Next Article in Journal
N-Glycosylation Patterns across the Age-Related Macular Degeneration Spectrum
Previous Article in Journal
Positive Effect of Ecological Restoration with Fabaceous Species on Microbial Activities of Former Guyanese Mining Sites
Previous Article in Special Issue
Dealing with the Ambiguity of Glycan Substructure Search
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational Modeling of O-Linked Glycan Biosynthesis in CHO Cells

1
Department of Bioinformatics, Graduate School of Engineering, Soka University, Tokyo 192-8577, Japan
2
Department of Cardiology, Keio University School of Medicine, Tokyo 160-8582, Japan
3
Proteomics Core Facility at Sahlgrenska Academy, University of Gothenburg, 405 30 Gothenburg, Sweden
4
Department of Life Sciences and Health, Faculty of Health Sciences, Oslo Metropolitan University, 0167 Oslo, Norway
5
Glycan & Life Systems Integration Center (GaLSIC), Soka University, Tokyo 192-8577, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(6), 1766; https://doi.org/10.3390/molecules27061766
Submission received: 30 October 2021 / Revised: 22 February 2022 / Accepted: 24 February 2022 / Published: 8 March 2022

Abstract

:
Glycan biosynthesis simulation research has progressed remarkably since 1997, when the first mathematical model for N-glycan biosynthesis was proposed. An O-glycan model has also been developed to predict O-glycan biosynthesis pathways in both forward and reverse directions. In this work, we started with a set of O-glycan profiles of CHO cells transiently transfected with various combinations of glycosyltransferases. The aim was to develop a model that encapsulated all the enzymes in the CHO transfected cell lines. Due to computational power restrictions, we were forced to focus on a smaller set of glycan profiles, where we were able to propose an optimized set of kinetics parameters for each enzyme in the model. Using this optimized model we showed that the abundance of more processed glycans could be simulated compared to observed abundance, while predicting the abundance of glycans earlier in the pathway was less accurate. The data generated show that for the accurate prediction of O-linked glycosylation, additional factors need to be incorporated into the model to better reflect the experimental conditions.

1. Introduction

The value of recombinant proteins as a main therapeutic approach, where human-like proteins are produced in host cells, has been increasing. The pharmacokinetics of these proteins in the human body vary according not only to the amino acid sequence, but also the glycans attached to the protein during the synthesis process. The enzyme-mediated process of adding glycans to a protein is a crucial post-translational modification (PTM) for its localization and trafficking, protein solubility, antigenicity, and cell–cell interaction [1,2,3,4]. For that reason, researchers have been studying the glycosylation process in order to understand and predict the produced glycan profiles of a protein in different biological systems.
One approach is computational modeling of the glycosylation process and studying the effect of changing variables on the final glycan profiles. The first mathematical model for N-glycan biosynthesis was developed in 1997 by Umana and Bailey (UB1997), simulating the initial stages of N-glycan biosynthesis from a single core structure [5]. Their model was built on the basic assumption that the four well-mixed compartments of the Golgi apparatus contain a set of eight glycosyltransferases. These enzymes catalyzed 33 reactions and generated 33 different oligosaccharide species of N-linked glycans attached to a specified protein. The model parameters were estimated from the literature describing Chinese Hamster Ovary (CHO) cells. The N-glycan profile predicted by the model was close to that of recombinant proteins produced by CHO cells. The UB1997 model laid the foundation for N-glycan modeling and was thereafter extended to many models.
In 2005, Krambeck and Betenbaugh extended UB1997 to include additional galactosylation, fucosylation, extension by adding N-acetyllactosamine repeats, and sialylation [6]. This KB2005 model generated 7565 oligosaccharide structures in a network of 22,871 reactions in terms of 11 enzymes and 20 reaction rules. The values of the model parameters were either derived from the UB1997 model or literature, or were adjusted in order to match the experimental N-glycan profiles in CHO cells. This model was further expanded and adjusted in 2009 by Krambeck et al. to estimate human N-glycans [7]. The KB2009 model included not only 19 enzymes with 26 reaction rules, but also a cutoff rule to stop the reactions after a specified molecular mass was reached. In 2017, another important model was developed by Krambeck and others, which successfully predicted the N-linked glycoform profiles of ten CHO cell lines, including one wild type and nine mutants [8].
By comparison, only a limited number of models were designed to simulate O-glycan synthesis. In 2008, Liu and co-workers modeled the first O-glycosylation reaction network through “object-oriented programming” [9]. The model started from core 2 structure Galβ1-3 (GlcNAcβ1-6) GalNAcol as the initial structure, while sialyl Lewis-X (sLex), a core 2 based O-linked glycan, was the only final product. In addition to core 2 and sLex, the reaction network contained another eighteen “intermediate” structures involved in 28 reactions. This model was fitted via sensitivity analysis to predict levels of sLex on (P-selectin glycoprotein ligand 1) PSGL-1 close to experimental data. On the other hand, McDonald et al. utilized a different methodology to develop their O-glycosylation network prediction software, called “O-Glycologue” [10]. This prediction tool included 25 enzymes and generated over 13 million possible unique structures in over 34 million reactions. O-Glycologue also has a feature to predict the synthesis pathway of any structure of the user’s choice. This feature traverses the reaction network in reverse starting from the end product, removing a monosaccharide with each reaction until it reaches the initial structure, a non-glycosylated Ser/Thr. This model did not include enzyme kinetics or mathematical modeling of the reactions. While the UB1997 model promoted N-glycan modeling and several models were built based on it, the lack of well-established studies modeling the kinetics of O-glycans’ synthesis slowed the advancement of O-glycan modeling.
In this study, we explore different methodologies for building a computational model of O-linked mucin-type glycan biosynthesis in CHO cells. We construct our models based on a set of experimental data in which the biological system of a cell line is modified by transiently transfecting the cells with different combinations of glycosyltransferases. Thereafter, the O-glycoform profile for each system is mapped and identified. We attempt to simulate these conditions and predict the subsequent glycan profile.

2. Results

2.1. Glycan Profiles Modeled in this Study

The experimental data utilized to build this model are based on our previous experiments on mapping the mucin-type O-glycan profile in CHO-K1 cells [11,12]. In total, 25 different combinations of glycosyltransferases were transiently expressed in CHO cells together with PSGL-1. Then, the mucin-type O-glycan profile was mapped for each cell line (Supplementary Materials). The recombinant protein PSGL-1 fused with the Fc portion of mouse immunoglobulin G2b (PSGL-1/mIgG2b) has more than 53 sites for mucin-type O-glycosylation [13]. The glycosyltransferases transiently expressed in CHO cells are shown in Table 1 [11,12].
Previous studies showed that the intensities of negatively charged glycan structures containing negatively charged residues, such as NeuAc, NeuGc, and/or sulfate, have a tendency to show higher signal in mass spectrometry studies [20]. Therefore, the intensities in our data were adjusted by a factor of 0.3 for singly charged glycans, and by a factor of 0.4 for doubly charged glycans. Consequently, the structures’ abundances across a glycan profile were recalculated.

2.2. Building the Models

The biosynthesis of GalNAc O-glycans is initiated in the Golgi apparatus, and carried out as the protein passes through the stacks of the Golgi apparatus and interacts with its content of glycosyltransferases (Figure 1) [21,22]. The Golgi apparatus stacks are functionally compartmentalized into the cis-Golgi network (CGN), cis-cisternae, medial-cisternae, trans-cisternae, and the trans-Golgi network (TGN) [23]. The glycosyltransferases’ distribution throughout the Golgi apparatus is not homogenous. Each compartment has a distinct composition of glycosyltransferases that reflects the sequence in which glycans are added to the proteins [24]. The cis and medial compartments contain core synthesizing enzymes, the medial and trans compartments mainly contain elongating glycosyltransferases, while capping and terminating enzymes are concentrated in trans and TGN compartments [25]. The biosynthesis of mucin-type glycans starts when an N-acetylgalactosamine (GalNAc) is attached to a polypeptide in the cis compartment [22,25]. A GalNAc linked to Ser/Thr is called a “Tn antigen”, and this reaction is catalyzed by a large family of GalNAc transferases [21]; however, it was not included in our model, since it necessitates considering the structure of the protein and the accessibility of the glycosylation sites. The Tn antigen is the initial structure in our model. This structure can be extended into one of the four cores (cores 1–4). The addition of a galactose (Gal) to the third carbon of GalNAc via a β-linkage forms “core 1 or T antigen”. This reaction is catalyzed by glycoprotein-N-acetylgalactosamine 3-beta-galactosyltransferase (C1GalT1) [26]. Competing with C1GalT1 for the same substrate (Tn Antigen) is beta-1,3-N-acetylglucosaminyltransferase (B3GNT6), which adds an N-acetylglucosamine (GlcNAc) at the same location to form “core 3” [27]. “Core 2” and “core 4” are formed by branching “core 1” and “core 3” via the attachment of a GlcNAc to the sixth carbon atom of the GalNAc, which is mediated by GCNT1 and/or GCNT3 (beta-1,6-N-acetylglucosaminyltransferase) [28,29]. All four cores are subsequently extended with LacNAc repeats, and are sulfated and/or sialylated before leaving the Golgi [21,22,25].
Considering that only transiently transfected glycosyltransferases were known, we computationally explored the reactions and glycosyltransferases required to produce the glycan structures found in each experiment. In total, 20 different enzymes were found to be active in producing all glycans, 12 of which were known transfected enzymes (Table 2). In case a family of enzymes catalyzed the same reaction, it was considered one enzyme for simplification reasons, as in B3GNT(s), B4GALT(s), ST3GAL(s), and ST6GALNAC(s). Four of the enzymes were O-glycan core synthesizing enzymes (C1GALT1, GCNT1/3, and B3GNT6). The list also included four fucosyltransferases (FUT2, 3, 4 and 7), three sialyltransferases (ST3GAL, ST6GAL1, and ST6GALNAC) and a sulfotransferase (CHST4). Six of the enzymes (B3GNT, B3GNT3, CHST4, GCNT1, ST3GAL, and ST6GALNAC) reacted with more than one substrate; therefore, multiple reaction rules were considered for these enzymes, and were differentiated by adding a small letter to the enzyme name. B3GNT_a and B3GNT_b are two reactions catalyzed by B3GNT(s). Our models included 30 reaction rules for 20 enzymes. The glycosyltransferases, their substrates, and the reactions they catalyze were derived from KEGG, and are listed in the Supplementary Materials [30].
Glycosylation reaction networks were generated computationally per experiment and later incorporated into the mathematical model. The model of an engineered CHO cell line included the glycosyltransferases transiently expressed in the engineered cells and the glycosyltransferases required for the reaction network of that model to pro-duce the final glycan profile (Figure 2). In case an enzyme catalyzed two different re-actions, the model included the reaction rule included in the reaction network of that model.
To modulate the biosynthesis of mucin-type O-glycans, we designed a four-compartment model: “cis”, “medial”, “trans”, and “TGN”. The four compartments were set at a similar size and each compartment was assumed to be homogenous [5,6]. The glycosyltransferases were distributed across the four compartments depending on their role and order in the biosynthesis pathway. The protein residence time in a compartment was 5.56 min, during which the protein interacted with the glycosyltransfer-ases in that compartment following the modified Michaelis–Menten kinetics equation used in previous N-glycosylation models [6,7,8]. Then, it was transported to the next compartment following the Golgi mathematical model used in the Umana and Baily model (1997) [5]. After 22.24 min, the simulation ended, and the concentrations of structures accumulated in the TGN compartment were calculated. The sugar donor concentrations across the Golgi shown in Table 3 were based on previous works by Krambeck et al. [7,8]. The PAP_S initial concentration in Golgi was estimated based on Krambeck 2005 [6,31]. The CMP_NeuGc donor concentration was assumed to be simi-lar to CMP_NeuAc. The nucleotide-sugar donors in previous models were distributed equally across the compartments; however, in our model, this was limited to the com-partments where they donated the sugar, which was based on the enzyme distribution.
The distribution of the enzymes across Golgi compartments shown in Figure 3 was set based on the reactions each enzyme catalyzes (core synthesis, elongation, capping etc.), and the localization of its substrates and the enzymes competing for that sub-strate. The core 1 and core 3 synthesizing enzymes compete for Tn Antigen and were localized in the “cis” compartment. Core 2 and core 4 synthesizing enzymes were as-sumed to peak in the second compartment. The elongating and capping glycosyltrans-ferases were localized in the last two compartments of trans and TGN. Fucosyltrans-ferases were distributed in the medial and trans compartments, while CHST4 was dis-tributed over all the compartments except cis.

2.3. Parameter Estimation and Simulations

The mathematical model based on a modified Michaelis–Menten kinetics equation requires enzyme concentrations, nucleotide-sugar donor concentrations and the kinetic parameters Km, Kmd, and Kf for each enzyme. Nucleotide-sugar donor concentrations and enzyme localizations are illustrated in Figure 3 and Table 3. However, finding the rest of the parameters in the literature on CHO was challenging. Therefore, the tool Parameter Estimation in COPASI was utilized to find the initial values of the parameters to run time-course simulations and compare the resulting glycan profiles with the experimental glycan profiles. This process was computationally demanding; therefore, only experiments with a small glycan profile were modeled. The models “CHO” and “CHO-WT” model the mucin-type O-glycan profile in non-modified CHO cells. The glycan profile for these cells included three glycan structures, shown in Figure 4: “Galβ1-3(NeuAcα2-6)GalNAcol”, “NeuAcα2-3Galβ1-3GalNAcol”, and “NeuAcα2-3Galβ1-3(NeuAcα2-6)GalNAcol”. The simulation results in this model are shown in Figure 4. The line chart (Figure 4A) shows the change in concentration of the three main glycan structures in the TGN over the course of the simulation. Comparing the glycan profile predicted by the model and the experimental glycan profile in Figure 4B shows that the two profiles were almost identical. This indicates that using Parameter Estimation is a useful tool to generate possible values for our models’ parameters. However, running the same process for larger models proved to be challenging due to the currently available computing power.
To validate these results, the parameter estimation and time-course simulation were repeated 20 times. The simulation results are identical to the experimental data; however, the generated parameters were found to vary, as shown in Figure 5. While there is wide variance in most of the parameters, such as the Kf value of C1GALT1 with a standard error of mean over 6000 min−1, other Kf values were less deviating, with standard errors of mean of less than 1700 min−1 for ST6GALNAC_b and ST6GALNAC_c. However, further investigation is needed.
As discussed earlier, applying parameter estimation in one experiment gave satisfactory simulation results, but the parameter values changed during each iteration. This indicates that individual parameter values were not that crucial when a limited number of reactions were considered, as in the CHO-WT model. However, this created issues down the line when expanding this to larger experimental datasets, where we had 20 transferases generating a spectrum of parameters, generating a wider spread of the simulated abundance of individual structures. The ultimate goal would be to find a single solution (a set of values for the parameters) that is applicable to all the models, and that results in satisfactory glycan profile prediction for all the models. From there, we considered combining all the models into one large model and applying parameter estimation to this model. To achieve this, we would need increased computing power, since this was not achievable in the computers available in our institute.
As an alternative approach to the computationally demanding large model, we designed a minimized model of a sample of relatively simple glycan profiles from four cell lines. Due to the limited number of glycans in their profiles, the following experiments were chosen: CHO-WT, Slex on Core 3, Slex on extended core 1, and A4GlcNAc on core 1. In this complex model, all the parameters in the four experiments were estimated simultaneously, then time-course simulations for each experiment were conducted separately. The calculation time for this model was about 4 h. A comparison of the simulation results with experimental results is illustrated in Figure 6, where the glycan profiles for each experiment are labeled with their experiment names. In CHO-WT and A4GLcNAc on core 1, the small number of glycans in these sets correlated well with the experimental data (Figure 6A,D). However, the simulation results for CHO-WT were less accurate when the parameters were estimated in the four-experiment model compared to the one-experiment model (Figure 4B and Figure 6A). On the other hand, there were larger disparities in the experiments generating Slex on core 3 and Slex on extended core 1. While Slex on core 1 overestimated the two most abundant glycans, it underestimated the abundances of NeuAcα2-3Galβ1-4(Fucα1-3)GlcNAcβ1-3Galβ1-3GalNAca1- and Galβ1-4GlcNAcβ1-3GalNAca1-. The disparities were even greater in Slex on core 3, where the levels of three structures were underestimated, including the final product NeuAcα2-3Galβ1-4(Fucα1-3)GlcNAcβ1-3GalNAcol. However, it predicted correctly that the galactosylated core 3 would be the dominating structure produced after transfection.

3. Discussion

Even though there were many technical challenges involved in performing large model simulations in this work, we were able to gain new insight into various aspects of O-glycan biosynthesis enzymes. We obtained estimations of the kinetic constants and concentrations for each reaction in a single experiment that gave simulation results close to the experimental data (Figure 4 and Figure 5).
On the other hand, Figure 6 illustrates that there were discrepancies between different experimental conditions when different enzymes were transiently expressed in separate experiments but on the same cell line. While our simulations were able to predict experimental glycome profiles per experiment, a combined model could not accurately predict all glycome profile across all experiments accurately. This might be due to the experimental factors involved that still need to be considered. Another possible factor that was not included in our model is simulating the initial step of the mucin-type O-linked biosynthesis, performed by the GalNAc-polypeptide transferases. In general, our simulation results could predict the more highly processed glycans more accurately than the structures that are synthesized earlier in the pathway. Indeed, it has been reported that the neutralization of pH in the Golgi apparatus causes glycosyltransferases to redistribute, causing the premature termination/elongation of O-linked glycans [32]. This suggests that the Golgi distributions of glycosyltransferases are not static. During cell culturing, increased confluence would cause some cells to experience hypoxia, affecting the cellular level of sugar nucleotides and the glycosylation machinery [33]. Therefore, it could very well be the case that such factors as the altered pH of the Golgi and hypoxia could cause the redistribution of glycosyltransferases during extended cell culturing. Altered sugar nucleotide levels would also need to be incorporated into the simulation of cell culture models to be able to make more accurate predictions about glycan biosynthesis. In addition, there is the possibility that transiently transfecting additional glycosyltransferases in CHO cells subsequently redistribute glycosyltransferases and nucleotide transporters in the Golgi compartments. The multi-assembly of glycosyltransferases and nucleotide transporters in Golgi has been shown for N-linked biosynthesis [34]. Whether O-linked glycoenzymes in the Golgi are also assembled in a similar manner, and how these assemblies are restructured as additional transferases are introduced, are unknown. Hence, for the accurate simulation of glycosylation, our initial approach of identifying a single set of parameters that could fit into one model for all 20 enzymes will have its limitations.

4. Materials and Methods

Engineering the CHO cell lines included the transient transfection of cells with different combinations of glycosyltransferases for each cell line, as described in our previous work [10,11].
The computer models were generated in extendable markup language “XML”, using our in-house developed tool “GlycoSim” [35,36]. Linear code was used to describe the glycan structures and the reaction rules. The localization of an enzyme in the four compartments was described as 0 and 1. Models were then imported into the complex pathway simulator “COPASI”, where the parameter estimations and time simulations were processed [37]. In COPASI, the volume of each compartment was changed to 2.5 µL and the initial glycan concentration was changed to 100 µmol/µL.
For the parameter estimation, we used the Evolution Strategy (SRES) method. Experimental data were matched to the concentration of the glycan structure at the TGN compartment. The weight method was set to standard deviation. Parameters to be estimated were added and cross-referenced to their relevant experiment. The lower and upper bounds were set to 10−6 and 105, respectively. In the case of the four-experiment model, we updated the model with the estimated parameters then saved it in four separate files relevant to the experiments. In each model, we set the concentrations of enzymes that were not relevant to the experiments to 0 µmol/µL. The time-course simulations deterministic (LSODA) method was used. The results were analyzed using GraphPad Prism 9 for macOS, GraphPad Software, San Diego, California USA, www.graphpad.com (accessed on 24 February 2022).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27061766/s1.

Author Contributions

Conceptualization, T.K.; formal analysis, T.K.; funding acquisition, K.F.A.-K.; investigation, I.S., C.J. and N.G.K.; methodology, T.K.; project administration, K.F.A.-K.; software, T.K., S.A. and I.S.; supervision, K.F.A.-K.; visualization, T.K.; writing—original draft, T.K.; writing—review and editing, C.J., N.G.K. and K.F.A.-K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Japan Science and Technology Agency (JST) Mirai Program “Construction of a simulation analysis framework for understanding glycan function”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available from [computational modeling of O-linked glycans biosynthesis in CHO cells].

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Rini, J.M.; Esko, J.D. Glycosyltransferases and Glycan-Processing Enzymes. In Essentials of Glycobiology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2017; pp. 65–77. ISBN 9780879697709. [Google Scholar]
  2. Varki, A.; Gagneux, P. Biological Functions of Glycans. In Essentials of Glycobiology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2017; pp. 77–89. [Google Scholar]
  3. Zhang, L.; Tian, E.; Ten Hagen, K.G. UDP-N-acetyl-alpha-DGalactosamine: Polypeptide N-Acetylgalactosaminyltransferases (ppGalNAc-Ts). In Handbook of Glycosyltransferases and Related Genes, 2nd ed.; Springer: Tokyo, Japan, 2014; Volume 1, pp. 495–511. ISBN 9784431542407. [Google Scholar]
  4. Kim, J.Y.; Kim, Y.G.; Lee, G.M. CHO cells in biotechnology for production of recombinant proteins: Current state and further potential. Appl. Microbiol. Biotechnol. 2012, 93, 917–930. [Google Scholar] [CrossRef]
  5. Umaña, P.; Bailey, J.E. A mathematical model of N-linked glycoform biosynthesis. Biotechnol. Bioeng. 1997, 55, 890–908. [Google Scholar] [CrossRef]
  6. Krambeck, F.J.; Betenbaugh, M.J. A mathematical model of N-linked glycosylation. Biotechnol. Bioeng. 2005, 92, 711–728. [Google Scholar] [CrossRef]
  7. Krambeck, F.J.; Bennun, S.V.; Narang, S.; Choi, S.; Yarema, K.J.; Betenbaugh, M.J. A mathematical model to derive N-glycan structures and cellular enzyme activities from mass spectrometric data. Glycobiology 2009, 19, 1163–1175. [Google Scholar] [CrossRef]
  8. Krambeck, F.J.; Bennun, S.V.; Andersen, M.R.; Betenbaugh, M.J. Model-based analysis of N-glycosylation in Chinese hamster ovary cells. PLoS ONE 2017, 12, e0175376. [Google Scholar] [CrossRef]
  9. Liu, G.; Marathe, D.D.; Matta, K.L.; Neelamegham, S. Systems-level modeling of cellular glycosylation reaction networks: O-linked glycan formation on natural selectin ligands. Bioinformatics 2008, 24, 2740–2747. [Google Scholar] [CrossRef]
  10. McDonald, A.G.; Tipton, K.F.; Davey, G.P. A Knowledge-Based System for Display and Prediction of O-Glycosylation Network Behaviour in Response to Enzyme Knockouts. PLoS Comput. Biol. 2016, 12, e1004844. [Google Scholar] [CrossRef]
  11. Liu, J.; Jin, C.; Cherian, R.M.; Karlsson, N.G.; Holgersson, J. O-glycan repertoires on a mucin-type reporter protein expressed in CHO cell pools transiently transfected with O-glycan core enzyme cDNAs. J. Biotechnol. 2015, 199, 77–89. [Google Scholar] [CrossRef]
  12. Cherian, R.M.; Jin, C.; Liu, J.; Karlsson, N.G.; Holgersson, J. A panel of recombinant mucins carrying a repertoire of sialylated O-glycans based on different core chains for studies of glycan binding proteins. Biomolecules 2015, 5, 1810–1831. [Google Scholar] [CrossRef]
  13. Moore, K.L.; Eaton, S.F.; Lyons, D.E.; Lichenstein, H.S.; Cummings, R.D.; McEver, R.P. The P-selectin glycoprotein ligand from human neutrophils displays sialylated, fucosylated, O-linked poly-N-acetyllactosamine. J. Biol. Chem. 1994, 269, 23318–23327. [Google Scholar] [CrossRef]
  14. Liu, J.; Qian, Y.; Holgersson, J. Removal of xenoreactive human anti-pig antibodies by absorption on recombinant mucin-containing glycoproteins carrying the Galα1,3Gal epitope. Transplantation 1997, 63, 1673–1682. [Google Scholar] [CrossRef]
  15. Löfling, J.; Holgersson, J. Core saccharide dependence of sialyl Lewis X biosynthesis. Glycoconj. J. 2009, 26, 33–40. [Google Scholar] [CrossRef]
  16. Löfling, J.; Diswall, M.; Eriksson, S.; Borén, T.; Breimer, M.E.; Holgersson, J. Studies of Lewis antigens and H. pylori adhesion in CHO cell lines engineered to express Lewis b determinants. Glycobiology 2008, 18, 494–501. [Google Scholar] [CrossRef]
  17. Mthembu, Y.H.; Jin, C.; Padra, M.; Liu, J.; Edlund, J.O.; Ma, H.; Padra, J.; Oscarson, S.; Borén, T.; Karlsson, N.G.; et al. Recombinant mucin-type proteins carrying LacdiNAc on different O-glycan core chains fail to support H. pylori binding. Mol. Omics. 2020, 16, 243–257. [Google Scholar] [CrossRef]
  18. Cherian, R.M.; Jin, C.; Liu, J.; Karlsson, N.G.; Holgersson, J. Recombinant mucin-type fusion proteins with a Galα1,3Gal substitution as Clostridium difficile toxin A inhibitors. Infect. Immun. 2016, 84, 2842–2852. [Google Scholar] [CrossRef]
  19. Liu, J.; Gustafsson, A.; Breimer, M.E.; Kussak, A.; Holgersson, J. Anti-pig antibody adsorption efficacy of α-Gal carrying recombinant P-selectin glycoprotein ligand-1/immunoglobulin chimeras increases with core 2 β1,6-N-acetylglucosaminyltransferase expression. Glycobiology 2005, 15, 571–583. [Google Scholar] [CrossRef]
  20. Olson, F.J. A MUC1 tandem repeat reporter protein produced in CHO-K1 cells has sialylated core 1 O-glycans and becomes more densely glycosylated if coexpressed with polypeptide-GalNAc-T4 transferase. Glycobiology 2005, 15, 177–191. [Google Scholar] [CrossRef]
  21. Bennett, E.P.; Mandel, U.; Clausen, H.; Gerken, T.A.; Fritz, T.A.; Tabak, L.A. Control of mucin-type O-glycosylation: A classification of the polypeptide GalNAc-transferase gene family. Glycobiology 2012, 22, 736–756. [Google Scholar] [CrossRef]
  22. Brockhausen, I.; Stanley, P. O-GalNAc Glycans. In Essentials of Glycobiology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2017; pp. 113–125. [Google Scholar]
  23. Hawkins, C.J. Golgi Apparatus: Structure, Functions and Mechanisms; Hawkins, C.J., Ed.; Cell Biology Research Progress; Nova Science Publishers: New York, NY, USA, 2011; ISBN 9781611220513. [Google Scholar]
  24. Tu, L.; Banfield, D.K. Localization of Golgi-resident glycosyltransferases. Cell. Mol. Life Sci. 2010, 67, 29–41. [Google Scholar] [CrossRef]
  25. Gill, D.J.; Clausen, H.; Bard, F. Location, location, location: New insights into O-GalNAc protein glycosylation. Trends Cell Biol. 2011, 21, 149–158. [Google Scholar] [CrossRef]
  26. Ju, T.; Cummings, R.D. Core 1 β3galactosyltransferase (C1GalT1, T-synthase) and its specific molecular chaperone cosmc (C1GalT1C1). In Handbook of Glycosyltransferases and Related Genes, 2nd ed.; Springer: Tokyo, Japan, 2014; Volume 1, pp. 149–169. ISBN 9784431542407. [Google Scholar]
  27. Togayachi, A.; Narimatsu, H. UDP-GlcNAc: Beta-gal beta1,3-NAcetylglucosaminyltransferase 6 (B3GNT6) (core 3 synthase, C3GnT). In Handbook of Glycosyltransferases and Related Genes, 2nd ed.; Springer: Tokyo, Japan, 2014; Volume 1, pp. 321–330. ISBN 9784431542407. [Google Scholar]
  28. Yeh, J.C.G.; Fukuda, M. Beta-1,3-galactosyl-O-glycosyl- glycoprotein beta-1,6-NAcetylglucosaminyltransferase 3 (GCNT3). In Handbook of Glycosyltransferases and Related Genes, 2nd ed.; Springer: Tokyo, Japan, 2014; Volume 1, pp. 367–377. ISBN 9784431542407. [Google Scholar]
  29. Tsuboi, S.; Fukuda, M. Beta-1,3-galactosyl-O-glycosyl-glycoprotein beta-1,6-NAcetylglucosaminyltransferase 1 (GCNT1) (C2GnT-L) and beta-1,3-galactosyl-oglycosyl-glycoprotein beta-1,6-NAcetylglucosaminyltransferase 3 (GCNT4) (C2GnT-T). In Handbook of Glycosyltransferases and Related Genes, 2nd ed.; Springer: Tokyo, Japan, 2014; Volume 1, pp. 355–366. ISBN 9784431542407. [Google Scholar]
  30. KEGG: Kyoto Encyclopedia of Genes and Genomes. Available online: https://www.genome.jp/kegg/ (accessed on 10 January 2020).
  31. James, M.O. Enzyme Kinetics of Conjugating Enzymes: PAPS Sulfotransferase. Methods Mol. Biol. 2014, 1113, 187–201. [Google Scholar] [CrossRef]
  32. Axelsson, M.A.B.; Karlsson, N.G.; Steel, D.M.; Ouwendijk, J.; Nilsson, T.; Hansson, G.C. Neutralization of pH in the Golgi apparatus causes redistribution of glycosyltransferases and changes in the O-glycosylation of mucins. Glycobiology 2001, 11, 633–644. [Google Scholar] [CrossRef]
  33. Kellokumpu, S. Golgi pH, ion and redox homeostasis: How much do they really matter? Front. Cell Dev. Biol. 2019, 7, 93. [Google Scholar] [CrossRef]
  34. Khoder-Agha, F.; Sosicka, P.; Escriva Conde, M.; Hassinen, A.; Glumoff, T.; Olczak, M.; Kellokumpu, S. N-acetylglucosaminyltransferases and nucleotide sugar transporters form multi-enzyme–multi-transporter assemblies in golgi membranes in vivo. Cell. Mol. Life Sci. 2019, 76, 1821–1832. [Google Scholar] [CrossRef]
  35. Akase, S. Simulation Analysis and Web Tool Development for Glycan Synthesis Prediction; Graduate School of Engineering, Soka University: Tokyo, Japan, 2020. [Google Scholar]
  36. GlycoSim. Available online: https://rings.glycoinfo.org/glycosims/index (accessed on 10 January 2020).
  37. Hoops, S.; Sahle, S.; Gauges, R.; Lee, C.; Pahle, J.; Simus, N.; Singhal, M.; Xu, L.; Mendes, P.; Kummer, U. COPASI—A COmplex PAthway SImulator. Bioinformatics 2006, 22, 3067–3074. [Google Scholar] [CrossRef]
Figure 1. Mucin-type O-glycan biosynthesis pathways. Structures are depicted using the Symbol Nomenclature for Glycans (NSFG) symbol nomenclature (yellow square = GalNAc, yellow circle = Gal, blue square = GlcNAc, red triangle = Fuc, purple diamond = sialic acid).
Figure 1. Mucin-type O-glycan biosynthesis pathways. Structures are depicted using the Symbol Nomenclature for Glycans (NSFG) symbol nomenclature (yellow square = GalNAc, yellow circle = Gal, blue square = GlcNAc, red triangle = Fuc, purple diamond = sialic acid).
Molecules 27 01766 g001
Figure 2. Glycosyltransferases included in modeling the O-glycan profiles of the 25 CHO cell lines. Dark grey represents the glycosyltransferases transiently expressed in the cells. Light grey represents the reactions of the glycosyltransferases required to produce a glycan profile. White cells represent enzymes that were neither transfected nor required for the glycan profile synthesis. C1: core 1; exC1: extended core 1; C2: core 2; C3: core 3; C4: core 4. “_a, _b” represents different reactions (reaction rules) catalyzed by a glycosyltransferase.
Figure 2. Glycosyltransferases included in modeling the O-glycan profiles of the 25 CHO cell lines. Dark grey represents the glycosyltransferases transiently expressed in the cells. Light grey represents the reactions of the glycosyltransferases required to produce a glycan profile. White cells represent enzymes that were neither transfected nor required for the glycan profile synthesis. C1: core 1; exC1: extended core 1; C2: core 2; C3: core 3; C4: core 4. “_a, _b” represents different reactions (reaction rules) catalyzed by a glycosyltransferase.
Molecules 27 01766 g002
Figure 3. Glycosyltransferase localization across the four Golgi compartments. Filled cells represent the presence of the enzymes in a compartment, while white cells represent enzymes absence. “_a/b/c/d” represent different reactions (reaction rules) for the same enzyme.
Figure 3. Glycosyltransferase localization across the four Golgi compartments. Filled cells represent the presence of the enzymes in a compartment, while white cells represent enzymes absence. “_a/b/c/d” represent different reactions (reaction rules) for the same enzyme.
Molecules 27 01766 g003
Figure 4. Simulation results of the CHO-WT cell line. (A) The simulation results in COPASI shows the glycan structures and their abundances in the last compartment “TGN”. (B) The abundance of each predicted structure at the end of the simulation compared to the concentrations measured experimentally. Structures are depicted using the Symbol Nomenclature for Glycans (NSFG) symbol nomenclature (yellow square = GalNAc, yellow circle = Gal, purple diamond = sialic acid).
Figure 4. Simulation results of the CHO-WT cell line. (A) The simulation results in COPASI shows the glycan structures and their abundances in the last compartment “TGN”. (B) The abundance of each predicted structure at the end of the simulation compared to the concentrations measured experimentally. Structures are depicted using the Symbol Nomenclature for Glycans (NSFG) symbol nomenclature (yellow square = GalNAc, yellow circle = Gal, purple diamond = sialic acid).
Molecules 27 01766 g004
Figure 5. The 20 sets of parameters estimated for the CHO-WT model. (A) Estimated concentrations of the C1GALT1, ST3GAL, and ST6GALNAC in each compartment of Golgi. (B) Mean estimated Km values for ST6GALNAC. _a/b/c represents the three reactions catalyzed by the enzymes. (C) Mean estimated Kf values for C1GALT1 and three reactions catalyzed by ST6GALNAC. Values are reported as mean with SEM.
Figure 5. The 20 sets of parameters estimated for the CHO-WT model. (A) Estimated concentrations of the C1GALT1, ST3GAL, and ST6GALNAC in each compartment of Golgi. (B) Mean estimated Km values for ST6GALNAC. _a/b/c represents the three reactions catalyzed by the enzymes. (C) Mean estimated Kf values for C1GALT1 and three reactions catalyzed by ST6GALNAC. Values are reported as mean with SEM.
Molecules 27 01766 g005
Figure 6. Predicted O-glycan profiles of four-experiment model. The abundance of predicted structures at the end of each model simulation is compared to the abundance measured experi-mentally. (A) CHO-WT model. (B) sLeX on extended Core 1 model. (C) sLeX on Core 3 model. (D) A4GlcNAc on Core1 model.
Figure 6. Predicted O-glycan profiles of four-experiment model. The abundance of predicted structures at the end of each model simulation is compared to the abundance measured experi-mentally. (A) CHO-WT model. (B) sLeX on extended Core 1 model. (C) sLeX on Core 3 model. (D) A4GlcNAc on Core1 model.
Molecules 27 01766 g006
Table 1. 25 CHO cells transfected with combinations of glycosyltransferases. CHO/CHO-WT: CHO-K1; C1: core 1; exC1: extended core 1; C2: core 2; C3: core 3; C4: core4; type 1: terminal type 1.
Table 1. 25 CHO cells transfected with combinations of glycosyltransferases. CHO/CHO-WT: CHO-K1; C1: core 1; exC1: extended core 1; C2: core 2; C3: core 3; C4: core4; type 1: terminal type 1.
Engineered CHO Cells’ NamesTransfected GlycosyltransferasesReference
CHO/CHO-WT-[11,14]
Leb on C2GCNT1, B3GalT5, FUT2, FUT4[15,16], Unpublished data
Leb on C3B3GNT6, B3GalT5, FUT2, FUT4[15,16], Unpublished data
Leb on exC1B3GNT3, B3GalT5, FUT2, FUT4[15,16], Unpublished data
Slex on C2GCNT1, FUT7[15], Unpublished data
Slex on C3B3GNT6, FUT7[15], Unpublished data
Slex on exC1B3GNT3, FUT7[15], Unpublished data
A4GlcNAc on C1A4GNTUnpublished data
LacdiNAc on C2GCNT1, B4GALNT[17]
A4GlcNAc and LacdiNAc on C2GCNT1, A4GNT, B4GALNTUnpublished data
α1,3Gal on C2GCNT1, A3GALT[18]
α1,4Gal on C2GCNT1, A4GALT[18]
ExC1B3GNT3[11]
C2GCNT1[11,19]
Terminal type 1 on C3B3GNT6, B3GalT5[12,16]
Terminal type 2 on C3B3GNT6[11,16]
Terminal type 1 on C4GCNT1, B3GNT6, B3GalT5[12,16]
C4GCNT1, B3GNT6[12]
α2,6sialic acidST6GAL1[11,12]
α2,6sialic acid with type 1 on exC1ST6GAL1, B3GNT3[11,12]
α2,6sialic acid on C2ST6GAL1, GCNT1[11,12]
α2,6sialic acid with type 1 on C3ST6GAL1, B3GNT6, B3GALT5[12,16]
α2,6sialic acid on C3ST6GAL1, B3GNT6[11,12]
α2,6sialic acid with type 1 on C4ST6GAL1, B3GNT3, GCNT1, B3GALT5[12,16]
α2,6sialic acid on C4ST6GAL1, B3GNT3, GCNT1[12]
Table 2. Glycosyltransferases used in CHO O-glycan profile model.
Table 2. Glycosyltransferases used in CHO O-glycan profile model.
AbbreviationGlycosyltransferaseEC NumberCores or Terminal Epitopes
A3GALT *Alpha-1,3-galactosyltransferase2.4.1.-αGal
A4GALT *Alpha-1,4-galactosyltransferase2.4.1.228P1 antigen
A4GNT *Alpha-1,4-N-Acetylglucosaminyltransferase2.4.1.-α1,4GlcNAc
B3GALT5 *Beta-1,3-galactosyltransferase 52.4.1.-Type 1 chain
B3GNTBeta-1,3-N-Acetylglucosaminyltransferase2.4.1.149
B3GNT3 *Beta-1,3-N-Acetylglucosaminyltransferase 32.4.1.146Extended core 1
B3GNT6 *Beta-1,3-N-Acetylglucosaminyltransferase 62.4.1.147Core 3
B4GALNT3 *Beta-1,4-N-acetylgalactosaminyltransferase 32.4.1.244LacdiNAc
B4GALTBetBeta-1,4-galactosyltransferase2.4.1.38Type 2 chain
C1GALT1Glycoprotein-N-acetylgalactosamine beta-1,3-galactosyltransferase2.4.1.122Core 1
CHST4Carbohydrate sulfotransferase 42.8.2.-6SGlcNAc
FUT2 *Fucosyltransferase 22.4.1.-α1,2Fuc
FUT3Fucosyltransferase 32.4.1.65α1,3/4Fuc
FUT4 *Fucosyltransferase 42.4.1.α1,3Fuc
FUT7 *Fucosyltransferase 72.4.1.α1,3Fuc
GCNT1 *Beta-1,6-N-acetylglucosaminyltransferase 12.4.1.102Core 2/4
GCNT3Beta-1,6-N-acetylglucosaminyltransferase 32.4.1.148Core 2/4
ST3GALBeta-galactoside alpha-2,3-sialyltransferase2.4.99.4, 2.4.99.6α2,3Sia
ST6GAL1 *Beta-galactoside alpha-2,6-sialyltransferase 12.4.99.1α2,6Sia
ST6GALNACST6 N-acetylgalactosaminide alpha-2,6-sialyltransferase2.4.99.3α2,6Sia
* Glycosyltransferases transiently expressed in one or more experiments. EC number: Enzyme Commission number.
Table 3. Donor concentrations in µM and distribution across the Golgi compartments.
Table 3. Donor concentrations in µM and distribution across the Golgi compartments.
DonorcisMedialTrans TGN
CMP_NeuAc0030003000
CMP_NeuGc0030003000
GDP_Fuc0500050000
UDP_GlcNAc9143914391430
UDP_Gal3810038103810
UDP_GalNAc0030000
PAP_S0920920920
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kouka, T.; Akase, S.; Sogabe, I.; Jin, C.; Karlsson, N.G.; Aoki-Kinoshita, K.F. Computational Modeling of O-Linked Glycan Biosynthesis in CHO Cells. Molecules 2022, 27, 1766. https://doi.org/10.3390/molecules27061766

AMA Style

Kouka T, Akase S, Sogabe I, Jin C, Karlsson NG, Aoki-Kinoshita KF. Computational Modeling of O-Linked Glycan Biosynthesis in CHO Cells. Molecules. 2022; 27(6):1766. https://doi.org/10.3390/molecules27061766

Chicago/Turabian Style

Kouka, Thukaa, Sachiko Akase, Isami Sogabe, Chunsheng Jin, Niclas G. Karlsson, and Kiyoko F. Aoki-Kinoshita. 2022. "Computational Modeling of O-Linked Glycan Biosynthesis in CHO Cells" Molecules 27, no. 6: 1766. https://doi.org/10.3390/molecules27061766

Article Metrics

Back to TopTop