throbber
HHS Public Access
`Author manuscript
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Published in final edited form as:
`Annu Rev Biomed Eng. 2015 ; 17: 191–216. doi:10.1146/annurev-bioeng-071114-040733.
`
`Advances in Antibody Design
`
`Kathryn E. Tiller and Peter M. Tessier
`Center for Biotechnology and Interdisciplinary Studies, Isermann Department of Chemical and
`Biological Engineering, Rensselaer Polytechnic Institute, Troy, New York 12180
`
`Abstract
`The use of monoclonal antibodies as therapeutics requires optimizing several of their key
`attributes. These include binding affinity and specificity, folding stability, solubility,
`pharmacokinetics, effector functions, and compatibility with the attachment of additional antibody
`domains (bispecific antibodies) and cytotoxic drugs (antibody–drug conjugates). Addressing these
`and other challenges requires the use of systematic design methods that complement powerful
`immunization and in vitro screening methods. We review advances in designing the binding loops,
`scaffolds, domain interfaces, constant regions, post-translational and chemical modifications, and
`bispecific architectures of antibodies and fragments thereof to improve their bioactivity. We also
`highlight unmet challenges in antibody design that must be overcome to generate potent antibody
`therapeutics.
`
`Keywords
`IgG; scFv; VH; Fab; CDR; complementarity-determining region
`
`1. INTRODUCTION
`
`Antibodies are affinity proteins that play a central role in humoral immunity. Their ability to
`bind to foreign invaders with high affinity and specificity is central to their function. Equally
`important is their ability to serve as adaptor molecules and recruit immune cells for various
`effector functions. There are five main classes of antibodies with diverse functions:
`immunoglobulin (Ig)A IgD, IgE, IgG, and IgM (1). IgGs are the most abundant class of
`antibodies, as they constitute approximately 75% of the serum immunoglobulin repertoire.
`There are four subclasses of IgGs, which vary in their abundance and ability to elicit specific
`effector functions.
`
`The overall architecture of IgGs is conserved across its four subclasses, and consists of two
`light chains and two heavy chains (Figure 1). The light chains contain variable (VL) and
`constant (CL) domains, and the heavy chains contain one variable (VH) and three constant
`(CH1, CH2, and CH3) domains. One notable difference between IgG subclasses is the
`location of the disulfide bonds between CH1 and CL (which link the heavy and light chains)
`
`DISCLOSURE STATEMENT
`P.M.T. has received honorariums or consulting fees, or both, from MedImmune, Eli Lilly, Bristol-Myers Squibb, Janssen Biotech,
`Merck, Genentech, Amgen, Pfizer, Adimab, Abbott, AbbVie, Bayer, Roche, DuPont, and Schrödinger.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 1
`
`

`

`Tiller and Tessier
`
`Page 2
`
`and the number of disulfide bonds in the hinge region (which link the heavy chains). The
`multidomain nature of IgGs elegantly divides their bioactivity into different subdomains.
`The antigen-binding fragment (Fab) contains both variable domains, and mediates antigen
`recognition via six peptide loops known as the complementarity-determining regions
`(CDRs). In contrast, the crystallizable fragment (Fc) contains the constant domains (CH2
`and CH3) that mediate effector function by binding to immunological receptor molecules
`such as complement proteins and Fc receptors.
`
`The multifunctional nature of IgGs is only one of the many reasons for the widespread
`interest in using monoclonal antibodies (mAbs) as therapeutics. The availability and
`refinement of robust methods for identifying and generating human mAbs, such as
`immunization and in vitro screening methods (2), have also contributed greatly to the
`interest in antibody therapeutics. In addition, mAbs typically display excellent
`pharmacokinetics (long circulation times), low toxicity and immunogenicity (for human or
`humanized mAbs), and high stability and solubility. It is also notable that the simplicity of
`expressing and purifying many different mAbs using similar platform processes is highly
`attractive from a manufacturing perspective and has enabled the production of a staggering
`number of different mAbs that are in clinical trials (3).
`
`Nevertheless, there are many challenges in generating mAbs for therapeutic applications. At
`the discovery stage, immunization affords limited control over antibody affinity and
`specificity due to the difficulty in controlling antigen presentation to the immune system. In
`vitro methods, such as phage and yeast surface display, enable improved control over
`antigen presentation. However, these display methods are limited by their need to screen
`large libraries, their typical use of antibody fragments instead of full-length antibodies, and
`their reduced quality-control mechanisms relative to mammalian systems. Moreover,
`antibodies identified via either immunization or display methods have variable and difficult-
`to-predict solubilities and viscosities at the high concentrations required for subcutaneous
`delivery (4, 5). Antibody aggregation is particularly concerning due to the potential
`immunogenicity of such aggregates (6), and abnormally high viscosity can prevent mAbs
`from being delivered via the subcutaneous route (7). It is also challenging to optimize
`bispecific antibodies that typically combine binding domains from different parent
`antibodies, given the large number of possible molecular architectures as well as the
`complex effects that these nonstandard antibody formats can have on antibody stability.
`Moreover, developing effective antibody–drug conjugates is extremely challenging due to
`the need to optimize the linker and conjugation chemistry as well as the location and number
`of attached drug molecules. Finally, it is difficult to engineer antibodies with the specific
`types and levels of effector functions that are optimal for a given therapeutic application.
`
`Although each of these challenges can be addressed through screening a large number of
`antibody variants, it is impractical to use such screening methods alone to address many of
`the challenges encountered in developing potent therapeutic antibodies. Attempts to
`optimize each antibody property sequentially are limited by the fact that improving one
`antibody attribute (such as binding affinity) can lead to defects in other attributes (such as
`solubility). Attempts to simultaneously optimize multiple antibody properties using
`mutagenesis and screening methods require libraries that are prohibitively large.
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 2
`
`

`

`Tiller and Tessier
`
`Page 3
`
`2. OVERVIEW OF APPROACHES FOR DESIGNING ANTIBODIES
`
`The complexity of optimizing several different antibody attributes (summarized in Figure 2)
`using traditional immunization and screening methods has led to intense interest in
`developing antibody-design methods. The most important antibody attributes are binding
`affinity and specificity, which involve optimizing the variable domains and the CDRs in
`particular. Colloidal stability (solubility) and conformational (folding) stability are also
`critical attributes of antibodies because therapeutic mAbs must be soluble for high-
`concentration delivery and stable for long-term storage. This typically requires optimizing
`solvent-exposed residues for solubility and solvent-shielded residues for conformational
`stability. The effector functions of antibodies are also critical to their bioactivity, and can be
`tailored by manipulating the hinge and Fc regions.
`
`Another increasingly important antibody attribute—which is uncommon in natural
`antibodies—is bispecificity for either multiple antigens or multiple epitopes on the same
`antigen. Achieving bispecificity requires methods for combining multiple antibodies into a
`single one as well as optimizing the key attributes of conventional antibodies. A second
`nonconventional attribute of antibodies that continues to grow in importance is their
`bioactivity when attached to small-molecule drugs. Developing antibody–drug conjugates
`(ADCs) requires optimizing many aspects of the chemistries and linkers used to derivatize
`antibodies in addition to the other key attributes of conventional antibodies.
`
`This review highlights progress in designing and optimizing each of these key antibody
`attributes. Given the large size and complexity of antibodies, most design efforts have
`focused on redesigning or optimizing existing antibodies rather than on de novo design of
`new antibodies. These design methods vary greatly, and range from knowledge-based
`methods based on previous mutagenesis results to advanced computational methods based
`on first principles. A commonality of these diverse methods is that they attempt to guide the
`design of antibodies in a systematic manner to reduce the need for screening and
`immunization methods. We discuss these design methods and their application to improve
`the properties of antibodies that are critical for their activity and stability.
`
`3. ANTIBODY BINDING AFFINITY AND SPECIFICITY
`
`The most important property of antibodies is their ability to recognize targets with high
`affinity and specificity. This binding activity is largely mediated by the CDRs. Several
`innovative approaches have been developed for designing CDRs that range from de novo
`design methods to those that involve the redesign of existing antibodies. Some of these
`design methods have used motif-grafting approaches to mimic natural protein interactions,
`and directed evolution approaches to achieve specificities for difficult-to-target antigens.
`
`3.1. De Novo Design
`
`The holy grail of antibody design is to accurately and reliably predict the sequences of
`antibodies that will bind with high affinity and specificity based solely on the sequence or
`composition of the antigen. Toward this ambitious goal, a computational approach named
`OptCDR (Optimal Complementarity Determining Regions) has been developed for
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 3
`
`

`

`Tiller and Tessier
`
`Page 4
`
`designing the CDRs of antibodies to recognize specific epitopes on a target antigen (8). This
`method uses canonical structures to generate CDR backbone conformations that are
`predicted to interact favorably with the antigen. Amino acids are then chosen for each
`position in the CDRs using rotamer libraries, and this process is repeated many times to
`refine the backbone structures and amino acid sequences. This leads to the prediction of
`several sets of CDR sequences, which can be grafted onto antibody scaffolds for evaluation.
`
`This approach has been tested for developing antibody–antigen complexes involving a
`hepatitis C virus capsid peptide, fluorescein, and vascular endothelial growth factor (VEGF)
`(8). The investigators predicted mutations that are expected to increase binding affinity
`(which were not evaluated experimentally) as well as evaluated mutations that had been
`reported previously to improve binding affinity for fluorescein antibodies. There is some
`correlation between predictions by OptCDR and experimental data for fluorescein
`antibodies. It will be important to further evaluate the ability of OptCDR and closely related
`methods (9) to make de novo predictions of CDR mutations as well as entire CDR
`sequences that either generate or improve antibody binding.
`
`3.2. Design by Mimicking Natural Protein Interactions
`
`Another fruitful approach for designing antibodies with specific binding activities has been
`to mimic natural protein interactions. For example, Williamson and colleagues (10) designed
`antibodies to recognize misfolded conformers of the prion protein (PrP) by mimicking
`natural interactions between cellular PrP (PrPC) and its misfolded counterpart (PrPSc).
`Previous studies had found that PrP residues 96–104 and 133–158 govern the ability of
`PrPSc to catalyze misfolding of soluble PrPC (11, 12). This led to the hypothesis that grafting
`such PrP peptides into heavy chain CDR3 (HCDR3) of an IgG—which originally lacks PrP-
`binding activity—would create antibodies that specifically recognized PrPSc (10). Indeed,
`they found that antibodies with PrP residues 89–112 or 136–158 in HCDR3 bound to PrPSc
`with apparent affinities in the low nanomolar range (2–25 nM), and these same antibodies
`weakly interacted with PrPC. Follow-up studies also identified a third region near the N
`terminus of PrP (residues 19–33) that resulted in binding activity for grafted PrP antibodies
`(13). Interestingly, grafting peptides from other PrP regions (such as those from the C-
`terminal domain) into the same CDR loop failed to generate binding activity. Moreover, the
`antibodies grafted with PrP residues 19–33 and 89–112 appear to bind via electrostatic
`interactions because mutating positively charged residues to alanine in these grafted peptides
`eliminated binding.
`
`This exciting study raises the question of whether grafting peptides from other aggregation-
`prone proteins into the CDRs of antibodies would also lead to specific binding activity. Our
`lab recently tested this question using the Alzheimer’s Aβ42 peptide (14). There are two
`hydrophobic segments within Aβ (residues 17–21 and 30–42) that mediate amyloid
`formation and are located within the β-sheet core of Aβ fibrils (15). We posited that grafting
`these peptide segments into CDR3 of a single-domain (VH) antibody would lead to antibody
`domains with Aβ-specific binding activity. Indeed, we found that grafted VH domains
`displaying the central hydrophobic region of Aβ (residues 17–21) in CDR3 bound to Aβ
`fibrils with submicromolar affinity (300–400 nM), and they weakly bound to Aβ monomers
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 4
`
`

`

`Tiller and Tessier
`
`Page 5
`
`or oligomers (14). Interestingly, VH domains grafted with the hydrophobic C terminus of Aβ
`(residues 30–42) bound both Aβ fibrils and oligomers with submicromolar affinity (300–700
`nM) and weakly recognized Aβ monomers. We refer to these grafted antibodies as
`gammabodies (Grafted AMyloid-Motif AntiBODIES). We have also verified that this
`grafting approach can be applied to other amyloid-forming proteins, including α-synuclein
`(associated with Parkinson’s disease) and IAPP (associated with type 2 diabetes) (16).
`Nevertheless, future work will need to develop more systematic methods for selecting
`amyloidogenic peptides for grafting because we currently do not understand why some
`sequences mediate binding and others do not. Moreover, it will be important to evaluate how
`multiple CDR loops can be engineered to display amyloidogenic peptides on the surface of
`single- and multidomain antibodies to improve the affinities of these grafted antibodies.
`
`3.3. Semirational Design Combined with Directed Evolution Methods
`
`Despite these key advances in the de novo design of CDRs, it has been extremely
`challenging to use such rational approaches to generate antibodies with subnanomolar (or
`lower) dissociation constants. Nevertheless, several innovative approaches have been
`developed that involve designing some CDR residues while randomizing others, and
`screening such libraries using in vitro display methods to select variants with high binding
`affinity and specificity. One of the first examples of this hybrid approach was the design of
`antibody libraries specific for integrins (17). The RGD sequence (arginine-glycine-aspartate)
`was inserted in the middle of HCDR3, and three flanking residues were randomized on each
`side of the RGD sequence. In addition, cysteines were introduced at each edge of HCDR3 to
`constrain the loop, which was posited to be necessary to generate high affinity for antibody
`binding that is mediated primarily via a single CDR. The investigators displayed a Fab
`library with these HCDR3 sequences on the surface of phage and screened for binding to
`integrins. Impressively, several antibody variants were identified with subnanomolar binding
`affinities, and these antibodies retained the same binding epitope as natural integrin ligands.
`This and related work (18, 19) has demonstrated the potential of using natural protein
`interactions to guide the design of high quality antibody libraries.
`
`Another example of this hybrid approach is a method for generating antibodies that
`recognize post-translational modifications (20, 21). It is difficult to isolate antibodies that
`recognize chemical modifications such as phosphorylation, especially for phosphoserine and
`phosphothreonine, because of the relatively small size of their side chains. Therefore, the
`investigators sought to introduce a common phosphate-binding motif from proteins such as
`kinases into the CDRs of antibodies (Figure 3). By identifying an antibody with a CDR loop
`(HCDR2) that naturally displays a similar anion-binding motif, the investigators first
`confirmed that this antibody bound weakly to phosphorylated peptides. Next, the affinities
`and specificities of such antibodies for phosphorylated serine, threonine, and tyrosine were
`evolved by randomizing sites within the anion-binding motif. After mutants were identified
`by phage display with selective and improved affinity for each type of modification, CDR
`residues outside the phospho-binding pocket in HCDR2 as well as in LCDR3 and HCDR3
`were randomized, and antibodies were selected for binding to different phosphorylated
`targets. Impressively, this approach generated many phospho-specific antibodies for a wide
`range of target peptides with modified serine and threonine in addition to tyrosine. Although
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 5
`
`

`

`Tiller and Tessier
`
`Page 6
`
`the binding affinities are modest (40–5,000 nM), they are similar to or better than those for
`previously reported phospho-specific antibodies [see (20) and references therein]. More
`importantly, this innovative approach to designing antibody libraries addresses the
`challenging problem of achieving binding specificities to subtly different antigens that are
`difficult to obtain using immunization.
`
`3.4. Antibody Redesign and Optimization
`
`Much effort has focused on using design methods to improve the binding affinity of existing
`antibodies due to the complexity of de novo design. These redesign efforts are important
`because immunization typically yields antibodies with affinities that are not high enough for
`therapeutic applications, and directed evolution approaches are limited in their ability to
`identify multiple synergistic mutations, given the unrealistically large libraries that would be
`required.
`
`One interesting study demonstrated the potential of optimizing electrostatic interactions to
`improve the binding affinity of antibodies (22). The authors used physics-based methods to
`initially evaluate the binding energy of an anti-lysozyme antibody (D1.3); every possible
`mutation was evaluated for 60 CDR positions using the crystal structure of the complex as a
`starting point. Interestingly, the mutations predicted to be most favorable were those that
`introduced large side chains into the binding interface to maximize van der Waals
`interactions, yet most of these mutations failed to improve affinity when evaluated
`experimentally. Instead, the electrostatic component of the binding energy was found to be a
`better predictor of mutations that improve affinity. This approach led to significant
`improvements in affinity (of one to two orders of magnitude) for cetuximab and an anti-
`lysozyme antibody (D44.1).
`
`The success of this study stems largely from two key types of mutations within the CDRs.
`One is the elimination of residues with unsatisfied polar groups (e.g., side chains of
`asparagine or threonine) in which desolvation is not compensated for by favorable
`interactions (e.g., hydrogen bonds) in the bound state (22). By mutating such residues to
`small hydrophobic ones, the authors observed increased binding affinity. A second key type
`of mutation is either the introduction or removal of charged residues at sites within the
`CDRs that are peripheral to the residues that contact the antigen in the nonoptimized
`complex. This finding builds on previous work demonstrating that charged residues outside
`the antibody–antigen interface but within the CDRs can increase the on-rate (and thereby the
`affinity) of an anti-VEGF antibody (23). More generally, these findings are consistent with
`other studies revealing that the likelihood of identifying beneficial mutations is higher
`outside the initial binding interface (24–27), likely due to the reduced risk of disrupting
`existing antibody–antigen interactions. This (22) and related work (24–26, 28, 29) have
`demonstrated the potential of using existing methods for calculating electrostatic
`interactions to identify mutations that improve antibody affinity.
`
`Another study elegantly showed that it is not necessary to use the crystal structures of
`antibodies or antibody–antigen complexes to guide efforts to improve affinity or alter
`binding specificity (25). The investigators sought to redesign a dengue virus antibody (4E11)
`to be broadly neutralizing. This is particularly challenging because the starting structures
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 6
`
`

`

`Tiller and Tessier
`
`Page 7
`
`were unknown for both the 4E11 antibody alone or when bound to its antigen (domain III of
`the dengue E protein). Nevertheless, the authors used computational docking methods to
`generate structures of the 4E11 antibody bound to four variants of the dengue E protein.
`Notably, the poor interaction between 4E11 and its antigen from one serotype (type 4) stems
`from the loss of key interactions (such as salt bridges and hydrogen bonds) that are predicted
`to be present in binding interfaces of the antibody with antigens from types 1–3. These
`insights led to the identification of several mutations that improve 4E11 binding to type 4
`without reducing binding to types 1–3. By combining successful mutations, the affinity for
`type 4 was enhanced by more than two orders of magnitude without reducing affinity for
`other serotypes. Most of the successful mutations were charged or polar, and located at the
`periphery of the binding interface, which is generally consistent with findings from other
`studies (22–24, 26). This exciting study highlights the potential of using design methods to
`achieve unique binding specificities that are difficult to achieve using conventional discovery
`methods.
`
`4. ANTIBODY CONFORMATIONAL (FOLDING) STABILITY
`
`Another key attribute of antibodies is their folding stability (30, 31), which is critical for
`maintaining long-term activity. Folding stability is a particularly important consideration
`when altering the CDRs of antibodies to improve binding affinity and specificity. Several
`studies have demonstrated that either grafting CDRs from one antibody to another or
`mutating CDRs can significantly affect folding stability (32–36). More generally, mutations
`that are located outside of the CDRs can also significantly impact folding stability (37–40).
`These challenges necessitate design methods that can rationally stabilize antibodies without
`compromising binding affinity and other key attributes.
`
`The high degree of sequence similarity between different antibodies, as well as the large
`number of available antibody sequences and structures, has led to significant understanding
`in how to stabilize antibodies. These approaches can be classified as (a) knowledge-based,
`(b) statistical, and (c) structure-based methods. Knowledge-based methods are those that
`rely on previous experimental studies in which stabilizing mutations or scaffolds have been
`identified (31, 36, 37, 41). Importantly, this general approach does not depend on natural
`antibody sequences and also includes mutations that are rare or absent in conventional
`antibody repertoires. Statistical methods use consensus approaches to identify stabilizing
`mutations based on the assumption that the most common antibody sequences are optimal
`(31, 39, 42). This approach can be used in powerful ways to evaluate not only the consensus
`of individual positions but also pairwise and higher order conservation at noncontiguous
`sites, such as interfaces between antibody domains (43). Structure-based methods are
`computational approaches that use either existing or predicted antibody structures to identify
`stabilizing mutations (24, 27, 44).
`
`The combination of the three approaches is especially powerful, which was demonstrated in
`a study of an unstable single-chain variable fragment (scFv) (44). The investigators sought to
`stabilize this scFv (initial melting temperature of 51°C) to the point that it could be fused to
`an IgG to generate a bispecific antibody with both high activity and stability. They used (a)
`knowledge-based approaches (31); (b) statistical methods, such as covariation and frequency
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 7
`
`

`

`Tiller and Tessier
`
`Page 8
`
`analysis (37, 43); and (c) structure-based methods, such as Rosetta (45) and molecular
`simulations (46), to predict positions within the scFv that are most important to stability.
`This led to the identification of 18 stabilizing mutations at 10 different positions (44). These
`included single mutations that increased the melting temperature significantly (67°C for a
`variant with P101D in VH) as well as combinations of mutations that were even more
`stabilizing (melting temperature of 82°C for a variant with S16E, V55G, and P101D in VH,
`and S46L in VL).
`
`Four of these stabilizing mutations are featured in Figure 4. Two are located at the VH–VL
`interface (VH P101D and VL S46L), which has been shown to be critical in determining the
`overall stability of Fv and scFv antibody fragments (37, 40, 43, 47, 48). The negatively
`charged mutation (VH P101D) forms a salt bridge with a neighboring residue (VH R98), and
`the hydrophobic mutation (VL S46L) increases van der Waals contacts and induces favorable
`structural changes in neighboring residues. The other two stabilizing mutations in the VH
`domain are located far from the VH–VL interface. The VH S16E mutation appears to be
`stabilizing due to electrostatic interactions (it is located in a positively charged region) as
`well as favorable van der Waals interactions involving the aliphatic side chain (which is
`longer for glutamic acid than serine). The V55G mutation is located in a turn near HCDR2
`and appears to stabilize the scFv by eliminating strain caused by the unfavorable φ and ψ
`angles for valine at this position.
`
`Many other impactful studies have also demonstrated rational approaches for stabilizing
`antibodies. One approach is to introduce additional intramolecular disulfide bonds within
`single-domain antibodies as well as interdomain disulfide bonds within Fvs and scFvs to
`increase folding stability (49–52). These methods have yielded significant improvements in
`stability, although in some cases they have resulted in reduced expression. In addition, much
`effort has focused on optimizing the VH–VL interfaces using noncysteine mutations (40, 43,
`48). This is important to improve both thermodynamic and kinetic folding stability
`(especially for Fvs and scFvs) (30, 47), and to reduce the complexity of the resulting
`antibodies by avoiding additional disulfide bonds. These and related studies (38, 41, 53, 54)
`are improving the systematic and robust optimization of antibody conformational stability.
`
`5. ANTIBODY COLLOIDAL STABILITY (SOLUBILITY)
`
`The colloidal stability of antibodies—which is governed by solvent-exposed residues in their
`native folded structure—is not as well understood as conformational stability. Nevertheless,
`colloidal stability is also a critical attribute of antibodies, especially for those antibodies with
`high conformational stability (as observed for many IgGs) (4, 5, 55). There are three key
`elements of antibodies that impact their solubility, namely the (a) CDRs, (b) frameworks of
`the variable and constant domains, and (c) glycans. CDRs commonly contain hydrophobic
`and charged residues to mediate high-affinity binding, yet these same residues can also
`mediate antibody self-association and aggregation (35, 56–60). Therefore, mutations in the
`CDRs can significantly impact antibody solubility (24, 35, 56–59). The frameworks of
`antibodies are also important determinants of their solubility (24, 56, 58, 61, 62). These
`regions typically contain hydrophobic patches (e.g., Fc receptor-binding sites) and
`oppositely charged domains that can interact with themselves or with the CDRs, leading to
`
`Annu Rev Biomed Eng. Author manuscript; available in PMC 2017 February 02.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Lassen - Exhibit 1053, p. 8
`
`

`

`Tiller and Tessier
`
`Page 9
`
`poor solubility. Glycans also significantly impact solubility, typically in a positive manner
`(24, 63, 64).
`
`An excellent study of the impact of each of these factors on antibody solubility is
`summarized in Figure 5 (61). The investigators sought to improve the poor solubility of an
`antibody specific for the glycoprotein LINGO-1. Because this antibody has excellent
`(subnanomolar) binding affinity and high bioactivity, multiple strategies were pursued to
`increase its solubility without reducing binding activity. The first approach was to switch the
`antibody framework from an IgG1 framework to IgG2 and IgG4 frameworks. Surprisingly,
`this simple change resulted in dramatic increases in solubility for both frameworks (increase
`from <1 to >30 mg/mL). The origin of these improvements is not clear given that the
`isoelectric points of the three variants are all high (>pH 8.2), and the IgG2 and IgG4 variants
`display folding stabilities that are similar to or lower than that of the parent IgG1 antibody.
`
`Given the preference of specific IgG subtypes for different therapeutic applications, the
`investigators also evaluated the impact of CDR mutations and glycans on the solubility of
`the parent IgG1 (anti-LINGO-1) antibody (Figure 5). Hydrophobic residues in HCDR2
`(isoleucine 57), HCDR3 (tryptophan 104), and LCDR3 (tryptophan 94) were mutated to be
`less hydrophobic or polar residues. Multiple mutations at these sites improved solubility by
`more than an order of magnitude without reducing binding affinity. Glycosylation was also
`found to significantly impact the solubility of the wild-type antibody. For example,
`removing the glycans from the IgG2 and IgG4 variants reduced the solubility to levels
`similar to those of the parent IgG1 antibody (either with or without glycans), suggesting that
`the glycans of the non-IgG1 variants are critical to their superior solubility. The authors also
`introduced glycosylation sites within the CH1 domain of the IgG1 antibody at four different
`positions. Interestingly, the solubilizing activity of the glycosylation variants was high for
`two mutants (>50 mg/mL) and modest for the others (3–5 mg/mL), and the differences were
`not predictable based on the proximity of the glycosylation sites to the variable domains.
`
`Several additional studies have used related approaches to improve antibody solubility. For
`example, multiple studies have reported how the sequences for CDRs, both of antibody
`fragments and full-length antibodies, can be engineered to increase solubility without
`compromising binding activity (35, 46, 56–59). These studies have also revealed that the
`location of aggregation hot spots within the CDRs is variable, and identifying effective sites
`for mutation requires the use of systematic design approaches (46, 57, 59, 65–67).
`Significant progress has also been made in engineering the fram

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