throbber
iMH·ihi
`
`AIChE
`
`~ WILEY
`lnterScience•
`
`Engineering Challenges of
`Protein Formulations
`
`Theodore W. Randolph
`Dept. of Chemical and Biological Engineering, University of Colorado, Boulder, CO 80309
`
`John F. Carpenter
`Dept. of Pharmaceut ical Sciences, University of Co lorado Healt h Sciences Center, Denver, CO 80262
`
`DOI I0.!002/aic.11252
`Published online June 25, 2007 in Wiley lnterScience (www.interscience.wiley.com).
`
`Keywords: protein stability, protein aggregation, shelf life, accelerated degradation
`
`Introduction
`
`Protein based pharmaceuticals are the fastest growing
`
`class of new drugs. They not only offer promise for
`treatments to address major health challenges, such as
`cancer, but also a wealth of new engineering problems to
`solve. Chemical engineers have long been proficient at pro
`ducing products that meet exacting specifications for chemical
`purity, but therapeutic proteins now bring additional chal
`lenges: these products must not only be highly chemically
`pure, but also conformationally pure, and must remain so dur
`ing manufacturing and through the drug's entire shelf life and
`delivery to patients.
`Proteins degrade through a variety of mechanisms. These are
`usually classified as either physical or chemical degradation
`pathways. 1 Physical degradation pathways include unfolding,
`misfolding, and aggregation of the protein molecules. Chemical
`degradation pathways encompass a myriad of unwanted
`chemical reactions that proteins commonly undergo, such as
`oxidation, dearnidation, racemization, hydrolysis, disulfide
`exchange, and carbamylation. Classification of degradation
`pathways as physical or chemical is somewhat artificial,
`because the two types of degradation often are closely linked.
`For example, we have shown that an oxidation process result
`ing in crosslinking of tyrosine residues in a synuclein ( a protein
`that forms characteristic fibrils in Parkinson's disease), is a pre
`cursor to aggregation and fibrillogenesis.2 Addition of radical
`scavenging molecules, such as methionine to a synuclein for
`mulations delays onset of in vitro fibril formation by reducing
`the rate of tyrosine oxidation formation. Conversely, oxidation
`of the serine protease subtilisin can be inhibited by adding for
`mulation excipients, such as sucrose that act to increase the
`thermodynamic conformational stability of the protein. 3
`
`Correspondence concerning this article should be addre~ ed to T. W. Randolph
`at lheodore.randolpl-@colorado.edu.
`
`© 2007 American Institute of Chemical Engineers
`
`Formulation Challenge
`
`To allow proteins to be used as therapeutic agents, proteins
`must be placed in a formulation that confers suitable stability
`against physical and chemical degradation. In addition to sta
`bilizing the pharmaceutically active protein ingredients, for
`mulation components, or excipients, also must be compatible
`with their intended use. For example, a formulation intended
`for parenteral use (e.g., subcutaneous injection) must be ster
`ile, nontoxic, and exhibit acceptable viscosity and tonicity.
`Although these requirements place limits on the types and
`concentrations of excipients that practically can be used, there
`are still far too many possible sets of formulations to allow a
`purely empirical screening approach to be successful.
`For economic viability, therapeutic protein formulations
`typically require a shelf life of 18 24 months.4 Over the
`course of this time, when stored as directed on the product
`label (usually refrigerated at 2 8 °C), the protein must retain
`adequate chemical and conformational purity. Meeting the
`stringent requirements for stability during shelf life is a daunt
`ing task. Most of the common chemical degradation products
`(especially hydrolysis and oxidation byproducts) are signifi
`cantly thermodynamically favored vs. the desired native state
`of the protein. Furthermore, the properly folded native state of
`most proteins is only marginally more stable (the free energy
`of unfolding .1Gumf, is about 20 (,() kJ/mol) than the folded
`state,5 and appears to be unstable under most conditions with
`7
`respect to aggregated forms of the protein.6
`•
`Required chemical and conformational purity levels are die
`tated by the individual protein's safety and efficacy profile,
`but frequently levels of chemical impurities >5%, or confor
`mational impurities > l % at the end of the labeled shelf life
`might be considered unacceptable. Given that typical concen
`trations of protein in therapeutic formulations are near
`10 micromolar, this suggests that levels of degradation prod
`ucts typically must be held below 100 nanomolar over the
`course of two year storage. Thus, an average rate of degrada
`
`1902
`
`August 2007 Vol. 53, No. 8
`
`AIChE Journal
`
`Bausch Health Ireland Exhibit 2011, Page 1 of 6
`Mylan v. Bausch Health Ireland - IPR2022-01104
`
`

`

`tion of 1 nanomolar/week may indicate an unacceptable level
`of product lability.
`As part of the approval process for protein therapeutics, the
`US Food and Drug Administration requires that protein drug
`stability be demonstrated in real time, under conditions mim
`icking the proposed labeled storage conditions, i.e., in the pro
`posed container/closure system, at the proposed protein con
`centration in a final formulation and under specified tempera
`ture conditions. This requires that enormous resources be
`dedicated years before a product can be sold, and represents a
`bottleneck for the entire therapeutic protein development pro
`cess. For products with anticipated annual sales often of more
`than $500M, delays in development of suitably stable formu
`lations may, thus, represent lost sales of $1M or more per day.
`Clearly, a goal is to make sure that the formulations that enter
`real time stability testing have a high probability of success.
`For many kinds of protein degradation, acceptable levels of
`degradation products at the end of shelf life are very low. This
`creates a quandary. We wish to be reasonably sure at the onset
`of a real time stability study that at the study’s completion 18
`or 24 months later we will have acceptably low levels of deg
`radation products. However, in part, because of analytical lim
`itations, it is difficult or impossible to conduct a relatively
`short (e.g., one week) experiment under proposed storage con
`ditions that can be extrapolated to an 18 or 24 month storage
`life. Thus, in order to predict which set of excipients are likely
`to provide an adequate shelf life, accelerated degradation
`experiments must be conducted. In these experiments, formu
`lations are subjected to an additional ‘‘stress’’, such as ele
`vated temperature, freeze thawing, the presence of air water
`interfaces or high or low ionic strengths, and the kinetics of
`protein degradation measured. Combinations of excipients
`that protect against degradation under ‘‘stressed’’ conditions
`are then assumed to be most likely to confer stability under
`more benign storage conditions. Usually, the result of such
`studies is a formulation that provides the greatest relative sta
`bility. However, there is no assurance that this level of stabili
`zation will be sufficient for the shelf life.
`By definition, accelerated degradation studies are conducted
`under conditions that are different from anticipated actual
`storage conditions. How predictive are these studies of protein
`behavior at actual storage conditions? fortunately, the answer
`(at least for relatively simple accelerated stability studies) is
`often ‘‘not very’’. Discrepancies between the predictions of
`simple accelerated studies and actual behavior might not be
`surprising, given the complicated structure of proteins, and the
`likely presence of multiple degradation pathways, but they
`serve to emphasize the need for better models and mechanistic
`understanding of protein degradation. Significant progress has
`been made in the ‘‘rational design’’ of protein formulations,4,8,
`but there remain lacunae in the mechanistic understanding of
`the protein degradation pathways, and their responses to accel
`erated stability protocols. Some selected examples of chal
`lenges and progress in the protein formulation arena follow.
`
`Thermally accelerated protein degradation
`
`Elevated temperature is perhaps the most obvious ‘‘stress’’
`condition that might be applied to accelerate protein degrada
`tion. Intuitively, one might expect that a simple Arrhenius
`analysis might allow data obtained on protein degradation at
`
`elevated temperatures (e.g., rate constants for protein aggrega
`tion measured at 50 80 8C) to be extrapolated to typical refri
`gerated storage conditions. However, often it is found empiri
`cally that predictions made using such an approach are poor.
`A reason for the observed discrepancy between predictions of
`degradation rates, based on simple Arrhenius analysis of ther
`mally accelerated stability studies and actual behavior, lies in
`the coupling of thermodynamic equilibria between various
`protein conformations, each with a characteristic reactivity for
`a given degradation pathway and the kinetics of reactions on
`that pathway. Some protein degradation pathways, notably
`those leading to aggregation, occur through partially unfolded
`intermediates or through reactive subpopulations of the pro
`tein’s native state ensemble. Because of the marginal confor
`mational stability of proteins, relatively small changes in tem
`perature can significantly perturb conformational stability,
`which in turn alters the population of aggregation prone pro
`tein molecules. For example, Roberts9 has shown that the rate
`of aggregation of recombinant bovine granulocyte colony
`stimulating factor as a function of temperature, shows strik
`ingly non Arrhenius behavior and a simple prediction of
`shelf life based on a simple Arrhenius extrapolation of data
`taken above room temperature to 5 8C storage conditions,
`would lead to an overestimation of shelf life by orders of
`magnitude. However, when the temperature dependency of
`recombinant bovine granulocyte colony stimulating factor
`conformational stability, and its effect on the observed rate
`constants were taken into account, the underlying rate con
`stants for aggregation showed Arrhenius behavior.9
`A commonly used approach to screen excipients for protein
`formulations is to use differential scanning calorimetry to
`measure the apparent ‘‘melting temperature’’ Tm, or tempera
`ture at which the protein olds in a given formulation. Formula
`tions that yield elevated values are those in which the protein
`is presumed to be the most stable under real time storage con
`ditions.10 However, there are some excipients (notably non
`ionic surfactants and some preservatives) that lower values,
`but have little detrimental effect or act to increase stability at
`lower temperature storage conditions. For example,
`in the
`presence of the preservative benzyl alcohol, the apparent of
`recombinant human interleukin 1 receptor antagonist
`is
`depressed by about 8 8C, and the protein aggregates rapidly at
`37 8C. However, little effect of benzyl alcohol is seen on the
`aggregation rate 25 8C.11 In part, the contradiction can be
`explained on the basis of the temperature dependency of
`hydrophobic interactions, which are strengthened at higher
`temperatures. At higher temperatures, increased hydrophobic
`interactions favor binding of preservative to the exposed
`hydrophobic regions of olded protein molecules, which,
`according to the Wyman linkage function,12 should result in a
`greater population of olded species, and, hence, a lower Tm
`than in the absence of preservative.11
`Protein aggregation frequently appears to result from multi
`step and/or multipathway reactions.13–15 Because the activa
`tion energies for each step in the reaction pathway may be dif
`ferent, the rate limiting step for the degradation process may
`change with temperature. In the case of proteins that exhibit
`multiple degradation pathways,
`the dominant degradation
`product that is formed during storage at refrigerated condi
`tions may be different than that formed at temperatures used
`for accelerated stability studies. For example, when stored at
`
`AIChE Journal
`
`August 2007 Vol. 53, No. 8
`
`Published on behalf of the AIChE
`
`DOI 10.1002/aic
`
`1903
`
`Bausch Health Ireland Exhibit 2011, Page 2 of 6
`Mylan v. Bausch Health Ireland - IPR2022-01104
`
`

`

`interleukin 1 receptor
`temperatures near room temperature,
`antagonist forms irreversible, soluble aggregates with nearly
`native secondary structure that are crosslinked through disul
`fide bonds.16 In contrast, under accelerated degradation condi
`tions at 55 8C, the protein forms aggregates that are not cross
`linked through disulfides, but that contain significant non
`native b sheet structures.17 Clearly, in this case extrapolations
`of protein aggregation kinetics based on high temperature
`studies would not be expected to be predictive of low temper
`ature storage behavior.
`Pressure is a variable that may provide a useful alternative
`to temperature for accelerated stability studies. Analogous to
`using Arrhenius plots to determine activation energies, semilo
`garithmic plots of reaction rate constants vs. pressure may be
`used to determine the activation volume for a reaction. Acti
`vation volumes may then be used to predict rate constants at
`other pressures of interest. Webb et al.18 measured activation
`volumes for aggregation of interferon g, and found that the
`volume change required for aggregation 41 mL/mol, was
`unfolding of interferon g (209 mL/mol), suggesting that the
`reactive species involved in the aggregation of interferon g is
`a partially unfolded,19 rather than a completely olded spe
`cies.18 Additional measurements of folding equilibria and
`aggregation kinetics made as a function of temperature and
`surface tension also show that the transition state for aggrega
`tion of interferon g is partially, rather than fully unfolded
`(Figure 1). In the case of interleukin 1 receptor antagonist,
`activation volumes for aggregation were nearly identical to
`those required to old the protein, suggesting that a nearly com
`pletely olded species was required for aggregation. Interest
`ingly, the degradation products were disulfide bonded dimers
`similar to those seen after long term storage at atmospheric
`pressure and temperatures near room temperature.17
`
`only about 20% of the volume change required for complete
`
`N
`
`Reaction Coordinate
`
`Figure 1. Reaction coordinate for aggregation of interferon-
`at 32 8C.
`Interferon g, a protein whose native state is a homo
`dimer, unfolds and aggregates rapidly upon dissocia
`tion into monomers. When the transition state is
`formed from the native state,
`the protein’s partial
`molar volume (magenta) decreases by 41 mL/mol, the
`partial molar solvent exposed surface area increases
`(black) by 3.5 nm2/molecule, and the associated activa
`tion energy Ea (blue) is 130 kJ/mol. In contrast, com
`plete dissociation and unfolding of the native state
`dimer
`is associated with a partial molar volume
`decrease of
`209 mL/mol, a partial molar solvent
`exposed surface area increase of 12.7 nm2/molecule, a
`free energy change of 27.2 kJ/mol, and an enthalpy
`change DH of 460 kJ/mol.1
`
`Mechanisms of Protein Aggregation
`
`Proteins are highly susceptible to the formation of non
`native aggregates and precipitates.20,21 Irreversible, non native
`protein aggregation is a ubiquitous concern for biopharma
`ceuticals and biotechnology products,22 because the biological
`activity of a protein in an aggregate is usually greatly reduced.
`More importantly, non native protein aggregates can cause
`adverse reactions in patients, ranging from immune response
`to anaphylactic shock and even death.23–25 Adverse responses
`to aggregates of a given protein cannot be predicted, nor can
`the maximal level of aggregates that can be safely tolerated
`be determined without costly and time consuming clinical tri
`als.4 Unfortunately, the link between immunogenicity and pro
`tein aggregates is often not discovered until side effects are
`noted, following either long term administration or increases
`in the patient population size after
`the drug has been
`approved. Thus, it is essential during product development to
`design proteins and protein formulations that minimize protein
`aggregation.
`Protein aggregates generally exhibit secondary structures
`that are rich in b sheet structures, and that are dramatically
`perturbed from the protein’s native secondary structure.26 Pro
`tein aggregation rates depend strongly on protein conforma
`tion,27 and even relatively small perturbations in protein struc
`ture can be sufficient to form transition state species on the
`aggregation pathway.18,19 The kinetics of protein aggregation
`are controlled by both the concentration and the reactivity of
`these partially unfolded, transient intermediate species. If we
`assume that the free energy change associated with partial
`unfolding to form an aggregation competent transition state
`(DG*) is of the same order of magnitude as that for complete
`unfolding (20 60 kJ/mol), on average fewer than 1/10,000 of
`the protein molecules exist in the transition state, or about 3
`nM at typical formulation conditions. This creates an experi
`mental challenge, because the concentrations of these transient
`species are too low for direct measurement.
`Although the properties of proteins in the aggregation
`competent transition state cannot be accessed spectroscopi
`cally, some insight into how formulation conditions affect
`aggregation rates can be gained by making two assumptions.
`The first assumption is that DG* and DGumf, are positively
`correlated. Thus, measurements of excipient effects on DGumf,
`which can be made using various calorimetric and spectro
`scopic techniques.28 should allow at least a qualitative predic
`tion of excipient effects on DG*, and excipients that stabilize
`the native state against unfolding and increase DGumf, should
`also reduce the equilibrium concentrations of partially
`unfolded aggregation competent species. A second assump
`tion is that the dominant protein protein interactions between
`native proteins are similar to those between protein molecules
`in the transition state. Protein protein interactions can be
`quantified by measurement of the osmotic second virial coeffi
`cient B22.14 Large, positive B22 values reflect net repulsive
`interactions between protein molecules. Formulation strat
`egies may, thus, be designed so as to reduce protein aggrega
`tion by adding using solution conditions (e.g., pH), and exci
`pients that increase DGumf and/or B22 values.14 Examples of
`protein formulations that have been stabilized by addition of
`agents that increased DGumf include stabilization of acidic
`fibroblast growth factor,29 and recombinant keratinocyte
`
`1904
`
`DOI 10.1002/aic
`
`Published on behalf of the AIChE
`
`August 2007 Vol. 53, No. 8
`
`AIChE Journal
`
`Bausch Health Ireland Exhibit 2011, Page 3 of 6
`Mylan v. Bausch Health Ireland - IPR2022-01104
`
`

`

`growth factor30 by polyanionic excipients, stabilization of
`recombinant human growth hormone and recombinant human
`nerve growth factor by addition of zinc,31 and stabilization of
`recombinant human interferon g by addition of sucrose.27
`Commercial formulations of recombinant human granulocyte
`colony stimulating factor, in contrast, minimize aggregation
`by adopting the strategy of maximizing protein protein repul
`sive interactions by formulating at low pH, where B22 values
`are large and positive.14
`
`Formulations at High-Protein
`Concentrations
`
`Many of the first generation of recombinant protein ther
`erythropoietin and interferon g,
`apeutics,
`such as
`are
`extremely potent molecules
`that
`required only minute
`amounts of protein per dose. For example, erythropoietin is
`administered in a dosage form containing about 60 mg/mL
`erythropoietin. In contrast, newer, antibody based products
`are less potent, and, hence, require much larger doses. For
`1
`example, the monoclonal antibody Herceptin
`is sold in a
`vial containing 440 mg protein. The requirement for nearly
`10,000 fold increases in protein dosages, combined with
`practical limitations on the volume (<1.5 mL) that can be
`delivered in a single subcutaneous dose has led to the need
`to develop formulations that are highly concentrated in
`protein.
`Development of these formulations poses a number of seri
`ous obstacles to commercialization.32,33 Although protein con
`centrations rarely exceed 10 mM, even in highly concentrated
`formulations, due to the relatively large molecular weight of
`proteins,
`this may represent a substantial volume fraction
`(10 15%) of the formulation. Solution nonidealities caused
`by protein protein interactions in these solutions may result
`in undesirably high solution viscosities,32 opalescence,34 and
`increased rates of aggregation.35 High viscosities are problem
`atic because they can make manufacturing operations, such as
`filtration impractical, or limit the ability to deliver doses via
`narrow bore syringe needles. Opalescence, although not nec
`essarily harmful in itself, compromises the ability to detect
`product aggregation or particulate contamination within a vial,
`and creates difficulties during clinical trials, because of the
`lack of availability of opalescent placebo solutions. Most com
`mon analytical techniques used to examine protein protein
`interactions have been developed for use with much lower
`protein concentrations, and so current understanding of the
`interactions that cause high viscosities or opalescence concen
`tration is limited in part by the lack of appropriate analytical
`technologies.
`In recent unpublished studies, we have shown that sim
`ple Carnahan Starling hard sphere models36 of protein ac
`tivity coefficients accurately predict the protein concentra
`tion dependence of apparent rate constants for dimerization
`of recombinant human interleukin 1 receptor antagonist. In
`contrast, the effect of protein concentration on solution vis
`cosities for the same protein are poorly predicted from
`hard sphere models. A detailed understanding of the pro
`tein protein interactions that cause viscosity to vary from
`those predicted form hard sphere models is not available at
`this time.
`
`Heterogeneous Nucleation during
`Processing and in Final
`Product Containers
`
`Even under solution conditions where protein physical sta
`bility appears to be optimized so as to minimize protein
`aggregation in the bulk solution there can be formation of
`visible and subvisible protein particles that may constitute
`only a minute fraction of the total protein population. The
`presence of even a small number of protein particles can
`render a product clinically unacceptable. Particle formation
`can occur routinely during processing steps, such as pumping
`of protein solution during vial/syringe filling. In other cases,
`particle formation may appear to be random. For example,
`particles will be seen in a small fraction of vials or prefilled
`syringes in a given product lot. Unfortunately, these particles
`formed during vial filling operations appear downstream of
`sterile filtration steps and practically cannot be removed by fil
`tration.
`We hypothesize that protein particle formation can arise
`from heterogeneous nucleation of protein aggregation on
`the surfaces of microparticles of foreign materials. These
`particulate contaminants can include metals shed from vial
`filling pumps, tungsten microparticles produced during man
`ufacture of glass syringes, and glass microparticles shed
`from vials as a result of high temperature depyrogenation
`procedures. Although such particles and the protein aggre
`gates that we hypothesize result from them are ubiquitous,
`virtually no systematic characterization of the problem, and
`the mechanisms governing it have been addressed in the
`literature. Aggregation at microparticle surfaces has been
`studied only for a limited number of surfaces and proteins,
`and under an even more limited range of solution condi
`tions, with few of the tested conditions being relevant for
`parenteral formulations of therapeutics. Furthermore, only
`two published studies have focused on therapeutic pro
`teins,37,38 and none have focused on monoclonal antibodies,
`which are the largest class of therapeutic products currently
`being tested clinically.
`It should be noted that for a commercial pharmaceutical
`product, usually it is not economically practical to eliminate
`the risk of heterogeneous nucleation by re engineering the sur
`face properties of containers, pumps or tubing to completely
`eliminate shedding of particles, or by re engineering the pro
`tein to reduce interactions with a surface. In fact in a recent
`review chapter Akers and Nail state, ‘‘Regardless of the qual
`ity of glass, the reputation of the manufacturer, the method of
`manufacture, or the method of cleaning, glass particulates are
`unavoidable.’’39 Thus, development of a safe, effective formu
`lation of a protein therapeutic depends on determining solu
`tion conditions that prevent the interactions of proteins with
`microparticulate contaminants that nucleate formation of pro
`tein particles, while still maintaining protein stability in bulk
`solution. However, current commercial formulation develop
`ment8 does not include testing for heterogeneous nucleation
`during processes, such as vial filling. This problem is not triv
`ial. For example, relatively high concentrations of nonionic
`surfactant may reduce protein binding to contaminants, but
`could also foster unacceptably rapid aggregation of the protein
`in bulk solution.
`
`AIChE Journal
`
`August 2007 Vol. 53, No. 8
`
`Published on behalf of the AIChE
`
`DOI 10.1002/aic
`
`1905
`
`Bausch Health Ireland Exhibit 2011, Page 4 of 6
`Mylan v. Bausch Health Ireland - IPR2022-01104
`
`

`

`Numerous studies have investigated the effects of surface
`chemistry on protein adsorption (see, for example40, and refer
`ences therein). A major motivation of these earlier studies has
`been trying to understand the roles of protein interactions with
`surfaces associated with implantable medical devices. Conse
`quently, the solution conditions for these investigations have
`generally been chosen to mimic physiological conditions, e.g.,
`phosphate buffered saline, and the surfaces tested have often
`been those characteristic of implantable devices or natural sur
`faces found in vivo, such as bone. In contrast, for formulations
`of therapeutic proteins, solution conditions are typically cho
`sen that are not physiological, but rather optimized to provide
`long term storage stability to the protein. For example,
`Amgen’s recombinant human granulocyte colony stimulating
`factor product is formulated in HCl, pH 4.0, a solution condi
`tion that provides two year shelf life for the protein (in con
`trast, the protein forms aggregates within a week if stored in
`phosphate buffered saline at pH 7). Thus, the effects of solu
`tion conditions (especially, the role of pharmaceutical exci
`pients required for control of tonicity, antimicrobial activity,
`or protein stability in bulk solutions) on protein interactions
`with foreign microparticles have received limited attention in
`the literature.
`In previous studies of protein aggregation in solution, we
`found that the rate of aggregation can be manipulated by alter
`ing solution conditions to modify protein conformational and
`colloidal stability.14 However, for cases where heterogeneous
`nucleation is operative, it is unclear whether control of these
`factors is sufficient to prevent protein aggregation. For example,
`in our studies of protein recombinant human platelet activating
`factor acetylhydrolyase, we observed significant protein particle
`formation, even in formulations that conferred both conforma
`tional and colloidal stabilities.37 We traced the cause of particle
`formation to the presence of small numbers of glass micropar
`ticles that were present in the drug product vials, presumably
`created during commercial depyrogenation procedures.
`
`Conclusions
`
`The remarkable advances in proteomics, development of
`fully humanized monoclonal antibodies and rapid drug candi
`date screening have led to a large number of proteins that are
`under development as possible therapeutics. Development of
`stable, pharmaceutically acceptable formulations now poses a
`bottleneck that must be addressed if we are to take full
`advantage of these remarkable new drug candidates. The cou
`pling of conformational equilibria with reaction kinetics,
`under solution conditions that tax existing analytical techni
`ques will provide chemical engineers and pharmaceutical sci
`entists with challenges for some time to come.
`
`Literature Cited
`
`1. Manning, M, Patel K, Borchardt R. Stability of protein
`pharmaceuticals. Pharmaceutical Res. 1989;6:903 918.
`2. Krishnan S, Chi EY, Wood SJ, et al. Oxidative dimer
`formation is the critical rate limiting step for Parkin
`son’s disease alpha synuclein fibrillogenesis. Biochem.
`2003;42(3):829 837.
`3. DePaz RA, Barnett CC, Dale DA, Carpenter JF, Gaert
`ner AL, Randolph TW. The excluding effects of sucrose
`
`on a protein chemical degradation pathway: Methionine
`oxidation in subtilisin. Archives of Biochem and Bio
`phys. 2000;384(1):123 132.
`4. Chang BS, Hershenson S. Practical approaches to pro
`tein formulation development. In: Carpenter JF, Man
`ning MC, eds. Rational Design of Stable Protein Formu
`lations. Kluwer: New York; 2002:1 25.
`Jaenicke, R., Protein Stability and Protein Folding. Ciba
`Foundation Symposia. 1991;161:206 221.
`6. Gazit E. The ‘‘Correctly folded’’ state of proteins: Is it a
`metastable state. Angewandte Chemie Intl Ed. 2001;41
`(2):257 þ.
`
`5.
`
`7. Onuchic JN, LutheySchulten Z, Wolynes PG. Theory of
`protein folding: The energy landscape perspective.
`Annual Rev of Phys Chem. 1997;48:545 600.
`8. Carpenter JF, Pikal MJ, Chang BS, Randolph TW.
`Rational design of Stable Lyophilized Protein Formula
`tions: Some Practical Advice. Pharmaceutical Res.
`1997;14(8):969 975.
`9. Roberts CJ. Kinetics of irreversible protein aggregation:
`Analysis of extended Lumry Eyring models and impli
`cations for predicting protein shelf life. J of Phys Chem
`B. 2003;107(5):1194 1207.
`10. Remmele RL, Jr, Nightlinger NS, Srinivasan S, Gombotz
`WR. Interleukin 1 receptor (IL 1R) liquid formulation
`development using differential scanning calorimetry.
`Pharmaceutical Res. 1998;15(2):200 8.
`11. Roy S, Katayama D, Dong AC, Kerwin BA, Randolph
`TW, Carpenter JF. Temperature dependence of benzyl
`alcohol
`and 8 anilinonaphthalene 1 sulfonate induced
`aggregation of recombinant human interleukin 1 recep
`tor antagonist. Biochem. 2006;45(12):3898 3911.
`12. Wyma J, Jr, Gill SJ. Binding and Linkage. Functional
`Chemistry of Biological Macromolecules. Mill Valley,
`CA:University Science Books; 1990.
`13. Krishnan S, Chi EY, Webb JN, et al. Aggregation of
`granulocyte colony stimulating factor under physiologi
`cal conditions: Characterization and thermodynamic in
`hibition. Biochem. 2002;41(20):6422 6431.
`14. Chi EY, Krishnan S, Kendrick BS, Chang BS, Carpenter
`JF, Randolph TW. Roles of conformational stability and
`colloidal stability in the aggregation of recombinant
`human granulocyte colony stimulating factor. Protein
`Sci. 2003;12(5):903 913.
`15. Chi EY, Krishnan S, Randolph TW, Carpenter JF. Phys
`ical stability of proteins in aqueous solution: Mechanism
`and driving forces in nonnative protein aggregation.
`Pharmaceutical Res. 2003;20(9):1325 1336.
`16. Chang BS, Beauvais RM, Arakawa T, et al. Formation
`of an active dimer during storage of interleukin 1 recep
`tor antagonist in aqueous solution. Biophysical J. 1996;
`71(6):3399 3406.
`17. Seefeldt MB, Kim YS, Tolley KP, Seely J, Carpenter
`JF, Randolph TW. High pressure studies of aggregation
`of recombinant human interleukin 1 receptor antagonist:
`Thermodynamics, kinetics, and application to acceler
`ated formulation studies. Protein Sci. 2005;14(9):2258
`2266.
`18. Webb JN, Webb SD, Cleland JL, Carpenter JF, Ran
`dolph TW. Partial Molar Volume, Surface Area, and
`Hydration Changes for Equilibrium olding and Forma
`
`1906
`
`DOI 10.1002/aic
`
`Published on behalf of the AIChE
`
`August 2007 Vol. 53, No. 8
`
`AIChE Journal
`
`Bausch Health Ireland Exhibit 2011, Page 5 of 6
`Mylan v. Bausch Health Ireland - IPR2022-01104
`
`

`

`tion of Aggregation Transition State: High Pressure and
`Co Solute Studies on Recombinant Human Interferon
`gamma. Proceedings of the National Academy of Scien
`ces. 2001;98(13):7259 7264.
`19. Kendrick BS, Carpenter JF, Cleland JL, Randolph TW.
`A transient expansion of the native state precedes aggre
`gation of recombinant human interferon g. Proceedings
`of the National Academy of the Sciences USA. 1998;95:
`14142 14146.
`20. Cleland JL. Impact of Protein Folding on Biotechnol
`ogy, in Protein Folding. In Vivo and In Vitro. Cleland
`JL, ed. Washington, DC: American Chemical Society;
`1993:1 21.
`21. Carpenter JF, Kendrick BS, Chang BS, Manning MC,
`Randolph TW. Inhibition of stress ind

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket