- actions
sidering the possible induction of allergic toxicity due to particles or Different parameters can be studied to estimate the degree of
aggregates, even at a very low level, which can be initially present internal or external stabilization, such as the changes in unfolding
or formed during biological drug handling (Sharma, 2007a,b). (Tm ) or aggregation (Tagg ) temperature. As shown in Table 2, the Tagg
A weak fragmentation of the RTX molecule was demonstrated values obtained in our study were as high as those found in the lit-
by SEC (Fig. 5), but only after six months at 40 ◦ C. SEC revealed erature (Tm IgG: 73.6 ◦ C ± 0.4 (Liu et al., 2009), 72.6 ◦ C (Vermeer and
one additional peak downstream of the main peak, corresponding Norde, 2000b), 74 ◦ C (Wang, 1999; Matheus et al., 2006), indicat-
to a lower molecular weight than IgG as previously described by ing a high resistance to temperature (the higher the Tm , the greater
Cordoba (Cordoba et al., 2005). The molecular weight of this frag- the thermal resistance of a protein). The storage of RTX during a
ment was estimated at 22.1 ± 1.2 kDa, which is coherent with the long period of time and at a high temperature (40 ◦ C) led to a slight
release of light chains, as previously reported by Liu (23.5 kDa) after decrease in Tagg (−0.8 ◦ C), suggesting a low but significant alter-
6 months at 40 ◦ C (Liu et al., 2006). ation of the protein integrity, because when the integrity is strongly
Regarding chemical stability after storage at 40 ◦ C, significant altered (breaking of disulfide bonds), the decrease of this parameter
modifications were observed with all methods employed. CEX is is more important (>1 ◦ C) (Wang, 1999).
a useful tool to detect subtle protein alterations such as deamida- We observed that all thermodynamic parameters were posi-
tion or oxidation (Moorhouse et al., 1997). These alterations could tive, thus corresponding to an endothermic process as described
modify the primary and secondary structure of RTX. Indeed, two for other proteins. Because H and G are positive, aggregation
increasing distinct peaks with a shorter retention time (2.9 and is not possible without an increase in temperature. As shown in
3.3 min, respectively) than the reference sample (4.27 min) were Table 2, the initial value of H is high (about 500 kcal mol−1 ) and
observed in samples stored at 40 ◦ C, demonstrating a modification close to values found by Matheus (578 kcal mol−1 ) or Vermeer
of the charge of RTX molecules, and the formation of more acidic (556 kcal mol−1 ) (Vermeer and Norde, 2000b; Matheus et al., 2006).
derivatives, which can reflect both global charge or local charge After a long storage period at 40 ◦ C, a major decrease in all thermo-
modifications, as shown in Fig. 4. Furthermore, according to the lit- dynamic parameters (S, H) is noticed. The extended stability of
erature on deamidation and isomerization, a C-terminal position or diluted RTX stored at 4 ◦ C was finally ascertained by verifying that
succession of Gly, His or Ser, or Asp and Asn residues can influence its cytotoxic properties were maintained. Indeed, a major part of
the protein degradation rate (Wakankar and Borchardt, 2006). Con- the biological activity of RTX is related to its ability to bind CD20-
sidering the primary structure of RTX, both of these conditions are expressing cells. After storage of diluted RTX for six months at 4 ◦ C,
met, as Gly and Ser succeed to Asp inside the heavy chain, and Asp using a human lymphoma cell line, we demonstrated that dose-
is in the C-terminal position inside the light chain. Finally, at pH dependent cytotoxicity curves were fully superimposable with no
5 (pH 6.5 in the RTX formulation), an increase in Asn deamidation difference between DL50 values. In the opposite, RTX maintained in
rates was observed (Manning et al., 2010). stressed condition (30 days at 40 ◦ C) have lost about 50% of its bio-
Furthermore, peptide mapping, which is a sensitive finger- logical activity. Interestingly, this dramatic lost of biological activity
print process method including enzymatic and chemical cleavage, was observed even analytical method indicated indeed detectable
revealed changes in the chromatographic profiles (Fig. 6) with alterations but which seemed limited. These results suggest that if
four additional peaks located within the 70–110-min elution zone. several analytical methods are mandatory to assess long-term sta-
These additional peaks could be the result of Asn or Gln deami- bility of mAbs, they should be completed by an estimation of their
dation (Liu et al., 2006), and could induce conformational and/or biological activity.
290 M. Paul et al. / International Journal of Pharmaceutics 436 (2012) 282–290

5. Conclusions Kong, J., Yu, S., 2007. Fourier transform infrared spectroscopic analysis of protein
secondary structures. Acta Biochim. Biophys. Sin (Shanghai) 39, 549–559.
Kueltzo, L.A., Wang, W., Randolph, T.W. Carpenter, J.F. 2007. Effects of solution condi-
Our results demonstrate that diluted RTX (1 mg/mL in saline) tions, processing parameters, and container materials on aggregation of a mon-
remains stable for at least six months when stored in polyolefin oclonal antibody during freeze–thawing, http://dx.doi.org/10.1002/jps.21110,
bags at 4 ◦ C. Indeed, no modification of its chemical, physical and 1-12.
Lahlou, A., Blanchet, B., Carvalho, M., Paul, M., Astier, A., 2009. Mechanically-induced
structural characteristics and no aggregation were observed. This aggregation of the monoclonal antibody cetuximab. Ann. Pharm. Fr. 67, 340–352.
conclusion is supported by the results of several orthogonal meth- Liu, C., Vailhe, C., Samuel, N., Min, Q., 2009. A simple and reproducible approach to
ods, as recommended in the stability assessment of biological characterize protein stability using rheology. Int. J. Pharm. 374, 1–4.
Liu, H., Gaza-Bulseco, G., Sun, J., 2006. Characterization of the stability of a fully
products. Moreover, the pharmacological properties of RTX were
human monoclonal IgG after prolonged incubation at elevated temperature. J.
fully conserved. These results indicate that the time frame for the Chromatogr. B: Analyt. Technol. Biomed. Life Sci. 837, 35–43.
practical use of diluted RTX can be safely extended, allowing, for Mahler, H.C., Friess, W., Grauschopf, U., Kiese, S., 2009. Protein aggregation: path-
ways, induction factors and analysis. J. Pharm. Sci. 98, 2909–2934.
example, early reconstitution of workload optimization or safe
Mahler, H.C., Muller, R., Friess, W., Delille, A., Matheus, S., 2005. Induction and analy-
storage of a patient’s bags for several days if the RTX cannot be sis of aggregates in a liquid IgG1-antibody formulation. Eur. J. Pharm. Biopharm.
immediately administered. Our results also support the possibility 59, 407–417.
of preparing standardized batches of “ready-to-use” diluted RTX Manning, M.C., Chou, D.K., Murphy, B.M., Payne, R.W., Katayama, D.S., 2010. Stability
of protein pharmaceuticals: an update. Pharm. Res. 27, 544–575.
under good manufacturing procedures to ensure maximal secu- Matheus, S., Friess, W., Mahler, H.C., 2006. FTIR and nDSC as analytical tools for
rity in terms of sterile conditions and quality controls. Finally, this high-concentration protein formulations. Pharm. Res. 23, 1350–1363.
extended in-use stability can help pharmacists to optimize hospital Moorhouse, K.G., Nashabeh, W., Deveney, J., Bjork, N.S., Mulkerrin, M.G., Ryskamp, T.,
1997. Validation of an HPLC method for the analysis of the charge heterogeneity
expenses due to this costly drug. of the recombinant monoclonal antibody IDEC-C2B8 after papain digestion. J.
Pharm. Biomed. Anal. 16, 593–603.
Nicolaides, N.C., Sass, P.M., Grasso, L., 2006. Monoclonal antibodies: a
morphing landscape for therapeutics. Drug Dev. Res. 67, 781–789,
References http://dx.doi.org/10.1002/ddr.
Pabari, R.M., Ryan, B., Mccarthy, C., Ramtoola, Z., 2011. Effect of microencapsula-
Bakri, S.J., Snyder, M.R., Pulido, J.S., Mccannel, C.A., Weiss, W.T., Singh, R.J., 2006. Six- tion shear stress on the structural integrity and biological activity of a model
month stability of bevacizumab (Avastin) binding to vascular endothelial growth monoclonal antibody, trastuzumab. Pharmaceutics 3, 510–524.
factor after withdrawal into a syringe and refrigeration or freezing. Retina 26, Pedersen, I.M., Buhl, A.M., Klausen, P., Geisler, C.H., Jurlander, J., 2002. The chimeric
519–522. anti-CD20 antibody rituximab induces apoptosis in B-cell chronic lympho-
Balestrieri, C., Colonna, G., Giovane, A., Irace, G., Servillo, L., 1980. Second-derivative cytic leukemia cells through a p38 mitogen activated protein-kinase-dependent
spectroscopy of proteins: studies on tyrosyl residues. Anal. Biochem. 106, 49–54. mechanism. Blood 99, 1314–1319.
Bardin, C., Astier, A., Vulto, A., Sewell, G., Vigneron, J., Trittler, R., Daouphars, M., Schellekens, H., 2002. Bioequivalence and the immunogenicity of biopharmaceuti-
Paul, M., Trojniak, M., Pinguet, F., 2011. Guidelines for the practical stability cals. Nat. Rev. Drug Discov. 1, 457–462.
studies of anticancer drugs: a European consensus conference. Ann. Pharm. Fr. Seefeldt, M.B., Kim, Y.S., Tolley, K.P., Seely, J., Carpenter, J.F., Randolph, T.W., 2005.
69, 221–231. High-pressure studies of aggregation of recombinant human interleukin-1
Cheung, M.C., Haynes, A.E., Meyer, R.M., Stevens, A., Imrie, K.R., 2007. Rituximab in receptor antagonist: thermodynamics, kinetics, and application to accelerated
lymphoma: a systematic review and consensus practice guideline from Cancer formulation studies. Protein Sci. 14, 2258–2266.
Care Ontario. Cancer Treat. Rev. 33, 161–176. Servillo, L., Colonna, G., Balestrieri, C., Ragone, R., Irace, G., 1982. Simultaneous deter-
Chi, E.Y., Krishnan, S., Randolph, T.W., Carpenter, J.F., 2003. Physical stability of pro- mination of tyrosine and tryptophan residues in proteins by second-derivative
teins in aqueous solution: mechanism and driving forces in nonnative protein spectroscopy. Anal. Biochem. 126, 254–257.
aggregation. Pharm. Res. 20, 1325–1336. Sharma, B., 2007a. Immunogenicity of therapeutic proteins. Part 1: impact of product
Chumsae, C., Gaza-Bulseco, G., Sun, J., Liu, H., 2007. Comparison of methionine oxi- handling. Biotechnol. Adv. 25, 310–317.
dation in thermal stability and chemically stressed samples of a fully human Sharma, B., 2007b. Immunogenicity of therapeutic proteins. Part 3: impact of man-
monoclonal antibody. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 850, ufacturing changes. Biotechnol. Adv. 25, 325–331.
285–294. Speed, M.A., King, J., Wang, D.I., 1997. Polymerization mechanism of polypeptide
Cordoba, A.J., Shyong, B.J., Breen, D., Harris, R.J., 2005. Non-enzymatic hinge region chain aggregation. Biotechnol. Bioeng. 54, 333–343.
fragmentation of antibodies in solution. J. Chromatogr. B: Anal. Technol Biomed Sreedhara, A., Glover, Z.K., Piros, N., Xiao, N., Patel, A., Kabakoff, B., 2011. Stability of
Life Sci 818, 115–121. IgG1 monoclonal antibodies in intravenous infusion bags under clinical in-use
Dick Jr., L.W., Mahon, D., Qiu, D., Cheng, K.C., 2009. Peptide mapping of therapeutic conditions. J. Pharm. Sci. 101, 21–30.
monoclonal antibodies: improvements for increased speed and fewer artifacts. Thirumangalathu, R., Krishnan, S., Brems, D.N., Randolph, T.W., Carpenter, J.F., 2006.
J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 877, 230–236. Effects of pH, temperature, and sucrose on benzyl alcohol-induced aggregation
Dong, A., Kendrick, B., Kreilgard, L., Matsuura, J., Manning, M.C., Carpenter, J.F., 1997. of recombinant human granulocyte colony stimulating factor. J. Pharm. Sci. 95,
Spectroscopic study of secondary structure and thermal denaturation of recom- 1480–1497.
binant human factor XIII in aqueous solution. Arch. Biochem. Biophys. 347, Usami, A., Ohtsu, A., Takahama, S., Fujii, T., 1996. The effect of pH, hydrogen perox-
213–220. ide and temperature on the stability of human monoclonal antibody. J. Pharm.
Duddu, S.P., Dal Monte, P.R., 1997. Effect of glass transition temperature on the sta- Biomed. Anal. 14, 1133–1140.
bility of lyophilized formulations containing a chimeric therapeutic monoclonal Vermeer, A.W., Bremer, M.G., Norde, W., 1998. Structural changes of IgG induced by
antibody. Pharm. Res. 14, 591–595. heat treatment and by adsorption onto a hydrophobic Teflon surface studied by
Edwards, J.C., Szczepanski, L., Szechinski, J., Filipowicz-Sosnowska, A., Emery, P., circular dichroism spectroscopy. Biochim. Biophys. Acta 1425, 1–12.
Close, D.R., Stevens, R.M., Shaw, T., 2004. Efficacy of B-cell-targeted therapy with Vermeer, A.W., Norde, W., 2000a. The influence of binding of low molecular weight
rituximab in patients with rheumatoid arthritis. N. Engl. J. Med. 350, 2572–2581. surfactants on the thermal stability and secondary structure of IgG. Colloids Surf.
Farruggia, B., Garcia, G., D’angelo, C., Pico, G., 1997. Destabilization of human serum A 161, 139–150.
albumin by polyethylene glycols studied by thermodynamical equilibrium and Vermeer, A.W., Norde, W., 2000b. The thermal stability of immunoglobulin: unfold-
kinetic approaches. Int. J. Biol. Macromol. 20, 43–51. ing and aggregation of a multi-domain protein. Biophys. J. 78, 394–404.
Harn, N., Allan, C., Oliver, C., Middaugh, C.R., 2007. Highly concentrated monoclonal Wakankar, A.A., Borchardt, R.T., 2006. Formulation considerations for proteins sus-
antibody solutions: direct analysis of physical structure and thermal stability. J. ceptible to asparagine deamidation and aspartate isomerization. J. Pharm. Sci.
Pharm. Sci. 96, 532–546. 95, 2321–2336.
Harris, R.J., Kabakoff, B., Macchi, F.D., Shen, F.J., Kwong, M., Andya, J.D., Shire, S.J., Wan, M., Shiau, F.Y., Gordon, W., Wang, G.Y., 1999. Variant antibody identification
Bjork, N., Totpal, K., Chen, A.B., 2001. Identification of multiple sources of charge by peptide mapping. Biotechnol. Bioeng. 62, 485–488.
heterogeneity in a recombinant antibody. J. Chromatogr. B: Biomed. Sci. Appl. Wang, W., 1999. Instability, stabilization, and formulation of liquid protein pharma-
752, 233–245. ceuticals. Int. J. Pharm. 185, 129–188.
Hawe, A., Kasper, J.C., Friess, W., Jiskoot, W., 2009. Structural properties of mono- Wang, W., 2005. Protein aggregation and its inhibition in biopharmaceutics. Int. J.
clonal antibody aggregates induced by freeze–thawing and thermal stress. Eur. Pharm. 289, 1–30.
J. Pharm. Sci. 38, 79–87. Wang, W., Singh, S., Zeng, D.L., King, K., Nema, S., 2007. Antibody structure, instabil-
Hermeling, S., Crommelin, D.J., Schellekens, H., Jiskoot, W., 2004. Structure- ity, and formulation. J. Pharm. Sci. 96, 1–26.
immunogenicity relationships of therapeutic proteins. Pharm. Res. 21, 897–903. Yang, M.X., Shenoy, B., Disttler, M., Patel, R., Mcgrath, M., Pechenov, S., Margolin, A.L.,
Hwang, W.Y., Foote, J., 2005. Immunogenicity of engineered antibodies. Methods 36, 2003. Crystalline monoclonal antibodies for subcutaneous delivery. Proc. Natl.
3–10. Acad. Sci. U. S. A. 100, 6934–6939.
Jiskoot, W., Beuvery, E.C., De Koning, A.A., Herron, J.N., Crommelin, D.J., 1990. Ana- Zheng, J.Y., Janis, L.J., 2006. Influence of pH, buffer species, and storage temperature
lytical approaches to the study of monoclonal antibody stability. Pharm. Res. 7, on physicochemical stability of a humanized monoclonal antibody LA298. Int. J.
1234–1241. Pharm. 308, 46–51.

Das könnte Ihnen auch gefallen