throbber
International Journal of Pharmaceutics 203 (2000) 1–60
`
`www.elsevier.com:locate:ijpharm
`
`Review
`Lyophilization and development of solid protein
`pharmaceuticals
`
`Wei Wang*
`
`Biotechnology, Bayer Corporation, 800 Dwight Way, Berkeley, CA 94701, USA
`
`Received 9 September 1999; received in revised form 21 December 1999; accepted 4 April 2000
`
`Abstract
`
`Developing recombinant protein pharmaceuticals has proved to be very challenging because of both the complexity
`of protein production and purification, and the limited physical and chemical stability of proteins. To overcome the
`instability barrier, proteins often have to be made into solid forms to achieve an acceptable shelf life as pharmaceu-
`tical products. The most commonly used method for preparing solid protein pharmaceuticals is lyophilization
`(freeze-drying). Unfortunately, the lyophilization process generates both freezing and drying stresses, which can
`denature proteins to various degrees. Even after successful lyophilization with a protein stabilizer(s), proteins in solid
`state may still have limited long-term storage stability. In the past two decades, numerous studies have been
`conducted in the area of protein lyophilization technology, and instability:stabilization during lyophilization and
`long-term storage. Many critical issues have been identified. To have an up-to-date perspective of the lyophilization
`process and more importantly, its application in formulating solid protein pharmaceuticals, this article reviews the
`recent investigations and achievements in these exciting areas, especially in the past 10 years. Four interrelated topics
`are discussed: lyophilization and its denaturation stresses, cryo- and lyo-protection of proteins by excipients, design
`of a robust lyophilization cycle, and with emphasis,
`instability, stabilization, and formulation of solid protein
`pharmaceuticals. © 2000 Elsevier Science B.V. All rights reserved.
`
`Keywords: Aggregation; Cryoprotection; Denaturation; Excipient; Formulation; Freeze-drying; Glass transition; Stability; Lyopro-
`tection; Residual moisture
`
`1. Introduction
`
`Developing recombinant protein pharmaceuti-
`cals has proved to be very challenging because of
`
`* Tel.: (cid:27)1-510-7054755; fax: (cid:27)1-510-7055629.
`E-mail address: wei.wang.b@bayer.com (W. Wang).
`
`both the complexity of protein production and
`purification, and the limited physical and chemi-
`cal stability of proteins. In fact, protein instability
`is one of the two major reasons why protein
`pharmaceuticals are administered traditionally
`through injection rather than taken orally like
`most small chemical drugs (Wang, 1996). To over-
`
`0378-5173:00:$ - see front matter © 2000 Elsevier Science B.V. All rights reserved.
`PII: S 0 3 7 8 - 5 1 7 3 ( 0 0 ) 0 0 4 2 3 - 3
`
`Amneal v. Cubist
`IPR2020-00193
`Cubist Ex. 2004
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`cal and chemical instabilities and stabilization of
`proteins in aqueous and solid states (Manning et
`al., 1989; Cleland et al., 1993); chemical instability
`mechanisms of proteins in solid state (Lai and
`Topp, 1999); various factors affecting protein sta-
`bility during freeze-thawing, freeze-drying, and
`storage
`of
`solid
`protein
`pharmaceuticals
`(Arakawa et al., 1993); and application of
`lyophilization in protein drug development (Pikal,
`1990a,b; Skrabanja et al., 1994; Carpenter et al.,
`1997; Jennings, 1999). Nevertheless,
`it appears
`that several critical issues in the development of
`solid protein pharmaceuticals have not been fully
`examined,
`including various instability factors,
`stabilization, and formulation of solid protein
`pharmaceuticals.
`the
`To have an up-to-date perspective of
`lyophilization process and more importantly, its
`application in formulating solid protein pharma-
`ceuticals, this article reviews the recent investiga-
`tions and achievements in these exciting areas,
`especially in the past 10 years. Four interrelated
`topics are discussed sequentially,
`lyophilization
`and its denaturation stresses; cryo- and lyo-pro-
`tection of proteins by excipients; design of a ro-
`bust
`lyophilization cycle; and with emphasis,
`instability, stabilization, and formulation of solid
`protein pharmaceuticals.
`
`2. Lyophilization and its denaturation stresses
`
`2.1. Lyophilization process
`
`Lyophilization (freeze-drying) is the most com-
`mon process for making solid protein pharmaceu-
`ticals (Cleland et al., 1993; Fox, 1995). This
`process consists of two major steps: freezing of a
`protein solution, and drying of the frozen solid
`under vacuum. The drying step is further divided
`into two phases: primary and secondary drying.
`The primary drying removes the frozen water and
`the secondary drying removes the non-frozen
`‘bound’ water (Arakawa et al., 1993). The amount
`of non-frozen water for globular proteins is about
`0.3–0.35 g g(cid:28) 1 protein, slightly less than the
`proteins’ hydration shell
`(Rupley and Careri,
`1991; Kuhlman et al., 1997). More detailed analy-
`
`2 c
`
`ome the instability barrier, proteins often have to
`be made into solid forms to achieve an acceptable
`shelf life.
`The most commonly used method for preparing
`solid protein pharmaceuticals is lyophilization
`(freeze-drying). However, this process generates a
`variety of freezing and drying stresses, such as
`solute concentration, formation of ice crystals, pH
`changes, etc. All of these stresses can denature
`proteins to various degrees. Thus, stabilizers are
`often required in a protein formulation to protect
`protein stability both during freezing and drying
`processes.
`Even after successful lyophilization, the long-
`term storage stability of proteins may still be very
`limited, especially at high storage temperatures. In
`several cases, protein stability in solid state has
`been shown to be equal to, or even worse than,
`that in liquid state, depending on the storage
`temperature and formulation composition. For
`example, a major degradation pathway of human
`insulin-like growth factor I (hIGF-I) is oxidation
`of Met59 and the oxidation rate in a freeze-dried
`formulation in air-filled vials is roughly the same
`as that
`in a solution at either 25 or 30°C
`(Fransson et al., 1996). Similarly, the oxidation
`rate of lyophilized interleukin 2 (IL-2) is the same
`as that in a liquid formulation containing 1 mg
`ml(cid:28) 1 IL-2, 0.5% hydroxypropyl-b-cyclodextrin
`(HP-b-CD), and 2% sucrose during storage at 4°C
`(Hora et al., 1992b). At a high water content
`(\50%), the degradation rate of insulin is higher
`in a lyophilized formulation than in a solution
`with similar pH-rate profiles
`in both states
`(Strickley and Anderson, 1996). The glucose-in-
`duced formation of des-Ser
`relaxin in a
`lyophilized formulation is faster than in a solution
`during storage at 40°C (Li et al., 1996). These
`examples indicate that stabilizers are still required
`in lyophilized formulations to increase long-term
`storage stability.
`In the past two decades, numerous studies have
`been conducted in the areas of protein freezing
`and drying, and instability and stabilization of
`proteins during lyophilization and long-term stor-
`age. Many critical issues have been identified in
`this period. These studies and achievements have
`been reviewed elsewhere with emphasis on physi-
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`3
`
`sis of each lyophilization step is provided in Sec-
`tion 4.
`Lyophilization generates a variety of stresses,
`which tend to destabilize or unfold:denature an
`unprotected protein. Different proteins tolerate
`freezing and:or drying stresses to various degrees.
`Freeze-thawing of ovalbumin at neutral pH did
`not cause denaturation (Koseki et al., 1990). Re-
`peated (three times) freeze-thawing of tissue-type
`plasminogen activator (tPA) did not cause any
`decrease in protein activity (Hsu et al., 1995).
`Some proteins can keep their activity both during
`freezing and drying processes, such as a1-an-
`titrypsin in phosphate–citrate buffer (Vemuri et
`al., 1994), porcine pancreatic elastase without ex-
`cipients (Chang et al., 1993), and bovine pancre-
`atic ribonuclease A (RNase A, 13.7 kD) in the
`presence or absence of phosphate (Townsend and
`DeLuca, 1990).
`However, many proteins cannot stand freezing
`and:or drying stresses. Freeze-thawing caused loss
`of activity of
`lactate dehydrogenase
`(LDH)
`(Nema and Avis, 1992; Izutsu et al., 1994b; An-
`dersson and Hatti-Kaul, 1999), 60% loss of L-as-
`paraginase (10 mg ml(cid:28) 1) activity in 50 mM
`sodium phosphate buffer (pH 7.4) (Izutsu et al.,
`1994a),
`and
`aggregation
`of
`recombinant
`hemoglobin (Kerwin et al., 1998). Freeze-drying
`caused 10% loss of the antigen-binding capacity
`of a mouse monoclonal antibody (MN12) (Ress-
`ing et al., 1992), more than 40% loss of bilirubin
`oxidase (BO) activity in the presence of dextran or
`polyvinylalcohol (PVA) (Nakai et al., 1998), loss
`of most b-galactosidase activity at 2 or 20 mg
`ml(cid:28) 1 (Izutsu et al., 1993, 1994a), complete loss of
`phosphofructokinase (PFK) and LDH activity in
`the absence of stabilizers (Carpenter et al., 1986,
`1990; Prestrelski et al., 1993a; Anchordoquy and
`Carpenter, 1996), and dissociation of Erwinia L-
`asparaginase tetramer (135 kD) into four inactive
`subunits (34 kD each) in the absence of any
`protectants (Adams and Ramsay, 1996).
`
`2.2. Denaturation stresses during lyophilization
`
`The lyophilization process generates a variety
`of stresses to denature proteins. These include (1)
`low temperature stress; (2) freezing stresses, in-
`
`in-
`cluding formation of dendritic ice crystals,
`creased ionic strength, changed pH, and phase
`separation; and (3) drying stress (removing of the
`protein hydration shell).
`
`2.2.1. Low temperature stress
`The first quantitative study on low-temperature
`denaturation of a model protein was conducted
`presumably by Shikama and Yamazaki (1961).
`They demonstrated a specific temperature range
`in which ox liver catalase was denatured during
`freeze-thawing. Cold denaturation of catalase at
`8.4 mg ml (cid:28) 1 in 10 mM phosphate buffer (pH 7.0)
`started at (cid:28)6°C. Loss of catalase activity reached
`20% at (cid:28)12°C, remained at this level between
`(cid:28)12°C and near (cid:28)75°C, then decreased gradu-
`ally from (cid:28)75 to (cid:28)120°C. There was almost no
`activity loss between (cid:28)129 and (cid:28)192°C. Similar
`results were also obtained for ovalbumin by
`Koseki et al. (1990). Incubation of frozen ovalbu-
`min solution caused structural change of ovalbu-
`min, as monitored by UV difference spectra,
`which increased with decreasing temperature be-
`tween (cid:28)10 and (cid:28)40°C. Further decrease in
`incubation temperature to (cid:28)80°C caused less
`structural change, and no change at (cid:28)192°C.
`Perlman and Nguyen (1992) reported that inter-
`feron-g(IFN-g) aggregation in a liquid mannitol
`formulation was more severe at (cid:28)20°C than at
`(cid:28)70, 5 and 15°C during storage. To prevent
`freezing-induced complication in studying cold
`protein denaturation, cold and heat denaturation
`of RNase A has been conducted under high pres-
`sure (3 kbar). Under this condition, RNase A
`denatured below (cid:28)22°C and above 40°C (Zhang
`et al., 1995). All these examples are clear indica-
`tion of low temperature denaturation rather than
`a freezing or thawing effect.
`The nature of cold denaturation has not been
`satisfactorily delineated. Since solubility of non-
`polar groups in water increases with decreasing
`temperature due to increased hydration of the
`non-polar groups,
`solvophobic
`interaction in
`proteins weakens with decreasing temperature
`(Dill et al., 1989; Graziano et al., 1997). The
`decreasing solvophobic interaction in proteins can
`reach a point where protein stability reaches zero,
`causing cold denaturation (Jaenicke, 1990). While
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`4 n
`
`ormal or thermal denaturation is entropy-driven,
`cold denaturation is enthalpy-driven (Dill et al.,
`1989; Shortle, 1996). Oligomeric proteins typically
`show cold denaturation, i.e. dissociation of sub-
`unit oligomers, since association is considered to
`be a consequence of hydrophobic interaction
`(Jaenicke, 1990; Wisniewski, 1998). Theoretically,
`the calculated free energy of unfolding (DGunf) for
`proteins has a parabolic relationship with temper-
`ature. This means that a temperature of maximum
`stability exists, and both high and low tempera-
`ture can destabilize a protein (Jaenicke, 1990;
`Kristja´nsson and Kinsella, 1991).
`
`2.2.2. Concentration effect
`Freezing a protein solution rapidly increases the
`concentration of all solutes due to ice formation.
`For example, freezing a 0.9% NaCl solution to its
`eutectic temperature of (cid:28)21°C can cause a 24-
`fold increase in its concentration (Franks, 1990).
`The calculated concentration of small carbohy-
`drates in the maximally freeze-concentrated ma-
`trices (MFCS) is as high as 80% (Roos, 1993).
`Thus, all physical properties related to concentra-
`tion may change, such as ionic strength and rela-
`tive composition of
`solutes due to selective
`crystallization. These changes may potentially
`destabilize a protein.
`Generally, lowering the temperature reduces the
`rate of chemical reactions. However, chemical
`reactions may actually accelerate in a partially
`frozen aqueous solution due to increased solute
`concentration (Pikal, 1999). Due to solute concen-
`tration, the rate of oligomerization of b-glutamic
`acid at (cid:28)20°C was much faster than at 0 or 25°C
`in the presence of a water-soluble carbodiimide,
`1-ethyl-3-(3-dimethylaminopropyl)
`carbodiimide
`(EDAC) (Liu and Orgel, 1997).
`The increase in the rate of a chemical reaction
`in a partially frozen state could reach several
`orders of magnitude relative to that in solution
`(Franks, 1990, 1994).
`The reported oxygen concentration in a par-
`tially frozen solution at (cid:28)3°C is as high as 1150
`times that in solution at 0°C (Wisniewski, 1998).
`The increased oxygen concentration can readily
`oxidize sulphydryl groups in proteins. If a protein
`solution contains any contaminant proteases, con-
`
`centration upon freezing may drastically acceler-
`ate protease-catalyzed protein degradation.
`
`2.2.3. Formation of ice-water interface
`Freezing a protein solution generates an ice-wa-
`ter interface. Proteins can be adsorbed to the
`interface, loosening the native fold of proteins and
`resulting
`in
`surface-induced
`denaturation
`(Strambini and Gabellieri, 1996). Rapid (quench)
`cooling generates a large ice-water interface while
`a smaller interface is induced by slow cooling
`(also see Section 4.2). Chang et al.
`(1996b)
`demonstrated that a single freeze–thaw cycle with
`quench cooling denatured six model proteins, in-
`cluding ciliary neurotrophic factor (CNTF), gluta-
`mate
`dehydrogenase
`(GDH),
`interleukin-1
`receptor antagonist
`(IL-1ra), LDH, PFK, and
`tumor necrosis factor binding protein (TNFbp).
`The denaturation effect of quench cooling was
`greater or equivalent to that after 11 cycles of
`slow cooling, suggesting surface-induced denatu-
`ration. This denaturation mechanism was sup-
`ported by a good correlation (r(cid:30)0.99) found
`between the degree of freeze-induced denaturation
`and that of artificially surface-induced denatura-
`tion. The surface was introduced by shaking the
`protein solution containing hydrophobic Teflon
`beads. In a similar study, a correlation coefficient
`of 0.93 was found between the tendency of freeze
`denaturation and surface-induced denaturation
`for eight model proteins, including aldolase, basic
`fibroblast growth factor (bFGF), GDH, IL-1ra,
`LDH, maleate dehydrogenase (MDH), PFK, and
`TNFbp (Kendrick et al., 1995b). However, there
`was no significant correlation (r(cid:30)0.78) between
`freeze denaturation and thermal denaturation
`temperature (Chang et al., 1996b).
`
`2.2.4. pH changes during freezing
`Many proteins are stable only in a narrow pH
`range, such as low molecular weight urokinase
`(LMW-UK) at pH 6–7 (Vrkljan et al., 1994). At
`extreme pHs, increased electrostatic repulsion be-
`tween like charges in proteins tends to cause
`protein unfolding or denaturation (Goto and
`Fink, 1989; Volkin and Klibanov, 1989; Dill,
`1990). Thus, the rate of protein aggregation is
`strongly affected by pH, such as aggregation of
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`5
`
`interleukin 1b (IL-1b) (Gu et al., 1991), human
`relaxin (Li et al., 1995a), and bovine pancreatic
`RNase A (Townsend and DeLuca, 1990; Tsai et
`al., 1998). Moreover, the solution pH can signifi-
`cantly affect the rate of many chemical degrada-
`tions in proteins (Wang, 1999).
`Freezing a buffered protein solution may selec-
`tively crystallize one buffering species, causing pH
`changes. Na2HPO4 crystallizes more readily than
`NaH2PO4 because the solubility of the disodium
`form is considerably lower than that of
`the
`monosodium form. Because of this, a sodium
`phosphate buffer at pH 7 has a molar [NaH2PO4]:
`[Na2HPO4] ratio of 0.72, but this ratio increases
`to 57 at the ternary eutectic temperature during
`freezing (Franks, 1990, 1993). This can lead to a
`significant pH drop during freezing, which then
`denatures pH-sensitive proteins. For example,
`freezing of a LDH solution caused protein denat-
`uration due to a pH drop from 7.5 to 4.5 upon
`selective crystallization of Na2HPO4 (Anchordo-
`quy and Carpenter, 1996). LDH is a pH-sensitive
`protein and a small drop in pH during freezing
`can partially denature the protein even in the
`presence of stabilizers such as sucrose and tre-
`halose (Nema and Avis, 1992). The pH drop
`during freezing may also explain why freezing
`bovine and human IgG species in a sodium phos-
`phate buffer caused formation of more aggregates
`than in potassium phosphate buffer, because
`potassium phosphate buffer does not show signifi-
`cant pH changes during freezing (Sarciaux et al.,
`1998).
`The pH drop during freezing can potentially
`affect storage stability of
`lyophilized proteins.
`Lyophilized IL-1ra in a formulation containing
`phosphate buffer at pH 6.5 aggregated more
`rapidly than that containing citrate buffer at the
`same pH during storage at 8, 30 and 50°C (Chang
`et al., 1996c). Similarly, the pH drop of a succi-
`nate-containing formulation from 5 to 3–4 during
`freezing appeared to be the cause of less storage
`stability for lyophilized IFN-g than that contain-
`ing glycocholate buffer at the same pH (Lam et
`al., 1996).
`
`2.2.5. Phase separation during freezing
`Freezing polymer solutions may cause phase
`
`separation due to polymers’ altered solubilities at
`low temperatures. Freezing-induced phase separa-
`tion can easily occur in a solution containing two
`incompatible polymers such as dextran and Ficoll
`(Izutsu et al., 1996). During freezing of recombi-
`nant hemoglobin in a phosphate buffer containing
`4% (w:w) PEG 3350, 4% (w:w) dextran T500, and
`150 mM NaCl,
`liquid–liquid phase separation
`occurred and created a large excess of interface,
`denaturing the protein (Heller et al., 1997). Addi-
`tion of 5% sucrose or trehalose could not reverse
`the denaturation effect in the system (Heller et al.,
`1999a).
`Several strategies have been proposed to miti-
`gate or prevent phase separation-induced protein
`denaturation during freezing. These include use of
`alternative salts (Heller et al., 1999a), adjustment
`of the relative composition of polymers to avoid
`or to rapidly pass over a temperature region
`where the system may result
`in liquid–liquid
`phase separation (Heller et al., 1999c), and chemi-
`cal modification of the protein such as pegylation
`(Heller et al., 1999b).
`
`2.2.6. Dehydration stresses
`Proteins in an aqueous solution are fully hy-
`drated. A fully hydrated protein has a monolayer
`of water covering the protein surface, which is
`termed the hydration shell (Rupley and Careri,
`1991). The amount of water in full hydration is
`0.3–0.35 g g (cid:28) 1 protein (Rupley and Careri, 1991;
`Kuhlman et al., 1997). Generally, the water con-
`tent of a lyophilized protein product is less than
`10%. Therefore,
`lyophilization removes part of
`the hydration shell. Removal of the hydration
`shell may disrupt the native state of a protein and
`cause denaturation. A hydrated protein, when
`exposed to a water-poor environment during de-
`hydration, tends to transfer protons to ionized
`carboxyl groups and thus abolishes as many
`charges as possible in the protein (Rupley and
`Careri, 1991). The decreased charge density may
`facilitate protein–protein hydrophobic interac-
`tion, causing protein aggregation.
`Water molecules can also be an integral part of
`an active site(s) in proteins. Removal of these
`functional water molecules during dehydration
`easily inactivates proteins. For example, dehydra-
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`6 t
`
`ion of lysozyme caused loss of activity apparently
`due to removal of those water molecules residing
`functionally in the active site (Nagendra et al.,
`1998).
`Lastly, dehydration during lyophilization may
`cause significant difference in moisture distribu-
`tion in different locations of a product cake. The
`uneven moisture distribution may lead to possible
`localized overdrying, which may exacerbate dehy-
`dration-induced protein denaturation (Pikal and
`Shah, 1997).
`
`2.3. Monitoring protein denaturation upon
`lyophilization
`
`The most common method for monitoring
`protein denaturation upon lyophilization appears
`to be infrared (IR) spectroscopy, although other
`methods have been used such as mass spec-
`troscopy (Bunk, 1997), and Raman spectroscopy
`(Belton and Gil, 1994). In the following section,
`IR methodology is discussed in monitoring
`protein denaturation upon lyophilization followed
`by a discussion on reversibility of protein
`denaturation.
`
`2.3.1. Infrared (IR) spectroscopy
`IR (or FTIR) is probably the most extensively
`used technique today for
`studying structural
`changes in proteins upon lyophilization (Susi and
`Byler, 1986; Dong et al., 1995; Carpenter et al.,
`1998, 1999). The lyophilization-induced structural
`changes can be monitored conveniently in the
`amide I, II, or III region. For lyophilized protein
`samples, residual water up to 10% (w:w) does not
`interfere significantly in the amide I region, a
`frequently used sensitive region for determination
`of secondary structures (Dong et al., 1995). How-
`ever, IR studies on proteins in an aqueous solu-
`tion need either subtraction of water absorption
`or solvent replacement with D2O (Goormaghtigh
`et al., 1994). To make reliable subtraction, high
`protein concentrations (\10 mg ml (cid:28) 1) are rec-
`ommended to increase protein absorption signal,
`and a CaF2 (or BaF2) cell with a path length of 10
`mm or less should be used to control the total
`sample absorbance within 1 (Cooper and Knut-
`son, 1995).
`
`Lyophilization may induce several potential
`changes in the IR spectra of proteins. Disruption
`of hydrogen bonds in proteins during lyophiliza-
`tion generally leads to an increase in frequency
`and a decrease in intensity of hydroxyl stretching
`bands (Carpenter and Crowe, 1989). Unfolding of
`proteins during lyophilization broadens and shifts
`(to higher wave numbers) amide I component
`peaks (Prestrelski et al., 1993b; Allison et al.,
`1996). Lyophilization often leads to an increase in
`b-sheet content with a concomitant decrease in
`a-helix content. Conversion of a-helix to b-sheet
`during lyophilization has been observed in many
`proteins such as tetanus toxoid (TT) in 10 mM
`sodium phosphate buffer (pH 7.3) (Costantino et
`al., 1996), recombinant human albumin (rHA) in
`different buffer
`solutions
`at different pHs
`(Costantino et al., 1995a), hGH at pH 7.8, and
`seven model proteins in water, including bovine
`pancreatic trypsin inhibitor (BPTI), chymotrypsi-
`nogen, horse myoglobin (Mb), horse heart cy-
`tochrome c (Cyt c), rHA, porcine insulin, and
`RNase A (Griebenow and Klibanov, 1995).
`b-sheet
`An
`increase
`in
`content
`during
`lyophilization is often an indication of protein
`aggregation and:or increased intermolecular inter-
`action (Yeo et
`al.,
`1994; Griebenow and
`Klibanov,
`1995; Overcashier
`et
`al.,
`1997).
`Lyophilization-induced increase in b-sheet content
`seems to be a rather general phenomenon as
`lyophilization or air-drying of unordered poly-L-
`lysine induced structural transition to a highly
`ordered b-sheet (Prestrelski et al., 1993b; Wolkers
`et al., 1998b). Such transition has also been ob-
`served in proteins during lyophilization such as
`human insulin in water (pH 7.1) (Pikal and Rigs-
`bee, 1997). The b-sheet structure after lyophiliza-
`tion shows a higher degree of
`intermolecular
`hydrogen bonding because polar groups must sat-
`isfy their H-bonding requirement by intra- or
`intermolecular interaction upon removal of water.
`The intermolecular b-sheet is characterized by two
`major IR bands at about 1617 and 1697 cm(cid:28) 1 in
`solid state, which can be used to monitor protein
`denaturation (Allison et al., 1996). Similarly, the
`relative intensity of a-helix band also can be used
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`7
`
`in this regard (Yang et al., 1999; Heller et al.,
`1999b).
`The extent of changes in overall IR spectrum of
`a protein upon lyophilization reflects the degree of
`protein denaturation. The changes relative to a
`reference spectrum can be measured using a corre-
`lation coefficient (r) as defined by Prestrelski et al.
`(1993a), or the extent of spectral area overlap
`(Heimburg and Marsh, 1993; Allison et al., 1996;
`Kendrick et al., 1996). Using the correlation co-
`efficient, Prestrelski et al. (1993b) were able to
`measure the relative freeze-drying stability of sev-
`eral model proteins, including bFGF, bovine a-
`a-casein,
`IFN-g,
`lactalbumin,
`bovine
`and
`recombinant granulocyte colony-stimulating fac-
`tor (rG-CSF) in the presence of different sugars.
`Nevertheless, Griebenow and Klibanov (1995),
`after analyzing secondary structures of seven
`model proteins upon lyophilization, concluded
`that the correlation coefficient was not highly
`sensitive to structural alterations in proteins. In-
`stead, comparison of overlapping area-normalized
`second-derivative or deconvoluted spectra seemed
`more reliable and objective.
`Recently, IR has been used in real-time moni-
`toring of freezing and dehydration stresses on
`proteins during lyophilization. By this method,
`glucose at 10% was shown to protect lysozyme
`both during the freezing and drying processes
`(Remmele et al., 1997).
`
`2.3.2. Re6ersibility of freezing- or
`lyophilization-induced protein denaturation
`Many proteins denature to various extents
`upon freezing, especially at low concentrations
`(B0.1 mg ml(cid:28) 1). Freezing-induced denaturation
`may or may not be reversible. Freezing lysozyme
`or
`IL-1ra
`caused
`reversible
`denaturation
`(Kendrick et al., 1995a). In contrast, recombinant
`factor XIII (rFXIII, 166 kD) was irreversibly
`denatured upon freezing, and loss of native
`rFXIII at 1 mg ml (cid:28) 1 increased linearly with the
`number of freeze–thaw cycles (Kreilgaard et al.,
`1998b).
`lyophilization-induced denaturation
`Similarly,
`can be either reversible or irreversible. In the
`absence of stabilizers, PFK at 25 mg ml (cid:28) 1 at pH
`7.5 and 8.0 was fully and irreversibly inactivated
`
`upon lyophilization (Carpenter et al., 1993), while
`loss of BO activity in a PVA-containing formula-
`tion was at least partially reversible (Nakai et al.,
`1998). Using IR spectroscopy, Prestrelski et al.
`(1993b) demonstrated that lyophilization-induced
`structural changes were irreversible for bFGF,
`IFN-g, and bovine a-casein, but essentially re-
`versible for G-CSF and bovine a-lactalbumin.
`The extensive aggregation and precipitation of
`IFN-g and casein upon rehydration confirmed the
`irreversibility in structural changes. Therefore,
`lyophilization of proteins may lead to three types
`of behavior, (1) no change in protein conforma-
`tion; (2) reversible denaturation; or (3) irreversible
`denaturation.
`In many cases, IR-monitored structural changes
`during lyophilization seem to be
`reversible.
`Griebenow and Klibanov (1995) demonstrated
`that lyophilization (dehydration) caused signifi-
`cant changes in the secondary structures of seven
`model proteins in the amide III region (1220–
`1330 cm(cid:28) 1), including BPTI, chymotrypsinogen,
`Mb, Cyt c, rHA,
`insulin, and RNase A. The
`structure of almost all proteins became more or-
`dered upon lyophilization with a decrease in the
`unordered structures. Nevertheless, all these struc-
`tural changes were reversible upon reconstitution.
`Other examples of reversible changes in the sec-
`ondary structures of proteins upon lyophilization
`include rHA (Costantino et al., 1995a), Humicola
`lanuginosa lipase (Kreilgaard et al., 1999), IL-2
`(Prestrelski et al., 1995), and lysozyme (Allison et
`al., 1999).
`
`3. Cryo- and lyo-protection of proteins by
`stabilizers
`
`As discussed before, both freezing and dehydra-
`tion can induce protein denaturation. To protect a
`protein from freezing (cryoprotection) and:or de-
`hydration (lyoprotection) denaturation, a protein
`stabilizer(s) may be used. These stabilizers are
`also referenced as chemical additives (Li et al.,
`1995b), co-solutes (Arakawa et al. 1993), co-sol-
`vents
`(Timasheff, 1993, 1998), or
`excipients
`(Wong and Parascrampuria, 1997; Wang, 1999).
`In the following section, a variety of protein
`
`

`

`W. Wang :International Journal of Pharmaceutics 203 (2000) 1–60
`
`8 s
`
`tabilizers are presented for cryo- and lyo-protec-
`tion, followed by discussions of their possible
`stabilization mechanisms.
`
`3.1. Stabilizers for cryo- and lyo-protection
`
`Nature protects life from freezing or osmotic
`shock by accumulating selected compounds to
`high concentrations (\1 M) within organisms.
`These accumulated compounds are known as cry-
`oprotectants and osmolytes, which are preferen-
`tially excluded from surfaces of proteins and act
`as structure stabilizers (Timasheff, 1993). How-
`ever, since the dehydration stress is different from
`those of freezing, many effective cryoprotectants
`or protein stabilizers in solution do not stabilize
`proteins during dehydration (drying). Some even
`destabilize proteins during lyophilization. For ex-
`ample, CaCl2 stabilized elastase (20 mg ml (cid:28) 1) in
`10 mM sodium acetate (pH 5.0), but caused the
`lyophilized protein cake to collapse and lose activ-
`ity (Chang et al., 1993).
`Similarly, effective lyophilization stabilizers (ly-
`oprotectants) may or may not stabilize proteins
`effectively during freezing. Therefore,
`in cases
`when a single stabilizer does not serve as both a
`cryoprotectant and a lyoprotectant, two (or more)
`stabilizers may have to be used to protect proteins
`from denaturation during lyophilization.
`
`3.1.1. Sugars:polyols
`Many sugars or polyols are frequently used
`nonspecific protein stabilizers in solution and dur-
`ing freeze-thawing and freeze-drying. They have
`been used both as effective cryoprotectants and
`remarkable lyoprotectants. In fact, their function
`as lyoprotectants for proteins has long been
`adopted by nature. Anhydrobiotic organisms (wa-
`ter content B1%) commonly contain high con-
`centrations
`(up
`to
`50%)
`of
`disaccharides,
`particularly sucrose or trehalose, to protect them-
`selves (Crowe et al., 1992, 1998).
`The level of stabilization afforded by sugars or
`polyols generally depends on their concentrations.
`A concentration of 0.3 M has been suggested to
`be the minimum to achieve significant stabiliza-
`tion (Arakawa et al., 1993). This has been found
`to be true in many cases during freeze-thawing.
`
`For example, freezing rabbit muscle LDH in wa-
`ter caused 64% loss of protein activity, and in the
`presence of 5, 10 or 34.2% sucrose, the respective
`losses were 27, 12, and 0% (Nema and Avis,
`1992). Other sugars or polyols that can protect
`LDH during freeze-thawing to different degrees
`include lactose, glycerol, xylitol, sorbitol, and
`mannitol, at 0.5–1 M (Carpenter et al., 1990).
`Increasing trehalose concentration gradually in-
`creased the recovery of PFK activity during
`freeze-thawing and the recovery reached a maxi-
`mum of 90% at about 300 mg ml(cid:28) 1 (Carpenter et
`al., 1990). A similar stabilizing trend was also
`observed for sucrose, maltose, glucose, or inositol
`(Carpenter et al., 1986).
`Since freezing is part of the freeze-drying pro-
`cess, high concentrations of sugars or polyols are
`often necessary for lyoprotection. These examples
`include the protection of chymotrypsinogen in the
`presence of 300 mM sucrose (Allison et al., 1996),
`complete inhibition of acidic fibroblast growth
`factor (aFGF) aggregation by 2% sucrose (Volkin
`and Middaugh, 1996), increase in glucose-6-phos-
`phate dehydrogenase (G6PDH) activity from 40
`to about 90% by 5.5% sugar mixture (glu-
`cose:sucrose(cid:30)1:10, w:w) (Sun et al., 1998), com-
`plete recovery of LDH by either 7% sucrose or 7%
`raffinose, a trisaccharide (Moreira et al., 1998),
`significant improvement of PFK recovery by 400
`mM trehalose (Carpenter et al., 1993), and com-
`plete pro

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