throbber
Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 1 of 17 PageID #: 30637
`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 1 of 17 PageID #: 30637
`
`(cid:40)(cid:59)(cid:43)(cid:44)(cid:37)(cid:44)(cid:55)(cid:3)(cid:23)(cid:3)
`
`EXHIBIT 4
`
`(cid:3)
`
`

`

`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 2 of 17 PageID #: 30638
`
`Appl Microbiol Biotechnol (2005) 68: 283–291
`DOI 10.1007/s00253-005-1980-8
`
`MINI-REVIEW
`
`Michael Butler
`Animal cell cultures: recent achievements and perspectives
`in the production of biopharmaceuticals
`
`Received: 11 February 2005 / Revised: 23 March 2005 / Accepted: 31 March 2005 / Published online: 16 April 2005
`# Springer-Verlag 2005
`
`Abstract There has been a rapid increase in the number
`and demand for approved biopharmaceuticals produced
`from animal cell culture processes over the last few years.
`In part, this has been due to the efficacy of several hu-
`manized monoclonal antibodies that are required at large
`doses for therapeutic use. There have also been several
`identifiable advances in animal cell technology that has
`enabled efficient biomanufacture of these products. Gene
`vector systems allow high specific protein expression and
`some minimize the undesirable process of gene silencing
`that may occur in prolonged culture. Characterization of
`cellular metabolism and physiology has enabled the design
`of fed-batch and perfusion bioreactor processes that has
`allowed a significant improvement in product yield, some
`of which are now approaching 5 g/L. Many of these pro-
`cesses are now being designed in serum-free and animal-
`component-free media to ensure that products are not
`contaminated with the adventitious agents found in bovine
`serum. There are several areas that can be identified that
`could lead to further improvement in cell culture systems.
`This includes the down-regulation of apoptosis to enable
`prolonged cell survival under potentially adverse condi-
`tions. The characterization of the critical parameters of
`glycosylation should enable process control to reduce the
`heterogeneity of glycoforms so that production processes
`are consistent. Further improvement may also be made by
`the identification of glycoforms with enhanced biological
`activity to enhance clinical efficacy. The ability to produce
`the ever-increasing number of biopharmaceuticals by ani-
`mal cell culture is dependent on sufficient bioreactor capacity
`in the industry. A recent shortfall in available worldwide
`culture capacity has encouraged commercial activity in
`contract manufacturing operations. However, some ana-
`
`M. Butler (*)
`Department of Microbiology,
`University of Manitoba,
`Buller Building,
`Winnipeg, Manitoba, Canada, R3T 2N2
`e-mail: butler@cc.umanitoba.ca
`Tel.: +1-204-4746543
`Fax: +1-204-4747603
`
`lysts indicate that this still may not be enough and that future
`manufacturing demand may exceed production capacity as
`the number of approved biotherapeutics increases.
`
`Introduction
`
`Although animal cell cultures have been important at a
`laboratory scale for most of the last 100 years, it was the
`initial need for human viral vaccines in the 1950s (par-
`ticularly for poliomyelitis) that accelerated the design of
`large-scale bioprocesses for mammalian cells (Kretzmer
`2002). These processes required the use of anchorage-de-
`pendent cells and the modern version of this viral vaccine
`technology currently employs microcarrier support sys-
`tems that can be used in pseudosuspension cultures de-
`signed in stirred tank bioreactors.
`However, more recently the enhanced interest in mam-
`malian cell culture bioprocesses is associated with re-
`combinant protein technology developed in the 1970s and
`1980s. The first human therapeutic protein to be licensed
`from this technology in 1982 was recombinant insulin (Hu-
`mulin from Genentech) but the relative structural simplicity
`of this molecule allowed its large-scale production to be
`developed in Escherichia coli, which is fast growing and
`robust compared to mammalian cells. It was soon realised
`that the subsequent targets for recombinant therapeutics
`were more complex and required the post-translational
`metabolic machinery only available in eukaryotic cells. At
`the present time there are up to 30 licensed biopharmaceu-
`ticals produced from mammalian cell bioprocesses (Walsh
`2003; Pavlou 2003, Molowa and Mazanet 2003). These are
`defined as recombinant proteins, monoclonal antibodies
`and nucleic acid-based products. Since 1996, the chimeric
`and humanized monoclonal antibodies have dominated this
`group with such blockbuster products as Rituxan, Remi-
`cade, Synagis and Herceptin (Brekke and Sandie 2003;
`Pavlou 2004). A chimeric antibody (e.g. Rituxan) consists
`of a molecular construct in which the mouse variable region
`is linked to the human constant region. A further step to
`humanizing an antibody can be made by replacement of the
`APPX 0220
`
`

`

`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 3 of 17 PageID #: 30639
`
`284
`
`murine framework region, leaving only the complementar-
`ity determining regions (CDRs) that are of murine origin.
`These hybrid construct molecules are far less immunogenic
`than their murine counterparts and have serum half-lives of
`up to 20 days.
`In this review, key factors for the recent achievements
`in the production of biopharmaceuticals from animal cell
`culture processes are discussed.
`
`Cell line transfection and selection
`
`The ability to produce and select for a high-producing
`animal cell line is key to the initial stages of the devel-
`opment of a cell culture bioprocess (Wurm 2004; Andersen
`and Krummen 2002). Chinese hamster ovary (CHO) cells
`have become the standard mammalian host cells used in the
`production of recombinant proteins, although the mouse
`myeloma (NS0), baby hamster kidney (BHK), human em-
`bryonic kidney (HEK-293) or human-retina-derived (PER-
`C6) cells are alternatives. All these cell lines have been
`adapted to grow in suspension culture and are well suited
`for scale-up in stirred tank bioreactors. The advantage of
`CHO and NSO cells is that there are well-characterised
`platform technologies that allow for transfection, amplifi-
`cation and selection of high-producer clones. Transfection
`of cells with the target gene along with an amplifiable gene
`such as dihydrofolate reductase (DHFR) or glutamine syn-
`thetase (GS) has offered effective platforms for expression
`of the required proteins. In these systems, selective pressure
`is applied to the cell culture with an inhibitor of the DHFR
`or GS enzymes that causes an increase in the number of
`copies of the transfected genes including the target gene.
`The DHFR system is routinely used with CHO cells
`−
`). The target gene
`deficient in the DHFR activity (DHFR
`is delivered to the cells along with the DHFR marker gene,
`usually on the same plasmid vector (Gasser et al. 1982;
`Lucas et al. 1996). The expression vector normally con-
`tains a strong viral promoter to drive transcription of the
`recombinant gene and this is delivered into the cells by
`one of a number of possible non-viral transfer techniques.
`These include calcium phosphate, electroporation, lipofec-
`tion or polymer-mediated gene transfer. The transfected
`cells are selected in media requiring the activity of DHFR
`for nucleotide synthesis and cell growth. Exposure of the
`cells to several rounds of gradually increased concentra-
`tions of the DHFR enzyme inhibitor, methotrexate, pro-
`motes amplification of the DHFR and the co-transfected
`target gene. Methotrexate treatment enhances specific pro-
`tein production following an increased gene copy number,
`which can be up to several hundred in selected cells.
`The glutamine synthetase expression system is an alter-
`native that works as a dominant selectable marker, which
`is an advantage because this does not require the use of
`specific mutant cells (Bebbington et al. 1992). CHO cells
`that contain endogenous enzyme activity can be used al-
`though NS0 cells are preferred because of the absence of
`GS. This means that a lower activity of an enzyme inhib-
`
`itor can be used for selection and amplification in NS0. GS
`enzyme allows the synthesis of glutamine intracellularly
`and so the transfected cells are selected in a glutamine-free
`media. The added advantage of this is that the cell cultures
`produce less ammonia, which is a potentially toxic met-
`abolic by-product of mammalian cells that affects protein
`glycosylation and may inhibit cell growth. Gene amplifi-
`cation in this system is mediated by methionine sulphox-
`imine (MSX), which is required at concentrations of 10–
`100 μM to provide clones with amplified genes and suf-
`ficiently high specific productivities. Typically copy num-
`bers of only four to ten genes per cell are found in these
`cells but they give as high expression levels as the cells
`from the DHFR system. The advantage of this is that the
`GS high-producer clones can be produced in around 3
`months, which is half the time it takes for the selection of
`DHFR clones.
`High yields of recombinant proteins can also be pro-
`duced from a human cell line, notably PER.C6, which was
`created by immortalizing healthy human embryonic retina
`cells with the E1 gene of adenovirus (Jones et al. 2003).
`This cell line has been well characterized and has been
`shown to be able to produce high levels of recombinant
`protein with relatively low gene copy numbers and with-
`out the need for amplification protocols. The added value
`of these cells is that they ensure the recombinant proteins
`produced receive a human profile of glycosylation.
`Screening for a high-producer clone can be a lengthy
`process that depends upon assaying the secreted proteins
`to determine productivities of all candidate clones. High-
`throughput selection systems have been devised based on
`rapid assays or the use of flow cytometry to identify clones
`that have an appropriate product marker on the cell surface
`(Borth et al. 2001; Carroll and Al-Rubeai 2004). A cell clone
`−1 day
`−1 can
`with a specific productivity of up to 10 pg cell
`be produced fairly routinely for recombinant protein pro-
`duction. However, higher specific productivity (up to 90 pg
`
`−1 day−1) may be possible with improvements in vector
`cell
`technology and further understanding of the parameters
`that control protein expression in the cell (Wurm 2004).
`
`Stability of gene expression
`
`The stability of selected clones over long-term culture is a
`critical parameter for commercial production (Kim et al.
`1998). The application of selective pressure such as meth-
`otrexate in the case of the DHFR selection system causes
`gene amplification but a proportion of these genes are
`unstable and removal of the selective agent, as is necessary
`in production cultures, results in a gradual loss of the gene
`copy number. Fann et al. (2000) reported the stepwise ad-
`aptation of tissue plasminogen activator-producing CHO
`cell lines to 5 μM of methotrexate, which resulted in a
`maximum specific recombinant protein production of 43 pg
`
`−1 day−1, but on removal of the methotrexate the max-
`cell
`
`−1 day−1 within
`imum productivity decreased to 12 pg cell
`40 days. This decrease in productivity could be correlated
`
`APPX 0221
`
`

`

`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 4 of 17 PageID #: 30640
`
`with a reduction of gene copy number for individual clones.
`Barnes et al. (2004) studied the stability of antibody ex-
`pression from NS0 cells amplified with the GS system.
`They reported that there was a loss of mRNA for the re-
`combinant protein over long-term culture but this was only
`reflected in a decrease in protein expression if the mRNA
`was below a threshold level. This indicates that selection for
`clones for high levels of recombinant mRNA may be useful
`as a predictor of stable protein production. Above a sat-
`uration level of mRNA it is argued that the limitation to
`protein expression resides in the translational/ secretory
`machinery of the cell.
`Production of a high-producer clone is dependent on the
`integration of the expression vector in the host cell genome.
`The site of integration has a major effect on the transcription
`rate of the recombinant gene and progressive gene silencing
`can occur over successive culture passages after clonal
`isolation. This is thought to be associated with the spread of
`heterochromatin structure, which is condensed and tran-
`scriptionally silent. Most expression systems cause random
`transgene integration into the host cell and this leads to
`positional effects that cause variable expression and stabil-
`ity. However, the genetic control elements that are respon-
`sible for establishing a transcriptionally active transgene are
`not fully understood. Vectors in which the genes are flanked
`with insulators, boundary elements or ubiquitous chromatin
`opening elements may promote stable expression by insu-
`lating the transgene from positional effects of the chroma-
`tin. These elements that can be incorporated into expression
`vectors include matrix (or scaffold) attachment regions
`(MAR or SAR) that allow open chromatin structure to be
`maintained. This can allow higher efficiency of expression
`of the integrated genes (Kim et al. 2004). Gene regulatory
`elements associated with ubiquitously expressed house-
`keeping genes have been recently isolated (Antoniou et al.
`2003). These regulatory elements appear to confer a dom-
`inant chromatin opening function and give rise to an ability
`to resist transgene silencing. These ubiquitous chromatin
`opening elements (UCOE) have also been incorporated into
`transgene vectors to prevent gene silencing and give con-
`sistent, stable and high-level gene expression irrespective
`of the chromosomal integration site (Haines et al. 2004).
`
`Culture modes
`
`A producer cell clone may be grown in batch cultures to
`above 106 cells/mL over 3–4 days to allow synthesis and
`product secretion. The limits for growth and production
`are related to the accumulation of metabolic by-products,
`such as ammonia and lactate, or the depletion of nutrients,
`such as glucose or glutamine (Butler and Jenkins 1989).
`The growth of cells can be extended if these limitations are
`addressed through perfusion culture in which the constant
`supply of nutrients and the removal of media can lead to
`cell densities of at least 107 cells/mL (Butler et al. 1983).
`These principles have been extended to fed-batch cultures,
`which have been shown to be operationally simple, reli-
`able and flexible for multi-purpose facilities (Bibila and
`
`285
`
`Robinson 1995, Cruz et al. 2000; Xie and Wang 1997). The
`most successful strategies involve feeding concentrates of
`nutrients based upon the predicted requirements of the cells
`for growth and production. This can involve slow feeding of
`low concentrations of key nutrients. The maintenance of
`low concentration set points of the major carbon substrates
`enables a more efficient primary metabolism with leads to
`lower rates of production of metabolic by-products, such as
`ammonia and lactate. As a result the cells remain in a pro-
`ductive state over extended time frames. The strategic use of
`fed-batch cultures has enabled considerable enhancement
`of yields from these processes. This is often combined with
`a biphasic strategy of production in which cell proliferation
`is allowed in the first phase so that high cell densities ac-
`cumulate, followed by a phase in which cell division is
`arrested to allow cells to attain a high specific productivity.
`In this type of strategy, growth can be arrested by a decrease
`in culture temperature (Fox et al. 2004; Yoon et al. 2003).
`By directly supplying cells with a balanced nutrient feed, a
`fed-batch culture can now be expected to yield upwards of
`2 g/L of recombinant protein, which is probably at least
`tenfold higher than the maximum that could be expected by
`a simple batch culture in standard culture medium.
`Animal cell cultures are normally grown in stainless
`steel, stirred tank bioreactors that are designed with im-
`pellers that minimize shear forces (Kretzmer 2002). Pro-
`ducer cells can be made to be sufficiently robust in this
`environment if they are provided with suitable growth
`media and gas sparging is carefully controlled. The capacity
`of commercial bioreactors for animal cells has gradually
`increased over the past two decades, with capacities now
`reported up to 20,000 L from some of the larger biopharma-
`ceutical companies. Airlift bioreactors have also been ap-
`plied to large-scale animal cells and these have been shown
`to be efficient for protein production.
`Perfusion cultures are more demanding to set up at a
`large scale but they have the potential advantage of allow-
`ing a continuous stream of product over several weeks or
`even months (Mercille et al. 2000). A further advantage is
`the rapid removal of any potentially labile products from
`the culture environment. An effective cell separator will
`allow the protein-containing media to be fed directly into a
`chromatography column suitable for extraction and down-
`stream processing (Shirgaonkar et al. 2004; Castilho and
`Medronho 2002; Wen et al. 2000). A further advantage of
`this mode of culture is that the bioreactor may be up to ten
`times smaller for the production of the same quantity of
`product (Ryll et al. 2000).
`This area of bioprocess design will become of even
`greater importance as some of the first-generation block-
`buster drugs (e.g. erythropoietin, human growth hormone
`and α-interferon) start being produced as generics (Walsh
`2003). Eleven biopharmaceuticals with combined annual
`sales of $13.5 billion lose patent protection in 2006 (Walsh
`2003). The challenge then will be to produce bioequiva-
`lents in efficient low-cost bioprocesses.
`There are several challenging areas of bioprocess de-
`velopment that are required to be addressed to ensure the
`future success of animal cell culture processes. These in-
`
`APPX 0222
`
`

`

`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 5 of 17 PageID #: 30641
`
`286
`
`clude serum-free media, apoptosis, glycomics and the ca-
`pacity crunch.
`
`Apoptosis
`
`Serum-free media
`
`Bovine serum was used as a supplement of cell culture
`media for several decades. It is a rich source of hormones,
`growth factors and trace elements that promote rapid cell
`growth and also its high albumin content ensures that the
`cells are well protected from potentially adverse condi-
`tions such as pH fluctuations or shear forces. However, the
`composition of serum is variable and undefined, which
`leads to inconsistent growth and productivity. Early at-
`tempts to develop serum-free substitutes incorporated such
`components as insulin, transferrin, albumin and cholesterol.
`However, the mad cow crisis in the beef industry alerted a
`concern for the use of animal serum and any other animal-
`derived components in the production of biotherapeutics.
`This has now led to a strong demand for cell culture for-
`mulations that are free of all animal components. The chal-
`lenge that this demand poses is to be able to identify effective
`substitutes for all the growth-promoting factors that are
`present in serum. It turns out that producer cell lines are quite
`fastidious in their growth requirements and that such re-
`quirements vary considerably from one cell line to another.
`Therefore, it has not been possible to design a single serum-
`free formulation to act as a serum substitute suitable for the
`growth of all cell lines. In fact even different clones of CHO
`cells may require different formulations for optimal growth.
`This has given rise to a strong drive for the development of
`serum-free and animal-component-free formulations that are
`tailored to the needs of specific producer cell lines.
`There are several strategies that can be used to design
`these formulations. Combinations of standard basal media
`may be tested to determine those that result in good cell
`growth and productivity at minimal serum levels. In some
`cases metabolic analysis may help in media design. For
`example, NS0 myeloma cells lack a functional pathway
`for cholesterol synthesis and so cholesterol is required as a
`lipoprotein supplement in the medium (Gorfien et al. 2000).
`Protein hydrolysates from non-animal sources have been
`found to provide good growth promotion in some culture
`systems (Sung et al. 2004). Analysis of the depletion of
`media components may lead to the identification of specific
`nutrients that may be required at higher supplement levels
`or for inclusion in feeding regimes. Statistical approaches
`can be used, such as the Plackett–Burman experimental
`design, so that mixtures of components may be tested si-
`multaneously in matrix experiments of growth and produc-
`tivity (Castro et al. 1992). Another original approach is the
`identification by microarray analysis of specific receptors
`expressed during cell growth, so that corresponding ligands
`may be incorporated into the medium (Donahue 2004).
`These approaches are presently being used by various
`specialized media companies to customise animal-compo-
`nent-free formulations for the production of the plethora of
`recombinant proteins that are being introduced into the
`market.
`
`Cell cultures are often terminated because of cell death
`that may be caused by one of several factors including nu-
`trient depletion, metabolic by-product accumulation, exces-
`sive shear forces or hypoxia. Cell death may be by necrosis
`caused by extreme conditions resulting in physical damage
`to the cells. Alternatively and more commonly, cell death in
`a bioreactor occurs by apoptosis, which is a form of pro-
`grammed cell death regulated through a cellular cascade of
`activities in response to one of the factors mentioned above
`(Arden and Betenbaugh 2004). Characteristic changes in-
`clude chromatin shrinkage followed by membrane bleb-
`bing and the formation of apoptotic bodies. The DNA of
`the cell is fragmented and this can be the basis of an assay
`to quantify apoptosis in a cell population.
`It is of considerable value to be able to prevent or inhibit
`apoptosis in culture in order to extend the time of high cell
`viability and prolong protein production. There are two
`strategies that can be used for this. The cellular environment
`can be manipulated through media supplementation or
`the intracellular environment can be modified by genetic
`engineering.
`Nutrient feeding can provide protection and this is nor-
`mally used as the first preventive measure to control the
`cellular environment to delay apoptosis. Serum is known to
`contain unidentified anti-apoptotic factors that can offer
`protection (Zanghi et al. 1999). However, the serum-free
`formulations that are required for production processes
`make the cells more vulnerable to apoptosis. Some supple-
`ments such as suramin (Zanghi et al. 2000) or insulin growth
`factor (Sunstrom et al. 2000) may provide independent anti-
`apoptotic protection in serum-free cultures. There are also
`other specific caspase inhibitors available to suppress apo-
`ptosis (Tinto et al. 2002) but their expense in large-scale
`cultures is likely to be prohibitive.
`Genetic strategies involve the transfection of anti-apo-
`ptotic genes such as bcl-2 or bcl-x2 into a host cell. The
`expression of the corresponding proteins inhibits the release
`of pro-apoptotic molecules from the mitochondria and may
`prolong the viability of the cell. This strategy has been
`shown to work for several cell lines, which have shown
`higher viabilities and improved robustness under conditions
`that would normally be expected to cause apoptosis (Tey
`et al. 2000; Mastrangelo et al. 2000; Kim and Lee 2002).
`
`Glycosylation
`
`Animal cells are used for biomanufacture because of their
`capabilities of adding carbohydrates (glycans) to synthe-
`sised proteins (Butler 2004). These are produced as pools of
`different glycoforms with varying glycan structures at-
`tached to a single peptide backbone with a known amino
`acid sequence. The basic protein structures can be controlled
`and directed by the expression of appropriate genetic se-
`quences. However, controlling the pool of glycan structures
`(glycomics) that occupy a recombinant protein is still dif-
`ficult. Variations may be found in the site occupancy (mac-
`
`APPX 0223
`
`

`

`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 6 of 17 PageID #: 30642
`
`roheterogeneity) or in the structure of attached glycans
`(microheterogeneity).
`For the production of a recombinant protein as a bio-
`therapeutic it is essential to ensure that a consistent glyco-
`sylation profile is maintained between batches (Restelli
`and Butler 2002). However, this may not be so easy to con-
`trol given that the extent of glycosylation may decrease over
`time in a batch culture (Curling et al. 1990). This is likely to
`be due to the depletion of nutrients, particularly glucose or
`glutamine, which have been shown to limit the glycosyla-
`tion process (Hayter et al. 1992; Nyberg et al. 1999). Fed-
`batch strategies should also be designed to ensure that the
`concentrations of these key nutrients do not decrease to
`a critical level that could compromise protein glycosyla-
`tion (Xie and Wang 1997). These lower levels were found to
`be <0.1 mM glutamine and <0.7 mM glucose for the pro-
`duction of γ-interferon from CHO cells (Chee et al. 2005).
`In one report it is suggested that site occupancy could vary
`with the growth state of cells and correlates with the frac-
`tion of cells in the G0/G1 phase of the cell cycle (Andersen
`et al. 2000). This suggests a mechanism by which glyco-
`sylation efficiency improves at a reduced rate of protein
`translation.
`Culture parameters such as nutrient content, pH, tem-
`perature, oxygen or ammonia, may have a significant effect
`on the distribution of glycan structures found on the re-
`sulting recombinant protein (microheterogeneity). This of
`course is of major concern in trying to produce consistent
`biopharmaceuticals. Decreased sialylation or altered pat-
`terns of glycan branching occur when the ammonia level
`accumulates in culture (Andersen and Goochee 1994;
`Zanghi et al. 1998; Yang and Butler 2000). Non-optimal pH
`conditions (<6.9 and >8.2) have also been shown to alter the
`pattern of glycosylation (Rothman et al. 1989; Borys et al.
`1993). Reduced terminal galactosylation has been shown in
`the glycans of immunoglobulin (IgG) produced under low
`oxygen conditions (Kunkel et al. 1998). Nabi and Dennis
`(1998) observed an increase in the polylactosamine content
`of a protein produced at lower temperatures and attributed
`this to changes in the transit time through the Golgi.
`The pattern of protein glycosylation is dependent on the
`expression of various glycosyltransferase enzymes that are
`present in the Golgi of the cell. Differences in the relative
`activity of these enzymes among species can account for
`significant variations in structure. In one systematic study
`of glycan structures of IgG produced from cells of 13
`different species significant variation was found in the
`proportion of terminal galactose, core fucose and bisecting
`GlcNAc (Raju et al. 2000). The structure of sialic acid
`may also vary, with N-glycolyl-neuraminic acid (NGNA)
`found in goat, sheep and cows rather than the N-acetyl-
`neuraminic acid (NANA) found in humans. NGNA is the
`predominant sialic acid in mice, but CHO-produced gly-
`coproteins have predominantly NANA, although a small
`proportion (up to 15%) of NGNA can occur (Baker et al.
`2001). These differences in glycan structure are important
`tential immunogenicity of these structures in humans. Mouse
`cells express the enzyme α1,3 galactosyltransferase, which
`
`287
`generates Galα1,3-Galβ1,4-GlcNAc residues that are high-
`ly immunogenic in humans (Jenkins et al. 1996). Fortu-
`nately, this enzyme appears to be inactive in CHO and BHK
`cells, which are the most commonly used cell lines for the
`production of recombinant proteins. However, both CHO
`and BHK show differences in their potential for glycosy-
`lation compared to human cells. The sialyl transferase en-
`zyme (α2,6 ST) that normally provides an α2,6 linkage for
`terminal sialic acid in glycoproteins produced in humans is
`absent in the hamster and thus CHO and BHK cells pro-
`duce exclusively α2,3 terminal sialic acid residues. Further-
`more, the absence of a functional α1,3 fucosyltransferase
`in CHO cells prevents the addition of peripheral fucose
`residues and the absence of N-acetylglucosaminyltransfer-
`ase III (Gn TIII) prevents the addition of bisecting GlcNAc
`to glycan structures (Jenkins and Curling 1994). However,
`these differences in glycosylation potential between CHO
`and human cells do not appear to result in glycoproteins
`that are immunogenic. Natural human erythropoietin (EPO)
`consists of a mixture of sialylated forms: 60% are 2,3 linked
`and 40% are 2,6 linked. Because of the restricted sialyla-
`tion capacity of CHO cells, the commercially available
`EPO is sialylated entirely via the α2,3 linkages. Never-
`theless, recombinant EPO produced from CHO cells has
`proven to be a highly effective therapeutic agent with no
`evidence of an adverse physiological effect due to the
`structural differences in terminal sialylation.
`The production of specific protein glycoforms may allow
`the possibility of even more efficacious drugs (Shriver et al.
`2004). Functional glycomics is an expanding area of sci-
`ence that attempts to understand the physiological function
`of specific carbohydrate groups. This approach established
`the importance of the sialylation of EPO with the discovery
`that the removal of sialic acid groups from the glycans re-
`sulted in a significantly reduced half-life in the blood stream
`(Erbayraktar et al. 2003). Protein engineering has allowed
`the creation of a modified EPO with two extra glycan at-
`tachment sites and with the potential to incorporate eight
`extra sialic acid groups per molecule. This has led to a new-
`generation EPO called darbepoetin, which has a three times
`higher drug half-life (Egrie et al. 2003). This strategy of
`enhancing the half-life of a biotherapeutic has also been
`successful for other recombinant proteins such as follicle-
`stimulating hormone (Perlman et al. 2003) and thyroid-
`stimulating hormone (Thotakura et al. 1991).
`Structural changes of glycans can also be brought about
`by metabolic engineering of the host cell line. This in-
`cludes gene knockout of already expressed glycosyltrans-
`ferases or the insertion of novel activities (Weikert et al.
`1999). The presence of a bisecting N-acetylglucosamine
`(Umana et al. 1999; Davies et al. 2001) or the absence of
`fucose (Shields et al. 2002; Shinkawa et al. 2003; Okazaki
`et al. 2004) in the conserved glycan of an IgG antibody
`has been shown to enhance attachment to Fc receptors and
`result in an increase in antibody-dependent, cell-mediated
`cytotoxicity (ADCC). This has been of value in the design
`of antibody therapeutics. For example, recent work with
`Herceptin, which is a novel humanized antibody approved
`
`APPX 0224
`
`

`

`Case 1:18-cv-00924-CFC Document 399-4 Filed 10/07/19 Page 7 of 17 PageID #: 30643
`
`288
`
`for the treatment of breast cancer, has shown that a glyco-
`form with no fucose has a 53 times higher binding ca-
`pacity to an Fc receptor that triggers its therapeutic activity
`(Shinkawa et al. 2003). This enhancement of ADCC al-
`lows the antibody to be effective at lower doses. Afuco-
`sylated antibodies can be produced from cells in which the
`gene for fucosyl transferase has been removed by gene
`knockout technology.
`Complete glycosylation of recombinant proteins is usu-
`ally associated with maximisation of galactosylation and
`sialylation. Often these two processes are incomplete and
`this gives rise to considerable glycan structural variation.
`CHO cells can be engineered with a combination of human
`β1,4-galactosyltransferase and α2,3-sialyltransferase to en-
`sure high activities of these enzymes. The recombinant pro-
`teins produced by these cells exhibited greater homogeneity
`compared to controls and increased terminal sialic acid re-
`sidues (Weikert et al. 1999). An alternative approach in-
`volves glycoengineering of the proteins in vitro (Raju et al.
`2001). Preparations of these terminal transferase enzymes
`can be immobilized so that glycoproteins can be galacto-
`sylated and sialylated in the presence of appropriate ga-
`lactose and sialic acid donors.
`
`The capacity crunch
`
`With an increase in the number and demand for recombi-
`nant biopharmaceuticals, there is a requirement for greater
`biomanufacturing capacity. This created a major problem in
`2001 when the demand for Enbrel, a recombinant fusion
`protein commercialized by Immunex for the treatment of
`rheumatoid arthritis, exceeded expectations. However, there
`was insufficient large-scale culture manufacturing capacity
`to meet this clinical demand, even by contract manufactur-
`ers available

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