throbber
J. F. Carpenter et al. (eds.), Rational Design of Stable Protein Formulations
`© Kluwer Academic/Plenum Publishers, New York 2002
`
`Apotex Exhibit 1027
`Page 1 of 17
`
`

`

`160
`
`Theodore W. Randolph and LaToya S. Jones
`
`The mechanisms of degradation of protein structure and activity are often
`categorized in two broad classes, chemical and physical. Chemical degradation
`refers to those modifications involving covalent bonds, such as deamidation,
`oxidation and disulfide bond shuffling. Physical degradation includes unfolding
`of the protein, undesired adsorption of the protein to surfaces, and aggregation.
`The two categories are not completely independent of one another. For example,
`protein oxidation may result in a greater proclivity to aggregate, and the rate of
`non-native disulfide bond formation may be higher in aggregated proteins.
`Surface-active agents, or surfactants, are often added to protein solutions to
`prevent physical damage during purification, filtration, transportation, freeze-
`drying, spray-drying and storage. Surfactants are amphiphilic, containing a polar
`head group and a non-polartail. This dual nature causes surfactants to adapt spe-
`cific orientations at interfaces and in aqueoussolutions. It is this characteristic
`that lies at the root of the mechanisms by which surfactants affect the physical
`stability of proteins.
`A well-known exampleis the anionic surfactant sodium dodecyl sulfate, or
`SDS. The sulfate anion is the hydrophilic head group of SDS, while the long
`aliphatic dodecyl chain formsthe tail group. Ionic surfactants such as SDS have
`been knownsincethe late 1930’s as effective protein denaturants (Anson, 1939),
`and are commonly used for this purpose, e.g., as a pre-treatment for proteins in
`polyacrylamide gel electrophoresis (SDS-PAGE). In contrast, surfactants used as
`stabilizing agents in protein formulations are typically non-ionic (Loughheed et
`al., 1983; Twardowskiet al., 1983; Chawla et al., 1985). This chapter will focus
`on non-ionic surfactants; protein interactions with ionic surfactants have been
`reviewed elsewhere (Jones, 1996). An example non-ionic surfactant is poly-
`oxyethylene sorbitan monolaurate (Tween 20°), shown in Figure 1. In this
`molecule, the hydrophilic polyoxyethylene units form the head group, while the
`hydrophobic monolaurate group is the tail. Tween 20 is often added to formula-
`tions due to its ability to protect proteins from surface-induced denaturation
`(Changetal., 1996; Jones et al., 1997; Bam et al., 1998; Kreilgaard et al., 1998;
`Maaetal., 1998).
`There are a number of mechanisms by which surfactants can prevent or
`promote damage to proteins. Some of these mechanisms are generic to all
`excipients, and can be explained in the solution thermodynamic framework ofthe
`Wyman linkage theory (Wyman and Gill, 1990) and the preferential exclusion
`mechanisms developed by Timasheff and colleagues (Arakawa and Timasheff,
`1982, 1983, 1984a,b, 1985a,b,c; Arakawaet al., 1990; Timasheff, 1998). Others
`derive from the amphiphilicity of surfactants and the resulting effect of micro-
`scopic ordering of surfactant molecules at interfaces, which in turn affects the
`kinetics and thermodynamics of protein interfaces. In this chapter, we discuss a
`numberof these mechanismsandtheir implications for proteinstability.
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 2 of 17
`Page 2 of 17
`
`

`

`Surfactant-Protein Interactions
`
`161
`
`O(CH2CH,)x-H
`
`Polyethylene glycol ether
`Triton X-100, x=9-10 (average)
`Triton X-114, x=7-8 (average)
`
`2
`
`_-— (CH2CH20),H
`oo(CH2CH20)yH
`
`H(OCH2CH2)w
`
`(CH2CH,0),H
`
`wtxty+z=20
`Polysorbate
`Tween 20, R=C,,H53CO,
`Tween 80, R=C7H33CO,
`
`Figure 1. Example non-ionic surfactants.
`
`PROTEINS AND SURFACTANTSAT SURFACES
`
`Because of their dual hydrophobic/hydrophilic nature, surfactants in solu-
`tion tend to orient themselves so that the exposure of the hydrophobic portion of
`the surfactant to the aqueous solution is minimized. Thus, in systems containing
`air/water interfaces, surfactants will
`tend to accumulate at
`these interfaces,
`forming a surface layer of surfactant oriented in such a fashion that only their
`hydrophilic ends are exposed to water. Such orientation and surface adsorption
`can also occurat solid/water interfaces such as those foundin vials, syringes, and
`other containers. Protein molecules also exhibit surface activity (for a review see
`(Magdassi, 1996), and references therein) and as such will also tend to adsorb to
`and orient at these interfaces.
`From classical thermodynamics, the excess surface internal energy dUj of
`a surface with area A at a temperature T is related to the excess surface entropy
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 3 of 17
`Page 3 of 17
`
`

`

`162
`
`Theodore W. Randolph and LaToyaS. Jones
`
`S? and the chemical potential and number of surface excess moles of each
`adsorbed species:
`
`dU? = TdS? + 0dA+ DY dny,
`i=2
`
`(1)
`
`Here the subscript 1 refers to a dividing surface chosen so that there is no excess
`adsorption of species 1, the solvent (water) (Gibbs, 1961), and o is the surface
`tension. The equilibrium criterion, 5S = 0, requires that 6 be constant across the
`surface. Thus, if a protein adsorbs to an interface, at equilibrium the surface
`tension forces must be continuousand constant across the wholeinterface, includ-
`ing acrossthe protein. Thestability criterion at equilibrium requires that:
`
`00(3). >0
`
`(2)
`
`If the surface tension of the interface is greater than the internal tension in the
`protein, then in order to meet these two conditions, the surface area ofthe protein
`must increase until the two tensions are equal, i.e., the protein must unfold. In
`some cases, nearly complete loss of native activity is lost upon adsorbing to
`the interface (Rothen, 1947; Verger et al., 1973). The Gibbs adsorption equation
`relates the surface tension to the concentration of adsorbed species at an
`interface:
`
`-do = sfdT + YTisdy,
`i=2
`
`(3)
`
`where [;, and s?dT are, respectively, the excess surface adsorption and excess
`surface entropy of component/, bothrelative to a dividing surface with no surface
`excess of solvent (1), and 1; is the chemical potential of species i. Adsorption of
`protein to the interface thus lowers the interfacial tension, making unfolding less
`likely as adsorption progresses.
`If the process of surface adsorption and unfolding of protein were to stop
`after the formation of an equilibrium monolayer, the amount of adsorbed protein
`would be so small as to be generally of no consequence. Indeed, a significant
`amountof work has been dedicated to development of methods with sensitivities
`high enoughto characterize the minute amount of protein adsorbedto the inter-
`face (Tupyet al., 1998; Vermeer and Norde, 2000). However, depending on the
`degree of surface hydrophobicity and characteristics of the protein in question,
`additional processes can occurin the adsorbedfilms, leading to behavior that is
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 4 of 17
`Page 4 of 17
`
`

`

`Surfactant-Protein Interactions
`
`163
`
`no longer described by the Gibbs’ adsorption equation. Among the possible
`processes are gas-to-liquid surface phase transitions, surface precipitation, and
`the formation of surface sublayers. In many cases, the adsorbed molecules can
`be rapidly released from the surface and surface sublayers, and may exchange
`with bulk protein molecules. Alternatively, because of slow refolding kinetics,
`proteins can become irreversibly adsorbed at
`the interface, or have a rate
`of exchange somewhere between these extremes (Dickinson, 1999; Norde and
`Giacomelli, 1999). These processes of adsorption and release of structurally-
`perturbed protein molecules into the bulk solution have been implicated as one
`of the causes of protein aggregation and denaturation.
`When discussing protein adsorption at interfaces, globular proteins are
`typically characterized as being either “hard” or “soft,” having a low or high
`degree of flexibility, respectively, and by their degrees of hydrophobicity. Soft,
`hydrophobic proteins, attaining monolayer coverage of the air/water interface in
`the matter of minutes, are generally more surface reactive at hydrophobic sur-
`faces than hard, hydrophilic proteins, attaining coverage of the same surfaces in
`a matter of hours (Tripp et al., 1995). The driving force for protein adsorption is
`the decrease in the entropy of the water molecules that are ordered around the
`hydrophobic protein domains whenthe protein is in the bulk solution. Thus, the
`role of the relative degree of hydrophobicity on protein surface adsorption is
`rather straightforward: given two proteins only differing in their hydrophobici-
`ties, the more hydrophobic protein will have a greater number of productive
`interactions with the surface and will form the monolayer more quickly.
`Middelberg et al. (2000) proposed that the difference in adsorption kinetics of
`Lac21 and Lac28 peptides is because the monomeric Lac21 has more hydropho-
`bic residues exposed than its tetratmeric counterpart, Lac28; thus, Lac21 more
`readily forms a monolayer at an octane-water interface. Protein flexibility
`is important in protein spreading that occurs at the interface. A flexible protein
`can expose additional non-polar residues, leading to an increased strength in
`binding to the surface. Finally, the protein flexibility dictates the number of
`proteins that can adsorb at the interface, and their spreading rate (Norde and
`Giacomelli, 1999).
`Protein adsorption to a hydrophobic surface does not necessarily lead to a
`complete loss of “native” structure: some proteins actually gain structure. For
`example, melittin, a honeybee venom peptide, increases its o-helical content
`slightly when adsorbed to hydrophobic quartz. In contrast, adsorption of a
`tetramer of the same peptide is thought to require a loosening of the o-helical
`content. The orientation of the adsorbed helices in the peptide is parallel to the
`quartz plane, with the hydrophobic moieties facing the plane (Smith and Clark,
`1992). Caessens et al. (1999) also report an increase in helical content on the
`adsorption of the predominately random coil B-casein and B-casein peptides at
`the teflon/waterinterface.
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 5 of 17
`Page 5 of 17
`
`

`

`164
`
`Theodore W. Randolph and LaToyaS. Jones
`
`Surfactants will also adsorb to the interface. The Gibbs adsorption isotherm
`again predicts that this interaction will lower the surface tension. Thus, if sur-
`factants co-adsorb to an interface together with proteins, there will be a smaller
`driving force for surface adsorption. Furthermore, competition between the
`protein and the surfactant for the interface may reduce equilibrium protein
`adsorption. For example, Tween 20 addition displaces beta-lactoglobulin films
`from air-water interfaces (Roth et al., 2000). This effect is not universal: the ionic
`surfactant sodium dodecyl sulfate forms a complex with high surface activity that
`exhibits enhanced surface adsorption (Green et al., 2000). Finally, because the
`surface tension is lowered after surfactant is adsorbed, less damageto the protein
`that does adsorbed may occur. For example, a loss of o-helix and an increase in
`B-turn structures occur when bovine serum albumin adsorbs to polystyrene par-
`ticles. This effect is decreased when the surface is more crowded (Norde and
`Giacomelli, 2000).
`Protein stabilization by nonionic surfactants can often be observed by for-
`mulating with micromolar concentrations of surfactant. This is due to the high
`surface-activity of this class of excipients, which renders a higher effective con-
`centration of surfactant molecules at interfaces than in the bulk solution. When
`the concentration of the surfactant is much lower than its critical micelle con-
`centration (CMC), surfactant moleculeslie flat at the air/water and hydrophobic
`solid/water interfaces (Figure 2A)
`(Porter, 1994). As the concentration is
`increased, more molecules adsorb to these interfaces, such that the surface con-
`centration remains linearly proportional
`to the bulk concentration (Tanford,
`1973). This crowding forces the surfactant molecules to order themselves such
`that the hydrophilic groups are oriented towards the bulk water and the hydro-
`carbon chains are pointed towards the air or hydrophobic solid (Porter, 1994;
`Fainerman et al., 2000) (Figure 2B). At sufficiently high surfactant concentra-
`tions (i.e., at or above the CMC), there is an oriented monolayer of surfactant
`molecules and maximum surfactant absorption, at the interface (Figures 2C &
`D). The surface saturation is responsible for the sharp slope change (to essen-
`tially zero) observed in experimental plots of surface properties (surface tension,
`osmotic pressure) versus surfactant concentration (Porter, 1994) (Figure 3). It
`should be notedthat the linear variations in the surface properties shown in Figure
`3 are indicative that the surface activity of a surfactant is due to the hydropho-
`bic effect: the variation would be cooperative if hydrocarbon self-affinity was the
`appropriate explanation (Tanford, 1973). Surfactant micelles are formed in the
`bulk phase whenthe concentration is of the surfactant is above the CMC (Figure
`2D). Thus, a range of surfactant concentrations could inhibit protein denaturation
`at an interface; however, the necessity of CMClevels of surfactant to completely
`inhibit protein damage would be a strong indication that the damage is caused by
`adsorption of protein at the interface and that inhibiting protein adsorption at the
`interface plays a role in preventing protein aggregation.
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 6 of 17
`Page 6 of 17
`
`

`

`Surfactant-Protein Interactions
`
`165
`
`A
`
`B
`
`C
`\
`YN 2Xad
`D
`— deaneeia| SSS,|HART|TY
`AION
`
`a
`
`&
`
`SP ay ah’
`
`Figure 2. Simplified models of the interfacial behavior of a nonionic surfactant at several concen-
`trations in water. Circles—polar head groups(e.g., polyoxyethelenes). Rectangles—hydrophobictails
`(e.g., hydrocarbon chains). Models abovethe dotted line is for the hydrophobicsolid (ice)/waterinter-
`face and those aboveare for the air/water interface. (A) Surfactant concentration is well below the
`CMC(B) Surfactant concentration is greater than in A, but still below the CMC.(C) Surfactant con-
`centration is at the CMC.(D) The surfactant concentration is above the CMC. (This model is adapted
`from Figures 4.4 and 4.8 of Porter (Porter, 1994)).
`
`TOnnantpahian
`S
`concentration
`oO
`
`Surface tension
`Osmotic pressure
`
`Figure 3. Representation of changes two properties in determining the CMCofa surfactant.
`
`Decreased protein adsorption at interfaces (e.g., air/liquid, ice/liquid) in the
`presence of nonionic surfactants such as Tween 20 can beattributed to the surface
`activity of nonionic surfactants (Changet al., 1996; Kreilgaard etal., 1998; Miller
`et al., 2000a,b) and, in somecases,direct interactions between the surfactant and
`protein molecules (Dickinson, 1998; Bam et al., 1998; Miller et al., 2000a,b). In
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 7 of 17
`Page 7 of 17
`
`

`

`166
`
`Theodore W. Randolph and LaToyaS. Jones
`
`mixed protein/surfactant systems in which the surfactant binds to hydrophobic
`regionsofthe protein, the protein is less surface-active than it would be in a solu-
`tion devoid of the nonionic surfactant. This explains the increase in surface
`tension relative to that of the pure protein solution that can be observed at
`extremely low surfactant concentrations. Adsorption of surfactant and protein
`molecules in the mixed system is competitive. Nonionic surfactants usually bind
`tighter than proteins or protein—-surfactant complexes at interfaces (Dickinson,
`1998). Thus, above a critical concentration of the surfactant, protein adsorption
`becomes negligible and the adsorption isotherms for mixed surfactant/protein
`systems can be roughly identical that of a pure surfactant solution, as observed
`by Miller et al. (2000a) for the HSA/C,)DMPO when the C;,DMPOconcentra-
`tion exceeds 107 mol/cm’.
`
`PROTEIN-SURFACTANT INTERACTIONS IN SOLUTION
`
`In addition to altering the interaction of proteins with surfaces, non-ionic
`surfactants can also interact directly with proteins in solution. For example,
`Tween 20 acts as a chemical chaperone,aiding in the refolding of proteins (Bam
`et al., 1998; Kreilgaard et al., 1999). In vivo, proteins fold while at average con-
`centrations of approximately 35 mg/ml (Hartl, 1996): in vitro, non-native protein
`moleculesat this concentration (e.g., due to freeze concentration) usually aggre-
`gate. Inside cells, protein folding is aided by naturally occurring molecular chap-
`erones. Unlike folding catalysts that have steric information to guide the protein
`folding, molecular chaperones act by non-covalently binding to partially-folded
`proteins, usually via hydrophobic interactions, to prevent misfolding or aggrega-
`tion while the protein is attempting to adopt its native conformation (Gatenby
`and Ellis, 1990; Hartl, 1996). Chaperone-assisted protein refolding can prevent
`the protein from falling into kinetic traps or simply allow more time for the
`protein to refold (Gatenby and Ellis, 1990). The hydrophobic effect, the driving
`force of protein folding and surface activity of Tween 20, is also implicated in
`the interactions between the exposed hydrophobic regions of the partially folded
`protein and the hydrocarbontail of the surfactant. These mixed surfactant/protein
`complexes and protein folding are dynamic processes, eventually the protein
`attains a conformation in which the hydrophobic groups are not as surface
`exposed. Unless there are hydrophobic patches in the native protein conforma-
`tion, the surfactant molecules will not necessarily bind to the protein, as in the
`case of molecular chaperones. In vitro, surfactants such as Tweens (Bam etal.,
`1996), polyethylene glycol (Cleland and Wang, 1990; Cleland et al., 1992;
`Cleland and Randolph, 1992; Cleland 1993; Cleland and Wang, 1993), Triton
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 8 of 17
`Page 8 of 17
`
`

`

`Surfactant-Protein Interactions
`
`167
`
`X-100 (Donate et al., 1998), and lubrol (Donate et al., 1998), have been impli-
`cated in aiding protein refolding by acting as chemical chaperones.
`Surfactants, such as Tween 20, can also affect the thermodynamic confor-
`mational stability of a protein. As discussed in other chapters in this book, and
`in detail in references from Timasheff and co-workers (Lee and Timasheff, 1981;
`Arakawa and Timasheff, 1982; Arakawa and Timasheff, 1985; Timasheff, 1998),
`thermodynamicstability is increased if a ligand exhibits greater binding to the
`native state of a protein than to a non-native state. However, with many excipi-
`ents (e.g., sucrose) excluded volume effects produce a non-specific negative
`binding to the native state, and a concomitantly larger negative binding to
`expanded, non-native conformations. This differential negative binding also
`results in a stabilization of the native state. At the low concentrations of surfac-
`tant (ca. 100 micromolar) typically used in formulations of therapeutic proteins,
`thermodynamic effects due to excluded volume can usually be neglected. More
`important is specific binding of surfactants to either the native or unfolded states
`of a protein. Randolph and colleagues report that some proteins and nonionic sur-
`factants, including Tween 20, form mixedprotein: detergent complexes (Bam et
`al., 1995, 1996, 1998; Jones, et al., 1999). In the presence of Tween 40 at a 4:1
`surfactant: protein molar ratio, native recombinant human growth hormoneis sig-
`nificantly stabilized: the denaturation midpoint in guanidine hydrochloride solu-
`tions increases from 4.6M guanidine hydrochloride in the absence of Tween 40
`to 5.9M in the presence of Tween 40 (Bam etal., 1996). Likewise, 4:1 Tween
`40 increases the AG of unfolding of recombinant human growth hormone by
`4.1 kcal/mol (Bam et al., 1996), and 10:1 Tween 40 increases its melting point
`slightly from 88.8 to 89.4° C (Bam etal., 1998). Stabilization of recombinant
`human growth hormoneby surfactants results in reduced aggregation in agitated
`solutions (Bam et al., 1998). With bovine serum albumin in the presence of sur-
`factant there. are decreased amounts of thermally-induced protein aggregates,
`relative to surfactant-free controls (Arakawaand Kita, 2000). Conformationalsta-
`bilization of proteins by non-ionic surfactants is not universal; stability of IgG is
`unaffected by low concentrations of Tween 20 (Vermeer and Norde, 2000), and
`recombinant human interferon-y shows lower free energies of unfolding in the
`presence of Tween 20 (Webbetal., 2000).
`
`SURFACTANT EFFECTS ON PROTEIN ASSEMBLY STATE
`
`The hydrophobic portion of non-ionic surfactants can bind to hydrophobic
`patches on proteins. This naturally causes the surfactant to order itself so that
`more hydrophilic groups are solvent exposed, resulting in a “hydrophobicity
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 9 of 17
`Page 9 of 17
`
`

`

`168
`
`Theodore W. Randolph and LaToya S. Jones
`
`the protein-surfactant
`reversal”. This “hydrophobicity reversal” means that
`complex is more hydrophilic that either the surfactant or protein alone, and effec-
`tively increases the solubility of the complex. This, in turn can reduce the pro-
`pensity of the protein to form higher-order aggregates. For example, bovine
`mitochondrial cytochrome bc1 is dimeric in solutions at low ionic strength and
`low surfactant levels. At Tween 20 concentrations above 5mg/mgprotein, a
`homogeneous, monomeric,reversible and enzymatically active protein-surfactant
`complex is formed (Musatov and Robinson, 1994). Likewise, in freeze-thaw
`studies of recombinant factor XIII, addition of Tween 20 at concentrations near
`the CMCblockedthe progression of aggregates from a relatively low molecular
`weight, soluble fraction to insoluble aggregates. Figure 4 shows levels of
`
`=nOaoOo
`
`
`
`aggregates
`
`
`nho
`
`%nativeprotein aOo
`
`
`%solubleaggregates
`%insoluble
`
`= oOo
`
`o
`
`0
`
`50
`
`100
`
`150
`
`200
`
`250
`
`Tween 20 concentration [iM]
`
`Figure 4. Recovery of native rFXIII (A) and formation of soluble (B) and insoluble aggregates (C)
`following 10 freeze-thaw cycles of 1 mg/ml (—e®—, 5 mg/ml (.---O----) and 10 mg/ml rFXIII (—¥
`as a function of Tween 20. Results are plotted as mean values +/— standard deviation for duplicate
`samples (reproduced from Kreilgaard et al., 1998).
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 10 of 17
`Page 10 of 17
`
`

`

`Surfactant-Protein Interactions
`
`169
`
`aggregation for factor XIII after freeze thaw cycling in the presence and absence
`of Tween 20. Note that Tween addition did not completely block aggregation,
`but was very effective at preventing the formation of insoluble aggregates
`(Kreilgaard et al., 1998).
`Non-ionic surfactants can also have the opposite effect on protein assembly
`state. In cases where a non-ionic surfactant destabilizes the conformation of
`
`a protein, this effect may compete against the solubilizing effect of surfactant
`binding and hydrophobicity reversal. For example, Bax, a monomeric protein that
`regulates apoptosis, readily forms dimers in the presence of Tween 20. However,
`these dimers are apparently non-native, as they do not expose the characteristic
`N-terminal Bax epitope (Hsu and Youle, 1998). In the case of the hydrophobic
`lipase from Humicola lanuginose, Tween 20 addition caused the formation of
`large, insoluble non-native aggregates (Kreilgaard et al., 1999).
`
`SURFACTANT EFFECTS ON PROTEINS DURING FREEZING,
`FREEZE-DRYING AND RECONSTITUTION
`
`The processes of freezing, drying, and reconstitution of protein solutions
`present a numberof stresses that may denature proteins. Many ofthese stresses
`are associated with surfaces: new ice-water and ice-glassy solid interfaces are
`formed during freezing, drying replaces ice-glass interfaces with air-glass inter-
`faces, and reconstitution exposes the glassy solid surfaces to aqueous solution.
`In each of these steps protein adsorption to surfaces is potentially damaging.
`The ice-water interface has been implicated as a source of damage to proteins
`(Strambini and Gabellieri, 1996), as has the solid-air interface (Hsuetal., 1995).
`Addition of nonionic surfactants can reduce this damage, presumably by com-
`peting with the protein for the ice-water interface (Chang, 1996). For example,
`addition of Tween 80 to solutions of recombinant hemoglobin reduced aggrega-
`tion seen during freeze thaw studies (Kerwin etal., 1998). Interestingly, Tween
`80 did not offer protection against methemoglobin formation or hemoglobin
`aggregation during long-term frozen storage.
`Non-ionic surfactants also have been shownto affect the recovery of native
`protein from lyophilized formulations. Sarciaux et al. (1999) showed that the
`addition of Tween 80 to the formulation solution or the reconstitution medium
`for lyophilized formulations resulted in reduced levels of aggregates. Likewise,
`Zhanget al. (1995, 1996) have demonstrated that, following long-term storage,
`surfactants in the reconstitution medium can affect protein recovery. We have
`recently shown (Webbetal., 2000) that addition of Tween 20 to the reconstitu-
`tion medium for lyophilized preparations of recombinant human interferon-y
`results in decreased levels of aggregates. The mechanism for such reduced
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 11 of 17
`Page 11 of 17
`
`

`

`170
`
`Theodore W. Randolph and LaToya S. Jones
`
`aggregation was shownto be a surfactant effect on dissolution rates. Addition of
`Tween 20 slowedthe dissolution of the lyophilized solid, allowing protein that
`partially unfolded during freeze-drying to refold before aggregating. Interest-
`ingly, Tween 20 in aqueoussolutions destabilizes recombinant human interferon-
`Y against urea-induced unfolding, impedes refolding during rapid dilution from
`urea solutions, and actually increases aggregation during agitation.
`
`ENZYMATIC DEGRADATION OF NON-IONIC SURFACTANTS
`
`Although most non-ionic surfactants are thought of as chemically inert com-
`ponents of a formulation, specific chemical interactions between proteins can
`occur. For example, some enzymes show hydrolytic activity toward Tweens.
`Smegmatocin, an esterase from Mycobacterium smegmatis, shows a broad
`thermal and pHstability in its activity against Tween 80 (Tomioka, 1983). The
`byproduct of Tween 80 hydrolysis, oleic acid, is toxic to some bacteria. Similar
`bacteriocins, which require Tween their expression, have been ascribed to other
`mycobacteria (Saito et al., 1983). It is not clear how widespread esterase activ-
`ity is against Tweens. However,it is clear that caution should be used whenfor-
`mulating proteins with esterase activities in Tween solutions.
`
`RECOMMENDATIONS FOR PROTEIN FORMULATION
`
`Clearly, it is desirable to minimize the addition of any excipient to a for-
`mulation. This rule of thumb is even more pertinent for surfactants, because there
`is ample evidence that high concentrations of surfactants can be destabilizing to
`protein structure. On the other hand, small amounts of surfactant often provide
`benefits in preventing aggregation that greatly outweigh any conformationally
`destabilizing effect. How then should surfactant levels be chosen for optimal for-
`mulation? The answer appears to depend on the mechanism(s) by which a par-
`ticular protein is protected from damage by surfactant addition. In cases where
`surfactants act to stabilize the native state of a protein by bindingto the protein,
`a specific surfactant: protein stoichiometry may need to be maintained in order
`to provide optimal protection. In these cases, changes in the protein concentra-
`tion within a formulation will dictate proportional changes in surfactant concen-
`tration to maintain a fixed molar ratio. This appears to be the case for recombinant
`human growth hormone, where protection against agitation-induced damagecor-
`related with the molar ratio of surfactant to protein rather than to the surfactant’s
`CMC(Bam etal., 1998). In the case of specific binding, the choice of nonionic
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 12 of 17
`Page 12 of 17
`
`

`

`Surfactant-Protein Interactions
`
`171
`
`surfactant may be important. In the case of recombinant human growth hormone,
`for example, different binding stoichiometries and degrees of protein stabiliza-
`tion were seen for a variety of common surfactants (Bametal., 1995). A general
`recommendation for proteins that show specific binding to the native state of the
`protein is to formulate so that the ratio of surfactant to protein is slightly above
`the binding stoichiometry for a particular surfactant. The choice of surfactant may
`be dictated by the degree of stabilization (which should correlate with the degree
`of binding) provided to a protein by a particular surfactant.
`In contrast, if no specific binding is seen, then maximum levels of protec-
`tion generally correlate with the CMC of the surfactant. In this case, surfactant
`should be addedat levels slightly above the CMC. The choice of surfactant is
`often dictated by a trade-off: surfactants with lower CMC’s will require less sur-
`factantin solution to saturate surfaces and reduce surface-induced damageto pro-
`teins. However, surfactants with low CMC’s are much moredifficult to remove
`from solution (e.g., by dialysis) if necessary, and also tend to be less soluble than
`surfactants with higher CMC’s, raising the possibility of undesirable phase
`separation during processes such as freezing or lyophilization.
`
`_
`
`REFERENCES
`
`Adler, M., and Lee, G., 1999. Stability and surface activity of lactate dehydrogenase in
`spray-dried trehalose. Journal of Pharmaceutical Sciences 88:199.
`Anson,M., 1939. The denaturation ofproteins by detergents and bile salts. Science 90:256.
`Arakawa,T., Bhat, R. et al., 1990. Preferential interactions determine protein solubility in
`three-componentsolutions: the MgCl, system. Biochemistry 29:1914.
`Arakawa, T., Bhat, R. et al., 1990. Why preferential hydration does not always stabilize
`the native structure of globular proteins. Biochemistry 29:1924.
`Arakawa,T., and Kita, Y., 2000. Protection of bovine serum albumin from aggregation by
`Tween 80. J. Pharm. Sci. 89:646.
`
`Arakawa,T., and Timasheff. S.N., 1982. Preferential interactions of proteins with salts in
`concentrated solutions. Biochemistry 21:6545.
`Arakawa, T., and Timasheff, S.N., 1982. Stabilization of protein structure by sugars.
`Biochemistry 21:6536.
`Arakawa,T., and Timasheff, S.N., 1983. Preferential interactions of proteins with solvent
`components in aqueous amino acid solutions. Arch. Biochem. Biophys. 224:169.
`Arakawa,T., and Timasheff, S.N., 1984a. Mechanism ofprotein salting in and salting out
`by divalent cation salts: balance between hydration and salt binding. Biochemistry
`23:5912.
`
`Arakawa, T., and Timasheff, S.N., 1984b. Protein stabilization and destabilization by
`guanidinium salts. Biochemistry 23:5924.
`Arakawa,T., and Timasheff, S.N., 1985a. Mechanism of poly(ethylene glycol) interaction
`with proteins. Biochemistry 24:6756.
`
`Apotex Exhibit 1027
`Apotex Exhibit 1027
`Page 13 of 17
`Page 13 of 17
`
`

`

`172
`
`Theodore W. Randolph and LaToyaS. Jones
`
`Arakawa, T., and Timasheff, S.N., 1985b. The stabilization of proteins by osmolytes.
`Biophys. J. 47:A11.
`Arakawa, T., and Timasheff, S.N., 1985c. Theory of protein solubility. Meth. Enzymol.
`114:49.
`
`Bam, N.B., Randolph, T.W., and Cleland, J.L., 1995. Stability of protein formulations:
`Investigation of surfactant effects by a novel EPR spectroscopic technique. Pharm.
`Res, 12:2.
`
`Bam, N.B., Cleland, J.L. et al., 1996. Molten globule intermediate of recombinant human
`growth hormone:stabilization with surfactants. Biotech. Prog. 12:801.
`Bam, N.B., Cleland, J.L.et al., 1998. Tween protects recombinant human growth hormone
`against agitation-induced damage via hydrophobic interactions. J. Pharm. Sci.
`87:1554.
`
`Bam,N.B., Randolph, T.W.et al., 1995. Stability of protein formulations: investigation of
`surfactant effects by a novel EPR spectroscopic technique. Pharm. Res. 12:2.
`Broadhead, J., Rouan, S.K. et al., 1994. The effect of process and formulation variables
`on the properties of spray-dried beta-galactosidase. J. Pharmaceut. Pharmacol.
`46:458.
`
`Caessens, P.W., De Jongh, H.et al., 1999. The adsorption-induced secondary structure
`of B-casein and of distinct parts of its sequence in relation to foam and emulsion
`properties. Biochim. Biophys. Acta 1430:73.
`in:
`Carpenter, J.F., and Chang, B.S., 1996. Lyophilization of protein pharmaceuticals,
`Biotechnology and Biopharmaceutical Manufacturing, Processing, and Preservation,
`K_E. Avis and V.L. Wu, eds., Interpharm Press, Buffalo Grove, IL.
`Carpenter, J.F., Pikal, M.J., Chang, B.S., and Randolph, T.W., 1997. Rational design of
`stable lyophilized protein formulations: some practical

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