Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Nov 1.
Published in final edited form as: J Pharm Sci. 2014 Sep 10;103(11):3356–3363. doi: 10.1002/jps.24130

Improving monoclonal antibody selection and engineering using measurements of colloidal protein interactions

Steven B Geng 1, Jason K Cheung 2, Chakravarthy Narasimhan 2, Mohammed Shameem 2, Peter M Tessier 1,*
PMCID: PMC4206634  NIHMSID: NIHMS620347  PMID: 25209466

Abstract

A limitation of using monoclonal antibodies as therapeutic molecules is their propensity to associate with themselves and/or with other molecules via non-affinity (colloidal) interactions. This can lead to a variety of problems ranging from low solubility and high viscosity to off-target binding and fast antibody clearance. Measuring such colloidal interactions is challenging given that they are weak and potentially involve diverse target molecules. Nevertheless, assessing these weak interactions – especially during early antibody discovery and lead candidate optimization – is critical to preventing problems that can arise later in the development process. Here we review advances in developing and implementing sensitive methods for measuring antibody colloidal interactions as well as using these measurements for guiding antibody selection and engineering. These systematic efforts to minimize non-affinity interactions are expected to yield more effective and stable monoclonal antibodies for diverse therapeutic applications.

Introduction

There are a daunting number of factors that influence the effectiveness and success of therapeutic monoclonal antibodies (mAbs). The most important issues relate to the specific biological pathways being targeted. For example, the optimal binding affinity and epitope on a target antigen (mediated by the antibody variable domains) as well as the optimal type and level of effector function (mediated by the antibody constant domains) are dependent on the specific therapeutic target. The pharmacokinetics and biodistribution of therapeutic mAbs, which are influenced by the target antigen and recycling Fc receptors, also significantly impact their effectiveness. Thus, specific (affinity) interactions involving the variable and constant domains of mAbs are key determinants of their therapeutic activity.

Nevertheless, the same variable and constant regions of mAbs that mediate affinity interactions can also participate in non-affinity interactions with either themselves (self-interactions) or with other molecules (polyspecific interactions). The potential negative ramifications of these colloidal interactions are significant and are also important determinants of the success of therapeutic mAbs. Attractive self-interactions between antibodies (either in their native or non-native conformations) can lead to aggregation, abnormally high viscosity, liquid-liquid phase separation and opalescence.16 Aggregation is particularly concerning because of the suspected immunogenicity of antibody aggregates,710 while high viscosity is problematic for subcutaneous delivery applications.11,12 Polyspecific antibody interactions are also concerning because they can lead to off-target effects as well as fast antibody clearance.13,14

Therefore, it is critical to evaluate non-affinity antibody interactions early in therapeutic discovery and lead candidate optimization to minimize problems that can occur later in development. However, these interactions are difficult to measure because they are relatively weak and can involve a large number of potential molecules. This is particularly challenging during early antibody discovery because of the large number of candidate mAbs (tens to thousands), as well as their low concentrations (<100 μg/mL) and purities (unpurified cell supernatants). Here we review important recent progress in characterizing antibody colloidal interactions using biophysical methods during early antibody discovery and discuss how these measurements are being used to improve antibody selection and engineering.

Antibody self-interactions

Antibody self-association is the most fundamental and widely studied type of non-affinity antibody interaction. It is logical that mAbs can self-associate based on their multidomain architecture, symmetry, and non-uniform distribution of solvent-exposed hydrophobic and charged residues (Fig. 1). The variable heavy (VH) and light (VL) domains each display three solvent-exposed peptide loops (complementarity determining regions or CDRs) that commonly contain hydrophobic and charged residues to mediate high-affinity binding. Several studies have confirmed that hydrophobic and electrostatic interactions involving CDRs can mediate antibody self-association and aggregation.1523 More generally, attractive electrostatic interactions involving the Fab2325 and Fc26 regions of some antibodies have been shown to mediate self-association. The Fc regions of antibodies contain solvent-exposed hydrophobic residues involved in binding to Fc receptors and can also influence antibody self-association.22,2631 The bivalency of mAbs naturally amplifies these self-interactions.16,32

Figure 1.

Figure 1

Monoclonal antibodies are complex multidomain proteins that can associate with themselves or other molecules via non-affinity (colloidal) interactions. The Fab crystal structure of a poorly behaved antibody (PDB 3G6A) is highlighted with the heavy chain complementarity-determining regions (CDRs) in green and the light chain CDRs in blue. For this antibody, three consecutive aromatic residues (red) in heavy chain CDR3 mediate attractive colloidal interactions and poor antibody solubility.18

There are several valuable methods for measuring antibody self-association, including static3336 and dynamic light scattering,34,3741 neutron42,43 and X-ray4447 scattering, analytical ultracentrifugation,6,38,48,49 membrane osmometry,3,50 self-interaction chromatography,5157 self-interaction nanoparticle spectroscopy,5863 and biolayer interferometry.64 The most powerful and insightful methods such as neutron scattering, membrane osmometry, and analytical ultracentrifugation are generally the lowest throughput and most difficult to use during early antibody candidate selection. These methods are outside the scope of this review and have been reviewed previously.48,65,66

Methods such as dynamic light scattering, self-interaction nanoparticle spectroscopy and biolayer interferometry provide lower resolution information but afford increased throughput and flexibility for characterizing mAb candidates during early antibody discovery. The most widely used method is dynamic light scattering, which can be performed in a microplate format enabling tens to hundreds of samples to be evaluated.37,3941 As a measure of antibody self-association, mutual diffusion coefficients (D) are evaluated as a function of antibody concentration (c). This enables evaluation of the diffusion interaction parameter (kD):

D=D0(1+kDc)

where D0 is the self-diffusion coefficient. kD is not a direct measure of antibody self-association because it has both thermodynamic and hydrodynamic contributions. Therefore, it must be interpreted carefully, as discussed elsewhere.38 Nevertheless, multiple studies have confirmed that kD is generally well correlated with the second virial coefficient,37,39,40 a thermodynamic measure of pairwise protein interactions. Measurements of kD are typically conducted at antibody concentrations of 1–20 mg/mL, which typically requires approximately one mg of purified mAb per kD measurement. Although it is undesirable from the prospective of early antibody discovery that dynamic light scattering requires purified and concentrated mAbs, advances in high-throughput expression and purification methods are reducing this bottleneck.67

Several studies have demonstrated the power of dynamic light scattering at improving mAb candidate selection and characterization.23,3740,6872 One of the most comprehensive studies to date measured kD values for 29 different mAbs at several different solution conditions and compared these measurements conducted at 1–20 mg/mL to viscosity measurements conducted at much higher concentrations (up to 175 mg/mL).37 Impressively, the kD measurements are strongly correlated with viscosity measurements (Fig. 2), and demonstrate the utility of dynamic light scattering for efficiently assessing properties of concentrated solutions at modest antibody concentrations. Other studies have also demonstrated that kD measurements are typically correlated with antibody solubility and aggregation rates.38,40,59,73

Figure 2.

Figure 2

Measurements of diffusion coefficients as a function of antibody concentration are correlated with viscosities of concentrated antibody solutions. Mutual diffusion coefficients are measured using dynamic light scattering for moderately concentrated antibody solutions (typically 1–20 mg/mL). Measurements of diffusion coefficients as a function of antibody concentration are used to calculate the diffusion interaction parameter (kD). The exponential viscosity coefficients are obtained from viscosity measurements as a function of antibody concentration (50–175 mg/mL). The kD and viscosity measurements are reproduced from a previous report.37

To improve throughput of antibody self-interaction measurements during early antibody discovery, it is important to reduce the need for purifying and/or concentrating mAb candidates. This is challenging because typical mAb concentrations in unpurified cell culture supernatants are <0.1 mg/mL using small scale, transient transfections. Components in the media can also interfere with measurements of antibody self-association. One logical approach for characterizing unpurified mAb samples is to use a method that involves an affinity capture step in which antibodies are selectively adsorbed on a solid support and assayed. Although surface plasmon resonance methods such as Biacore have been used for this application,26,74 these methods are limited in their ability to assay weak interactions due to the low surface area for interaction.

A related approach – namely bio-layer interferometry75 – has been reported recently for characterizing antibody self-association that holds potential for assaying unpurified mAb samples.64 This method measures the interference pattern of light reflected from a surface with immobilized protein relative to a control surface to evaluate the wavelength shift, which is proportional to the thickness of the protein coating. By first coating the biosensor surface with an affinity capture antibody (e.g., goat anti-human antibody), dilute mAbs (~0.1 mg/mL) are adsorbed and the propensity of non-adsorbed mAbs to self-associate with immobilized mAbs is assayed. This approach has been shown to differentiate between highly and poorly soluble mAbs,64 and it shows good correlation with lower throughput methods for measuring self-association such as self-interaction chromatography (which has been used for optimizing antibody solubility).5456 Bio-layer interferometry experiments have been conducted in the presence of a non-specific protein (BSA), suggesting that it may be compatible with unpurified mAb samples that are common during early antibody discovery. This approach has good throughput because it can be conducted in 384-well plates. Minor limitations include low sensitivity at highly dilute mAb concentrations (~0.01 mg/mL) and the need for specialized equipment. Nevertheless, this powerful method holds significant potential for use during early antibody candidate selection.

Another promising approach for characterizing unpurified mAbs during early antibody discovery is affinity-capture self-interaction nanoparticle spectroscopy (AC-SINS).59,63 This method involves coating gold nanoparticles with different ratios of capture (goat anti-human) to non-capture (goat non-specific) antibodies, and then using these conjugates to selectively immobilize mAbs from dilute antibody solutions (0.001–0.05 mg/mL; Fig. 3). The colloidal stability of the mAb-gold conjugates is governed by the self-association propensities of the immobilized mAbs. Associative mAbs lead to reduced separation distances between mAb-gold conjugates and red-shifting of the wavelength of maximum absorbance (plasmon wavelength, λp). Evaluation of the plasmon wavelength as a function of the capture to non-capture antibody ratio enables sensitive discrimination between different mAbs.

Figure 3.

Figure 3

Identification of mutations that reduce antibody self-association and increase solubility using affinity-capture self-interaction nanoparticle spectroscopy (AC-SINS). (A) Gold nanoparticles are coated with mixtures of non-capture (gray) and capture (red) polyclonal antibodies, the latter of which is specific for human mAbs (blue). After capturing human mAbs, attractive mAb self-interactions are detected due to reduced interparticle separation distances and increased plasmon wavelengths. (B) Plasmon wavelength measurements as a function of the ratio of non-capture to capture antibody reveal that two mAbs with single mutations in their complementarity determining regions have reduced self-association. The inset images show that the mAbs that display reduced self-association at dilute concentrations (1–50 μg/mL) also do not phase separate when concentrated orders of magnitude (50 mg/mL). The red boxes highlight the phase separation for the three mAbs with elevated self-association. The plasmon wavelengths and images of antibody solutions are reproduced from a previous report.59

AC-SINS has been used to discriminate between high and low solubility mAbs that differ only by single mutations in their CDRs (Fig. 3).59 Interestingly, the AC-SINS measurements are weakly impacted by cell culture media, suggesting AC-SINS may be compatible with unpurified cell culture supernatants. Moreover, the AC-SINS findings obtained at dilute mAb concentrations (0.05 mg/mL) are strongly correlated with dynamic light scattering and solubility experiments conducted at orders of magnitude higher concentrations (1–50 mg/mL). In addition, AC-SINS has good throughput because it is typically conducted in 384-well plates and requires only a standard absorbance plate reader. Weaknesses of AC-SINS include the possible impact of antibody immobilization on mAb self-association, and the potential interference of additives and contaminants that interact strongly with gold. Nevertheless, this approach appears promising for guiding initial mAb candidate selection.

Antibody cross-interactions (polyspecificity)

The potential for antibodies to interact non-specifically with multiple targets has long been recognized.76 For example, antibodies generated from immature B-cells are typically polyspecific, while those from mature B-cells are generally much more specific.77 Quantifying polyspecificity is challenging given the large number of possible target molecules and the wide range of possible affinities.77,78 Polyspecificity is especially concerning for antibody therapeutics because of potential off-target effects and the negative impact that such cross-interactions may have on antibody pharmacokinetics.13,14

Therefore, measurements of antibody polyspecificity are increasingly being incorporated into the analysis of therapeutic mAb candidates.13,14,79 A particularly exciting study reported the characterization of a large panel of mAbs that displayed a wide range of antibody clearance rates in monkeys and humans.14 Importantly, the differences in clearance rates were unrelated to their target antigens or interactions with recycling Fc receptors. Therefore, the investigators reasoned that non-specific interactions between mAbs and off-target molecules may lead to fast clearance. To test this hypothesis, they evaluated the binding of mAbs to baculovirus particles that present a complex mixture of phospholipids, carbohydrates, glycoproteins, nucleic acids and other types of molecules. Interestingly, the relative binding of mAbs to immobilized baculovirus particles in an ELISA assay was positively correlated with the rate of mAb clearance both in monkeys and humans (Fig. 4). Similar results for a subset of the mAbs were obtained using mammalian cells, but the use of cells as a non-specific binding reagent is problematic because some of the antigens are endogenous cell-surface proteins. Mutations that reduced mAb binding to baculovirus particles also reduced the rate of antibody clearance. Other physical properties of antibodies – such as net charge, isoelectric point and hydrophobicity – showed little correlation with the rate of antibody clearance. Strengths of this approach include good throughput and the potential to evaluate unpurified mAbs early in antibody discovery due to the ELISA format. Future work will need to evaluate if the correlation with antibody clearance rates can be further improved using other types of non-specific reagents alone or in combination with baculovirus particles. Nevertheless, these findings highlight the value of evaluating non-specific interactions for guiding the selection of mAb candidates with desirable (long) circulation times.

Figure 4.

Figure 4

Non-affinity interactions between monoclonal antibodies and baculovirus particles are correlated with the rate of antibody clearance in humans. Baculovirus particles are immobilized in microtiter plates, and the relative binding of antibody variants is evaluated. Increased antibody binding is correlated with increased rate of antibody clearance. The data is reproduced from a previous report.14

Another approach for incorporating measurements of polyspecificity into the early discovery process was reported using an alternative polyspecificity reagent.13 Instead of using baculovirus particles, the investigators generated a soluble membrane protein reagent from homogenized mammalian cells prepared with a mild surfactant. The binding of biotinylated membrane proteins to antibody libraries displayed on the surface of yeast was evaluated with fluorescence-activated cell sorting (FACS) to select antibody variants with low non-specific binding relative to high specific binding to the target antigen. Antibody binding to the membrane protein reagent was correlated with binding to baculovirus particles, which suggests the FACS measurements may also be correlated with antibody clearance rates. The correlation between different assays also suggests that similar types of non-specific interactions are detected using different polyspecificity reagents. The strengths of this approach include the excellent throughput of FACS and the ability to identify antibodies with low polyspecificity at an unusually early stage in the antibody discovery process. Future work will need to evaluate if other polyspecificity reagents (including baculovirus particles) further improve the detection of non-specific antibody interactions either alone or in combination with soluble membrane proteins.

A simpler, complementary method for evaluating antibody polyspecificity is to evaluate the interaction of mAbs with polyclonal antibodies.79 This approach has been elegantly demonstrated in the form of cross-interaction chromatography in which human polyclonal antibodies are immobilized on chromatography particles and the retention times of human mAbs are evaluated from a polyclonal antibody column (Fig. 5).13,16,63,64,79 Longer retention times correspond to increased non-specific interactions between mAbs and polyclonal antibodies. There is some correlation between the propensity of human mAbs to interact with polyclonal antibodies and the propensity of mAbs to interact with baculovirus particles and soluble membrane proteins,13 suggesting that the three assays measure similar types of non-affinity interactions. Nevertheless, the baculovirus particle and membrane protein assays are more strongly correlated than are the cross-interaction chromatography and soluble membrane protein assays, which is consistent with the less complex nature of polyclonal antibodies relative to the other reagents. Strengths of cross-interaction chromatography include its simplicity, moderate throughput using a liquid chromatography system with an autosampler, relatively low mAb consumption (<10 μg per measurement), and the potential to evaluate unpurified mAbs if such samples can be modestly concentrated to ~100 μg/mL. Future work will need to further improve the throughput of the assay, reduce the required mAb concentrations for use in early antibody discovery, and evaluate the use of other immobilized molecules in addition to human polyclonal antibodies to improve detection of non-specific antibody interactions.

Figure 5.

Figure 5

Detection of monoclonal antibodies that interact non-specifically with polyclonal antibodies using cross-interaction chromatography. Polyclonal antibody is immobilized on chromatography particles, and the relative retention of different monoclonal antibodies is evaluated. Increased retention signifies attractive antibody interactions. The chromatograms are reproduced from a previous report.79

An interesting question is whether antibodies that are prone to self-associate (and which are poorly soluble and/or viscous) are also prone to interact promiscuously with other molecules. Indeed, cross-interaction chromatography was originally developed for identifying mAbs with low solubility.79 A closely related method (self-interaction chromatography52,56,57) could be used in which self-interactions between immobilized and non-immobilized mAbs are measured. However, the throughput of this method is low for screening different mAbs because a separate column must be prepared for each antibody variant. Nevertheless, cross-interaction chromatography measurements are correlated with mAb solubility (which is governed by antibody self-association).16,20,79,80 This suggests that interactions leading to non-specific antibody cross-association are similar to those that mediate antibody self-association. It is likely that the basis for the success of using cross-interaction chromatography measurements as a surrogate for self-interaction measurements is the similarity in the size, structure and composition of human mAbs and human polyclonal antibodies. Interestingly, a poorly soluble human mAb that interacts strongly with human polyclonal antibodies is less interactive with murine polyclonal antibodies,80 suggesting that the extreme similarity between human mAbs and human polyclonal antibodies is important for detection of non-affinity antibody interactions (at least for those that govern antibody solubility). Nevertheless, exceptions between mAb self-interactions and cross-interactions with human polyclonal antibodies are expected and have been reported.63 These findings caution against over-interpreting cross-interaction measurements without evaluating antibody self-association, and highlight the need for additional research to understand similarities and differences between antibody self- and cross-interactions.

A related intriguing question is whether selection and/or evolution of high-affinity antibodies generally increase the risk for non-specific antibody interactions. This seems logical for antibodies that use hydrophobic residues in their CDRs to mediate high-affinity binding. Indeed, the hydrophobic CDRs of multiple high-affinity antibodies obtained from phage libraries have been shown to also mediate non-specific antibody interactions.16,18,20,32 Elegant work has shown how cross-interaction chromatography can be used to detect these colloidal interactions and how directed evolution methods can be used to reduce these interactions without compromising binding affinity.20 Conventional wisdom suggests that high-affinity antibodies obtained from animals and/or eukaryotic cell display methods have a lower risk of non-specific antibody interactions due to the increased quality control mechanisms during evolution and selection. Indeed, a recent survey of >90 murine hybridoma antibodies revealed that they all display low non-specific interactions via cross-interaction chromatography.80 Nevertheless, several antibodies derived from animals and/or eukaryotic cell display methods have been reported to be associative and/or display undesirable solution properties at high antibody concentrations required for subcutaneous delivery.13,14,63,64 Additional work is needed to better define the relative risks of obtaining associative antibodies using different antibody discovery methods. Nevertheless, these findings suggest that diverse antibody discovery methods can be used to successfully identify antibodies with both high affinity and low propensity to interact non-specifically if effective assays of colloidal interactions can be incorporated into the early discovery process.

Conclusions and Future Directions

The intense interest in developing antibodies as therapeutics has led to many innovative approaches for detecting weak protein interactions that can negatively impact antibody developability and therapeutic activity. It will be important to refine these methods to be compatible with the extreme requirements of early antibody discovery during which large numbers of dilute and unpurified mAbs must be characterized. Important next steps include defining the accuracy and precision of these methods, as well as comparing several of them at the same time to better evaluate their similarities and differences. It will also be critical to better define how different types of colloidal interactions influence properties such as antibody solubility, viscosity and antibody clearance rate. Finally, it will be important to evaluate whether measurements of antibody colloidal interactions prior to purification and characterization of mAbs can improve the discovery and development of new antibody candidates with desirable biophysical and pharmacokinetic properties.

Acknowledgments

We thank members of the Tessier lab for helpful comments on the manuscript. This work was supported by NSF (CBET grants 0954450 and 1159943) and NIH (R01GM104130).

Footnotes

Conflicts of interest

PMT has received honorariums and/or consulting fees from MedImmune, Eli Lilly, Bristol-Myers Squibb, Janssen Biotech, Merck, Genentech, Amgen, Pfizer, Adimab, Abbott, Bayer, Roche and DuPont.

References

  • 1.Perchiacca JM, Tessier PM. Engineering aggregation-resistant antibodies. Annu Rev Chem Biomol Eng. 2012;3:263–286. doi: 10.1146/annurev-chembioeng-062011-081052. [DOI] [PubMed] [Google Scholar]
  • 2.Vazquez-Rey M, Lang DA. Aggregates in monoclonal antibody manufacturing processes. Biotechnol Bioeng. 2011;108(7):1494–1508. doi: 10.1002/bit.23155. [DOI] [PubMed] [Google Scholar]
  • 3.Salinas BA, Sathish HA, Bishop SM, Harn N, Carpenter JF, Randolph TW. Understanding and modulating opalescence and viscosity in a monoclonal antibody formulation. J Pharm Sci. 2010;99(1):82–93. doi: 10.1002/jps.21797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Nishi H, Miyajima M, Nakagami H, Noda M, Uchiyama S, Fukui K. Phase separation of an IgG1 antibody solution under a low ionic strength condition. Pharm Res. 2010;27(7):1348–1360. doi: 10.1007/s11095-010-0125-7. [DOI] [PubMed] [Google Scholar]
  • 5.Cromwell ME, Hilario E, Jacobson F. Protein aggregation and bioprocessing. AAPS J. 2006;8(3):E572–579. doi: 10.1208/aapsj080366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Liu J, Nguyen MD, Andya JD, Shire SJ. Reversible self-association increases the viscosity of a concentrated monoclonal antibody in aqueous solution. J Pharm Sci. 2005;94(9):1928–1940. doi: 10.1002/jps.20347. [DOI] [PubMed] [Google Scholar]
  • 7.Singh SK. Impact of product-related factors on immunogenicity of biotherapeutics. J Pharm Sci. 2011;100(2):354–387. doi: 10.1002/jps.22276. [DOI] [PubMed] [Google Scholar]
  • 8.De Groot AS, Scott DW. Immunogenicity of protein therapeutics. Trends Immunol. 2007;28(11):482–490. doi: 10.1016/j.it.2007.07.011. [DOI] [PubMed] [Google Scholar]
  • 9.Rosenberg AS. Effects of protein aggregates: an immunologic perspective. AAPS J. 2006;8(3):E501–507. doi: 10.1208/aapsj080359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Lowe D, Dudgeon K, Rouet R, Schofield P, Jermutus L, Christ D. Aggregation, stability, and formulation of human antibody therapeutics. Adv Protein Chem Struct Biol. 2011;84:41–61. doi: 10.1016/B978-0-12-386483-3.00004-5. [DOI] [PubMed] [Google Scholar]
  • 11.Shire SJ. Formulation and manufacturability of biologics. Curr Opin Biotechnol. 2009;20(6):708–714. doi: 10.1016/j.copbio.2009.10.006. [DOI] [PubMed] [Google Scholar]
  • 12.Daugherty AL, Mrsny RJ. Formulation and delivery issues for monoclonal antibody therapeutics. Adv Drug Deliv Rev. 2006;58(5–6):686–706. doi: 10.1016/j.addr.2006.03.011. [DOI] [PubMed] [Google Scholar]
  • 13.Xu Y, Roach W, Sun T, Jain T, Prinz B, Yu TY, Torrey J, Thomas J, Bobrowicz P, Vasquez M, Wittrup KD, Krauland E. Addressing polyspecificity of antibodies selected from an in vitro yeast presentation system: a FACS-based, high-throughput selection and analytical tool. Protein Eng Des Sel. 2013;26(10):663–670. doi: 10.1093/protein/gzt047. [DOI] [PubMed] [Google Scholar]
  • 14.Hotzel I, Theil FP, Bernstein LJ, Prabhu S, Deng R, Quintana L, Lutman J, Sibia R, Chan P, Bumbaca D, Fielder P, Carter PJ, Kelley RF. A strategy for risk mitigation of antibodies with fast clearance. mAbs. 2012;4(6):753–760. doi: 10.4161/mabs.22189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Perchiacca JM, Ladiwala AR, Bhattacharya M, Tessier PM. Aggregation-resistant domain antibodies engineered with charged mutations near the edges of the complementarity-determining regions. Protein Eng Des Sel. 2012;25(10):591–601. doi: 10.1093/protein/gzs042. [DOI] [PubMed] [Google Scholar]
  • 16.Bethea D, Wu SJ, Luo J, Hyun L, Lacy ER, Teplyakov A, Jacobs SA, O’Neil KT, Gilliland GL, Feng Y. Mechanisms of self-association of a human monoclonal antibody CNTO607. Protein Eng Des Sel. 2012;25(10):531–537. doi: 10.1093/protein/gzs047. [DOI] [PubMed] [Google Scholar]
  • 17.Perchiacca JM, Bhattacharya M, Tessier PM. Mutational analysis of domain antibodies reveals aggregation hotspots within and near the complementarity determining regions. Proteins. 2011;79(9):2637–2647. doi: 10.1002/prot.23085. [DOI] [PubMed] [Google Scholar]
  • 18.Wu SJ, Luo J, O’Neil KT, Kang J, Lacy ER, Canziani G, Baker A, Huang M, Tang QM, Raju TS, Jacobs SA, Teplyakov A, Gilliland GL, Feng Y. Structure-based engineering of a monoclonal antibody for improved solubility. Protein Eng Des Sel. 2010;23(8):643–651. doi: 10.1093/protein/gzq037. [DOI] [PubMed] [Google Scholar]
  • 19.Wang X, Das TK, Singh SK, Kumar S. Potential aggregation prone regions in biotherapeutics: A survey of commercial monoclonal antibodies. mAbs. 2009;1(3):254–267. doi: 10.4161/mabs.1.3.8035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kehoe JW, Whitaker B, Bethea D, Lacy ER, Boakye K, Santulli-Marotto S, Ryan MH, Feng Y, Wheeler JC. Isolation and optimization for affinity and biophysical characteristics of anti-CCL17 antibodies from the VH1-69 germline gene. Protein Eng Des Sel. 2014 doi: 10.1093/protein/gzu012. ePub Arp 16. [DOI] [PubMed] [Google Scholar]
  • 21.Dudgeon K, Rouet R, Kokmeijer I, Schofield P, Stolp J, Langley D, Stock D, Christ D. General strategy for the generation of human antibody variable domains with increased aggregation resistance. Proc Natl Acad Sci U S A. 2012;109(27):10879–10884. doi: 10.1073/pnas.1202866109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Chennamsetty N, Voynov V, Kayser V, Helk B, Trout BL. Design of therapeutic proteins with enhanced stability. Proc Natl Acad Sci U S A. 2009;106(29):11937–11942. doi: 10.1073/pnas.0904191106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Yadav S, Laue TM, Kalonia DS, Singh SN, Shire SJ. The influence of charge distribution on self-association and viscosity behavior of monoclonal antibody solutions. Mol Pharm. 2012;9(4):791–802. doi: 10.1021/mp200566k. [DOI] [PubMed] [Google Scholar]
  • 24.Kanai S, Liu J, Patapoff TW, Shire SJ. Reversible self-association of a concentrated monoclonal antibody solution mediated by Fab-Fab interaction that impacts solution viscosity. J Pharm Sci. 2008;97(10):4219–4227. doi: 10.1002/jps.21322. [DOI] [PubMed] [Google Scholar]
  • 25.Wang T, Kumru OS, Yi L, Wang YJ, Zhang J, Kim JH, Joshi SB, Middaugh CR, Volkin DB. Effect of ionic strength and pH on the physical and chemical stability of a monoclonal antibody antigen-binding fragment. J Pharm Sci. 2013;102(8):2520–2537. doi: 10.1002/jps.23645. [DOI] [PubMed] [Google Scholar]
  • 26.Nishi H, Miyajima M, Wakiyama N, Kubota K, Hasegawa J, Uchiyama S, Fukui K. Fc domain mediated self-association of an IgG1 monoclonal antibody under a low ionic strength condition. J Biosci Bioeng. 2011;112(4):326–332. doi: 10.1016/j.jbiosc.2011.06.017. [DOI] [PubMed] [Google Scholar]
  • 27.Chennamsetty N, Voynov V, Kayser V, Helk B, Trout BL. Prediction of aggregation prone regions of therapeutic proteins. J Phys Chem B. 2010;114(19):6614–6624. doi: 10.1021/jp911706q. [DOI] [PubMed] [Google Scholar]
  • 28.Kayser V, Chennamsetty N, Voynov V, Forrer K, Helk B, Trout BL. Glycosylation influences on the aggregation propensity of therapeutic monoclonal antibodies. Biotechnol J. 2011;6(1):38–44. doi: 10.1002/biot.201000091. [DOI] [PubMed] [Google Scholar]
  • 29.Brummitt RK, Nesta DP, Chang L, Chase SF, Laue TM, Roberts CJ. Nonnative aggregation of an IgG1 antibody in acidic conditions: part 1. Unfolding, colloidal interactions, and formation of high-molecular-weight aggregates. J Pharm Sci. 2011;100(6):2087–2103. doi: 10.1002/jps.22448. [DOI] [PubMed] [Google Scholar]
  • 30.Latypov RF, Hogan S, Lau H, Gadgil H, Liu D. Elucidation of acid-induced unfolding and aggregation of human immunoglobulin IgG1 and IgG2 Fc. J Biol Chem. 2012;287(2):1381–1396. doi: 10.1074/jbc.M111.297697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Fast JL, Cordes AA, Carpenter JF, Randolph TW. Physical instability of a therapeutic Fc fusion protein: domain contributions to conformational and colloidal stability. Biochemistry. 2009;48(49):11724–11736. doi: 10.1021/bi900853v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Pepinsky RB, Silvian L, Berkowitz SA, Farrington G, Lugovskoy A, Walus L, Eldredge J, Capili A, Mi S, Graff C, Garber E. Improving the solubility of anti-LINGO-1 monoclonal antibody Li33 by isotype switching and targeted mutagenesis. Protein Sci. 2010;19(5):954–966. doi: 10.1002/pro.372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Scherer TM, Liu J, Shire SJ, Minton AP. Intermolecular interactions of IgG1 monoclonal antibodies at high concentrations characterized by light scattering. J Phys Chem B. 2010;114(40):12948–12957. doi: 10.1021/jp1028646. [DOI] [PubMed] [Google Scholar]
  • 34.Arzensek D, Kuzman D, Podgornik R. Colloidal interactions between monoclonal antibodies in aqueous solutions. J Colloid Interface Sci. 2012;384(1):207–216. doi: 10.1016/j.jcis.2012.06.055. [DOI] [PubMed] [Google Scholar]
  • 35.Some D. Light-scattering-based analysis of biomolecular interactions. Biophysical reviews. 2013;5(2):147–158. doi: 10.1007/s12551-013-0107-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Bajaj H, Sharma VK, Badkar A, Zeng D, Nema S, Kalonia DS. Protein structural conformation and not second virial coefficient relates to long-term irreversible aggregation of a monoclonal antibody and ovalbumin in solution. Pharm Res. 2006;23(6):1382–1394. doi: 10.1007/s11095-006-0018-y. [DOI] [PubMed] [Google Scholar]
  • 37.Connolly BD, Petry C, Yadav S, Demeule B, Ciaccio N, Moore JM, Shire SJ, Gokarn YR. Weak interactions govern the viscosity of concentrated antibody solutions: high-throughput analysis using the diffusion interaction parameter. Biophys J. 2012;103(1):69–78. doi: 10.1016/j.bpj.2012.04.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Saluja A, Fesinmeyer RM, Hogan S, Brems DN, Gokarn YR. Diffusion and sedimentation interaction parameters for measuring the second virial coefficient and their utility as predictors of protein aggregation. Biophys J. 2010;99(8):2657–2665. doi: 10.1016/j.bpj.2010.08.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Lehermayr C, Mahler HC, Mader K, Fischer S. Assessment of net charge and protein-protein interactions of different monoclonal antibodies. J Pharm Sci. 2011;100(7):2551–2562. doi: 10.1002/jps.22506. [DOI] [PubMed] [Google Scholar]
  • 40.Saito S, Hasegawa J, Kobayashi N, Kishi N, Uchiyama S, Fukui K. Behavior of monoclonal antibodies: relation between the second virial coefficient (B(2)) at low concentrations and aggregation propensity and viscosity at high concentrations. Pharm Res. 2012;29(2):397–410. doi: 10.1007/s11095-011-0563-x. [DOI] [PubMed] [Google Scholar]
  • 41.Esfandiary R, Hayes DB, Parupudi A, Casas-Finet J, Bai S, Samra HS, Shah AU, Sathish HA. A systematic multitechnique approach for detection and characterization of reversible self-association during formulation development of therapeutic antibodies. J Pharm Sci. 2013;102(9):3089–3099. doi: 10.1002/jps.23654. [DOI] [PubMed] [Google Scholar]
  • 42.Yearley EJ, Zarraga IE, Shire SJ, Scherer TM, Gokarn Y, Wagner NJ, Liu Y. Small-angle neutron scattering characterization of monoclonal antibody conformations and interactions at high concentrations. Biophys J. 2013;105(3):720–731. doi: 10.1016/j.bpj.2013.06.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Liu Y, Fratini E, Baglioni P, Chen WR, Chen SH. Effective long-range attraction between protein molecules in solutions studied by small angle neutron scattering. Phys Rev Lett. 2005;95(11):118102. doi: 10.1103/PhysRevLett.95.118102. [DOI] [PubMed] [Google Scholar]
  • 44.Iacob RE, Bou-Assaf GM, Makowski L, Engen JR, Berkowitz SA, Houde D. Investigating monoclonal antibody aggregation using a combination of H/DX-MS and other biophysical measurements. J Pharm Sci. 2013;102(12):4315–4329. doi: 10.1002/jps.23754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lilyestrom WG, Yadav S, Shire SJ, Scherer TM. Monoclonal antibody self-association, cluster formation, and rheology at high concentrations. J Phys Chem B. 2013;117(21):6373–6384. doi: 10.1021/jp4008152. [DOI] [PubMed] [Google Scholar]
  • 46.Lilyestrom WG, Shire SJ, Scherer TM. Influence of the cosolute environment on IgG solution structure analyzed by small-angle X-ray scattering. J Phys Chem B. 2012;116(32):9611–9618. doi: 10.1021/jp303839t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Mosbaek CR, Konarev PV, Svergun DI, Rischel C, Vestergaard B. High concentration formulation studies of an IgG2 antibody using small angle X-ray scattering. Pharm Res. 2012;29(8):2225–2235. doi: 10.1007/s11095-012-0751-3. [DOI] [PubMed] [Google Scholar]
  • 48.Howlett GJ, Minton AP, Rivas G. Analytical ultracentrifugation for the study of protein association and assembly. Curr Opin Chem Biol. 2006;10(5):430–436. doi: 10.1016/j.cbpa.2006.08.017. [DOI] [PubMed] [Google Scholar]
  • 49.Jimenez M, Rivas G, Minton AP. Quantitative characterization of weak self-association in concentrated solutions of immunoglobulin G via the measurement of sedimentation equilibrium and osmotic pressure. Biochemistry. 2007;46(28):8373–8378. doi: 10.1021/bi7005515. [DOI] [PubMed] [Google Scholar]
  • 50.Binabaji E, Rao S, Zydney AL. The osmotic pressure of highly concentrated monoclonal antibody solutions: Effect of solution conditions. Biotechnol Bioeng. 2014;111(3):529–536. doi: 10.1002/bit.25104. [DOI] [PubMed] [Google Scholar]
  • 51.Tessier PM, Vandrey SD, Berger BW, Pazhianur R, Sandler SI, Lenhoff AM. Self-interaction chromatography: a novel screening method for rational protein crystallization. Acta Crystallogr D Biol Crystallogr. 2002;58(10):1531–1535. doi: 10.1107/s0907444902012775. [DOI] [PubMed] [Google Scholar]
  • 52.Tessier PM, Lenhoff AM, Sandler SI. Rapid measurement of protein osmotic second virial coefficients by self-interaction chromatography. Biophys J. 2002;82(3):1620–1631. doi: 10.1016/S0006-3495(02)75513-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Binabaji E, Rao S, Zydney AL. Improved method for evaluating the dead volume and protein-protein interactions by self-interaction chromatography. Anal Chem. 2013;85(19):9101–9106. doi: 10.1021/ac4017242. [DOI] [PubMed] [Google Scholar]
  • 54.Sule SV, Cheung JK, Antochshuk V, Bhalla AS, Narasimhan C, Blaisdell S, Shameem M, Tessier PM. Solution pH that minimizes self-association of three monoclonal antibodies is strongly dependent on ionic strength. Mol Pharm. 2012;9(4):744–751. doi: 10.1021/mp200448j. [DOI] [PubMed] [Google Scholar]
  • 55.Le Brun V, Friess W, Bassarab S, Muhlau S, Garidel P. A critical evaluation of self-interaction chromatography as a predictive tool for the assessment of protein-protein interactions in protein formulation development: a case study of a therapeutic monoclonal antibody. Eur J Pharm Biopharm. 2010;75(1):16–25. doi: 10.1016/j.ejpb.2010.01.009. [DOI] [PubMed] [Google Scholar]
  • 56.Ahamed T, Esteban BN, Ottens M, van Dedem GW, van der Wielen LA, Bisschops MA, Lee A, Pham C, Thommes J. Phase behavior of an intact monoclonal antibody. Biophys J. 2007;93(2):610–619. doi: 10.1529/biophysj.106.098293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Patro SY, Przybycien TM. Self-interaction chromatography: a tool for the study of protein-protein interactions in bioprocessing environments. Biotechnol Bioeng. 1996;52(2):193–203. doi: 10.1002/(SICI)1097-0290(19961020)52:2<193::AID-BIT2>3.0.CO;2-L. [DOI] [PubMed] [Google Scholar]
  • 58.Jayaraman J, wu J, Brunelle MC, Cruz AMM, Goldberg DS, Lobo BL, Shah A, Tessier PM. Plasmonic measurements of monoclonal antibody self-association using self-interaction nanoparticle spectroscopy. Biotechnol Bioeng. 2014 doi: 10.1002/bit.25221. ePub Feb 20. [DOI] [PubMed] [Google Scholar]
  • 59.Sule SV, Dickinson CD, Lu J, Chow CK, Tessier PM. Rapid analysis of antibody self-association in complex mixtures using immunogold conjugates. Mol Pharm. 2013;10(4):1322–1331. doi: 10.1021/mp300524x. [DOI] [PubMed] [Google Scholar]
  • 60.Sule SV, Sukumar M, Weiss WF, Marcelino-Cruz AM, Sample T, Tessier PM. High-throughput analysis of concentration-dependent antibody self-association. Biophys J. 2011;101(7):1749–1757. doi: 10.1016/j.bpj.2011.08.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Bengali AN, Tessier PM. Biospecific protein immobilization for rapid analysis of weak protein interactions using self-interaction nanoparticle spectroscopy. Biotechnol Bioeng. 2009;104(2):240–250. doi: 10.1002/bit.22392. [DOI] [PubMed] [Google Scholar]
  • 62.Tessier PM, Jinkoji J, Cheng YC, Prentice JL, Lenhoff AM. Self-interaction nanoparticle spectroscopy: a nanoparticle-based protein interaction assay. J Am Chem Soc. 2008;130(10):3106–3112. doi: 10.1021/ja077624q. [DOI] [PubMed] [Google Scholar]
  • 63.Liu Y, Caffry I, Wu J, Geng SB, Jain T, Sun T, Reid F, Cao Y, Estep P, Yu Y, Vasquez M, Tessier PM, Xu Y. High-throughput screening for developability during early-stage antibody discovery using self-interaction nanoparticle spectroscopy. mAbs. 2014;6(2):483–492. doi: 10.4161/mabs.27431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Sun T, Reid F, Liu Y, Cao Y, Estep P, Nauman C, Xu Y. High throughput detection of antibody self-interaction by bio-layer interferometry. mAbs. 2013;5(6) doi: 10.4161/mabs.26186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Gabel F, Bicout D, Lehnert U, Tehei M, Weik M, Zaccai G. Protein dynamics studied by neutron scattering. Q Rev Biophys. 2002;35(4):327–367. doi: 10.1017/s0033583502003840. [DOI] [PubMed] [Google Scholar]
  • 66.Saluja A, Kalonia DS. Nature and consequences of protein-protein interactions in high protein concentration solutions. Int J Pharm. 2008;358(1–2):1–15. doi: 10.1016/j.ijpharm.2008.03.041. [DOI] [PubMed] [Google Scholar]
  • 67.Bhambure R, Kumar K, Rathore AS. High-throughput process development for biopharmaceutical drug substances. Trends Biotechnol. 2011;29(3):127–135. doi: 10.1016/j.tibtech.2010.12.001. [DOI] [PubMed] [Google Scholar]
  • 68.Saluja A, Badkar AV, Zeng DL, Kalonia DS. Ultrasonic rheology of a monoclonal antibody (IgG2) solution: implications for physical stability of proteins in high concentration formulations. J Pharm Sci. 2007;96(12):3181–3195. doi: 10.1002/jps.20970. [DOI] [PubMed] [Google Scholar]
  • 69.Yadav S, Shire SJ, Kalonia DS. Factors affecting the viscosity in high concentration solutions of different monoclonal antibodies. J Pharm Sci. 2010;99(12):4812–4829. doi: 10.1002/jps.22190. [DOI] [PubMed] [Google Scholar]
  • 70.Yadav S, Shire SJ, Kalonia DS. Viscosity behavior of high-concentration monoclonal antibody solutions: correlation with interaction parameter and electroviscous effects. J Pharm Sci. 2012;101(3):998–1011. doi: 10.1002/jps.22831. [DOI] [PubMed] [Google Scholar]
  • 71.Saluja A, Badkar AV, Zeng DL, Nema S, Kalonia DS. Ultrasonic storage modulus as a novel parameter for analyzing protein-protein interactions in high protein concentration solutions: correlation with static and dynamic light scattering measurements. Biophys J. 2007;92(1):234–244. doi: 10.1529/biophysj.106.095174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Yadav S, Liu J, Shire SJ, Kalonia DS. Specific interactions in high concentration antibody solutions resulting in high viscosity. J Pharm Sci. 2010;99(3):1152–1168. doi: 10.1002/jps.21898. [DOI] [PubMed] [Google Scholar]
  • 73.He F, Woods CE, Becker GW, Narhi LO, Razinkov VI. High-throughput assessment of thermal and colloidal stability parameters for monoclonal antibody formulations. J Pharm Sci. 2011;100(12):5126–5141. doi: 10.1002/jps.22712. [DOI] [PubMed] [Google Scholar]
  • 74.Yagi H, Takahashi N, Yamaguchi Y, Kato K. Temperature-dependent isologous Fab-Fab interaction that mediates cryocrystallization of a monoclonal immunoglobulin G. Mol Immunol. 2004;41(12):1211–1215. doi: 10.1016/j.molimm.2004.06.003. [DOI] [PubMed] [Google Scholar]
  • 75.Concepcion J, Witte K, Wartchow C, Choo S, Yao D, Persson H, Wei J, Li P, Heidecker B, Ma W, Varma R, Zhao LS, Perillat D, Carricato G, Recknor M, Du K, Ho H, Ellis T, Gamez J, Howes M, Phi-Wilson J, Lockard S, Zuk R, Tan H. Label-free detection of biomolecular interactions using BioLayer interferometry for kinetic characterization. Comb Chem High Throughput Screen. 2009;12(8):791–800. doi: 10.2174/138620709789104915. [DOI] [PubMed] [Google Scholar]
  • 76.Silverstein AM. Autoimmunity versus horror autotoxicus: the struggle for recognition. Nature immunology. 2001;2(4):279–281. doi: 10.1038/86280. [DOI] [PubMed] [Google Scholar]
  • 77.Wardemann H, Yurasov S, Schaefer A, Young JW, Meffre E, Nussenzweig MC. Predominant autoantibody production by early human B cell precursors. Science. 2003;301(5638):1374–1377. doi: 10.1126/science.1086907. [DOI] [PubMed] [Google Scholar]
  • 78.Lueking A, Beator J, Patz E, Mullner S, Mehes G, Amersdorfer P. Determination and validation of off-target activities of anti-CD44 variant 6 antibodies using protein biochips and tissue microarrays. BioTechniques. 2008;45(4):1–5. doi: 10.2144/000112898. [DOI] [PubMed] [Google Scholar]
  • 79.Jacobs SA, Wu SJ, Feng Y, Bethea D, O’Neil KT. Cross-interaction chromatography: a rapid method to identify highly soluble monoclonal antibody candidates. Pharm Res. 2010;27(1):65–71. doi: 10.1007/s11095-009-0007-z. [DOI] [PubMed] [Google Scholar]
  • 80.Spencer S, Bethea D, Raju TS, Giles-Komar J, Feng Y. Solubility evaluation of murine hybridoma antibodies. mAbs. 2012;4(3):319–325. doi: 10.4161/mabs.19869. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES