Open access peer-reviewed chapter

Process Optimization for Recombinant Protein Expression in Insect Cells

Written By

Jan Zitzmann, Gundula Sprick, Tobias Weidner, Christine Schreiber and Peter Czermak

Reviewed: 13 February 2017 Published: 10 May 2017

DOI: 10.5772/67849

From the Edited Volume

New Insights into Cell Culture Technology

Edited by Sivakumar Joghi Thatha Gowder

Chapter metrics overview

5,301 Chapter Downloads

View Full Metrics

Abstract

Insect cells can be used for the efficient production of heterologous proteins. The baculovirus expression vector system (BEVS) in Spodoptera frugiperda cells and the stable transformation of Drosophila melanogaster S2 cells are widely used for this purpose. Whereas BEVS is a transient expression system for rapid protein production, stable D. melanogaster cell lines are compatible with more complex processes modes. This chapter describes the setup of both systems, including steps for the generation of expression vectors and comprehensive optimization approaches. The genetic elements available in each system are described, as well as the use of different cloning and transfection methods and advanced process monitoring to achieve robust protein expression in larger-scale bioreactors.

Keywords

  • heterologous protein expression
  • BEVS
  • stably transformed Drosophila melanogaster S2 cells
  • recombinant protein expression
  • insect cell

1. Introduction

The cultivation of recombinant insect cell lines has been the subject of intense research since the 1980s and also allows the industrial production of recombinant proteins, vaccines and insecticides [14]. Since the first insect cell line was isolated in 1963, more than 500 different cell lines have become available [5, 6]. The most common expression systems are based on cell lines derived from Spodoptera frugiperda (Sf-9 and Sf-21), Trichoplusia ni (BTI-TN-5B1-4, marketed as High Five™) and Drosophila melanogaster (S2). The S. frugiperda and T. ni cell lines are used with the baculovirus expression vector system (BEVS) [7], which is the gold standard for protein production in insect cells. More recently, stably transformed Drosophila S2 cell lines have been used to express a wide variety of proteins [8]. In all cases, it is necessary to optimize protein expression as well as the bioprocess conditions to achieve a robust and efficient upstream process. This includes the selection of suitable genetic elements, aspects of glyco-engineering, comprehensive screening for highly productive cell lines, appropriate aeration and mixing strategies and the selection of a robust process mode. Online analytical methods can be used to gain fundamental insights into the physiological state of the cells during the production process, with dielectric spectroscopy and online optical density measurements providing good results. In this chapter, we provide comprehensive, interdisciplinary guidance for the optimization of each process step and the upstream process as a whole.

Advertisement

2. Optimizing protein expression at the cellular level

2.1. Engineering the glycosylation profiles of insect cells

Many therapeutic proteins require a specific glycosylation pattern, so this aspect is a major issue when using insect cell lines for protein production. Generally, cell lines from S. frugiperda, T. ni and D. melanogaster can synthesize N-linked glycans. However, insect cells form shorter and less complex N-glycan structures than mammalian cells (Figure 1).

Figure 1.

N-glycan profiles in different expression hosts. Mammalian and insect glycans share a common precursor. Mammalian cells (left branch) elaborate this precursor using N-acetylglucosaminyltransferase (I) and other enzymes. In the middle branch, the native paucimannose glycans of insect cells are synthesized by N-acetylglucosaminidase. Insect proteins typically contain both 1,3-linked and 1,6-linked core fucose. Genetically engineered insect cell lines (right branch) can express α-1,6-mannosylglycoprotein 2-β-N-acetylglucosaminyltransferase (III), β-galactosyltrasferase 1 (VI) and β-galactoside α-2,6-sialyltransferase. The symbols were used as previously defined [25]. The figure is based on earlier publications [13, 18].

At least in some part, the alterations in the glycan patterns are of evolutionary origin. Studies revealed mutations in some enzymes of the apparatus for the protein glycosylation (e.g. transferases), occurring during the evolutionary split of vertebrates from invertebrates, 500 million years ago. This led to the future formation of distinct glycan patterns for the species [9].

In general, glyco-structures in insects thus formed in different patterns than in mammals. However, the complexity of the glycans is altered over the developmental stages in the insect [10], clearly hinting, that the glycans are involved in the development. The insect-derived cell lines do not express the respective enzymes in quantity and show therefore less complex glycosylation patterns [11].

The glycan structures of insect cells are mostly oligomannose or paucimannose forms with core fucose structures but no terminal sialic acid residues [12, 13]. In addition to human-like α1,6-linked fucose, insect-derived proteins may also carry a α1,3-linked fucose, which can induce immunogenic and allergic responses in humans [14]. About 70% of therapeutic proteins contain N-glycans and these structures can influence protein activity and tolerability, so the inability of insect cells to synthesize human-like glycosylation profiles is a disadvantage [15]. Several strategies have therefore been developed to address the issue of incomplete or incompatible N-glycosylation in insect cell expression systems.

The truncated paucimannose structures in D. melanogaster S2 cells were found to originate from an acetylglucosamidase (GlcNAcase) activity, which removes terminal N acetylglucosamine (GlcNAc) residues. The GlcNAcase was suppressed [16] and the simultaneous expression of a galactosyltransferase (GalT) resulted in more complex but still unsialylated and heterogeneous glycans [17].

The BVES can be improved by using genetically engineered host cells and/or baculoviruses (Figure 1). Cell lines have been modified to express recombinant glucosaminyltransferases, galactosyltransferases and sialyltransferases. In the presence of the sialic acid precursor acetyl-D-mannosamine, human-like glycans can be synthesized in these cells [13, 18]. In order to reduce the metabolic burden, the cells can also be equipped with inducible promoters so that the mammalian glycosylation machinery can easily be switched on and off [19]. Initial attempts to produce mammalian glycans in the BEVS required the co-infection of insect cells with baculoviruses encoding the target protein and the enzymes required for glycosylation [20]. The yields of protein were low due to the relatively small statistical likelihood of co-infection, so alternative BEVSs have been developed such as the SweetBac® technique, in which the glycan-generating enzymes are carried on a separate ‘glyco-module’. Although this simultaneous expression reduces the yield compared to traditional insect expression systems, the previously established cell line Tnao38 can achieve results comparable to transient expression in mammalian cells [21, 22]. Additionally, the synthesis of core fucose can be prevented by using a baculovirus encoding an enzyme that diverts the precursor molecule into other pathways [23].

The ability of insect cells to synthesize O-linked glycan structures has not been explored in detail. Insect O-glycosylation profiles are less diverse than those produced by mammals, but the precise nature of these structures depends on the cell line and culture medium [24].

2.2. Transient expression using BEVS

2.2.1. General procedure for recombinant protein expression using BEVS

BEVS is based on the use of insect viruses known as baculoviruses (family baculoviridae) that are rod-shaped dsDNA viruses that infect lepidopteran species. The best characterized baculovirus is Autographa californica multiple nucleopolyhedrovirus (AcMNPV). Its life cycle comprises two phases leading to different phenotypes [26]. Type 1 is known as a budded virus, which is enveloped with parts of the host cell membrane. After its release from the host cell between the early and late phases of the infection, the budded virus can spread and infect neighbouring cells. Type 2 is known as the occlusion-derived virus and is produced in the very late infection phase when the viral protein polyhedrin accumulates in the host cell and forms so-called occlusion bodies. Polyhedrin expression is controlled by the strongest viral promoter, the polh promoter. The promoters used for heterologous gene expression with BEVS are usually viral promoters, and these can be classed as early, late and very late promoters according to the timing of their activity post-infection. Early promoters are active directly after infection because they only require the host cell RNA polymerase. Early promoters drive the expression of genes encoding viral transcription factors and polymerases, which are in turn necessary for the expression of late genes. Very late genes, like polyhedrin, are even more active than the late genes, and are necessary for virus packaging. The polh promoter is the most widely used in BEVS because of its very high activity and due to the fact that occlusion body formation (and thus polyhedrin itself) is not necessary for baculovirus propagation in cell culture.

Recombinant protein expression using BEVS involves five different states, which are summarized in Figure 2 and explained in more detail in the following sections. Many different products used in (veterinary) medicine have been produced using BEVS and selected examples are listed in Table 1.

Figure 2.

The process chain for recombinant protein production using BEVS.

ApplicationProduct nameCompanyStageReferences
For human use
Cervical cancerCERVARIX®GSKApproved[27]
Prostate cancerPROVENGE®DendreonApproved[28]
InfluenzaFluBlok®Protein SciencesApproved[29, 30]
InfluenzaA/H5N1 Virus-like particleNovavaxPhase I (NCT01596725)[31]
For veterinary use
Procrine circovirus 2 (PCV2)Porcilis® PCVMerckApproved[32]
PCV2CircoFLEX®Boehringer IngelheimApproved[33]
Swine feverPorcilis Pesti®MerckApproved[34]

Table 1.

Selected human and veterinary vaccines produced using BEVS.

2.2.1.1. Cloning the transfer vector

In the first step, the gene of interest (GOI) must be integrated into a transfer vector. This can be achieved using classic insertion-ligation technology, or state-of-the-art techniques such as golden gate cloning. Depending on which kit is used to generate the recombinant baculovirus, different transfer vectors are provided by the manufacturers. The most important differences among these vectors are the promoter system, the protein tag and the secretion signal. More information about the different kits and genetic elements can be found in Section 2.2.2.

2.2.1.2. Generation of a recombinant bacmid

When the transfer vector is ready, the desired parts need to be integrated into the baculovirus genome, resulting in a recombinant bacmid. Several commercial kits are available for this step, including the Bac-to-Bac® Baculovirus Expression System (Thermo Fisher Scientific), flashBAC (Oxford Expression Technologies), the BaculoDirect™ Baculovirus Expression System (Thermo Fisher Scientific) and BaculoGold™ (BD Biosciences).

2.2.1.3. Virus production

The recombinant bacmid DNA is then transfected into insect cells (Sf9, Sf21 or HighFive™) for virus production. The cells are usually transfected in a six-well plate (2 mL culture volume) to prepare the initial virus stock (P1 stock), which typically has a titre of 103–105 plaque forming units (pfu)/mL. Transfection can be achieved using the chemical and physical methods described in the following sections.

2.2.1.3.1. Calcium phosphate–DNA co-precipitation

One of the oldest transfection methods is calcium phosphate-DNA co-precipitation [35] which was adapted for insect cell lines in the 1980s [36]. Mixing calcium chloride with a phosphate-buffered DNA-containing solution results in the formation of a fine calcium phosphate/DNA co-precipitate that binds to the cell surface and penetrates the cells by endocytosis.

2.2.1.3.2. Lipid-mediated and polymer-mediated transfection

In the late 1980s, a transfection method was developed based on the synthetic positively charged lipid N-[1-(2, 3-dioleyloxy)propyl]-N,N,N-trimethylammonium chloride (DOTMA). The cationic head groups of DOTMA interact with the anionic phosphate backbone of DNA to form a complex can bind to the cell membrane and probably taken up by endocytosis [37]. Many different lipid formulations are available to achieve highly efficient transduction, including BaculoPORTER (Biocat), Cellfectin, Cellfectin II and Lipofectin (all Thermo Fisher Scientific).

Various non-lipid transfection reagents can also form complexes with DNA, including baculoFECTIN II and flashFECTIN (Oxford Expression Technologies), FuGENE 6 (Promega), GeneJuice® (Merck Millipore) and TransIT®-Insect (Mirus Bioscience).

2.2.1.3.3. Electroporation

Electroporation is a convenient and efficient transfection method, but specialized equipment is required [38]. It is based on a short electrical pulse at an optimal voltage (specific for each cell line) to from transient pores in the plasma membrane. This in turn facilitates the intake of small molecules such as DNA, RNA or proteins [39].

2.2.1.4. Amplification

Following the generation of the P1 virus stock, the virus needs to be amplified to increase the titre and culture volume. The titre of the P2 stock is typically ≥108 pfu/mL. Insect cells are cultivated in shaker flasks with a culture volume of 50 mL. For industrial applications, virus stocks with an even higher volume are necessary, so P3 stocks are generated using bioreactors. In the authors’ laboratory, a low multiplicity of infection (MOI) of 0.01 pfu/mL and a cell density of 1 × 106 cells/mL are used for virus amplification.

2.2.1.5. Virus titre determination

The titre of infectious virus particles must be determined before the viruses can be used for protein expression. Two methods have been established for this purpose, namely the plaque assay and end-point dilution.

2.2.1.5.1. Plaque assay

The plaque assay determines the number of plaque-forming units, each of which is equivalent to a single infectious particle [40]. The infection of adherent insect cells with a highly diluted virus solution leads to plaque formation in the cell monolayer due to cell lysis caused by one infective virus particle. With an agarose overlay virus spreading is circumvented. Using neutral red, viable cells can be stained and white plaques appear when a cell has been lysed due to infection.

2.2.1.5.2. End-point dilution

End-point dilution is used to determine the infectious dose that leads to the infection of 50% of the cells, known as the 50% tissue culture infective dose (TCID50) [40]. Cultured cells are infected with different virus concentrations and the number of infected and non-infected cultures is counted. For accurate titre determinations, 12 replicates of eight virus dilutions are analysed in the authors’ laboratory. The readout can be conducted at 5 or 7 days post-infection. To facilitate evaluation, every baculovirus constructed and prepared in the authors’ laboratory carries a green fluorescent protein (GFP) marker cassette driven by the OpIE1 promoter.

2.2.1.5.3. Comparison of plaque assay and end-point dilution

When choosing a titration method, different points need to be considered. Generally, the plaque assay is performed in six-well plates resulting in a high cell concentration needed to seed the cells confluent. Due to the plate format, only a few virus dilutions can be analysed and a low number of replicates are possible using one six-well plate resulting in the need of experienced experimenters knowing the suitable virus dilutions. In comparison to that, the end-point dilution method can be done in 96-well plates resulting in more virus dilutions and replicates, which can be analysed using one plate. Moreover, the cells can be seeded sub-confluent as they are not covered with agarose, which limits the cell growth. For the titre evaluation using the end-point dilution, it is important to remember that genomic integrations of marker proteins into the viral genome simplify the evaluation.

2.2.1.6. Protein expression

The protein expression step can be divided into two stages: cell expansion and infection. In the first step, the bioreactor is inoculated with a low cell concentration, and at the desired time of infection (TOI), the baculovirus stock is added to the cell suspension.

Industrial baculovirus-based processes use a volume of 10–100 m3 [41]. Because of this size, the MOI strongly influences the virus volume required for one production cycle. Ideally the baculovirus stocks last for the whole process lifetime. A MOI > 1 statistically ensures the infection of all insect cells immediately after virus addition (see chapter 2.2.3). Using a lower MOI results in a two-stage infection process. In the first phase, only a proportion of cells are infected. These cells start to produce the recombinant protein and also produce new virus particles. In the second infection phase, these new virus particles infect further cells (Figure 3). A second effect of a low MOI is that the uninfected of the first infection phase continue to proliferate, resulting in a higher cell density at the start of the second infection phase.

Figure 3.

A comparison of BEVS processes with low and high MOI, respectively. The low MOI process is divided into two phases, whereas the high MOI process has only one phase.

2.2.1.7. Protein purification

The final step of the BEVS process is protein purification. Depending on the secretion signal (see Section 2.2.2.4), tag and characteristics such as protein size and pI, different purification methods can be used, many based on various forms of chromatography. Detailed coverage of this topic is beyond the scope of this chapter and the reader is referred to previous review articles [42, 43].

2.2.2. Historical overview and application of current BEVS kits

The BEVS patented by Max Summers and Gale Smith in 1983 (US Patent 4,745,051) used homologous recombination to integrate the GOI into a polyhedrin locus of the baculovirus genome. As stated above, polyhedrin is not essential for virus replication in cultured cells because its function is to form the occlusion bodies that protect the virus against UV light and high temperatures during the natural infection cycle. The original recombination-based method required two crossovers to integrate the GOI, which occurred at a low frequency (~0.1%). This drawback was addressed by inserting three BsuI36 restriction sites and a lacZ cassette into the baculovirus genome [44]. Digestion with Bsu36I linearized the virus DNA (called BacPAK6) and co-transfection of the linearized virus DNA and transfer plasmid was followed by homologous recombination to restore the circularity of the virus DNA, leading to the replication of recombinant viruses. BacPAK6 was optimized by deleting the chitinase gene (chiA) that inhibits the secretory pathway in insects, resulting in higher protein concentrations when the product is membrane targeted or secreted. This kit is marketed by Oxford Expression Technologies as BacPAK6-Sec+.

To enable the replication of baculovirus DNA in bacteria, a bacterial artificial chromosome (BAC) was integrated into BacPAK6 to produce a bacmid vector. This was initially marketed by Oxford Expression Technologies as the flashBAC™ system (Patent no. EP1144666). After homologous recombination with the transfer plasmid, the BAC is replaced by the GOI, the essential ORF 1629 is restored and the GOI is expressed under the control of the polh promoter. Variants of the flashBAC system are available in the flashBAC GOLD, flashBAC ULTRA and flashBAC PRIME kits. If less protease activity is required, the flashBAC GOLD system is most suitable because the v-cath gene is deleted in addition to chiA. The flashBAC ULTRA system improves the protein yield and quality by deleting the p10, p26 and p74 genes in addition to v-cath and chiA. The deletion of p10 increases the polh promoter activity, the function of p26 is unknown but deletion does not inhibit viral replication and p74 is only required for virus attachment and fusion in midgut cells in vivo so its deletion improves biosafety [45]. In contrast to the kits described above, flashBAC PRIME does not carry any genetic modifications in the viral backbone resulting in cell lysis due to infection, facilitating the release and subsequent purification of more complex products such as virus-like particles (VLPs) or proteins formed in the cytoplasm or nucleus.

BD Biosciences markets linearized baculovirus DNA kits as BaculoGold™ and BaculoGold™ Bright. The manual that comes with each kit states that an essential gene is deleted (the gene is not specified), and BaculoGold™ Bright also carries a GFP marker gene allowing the detection of infected insect cells by fluorescence analysis. Another variant is the vEHuni baculovirus DNA, which features a D. melanogaster hsp70 promoter and a multiple cloning site with two Bsu36I sites integrated into the non-essential ecdysteroid UDP-glucosyltransferase (egt) gene, allowing the expression of diverse products and the production of baculovirus expression libraries [46]. Similarly, the vECuni baculovirus DNA carries a hybrid promoter consisting of PcapminXIV and polh elements.

Thermo Fisher Scientific distributes the BaculoDirect™ kit that incorporates Gateway® cloning technology. Following the integration of the GOI into the Gateway® Entry Clone, in vitro LR recombination leads to the integration of the GOI into the BaculoDirect™ DNA, replacing the herpes simplex virus type 1 thymidine kinase negative selection marker. When insect cells are transfected with the recombinant vector and cultivated in the presence of ganciclovir, only recombinant baculoviruses are produced [47]. The BaculoDirect™ DNA can be combined with C-terminal or N-terminal V5 and His6 tags. In the latter case, the tags are followed by a tobacco etch virus (TEV) protease cleavage site. BaculoDirect™ DNA is also available with an N-terminal glutathione-S-transferase (GST) fusion tag.

Recombinant baculoviruses can also be generated using transposon activity which is marketed as the Bac-to-Bac® system (Thermo Fisher Scientific). The baculovirus DNA in this kit also contains an integrated BAC, an antibiotic resistance gene, a lacZ cassette and an attachment site for the bacterial transposon Tn7. The cloning and amplification of recombinant viral DNA is therefore carried out in Escherichia coli. The corresponding transfer plasmid consists of the GOI flanked by two mini-transposon sites (Tn7R and Tn7L). The E. coli strain provided in the kit (DH10Bac) carries a helper plasmid providing all necessary transposon system elements. Following the transformation of the DH10Bac strain, clones carrying recombinant bacmid can be identified by blue/white screening and PCR analysis (The different kits discussed (see also Table 2) above are compatible…) [205].

Kit nameMain componentIntegration siteSpecial featuresManufacturer
BacPAK6Linearized baculovirus DNApolh locusBsu36I restriction sites within ORF 603, ORF 1629 (essential gene) and lacZ cassetteOxford Expression Technologies
BacPAK6-Sec+Linearized baculovirus DNApolh locusDeletion of chitinase gene (chiA) → expression of membrane targeted and secreted proteinsOxford Expression Technologies
flashBACLinearized baculovirus DNApolh locusDeletion of chitinase gene (chiA), integration of BACOxford Expression Technologies
flashBAC GOLDLinearized baculovirus DNApolh locusAdditional deletion of cathepsin protease (v-cath) → less protease activityOxford Expression Technologies
flashBAC ULTRALinearized baculovirus DNApolh locusDeletion of chiA, v-cath, p10, p26 and p74 → improved yield and quality, expression difficult proteinsOxford Expression Technologies
flashBAC PRIMELinearized baculovirus DNApolh locusNo gene deletion in viral backboneOxford Expression Technologies
BaculoDirectLinearized baculovirus DNApolh locusNegative selection of non-recombinant baculoviruses using thymidine kinaseThermo Fisher Scientific
BaculoDirect™ C term linear DNApolh locusC-terminal V5 and His6tagsThermo Fisher Scientific
BaculoDirect™ N term linear DNApolh locusTobacco etch virus (TEV) cleavage site, N-terminal V5 and His6 tagsThermo Fisher Scientific
BaculoDirect™ GST GatewayLinearized baculovirus DNApolh locusN-terminal GST tagThermo Fisher Scientific
BaculoGoldLinearized baculovirus DNApolh locusPositive survival selection for recombinant baculovirusBD Biosciences
BD BaculoGold Bright linearized baculovirus DNApolh locusBD Biosciences
AcRP23.lacZ linearized baculovirus DNApolh locusIntact lacZ gene after integration of GOI behind polh promoterBD Biosciences
AcUW1.lacZ linearized baculovirus DNAp10 locusIntact lacZ gene after integration of GOI behind p10 promoterBD Biosciences
vEHuni baculovirus DNAegt locushsp70 promoterBD Biosciences
vECuni Baculovirus DNAegt locusPcapminXIV hybrid late/very late polyhedrin promoterBD Biosciences
Bac-to-Bac®polh locusSite-specific transposition used to integrate genes from transfer vector into bacmid DNA using E. coli DH10Bac cellsThermo Fisher Scientific

Table 2.

Commercial kits for the production of recombinant baculoviruses.

Abbreviations: polh = polyhedrin; egt = ecdysteroid UDP-glucosyltransferase.

The different kits discussed above are compatible with different transfer plasmids that are available from the same suppliers. Because most baculovirus DNA modifications are integrated at the polh locus by homologous recombination, common transfer plasmids facilitate gene expression under polh promoter control. Some transfer plasmids designed by BD Biosciences and Thermo Fisher Scientific also incorporate more than one promoter to allow multiple protein expression from the same recombinant virus. It is also possible to integrate the promoter into the transfer plasmid together with the GOI. Some of the commercially available transfer plasmids also include protein tags (Section 2.2.2.2) to facilitate protein purification and detection. If the tag is located upstream of the GOI, a protease cleavage site is integrated (Section 2.2.2.3). Secretion signals may also be integrated into the transfer plasmids (Section 2.2.2.4).

2.2.2.1. Promoters suitable for BEVS

The AcMNPV-derived immediate early (IE-1) promoter is used in several commercially available kits. The activity of this promoter is rather low but protein expression starts immediately after infection, allowing more time for post-translational modifications that are necessary for the function of many complex proteins. This promoter can also be used for transient expression because it is active in the absence of other viral factors. To enhance recombinant protein expression, the IE-1 promoter can be combined with the homologous region 5 (hr5) enhancer [48].

The late protein p6.9, also known as basic protein, core protein or VP12 [40], is essential for the production of infectious baculoviruses because it mediates viral DNA condensation and packaging [49]. The expression of enhanced GFP under the control of the p6.9 promoter could be detected as early as 6 hours post-infection, therefore also allowing more time for post-translational modifications than the very-late promoters.

The very-late p10 and polh genes produce the very-late 10 kDa protein and polyhedrin, respectively. These are highly active promoters and are used in many BEVS kits, but they depend on proteins translated in earlier phases of the infection for their activity. The activity of the polh promoter can be increased even further if p10 expression is abolished [50].

2.2.2.2. Purification tags used in BEVS

Protein tags can be used to simplify the detection and purification of recombinant proteins produced using BEVS. Common tags include His6 and GST. The His6 tag comprises six histidine residues and can be fused to either the N-terminus or C-terminus of proteins. This tag facilitates protein purification using nickel or cobalt ions or anti-histidine antibodies immobilized on a chromatography resin [51]. GST is a 26 kDa protein which is highly soluble and folds rapidly after translation. The tag is often used to increase protein solubility in prokaryotes. The affinity between this enzyme and its substrate means that immobilized glutathione can also be used for protein purification by affinity chromatography.

2.2.2.3. Common cleavage sites in BEVS

The purification tag is often removed from the recombinant protein after purification because its size and unique chemical properties can interfere with protein functions. For this purpose, a TEV protease cleavage site is often placed between the tag and the mature recombinant protein, allowing the tag to be released in vitro [52, 53]. Transfer vectors containing a TEV site are marketed by Thermo Fisher Scientific. Alternatively, transfer plasmids with thrombin cleavage sites are marketed by BD Biosciences and Oxford Expression Technologies.

2.2.2.4. Lead sequences to enhance protein secretion in BEVS

To improve the secretion of recombinant proteins produced in insect cells, secretion signals consisting of 15–30 amino acids can be fused to the N-terminus. Common secretion signals include those native to honeybee melittin (HBM) or the baculovirus envelope surface glycoprotein 67 (gp67). Transfer plasmids containing these secretion signals are available from BD Biosciences and LifeSensors. Further, signal peptides that enhance protein secretion have also been described [54].

2.2.3. Enhancing protein yields by optimizing the time and multiplicity of infection

The yields of recombinant proteins produced using BEVS can be enhanced by optimizing parameters such as the inoculum cell concentration, TOI, MOI and time of harvest. Both MOI and TOI are related to the cell concentration and therefore show significant correlation. The effect of low and high MOI on protein expression has been addressed in multiple experiments [5559] and simulations [6063]. The infection of a cell by a virus particle can be modelled using a Poisson distribution [61, 64]. The probability that a cell will absorb an infectious particle is therefore given in Equation 1, with n representing the number of absorbed baculovirus particles:

p (n,MOI)=MOIneMOIn!E1

As shown in Equation 2, the proportion of uninfected cells can be estimated when n = 0. The proportion of cells infected with at least one virus particle can then be estimated by subtracting this value from 100% (Equation 3).

Funinfectedn=0 (MOI)=eMOIE2
Finfectedwithn1 (MOI)=1eMOIE3

The simplified Poisson approach shown in Figure 4 indicates that even with an MOI of 1, approximately 37% of the cells remain uninfected. To guarantee the simultaneous infection of all cells, the MOI should be 5 (99.326% infected cells) or 10 (99.995% infected cells).

Figure 4.

a) The impact of MOI on the number of absorbed virus particles per cell, according to the Poission model. b) The proportion of cells infected with at least one virus particle at different MOI values.

As described in Section 2.2.1.6, the use of a low MOI for industrial processes has several benefits. A low MOI is easier to achieve and requires smaller volumes of virus stock, which is advantageous in large-scale cell cultures [56, 57]. Fewer virus amplification steps are required thus limiting the negative effects of passaging, such as the increasing proportion of defective viruses after each round of amplification [58]. Infection with a low MOI also results in the proliferation of cells not infected during primary infection, increasing the number of cells available for secondary infection and thus the number of cells producing the recombinant protein [58, 65]. On the other hand, the need for secondary infection also prolongs the process, but this drawback is outweighed by the advantages listed above. The final protein yield is not necessarily lower when starting with a low MOI compared to a high MOI [41, 62, 63, 66]. The early or mid-exponential phase is the optimal TOI when using a low MOI [56, 60, 61] because infection during the late exponential phase can lead to substrate limitation [67].

2.2.4. Virus purification and concentration methods

For laboratory-scale processes, centrifugation as a clarification step can produce virus stocks of sufficient purity and quantity. However, a virus concentration step is necessary for larger-scale processes, or for processes featuring a high MOI or high cell density. Very pure virus stocks are required for pharmaceutical applications such as the manufacture of vaccines [68] or the use of baculovirus vectors for in vivo gene therapy [69, 70], and it is particularly important to reduce host cell proteins and DNA to acceptable levels [71].

If the objective is to purify active virus particles with minimal loss, then the purification method must consider the stability of the virus. Baculovirus stability/activity has been tested against key parameters such as temperature, shear stress, ionic conductivity and pH. The virus is sensitive to high temperatures, i.e. it can be stored at 4°C for several months but higher temperatures (especially above 40°C) induce a rapid decline in activity. The virus is insensitive to shear forces in a peristaltic pump and is stable over the pH range 6–8. The virus is also stable within the conductivity range 8–15 mS/cm (Figure 5). However, its activity decreases rapidly when the conductivity falls below 8 mS/cm and higher conductivities up to 80 mS/cm cause an immediate 10-fold loss of infectivity rising to more than 1000-fold if the conditions are prolonged, e.g. for 24 h [72, 73]. This suggests that the virus is sensitive to osmotic pressure but shows that high-conductivity environments up to 80 mS/cm can be tolerated briefly without total loss of activity.

Figure 5.

Infectivity of baculovirus particles under conditions of differing conductivity [73].

The stability of the virus against shear forces allows it to be concentrated by tangential filtration. Polyethersulfone membranes with cut-offs in the range 100-1,000 kDa can be used to achieve a 20-fold concentration of virus particles at an average of 0.15 bar transmembrane pressure and 25 °C, but there is a considerable loss of flux due to fouling [74]. For medical applications, virus concentration alone does not meet the requirement to reduce the levels of host cell protein and DNA. These contaminants can be removed using ion exchange membranes, e.g. a polyethersulfone membrane with quaternary ammonium ligands achieved a three-fold concentration of virus particles while simultaneously reducing host cell DNA levels by nearly 90% and reducing host cell protein levels below the limit of detection [73]. Figure 6 shows the virus and protein concentrations at each step of the purification process: adsorption of the viral particle to the column, washing and elution [73].

Figure 6.

Steps in the purification of AcMNPV using membrane chromatography, showing the virus titer and host cell protein concentration [73].

Another alternative method is monolithic anion exchange chromatography, which can achieve a 50-fold virus concentration while reducing the amount of host cell protein and host cell DNA by 90 and 60%, respectively [75]. The most suitable method is therefore a matter of the required purity, process duration and scale.

2.3. Stable protein expression in D. melanogaster S2 cells

Although the BEVS platform is highly versatile and probably the most popular insect-based expression system, it is not the best choice for all products. Factors such as protein complexity, post-translational modifications and process mode must be considered during process development [70]. Recent studies show that recombinant D. melanogaster S2 cells (rS2 cells) can be used as an alternative and equally powerful expression platform [76, 77]. This system is based on an embryonic D. melanogaster cell line derived in 1972 [78]. Stable transformation of S2 cells with plasmid vectors facilitates the production of heterologous proteins. Since the 1980s, this system has been steadily refined and is now incorporated into commercial packages such as the DES® system (Thermo Fisher Scientific) and the ExpreS2 platform (ExpreS2ion Biotechnologies). The general procedure used to generate stable rS2 cell lines is the same for all packages and is summarized in Figure 7. A plasmid carrying the GOI is used to transfect S2 cells, and stable transformants that have integrated the expression construct are propagated under selection to yield a stable cell line. This line can be used for the isolation of a highly productive clone. Depending on the amount of protein required and the time available, several starting points can be used for protein production. For high-throughput screening or when small amounts of protein are sufficient, transient expression may produce enough and the selection of cell lines is unnecessary [79]. If higher protein yields are required, stable cell lines can be established [80] and even single cell cloning may be necessary [81]. Although protein production in stable cell lines usually takes longer than the BEVS platform, the rS2 system retains some flexibility.

Figure 7.

Overview of the general procedure to produce stably transformed D. melanogaster S2 cell lines for recombinant protein expression. Protein expression can be initiated at different points, starting with transient expression immediately after transfection followed by stable expression in a polyclonal cell line and finally the selection of a highly productive monoclonal cell line.

Stable rS2 cell lines are recommended for protein expression when the following aspects are important for the production process [76, 82]:

  • Both, stable rS2 cells and BEVS share the advantage of minimal risk of contamination with human viruses because most human viruses cannot replicate in insect cells. This is particularly important for the production of pharmaceutical proteins intended for administration to humans.

  • Stable rS2 cells are ideal when bacterial expression systems yield an inactive protein [83, 84] and where even BEVS is not efficient [8486].

  • Because of the non-lytic and stable nature of protein production, different bioprocessing modes such as batch or fed-batch cultures [87], chemostat cultures [88] and perfusion cultures [81] can be used. Perfusion mode in particularly achieves high protein yields [81]

    Definitions: A batch process is cell growth and protein production in a fix amount of growth medium. The fed-batch operation augments a batch processes by continuous or intermittent addition of growth medium to prevent nutrient depletion and to increase cell density and productivity. Chemostat processes involve the continuous replacement of culture medium including a withdrawal of medium with cells and product. This mode of operation is more suitable for kinetic studies rather than protein production. Perfusion is the continuous replacement of culture medium with an additional cell retention system. This mode of operation prevents nutrient depletion and cell drainage while maintaining a constant reaction volume. It allows high cell densities and product titers.

    .

  • Stable rS2 cells can grow to considerably higher cell densities than other insect cell lines (20–70 × 106 cells/mL) and are robust concerning their hydrodynamic environment due to their small size (6–10 µm).

  • No cell lysis occurs, so less host cell protein is released and the recombinant target protein is protected from proteolytic degradation. Therefore process-integrated product recovery is also conceivable.

  • Stable rS2 cells also achieve high batch-to batch reproducibility between manufacturing runs, and generate a homogeneous glycan profile.

Several rS2-derived products have already entered clinical development, confirming that rS2 cells are not only used routinely in research but also for the commercial production of high value pharmaceutical proteins (Table 3).

2.3.1. Plasmids used to generate stable cell lines for recombinant protein expression

Stable rS2 cell lines are produced by transformation with suitable plasmid vectors carrying the GOI in an expression cassette and a selectable marker. Five general strategies have been developed, as summarized in Figure 8.

Figure 8.

The different plasmid sets that can be used to generate stable rS2 cell lines.

Classically, separate expression and selection cassettes with their own promoters can be combined in a single plasmid [97, 98] (Figure 8b). Alternatively, both features can be placed in one expression cassette, separated by an internal ribosome entry site (IRES) or a 2A-like sequence (T2A), resulting in bicistronic vectors with heterologous protein production and antibiotic resistance under the control of the same (constitutive) promoter [99, 100] (Figure 8c). However, the most common approach is the use of two separate plasmids, the first containing the expression cassette and the second containing the selectable marker (Figure 8a). This exploits the ability of dipteran cell lines to recombine different plasmids in long tandem arrays [82, 101]. The transfection or co-transfection of S2 cells with such plasmids followed by selection generally leads to the integration of multiple plasmid copies into the genome [102, 82]. The co-transfection of two plasmids using the calcium phosphate precipitation method (Section 2.2.1) can result in the integration of up to 4000 copies per S2 cell [82, 102]. The ratio of integrated plasmid sequences is similar to their proportions in the transfection mixture. When using two separate plasmids, it is therefore advisable to increase the initial ratio in favor of the expression plasmid. The commercial DES® system recommends a 1:19 ratio of selection plasmid to expression plasmid, but ratios ranging from 1:1 to 1:100 have been successful [82, 103, 104]. The random integration of multiple plasmids generates heterogeneous cell populations and it is possible that rearrangements occur within the integrated array [102]. Nevertheless, this system can achieve high protein yields of 5–100 mg/L.

In contrast to the random transgene arrays described above, transposable elements can be used to insert a single copy of the GOI and selectable marker into the S2 cell genome (Figure 8d) [105, 106]. For this purpose, the expression construct and selectable marker are flanked by transposase recognition sites, i.e. Minos inverted repeats [107] or P element terminal repeats [108, 109]. Co transfection together with a helper plasmid encoding the corresponding transposase causes the GOI and marker to be inserted at a more or less random site. Transposition-mediated insertion events can occur more than once in the same genome, but the mutagenic nature of each insertion generally limits the number of integration events to between one and 10 copies [82]. This method is cumbersome because it is necessary to map the transposon insertion site and identify clones with single-copy insertions, and the low copy number limits the yield of recombinant protein. This method is more suitable when the goal is functional analysis rather than protein production, and can be very useful when combined with the technique of recombinase mediated cassette exchange (RMCE) as shown in Figure 8e. Two recent reports describe an RMCE system for D. melanogaster cell-lines based on integrase ϕC31 and its recognition sites attP and attB [108110]. In this system, a single docking cassette flanked by the first recognition site (attP) is stably integrated into the genome by transposition. A second helper plasmid is then used to transiently express the integrase. The subsequent introduction of a plasmid with an expression cassette flanked by the corresponding recognition site (attB) promotes cassette exchange. Based on one parental cell line containing the docking cassette, different comparable clones can easily be generated, which is particular helpful in comparative studies (e.g. for promoter screening).

Although the classification of different plasmid types can simplify the principles underlying the generation of rS2 cells, it is not a fixed dogma. For example, the bicistronic system can be used to express two different proteins of interest, while co-transfection with a second plasmid provides the selection cassette [100]. Furthermore, co-transfection is not restricted to two plasmids. Indeed, up to four different proteins have been expressed simultaneously by co transfecting S2 cells with multiple vectors [111]. Finally, rS2 cells can be even combined with BEVS. Although baculoviruses cannot replicate in S2 cells, infection achieves successful protein production from the non-replicating vector [112115]. Accordingly, rS2 cells appear to be a powerful and versatile tool for protein expression. Detailed protocols and more background information on the different techniques have been published [116118, 82].

2.3.1.1. Genetic elements in the expression cassette

As described above, the expression plasmid contains various features required for protein production and the best combinations must be assessed for each process.

2.3.1.1.1. Promoter systems for rS2 cells

Several constitutive and inducible promoters have been used for the production of recombinant proteins in rS2 cells (Table 4). Strong constitutive promoters are generally favored for transient expression, because the protein must be expressed immediately after transfection. Inducible promoters can be more suitable in stable rS2 cell lines, particularly if the overexpressed protein is toxic to the host cell. Inducible promoters allow the decoupling of cell proliferation and protein expression, which can be appropriate for advanced process designs (i.e. two-step processes). The most widely used promoters in rS2 cells are the constitutive actin 5C (Ac5) promoter [119] and the copper-inducible metallothionein (Mt) promoter [120]. Other constitutive promoters, such as the copia long terminal repeat (LTR) promoter, can be used for the GOI but are usually paired with the resistance marker gene. The inducible hsp70 promoter can also be used, but unlike the preferred Mt promoter it has a relatively high basal activity, and the heat shock required for induction can also induce endogenous heat-shock genes causing changes in gene expression and cell behavior that inhibit protein production [121]. Even so, the Mt promoter is usually induced with CuSO4 or CdCl2, both of which are cytotoxic at high concentrations [122] and Cd2+ also activates endogenous heat-shock promoters. Efficient protein production therefore requires a balance between promoter induction and toxicity, which varies according to the exact experimental conditions between 200 and 1000 µM for CuSO4 [95, 96, 123] and between 1 and 10 µM for CdCl2 [95, 92, 85]. The addition of other divalent ions such as Zn2+ can improve protein production [85]. To rationalize all the different factors that affect protein expression (inducer concentration, time of induction, culture medium), a structured approach such as statistically designed experiments (DoE) may yield valuable information. As a future prospect, the doxycycline-inducible TRE promoter may also be useful because it can achieve good yields in D. melanogaster rS2 cells, although it has not yet been used widely [124].

Proteins in clinical development
Recombinant placental malaria vaccinePhase I[76, 89, 90]
West Nile virus vaccinePhase I[76, 91]
HER-2 protein AutoVac™ (breast cancer)Phase II[76]
Proteins for research and process development
HIV-1 VLP and soluble HIV gp120VLP[92]
Arabidopsis thaliana sterol glycosyltransferaseEnzyme[93]
Psalmotoxin 1Small peptide toxin[83]
M2 muscarinic and glucagon receptorG-protein-coupled receptor[94]
Atlantic salmon serum C-type lectinLectin[95]
Monoclonal antibody against H5N1 influenza hemagglutininAntibody[81]
Enhanced green fluorescent protein (eGFP)Fluorescent marker protein[96]

Table 3.

A selection of proteins produced in rS2 cells.

PromoterCommentsExample plasmids (source)References
D. melanogaster actin 5C (Ac5)Strong constitutivepAc5.1/V5-His
(Thermo Fisher Scientific)
pUC-actGFP
(DGRC 1219)
[119]
D. melanogaster metallothionein (Mt)Strong inducible, induced by divalent metal ions such as Cu2+ and Cd2+pMT/Bip/V5-His
(Thermo Fisher Scientific)
pJACKS
(-)
[120]
[125]
D. melanogaster heat shock protein 70 (hsp70)Strong inducible with some basal activity, induced by heat shock (i.e. 30 min at 37°C) and also by Cd2+pHSPCat1
(-)
pHFHW
(i.e. DGRC 1121)
[126]
D. melanogaster copia LTRStrong constitutive, from copia retrotransposon; used in many selection cassettespCoBlast
(Thermo Fisher Scientific)
copia-CAT1
(-)
[127]
[126]
D. melanogaster DS47Moderate constitutivepDS47/V5-His
(Thermo Fisher Scientific)
[128]
OpIE1 or OpIE2Constitutive, derived from Orgyia pseudotsugata multiple
nucleopolyhedrovirus OpMNPV
PIZT/V5-His
(Thermo Fisher Scientific)
[48, 129, 130]
D. melanogaster adh; α1-tubulin; PGK; CMV; EF1A; UBC; CAGGPromoters with (lower) activity, not used routinely[127, 131]
SV40 early; fibroin; herpes simplex virus thymidine kinase; Rous sarcoma virus LTRThese promoters are inactive in D. melanogaster cell lines[106, 117, 127, 132]

Table 4.

Promoter systems used in D. melanogaster S2 cells.

2.3.1.1.2. Kozak sequence

The Kozak consensus sequence is required for the efficient initiation of translation. It is important to note that D. melanogaster has a different Kozak consensus sequence (cAAaATG) compared to vertebrates [133].

2.3.1.1.3. Signal peptides

Signal peptides are used to mediate protein trafficking, initiate proper folding or to ensure protein secretion to the supernatant. The signal peptides are cleaved off during or after translation. The most common signal peptide used in S2 cells is D. melanogaster BIP (homologous to the mammalian immunoglobulin heavy chain chaperone binding protein), which causes proteins to be secreted into the supernatant [134]. Signal peptides from human tissue plasminogen activator (tPa) [135] and from Galleria mellonella gloverin (GmGlv) also work in rS2 cells [136]. The proper folding of a dopamine receptor and its insertion into the cell membrane has been achieved using an influenza virus hemagglutinin signal sequence [125].

2.3.1.1.4. Fusions tags for protein detection and purification

As discussed above, fusion tags for protein detection and purification can be attached to either the C-terminus or N-terminus of a protein, with or without an additional protease cleavage site (e.g. enterokinase or thrombin). DES® plasmids usually contain a His6 tag and a V5 epitope tag [93, 96, 100, 137]. The His6 tag can be detected with an antibody, and purification can be achieved with the same antibody or by immobilized metal ion affinity chromatography (IMAC) [138, 136]. Other frequently used tags in the rS2 system include the BioEase™ tag (Thermo Fisher Scientific), the FLAG® and hemagglutinin epitope tags [139], the Myc tag [111)] and the S-tag™ [125].

2.3.1.1.5. Polyadenylation signals

As usual for eukaryotic organisms, mRNA must be polyadenylated in rS2 cells to maintain stability and support efficient protein synthesis [120, 132]. The late SV40 polyA signal (from simian virus 40) achieved the best performance, indicating that the polyadenylation mechanism is conserved between mammalian and insect cells. The early SV40 polyA signal and the polyA signals from D. melanogaster metallothionein (Mt) and alcohol dehydrogenase (adh) were also functional in rS2 cells.

2.3.1.2. Antibiotic selection markers

Several selection systems have been tested in rS2 cells as summarized in Table 5. Each system comprises a cytotoxic agent and a corresponding marker that confers resistance. The marker may encode an enzyme that catalyses the transformation of a selective agent into a harmless product, e.g. blasticidin and hygromycin resistance. Alternatively, the marker may encode a mutated enzyme that replaces an endogenous enzyme inhibited by the cytotoxic agent, e.g. α-amanitin and methotrexate resistance. Only cells that have integrated the selection cassette can survive and proliferate in the presence of the selective agent. The most appropriate selection system depends on time, cost and risk. For example, α-amanitin is fast and efficient but expensive and highly toxic to humans, making it less suitable for large-scale processes and pharmaceutical products. Blasticidin and puromycin are more expensive than hygromycin but also work faster. A high rate of spontaneous resistance has been reported for G418. It is always advisable to generate a kill curve in order to determine the 50% lethal dose (LD50) as a starting point for the optimization of the selection protocol, because factors such as the presence of fetal bovine serum (FBS) and the overall medium composition can influence the potency of these selective agents.

Selective agent (resistance marker)Working concentration of the antibioticPlasmids (source)References
Blasticidin S*
(blasticidin-S deaminase - bsd)
25 µg/mL, range: 5–100 µg/mLpCoBlasta
(Thermo Fisher Scientific)
[140]
Hygromycin B*
(hygromycin-B-phosphotransferase -hph)
300 µg/mL, range: 100–1000 µg/mLpCoHygroa
(Thermo Fisher Scientific)
pUC-HygroMTb
(DGRC 1059)
[103]
[141]
Puromycin
(puromycin N-acetyltransferase)
15–30 µg/mL, range: 2–30 µg/mLpCoPuroa
(RDB 08531)
pMT-PUROb
(RDB 08532)
[142]
[97]
Methotrexate
(resistant dihydrofolate reductase DHFR)
0.1 µg/ml, range 0.1–4 µg/mLp8HCOa
(DGRC 1003)
pHGCOa
(-)
[143]
[144]
α-Amanitin
(mutated RNA polymerase II)
5–10 µg/mLpPC4d
(DGRC 1217)
[145]
Geneticin (G418)
(neo-aminoglycoside phosphotransferase type II)
1000 µg/mlpUChsneoad
(-)
pAc5-STABLE1-Neoc
(Addgene 32425)
[106]
[99]
Zeocin™
(Streptoalloteichus hindustanus ble)
75 µg/mLPIZT/V5-Hisb
(Thermo Fisher Scientific)
[129]

Table 5.

Selection systems used for the development of rS2 cell.

co-transfection,


transfection with a single plasmid containing selection and expression cassettes,


transfection with a single bicistronic plasmid,


P-mediated transformation.


Commonly used selective agents are indicated with an asterisk. Plasmids are available from the Drosophila Genomics Resource Center (DGRC), Riken Bioresource Center DNA Bank (RDB), Addgene and Thermo Fisher Scientific. Lower case letters indicate delivery methods:

2.3.1.3. Transfection methods for S2 cell lines

Several strategies for the transfection of insect cells were described in Section 2.2.1.3. All of these methods are appropriate for the production of stable rS2 cell lines, but it is important to consider that transfection causes stress to the cells and a recovery period may be necessary before selection commences.

2.3.1.3.1. Calcium phosphate–DNA co-precipitation

As described above for Sf cells, calcium phosphate-DNA co-precipitation was also one of the first methods used to generate rS2 cells [117] and was used extensively in the past [83]. However, the limitations of this technique include the need of a fixed amount of ~20 µg DNA per mL precipitate [38]. Calcium phosphate-DNA co-precipitation has therefore been largely replaced by more flexible techniques that achieve greater reproducibility [102].

2.3.1.3.2. Electroporation

Electroporation is a convenient method for the transfection of S2 cells [38] that can achieve efficiencies comparable to calcium phosphate–DNA co-precipitation [102]. This method also allows the uptake of DNA over a very large concentration range, making it useful for the transposon-mediated generation of cells with single-copy inserts [102].

2.3.1.3.3. Lipid-based systems

Several cationic lipid reagents achieve the efficient transfection of S2 cells, including the Thermo Fisher Scientific products Cellfectin [96, 146], Cellfectin II [147] and Lipofectin [148], as well as DOTAP-Liposomes [149] and DDAB [150].

2.3.1.3.4. Other transfection reagents

Specialized proprietary formulations for S2 cells have recently been introduced, including the TransIT®-Insect transfection reagent (Mirus Bioscience), which is used routinely in the authors’ laboratory. Other non-lipid transfection reagents have been used successfully with rS2 cells including FuGene 6 (Roche Molecular Biochemical) [38], Effectene (Qiagen) [99] JetPEI (PolyPlus-transfection) [151] and DEAE dextran [117].

2.3.1.4. Case study for the development of a polyclonal rS2 cell line

To illustrate the generation of a stable rS2 cell line, this section describes the expression of the reporter protein GFP and its time resolved detection during cell line establishment. TransIT®-Insect-based transfection was carried out using a single plasmid with an expression cassette containing intracellular GFP under the control of the Ac5 promoter and a selection cassette containing hygromycin phosphotransferase driven by the copia LTR promoter. After transfection, the cells were allowed to recover for 3 days before 300 µg/mL hygromycin was added to the medium and renewed at each sub-culture interval. Figure 9 shows the GFP expression profile monitored by flow cytometry. Within the first 10 days, most of the non-transfected cells died and only low levels of GFP were detected. After 20 days under selection, the cell population became increasingly GFP-positive, and a population producing high levels of GFP was established within 30 days. Further sub-culturing revealed the existence of different sub-populations with varying GFP expression profiles and different growth properties. By day 77, a population characterized by moderate levels of GFP became prevalent, suggesting that polyclonality is detrimental for long-term protein production. This can be overcome by single cell cloning, as discussed in the next section.

Figure 9.

Time course showing the establishment of a stable rS2 cell line expressing GFP under the control of the constitutive Ac5 promoter. GFP activity was recorded by flow cytometry. The transfected cell population maintained under selection pressure with 300 µg/mL hygromycin (green) was compared to a non-transfected control (gray).

2.3.2. Single cell cloning for enhanced protein yields

Once a stable cell line is established, it can be used as a straightforward basis for subsequent up-scaling and protein expression even at the bioreactor scale. However, the expression profile within the polyclonal cell population is heterogeneous. Furthermore, high-copy-number transgenic loci confer an additional metabolic burden that may inhibit cell growth. Long-term subculturing therefore enriches subpopulations that have lost copies of the transgene, and protein expression declines [152] as illustrated with eGFP in Figure 9. As well as maintaining the cells under selection pressure, single cell cloning is necessary to minimize these effects and should commence before the highly-productive cells become overpopulated by their less-productive peers [118, 152]. However, the productive cells should only be chosen once they have recovered from the stress of transfection because the success of a single cell cloning is highly dependent on clones that proliferate well.

Another important consideration is that rS2 cells grow very slowly at low densities and may arrest completely if seeded at less than 5 × 105 cells/mL [77, 116]. This reflects the demand for autocrine growth factors, which accumulate to sufficient levels only at high cell densities [77]. For example, the adenosine deaminase-related growth factor (ADGF) family is known to promote the growth of S2 cells [153, 154] but no studies have yet shown that ADGF alone can stimulate growth at very low cell densities. Augmenting the culture with conditioned medium or heat inactivated FBS can improve proliferation, but is also not sufficient to stimulate propagation of single cells. Feeder cells are therefore required to facilitate the proliferation of a single rS2 cell. There are two traditional cloning methods: cloning in soft agar and cloning by dilution. Both make use of non-transfected feeder cells, which are exposed to a γ or X ray source (e.g. 24 kR per 50 cm? T-flask) before co-culture [82]. Irradiated feeder cells do not divide, but remain able to secrete growth-promoting substances and provide an additional source of nutrients when they die [116]. Cloning in soft agar requires the mixing of concentrated feeder cells (1–2 × 106 cells/mL) with much more dilute selected transformants (~25 cells/mL) in soft agar, which is then poured into a Petri dish. The cells grow in this semi-solid support and form colonies in the agar within 2 weeks. Cloning by dilution requires the mixing of feeder cells and transformants in such a way that approximately one clone per well can be seeded in a microtiter plate. The corresponding protocols have been described in detail [82, 38, 117, 118]. Both methods have been used successfully with rS2 cells for a long time, but their major drawback is the need for X-ray or γ sources which are not readily available routinely in cell culture laboratories. It is also necessary to ensure that the feeder cells are permanently unable to divide yet still survive at least 1 week post-irradiation to condition the medium during the early growth of the clones [117]. Robustness against ionizing radiation is cell line-dependent, so parameters such as radiation dose and distance from source must be empirically standardized to achieve the requirements described above. As an alternative to irradiation, feeder cells can be treated with mitomycin C to chemically block mitosis, but this method must fulfil the same requirements and empirical testing is still necessary [117].

A modified version of the limiting dilution protocol was recently reported that does not use radiation and therefore simplifies the cloning workflow [81, 95]. The method is based on the co-cultivation of single transformants with living, non-transfected feeder cells followed by antibiotic selection of the clones. Because the procedure is not yet well established, the steps in the protocol are summarized below:

  • Seeding of approximately one transformant per well in 100 µL medium, containing 5 × 105 non-transfected, living feeder cells/mL.

  • Co-cultivation for 1–3 days allows all cells to proliferate and ensures proper conditioning of the medium.

  • Adding the antibiotic (e.g. 15–25 µg/mL blasticidin or 100–1000 µg/mL hygromycin) to initiate the selection phase. Feeder cells will slowly decay, while colonies from antibiotic resistant clones will expand during the next 2 weeks. If necessary, medium can be added to renew the antibiotic and avoid desiccation.

  • Checking colony growth using a microscope: wells with multiple colonies should be discarded because they are not monoclonal (Figure 10).

  • Picking colonies and successively expanding them in 48-, 24-, 12- and 6-well plates, followed by cultivation in T-flasks or shake flasks until cryopreservation in freezing medium containing 50% spent medium and 7.5–10% dimethylsulfoxide (DSMO).

Figure 10.

Phase contrast image of wells used for single cell cloning. Multiple colonies (left) and a single cell colony (right) on a decaying layer of feeder cells.

For all methods, a negative control that only contains feeder cells is advisable. Feeder cells that have been irradiated, blocked with mitomycin C or treated with other antibiotics should show no evidence of proliferation after 2 weeks, otherwise the inactivation will not be successful and contamination of the single cell pool with less productive cells will remain possible (The method is based on the co-cultivation of single transformants with living, non-transfected feeder cells followed by antibiotic selection of the clones (Figure 11)).

Figure 11.

Time course for the establishment of a monoclonal population using the limiting dilution method and co-cultivation with non-transfected feeder cells under selection with 15 µg/mL blasticidin S.

To avoid extensive work during scale up, producer screening should be started soon after picking the single clones. The screening method depends on the expression strategy. If the recombinant protein is expressed constitutively then the cell lines can be screened directly, whereas inducible cell lines must first be sub-cultured in a medium containing the inducer. The location of the target protein should also be considered, i.e. whether the protein accumulates in the cell or is secreted to the medium. Methods to screen for specific proteins include flow cytometry, fluorescence microscopy, western blot analysis, the enzyme linked immunosorbent assay (ELISA) and specific functional assays, e.g. enzyme assays. For example, Figure 12 shows the comparison of an eGFP expressing parental polyclonal cell line and a highly productive monoclonal cell line by fluorescence microscopy. Note that the polyclonal cell line only contains a few highly productive cells, whereas the GFP fluorescence in the monoclonal cell line is much higher because all the cells produce large amount of the product. This indicates that a highly-productive cell line has been selected successfully. Although some authors report that single cell cloning does not always achieve enhanced protein production [116], others claim that the considerable effort is worthwhile [81, 95, 118, 155]. It is true that most clones do not show enhanced protein expression, but this is unsurprising because the frequency of highly-productive cells in the parental cell pool is usually low (Figure 12). Consequently, the likelihood of selecting a highly-productive cell reflects this initial ratio, and at least 500–1000 wells (i.e. 5–10 96-well plates) should be used for cloning and subsequent screening.

Figure 12.

Fluorescence microscopy images of a parental polyclonal cell line (left) and a highly productive monoclonal rS2 cell line (right), expressing an eGFP fusion protein under the control of the Mt promoter (induced with 900 µM CuSO4 at 1×106 cells/mL with viability > 97%). The images were captured using the same instrument settings 24 h post-induction.

Advertisement

3. Scale up of insect cell cultivation processes

3.1. Assessment of the cost-effectiveness

In order to produce recombinant proteins cost effectively, a satisfactory expression level has to be achieved in one of several species available for recombinant protein expression. Suitable hosts include bacteria (Escherichia coli), yeasts (such as Pichia pastoris) and cell lines of mammalian or insect origin. These expression systems differ in terms of complexity, space-time-yield and the ability to support protein folding and posttranslational modification [156]. The system of choice depends on the properties of the protein that is to be produced and insect cell lines are usually employed for the production of virus like particles and proteins that require folding and posttranslational modification [1, 156]. Generally the competitiveness of insect cell lines is demonstrated by the availability of different high value commercial products, such as the vaccines FluBlok®, CERVARIX® and PROVENGE® (se chapters 2.2.1 and 2.3), which are produced using baculovirus based systems. Furthermore a recent study showed that also stable rS2 cells can compete with established systems. Concretely rS2 cells were more suitable for the production of human coagulation factor IX than CHO cells [157] and also suitable to produce high titers of a monoclonal antibody [81].

3.2. General considerations on process design

Insect and mammalian cell lines both originate from tissues of multi-cellular organisms and therefore have comparable growth requirements, but insect cells offer some advantages in terms of process design [1]. They grow rapidly at lower temperatures (doubling time of approximately 24 h at 22–28°C) and tolerate higher levels of free amino acids and glucose without switching to overflow metabolism [158160]. This allows the use of rich media, incorporating the nutrients for complete batch processes. Although first-generation media (e.g. Grace insect medium, TMH-FH and TC100) required complex additives such as FBS or insect hemolymph [160], optimized protein-free media are now available, including Sf-900TM II SFM (Thermo Fisher Scientific), ExCell® 420 (Sigma Aldrich) and Insect-XPressTM (Lonza). Defined media offer better lot-to-lot reproducibility, as well as simplified qualification and validation, and even contribute to higher protein titers. The optimal pH for insect cell lines is slightly acidic (6.2–6.9) and is usually maintained by a phosphate buffer. Therefore no CO2 supply is required, unlike mammalian cell culture media which rely on the open bicarbonate buffer system. For industrial scale up, insect cells can be adapted to grow in suspension cultures, allowing the use of standard bioreactors resembling those typical for mammalian cell culture [158]. Beyond research, where more diverse culture devices are used [3, 161], stirred tank reactors (STRs) [87, 161] or wave bag reactors (e.g. GE Wave, AppliFlex or CultiBag) [81, 162, 163] are more suitable for the large-scale cultivation of insect cells. Both systems are scalable and are well established in the industry. Short reactor set-up times are important in particular for high-turnover baculovirus-based processes, which last 4–7 days. These processes are usually carried out in batch or fed-batch mode, because the virus-mediated lysis of the cells imposes a time limit on each production cycle. In contrast, continuous or perfusion mode is also compatible with stable rS2 cell lines. In each case, the main task of the bioreactor is to provide optimal growth and production conditions. The main challenge is therefore to ensure an adequate oxygen supply without generating destructive shear forces [164, 165].

3.3. Oxygen requirements of Sf and S2 cell lines

The essential nutrient oxygen must be delivered continuously because it is only sparingly soluble in cell culture medium. During scale up, it is not possible to achieve the high ratio of surface area to reaction volume, and in turn the high oxygen mass transfer through the headspace, which are characteristics of small-scale cell culture systems. It is therefore necessary to understand the oxygen requirements of the cultured cell line in detail in order to choose an appropriate oxygenation strategy. The cell-specific oxygen consumption rate q˙O2 is a key parameter for the physiological state of the cells, which provides useful information for scale up. It can be estimated by placing cells from a growing culture in a tightly sealed measurement chamber containing air-saturated medium. Given that oxygen transfer is negligible under these conditions, the saturation declines and a corresponding oxygen time course can be recorded (Figure 13). From the known cell concentration X and the slope of the curve ΔcO2Δt, the specific oxygen consumption rate q˙O2can be calculated according to Eq. (4). This method can also be adapted to determine q˙O2 directly in a bioreactor. In order to do this, the oxygen supply must be closed and a sample must be withdrawn to determine the current cell concentration.

Figure 13.

Time courses showing the declining oxygen saturation in a measurement chamber containing different concentrations of Sf 21 suspension cells in ExCell420 serum-free insect cell medium.

q˙O2=1XΔcO2ΔtE4

Table 6 summarizes specific oxygen consumption rates for the cell lines discussed in this chapter. The data show that, q˙O2does not remain constant but is strongly dependent on the physiological state of the cells. For example, the infection of Sf cells with baculovirus increases the oxygen demand [166]. Furthermore, a wide range of q˙O2 values has been reported for different stable rS2 cell lines, indicating that the expressed recombinant protein also affects the oxygen demand. In conclusion, q˙O2 seems to be a function of cell status, overall medium composition and the reactor set up. Therefore, it is advisable to verify the values reported in the literature in each new experimental setting.

Cell lineq˙ [10−12 mmol∙Z−1∙min−1]Growth phaseCultivation systemInfectedMediumFCS %References
Sf 92.58exp.ChambernoIPL-4110[167]
Sf 93.66-STRyesIPL-4110[168]
Sf 96.50-STRyesICSF-WB0[168]
Sf 91.20–3.00exp.STRnoSf-900 II0[169]
Sf 91.50–2.70-STRyesSf-900 II0[169]
Sf 95.50exp.STRnoExcell 4010[170]
Sf 910.0-STRyesExcell 4010[170]
Sf 93.13–3.35exp.AirliftnoIPL-4110[171]
Sf 91.33stat.AirliftnoIPL-4110[171]
Sf 94.44exp.STRnoSf-900II[172]
Sf 2110.5exp.ChambernoExcell 4200*
Sf 212.33exp.PerfusionnoIPL-415[171]
S2 wt1.5exp.STR-ExCell 4200*
S2 wt0.45exp.STR-Sf-900 II0[172]
S2AcGPV0.82exp.STR-Sf-900 II0[172]
S2AcGPV0.6–1.2exp.Spinner-IPL 41[173]
S2MtEGFP0.8–1.5exp.STR-Sf-900 II0[172]

Table 6.

Specific oxygen consumption rates for different insect cell lines and cultivation conditions.

Authors’ own data. Abbreviations: exp. = exponential growth phase; stat. = stationary phase.


Knowing q˙O2 and the current cell concentration X, the oxygen uptake rate (OUR) of the culture can be calculated according to Eq. (5):

OUR=q˙O2X E5

Similarly, the ability of a bioreactor to supply oxygen is characterized by its oxygen transfer rate (OTR), which is calculated according to Eq. (6):

OTR=kLa(cO2* cO2)E6

The volumetric mass-transfer coefficient kLa is a reactor-specific constant that describes the efficiency with which oxygen is transported to the medium under a given set of operating parameters [174]. The values cO2* and  cO2 are the maximum and actual oxygen concentration in the medium, respectively. To ensure a sufficient oxygen supply, the OTR must be equivalent to the OUR at the maximum cell density Xmax and the required kLa value of the cultivation system can thus be calculated according to Eq. (7):

kLa required=q˙O2XmaxcO2* cO2setE7

Eq. (7) also reveals why oxygen supply is a major concern for both expression systems. Whereas Sf cells show high specific oxygen consumption rates, S2 cells consume less oxygen per cell, but grow to higher densities. Both situations lead to a considerable oxygen demand. As a rule of thumb, the oxygen saturation in the medium  cO2set should not fall below 30% air saturation during cultivation [175, 158]. In order to avoid limitations at late process stages, the rate of oxygen transfer can be increased either by using highly efficient micro-porous spargers [165] or by using pure oxygen for aeration. Even with pure oxygen, sufficient gas flow must be maintained in order to strip out CO2 that inhibits cell growth at higher concentrations. Increasing the stirrer speed and gas flow are a second option for STRs, but caution should be exercised because insect cells (especially infected and swollen Sf cells) are susceptible to shear damage. To encounter the danger of shear related cell damage, insect media usually contain up to 0.1% of the non-ionic block co-polymer Pluronic F68 as a shear force protecting agent [176]. Pluronic F68 adheres to the cell surface and thus stabilizes the membrane. A detailed discussion on the assessment of stirring and bubble related shear damage with its consequences for animal cell culture process design has been published [164].

3.4. Characterization of STR oxygenation capabilities for insect cell culture

The kLa value for each reactor setup must be determined and compared with the calculated value required to achieve sufficient oxygen transfer. The kLa can easily be determined using standard methods such as the dynamic method or the sulphate method [177]. In combination with a structured experimental design, appropriate settings for different cell lines are easy to find. One way to structure the experiments is to use the response surface method (RSM). This statistical method explores the relationship between a response variable (i.e. the kLa value) and different input variables (e.g. stirrer speed and aeration rate). Based on a set of designed experiments, the RSM approximates the coherences between the variables using a polynomial model [178]. The resulting model can then be used to predict the kLa value for each factor combination so that further experiments are not required. Once a suitable model is established and verified, optimal settings for new reactor set-ups can be deduced from a simple readout.

As an example, a respective model was determined for a water-filled 2-L bioreactor system (working volume 1000 mL) equipped with a drilled pipe ring sparger and a pitched blade impeller (3 × 45°, d/D = 0.57). According to the typical cultivation conditions for insect cells, the temperature was set to 28°C. For RSM, the kLa values were determined by the dynamic method using a central composite design (CCD) for aeration rate v˙(0.010.09) and stirrer speed n (80260). The resulting data were used to fit the significant terms of the general interaction model as shown in Eq. (8):

kLa= β0+β1v˙+β2n+β12v˙nE8

The results can be visualized as a contour plot (Figure 14) that describes the kLa value as a function of both input variables and thus provides a straightforward way to define the most suitable bioreactor settings. However, information about homogenization is also important, so it is advisable to use the same experimental design to establish models for mixing time. Combining both models provides a comprehensive overview of the potential operating window for each process. The workflow described in this chapter is not limited to STRs but can also be adapted for wave reactors easily, where kLa and mixing time are functions of the shaking frequency and angle.

Figure 14.

Relationship between kLa, aeration rate and stirrer speed for a 2-L autoclaveable bioreactor filled with water and operated under standard conditions for the cultivation of insect cells (28°C, 0.01–0.09 vvm, 80–260 rpm).

Advertisement

4. Process monitoring

4.1. Process analytical technology as driving force for online monitoring

Complex recombinant proteins and vaccines are high-value products that are often intended for medical use. Such products must comply with high quality standards and recently the US Food and Drug Administration (FDA) released a guideline on process analytical technology (PAT), thus encouraging manufacturers to ensure product quality by improving their process understanding [179, 180]. This requires the measurement and control of critical process parameters, such as cell growth in the case of insect cells producing recombinant proteins. The dependency between cell growth and recombinant protein production exists because both processes require efficient protein synthesis machinery [181]. Optimal cell viability and comparable growth conditions during cultivation will therefore improve batch-to-batch reproducibility. Table 7 lists online techniques that have been developed to estimate cell biomass, among which optical probing and dielectric spectroscopy have proven to be robust and easy to use. The following section therefore summarizes the principles of these two techniques and their application in insect cell cultures.

MethodMeasured bioprocess variableDirect or indirect biomass quanti- ficationRepresentative commercial systemsReferences
Dielectric spectroscopyPermittivity and conductivityDirect viable cell densityAber Futura
Hamilton Incyte
[182]
Optical probingTurbidity or backscattered lightDirect whole cell density and particlesExner ExCell 230

Hamilton Dencyte
Optek ASD
Mettler InPro8100
Finesse TrueCell
Cerexinc
Wedgewood BT65
[183]
2D Fluorescence spectroscopyCellular fluorophores (NAD(P)H, flavins and aromatic amino acids)Indirect from metabolic activityDelta Light and Optics BioView[184]
BiocalorimetryHeat production from metabolic activityIndirect from metabolic activityMettler Toledo eRC1[185]
Off gas analysisRespiratory activity (OUR and CER)Indirect from respiratory activityBlue Sens CellinOne[186]
Soft sensorsDifferent process parametersIndirect from correlation/models[187]

Table 7.

Summary of online biomass monitoring techniques.

4.2. Dielectric spectroscopy to determine viable cell density

Dielectric or impedance spectroscopy is a noninvasive technique that is widely used to characterize materials in different research fields, including material testing, corrosion research and biological engineering [188, 189]. Historically, the application of dielectric spectroscopy for the characterization of cell suspension cultures dates back more than 150 years [190] and is now well established as a routine method [191, 192] even in industrial-scale processes [193]. Comprehensive reviews of the theory and application of this method have been published [194199, 200].

The method involves exposing cell suspensions to an alternating current of low magnitude. This stimulus results in a phase-shifted, frequency-dependent voltage response, which in turn is recorded and used to calculate the dielectric properties of the material (Figure 15a). The measurement is not that simple and modern devices make use of special bridge circuits, network analysers and other advanced methods to calculate the real and imaginary part of the complex impedance Z or admittance Y. Nevertheless, these measurements yield the two passive electrical properties of capacitance C and conductance G as function of the applied alternating current frequency. The capacitance C describes the ability of the material to store electrical charge, whereas the conductance G describes its ability to pass an electrical charge. Because these values depend on the electrode geometry, the relative permittivity ε and the conductivity κ are generally used to describe these electrical properties in an electrode-independent manner. The relationships are described in Eqs. (9) and (10), where C is the capacitance, ε is the relative permittivity, ε0 is the permittivity of the vacuum (8.854 pF/m), G is the conductance, κ is the conductivity and z is the specific electrode constant (i.e. area/plate distance for a parallel plate capacitor):

Figure 15.

(a) Schematic illustration of a capacitive sensor: only cells with intact plasma membranes are polarized, whereas cell debris, gas bubbles and non-polarizable solid particles do not influence the measurement. (b) Schematic illustration of the β-dispersion spectrum of the observed permittivity: at low frequencies, the duration of the shift in the excitation field is sufficient to fully polarize the cells, whereas at high frequencies the rapid shift allows little polarization. Increasing the cell number increases the number of polarizable cell membranes and consequently leads to an overall increase in Δε. (c) Theoretical comparison of three cell suspension cultures with equal biovolumes and conductivities, but composed of cells with different cell sizes. With decreasing cell size, the critical frequency fc increases. (d) Theoretical comparison of two cell suspension cultures with equal biovolumes and average cell size, but different cell size distributions. A wider distribution leads to a flattened β-distribution. Panels (b) and (c) are modified from Ref. [202].

C=εε0zE9
G=κzE10

To simplify the explanation of the observed dielectric phenomena when ε is plotted against the applied frequency f, the cell suspension can be modelled as being composed of two parts [196]. The first part is the conducting aqueous cell culture medium that surrounds the second part (the cells), which are in turn composed of an insulating cell membrane and the conducting cytoplasm. Whereas the medium and cytoplasm are simple electrolytes with a certain resistance, the lipid cell membranes act as dielectric barriers and cells can therefore be regarded as small spherical capacitors. Electrically charged ions accumulate at the membranes of living cells in an alternating electric field, whereas leaking cells, cell debris, evolved gas bubbles, micro-carriers and other media components are essentially invisible to this method [195]. Only very high volume fractions of non-biomass materials close to the sensor may influence the measurement as the cells are replaced by non-reactive materials [198]. The build-up of electrical charge across the membranes is known as polarization and this occurs when the frequency of the excitation field is in the range 0.1–10 MHz (the radio frequency band). At low frequencies, the ions have sufficient time to reach the cell membrane, complete polarization takes place and a high permittivity signal is produced. At high frequencies, ions do not have enough time to move and they do not accumulate at the membranes before the electric field changes direction thus forcing them in the opposite direction. In this case, only a low permittivity is detected. This behaviour results in a sigmoid shaped εlog(f) relationship, the so called β-dispersion (Figure 15b). The plateau value of the permittivity εmax depends on the quantity of polarizable cell membranes in the system. Hence, Δε, the difference between εmax and the residual permittivity ε (resulting from the medium), is a good measure of cell density. In 1957, Schwan derived Equation 11 predicts Δε for suspensions of ideal spherical cells [199]. The validity of the equation has been confirmed experimentally [189, 193].

Δε=9PrCm4ε0withP=4πr33NE11

Here, P denotes the cell volume fraction (dimensionless), N denotes the cell density per unit volume, r is the cell radius and Cm is the capacitance per membrane area. It is clear that Δε is actually measuring the cell volume fraction instead of cell density, because it depends on the cell radius. However, if the radius does not change throughout cultivation, Δε is directly proportional to the viable cell density. Technically, the value of Δε can be determined either by scanning the whole frequency range or by measuring at two distinct frequencies. Recording a full β-dispersion spectrum is advantageous because it provides additional information. First, the so-called critical frequency fc (Eq. (12)) represents the frequency at which one half of the potential polarization is achieved. It reflects changes in the cell radius as well as the conductivity of the cytoplasm σc and the surrounding medium σm (Figure 15c).

fc=12πrCm(1σc+12σm)E12

Second, another variable known as the Cole-Cole α factor can be obtained by fitting the Cole-Cole model to the measured spectrum [197, 201]. The α factor has a value between 0 and 1 and describes the steepness of the characteristic drop in β-dispersion. The α factor is assumed to reflect the homogeneity of the cell population [194, 201] as shown in Figure 15d.

Dielectric spectroscopy has proven to be a valuable tool for the online monitoring of insect cell cultures [195, 196, 203]. It is particularly useful for BEVS because the arrest of cell growth after infection and the swelling and lysis of infected cells can be monitored in real-time. In the context of insect cell cultures, dielectric spectroscopy was first used to monitor the growth of uninfected and infected Sf9 cells [204]. This study revealed a linear correlation between relative permittivity and viable cell density during the growth phase. However synchronous infection with a recombinant baculovirus encoding β-galactosidase at a MOI of 10 resulted in growth arrest, but permittivity increased further, indicating the successful detection of the infection-related cell swelling. During the late process stages, cell lysis associated with a decrease in cell viability and size was detected as a drop in the permittivity signal. In terms of process intensification by the optimization of feeding and infection strategies, the same authors reported the use of dielectric spectroscopy to monitor Sf9 and High FiveTM cells [205, 206]. In both studies, physiological parameters correlated with the impedance signal. Interestingly a peak in the CO2 evolution rate (CER) observed at high MOI correlated with a temporary plateau in the permittivity signal. This was interpreted to represent the initial release of virus particles into the medium. In contrast, no permittivity plateau was detected for infections with a lower MOI (0.001). This behavior can be attributed to the fact that a lower MOI does not cause the simultaneous infection of all cells, hence there is no steep response in the cellular events caused by the infection, which are instead distributed throughout the cultivation process resulting in the delayed CER response (205). Dielectric spectroscopy has also been used for the optimization of a baculovirus-based production process for recombinant adeno-associated virus (rAAV) vectors at the 40-L scale [207]. The permittivity signal was used to pinpoint the optimal time of infection. An optimal time of harvest was also determined, because an increase in cell diameter was correlated to the yield of rAAV. Well-established off line methods (Vi-Cell® and CASY®) have been used to correlate the permittivity signal with the viable and total cell culture volume [208]. The authors observed an increase in the critical frequency fc which coincided with the cell swelling after infection, but the influence of dynamic cell properties on fc was not investigated.

Although there are only a few reports demonstrating the use of dielectric spectroscopy to monitor stable insect cell lines, this method is used routinely with a wide variety of mammalian cell lines [192, 195] and was proven to work with rS2 cells. In the production of an antimicrobial protein with rS cells under control of the Mt promoter, the dielectric spectroscopy was used to pinpoint the key transitional events of the process (induction and harvest) [209].

4.3. Optical density as a tool to determine absolute cell density

One of the most established methods for the online monitoring of cell suspension cultures is the measurement of optical density (OD) and its popularity is already reflected by the widespread availability of different commercial sensors (Table 7) [210]. Optical density probes exploit the ability of suspended particles (in this case cells) to scatter light in all directions. As shown in Figure 16, the scattering can be used in two main ways to derive information about cell density. Sensors of the first type (e.g. ExCell 230, Exner Process Equipment) measure the reduction in light intensity during its transmission through the medium. The availability of different optical path lengths for this type of sensor allows the analysis of samples containing a low density of particles. Sensors of the second type (e.g. InPro8100, Mettler Toledo) instead detect light that is backscattered at an angle of 180°. Both sensor types use near infrared (NIR) light in the range 780–1100 nm to avoid the signal being affected by changes in the medium colour [183].

Figure 16.

(a) Schematic illustration of an OD probe measuring light transmission. (b) Schematic illustration of an OD probe based on the measurement of 180° backscattered light.

In addition to their traditional use with microbial cells, online OD sensors have already been used to monitor mammalian and insect cells [175]. Accordingly, the density of uninfected, actively growing Sf9 cells has been shown to correlate well with the OD response, and the correlation persists until 24 h post-infection with a recombinant Baculovirus [202]. However, the OD signal becomes static or increases during cell lysis even though the viable cell number declines [202] This is because cell debris and dead cells contribute to light scattering to the same extent as viable cells, which is the major drawback of OD measurements: the method determines the overall abundance of light-scattering particles but cannot distinguish between cells and non-cellular particles nor between living and dead cells. Nevertheless, the method is highly robust for the quantification of viable cells and is therefore especially suitable for monitoring non-lytic expression systems such as rS2 cells [209]. Furthermore, some devices allow the simultaneous determination of dielectric properties and OD (e.g. the combined Hamilton Decyte and Incyte system). Combining both methods may improve process understanding because the ratio of OD and permittivity is an additional real-time parameter that can be used for process monitoring and control.

Advertisement

Acknowledgments

The authors would like to thank the Hessen State Ministry of Higher Education, Research and the Arts for financial support within the Hessen initiative for scientific and economic excellence (LOEWE Program). The authors acknowledge Dr. Richard M Twyman for editing this chapter.

References

  1. 1. Drugmand JC, Schneider YJ, Agathos SN. Insect cells as factories for biomanufacturing. Biotechnol Adv. 2012 Sep;30(5):1140–57.
  2. 2. Cox MMJ. Recombinant protein vaccines produced in insect cells. Vaccine [Internet]. 2012 Feb 27;30(10):1759–66. Available from: http://www.ncbi.nlm.nih.gov/pubmed/22265860
  3. 3. Ikonomou L, Schneider YJ, Agathos SN. Insect cell culture for industrial production of recombinant proteins. Appl Microbiol Biotechnol. 2003 Jul;62(1):1–20.
  4. 4. Altmann F, Staudacher E, Wilson IBH, März L. Insect cells as hosts for the expression of recombinant glycoproteins.pdf. Glycoconj J [Internet]. 1999;16(2):109–23. Available from: http://dx.doi.org/10.1023/A:1026488408951
  5. 5. Grace TD. Establishment of four strains of cells from insect tissues grown in vitro. Nature. 1962 Aug;195:788–9.
  6. 6. Lynn DE. Novel techniques to establish new insect cell lines. Vitr Cell Dev Biol – Anim [Internet]. 2001 Jun;37(6):319. Available from: http://www.bioone.org/perlserv/?request=get-abstract&doi=10.1290/1071-2690(2001)037%3C0319:NTTENI%3E2.0.CO;2
  7. 7. Kost TA, Condreay JP, Jarvis DL. Baculovirus as versatile vectors for protein expression in insect and mammalian cells. Nat Biotechnol [Internet]. 2005;23(5):567–75. Available from: http://www.ncbi.nlm.nih.gov/pubmed/15877075
  8. 8. McCarroll L, King LA. Stable insect cell cultures for recombinant protein production. Curr Opin Biotechnol. 1997;8(5):590–4.
  9. 9. Ramakrishnan B, Qasba PK. Structure-based evolutionary relationship of glycosyltransferases: a case study of vertebrate β1,4-galactosyltransferase, invertebrate β1,4-N-acetylgalactosaminyltransferase and α-polypeptidyl-N-acetylgalactosaminyltransferase. Curr Opin Struct Biol [Internet]. 2010 Oct [cited 2017 Jan 25];20(5):536–42. Available from: http://linkinghub.elsevier.com/retrieve/pii/S0959440X1000120X
  10. 10. Aoki K, Perlman M, Lim J-M, Cantu R, Wells L, Tiemeyer M. Dynamic developmental elaboration of N-linked glycan complexity in the Drosophila melanogaster embryo. J Biol Chem [Internet]. 2007 Mar 23 [cited 2017 Jan 25];282(12):9127–42. Available from: http://www.ncbi.nlm.nih.gov/pubmed/17264077
  11. 11. Tomiya N, Narang S, Lee YC, Betenbaugh MJ. Comparing N-glycan processing in mammalian cell lines to native and engineered lepidopteran insect cell lines. Glycoconj J [Internet]. 2004 [cited 2017 Jan 25];21(6):343–60. Available from: http://www.ncbi.nlm.nih.gov/pubmed/15514482
  12. 12. Nettleship J. Structural biology of glycoproteins. In: Biochemistry, Genetics and Molecular Biology [Internet]. 2012. p. 41–62. Available from: http://cdn.intechopen.com/pdfs/39447/InTech-Structural_biology_of_glycoproteins.pdf
  13. 13. Harrison RL, Jarvis DL. Protein N-glycosylation in the baculovirus–insect cell expression system and engineering of insect cells to produce “Mammalianized” recombinant glycoproteins. Adv Virus Res. 2006;68:159–91.
  14. 14. Altmann F. The role of protein glycosylation in allergy. Int Arch Allergy Immunol. 2007;142(2):99–115.
  15. 15. Durocher Y, Butler M. Expression systems for therapeutic glycoprotein production. Curr Opin Biotechnol. 2009;20(6):700–7.
  16. 16. Kim YK, Kim KR, Kang DG, Jang SY, Kim YH, Cha HJ. Suppression of β-N-acetylglucosaminidase in the N-glycosylation pathway for complex glycoprotein formation in Drosophila S2 cells. Glycobiology. 2009;19(3):301–8.
  17. 17. Kim YK, Kim KR, Kang DG, Jang SY, Kim YH, Cha HJ. Expression of beta-1,4-galactosyltransferase and suppression of beta-N-acetylglucosaminidase to aid synthesis of complex N-glycans in insect Drosophila S2 cells. J Biotechnol [Internet]. 2011;153(3–4):145–52. Available from: http://dx.doi.org/10.1016/j.jbiotec.2011.03.021
  18. 18. Contreras-Gómez A, Sánchez-Mirón A, García-Camacho F, Molina-Grima E, Chisti Y. Protein production using the baculovirus-insect cell expression system. Biotechnol Prog [Internet]. 2014 Jan;30(1):1–18. Available from: http://www.ncbi.nlm.nih.gov/pubmed/24265112
  19. 19. Aumiller JJ, Mabashi-Asazuma H, Hillar A, Shi X, Jarvis DL. A new glycoengineered insect cell line with an inducibly mammalianized protein N-glycosylation pathway. Glycobiology. 2012;22(3):417–28.
  20. 20. Jarvis DL, Finn EE. Modifying the insect cell N-glycosylation pathway with immediate early baculovirus expression vectors. Nat Biotechnol. 1996;14(10):1288–92.
  21. 21. Palmberger D, Wilson IBH, Berger I, Grabherr R, Rendic D. Sweetbac: A new approach for the production of mammalianised glycoproteins in insect cells. PLoS One. 2012;7(4):1–8.
  22. 22. Palmberger D, Klausberger M, Berger I, Grabherr R. MultiBac turns sweet. Bioengineered. [Internet]. 2013. Mar 4;4(2):78–83. Available from: http://www.tandfonline.com/doi/abs/10.4161/bioe.22327
  23. 23. Mabashi-Asazuma H, Kuo CW, Khoo KH, Jarvis DL. A novel baculovirus vector for the production of nonfucosylated recombinant glycoproteins in insect cells. Glycobiology. 2014;24(3):325–40.
  24. 24. Lopez M, Tetaert D, Juliant S, Gazon M, Cerutti M, Verbert A, et al. O-Glycosylation potential of lepidopteran insect cell lines. Biochim Biophys Acta – Gen Subj. 1999;1427(1):49–61.
  25. 25. Stanley P, Schachter H, Taniguchi N. Essentials of Glycobiology [Internet]. Vol. 10. 2009;10:552–553. Available from: http://www.ncbi.nlm.nih.gov/pubmed/20301239
  26. 26. O’Reilly DR, Lockow V, Miller LK. Baculovirus Expression Vectors: A Laboratory Manual. 1st ed. O’Reilly DR, Lockow V, Miller LK, editors. Oxford: Oxford University Press; New York. 1994 p. 368.
  27. 27. Kirnbauer R, Booy F, Cheng N, Lowy DR, Schiller JT. Papillomavirus L1 major capsid protein self-assembles into virus-like particles that are highly immunogenic. Proc Natl Acad Sci U S A [Internet]. 1992 Dec 15 [cited 2016 Jul 31];89(24):12180–4. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1334560
  28. 28. Small EJ, Fratesi P, Reese DM, Strang G, Laus R, Peshwa M V, et al. Immunotherapy of hormone-refractory prostate cancer with antigen-loaded dendritic cells. J Clin Oncol [Internet]. 2000 Dec 1 [cited 2016 Jul 31];18(23):3894–903. Available from: http://www.ncbi.nlm.nih.gov/pubmed/11099318
  29. 29. Cox MMJ, Hollister JR. FluBlok, a next generation influenza vaccine manufactured in insect cells. Biologicals [Internet]. 2009;37(3):182–9. Available from: http://www.sciencedirect.com/science/article/B6WBS-4VVN4T3-1/2/ee1607f7fc8c59aeeed5f8272b2d1a1f
  30. 30. Cox MMJ, Izikson R, Post P, Dunkle L. Safety, efficacy, and immunogenicity of Flublok in the prevention of seasonal influenza in adults. Ther Adv vaccines [Internet]. 2015 Jul [cited 2016 Jul 31];3(4):97–108. Available from: http://www.ncbi.nlm.nih.gov/pubmed/26478817
  31. 31. Nolan T, Roy-Ghanta S, Montellano M, Weckx L, Ulloa-Gutierrez R, Lazcano-Ponce E, et al. Relative efficacy of AS03-adjuvanted pandemic influenza A(H1N1) vaccine in children: results of a controlled, randomized efficacy trial. J Infect Dis [Internet]. 2014 Aug 15 [cited 2016 Jul 31];210(4):545–57. Available from: http://www.ncbi.nlm.nih.gov/pubmed/24652494
  32. 32. Blanchard P, Mahé D, Cariolet R, Keranflec’h A, Baudouard M., Cordioli P, et al. Protection of swine against post-weaning multisystemic wasting syndrome (PMWS) by porcine circovirus type 2 (PCV2) proteins. Vaccine. 2003;21(31):4565–75.
  33. 33. Fachinger V, Bischoff R, Jedidia S Ben, Saalm?ller A, Elbers K. The effect of vaccination against porcine circovirus type 2 in pigs suffering from porcine respiratory disease complex. Vaccine. 2008;26(11):1488–99.
  34. 34. van Aarle P. Suitability of an E2 subunit vaccine of classical swine fever in combination with the E(rns)-marker-test for eradication through vaccination. Dev Biol (Basel) [Internet]. 2003 [cited 2016 Jul 31];114:193–200. Available from: http://www.ncbi.nlm.nih.gov/pubmed/14677689
  35. 35. Graham FL, van der Eb AJ. A new technique for the assay of infectivity of human adenovirus 5 DNA. Virology [Internet]. 1973 Apr [cited 2016 Jun 29];52(2):456–67. Available from: http://www.ncbi.nlm.nih.gov/pubmed/4705382
  36. 36. Burand JP, Summers MD, Smith GE. Transfection with baculovirus DNA. [Internet]. Virology. 1980. Feb [cited 2016 Jun 29];101(1):286–90. Available from: http://linkinghub.elsevier.com/retrieve/pii/004268228090505X
  37. 37. Felgner PL, Gadek TR, Holm M, Roman R, Chan HW, Wenz M, et al. Lipofection: a highly efficient, lipid-mediated DNA-transfection procedure. Proc Natl Acad Sci [Internet]. 1987;84(21):7413–7. Available from: http://www.pnas.org/content/84/21/7413.abstract
  38. 38. Cherbas L, Cherbas P. Drosophila cell culture and transformation. CSH Protoc [Internet].2007(8):pdb.top6. Available from: http://cshprotocols.cshlp.org/content/2007/8/pdb.top6.full
  39. 39. Mann SG, King LA. Efficient transfection of insect cells with baculovirus DNA using electroporation. J Gen Virol [Internet]. 1989 Dec 1 [cited 2016 Jun 29];70(12):3501–5. Available from: http://jgv.microbiologyresearch.org/content/journal/jgv/10.1099/0022-1317-70-12-3501
  40. 40. O’Reilly DR, Miller LK, Luckow VA. Baculovirus expression vectors. A laboratory manual. 1st ed. O’Reilly DR, Miller LK, Luckow VA, editors. Oxford: Oxford University Press; 1994. 347 p.
  41. 41. Wong KT, Peter CH, Greenfield PF, Reid S, Nielsen LK. Low multiplicity infection of insect cells with a recombinant baculovirus: The cell yield concept. Biotechnol Bioeng. 1996 Mar;49(6):659–66.
  42. 42. Rosenberg IM. Protein Analysis and Purification: Benchtop Techniques: Second edition [Internet]. Boston, MA: Birkhäuser Boston; 2005 [cited 2016 Aug 9]. 1–520 p. Available from: http://link.springer.com/10.1007/b138330
  43. 43. Terpe K. Overview of tag protein fusions: from molecular and biochemical fundamentals to commercial systems. Appl Microbiol Biotechnol [Internet]. 2003 Jan [cited 2016 Aug 9];60(5):523–33. Available from: http://link.springer.com/10.1007/s00253-002-1158-6
  44. 44. Kitts PA, Possee RD. A method for producing recombinant baculovirus expression vectors at high frequency. Biotechniques [Internet]. 1993 May [cited 2016 Jul 26];14(5):810–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8512707
  45. 45. Hitchman RB, Possee RD, Crombie AT, Chambers A, Ho K, Siaterli E, et al. Genetic modification of a baculovirus vector for increased expression in insect cells. Cell Biol Toxicol [Internet]. 2010 Feb 5 [cited 2016 Jul 20];26(1):57–68. Available from: http://link.springer.com/10.1007/s10565-009-9133-y
  46. 46. Lu A, Miller LK. Generation of recombinant baculoviruses by direct cloning. Biotechniques [Internet]. 1996 Jul [cited 2016 Jul 28];21(1):63–8. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8816238
  47. 47. Jarvis DL. Baculovirus-insect cell expression systems. Methods Enzymol [Internet]. 2009;463:191–222. Available from: http://www.ncbi.nlm.nih.gov/pubmed/19892174
  48. 48. Pfeifer TA, Hegedus DD, Grigliatti TA, Theilmann DA. Baculovirus immediate-early promoter-mediated expression of the Zeocin? resistance gene for use as a dominant selectable marker in Dipteran and Lepidopteran insect cell lines. Gene. 1997 Apr;188(2):183–90.
  49. 49. Wang M, Tuladhar E, Shen S, Wang H, van Oers MM, Vlak JM, et al. Specificity of baculovirus P6.9 basic DNA-binding proteins and critical role of the C Terminus in virion formation. J Virol [Internet]. 2010 Sep 1 [cited 2016 Jul 29];84(17):8821–8. Available from: http://jvi.asm.org/cgi/doi/10.1128/JVI.00072-10
  50. 50. Chaabihi H, Ogliastro MH, Martin M, Giraud C, Devauchelle G, Cerutti M. Competition between baculovirus polyhedrin and p10 gene expression during infection of insect cells. J Virol [Internet]. 1993 May [cited 2016 Jul 29];67(5):2664–71. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8474166
  51. 51. Lindner P, Bauer K, Krebber A, Nieba L, Kremmer E, Krebber C, et al. Specific detection of his-tagged proteins with recombinant anti-His tag scFv-phosphatase or scFv-phage fusions. Biotechniques [Internet]. 1997 Jan [cited 2016 Jul 29];22(1):140–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8994661
  52. 52. Carrington JC, Dougherty WG. A viral cleavage site cassette: Identification of amino acid sequences required for tobacco etch virus polyprotein processing. Proc Natl Acad Sci U S A [Internet]. 1988 May [cited 2016 Jul 29];85(10):3391–5. Available from: http://www.ncbi.nlm.nih.gov/pubmed/3285343
  53. 53. Dougherty WG, Carrington JC, Cary SM, Parks TD. Biochemical and mutational analysis of a plant virus polyprotein cleavage site. EMBO J [Internet]. 1988 May [cited 2016 Jul 29];7(5):1281–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/3409865
  54. 54. Futatsumori-Sugai M, Tsumoto K. Signal peptide design for improving recombinant protein secretion in the baculovirus expression vector system. Biochem Biophys Res Commun. 2010 Jan;391(1):931–5.
  55. 55. Radford KM, Cavegn C, Bertrand M, Bernard AR, Reid S, Greenfield PF. The indirect effects of multiplicity of infection on baculovirus expressed proteins in insect cells: Secreted and non-secreted products. Cytotechnology [Internet]. 1997 May [cited 2016 Aug 16];24(1):73–81. Available from: http://www.ncbi.nlm.nih.gov/pubmed/22358599
  56. 56. Chico E, Jäger V. Perfusion culture of baculovirus-infected BTI-Tn-5B1-4 insect cells: A method to restore cell-specific β-trace glycoprotein productivity at high cell density. Biotechnol Bioeng [Internet]. 2000 Dec 5 [cited 2016 Aug 14];70(5):574–86. Available from: http://doi.wiley.com/10.1002/1097-0290%2820001205%2970%3A5%3C574%3A%3AAID-BIT12%3E3.0.CO%3B2-Q
  57. 57. Gotoh T, Miyazaki Y, Chiba K, Kikuchi K-I. Significant increase in recombinant protein production of a virus-infected Sf-9 insect cell culture of low MOI under low dissolved oxygen conditions. J Biosci Bioeng [Internet]. 2002 Jan;94(5):426–33. Available from: http://www.ncbi.nlm.nih.gov/pubmed/16233329
  58. 58. Maranga L, Brazão TF, Carrondo MJT. Virus-like particle production at low multiplicities of infection with the baculovirus insect cell system. Biotechnol Bioeng [Internet]. 2003 Oct 20 [cited 2016 Aug 16];84(2):245–53. Available from: http://doi.wiley.com/10.1002/bit.10773
  59. 59. Zhang YH, Enden G, Merchuk JC. Insect cells–baculovirus system: Factors affecting growth and low MOI infection. Biochem Eng J. 2005;27(1):8–16.
  60. 60. Power JF, Reid S, Radford KM, Greenfield PF, Nielsen LK. Modeling and optimization of the baculovirus expression vector system in batch suspension culture. Biotechnol Bioeng [Internet]. 1994 Sep 5 [cited 2016 Aug 16];44(6):710–9. Available from: http://doi.wiley.com/10.1002/bit.260440607
  61. 61. Licari P, Bailey JE. Modeling the population dynamics of baculovirus-infected insect cells: Optimizing infection strategies for enhanced recombinant protein yields. Biotechnol Bioeng [Internet]. 1992 Feb 20 [cited 2016 Aug 16];39(4):432–41. Available from: http://doi.wiley.com/10.1002/bit.260390409
  62. 62. Hu YC, Bentley WE. Effect of MOI ratio on the composition and yield of chimeric infectious bursal disease virus-like particles by baculovirus co-infection: deterministic predictions and experimental results. Biotechnol Bioeng [Internet]. 2001 Oct 5 [cited 2016 Aug 16];75(1):104–19. Available from: http://doi.wiley.com/10.1002/bit.1170
  63. 63. Roldão A, Carrondo MJT, Alves PM, Oliveira R. Stochastic simulation of protein expression in the baculovirus/insect cells system. Comput Chem Eng. 2008;32(1):68–77.
  64. 64. Murhammer DW. Baculovirus and Insect Cell Expression Protocols. [Internet]. Murhammer DW, editor. Totowa, NJ: Humana Press; 2007. (Methods in Molecular BiologyTM; vol. 388). Available from: http://link.springer.com/10.1007/978-1-59745-457-5
  65. 65. Schweizer HP. Optimized insect cell culture for the production of recombinant heterologous proteins and baculovirus particles. Biotechniques. 2001;31(6):18–21.
  66. 66. Bernard A, Payton M, Radford KR. Protein expression in the baculovirus system. In: Current Protocols in Protein Science [Internet]. John Wiley & Sons, Inc; Hoboken, NJ, USA. 1995. pp. 5.5.1–5.5.18. Available from: http://doi.wiley.com/10.1002/0471140864.ps0505s00
  67. 67. Bédard C, Kamen A, Tom R, Massie B. Maximization of recombinant protein yield in the insect cell/baculovirus system by one-time addition of nutrients to high-density batch cultures. Cytotechnology [Internet]. 1994 [cited 2016 Aug 16];15(1–3):129–38. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7765925
  68. 68. Krammer F, Grabherr R. Alternative influenza vaccines made by insect cells. Trends Mol Med. 2010/06/24. 2010;16(7):313–20.
  69. 69. Airenne KJ, Hu Y-C, Kost TA, Smith RH, Kotin RM, Ono C, et al. Baculovirus: an insect-derived vector for diverse gene transfer applications. Mol Ther [Internet]. 2013;21(4):739–49. Available from: http://dx.doi.org/10.1038/mt.2012.286
  70. 70. Felberbaum RS. The baculovirus expression vector system: A commercial manufacturing platform for viral vaccines and gene therapy vectors. Biotechnol J [Internet]. 2015 May [cited 2016 May 13];10(5):702–14. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25800821
  71. 71. Aucoin MG, Mena JA, Kamen AA. Bioprocessing of baculovirus vectors: a review. Curr Gene Ther [Internet]. 2010;10(3):174–86. Available from: http://www.ncbi.nlm.nih.gov/pubmed/20380645
  72. 72. Michalsky R, Pfromm PH, Czermak P, Sorensen CM, Passarelli AL. Effects of temperature and shear force on infectivity of the baculovirus Autographa californica M nucleopolyhedrovirus. J Virol Methods. 2008;153(2):90–6.
  73. 73. Grein TA, Michalsky R, Vega López M, Czermak P. Purification of a recombinant baculovirus of Autographa californica M nucleopolyhedrovirus by ion exchange membrane chromatography. J Virol Methods [Internet]. 2012;183(2):117–24. Available from: http://dx.doi.org/10.1016/j.jviromet.2012.03.031
  74. 74. Michalsky R, Passarelli AL, Pfromm PH, Czermak P. Purification of the baculovirus Autographa californica M nucleopolyhedrovirus by tangential flow ultrafiltration. Desalination. 2009 Sep;245(1–3):694–700.
  75. 75. Gerster P, Kopecky EM, Hammerschmidt N, Klausberger M, Krammer F, Grabherr R, et al. Purification of infective baculoviruses by monoliths. J Chromatogr A. 2013;1290:36–45.
  76. 76. de Jongh WA, Salgueiro S, Dyring C. The use of Drosophila S2 cells in R&D and bioprocessing. Pharm Bioprocess. 2013 Jun;1(2):197–213.
  77. 77. Moraes ÂM, Jorge SAC, Astray RM, Suazo CAT, Calderón Riquelme CE, Augusto EFP, et al. Drosophila melanogaster S2 cells for expression of heterologous genes: From gene cloning to bioprocess development. Biotechnol Adv. 2012;30(3):613–28.
  78. 78. Schneider I. Cell lines derived from late embryonic stages of Drosophila melanogaster. J Embryol Exp Morphol. 1972 Apr;27(2):353–65.
  79. 79. Suárez-Patiño SF, Mancini RA, Pereira CA, Suazo CAT, Mendonça RZ, Jorge SAC. Transient expression of rabies virus glycoprotein (RVGP) in Drosophila melanogaster Schneider 2 (S2) cells. J Biotechnol [Internet]. 2014 Dec 20;192 Pt A(Part A):255–62. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25011097
  80. 80. Ventini DC, Astray RM, Lemos MAN, Jorge SAC, Calderon Riquelme C, Suazo CAT, et al. Recombinant rabies virus glycoprotein synthesis in bioreactor by transfected Drosophila melanogaster S2 cells carrying a constitutive or an inducible promoter. J Biotechnol. 2010 Apr;146(4):169–72.
  81. 81. Wang L, Hu H, Yang J, Wang F, Kaisermayer C, Zhou P. High yield of human monoclonal antibody produced by stably transfected Drosophila Schneider 2 cells in perfusion culture using wave bioreactor. Mol Biotechnol. 2012 Oct;52(2):170–9.
  82. 82. Cherbas L, Cherbas P. Transformation of Drosophila Cell Lines. In: Murhammer D, editor. Baculovirus and Insect Cell Expression Protocols [Internet]. Humana Press; 2007. pp. 317–40. (Methods in Molecular BiologyTM). Available from: http://link.springer.com/10.1007/978-1-59745-457-5_16
  83. 83. Escoubas P, Bernard C, Lambeau G, Lazdunski M, Darbon H. Recombinant production and solution structure of PcTx1, the specific peptide inhibitor of ASIC1a proton-gated cation channels. Protein Sci [Internet]. 2003 Jul;12(7):1332–43. Available from: http://doi.wiley.com/10.1110/ps.0307003
  84. 84. Vogel CW, Fritzinger DC, Hew BE, Thorne M, Bammert H. Recombinant cobra venom factor. Mol Immunol. 2004 Jun;41(2–3):191–9.
  85. 85. Prosise WW, Yarosh-Tomaine T, Lozewski Z, Ingram RN, Zou J, Liu JJ, et al. Protease domain of human ADAM33 produced by Drosophila S2 cells. Protein Expr Purif. 2004 Dec;38(2):292–301.
  86. 86. Scotter AJ, Kuntz DA, Saul M, Graham LA, Davies PL, Rose DR. Expression and purification of sea raven type II antifreeze protein from Drosophila melanogaster S2 cells. Protein Expr Purif. 2006;47(2):374–83.
  87. 87. Srandergaard L. Drosophila cells can be grown to high cell densities in a bioreactor. Biotechnol Tech. 1996 Mar;10(3):161–6.
  88. 88. Vieira PB, da Costa BLV, de Fatima Pires Augusto E, Tonso A. Culturing Drosophila melanogaster (S2) in a chemostat. Biotechnol Lett. 2015 Nov;37(3):533–8.
  89. 89. De Jongh WA, Dos SMRM, Leisted C, Strobaek A, Berisha B, Nielsen MA, et al. Development of a Drosophila S2 insect-cell based placental malaria vaccine production process. BMC Proceedings Conf 23rd Meet Eur Soc Anim Cell Technol ESACT [Internet]. 2013 Dec;20130623(20130626):P20. Available from: http://ovidsp.ovid.com/ovidweb.cgi?T=JS&CSC=Y&NEWS=N&PAGE=fulltext&D=emed12&AN=71477960 http://eleanor.lib.gla.ac.uk:4550/resserv?sid=OVID:embase&id=pmid:&id=doi:&issn=1753-6561&isbn=&volume=7&issue=&spage=&pages=&date=2013&title=BMC+Proceedings&atitle=De
  90. 90. Dyring C. Development of Pregnancy- Associated Malaria Vaccine Using the ExpreS 2 Insect Cell Expression System. Bioprocess J. 2012 Oct;11(3):14–9.
  91. 91. Lieberman MM, Clements DE, Ogata S, Wang G, Corpuz G, Wong T, et al. Preparation and immunogenic properties of a recombinant West Nile subunit vaccine. Vaccine. 2007 Jan;25(3):414–23.
  92. 92. Yang L, Song Y, Li X, Huang X, Liu J, Ding H, et al. HIV-1 Virus-Like Particles Produced by Stably Transfected Drosophila S2 Cells: a Desirable Vaccine Component. J Virol. 2012 Jul;86(14):7662–76.
  93. 93. Chung HY, Hwang-Bo J, Kim SK, Baek NI, Lee YH, Chung IS, et al. Functional expression of Arabidopsis thaliana sterol glycosyltransferase from stably transformed Drosophila melanogaster S2 cells. Biotechnol Bioprocess Eng. 2011 Aug;16(4):801–7.
  94. 94. Graziano MP, Broderick DJ, Tota MR. Expression of G protein-coupled receptors in Drosophila Schneider 2 cells. In: Receptor biochemistry and methodology series, identification and expression of G protein-coupled receptors. John Wiley & Sons; 1998. p. 181–95.
  95. 95. Uribe E, Venkatesan M, Rose DR, Ewart KV. Expression of recombinant Atlantic salmon serum C-type lectin in Drosophila melanogaster Schneider 2 cells. Cytotechnology. 2013 Aug;65(4):513–21.
  96. 96. Santos MG, Jorge SAC, Brillet K, Pereira CA. Improving heterologous protein expression in transfected Drosophila S2 cells as assessed by EGFP expression. Cytotechnology. 2007 May;54(1):15–24.
  97. 97. Iwaki T, Umemura K. A single plasmid transfection that offers a significant advantage associated with puromycin selection, fluorescence-assisted cell sorting, and doxycycline-inducible protein expression in mammalian cells. Cytotechnology. 2011 May;63(4):337–43.
  98. 98. Jorge SAC, Santos AS, Spina A, Pereira CA. Expression of the hepatitis B virus surface antigen in Drosophila S2 cells. Cytotechnology [Internet]. 2008 May;57(1):51–9. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=2553638&tool=pmcentrez&rendertype=abstract
  99. 99. González M, Martín-Ruíz I, Jiménez S, Pirone L, Barrio R, Sutherland JD. Generation of stable Drosophila cell lines using multicistronic vectors. Sci Rep [Internet]. 2011 Aug;1:75. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3216562&tool=pmcentrez&rendertype=abstract
  100. 100. Lee JM, Chung HY, Kim K Il, Yoo KH, Hwang-Bo J, Chung IS, et al. Synthesis of double-layered rotavirus-like particles using internal ribosome entry site vector system in stably-transformed Drosophila melanogaster. Biotechnol Lett. 2011 Jan;33(1):41–6.
  101. 101. Monroe TJ, Muhlmann-Diaz MC, Kovach MJ, Carlson JO, Bedford JS, Beaty BJ. Stable transformation of a mosquito cell line results in extraordinarily high copy numbers of the plasmid. Proc Natl Acad Sci U S A [Internet]. 1992 Jul;89(13):5725–9. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=49369&tool=pmcentrez&rendertype=abstract
  102. 102. Cherbas L, Gong L. Cell lines. Methods [Internet]. 2014 Jun;(1):74–81. Available from: http://linkinghub.elsevier.com/retrieve/pii/S1046202314000073
  103. 103. van der Straten A, Johansen H, Rosenberg M, Sweet R. Introduction and constitutive expression of gene products in cultured Drosophila cells using hygromycin B selection. Methods Mol Cell Biol. 1989;1(1):1–8.
  104. 104. Ivey-Hoyle M. Recombinant gene expression in cultured Drosophila melanogaster cells. Curr Opin Biotechnol. 1991;2(5):704–7.
  105. 105. Segal D, Cherbas L, Cherbas P. Genetic transformation of Drosophila cells in culture by P element-mediated transposition. Somat Cell Mol Genet. 1996 Mar;22(2):159–65.
  106. 106. Steller H, Pirrotta V. A transposable P vector that confers selectable G418 resistance to Drosophila larvae. EMBO J [Internet]. 1985 Jan;4(1):167–71. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=554166&tool=pmcentrez&rendertype=abstract
  107. 107. Neumüller RA, Wirtz-Peitz F, Lee S, Kwon Y, Buckner M, Hoskins RA, et al. Stringent analysis of gene function and protein-protein interactions using fluorescently tagged genes. Genetics. 2012 Mar;190(3):931–40.
  108. 108. Cherbas L, Hackney J, Gong L, Salzer C, Mauser E, Zhang D, et al. Tools for targeted genome engineering of established Drosophila cell lines. Genetics. 2015 Dec;201(4):1307–18.
  109. 109. Groth AC, Fish M, Nusse R, Calos MP. Construction of transgenic Drosophila by using the site-specific integrase from phage phiC31. Genetics. 2004 Apr;166(4):1775–82.
  110. 110. Manivannan SN, Jacobsen TL, Lyon P, Selvaraj B, Halpin P, Simcox A. Targeted integration of single-copy transgenes in Drosophila melanogaster tissue-culture cells using recombination-mediated cassette exchange. Genetics. 2015 Dec;201(4):1319–28.
  111. 111. Makridou P, Burnett C, Landy T, Howard K. Hygromycin B-selected cell lines from GAL4-regulated pUAST constructs. Genesis. 2003 Jun;36(2):83–7.
  112. 112. Bryson TD, Weber CM, Henikoff S. Baculovirus-encoded protein expression for epigenomic profiling in Drosophila cells. Fly (Austin). 2010 Jul;4(3):258–65.
  113. 113. Kim KR, Kim YK, Cha HJ. Recombinant baculovirus-based multiple protein expression platform for Drosophila S2 cell culture. J Biotechnol. 2008 Jan;133(1):116–22.
  114. 114. Cho HS, Kim YK, Cha HJ. Expression of double foreign protein types following recombinant baculovirus infection of stably transfected Drosophila S2 cells. Enzyme Microb Technol. 2004;35(6–7):525–31.
  115. 115. Lee DF, Chen CC, Hsu T a, Juang JL. A baculovirus superinfection system: Efficient vehicle for gene transfer into Drosophila S2 cells. J Virol. 2000 Dec;74(24):11873–80.
  116. 116. Schetz JA, Shankar EPN. Protein Expression in the Drosophila Schneider 2 Cell System. In: Gerfen CR, Holmes A, Sibley D, Skolnick P, Wray S, editors. Current Protocols in Neuroscience [Internet]. New York, NY: John Wiley & Sons, Inc; 2004. p. Unit 4.16. Available from: http://doi.wiley.com/10.1002/0471142301.ns0416s27
  117. 117. Echalier G. Drosophila Continuous Cell Lines. In: Drosophila Cells in Culture. Academic Press; New York; 1997. pp. 129–86.
  118. 118. Cherbas L, Moss R, Cherbas P. Transformation Techniques for Drosophila Cell Lines. In: Goldstein LSB, Fyrberg EA, editors. Methods in Cell Biology [Internet]. San Diego, New York, Boston, London, Sydney, Tokyo, Toronto: Academic Press; 1994. pp. 161–79. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7707950
  119. 119. Chung YT, Keller EB. Positive and negative regulatory elements mediating transcription from the Drosophila melanogaster actin 5C distal promoter. Mol Cell Biol [Internet]. 1990 Dec;10(12):6172–80. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2123290
  120. 120. Bunch TA, Grinblat Y, Goldstein LSB. Characterization and use of the Drosophila metallothionein promoter in cultured Drosophila melanogaster cells. Nucleic Acids Res. 1988 Feb;16(3):1043–61.
  121. 121. Goldstein LSB, Fyrberg EA. Drosophila Melanogaster: Practical Uses in Cell and Molecular Biology. Goldstein LSB, Fyrberg EA, editors. Academic Press; San Diego. 1994. 755 p.
  122. 122. Lim HJ, Cha HJ. Observation and modeling of induction effect on human transferrin production from stably transfected Drosophila S2 cell culture. Enzyme Microb Technol. 2006 Jun;39(2):208–14.
  123. 123. Valle MA, Kester MB, Burns AL, Marx SJ, Spiegel AM, Shiloach J. Production and purification of human menin from Drosophila melanogaster S2 cells using stirred tank reactor. Cytotechnology. 2001 Mar;35(2):127–35.
  124. 124. Qin JY, Zhang L, Clift KL, Hulur I, Xiang AP, Ren B-Z, et al. Systematic comparison of constitutive promoters and the doxycycline-inducible promoter. PLoS One [Internet]. 2010 Jan 12 [cited 2014 Jul 11];5(5):e10611. Available from: http://www.plosone.org/article/info:doi/10.1371/journal.pone.0010611#s3
  125. 125. Schetz JA, Kim OJ, Sibley DR. Pharmacological characterization of mammalian D1 and D2 dopamine receptors expressed in Drosophila Schneider-2 cells. J Recept Signal Transduct Res [Internet]. 2003 Feb;23(1):99–109. Available from: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=12680592
  126. 126. Di Nocera PP, Dawid IB. Transient expression of genes introduced into cultured cells of Drosophila. Proc Natl Acad Sci U S A [Internet]. 1983;80(23):7095–8. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=389999&tool=pmcentrez&rendertype=abstract
  127. 127. Qin JY, Zhang L, Clift KL, Hulur I, Xiang AP, Ren BZ, et al. Systematic comparison of constitutive promoters and the doxycycline-inducible promoter. PLoS One. 2010 Dec;5(5):e10611.
  128. 128. Dobrosotskaya IY, Goldstein JL, Brown MS, Rawson RB. Reconstitution of sterol-regulated endoplasmic reticulum-to-Golgi transport of SREBP-2 in insect cells by co-expression of mammalian SCAP and Insigs. J Biol Chem. 2003 Sep;278(37):35837–43.
  129. 129. Chyb S, Dahanukar A, Wickens A, Carlson JR. Drosophila Gr5a encodes a taste receptor tuned to trehalose. Proc Natl Acad Sci U S A [Internet]. 2003 Nov;100 Suppl(Suppl 2):14526–30. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=304113&tool=pmcentrez&rendertype=abstract
  130. 130. Beck M, Strand MR. Glc1.8 from Microplitis demolitor bracovirus induces a loss of adhesion and phagocytosis in insect high five and S2 cells. J Virol. 2005 Feb;79(3):1861–70.
  131. 131. Benyajati C, Ewel A, McKeon J, Chovav M, Juan E. Characterization and purification of Adh distal promoter factor 2, Adf-2, a cell-specific and promoter-specific repressor in Drosophila. Nucleic Acids Res. 1992 Sep;20(17):4481–9.
  132. 132. Angelichio ML, Beck JA, Johansen H, Ivey-Hoyle M. Comparison of several promoters and polyadenylation signals for use in heterologous gene expression in cultured Drosophila cells. Nucleic Acids Res [Internet]. 1991 Sep;19(18):5037–43. Available from: http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=1656386&retmode=ref&cmd=prlinks%5Cnpapers3://publication/uuid/3CED37BF-4D13-4157-BCD7-060DDA05EC5A
  133. 133. Cavener DR. Comparison of the consensus sequence flanking translational start sites in Drosophila and vertebrates. Nucleic Acids Res. 1987 Feb;15(4):1353–61.
  134. 134. Kirkpatrick RB, Ganguly S, Angelichio M, Griego S, Shatzman A, Silverman C, et al. Heavy chain dimers as well as complete antibodies are efficiently formed and secreted from Drosophila via a BiP-mediated pathway. J Biol Chem. 1995 Aug;270(34):19800–5.
  135. 135. Culp JS, Johansen H, Hellmig B, Beck J, Matthews TJ, Delers A, et al. Regulated expression allows high level production and secretion of HIV-1 gp120 envelope glycoprotein in Drosophila Schneider cells. Biotechnology (N Y) [Internet]. 1991 Feb;9(2):173–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1369452
  136. 136. Kollewe C, Vilcinskas A. Production of recombinant proteins in insect cells. Am J Biochem Biotechnol. 2013 Mar;9(3):255–71.
  137. 137. Park JH, Hwang IS, Kim K Il, Lee JM, Park YM, Park CH, et al. Functional expression of recombinant human ribonuclease/angiogenin inhibitor in stably transformed Drosophila melanogaster S2 cells. Cytotechnology. 2008 May;57(1):93–9.
  138. 138. Shin HS, Cha HJ. Statistical optimization for immobilized metal affinity purification of secreted human erythropoietin from Drosophila S2 cells. Protein Expr Purif. 2003 Apr;28(2):331–9.
  139. 139. Braun H, Suske G. Vectors for inducible expression of dual epitope-tagged proteins in insect cells. Biotechniques. 1999 Jun;26(6):1038–42.
  140. 140. Kimura M, Takatsuki A, Yamaguchi I. Blasticidin S deaminase gene from Aspergillus terreus (BSD): a new drug resistance gene for transfection of mammalian cells. Biochim Biophys Acta. 1994 Nov;1219(3):653–9.
  141. 141. To A, Culture T. Drosophila Genomics Resource Center. Insulin. 1990. p. 10–2.
  142. 142. Iwaki T, Figuera M, Ploplis VA, Castellino FJ. Rapid selection of Drosophila S2 cells with the puromycin resistance gene. Biotechniques. 2003 Sep;35(3):482–6.
  143. 143. Rebay I, Fleming RJ, Fehon RG, Cherbas L, Cherbas P, Artavanis-Tsakonas S. Specific EGF repeats of Notch mediate interactions with Delta and Serrate: implications for Notch as a multifunctional receptor. Cell [Internet]. 1991 Nov;67(4):687–99. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1657403
  144. 144. Bourouis M, Jarry B. Vectors containing a prokaryotic dihydrofolate reductase gene transform Drosophila cells to methotrexate-resistance. EMBO J [Internet]. 1983;2(7):1099–104. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=555241&tool=pmcentrez&rendertype=abstract
  145. 145. Jokerst RS, Weeks JR, Zehring W a, Greenleaf a L. Analysis of the gene encoding the largest subunit of RNA polymerase II in Drosophila. Mol Gen Genet [Internet]. 1989 Jan;215(2):266–75. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2496296
  146. 146. Lemos MAN, Santos AS d, Astray RM, Pereira CA, Jorge SAC. Rabies virus glycoprotein expression in Drosophila S2 cells. I: Design of expression/selection vectors, subpopulations selection and influence of sodium butyrate and culture medium on protein expression. J Biotechnol. 2009 Aug;143(2):103–10.
  147. 147. Fan Q, Bohannon KP, Longnecker R. Drosophila Schneider 2 (S2) cells: A novel tool for studying HSV-induced membrane fusion. Virology. 2013 Mar;437(2):100–9.
  148. 148. Shin HS, Cha HJ. Facile and statistical optimization of transfection conditions for secretion of foreign proteins from insect Drosophila S2 cells using green fluorescent protein reporter. Biotechnol Prog. 2002 Jan;18(6):1187–94.
  149. 149. Deml L, Wolf H, Wagner R. High level expression of hepatitis B virus surface antigen in stably transfected Drosophila Schneider-2 cells. J Virol Methods. 1999;79(2):191–203.
  150. 150. Han K. An efficient DDAB-mediated transfection of Drosophila S2 cells. Nucleic Acids Res. 1996 Nov;24(21):4362–3.
  151. 151. Perret BG, Wagner R, Lecat S, Brillet K, Rabut G, Bucher B, et al. Expression of EGFP-amino-tagged human mu opioid receptor in Drosophila Schneider 2 cells: A potential expression system for large-scale production of G-protein coupled receptors. Protein Expr Purif. 2003 Sep;31(1):123–32.
  152. 152. Baum B, Cherbas L. Drosophila Cell Lines as Model Systems and as an Experimental Tool. In: Methods in molecular biology (Clifton, NJ) [Internet]. 2008 [cited 2017 Mar 20]. p. 391–424. Available from: http://www.ncbi.nlm.nih.gov/pubmed/18641962
  153. 153. Zurovec M, Dolezal T, Gazi M, Pavlova E, Bryant PJ. Adenosine deaminase-related growth factors stimulate cell proliferation in Drosophila by depleting extracellular adenosine. Proc Natl Acad Sci U S A [Internet]. 2002 Apr;99(7):4403–8. Available from: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=123660&tool=pmcentrez&rendertype=abstract
  154. 154. Maier SA, Podemski L, Graham SW, McDermid HE, Locke J. Characterization of the adenosine deaminase-related growth factor (ADGF) gene family in Drosophila. Gene. 2001 Dec;280(1–2):27–36.
  155. 155. Santos NGL, Rocca MP, Pereira CA, Ventini DC, Puglia ALP, Jorge SAC, et al. Impact of recombinant Drosophila S2 cell population enrichment on expression of rabies virus glycoprotein. Cytotechnology [Internet]. 2016 Dec 23;68(6):2605–11. Available from: http://link.springer.com/10.1007/s10616-016-9984-z
  156. 156. Schmidt FR. Recombinant expression systems in the pharmaceutical industry. Appl Microbiol Biotechnol [Internet]. 2004 Sep 24 [cited 2017 Jan 10];65(4):363–72. Available from: http://www.ncbi.nlm.nih.gov/pubmed/15480623
  157. 157. Vatandoost J, Bos MHA. Efficient expression of functional human coagulation factor IX in stably-transfected Drosophila melanogaster S2 cells; comparison with the mammalian CHO system. Biotechnol Lett [Internet]. 2016 Oct 26 [cited 2017 Jan 10];38(10):1691–8. Available from: http://www.ncbi.nlm.nih.gov/pubmed/27565667
  158. 158. Weber W, Fussenegger M. Insect cell-based recombinant protein production. In: Cell and Tissue Reaction Engineering. Springer, Berlin Heidelberg; 2009. pp. 263–77. (Principles and Practice).
  159. 159. Echalier G. Composition of the body fluid of Drosophila and the design of culture media for Drosophila cells. In: Echalier G, editor. Drosophila Cells in Culture. Academic Press; New York; 1997. pp. 1–67.
  160. 160. Schlaeger E-J. Medium design for insect cell culture. Cytotechnology. 1996 Jan;20(1–3):57–70.
  161. 161. Agathos SN. Insect cell bioreactors. Cytotechnology [Internet]. 1996;20(1):173–89. Available from: http://dx.doi.org/10.1007/BF00350398
  162. 162. Weber W, Weber E, Geisse S, Memmert K. Optimisation of protein expression and establishment of the wave bioreactor for Baculovirus/insect cell culture. Cytotechnology. 2002;38(1–3):77–85.
  163. 163. Weber W, Weber E, Geisse DS, Memmert DK. Catching the Wave: The BEVS and the Biowave. In: Lindner-Olsson DE, Chatzissavidou MN, Lüllau DE, editors. Animal Cell Technology: From Target to Market; Springer, Netherlands; 2001. pp. 335–7. (ESACT Proceedings).
  164. 164. Eibl R, Eibl D, Pörtner R, Catapano G, Czermak P. Cell and tissue reaction engineering. In: Cell an Tissue Reaction Engineering [Internet]; Springer, Berlin Heidelberg; 2009. p. 363. (Principles and Practice). Available from: http://link.springer.com/10.1007/978-3-540-68182-3
  165. 165. Druzinec D, Salzig D, Kraume M, Czermak P. Micro-bubble aeration in turbulent stirred bioreactors: Coalescence behavior in Pluronic F68 containing cell culture media. Chem Eng Sci. 2015 Apr;126:160–8.
  166. 166. Wong TKK, Nielsen LK, Greenfield PF, Reid S. Relationship between oxygen-uptake rate and time of infection of Sf9 insect cells infected with a recombinant baculovirus. Cytotechnology. 1994;15(1–3):157–67.
  167. 167. Maiorella B, Inlow D, Shauger A, Harano D. Large-Scale Insect Cell-Culture for Recombinant Protein Production. Nat Biotech. 1988 Dec;6(12):1406–10.
  168. 168. Reuveny S, Kemp CW, Eppstein L, Shiloach J. Carbohydrate Metabolism in Insect Cell Cultures during Cell Growth and Recombinant Protein Production. Ann N Y Acad Sci [Internet]. 1992 Oct [cited 2016 Aug 10];665(1 Biochemical E):230–7. Available from: http://doi.wiley.com/10.1111/j.1749-6632.1992.tb42587.x
  169. 169. Rhiel, M., Mitchell-Logan, C.M., Murhammer DW. Comparison of Trichoplusia ni BTI-Tn-5B1-4 (high fiveTM) and Spodoptera frugiperda Sf-9 insect cell line metabolism in suspension cultures. Biotechnol Bioeng. 1997;55:909–20.
  170. 170. Hensler WT, Agathos SN. Evaluation of monitoring approaches and effects of culture conditions on recombinant protein production in baculovirus-infected insect cells. Cytotechnology [Internet]. 1994;15(1):177–86. Available from: http://dx.doi.org/10.1007/BF00762392
  171. 171. Deutschmann SM, Jäger V. Optimization of the growth conditions of Sf21 insect cells for high-density perfusion culture in stirred-tank bioreactors. Enzyme Microb Technol [Internet]. 1994 Jun [cited 2016 Aug 10];16(6):506–12. Available from: http://linkinghub.elsevier.com/retrieve/pii/0141022994900221
  172. 172. Pamboukian MM, Jorge SAC, Santos MG, Yokomizo AY, Pereira CA, Tonso A. Insect cells respiratory activity in bioreactor. Cytotechnology. 2008 May;57(1):37–44.
  173. 173. Batista FRX, Moraes ÂM, B?ntemeyer H, Noll T. Influence of culture conditions on recombinant Drosophila melanogaster S2 cells producing rabies virus glycoprotein cultivated in serum-free medium. Biologicals. 2009 Apr;37(2):108–18.
  174. 174. Kane J. Measuring kLa for better bioreactor performance. BioProcess International [Internet]. 2012; 10(3):46–9. Available from: http://www.bioprocessintl.com/wp-content/uploads/2014/07/BPI_A_121003AR10_O_176218a.pdf
  175. 175. Weber W, Fussenegger M. Baculovirus-based production of biopharmaceuticals using insect cell culture processes. In: Knäblein J, editor. Modern Biopharmaceuticals [Internet]. Weinheim, Germany: Wiley-VCH, Verlag GmbH; 2005. pp. 1045–62. Available from: http://doi.wiley.com/10.1002/9783527620982.ch44
  176. 176. Palomares LA, Gonzalez M, Ramirez OT. Evidence of pluronic F-68 direct interaction with insect cells: Impact on shear protection, recombinant protein, and baculovirus production. Enzyme Microb Technol. 2000;26(5–6):324–31.
  177. 177. Doran PM. Bioprocess Engineering Principles: [Internet]. 2nd ed. Doran PM, editor. Amsterdam; Boston: Academic Press; 2012. [cited 2017 Mar 17]. 919 p. Available from: http://www.sciencedirect.com/science/book/9780122208515
  178. 178. Myers RH, Montgomery DC, Anderson-Cook CM. Response urface methodology: process and product optimization using designed experiments. 3rd ed. Hoboken, NJ, USA: John Wiley & Sons; 2009. 704. p.
  179. 179. United States Food and Drug Administration. Guidance for industry: PAT — A framework for innovative pharmaceutical development, manufacturing, and quality assurance. U.S. Department of Health and Human Services. 2004. pp. 1–16.
  180. 180. Junker BH, Wang HY. Bioprocess monitoring and computer control: Key roots of the current PAT initiative. Biotechnol Bioeng. 2006;95(2):226–61.
  181. 181. Gnoth S, Jenzsch M, Simutis R, Lübbert A. Process Analytical Technology (PAT): Batch-to-batch reproducibility of fermentation processes by robust process operational design and control. J Biotechnol. 2007;132(2):180–6.
  182. 182. Párta L, Zalai D, Borbély S, Putics Á. Application of dielectric spectroscopy for monitoring high cell density in monoclonal antibody producing CHO cell cultivations. Bioprocess Biosyst Eng [Internet]. 2014 Feb 26;37(2):311–23. Available from: http://link.springer.com/10.1007/s00449-013-0998-z
  183. 183. Wu P, Ozturk SS, Blackie JD, Thrift JC, Fugueroa C, Naveh D. Evaluation and applications of optical cell density probes in mammalian cell bioreactors. Biotechnol Bioeng. 1995;45(6):495–502.
  184. 184. Marose S, Lindemann C, Scheper T. Two-dimensional fluorescence spectroscopy: A new tool for on-line bioprocess monitoring. Biotechnol Prog. 1998 Jan;14(1):63–74.
  185. 185. Cole H, Demont A, Marison I. The Application of Dielectric Spectroscopy and Biocalorimetry for the Monitoring of Biomass in Immobilized Mammalian Cell Cultures. Processes [Internet]. 2015 May;3(2):384–405. Available from: http://www.mdpi.com/2227-9717/3/2/384/
  186. 186. Aehle M, Kuprijanov A, Schaepe S, Simutis R, L?bbert A. Simplified off-gas analyses in animal cell cultures for process monitoring and control purposes. Biotechnol Lett. 2011 Nov;33(11):2103–10.
  187. 187. Chen LZ, Nguang SK, Li XM, Chen XD. Soft sensors for on-line biomass measurements. Bioprocess Biosyst Eng. 2004 Apr;26(3):191–5.
  188. 188. Kremer F, Schonhals A. Broadband Dielectric Spectroscopy [Internet]. Kremer F, Schönhals A, editors. Springer-Verlag, Berlin Heidelberg. 2003. 91 p. Available from: http://www.springer.com/us/book/9783540434078
  189. 189. F.Lvovich V. Impedance spectroscopy application to electrochemical and dielectric phenomena: First edition. Wiley; Hoboken N.;J 2012. pg. 3.
  190. 190. McAdams ET, Jossinet J. Tissue impedance: a historical overview. Physiol Meas. 1995;16(3 Suppl A):A1–13.
  191. 191. Kaiser C, Carvell JP, Luttmann R. A Sensitive, compact, in situ biomass measurement system controlling and monitoring microbial fermentations using radio-frequency impedance. Bioprocess Int. 2007 Jan;(2007):52–5.
  192. 192. Carvell JP, Dowd JE. On-line measurements and control of viable cell density in cell culture manufacturing processes using radio-frequency impedance. Cytotechnology. 2006 Jun;50(1–3):35–48.
  193. 193. Ducommun P, Kadouri A, Von Stockar U, Marison IW. On-line determination of animal cell concentration in two industrial high-density culture processes by dielectric spectroscopy. Biotechnol Bioeng. 2001 Feb;77(3):316–23.
  194. 194. Druzinec D, Weiss K, Elseberg C, Salzig D, Kraume M, Pörtner R, et al. Process Analytical Technology (PAT) in Insect and Mammalian Cell Culture Processes: Dielectric Spectroscopy and Focused Beam Reflectance Measurement (FBRM). In: Pörtner R, editor. Animal Cell Biotechnology Methods in Molecular Biology [Internet]. Humana Press; 2014. pp. 313–41. (Methods in Molecular Biology; vol. 1104). Available from: http://link.springer.com/10.1007/978-1-62703-733-4_20
  195. 195. Justice C, Brix A, Freimark D, Kraume M, Pfromm P, Eichenmueller B, et al. Process control in cell culture technology using dielectric spectroscopy. Biotechnol Adv. 2011 May;29(4):391–401.
  196. 196. Yardley JE, Kell DB, Barrett J, Davey CL. On-line, real-time measurements of cellular biomass using dielectric spectroscopy. Biotechnol Genet Eng Rev. 2000;17:3–35.
  197. 197. Markx GH, Davey CL. The dielectric properties of biological cells at radiofrequencies: Applications in biotechnology. Enzyme Microb Technol. 1999 Aug;25(3–5):161–71.
  198. 198. Davey CL, Davey HM, Kell DB, Todd RW. Introduction to the dielectric estimation of cellular biomass in real time, with special emphasis on measurements at high volume fractions. Anal Chim Acta. 1993 Jul;279(1):155–61.
  199. 199. Schwan HP. Electrical Properties of Tissue and Cell Suspensions. TOBIAS JHL and CA, editor. Adv Biol Med Phys [Internet]. 1957;5:147–209. Available from: http://www.ncbi.nlm.nih.gov/pubmed/13520431
  200. 200. Harris CM, Kell DB. The radio-frequency dielectric properties of yeast cells measured with a rapid, automated, frequency-domain dielectric spectrometer. J Electroanal Chem. 1983 Jan;156(C):15–28.
  201. 201. Davey CL, Markx GH, Kell DB. On the dielectric method of monitoring cellular viability. Pure Appl Chem. 1993;65(9):1921–6.
  202. 202. Cannizzaro C, Gügerli R, Marison I, von Stockar U. On-line biomass monitoring of CHO perfusion culture with scanning dielectric spectroscopy. Biotechnol Bioeng [Internet]. 2003 Dec 5;84(5):597?610. Available from: http://doi.wiley.com/10.1002/bit.10809
  203. 203. Druzinec D, Salzig D, Brix A, Kraume M, Vilcinskas A, Kollewe C, et al. Optimization of insect cell based protein production processes – Online monitoring, expression systems, scale up. Adv Biochem Eng Biotechnol. 2013;136:65–100.
  204. 204. Zeiser a, Bédard C, Voyer R, Jardin B, Tom R, Kamen a a. On-line monitoring of the progress of infection in Sf-9 insect cell cultures using relative permittivity measurements. Biotechnol Bioeng [Internet]. 1999 Apr 5;63(1):122–6. Available from: http://www.ncbi.nlm.nih.gov/pubmed/10099588
  205. 205. Zeiser A, Elias CB, Voyer R, Jardin B, Kamen AA. On-line monitoring of physiological parameters of insect cell cultures during the growth and infection process. Biotechnol Prog. 2000 Jan;16(5):803–8.
  206. 206. Elias CB, Zeiser A, Bedard C, Kamen AA. Enhanced growth of Sf-9 cells to a maximum density of 5.2 x 107 cells per mL and production of beta galactosidase at high cell density by fed batch culture. Biotechnol Bioeng. 2000 May;68(4):381–8.
  207. 207. Negrete A, Esteban G, Kotin RM. Process optimization of large-scale production of recombinant adeno-associated vectors using dielectric spectroscopy. Appl Microbiol Biotechnol. 2007 Aug;76(4):761–72.
  208. 208. Ansorge S, Esteban G, Schmid G. On-line monitoring of infected Sf-9 insect cell cultures by scanning permittivity measurements and comparison with off-line biovolume measurements. Cytotechnology. 2007 Oct;55(2–3):115–24.
  209. 209. Zitzmann J, Weidner T, Czermak P. Optimized expression of the antimicrobial protein Gloverin from Galleria mellonella using stably transformed Drosophila melanogaster S2 cells. Cytotechnology [Internet]. 2017 Jan 28 [cited 2017 Jan 31];1–19. Available from: http://link.springer.com/10.1007/s10616-017-0068-5
  210. 210. Junker BH, Reddy J, Gbewonyo K, Greasham R. On-line and in-situ monitoring technology for cell density measurement in microbial and animal cell cultures. Bioprocess Eng. 1994;10(5–6):195–207.

Notes

  • Definitions: A batch process is cell growth and protein production in a fix amount of growth medium. The fed-batch operation augments a batch processes by continuous or intermittent addition of growth medium to prevent nutrient depletion and to increase cell density and productivity. Chemostat processes involve the continuous replacement of culture medium including a withdrawal of medium with cells and product. This mode of operation is more suitable for kinetic studies rather than protein production. Perfusion is the continuous replacement of culture medium with an additional cell retention system. This mode of operation prevents nutrient depletion and cell drainage while maintaining a constant reaction volume. It allows high cell densities and product titers.

Written By

Jan Zitzmann, Gundula Sprick, Tobias Weidner, Christine Schreiber and Peter Czermak

Reviewed: 13 February 2017 Published: 10 May 2017