Application of Advanced Proteomic and Mass Spectrometry Technologies to Enhance the Efficacy of Production of Biopharmaceuticals

Laura Bryan1 and Paula Meleady1,2*

  • 1 National Institute for Cellular Biotechnology
  • 2 School of Biotechnology, Dublin City University, Dublin 9, Ireland
  • *Corresponding Author

Abstract

The Chinese Hamster Ovary (CHO) cell line is the dominant mammalian expression system for biopharmaceutical production. Improving the efficiency of production of these biologics will be critical in controlling costs to healthcare systems as more of these drugs come to market. Proteomic and mass spectrometry analysis has been used to profile recombinant CHO cells to get a greater understanding of the molecular and cellular mechanisms that result in a high producing cellular phenotype. In this review we will focus on the application of mass spectrometry (MS) to characterize the proteome of CHO cells and also to delve deeper into the proteome by discussing how mass spectrometry analysis of cellular post translational modifications could potentially play a key role in understanding transcription, growth, cellular signalling events, protein degradation and apoptosis in recombinant CHO cells. Such events are likely to be key to understanding and controlling bioprocess relevant phenotypes to improve the efficacy of production of high quality biotherapeutic proteins.

Introduction

Chinese Hamster Ovary (CHO) cells are known as the workhorses of the biopharmaceutical industry for the production of biotherapeutics.1 In 2017, the annual sales of biopharmaceuticals were estimated at $188 billion worldwide.2 The number of approved products in Europe and the US had steadily increased to 374 in 2018, of which 45 have “blockbuster” status, i.e. sales over $1 billion per year, with monoclonal antibodies (Mabs) representing the most lucrative single product class.2 In April 2021, a landmark was reached with the approval by the FDA of the 100th Mab product, GlaxoSmithKline’s PD1 blocker dostarlimab.3 Currently over 84% of Mabs are produced in CHO cells.2 The CHO cell line, therefore, is still the mammalian cell expression system of choice for the production of recombinant therapeutic proteins due to their ability to produce grams/L of high quality protein, the availability of highly effective gene amplification systems, their ability to undertake human-like post-translational modifications (e.g. glycosylation) and protein folding, their robust nature (e.g. conducive to growth in large-scale suspension culture), their track record in industry, and their safety record.2,4 As a result, there is considerable scientific and commercial interest in research that could improve the efficiency of CHO cell production processes as this dominance is unlikely to change for the foreseeable future.2 Within biopharmaceutical companies, the drug discovery pipelines are also producing increasingly complex modalities including fusion proteins, antibody drug conjugates (ADCs), bispecifics, nanobodies and oligomeric structures, which are putting increasing pressure on cell line development groups to generate sufficient, high quality material to meet requirements for clinical trials.5

One of the primary aims of pharmaceutical companies is lowering costs associated with development and processes, which makes genetic engineering an appealing area of research as the results can potentially lead to cost reductions. Expression vector improvements have resulted in up to 20-fold improvements in product titers since the 1990s; however, most of these improvements are attributed to optimized feeding strategies and bioprocess regimes. Genetic engineering of CHO cells has become an area of interest amongst researchers in recent years; this is likely due to the publishing of the CHO genome in 2013.6 CHO cell lines have also been shown to demonstrate a high degree of heterogeneity6 which can lead to problems in variation on the biopharmaceutical product produced.7 This product variation needs to be extensively characterized to ensure the product is safe and efficacious for therapeutic use in patients.

High-producing CHO cell culture processes have been achieved to date using approaches such as optimization of media formulations, improvements in expression vector design and selection systems, and process improvements such as in the design of bioreactors.8,9 Genetic engineering of the host (CHO) cell itself is also increasingly being used to alter the expression of endogenous genes, in order to improve the efficiency of biotherapeutic protein production.10 This has been hugely advanced through the development and application of the gene editing tool, CRISPR/Cas,9 to address host- or biopharmaceutical-specific product quality problems. These bioprocessing innovations and cell engineering efforts have improved product titer and product quality; however, uncharacterized cellular processes and gene regulatory mechanisms still hinder cell growth, specific productivity, and protein quality.9 In order to enhance the production capabilities and efficiency of the host cell line, an increased understanding of cellular physiology of CHO cells, particularly using advanced ‘omic methods and systems biology, is of critical importance.9,11 Post-transcriptional bottlenecks suggest there is scope for improvement of secretory pathway machinery in CHO cells. The secretory machinery of CHO cells must also be optimized for the production of more difficult to produce recombinant proteins such as fusion proteins which are understood to be more prone to misfolding.12

There are substantial research efforts in progress focusing on systems biology approaches to extensively characterize CHO cell factories. The application of ‘omics-based approaches (e.g. genomics, transcriptomics, proteomics, glycomics, metabolomics, fluxomics, etc.) to CHO cells will enable a fundamental understanding of CHO cell physiology and as a result a better knowledge and understanding of recombinant protein production.9 In particular, as the available omics data for CHO cells continue to grow, genomic scale models will be increasingly used to unravel the biological drivers of protein production.13 Analysis at the protein level, rather than a the genomic or transcriptomic level, may more directly reflect cellular functions. It has also been shown in many studies that there is a low correlation between gene copy number in the genome and the relative changes at the protein level, meaning that many genomic variations are not or only partially translated to the protein level.14 In addition, there can be a lack of correlation between mRNA and protein levels, thus examination of the mRNA alone does not necessarily reflect the active cellular functions,15 hence the importance of studying the proteome in addition to the genome and transcriptome. In addition, proteins are subjected to a variety of post-translational modifications (PTMs) that affect biological activity. The post-translational modification of a protein (e.g. by ubiquitination, phosphorylation, glycosylation, methylation, etc.) is one of the most important mechanisms for regulating protein function by altering activity, cellular location, turnover, and interaction with other proteins;16 however, alterations to post translational modification of cellular proteins is understudied in recombinant CHO cells.

Proteomic Analyses of CHO Cells Using Mass Spectrometry

Most proteomic studies aim to identify, quantify and characterize expression levels in comparative analysis of proteins. Studies can also focus on determining subcellular location and determining the roles of certain proteins. Top down proteomics is the mass analysis of whole proteins and their isoforms. Advancements in MS technology along with protein chemistry have allowed for the structural analysis of proteins and their complexes, and requires less protein than other structural analysis techniques.17 Characterization of the protein product itself is a hugely important technique in the biopharmaceutical industry to ensure reproducibility of products between batches.7,18 Bottom-up proteomics identifies proteins using partial characterization of their amino acid sequence, such as through the use of peptide mapping. In order for this to be possible proteolytic enzyme digestion of intact proteins is carried out before mass spectrometry analysis. Peptides are identified by comparing the masses of proteolytic peptides with predicted proteolytic peptides from a known digested sequence database. Identification of multiple peptide sequences is used for protein identification. Liquid chromatography (LC)-MS/MS analysis requires bioinformatic tools to analyze the raw mass spectral data and convert it into peptide and protein identifications. These bioinformatic tools are used to match large quantities of mass data against predicted masses from protein sequence databases and to perform statistical analysis of samples.

Proteomic profiling of CHO cells themselves can be used to gain an improved understanding of the systems biology of CHO cells, which could enable cellular optimization of bioprocessing.19 The CHO cell genome was first sequenced in 2011 and reported in 2013.20 Also in recent years, data relating to the proteome, phosphoproteome, glycoproteome and miRNA profile of CHO cells have become available by the collaborative efforts of many researchers.21–29 The information gathered by these studies has been imperative to advancing the fundamental knowledge of CHO cell biology and understanding what creates advantageous phenotypes.

Quantitative proteomics allows us to measure changes in protein abundance between different types of cells or between the same cells in different conditions, e.g. high recombinant protein producing CHO cells and low recombinant protein producing CHO cells. Quantitative proteomics in combination with LC-MS/MS has become more popular in recent years due to advancements in mass spectrometry instrumentation allowing for the identification of hundreds or thousands of proteins per sample. Methods for both absolute and relative quantitation of peptide abundance have been developed. Proteomic techniques have been used to achieve relative and absolute quantification of CHO cell proteins. Quantitative label free LC-MS/MS methods as well as methods using metabolic and isobaric labels such as Stable Isotope Labeling using Amino Acids in Culture (SILAC), Tandem Mass Tags (TMT) and Isobaric Tags for Relative and Absolute Quantitation (iTRAQ) have been utilized in the comparison and quantification of proteins in high and low Qp CHO cell lines. For example, iTRAQ labelling has been used to quantitatively identify differentially expressed proteins in high and low Qp (specific productivity) DHFR-GFP fusion protein producing CHO cells and identified an upregulation of proteins associated with protein folding and metabolism.30 SILAC has also been used as a method to investigate the proteome of high producing CHO cell lines after supplementation with sodium butyrate and low temperature culture.31 The study found that an increased Qp is associated with sodium butyrate supplementation, and lower temperature culture was associated with an expansion of the cells secretory machinery. Carlage et al. used iTRAQ labeling to show that molecular chaperones such as glucose regulator protein 78 (GRP78) and protein disulfide isomerases were differentially expressed in a Bcl-xL overexpressing mAb producing CHO cell line during the stationary phase of growth.32 More recently, Sommeregger et al. used label-free LC-MS/MS to investigate the proteome of CHO cells expressing two similar single-chain variable fragment antibodies with variations in their thermal stability.33 They found that CHO cell clones expressing antibody fragments which were less thermally stable also had higher expression of proteins associated with proliferation, apoptosis and cellular stress.

Analysis of Cellular Post Translational Modifications in CHO Cells Using Mass Spectrometry

The vast majority of ‘omic based studies that have been carried out to date in CHO cells have failed to include data on the post-translational level of regulation; however, post translational modifications are crucial to the regulation of many cellular processes including transcription, cytoskeletal rearrangement, cell proliferation, differentiation, apoptosis, protein degradation and signal transduction pathways.

These cellular processes are likely to be hugely important and key to understanding and controlling bioprocess-relevant phenotypes. PTMs determine protein function by altering activity, cellular location, turnover, and interaction with other proteins.16 

However, there is little overall work done to date to characterize the role of cellular PTMs in regulating bioprocess-relevant phenotypes, particularly in CHO cells.34 Genomic and transcriptomic approaches are blind to PTMs, making proteomics and LC-MS technologies the only way to study such modifications on a large scale.35 PTMs can be static or dynamic, altering the chemical state of a protein resulting in increased diversity and complexity of the proteome.36 Recent cutting-edge advances in MS based approaches (i.e. high speed, high sensitivity and high resolution) has allowed PTMs to be studied in unprecedented detail in order to evaluate their functional role in the cell.37 

As previously outlined, there have been a number of proteomic studies in CHO to understanding phenotypes in relation to growth, productivity and product quality by our group and others.26,38–44 In contrast, very few studies have focused on post-translational modifications on cellular proteins (other than glycosylation 28) in CHO cells, which is surprising given their pivotal role in regulating cellular events such as growth, protein translation and apoptosis.45 Our group was first in field to publish global phosphoproteomic analyses of rCHO cells following temperature shift26 and growth phases in batch culture.46 In these studies, we have shown specific changes to the phosphorylation status of many proteins of the mTOR and autophagy pathways, 46 pathways which have previously been manipulated in CHO cells to boost growth and productivity.47–49 These studies show the additional beneficial information generated by including the analysis of cellular PTMs in CHO proteomic studies. Phosphoproteomic analysis also helped show that increased interaction of CREB1 with transgene promoter caused Mab production to be increased in CHO cell lines.50 Recently we have shown major changes in the phosphoproteome of industrially relevant high and low productivity IgG4 producing CHO cells.43

Other cellular post translational modifications that have the potential to be more extensively studied in CHO cells include glycosylation, methylation and ubiquitination. For example, ubiquitination involves the conjugation of the 8.5kD protein ubiquitin to target proteins marking them for proteasomal degradation.51 However, it is becoming increasingly clear that protein ubiquitination does not just target substrates for degradation but also has important regulatory functions similar to phosphorylation, including control of: cell cycle regulation; gene expression; apoptosis; signal transduction; cellular localisation of proteins; and modulating protein-protein interactions.51,52 Consequently, defects in the ubiquitin pathway are associated with various human diseases including cancer.53 As a result of rapid developments of affinity capture reagents and high-resolution MS it is now feasible to globally analyze the ubiquitinated proteome using quantitative strategies.54

Conclusions

Although great improvements have been made to the titer and Qp in CHO cells over the past few decades, we are only at the beginning of developing a true understanding of the intracellular biology of these cells upon which the production of human therapeutic proteins highly relies.

In order to engineer CHO cell lines which are capable of efficiently producing any recombinant protein to a high Qp, we must identify general effector genes. This goal, likely requires a change in the expression of multiple genes, ultimately meaning the targeting of intracellular pathways as opposed to individual genes with individual effects. For this reason, PTMs such as phosphorylation or ubiquitination may present as better potential targets for CHO cell engineering. For example, one phosphorylation event has the potential to trigger an entire signalling pathway. Identifying PTMs which effect bioprocess related phenotypes can lead to the identification of signalling pathways which affect bioprocess related phenotypes. Signalling pathways will likely be more effective targets for CHO cell engineering as these results are less likely to be product or cell line specific. This same principle also applies to miRNAs and transcription factors, which have the ability to alter entire signalling pathways if manipulated. However, targeting entire signalling pathways as opposed to single effector genes does have its disadvantages, such as adverse phenotypic effects from the undesired regulation of a subset of genes.55

Acknowledgments

We wish to acknowledge funding from the Irish Research Council Enterprise Partnership Scheme (Project ID EPSPG/2016/10) and from Science Foundation Ireland under Grant number [19/FFP/6759].

References

  1. O’Flaherty R, Bergin A, Flampouri E, et al. Mammalian cell culture for production of recombinant proteins: A review of the critical steps in their biomanufacturing. Biotechnol Adv. 2020;43:107552. doi:10.1016/j.biotechadv.2020.107552
  2. Walsh G. Biopharmaceutical benchmarks 2018. Nat Biotechnol. 2018;36(12):1136-1145. doi:10.1038/nbt.4305
  3. Mullard A. FDA approves 100th monoclonal antibody product. Nat Rev Drug Discov. 2021;20(7):491-495. doi:10.1038/d41573-021-00079-7
  4. Wurm FM. Production of recombinant protein therapeutics in cultivated mammalian cells. Nat Biotechnol. 2004;22(11):1393-1398. doi:10.1038/nbt1026
  5. Husain B, Ellerman D. Expanding the Boundaries of Biotherapeutics with Bispecific Antibodies. BioDrugs. 2018;32(5):441-464. doi:10.1007/s40259-018-0299-9
  6. Lewis NE, Liu X, Li Y, et al. Genomic landscapes of Chinese hamster ovary cell lines as revealed by the Cricetulus griseus draft genome. Nat Biotechnol. 2013;31(8):759-765. doi:10.1038/nbt.2624
  7. Beyer B, Schuster M, Jungbauer A, Lingg N. Microheterogeneity of Recombinant Antibodies: Analytics and Functional Impact. Biotechnol J. 2018;13(1):1700476. doi:10.1002/biot.201700476
  8. Hacker DL, De Jesus M, Wurm FM. 25 years of recombinant proteins from reactor-grown cells - where do we go from here? Biotechnol Adv. 2009;27(6):1023-1027. doi:10.1016/j. biotechadv.2009.05.008
  9. Kuo C-C, Chiang AW, Shamie I, Samoudi M, Gutierrez JM, Lewis NE. The emerging role of systems biology for engineering protein production in CHO cells. Curr Opin Biotechnol. 2018;51:64-69. doi:10.1016/j.copbio.2017.11.015
  10. Fischer S, Handrick R, Otte K. The art of CHO cell engineering: A comprehensive retrospect and future perspectives. Biotechnol Adv. 2015;33(8):1878-1896. doi:10.1016/j. biotechadv.2015.10.015
  11. Stolfa G, Smonskey MT, Boniface R, et al. CHO-Omics Review: The Impact of Current and Emerging Technologies on Chinese Hamster Ovary Based Bioproduction. Biotechnol J. 2018;13(3):e1700227. doi:10.1002/biot.201700227
  12. Aggarwal RS. What’s fueling the biotech engine-2012 to 2013. Nat Biotechnol. 2014;32(1):32-39. doi:10.1038/nbt.2794
  13. Antonakoudis A, Barbosa R, Kotidis P, Kontoravdi C. The era of big data: Genome-scale modelling meets machine learning. Comput Struct Biotechnol J. 2020;18:3287-3300. doi:10.1016/j.csbj.2020.10.011
  14. Geiger T, Cox J, Mann M. Proteomic changes resulting from gene copy number variations in cancer cells. PLoS Genet. 2010;6(9):e1001090. doi:10.1371/journal.pgen.1001090
  15. Schwanhäusser B, Busse D, Li N, et al. Global quantification of mammalian gene expression control. Nature. 2011;473(7347):337-342. doi:10.1038/nature10098
  16. Choudhary C, Mann M. Decoding signalling networks by mass spectrometry-based proteomics. Nat Rev Mol Cell Biol. 2010;11(6):427-439. doi:10.1038/nrm2900
  17. Artigues A, Nadeau OW, Rimmer MA, et al. Protein Structural Analysis via Mass Spectrometry-Based Proteomics. Adv Exp Med Biol. 2016;919:397-431. doi:10.1007/978- 3-319-41448-5_19
  18. Füssl F, Strasser L, Cari S, Bones J. Native LC-MS for capturing quality attributes of biopharmaceuticals on the intact protein level. Curr Opin Biotechnol. 2021;71:32-40. doi:10.1016/j.copbio.2021.05.008
  19. Datta P, Linhardt RJ, Sharfstein ST. An ’omics approach towards CHO cell engineering. Biotechnol Bioeng. 2013;110(5):1255-1271. doi:10.1002/bit.24841
  20. Xu X, Nagarajan H, Lewis NE, et al. The genomic sequence of the Chinese hamster ovary (CHO)-K1 cell line. Nat Biotechnol. 2011;29(8):735-741. doi:10.1038/nbt.1932
  21. Baycin-Hizal D, Tabb DL, Chaerkady R, et al. Proteomic analysis of Chinese hamster ovary cells. J Proteome Res. 2012;11(11):5265-5276. doi:10.1021/pr300476w
  22. Becker J, Hackl M, Rupp O, et al. Unraveling the Chinese hamster ovary cell line transcriptome by next-generation sequencing. J Biotechnol. 2011;156(3):227-235. doi:10.1016/j.jbiotec.2011.09.014
  23. Clarke C, Henry M, Doolan P, et al. Integrated miRNA, mRNA and protein expression analysis reveals the role of post-transcriptional regulation in controlling CHO cell growth rate. BMC Genomics. 2012;13:656. doi:10.1186/1471-2164-13-656
  24. Hackl M, Jakobi T, Blom J, et al. Next-generation sequencing of the Chinese hamster ovary microRNA transcriptome: Identification, annotation and profiling of microRNAs as targets for cellular engineering. J Biotechnol. 2011;153(1-2):62-75. doi:10.1016/j. jbiotec.2011.02.011
  25. Hefzi H, Ang KS, Hanscho M, et al. A Consensus Genome-scale Reconstruction of Chinese Hamster Ovary Cell Metabolism. Cell Syst. 2016;3(5):434-443.e8. doi:10.1016/j. cels.2016.10.020
  26. Henry M, Power M, Kaushik P, Coleman O, Clynes M, Meleady P. Differential Phosphoproteomic Analysis of Recombinant Chinese Hamster Ovary Cells Following Temperature Shift. J Proteome Res. 2017;16(7):2339-2358. doi:10.1021/acs. Jproteome.6b00868
  27. Li S, Cha SW, Heffner K, et al. Proteogenomic Annotation of Chinese Hamsters Reveals Extensive Novel Translation Events and Endogenous Retroviral Elements. J Proteome Res. 2019;18(6):2433-2445. doi:10.1021/acs.jproteome.8b00935
  28. Yang G, Hu Y, Sun S, et al. Comprehensive Glycoproteomic Analysis of Chinese Hamster Ovary Cells. Anal Chem. 2018;90(24):14294-14302. doi:10.1021/acs.analchem.8b03520
  29. Bryan L, Henry M, Barron N, et al. Differential expression of miRNAs and functional role of mir-200a in high and low productivity CHO cells expressing an Fc fusion protein. Biotechnol Lett. 2021;43(8):1551-1563. doi:10.1007/s10529-021-03153-7
  30. Nissom PM, Sanny A, Kok YJ, et al. Transcriptome and proteome profiling to understanding the biology of high productivity CHO cells. Mol Biotechnol. 2006;34(2):125-140. doi:10.1385/mb:34:2:125
  31. Kantardjieff A, Jacob NM, Yee JC, et al. Transcriptome and proteome analysis of Chinese hamster ovary cells under low temperature and butyrate treatment. J Biotechnol. 2010;145(2):143-159. doi:10.1016/j.jbiotec.2009.09.008
  32. Carlage T, Kshirsagar R, Zang L, et al. Analysis of dynamic changes in the proteome of a Bcl-XL overexpressing Chinese hamster ovary cell culture during exponential and stationary phases. Biotechnol Prog. 2012;28(3):814-823. doi:10.1002/btpr.1534
  33. Sommeregger W, Mayrhofer P, Steinfellner W, et al. Proteomic differences in recombinant CHO cells producing two similar antibody fragments. Biotechnol Bioeng. 2016;113(9):1902- 1912. doi:10.1002/bit.25957
  34. Bryan L, Clynes M, Meleady P. The emerging role of cellular post-translational modifications in modulating growth and productivity of recombinant Chinese hamster ovary cells. Biotechnol Adv. 2021;49:107757. doi:10.1016/j.biotechadv.2021.107757
  35. Angel TE, Aryal UK, Hengel SM, et al. Mass spectrometry-based proteomics: existing capabilities and future directions. Chem Soc Rev. 2012;41(10):3912-3928. doi:10.1039/ c2cs15331a
  36. Simon GM, Cravatt BF. Challenges for the “chemical-systems” biologist. Nat Chem Biol. 2008;4(11):639-642. doi:10.1038/nchembio1108-639
  37. Thygesen C, Boll I, Finsen B, Modzel M, Larsen MR. Characterizing disease-associated changes in post-translational modifications by mass spectrometry. Expert Rev Proteomics. 2018;15(3):245-258. doi:10.1080/14789450.2018.1433036
  38. Meleady P, Doolan P, Henry M, et al. Sustained productivity in recombinant Chinese hamster ovary (CHO) cell lines: proteome analysis of the molecular basis for a process- related phenotype. BMC Biotechnol. 2011;11:78. doi:10.1186/1472-6750-11-78
  39. Meleady P, Henry M, Gammell P, et al. Proteomic profiling of CHO cells with enhanced rhBMP-2 productivity following co-expression of PACEsol. Proteomics. 2008;8(13):2611- 2624. doi:10.1002/pmic.200700854
  40. Doolan P, Meleady P, Barron N, et al. Microarray and proteomics expression profiling identifies several candidates, including the valosin-containing protein (VCP), involved in regulating high cellular growth rate in production CHO cell lines. Biotechnol Bioeng. 2010;106(1):42-56. doi:10.1002/bit.22670
  41. Carlage T, Hincapie M, Zang L, et al. Proteomic profiling of a high-producing Chinese hamster ovary cell culture. Anal Chem. 2009;81(17):7357-7362. doi:10.1021/ac900792z
  42. Hausmann R, Chudobová I, Spiegel H, Schillberg S. Proteomic analysis of CHO cell lines producing high and low quantities of a recombinant antibody before and after selection with methotrexate. J Biotechnol. 2018;265:65-69. doi:10.1016/j.jbiotec.2017.11.008
  43. Bryan L, Henry M, Kelly RM, et al. Global phosphoproteomic study of high/low specific productivity industrially relevant mAb producing recombinant CHO cell lines. Current Research in Biotechnology. 2021;3:49-56. doi:10.1016/j.crbiot.2021.02.002
  44. Bryan L, Henry M, Kelly RM, et al. Mapping the molecular basis for growth related phenotypes in industrial producer CHO cell lines using differential proteomic analysis. BMC Biotechnol. 2021;21(1):43. doi:10.1186/s12896-021-00704-8
  45. Farrell A, McLoughlin N, Milne JJ, Marison IW, Bones J. Application of multi-omics techniques for bioprocess design and optimization in chinese hamster ovary cells. J Proteome Res. 2014;13(7):3144-3159. doi:10.1021/pr500219b
  46. Kaushik P, Henry M, Clynes M, Meleady P. The Expression Pattern of the Phosphoproteome Is Significantly Changed During the Growth Phases of Recombinant CHO Cell Culture. Biotechnol J. 2018;13(10):e1700221. doi:10.1002/biot.201700221
  47. Dreesen IAJ, Fussenegger M. Ectopic expression of human mTOR increases viability, robustness, cell size, proliferation, and antibody production of chinese hamster ovary cells. Biotechnol Bioeng. 2011;108(4):853-866. doi:10.1002/bit.22990
  48. Baek E, Kim CL, Kim MG, Lee JS, Lee GM. Chemical inhibition of autophagy: Examining its potential to increase the specific productivity of recombinant CHO cell lines. Biotechnol Bioeng. 2016;113(9):1953-1961. doi:10.1002/bit.25962
  49. Dadehbeigi N, Dickson AJ. Chemical manipulation of the mTORC1 pathway in industrially relevant CHOK1 cells enhances production of therapeutic proteins. Biotechnol J. 2015;10(7):1041-1050. doi:10.1002/biot.201500075
  50. Dahodwala H, Kaushik P, Tejwani V, et al. Increased mAb production in amplified CHO cell lines is associated with increased interaction of CREB1 with transgene promoter. Curr Res Biotechnol. 2019;1:49-57. doi:10.1016/j.crbiot.2019.09.001
  51. Ordureau A, Münch C, Harper JW. Quantifying ubiquitin signaling. Mol Cell. 2015;58(4):660- 676. doi:10.1016/j.molcel.2015.02.020
  52. Sadowski M, Suryadinata R, Tan AR, Roesley SNA, Sarcevic B. Protein monoubiquitination and polyubiquitination generate structural diversity to control distinct biological processes. IUBMB Life. 2012;64(2):136-142. doi:10.1002/iub.589
  53. Zou T, Lin Z. The Involvement of Ubiquitination Machinery in Cell Cycle Regulation and Cancer Progression. Int J Mol Sci. 2021;22(11):5754. doi:10.3390/ijms22115754
  54. van der Wal L, Bezstarosti K, Sap KA, et al. Improvement of ubiquitylation site detection by Orbitrap mass spectrometry. J Proteomics. 2018;172:49-56. doi:10.1016/j. jprot.2017.10.014
  55. Stearns FW. One hundred years of pleiotropy: a retrospective. Genetics. 2010;186(3):767-773. doi:10.1534/genetics.110.122549

Author Biographies

Laura Bryan is a PhD candidate in CHO cell engineering at the National Institute for Cellular Biotechnology, Dublin City University (DCU) under the supervision of Dr. Paula Meleady. She received an Irish Research Council Enterprise Partnership Scheme award to carry out her research project. She received a BSc in Cell and Molecular Biology in 2015 from University College Dublin, Ireland.

Dr. Paula Meleady is a Principal Investigator at the National Institute for Cellular Biotechnology and an Assistant Professor in the School of Biotechnology, DCU, Ireland. She received her BSc in Biotechnology from DCU in 1993 and a PhD from DCU in 1997. Her research interests are focused on the application of advanced proteomic and mass spectrometry methods to understand biological systems, in particular characterization of recombinant mammalian cell lines (Chinese hamster ovary) in order to gain insights to improving efficiency of production of biopharmaceuticals. She has co-authored over 120 peer-reviewed publications and 12 book chapters to date in research areas related to proteomics, bioprocessing and cancer.

Subscribe to our e-Newsletters
Stay up to date with the latest news, articles, and events. Plus, get special offers
from American Pharmaceutical Review – all delivered right to your inbox! Sign up now!

  • <<
  • >>

Join the Discussion