Cell Systems Review Cell Factory Engineering Anne Mathilde Davy,1 Helene Faustrup Kildegaard,2 and Mikael Rørdam Andersen1,* 1Department of Biotechnology and Biomedicine 2The Novo Nordisk Foundation Center for Biosustainability Technical University of Denmark, 2800 Kgs. Lyngby, Denmark *Correspondence: mr@bio.dtu.dk http://dx.doi.org/10.1016/j.cels.2017.02.010 Rational approaches to modifying cells to make molecules of interest are of substantial economic and scientific interest. Most of these efforts aim at the production of native metabolites, expression of heterologous biosynthetic pathways, or protein expression. Reviews of these topics have largely focused on individual strategies or cell types, but collectively they fall under the broad umbrella of a growing field known as cell factory engineering. Here we condense >130 reviews and key studies in the art into a meta-review of cell factory engineering. We identified 33 generic strategies in the field, all applicable to multiple types of cells and products, and proven successful in multiple major cell types. These apply to three major categories: production of native metabolites and/or bioactives, heterologous expression of biosynthetic pathways, and protein expression. This meta-review provides general strategy guides for the broad range of applications of rational engineering of cell factories. Introduction Cells engineered for the enhanced production of native compounds, or production of heterologous products is an established and economically important discipline. Serving as the basis of all product-oriented industrial biotechnology, the economic footprint of these cell factories ranges in the hundreds of billions of USD/year on the global markets: pharmaceutical proteins have been estimated at 140 billion USD in 2013 (Walsh, 2014); industrial enzymes in the range of 1.8 billion USD in 2009 (Waegeman and Soetaert, 2011); bio-derived non-protein pharmaceuticals 100 billion USD (Chemier et al., 2009); and bulk biochemicals (excluding biofuels) 58 billion USD (Nieuwenhuizen and Lyon, 2011). For comparison, the petrochemical industry is >3 trillion USD/year (2015), so there is still a large market to expand into. While industrial biotechnology has a long history, it was not until the arrival of genetic engineering that it became possible to modify the DNA of the cell factories to improve production (Figure 1), a process that hitherto had been based on clonal selection. Such developments gave rise to the discipline cellular engineering (Nerem, 1991), which covers both basic and applied cell research. The same year, Bailey defined metabolic engineering as a rational and directed process of engineering metabolism, rather than a cycle of trial and error (Bailey, 1991). Since then, the field of engineering cell factories has expanded in outlook and scope to include several ‘‘flavors’’ of cellular engineering specific to industrial biotechnology (Figure 1). Two examples are (1) inverse metabolic engineering (Bailey et al., 2002), in which one starts with the desired phenotype and works toward that goal by directed genetic or environmental manipulation, and (2) systems metabolic engineering as coined by the group of Sang-Yup Lee (Lee et al., 2007; Lee and Kim, 2015) as a term for large-scale holistic metabolic engineering. However, in many applications, the engineering efforts are not limited to the metabolic network of the cell, and therefore ‘‘metabolic engineering’’ does not fully encompass all activities. This is particularly true for the large sector of expression of protein, native and heterologous, which covers the range from bulk enzymes to formulated pharmaceutical proteins. Here, engineering targets can be within cellular machinery such as the protein secretion pathway. To include all of these activities, this meta-review defines cell factory engineering, encompassing all rational approaches to improve a cell factory. The objective of this meta-review is specifically not to present a comprehensive list of examples within the individual strategies, nor is it to present direct strategies for target identification, such as modeling in tandem with predictive algorithms (Ranganathan et al., 2010; Burgard et al., 2003; Pharkya and Maranas, 2006). For this, specialized reviews of high quality and information content already exist (see, e.g., the excellent and recent work of the group of Sang-Yup Lee; Lee and Kim, 2015). In this text, we provide a meta-review summarizing several years of cell engineering efforts, in essence an applicable list of strategies generally applicable across species and products suitable for the experienced scientist. In this, we have focused on strategies applied reproducibly across multiple cell factories, and chosen the most applied microbial cell factories from the entire Tree of life, spanning bacteria, yeast, filamentous fungi, and mammalian cells (in particular Chinese hamster ovary [CHO] cells), as well as some higher fungi. See also Box 1 for an overview of cell factory engineering methods in other fields. This meta-review provides representative examples of applications of the strategies for illustration. Meta-review Overview The analysis here draws on a long list of reviews supplemented by primary literature to provide an overview of cell factory engineering. Table 1 lists the reviews cited in this text and annotations on which types of strategies and organisms these reviews are most relevant for. It is the reductionist argument of this meta-review that nearly all cell factory engineering efforts can be classified in one of the following three categories or as combinations of them: (1) optimization of the production of a metabolite in the native 262 Cell Systems 4, March 22, 2017 ª 2017 Elsevier Inc. host; (2) production of a non-native metabolite by expression of a heterologous biosynthetic pathway; and (3) expression of a heterologous protein. Here, we condense the strategies of the reviews in Table 1 into these three generic categories, examine each in detail, and provide guidance on choosing and applying individual strategies. Production of Native Metabolites Native metabolites are here compounds naturally produced by the cell factory, either intracellularly or (preferably) a secreted compound. Examples are amino and nucleic acids, antibiotics, vitamins, enzymes, bioactive compounds, and proteins produced from anabolic pathways of cells (see details for protein products further below). Common for these are that they cannot be synthetically produced or for which it is not economical to do so (Stephanopoulos and Vallino, 1991). This has been examined for specific cells or products in a multitude of excellent reviews (see, e.g., Bailey et al., 2002; Pickens et al., 2011; Stephanopoulos and Vallino, 1991; Keasling, 2008, 2012; Hwang et al., 2014; Wu et al., 2014; Weber et al., 2015; Kiel et al., 2010; Xiao and Zhong, 2016). Here, we provide an overview of general strategies to increase the formation of native metabolites (Figure 2). The strategies one would apply to this problem can be reduced to ten types (Figure 2, 1A–1J). 1A1-2. Pathway overexpression: Using this strategy, one would typically overexpress one or more enzymes in the biosynthetic pathway. It is a common strategy and is often achieved by overexpressing the native genes (1A1). As an alternative to normal overexpression, enzymes could be engineered to have higher activity. In either case, it can be advantageous to identify enzymatic steps with particular control of the flux to the product, such as irreversible reactions, or the first steps in the pathway. Some steps in the pathway (often the latter) may have very little control over the flux, so multiple targets should be engineered and/or metabolic control analysis (Nielsen, 1998) should be employed. It has also been seen that heterologous expression of ortholog enzymes from related species (1A2) can have a larger effect than the native enzymes. The reason for this remains speculative, but one hypothesis could be a lower regulatory effect on the heterologous proteins. One example of the latter is enhanced citrate production in Aspergillus niger by heterologous expression of tricarboxylic acid cycle enzymes from Saccharomyces cerevisiae and Rhizopus oryzae (de Jongh and Nielsen, 2008), or improved ganoderic acid accumulation in Ganoderma lucidum (Xu et al., 2012). 1B1-2. Transporter engineering: Accumulation of the product in the cell can decrease the carbon flux by affecting enzyme kinetics, thus decreasing production rates and yields. Furthermore, accumulation of product can trigger feedback inhibition, which will severely limit the flux through the pathway. In some cases, the product may even be toxic. Overexpressing product efflux pumps can thus be an efficient way of increasing the flux (1B1) (Dunlop et al., 2011; Wu et al., 2014; Lee et al., 2012; Lee and Kim, 2015). One example is the improved production of biofuels in Escherichia coli by systems engineering of 43 efflux pumps (Dunlop et al., 2011). This strategy both increased production and lowered product toxicity. Alternatively, or in combination with 1B1, gene knockouts of uptake transporters specific for the product (1B2) can also be effective (Lee et al., 2012). 1C. De-branching: Branching or competing pathways can decrease the overall flux toward the product (Pickens et al., 2011; Lee et al., 2012; Pfleger et al., 2015). If these pathways are not lethal, deleting the first branching step may improve product formation. With essential pathways, decreasing the activity by knockdown or, e.g., tunable promoters can be an alternative option. This is a common strategy; one comprehensive example includes the knockout of L-lysine, L-methionine, and L-glycine biosynthesis for improved isoleucine production in E. coli (Park et al., 2012). 1D. Product degradation: Any non-essential reactions that convert the product to unwanted metabolites should be deleted, as these may degrade the product and decrease yields and titers. Such an example is the work of Lee et al. (2007), where threonine dehydrogenase was deleted in E. coli to increase the production of L-threonine. 1E1-3. Co-factor engineering: In some cases, it has been shown that a major limitation is the availability of co-factors (NADH/NAD+ , NADPH2, NADP+ , acetyl-CoA, etc.) (Lee and Kim, 2015; Ghosh et al., 2011; Lee et al., 2012; Pfleger et al., 2015; van Rossum et al., 2016). In these cases, one must make more co-factors available by engineering other pathways. Ideally, the deletion of a non-essential enzyme, which catabolizes large amounts of the co-factor, is preferred (1E1). In cases where this is not possible, an alternative might be replacing such an enzyme with a native or heterologous enzyme with the same function, but specific for another cofactor (1E2). An example of this is substitution of a native NADPH-dependent glutamate dehydrogenase with an overexpressed NADH-dependent glutamate dehydrogenase to enhance sesquiterpene production in S. cerevisiae (Asadollahi et al., 2009). A third option is the insertion (Yamauchi Figure 1. Ontology of Different Types of Cellular Engineering Covered in This Review Cell Systems 4, March 22, 2017 263 Cell Systems Review et al., 2014) or overexpression (Cui et al., 2014) of an E. coli transhydrogenase for interconversion of NADH and NAPDH (1E3). 1F. Removal of feedback inhibition: In many cases, especially with products that are a part of standard growth metabolism (e.g., amino acids), strong feedback inhibition exists to tightly regulate the concentrations of the product. When one wishes to produce such compounds in large amounts, it can be necessary to disable feedback inhibition. Often this is achieved by random or targeted mutagenesis of enzymes in the pathway known to be feedback inhibited (Lee and Kim, 2015). In some cases, analogs of the product, which bind tightly/near irreversibly to the regulated enzymes, can be used to screen for feedback-deregulated mutants. This has been used, e.g., for L-threonine (Lee et al., 2003), and L-tryptophan and L-serine (Rodrigues et al., 2013), both in E. coli. This strategy was also efficient for engineering acid production in A. niger (de Jongh and Nielsen, 2008) and for production of fatty acids in E. coli (Pfleger et al., 2015). 1G. By-product elimination: Several species produce varying amounts of by-products. Often these by-products, while not directly linked to the metabolic pathway of the product, compete with the product for available carbon and/or co-factors (Lee et al., 2012). If possible without making lethal deletions, the enzymatic activities producing such compounds should be deleted or reduced. Numerous successful examples of this strategy can be found, for instance removal of glycerol biosynthesis in S. cerevisiae for increased ethanol production (Wang et al., 2013). 1H. Precursor/substrate enrichment: It will often be advantageous to increase the availability of the substrate for the product biosynthesis (Lee and Kim, 2015; Pickens et al., 2011; Lee et al., 2012). This can be achieved by a multitude of strategies, essentially by considering the substrate as an intermediate product, and applying one or more of strategies 1A–1J to increase substrate formation. When considering substrates, one should also remember to take co-substrates such as acetyl-CoA in account (see, e.g., a recent review of acetyl-CoA engineering in S. cerevisiae; Nielsen, 2014). Other carbon donors can also become limiting, e.g., malonyl-CoA and glucose-1-phosphate in the production of an anti-cancer compound in Streptomyces argillaceus (Zabala et al., 2013). 1I. De-regulation of carbon catabolism: In some cases, the pathway of interest may be subject to general metabolic regulation of the cell, e.g., general regulators of carbon catabolism or nitrogen source-induced regulation. Examples of this is de-regulation of galactose metabolism in S. cerevisiae by deletion of negative regulators, leading to de-repression and increased galactose utilization (Ostergaard et al., 2000) or disruption of a global regulator in Pichia guilliermondii to trigger aerobic glucose catabolism for ethanol production (Qi et al., 2014). 1J. Signal transduction engineering: In some cases, the production of specific metabolites is not regulated by carbon or nitrogen sources (1I), but may be subject to signals from, e.g., micronutrients, or from other steps in the pathway. In these cases, engineering signal transduction can be a strong strategy (Kiel et al., 2010). Choosing a Strategy for Producing Native Metabolites Generally, there is a logical order in which to apply strategies 1A– 1J. The strategies can be sorted into three categories, which we suggest to apply in progression. Step 1: Direct Optimization of the Pathway in Any Way Possible. The main goal of this step is to ensure that neither enzymes nor intermediates of the pathway are limiting production. If this is not achieved, the other strategies may not be effective. This can be addressed by the following actions in roughly this order: i. Overexpression of the biosynthetic pathway using the strategies of 1A. This ensures that the concentrations of the enzymes are not limiting. ii. Enrichment for the substrates (1H) and for the co-factors (1E), thus ensuring that the required metabolites, precursors, and co-factors do not become limiting. iii. Ensuring that the product is removed from the cell by transporter engineering (1B) if possible. Accumulation of the product can seriously decrease product formation as enzyme kinetics are dependent on concentrations of the product. Furthermore, product accumulation can in some cases lead to feedback inhibition of the entire pathway. iv. If feedback inhibition is known for the pathway, this should be engineered out if possible, or removed by mutagenesis, screening and reverse genetics (1F).Again, this may not be a problem if actions i–iii are limiting. Step 2: Remove Competing Activities. Once the pathway itself is optimized, the next steps is to ensure that no other pathways are impairing the product formation, either directly by sharing metabolites or co-factors, or by using carbon which could be converted to product. The three main strategies here are as follows: i. De-branching (1C). Any pathway that shares intermediates or precursors with the pathway of interest should be deleted if possible. ii. Product degradation (1D). A particular case of 1C is pathways converting/degrading the product of interest. These should also be deleted if possible. iii. Removal of by-products (1G). While by-products are not often directly associated with the product pathway, by-products will use carbon, co-factors, and energy which could be converted into product. Step 3: Application of Global Regulation Engineering. This does not seem to be common strategies, as it will often be highly effective to perform the actions above. However, should this be in place, engineering carbon repression (1I) or similar signal transduction pathways (1J) can be a final approach. Clearly, the actions above can, and should be, combined for increased effects. Prime examples of this are large-scale rational design of metabolic pathways, which has been applied to great effect in several systems, in particular in bacterial hosts (Lee et al., 2007; Rodrigues et al., 2013; Becker et al., 2013) and yeast (Wu et al., 2014; Lee et al., 2012). Heterologous Expression of Biosynthetic Pathways When trying to produce an interesting compound, one of the most important decisions is the choice of production in the native 264 Cell Systems 4, March 22, 2017 Cell Systems Review host, and optimize this host, or transfer of the pathway to another well-known host. If the original host can be adapted to an industrial fermentation process, and there are no health-related risks in doing so (e.g., production of toxic by-products), this can be a preferred strategy (as was the case e.g., for penicillin). However, in many modern cases, the potential of using an industrially preferred cell factory and related platform processes out-weighs the difficulty of transferring the pathway. In some cases, transplanting a pathway also removes metabolic inhibition found in the original host (Martin et al., 2003). In this section, we examine how one may need to adapt the cell factory in order to accommodate production of a heterologous product. In general, several excellent specialized reviews exist within this area, and for additional details on specific cases for particular groups of compounds, we recommend the following studies: Pickens et al. (2011), Pfeifer and Khosla (2001), Lee et al. (2012), and Xiao and Zhong (2016). Here, we give an overview of common and general problems regarding heterologous expression of pathways. Innate differences between the native host and the cell factory of choice are major challenges in expressing a pathway in a new host. In general, the compatibility of the enzymes and metabolites with the new host should be considered. The more complicated the pathway, the larger the advantage of choosing a more closely related host. Major types of challenges are shown in Figure 3. These challenges can be condensed into points 2A–2F below. Note that these cover both eukaryotic and prokaryotic hosts and donors in combination, meaning that some of these are specific to certain types of cells (e.g., intracellular compartmentalization is seldom a problem in prokaryotes). 2A. Compartmentalization or steric proximity: In heterologous expression, a common pitfall is not making sure that the pathway is expressed in the same compartment as the substrate metabolite. If the heterologous enzymes do not contain targeting signals, in a eukaryotic host they will be expressed in the cytosol. In the case where the substrate is in another compartment, targeting sequences or gene fusions can be applied to direct the heterologous enzyme(s) to the correct compartment (Siddiqui et al., 2012). As an alternative, synthetic scaffolds have been made to bring biosynthetic enzymes together with great effect in both E. coli (Dueber et al., 2009) and S. cerevisiae (Wang and Yu, 2012). 2B1-2. Co-factor availability: Any overexpressed pathway will present a significant drain on available co-factors (van Rossum et al., 2016). It is advantageous to ensure that these are present in sufficient amounts in the host (2B1), as shown, e.g., in Streptomyces coelicolor (Borodina et al., 2008). This may be specific to the compartment (2B2). Alternatively, transhydrogenases may be engineered as described in 1E1-2. 2C1-2. Substrate and co-substrate availability: In addition to co-factors, one must also ensure that the host produces all substrates and co-substrates/precursors required for the pathway (2C1). It may also be the case that the host produces similar compounds, which may be competing for the substrate or precursors. In these cases, it can be advantageous to delete the competing pathways (Baltz, 2016). It has been demonstrated in, e.g., E. coli (Rodrigues et al., 2013, 2014) and Corynebacterium glutamicum (Becker et al., 2013), that high availability of the substrate in the heterologous host improves productivity. If all (co-)substrates are not available or present in low amounts, it is necessary to insert or overexpress biosynthetic genes for these as well (2C2), Examples of this are seen for, e.g., amino acids or oxaloacetate (Kind et al., 2010; Rodrigues et al., 2013) or adipic acid (Yu et al., 2014). 2D1-2. Product efflux pumps: When adding biosynthesis of a new compound to a cell, specific transporters for that compound may not exist. Accumulation of the product in the cell will decrease the flux through the biosynthetic pathway (Lee et al., 2012) and may also have toxic effects on the cell (Pfeifer and Khosla, 2001). Passive transport or unspecific transporters may be available, but if this is not the case a specific transporter must be added (2D1,) as seen for, e.g., flavonoids (Wu et al., 2014) or cadaverine (Qian et al., 2011). Should the pathway be compartmentalized, this also needs to be accounted for, possibly by expressing an organellespecific transporter (2D2), e.g., with mitochondrial products (Chen et al., 2015). 2E. Biosynthesis of functional groups: For a number of proteins, all functional groups are not encoded by the gene, Box 1. Research Applications of Cell Factory Engineering The focus of the current review is on cell factory engineering for biotechnological applications. However, there are many other applications of cell factory engineering in life sciences and medicine. Besides industrial applications, a main application of cell factory engineering is to study the biological function of genes and proteins in basic research. For this, genome engineering and synthetic biology tools can be applied to regulate and remove current gene function or introduce new function, followed by analyzing the effect on cellular functions including biochemical reactions, regulatory networks, or cell phenotypes (Hsu et al., 2014; Bashor et al., 2010). An example is engineering of genes involved in glycosylation to study their function in generating certain glycoforms that can be applied to achieve homogenized glycoforms on recombinant proteins for comparative studies of their biological effect (Yang et al., 2015). Cell factory engineering is also often applied in generating reagents for research. Examples include expression of antibodies to obtain reagents for genetics studies (Barnstable et al., 1978), hormones to obtain reagents for immunoassays (Ribela et al., 1996), and purified proteins for structural analysis by crystallographers and nuclear magnetic resonance spectroscopists (Edwards et al., 2000). In addition, the products from cell factory engineering can be applied in screening for drug activity or as a potential drug target for medical applications (Trosset and Carbonell, 2015). This also includes engineering of natural probiotics to produce valuable compounds for enhancement of their benefit to the host (Behnsen et al., 2013). Cell Systems 4, March 22, 2017 265 Cell Systems Review but require separate biosynthesis. One example is heme groups, found in multiple types of enzymes requiring oxygen as a co-factor. Heme groups are not found in all prokaryotes (Cavallaro et al., 2008), and may be limiting in some fungal systems (Franken et al., 2011). Another functional group is Fe-S clusters, which have several different biosynthetic pathways specific to the type of host organism. Fe-S clusters are synthesized in the cytoplasm of bacteria and in the mitochondrion of eukaryotic microbes, from where they are transported into the cytosol. In order for the heterologous pathway Table 1. Overview of Reviews Covered in This Text Reference Species Strategies Products Walsh (2014) bacteria, yeasts, fungi, mammalian cells protein expression biopharma * Waegeman and Soetaert (2011) E. coli protein expression enzymes and biopharma * Chemier et al. (2009) E. coli, S. cerevisiae native and heterologous pathways bioactives, biofuels ** Lee and Kim (2015) bacteria, yeasts native and heterologous pathways any *** Pickens et al. (2011) bacteria, fungi heterologous pathways bioactives *** Hwang et al. (2014) Streptomyces sp. native pathways bioactives ** Wu et al. (2014) E. coli, S. cerevisiae heterologous pathways flavonoids ** Anyaogu and Mortensen (2015) fungi heterologous pathways bioactives * Xiao and Zhong (2016) basidiomycetes heterologous pathways terpenoids * Nielsen (1998) bacteria, yeasts, fungi, mammalian cells all any ** van Rossum et al. (2016) S. cerevisiae native and heterologous pathways acetyl-CoA-derived compounds ** Lee et al. (2012) E. coli native and heterologous pathways small molecules *** Nielsen (2014) S. cerevisiae native and heterologous pathways acetyl-CoA-derived compounds * Pfleger et al. (2015) E. coli, S. cerevisiae native and heterologous pathways oleochemicals ** Pfeifer and Khosla (2001) S. coelicolor, E. coli, yeasts, fungi native and heterologous pathways polyketides ** Siddiqui et al. (2012) yeasts heterologous pathways bioactives *** Franken et al. (2011) fungi heterologous pathways heme * Baltz (2016) actinomycetes heterologous pathways bioactives ** Bekiesch et al. (2016) actinomycetes heterologous pathways bioactives * Fisher et al. (2014) bacteria, yeasts, fungi heterologous pathways any ** Mattanovich et al. (2012) yeasts protein expression biopharma, enzymes ** Hossler (2012); Hossler et al. (2009) mammalian cells protein expression biopharma ** Andersen et al. (2011) mammalian cells protein expression biopharma ** Damasceno et al. (2012) P. pastoris protein expression proteins *** Celik and Calık (2011) yeasts protein expression proteins *** Ward (2011) fungi protein expression proteins *** Wurm (2004) mammalian cells protein expression biopharma * Berkmen (2012) E. coli protein expression proteins with disulfide bonds * de Marco (2009) E. coli protein expression proteins with disulfide bonds * Schro¨ der (2008) eukaryotes protein expression biopharma ** Fleissner and Dersch (2010) Aspergillus protein expression enzymes ** Lubertozzi and Keasling (2009) Aspergillus native and heterologous pathways, protein expression small molecules, proteins * Vogl et al. (2013) P. pastoris protein expression biopharma * Rosano and Ceccarelli (2014) E. coli protein expression proteins ** Liu et al. (2013) Gram-positives protein expression proteins ** Gasser et al. (2008) bacteria, yeasts, fungi protein expression proteins ** Asterisks on the right denote the general relevance of the text for cell factory engineering. 266 Cell Systems 4, March 22, 2017 Cell Systems Review to be functional, it may be required to express the native biosynthesis pathway heterologously (2E). A specific type of Fe-S proteins (ferredoxins) mediates electron transfer. Cases exist where the expression of specific ferredoxins from the native host were necessary for optimal expression of the pathway (Molna´ r et al., 2006). 2F. Transcription engineering: With many secondary metabolites, several genes are required to act in concert to form the product. If only one or a few genes are active, the product may be absent or different from the expected product. For many of these pathways, one regulatory protein exists, which transcriptionally activates the entire pathway. If one uses native promoters to express the genes, it can be advantageous to overexpress the regulatory protein (if it can be identified), and thereby induce the entire set of genes (Pickens et al., 2011; Baltz, 2016; Bekiesch et al., 2016). Examples of this include the transplant of the geodin gene cluster from Aspergillus terreus into Aspergillus nidulans and substitution of the native promoter for the transcription factor for a strong constitutive promoter, thus allowing heterologous expression of geodin (Nielsen et al., 2013). Choosing a Strategy for Heterologous Pathway Expression For heterologous pathways, the strategy is a combination of the issues encountered in the expression of native pathways, and issues arriving from interaction with the new host. Roughly, the considerations can be sorted into two major steps. Step 1: Compatibility of the Pathway to the Host. The actions listed in this step are interesting in that they may not be needed, dependent on the interaction with the host. Appropriate host selection can thus be used already in the design phase to remove or minimize the problems (for a few reviews on host selection, see, e.g., Fisher et al., 2014; Lee and Kim, 2015; Bekiesch et al., 2016). However, if these are not considered, no other engineering strategies may be effective. The three main things to consider are thus: i. Compartmentalization (2A). Spatial co-localization of the inserted enzymes as well as availability of co-factors and precursors in the compartment(s) of choice. ii. Functional group biosynthesis (2E). Ensuring functionality of all enzymes. iii. Substrates, co-substrates (2C), and co-factors (2B). Ensuring that all required precursors are available in the host. Step 2: Optimization of the Pathway. Once it is ensured that the pathway is functional in the host, one can apply strategies to increase flux through the pathway. Here the following five steps should be investigated, sorted in order of perceived importance. Figure 2. Engineering Strategies for Optimizing the Production of Native Metabolites Metabolites are denoted: S, substrate; S2, alternative substrate; X1-9, pathway intermediates; P, product of interest; BP, by-product. Generic engineering strategies are marked in blue and annotated as 1A–1I (described in the main text). Cell Systems 4, March 22, 2017 267 Cell Systems Review i. Application of transcription engineering where possible (2B). Increased transcription of all enzymatic steps is an efficient way to increase enzyme levels. ii. Pathway overexpression strategies (1A1) and removal of feedback inhibition (1F) are equally applicable to heterologous pathways. iii. Removal of competing activities as described in 1C and 1D. This is particularly interesting when producing a compound, where the host produces several similar compounds competing for the precursors, e.g., within microbial bioactive compounds (Pickens et al., 2011). iv. Improving the product efflux by transporter engineering (2D and 1B). v. Removal of by-products (1G) can possibly be considered last, as the strategies above are more direct toward improving the pathway. However, by-products removal has been seen to have importance here (Wu et al., 2014; Pickens et al., 2011). In summary, the overview above provides a strategy guide for heterologous pathway expression encompassing many different reviews and studies. However, it is important to note that this does not cover host-specific or donor-specific problems. In these cases, we direct the reader to Table 1 to find suggestions for additional species-specific engineering challenges. Protein Expression The expression of proteins, both homologous and heterologous, is presently done in a wide variety of hosts from E. coli and Bacillis subtilis, over yeasts, e.g., Klyuveromyces lactis, Pichia pastoris, and S. cerevisiae, through filamentous fungi such as A. niger, to cells derived from multicellular organisms such as mammals and insects. The variety of proteins of commercial interest is great, ranging from bulk enzymes to complex biopharmaceuticals (Association of Manufacturers and Formulators of Enzyme Products, 2009; Walsh, 2014). Due to the diverse properties of proteins, it is currently not possible to use one platform organism for expression of all proteins. The scientist must thus choose the cell factory based on the properties and applications of the desired protein. The advantages and disadvantages of applying different cell factories are discussed in several excellent reviews specialized to particular expression systems (see Table 1). In particular we recommend the review of Waegeman and Soetaert (2011) for a very clear comparison of expression systems in addition to a thorough overview of E. coli expression. Despite the variety of employed systems, there are generic strategies applicable to high-yield expression of proteins. Not all of the strategies presented here are applicable in every host, but we focus on strategies, which are applicable in several hosts. For this reason, the present review does not discuss strategies related to the accumulation of protein in inclusion bodies, a feature encountered in some bacterial hosts such as E. coli for some proteins with particular folding. For this, we again direct the reader to specialized overviews (de Marco, 2009; Waegeman and Soetaert, 2011). Overall, the successful high-yield process for production of a given protein requires high transcription and translation of the Figure 3. Engineering Strategies for Heterologous Expression of Biosynthetic Pathways in a Eukaryotic Host Cell Metabolites are denoted: S, Substrate; X1-6, pathway intermediates; P, product of interest; CS, co-substrate; E1-10, enzymes; FG, functional group. The inserted heterologous pathway is marked in green. Generic engineering strategies are boxed in blue and annotated as 2A–2F (described in the main text). 268 Cell Systems 4, March 22, 2017 Cell Systems Review gene, successful targeting of the protein to the secretion pathway (if secretion is desired), correct folding and limited induction of secretion stress, the desired post-translational modifications, efficient secretion, and limited or no degradation of the product in the medium. In general, the major strategies for engineering increased protein expression can be found in Figure 4, and are summarized in points 3A–3F below. 3A1-2. Promoter Engineering: Nearly all systems aim at ensuring maximal availability of recombinant mRNA so that this is not a bottle neck for protein expression. The major strategy employed is addition of a highly expressed constitutive promoter (3A1). A selection of these are known for most hosts such as the native GAPDH promoter in yeasts (Mattanovich et al., 2012), the heterologous gdhA promoter in Aspergillus species (Fleissner and Dersch, 2010), or viral promoters in mammalian hosts (Wurm, 2004). It is also a common strategy to develop synthetic promoters (Dehli et al., 2012; Vogl et al., 2013; Fleissner and Dersch, 2010). Alternatively, one can employ strong inducible promoters (3A2) to have a biphasic process (Waegeman and Soetaert, 2011; Fleissner and Dersch, 2010), for instance methanol-inducible gene expression in the methylotrophic yeast P. pastoris (Mattanovich et al., 2012; Damasceno et al., 2012). Reviews with particularly good overviews of promoters for specific systems are available (Celik and Calık, 2011; Fleissner and Dersch, 2010). A complementary strategy to the use of strong promoters is the expression from high-copy plasmids (Rosano and Ceccarelli, 2014) or multigene insertions (Westwood et al., 2010; Wurm, 2004; Damasceno et al., 2012). Other transcriptional elements such as enhancers, transcription factor binding, and chromosomal elements should be considered dependent on expression systems (Liu et al., 2013; Fleissner and Dersch, 2010; Westwood et al., 2010). 3B1-2. Gene Fusion for Enhanced Secretion: For proteins with no inherent secretion signal, the gene sequence requires engineering to facilitate secretion of the protein. The predominant way is the addition of a signal peptide/secretion leader signal (3B1). This can also be applied to substitute the original signal peptide for improved secretion in the host. For the major hosts, efficient signal peptides are known from native secreted proteins, e.g., alpha-mating factor or acid phosphatase in yeasts (Mattanovich et al., 2012; Damasceno et al., 2012) or leader sequences from secreted proteins in Aspergillus (Fleissner and Dersch, 2010) or bacteria (Liu et al., 2013). In some combinations of host and protein, this may not be sufficient; in which case, the gene of interest is fused with the sequence for a carrier protein (3B2), which then has the effect of ushering the protein out of the cell. One example is the production of animal proteins in Aspergillus species, where a successful strategy for bovine chymosin production was fusion with the glucoamylase gene. This approach has since then become a preferred method (Ward et al., 1990; Ward, 2011; Fleissner and Dersch, 2010). 3C. Stability of Heterologous Gene Transcripts: Most eukaryotic genes contain introns. In many cases, their removal from the transcript is necessary to generate a functional gene product due to differences in (or absence of) splicing Figure 4. Overview of Generic Engineering Strategies for Expression of Proteins in a Eukaryotic Host Cell The inserted gene and its derived mRNA and polypeptide are marked in green. Generic engineering strategies are marked in blue and annotated as 3A–3J (described in the main text). Cell Systems 4, March 22, 2017 269 Cell Systems Review machinery between species (Hamann and Lange, 2006; Innis et al., 1985). In higher eukaryotic systems a single intron early in the transcript or in the promoter can, however, successfully enhance stability of mRNA and increase the final product titer (Borkovich et al., 2004). In many cases, codon optimization of heterologous transcripts are often needed due to incompatibility between the host and the protein codons, e.g., use of rare codons or difference in stop codons. In general, the half-life of a heterologous transcript might be different from related transcripts of the host. In a bacterial system, the importance of terminators and 30 UTR regions to transcript stability has been well established (Cambray et al., 2013; Pfleger et al., 2006; Curran et al., 2013). Often changing natural or adding new structures, e.g., hairpin structures, to the ends of transcripts, have been shown in bacteria to accumulate mRNA and increase product formation (Hienonen et al., 2007). In yeast and fungal systems, recent studies show that changing a terminator can effectively optimize the transcript stability and increase the product titer (Curran et al., 2013). 3D1-2. Improved Translocation to the Endoplasmic Reticulum: The induced pressure on the secretion machinery creates numerous rate-limiting steps. The first is already at the entrance of the secretion pathway through translocation (3D1). A successful approach for several systems is overexpression of signal peptidases cleaving the signal peptide by entrance to the endoplasmic reticulum (ER) (Meta et al., 2009; van Dijl et al., 1991; Ailor et al., 1999). Insufficient amount of proteolytic cleavage enzymes may also be limiting for secreted proteins with precursor domains (3D2). An example is for therapeutic protein produced in CHO cells, where overexpression of the cleaving enzyme PACE, increase the secretion capability for several different proteins (Sathyamurthy et al., 2012). 3E1-2. Protein Secretion Stress Engineering: It is generally found that overexpression of proteins induces protein secretion stress to some degree, which decreases productivity and overall cell fitness (Lubertozzi and Keasling, 2009; Gasser et al., 2008; Schro¨ der, 2008). One generally applied strategy is the overexpression of chaperones (3E1). This strategy has been proven to be successful in several studies in a multitude of systems: E. coli (Waegeman and Soetaert, 2011; Gasser et al., 2008; Rosano and Ceccarelli, 2014), other bacteria (Gasser et al., 2008), yeasts (Mattanovich et al., 2012; Gasser et al., 2008), fungi (Ward, 2011; Fleissner and Dersch, 2010), and CHO cells (Ailor and Betenbaugh, 1998; Josse´ et al., 2012; Pybus et al., 2013). It has also been broadly successful in regulating global activators of the ER or the unfolded protein response (3E2), in bacteria (Gasser et al., 2008), S. cerevisiae (Valkonen et al., 2003b; Mattanovich et al., 2012; Calfon et al., 2002), in A. niger var. awamori (Valkonen et al., 2003a; Carvalho et al., 2012; Fleissner and Dersch, 2010), and in mammalian cells (Ohya et al., 2008; Tigges and Fussenegger, 2006; Ku et al., 2008). 3F. Engineering the Post-translational Modification Machinery: In some cases, the bottlenecks are in the formation of disulfide bridges (Schro¨ der, 2008). This has been a problem in E. coli in particular (de Marco, 2009), but proteins involved in disulfide bridge formation have been seen to be limiting in many cases, as seen by the positive effect of protein disulfide isomerase in many other organisms, such as several yeasts, Aspergillus (Gasser et al., 2008; Fleissner and Dersch, 2010), and CHO (Borth et al., 2005; Davis et al., 2000; Mohan et al., 2007). 3G. Improved Vesicle Trafficking: Another rate-limiting step is the vesicle trafficking between ER-Golgi and Golgi-membrane. Overexpression of SNAREs and their key regulators can stimulate vesicular trafficking in yeast and enhance heterologous protein secretion (Hou et al., 2012; Ruohonen et al., 1997). Vacuolar protein sorting is complex, illustrated by disruption of the vacuolar protein sorting receptor, Vsp10p, which has a positive impact on secreted protein in both filamentous fungi and yeast (Yoon et al., 2010; Idiris et al., 2010). 3H. Protein Glycosylation Engineering: This discipline does not directly aim to improve production rate or titer of the product, but instead addresses protein quality, in the form of protein glycosylation. This has two branches, one where it is sought to optimize the native protein glycosylation, and one where the host organism does not have the required protein glycosylation features, and these are engineered into the cell factory (Mattanovich et al., 2012; Hossler, 2012; Hossler et al., 2009; Andersen et al., 2011; Vogl et al., 2013). E. coli, like most prokaryotes, does not have native protein glycosylation, but genes from other prokaryotes with protein glycosylation have successfully been engineered into the host (Waegeman and Soetaert, 2011). Protein glycosylation has also been engineered in filamentous fungi (Ward, 2011). A very ambitious example is the expression of major parts of the human glycosylation pathway in P. pastoris (Li et al., 2006; Hamilton et al., 2006; Choi et al., 2003; Damasceno et al., 2012), a technology later acquired by Merck (Walsh, 2010). 3I. Protease Deletions: The deletion of extracellular proteases has been pursued in many systems with significant effects (Ward, 2011; Fleissner and Dersch, 2010). Examples include the deletion of all 25 known proteases in E. coli (Meerman and Georgiou, 1994), S. cerevisiae (Tyo et al., 2014), and the deletion of five proteases in A. oryzae (Jin et al., 2007). Another strategy, with effects in several Aspergillus species, has been the identification and deletion of a global regulator of protease expression, PrtT. Deletion of this gene eliminates nearly all protease activity (Punt et al., 2008; Fleissner and Dersch, 2010). 3J. By-product Removal: A final general strategy, in all species, is the removal of by-products with negative effect on protein production. Examples include removal of acetate biosynthesis in E. coli (Waegeman and Soetaert, 2011), oxalic acid production in Aspergilli (Li et al., 2013), or lactate production in CHO cells (Kim and Lee, 2007). All of these have been shown to improve product formation, growth characteristics or both. Choosing a Strategy for Protein Expression Contrary to the strategies for production of smaller compounds, where the yield and titer of the product is the primary optimization criterion, it is more difficult to define a generalized order of engineering strategies for protein products. The main reason is, that for some proteins, in particular pharmaceuticals, quality is more important than quantity. In some cases, quantity even has a detrimental effect on quality, as it may elicit stress 270 Cell Systems 4, March 22, 2017 Cell Systems Review responses in the cell which degrade the product (Wurm, 2004; Damasceno et al., 2012; Hossler, 2012; Hossler et al., 2009). Therefore, we propose two strategies, one for optimizing titers (e.g., for enzymes and bulk products), and one for products focused on quality (i.e., pharmaceuticals). Strategy A: Optimal Expression of the Heterologous Gene. Here, multiple initiatives can be used separately, sequentially, or in parallel, to find the strategy that is the most efficient. The following six actions are thus applicable only in the cases where that factor is limiting. In general, actions 3A–3C in particular are relatively consistently applied in successful studies. i. Selection and engineering of optimal promoters (3A) are vital for high levels of transcript, so this does not become a limiting factor. ii. Engineering of the heterologous gene in regard to codon compatibility and optimality and removal or adaptation of introns (3C) are also found in nearly all studies. iii. Selection and/or engineering of the secretion signal (3B) is required to ensure secretion of the product, and appropriate trafficking of the peptide chain to the ER. This can affect the production by severalfold. iv. Protein secretion stress reduction (3E), in particular regarding the formation of disulfide bridges, generally increases product formation. v. As is seen for small molecules (1D), removal of product degradation improves productivity. For proteins, this is solved by protease deletions (3I). vi. Finally, it has been shown that engineering vesicle trafficking (3G) and translocation to the ER (3D) increases productivity. However, this is in only few cases, possibly due to the complexity of engineering these processes. It thus becomes difficult to evaluate how applicable this is in general. Strategy B: Protein Quality. For optimization of protein quality, the strategy depends on which quality criterion is suboptimal in the production process, and a first step should thus be the determination of this. Here, analytical biochemistry will be the primary tool, and thus not within the scope of this meta-review. Once it has been established, one can apply one or more of the following three engineering types: d Protein glycosylation engineering (3H) is generally very attractive for glycosylated biopharmaceuticals (Walsh, 2014; Ratner, 2014). d Engineering disulfide bridge formation (3F) and protein folding(3E)ingeneralcanhelpremoveerroneouslyfolded protein and decrease protein folding-associated stress. d Protease deletions (3I) are just as important for maintaining protein quality as quantity. In addition to the strategies of this section, one can also consider adding strategies of the previous sections where appropriate. In particular, by-product removal (3J) has been demonstrated to be efficient. Conclusions Considering the breadth and depth of the strategies discussed above, it is clear that the field of cell factory engineering as a whole has come a long way. Through tens of thousands of studies, a multitude of individual challenges have been solved across a broad range of expression systems and diverse types of compounds. New and interesting avenues are being opened, such as expansion of the substrate range of E. coli turning it into a synthetic methylotroph (M€uller et al., 2015), or achieving the biosynthesis of caffeine and other methylxanthines in yeast from plant biosynthetic genes (McKeague et al., 2016), or achieving biobased nylon through large-scale engineering in C. glutamicum (Kind et al., 2014). We are also now seeing engineering of a central metabolism for increased protein production (Nocon et al., 2014). It seems like there is no obvious limit to the possibilities in sight. Furthermore, increasingly advanced work is being published, opening the field up into the applications of synthetic biology. This impacts small parts of the cell factory engineering, such as improvements in synthetic promoters (Vogl et al., 2013) and, at larger scale, the building of artificial pathways rather than using ‘‘simple’’ heterologous expression. Examples here include the biosynthesis of gastrodin (Bai et al., 2016) and the impressive feat of the Smolke Lab to biosynthesize opioids in S. cerevisiae (Thodey et al., 2014; Galanie et al., 2015). Another interesting development is the use of engineered consortia of species for achieving particular activities and synergies from using multiple species. A recent example employs bacterial consortia for desulfurization of oil-based fuels (Martı´nez et al., 2016), thus improving the quality. Even so, there are still significant problems that need to be addressed. Despite the extensive size of the toolbox of strategies outlined above, it is still difficult to know a priori which modifications are required for a specific combination of product and cell factory. This is one of the main reasons why the development time for new cell factories remains the largest bottleneck for new bioproducts. In order to move the field forward, these challenges must be addressed in multiple ways. Currently we see a next major step is to predict how cells change dynamically over the course of cultivation. At the moment, the most successful modeling of biological systems has been for steady state; which is often not representative of the conditions in a bioreactor in a long production process. Another brick in the wall will be the new conceptual frameworks (e.g., systems biology or system metabolic engineering), which are moving toward holistic design of cell factories and biological networks (Nocon et al., 2014). To achieve these goals, the importance of efficient genetic engineering and genome-editing tools cannot be overstated. Every time genetic engineering technologies have improved, so has the sophistication of cell factory engineering. Synthetic biology and genome-editing technologies such as CRISPR will accelerate cell factory engineering as we know it (Jakociunas et al., 2016), and they also promise to facilitate more-rapid tests of new theories, permutations of solutions, and generally cell engineering at a systems level. In tandem, dynamic modeling, holistic design, synthetic biology, and genome editing hold great promises for rational design of biological systems. AUTHOR CONTRIBUTIONS All authors were a part of the literature analysis and wrote the manuscript. Cell Systems 4, March 22, 2017 271 Cell Systems Review ACKNOWLEDGMENTS The authors are indebted to past and present students on the Cell Factory Engineering course at the Technical University of Denmark for giving inspiration and motivation for this text. Furthermore, sage advice and valuable input from several colleagues in writing this text is gratefully acknowledged. In particular, the input from Ana Ley was highly appreciated. H.F. Kildegaard thanks the Novo Nordisk Foundation for generous funding. M.R. Andersen is partially funded by the Novo Nordisk Foundation (grant NNF13OC0004831), and the Villum Foundation (grant VKR0234037). REFERENCES Ailor, E., and Betenbaugh, M.J. (1998). Overexpression of a cytosolic chaperone to improve solubility and secretion of a recombinant IgG protein in insect cells. Biotechnol. Bioeng. 58, 196–203. Ailor, E., Pathmanathan, J., Jongbloed, J.D., and Betenbaugh, M.J. (1999). A bacterial signal peptidase enhances processing of a recombinant single chain antibody fragment in insect cells. Biochem. Biophys. Res. Commun. 255, 444–450. Andersen, M.R., Nam, J.H., and Sharfstein, S.T. (2011). Protein glycosylation: analysis, characterization, and engineering. In Encyclopedia of Industrial Biotechnology, Bioprocess, Bioseparation, and Cell Technology, M.C. Flickinger, ed. (John Wiley), pp. 1–50. Anyaogu, D.C., and Mortensen, U.H. (2015). Heterologous production of fungal secondary metabolites in Aspergilli. Front. Microbiol. 6, 77. Asadollahi, M.A., Maury, J., Patil, K.R., Schalk, M., Clark, A., and Nielsen, J. (2009). Enhancing sesquiterpene production in Saccharomyces cerevisiae through in silico driven metabolic engineering. Metab. Eng. 11, 328–334. Association of Manufacturers and Formulators of Enzyme Products, 2009. List of commercial enzymes. http://www.amfep.org/content/list-enzymes. Bai, Y., Yin, H., Bi, H., Zhuang, Y., Liu, T., and Ma, Y. (2016). De novo biosynthesis of gastrodin in Escherichia coli. Metab. Eng. 35, 138–147. Bailey, J.E. (1991). Toward a science of metabolic engineering. Science 252, 1668–1675. Bailey, J.E., Sburlati, A., Hatzimanikatis, V., Lee, K., Renner, W.A., and Tsai, P.S. (2002). Inverse metabolic engineering: a strategy for directed genetic engineering of useful phenotypes. Biotechnol. Bioeng. 79, 568–579. Baltz, R.H. (2016). Genetic manipulation of secondary metabolite biosynthesis for improved production in Streptomyces and other actinomycetes. J. Ind. Microbiol. Biotechnol. 43, 343–370. Barnstable, C.J., Bodmer, W.F., Brown, G., Galfre, G., Milstein, C., Williams, A.F., and Ziegler, A. (1978). Production of monoclonal antibodies to group A erythrocytes, HLA and other human cell surface antigen-new tools for genetic analysis. Cell 14, 9–18. Bashor, C.J., Horwitz, A.A., Peisajovich, S.G., Lim, W.A., et al. (2010). Rewiring cells: synthetic biology as a tool to interrogate the organizational principles of living systems. Annu. Rev. Biophys. 39, 515–537. Becker, J., Sch€afer, R., Kohlstedt, M., Harder, B.J., Borchert, N.S., Sto¨ veken, N., Bremer, E., and Wittmann, C. (2013). Systems metabolic engineering of Corynebacterium glutamicum for production of the chemical chaperone ectoine. Microb. Cell Fact. 12, 110. Behnsen, J., Deriu, E., Sassone-Corsi, M., and Raffatellu, M. (2013). Probiotics: properties, examples, and specific applications. Cold Spring Harb. Perspect. Med. 3, a010074. Bekiesch, P., Basitta, P., and Apel, A.K. (2016). Challenges in the heterologous production of antibiotics in Streptomyces. Arch. Pharm. (Weinheim) 349, 594–601. Berkmen, M. (2012). Production of disulfide-bonded proteins in Escherichia coli. Protein Expr. Purif. 82, 240–251. Borkovich, K.A., Alex, L.A., Yarden, O., Freitag, M., Turner, G.E., Read, N.D., Seiler, S., Bell-Pedersen, D., Paietta, J., Plesofsky, N., et al. (2004). Lessons from the genome sequence of Neurospora crassa: tracing the path from genomic blueprint to multicellular organism. Microbiol. Mol. Biol. Rev. 68, 1–108. Borodina, I., Siebring, J., Zhang, J., Smith, C.P., van Keulen, G., Dijkhuizen, L., and Nielsen, J. (2008). Antibiotic overproduction in Streptomyces coelicolor A3(2) mediated by phosphofructokinase deletion. J. Biol. Chem. 283, 25186–25199. Borth, N., Mattanovich, D., Kunert, R., and Katinger, H. (2005). Effect of increased expression of protein disulfide isomerase and heavy chain binding protein on antibody secretion in a recombinant CHO cell line. Biotechnol. Prog. 21, 106–111. Burgard, A.P., Pharkya, P., and Maranas, C.D. (2003). Optknock: a bilevel programming framework for identifying gene knockout strategies for microbial strain optimization. Biotechnol. Bioeng. 84, 647–657. Calfon, M., Zeng, H., Urano, F., Till, J.H., Hubbard, S.R., Harding, H.P., Clark, S.G., and Ron, D. (2002). IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415, 92–96. Cambray, G., Guimaraes, J.C., Mutalik, V.K., Lam, C., Mai, Q.A., Thimmaiah, T., Carothers, J.M., Arkin, A.P., and Endy, D. (2013). Measurement and modeling of intrinsic transcription terminators. Nucleic Acids Res. 41, 5139–5148. Carvalho, N.D., Jørgensen, T.R., Arentshorst, M., Nitsche, B.M., van den Hondel, C.A., Archer, D.B., and Ram, A.F. (2012). Genome-wide expression analysis upon constitutive activation of the HacA bZIP transcription factor in Aspergillus niger reveals a coordinated cellular response to counteract ER stress. BMC Genomics 13, 350. Cavallaro, G., Decaria, L., and Rosato, A. (2008). Genome-based analysis of heme biosynthesis and uptake in prokaryotic systems. J. Proteome Res. 7, 4946–4954. Celik, E., and Calık, P. (2011). Production of recombinant proteins by yeast cells. Biotechnol. Adv. 30, 1108–1118. Chemier, J.A., Fowler, Z.L., Koffas, M.A., and Leonard, E. (2009). Trends in microbial synthesis of natural products and biofuels. Adv. Enzymol. Relat. Areas Mol. Biol. 76, 151–217. Chen, X., Dong, X., Wang, Y., Zhao, Z., and Liu, L. (2015). Mitochondrial engineering of the TCA cycle for fumarate production. Metab. Eng. 31, 62. Choi, B.-K., Bobrowicz, P., Davidson, R.C., Hamilton, S.R., Kung, D.H., Li, H., Miele, R.G., Nett, J.H., Wildt, S., and Gerngross, T.U. (2003). Use of combinatorial genetic libraries to humanize N-linked glycosylation in the yeast Pichia pastoris. Proc. Natl. Acad. Sci. USA 100, 5022–5027. Cui, Y.-Y., Ling, C., Zhang, Y.Y., Huang, J., and Liu, J.Z. (2014). Production of shikimic acid from Escherichia coli through chemically inducible chromosomal evolution and cofactor metabolic engineering. Microb. Cell Fact. 13, 21. Curran, K.A., Karim, A.S., Gupta, A., and Alper, H.S. (2013). Use of expressionenhancing terminators in Saccharomyces cerevisiae to increase mRNA half-life and improve gene expression control for metabolic engineering applications. Metab. Eng. 19, 88–97. Damasceno, L.M., Huang, C.-J.J., and Batt, C.A. (2012). Protein secretion in Pichia pastoris and advances in protein production. Appl. Microbiol. Biotechnol. 93, 31–39. Davis, R., Schooley, K., Rasmussen, B., Thomas, J., and Reddy, P. (2000). Effect of PDI overexpression on recombinant protein secretion in CHO cells. Biotechnol. Prog. 16, 736–743. de Marco, A. (2009). Strategies for successful recombinant expression of disulfide bond-dependent proteins in Escherichia coli. Microb. Cell Fact. 8, 26. Dehli, T., Solem, C., and Jensen, P.R. (2012). Tunable promoters in synthetic and systems biology. Subcell Biochem. 64, 181–201. Dueber, J.E., Wu, G.C., Malmirchegini, G.R., Moon, T.S., Petzold, C.J., Ullal, A.V., Prather, K.L., and Keasling, J.D. (2009). Synthetic protein scaffolds provide modular control over metabolic flux. Nat. Biotechnol. 27, 753–759. Dunlop, M.J., Dossani, Z.Y., Szmidt, H.L., Chu, H.C., Lee, T.S., Keasling, J.D., Hadi, M.Z., and Mukhopadhyay, A. (2011). Engineering microbial biofuel tolerance and export using efflux pumps. Mol. Syst. Biol. 7, 487. Edwards, A.M., Arrowsmith, C.H., Christendat, D., Dharamsi, A., Friesen, J.D., Greenblatt, J.F., and Vedadi, M. (2000). Protein production: feeding the crystallographers and NMR spectroscopists. Nat. Struct. Biol. 7, 970–972. 272 Cell Systems 4, March 22, 2017 Cell Systems Review Fisher, A.K., Freedman, B.G., Bevan, D.R., and Senger, R.S. (2014). A review of metabolic and enzymatic engineering strategies for designing and optimizing performance of microbial cell factories. Comput. Struct. Biotechnol. J. 11, 91–99. Fleissner, A., and Dersch, P. (2010). Expression and export: recombinant protein production systems for Aspergillus. Appl. Microbiol. Biotechnol. 87, 1255–1270. Franken, A.C.W., Lokman, B.C., Ram, A.F., Punt, P.J., van den Hondel, C.A., and de Weert, S. (2011). Heme biosynthesis and its regulation: towards understanding and improvement of heme biosynthesis in filamentous fungi. Appl. Microbiol. Biotechnol. 91, 447–460. Galanie, S., Thodey, K., Trenchard, I.J., Filsinger Interrante, M., and Smolke, C.D. (2015). Complete biosynthesis of opioids in yeast. Science 349, 1095–1100. Gasser, B., Saloheimo, M., Rinas, U., Dragosits, M., Rodrı´guez-Carmona, E., Baumann, K., Giuliani, M., Parrilli, E., Branduardi, P., Lang, C., et al. (2008). Protein folding and conformational stress in microbial cells producing recombinant proteins: a host comparative overview. Microb. Cell Fact. 7, 11. Ghosh, A., Zhao, H., and Price, N.D. (2011). Genome-scale consequences of cofactor balancing in engineered pentose utilization pathways in Saccharomyces cerevisiae. PLoS One 6, e27316. Hamann, T., and Lange, L. (2006). Discovery, cloning and heterologous expression of secreted potato proteins reveal erroneous pre-mRNA splicing in Aspergillus oryzae. J. Biotechnol. 126, 265–276. Hamilton, S.R., Davidson, R.C., Sethuraman, N., Nett, J.H., Jiang, Y., Rios, S., Bobrowicz, P., Stadheim, T.A., Li, H., Choi, B.K., et al. (2006). Humanization of yeast to produce complex terminally sialylated glycoproteins. Science 313, 1441–1443. Hienonen, E., Romantschuk, M., Fenel, F., and Taira, S. (2007). Transcript stabilization by mRNA sequences from hrpA of Pseudomonas syringae. J. Biotechnol. 128, 258–267. Hossler, P. (2012). Protein glycosylation control in mammalian cell culture: past precedents and contemporary prospects. Adv. Biochem. Eng. Biotechnol. 127, 187–219. Hossler, P., Khattak, S.F., and Li, Z.J. (2009). Optimal and consistent protein glycosylation in mammalian cell culture. Glycobiology 19, 936–949. Hou, J., Tyo, K., Liu, Z., Petranovic, D., and Nielsen, J. (2012). Engineering of vesicle trafficking improves heterologous protein secretion in Saccharomyces cerevisiae. Metab. Eng. 14, 120–127. Hsu, P.D., Lander, E.S., and Zhang, F. (2014). Development and applications of CRISPR-Cas9 for genome engineering. Cell 157, 1262–1278. Hwang, K.-S., Kim, H.U., Charusanti, P., Palsson, B.Ø., and Lee, S.Y. (2014). Systems biology and biotechnology of Streptomyces species for the production of secondary metabolites. Biotechnol. Adv. 32, 255–268. Idiris, A., Tohda, H., Sasaki, M., Okada, K., Kumagai, H., Giga-Hama, Y., and Takegawa, K. (2010). Enhanced protein secretion from multiprotease-deficient fission yeast by modification of its vacuolar protein sorting pathway. Appl. Microbiol. Biotechnol. 85, 667–677. Innis, M.A., Holland, M.J., McCabe, P.C., Cole, G.E., Wittman, V.P., Tal, R., Watt, K.W., Gelfand, D.H., Holland, J.P., and Meade, J.H. (1985). Expression, glycosylation, and secretion of an Aspergillus glucoamylase by Saccharomyces cerevisiae. Science 228, 21–26. Jakociunas, T., Jensen, M.K., and Keasling, J.D. (2016). CRISPR/Cas9 advances engineering of microbial cell factories. Metab. Eng. 34, 44–59. Jin, F.J., Watanabe, T., Juvvadi, P.R., Maruyama, J., Arioka, M., and Kitamoto, K. (2007). Double disruption of the proteinase genes, tppA and pepE, increases the production level of human lysozyme by Aspergillus oryzae. Appl. Microbiol. Biotechnol. 76, 1059–1068. de Jongh, W.A., and Nielsen, J. (2008). Enhanced citrate production through gene insertion in Aspergillus niger. Metab. Eng. 10, 87–96. Josse´ , L., Smales, C.M., and Tuite, M.F. (2012). Engineering the chaperone network of CHO cells for optimal recombinant protein production and authenticity. Methods Mol. Biol. 824, 595–608. Keasling, J.D. (2008). Synthetic biology for synthetic chemistry. ACS Chem. Biol. 3, 64–76. Keasling, J.D. (2012). Synthetic biology and the development of tools for metabolic engineering. Metab. Eng. 14, 189–195. Kiel, C., Yus, E., and Serrano, L. (2010). Engineering signal transduction pathways. Cell 140, 33–47. Kim, S.H., and Lee, G.M. (2007). Down-regulation of lactate dehydrogenase-A by siRNAs for reduced lactic acid formation of Chinese hamster ovary cells producing thrombopoietin. Appl. Microbiol. Biotechnol. 74, 152–159. Kind, S., Jeong, W.K., Schro¨ der, H., and Wittmann, C. (2010). Systems-wide metabolic pathway engineering in Corynebacterium glutamicum for bio-based production of diaminopentane. Metab. Eng. 12, 341–351. Kind, S., Neubauer, S., Becker, J., Yamamoto, M., Vo¨ lkert, M., Abendroth, Gv, Zelder, O., and Wittmann, C. (2014). From zero to hero – production of biobased nylon from renewable resources using engineered Corynebacterium glutamicum. Metab. Eng. 25, 113–123. Ku, S.C.Y., Ng, D.T., Yap, M.G., and Chao, S.H. (2008). Effects of overexpression of X-box binding protein 1 on recombinant protein production in Chinese hamster ovary and NS0 myeloma cells. Biotechnol. Bioeng. 99, 155–164. Lee, S.Y., and Kim, H.U. (2015). Systems strategies for developing industrial microbial strains. Nat. Biotechnol. 33, 1061–1072. Lee, J.H., Lee, D.E., Lee, B.U., and Kim, H.S. (2003). Global analyses of transcriptomes and proteomes of a parent strain and an L-threonine-overproducing mutant strain. J. Bacteriol. 185, 5442–5451. Lee, K.H., Park, J.H., Kim, T.Y., Kim, H.U., and Lee, S.Y. (2007). Systems metabolic engineering of Escherichia coli for L-threonine production. Mol. Syst. Biol. 3, 149. Lee, J.W., Na, D., Park, J.M., Lee, J., Choi, S., and Lee, S.Y. (2012). Systems metabolic engineering of microorganisms for natural and non-natural chemicals. Nat. Chem. Biol. 8, 536–546. Li, H., Sethuraman, N., Stadheim, T.A., Zha, D., Prinz, B., Ballew, N., Bobrowicz, P., Choi, B.K., Cook, W.J., Cukan, M., et al. (2006). Optimization of humanized IgGs in glycoengineered Pichia pastoris. Nat. Biotechnol. 24, 210–215. Li, A., Pfelzer, N., Zuijderwijk, R., Brickwedde, A., van Zeijl, C., and Punt, P. (2013). Reduced by-product formation and modified oxygen availability improve itaconic acid production in Aspergillus niger. Appl. Microbiol. Biotechnol. 97, 3901–3911. Liu, L., Yang, H., Shin, H.D., Li, J., Du, G., and Chen, J. (2013). Recent advances in recombinant protein expression by Corynebacterium, Brevibacterium, and Streptomyces: from transcription and translation regulation to secretion pathway selection. Appl. Microbiol. Biotechnol. 97, 9597–9608. Lubertozzi, D., and Keasling, J.D. (2009). Developing Aspergillus as a host for heterologous expression. Biotechnol. Adv. 27, 53–75. Martin, V.J.J., Pitera, D.J., Withers, S.T., Newman, J.D., and Keasling, J.D. (2003). Engineering a mevalonate pathway in Escherichia coli for production of terpenoids. Nat. Biotechnol. 21, 796–802. Martı´nez, I., Mohamed Mel, S., Rozas, D., Garcı´a, J.L., and Dı´az, E. (2016). Engineering synthetic bacterial consortia for enhanced desulfurization and revalorization of oil sulfur compounds. Metab. Eng. 35, 46–54. Mattanovich, D., Branduardi, P., Dato, L., Gasser, B., Sauer, M., and Porro, D. (2012). Recombinant protein production in yeasts. Methods Mol. Biol. 824, 329–358. McKeague, M., Wang, Y.H., Cravens, A., Win, M.N., and Smolke, C.D. (2016). Engineering a microbial platform for de novo biosynthesis of diverse methylxanthines. Metab. Eng. 38, 191–203. Meerman, H.J., and Georgiou, G. (1994). Construction and characterization of a set of E. coli strains deficient in all known loci affecting the proteolytic stability of secreted recombinant proteins. Biotechnology (N Y) 12, 1107–1110. Meta, A., Nakatake, H., Imamura, T., Nozaki, C., and Sugimura, K. (2009). High-yield production and characterization of biologically active recombinant aprotinin expressed in Saccharomyces cerevisiae. Protein Expr. Purif. 66, 22–27. Cell Systems 4, March 22, 2017 273 Cell Systems Review Mohan, C., Park, S.H., Chung, J.Y., and Lee, G.M. (2007). Effect of doxycycline-regulated protein disulfide isomerase expression on the specific productivity of recombinant CHO cells: thrombopoietin and antibody. Biotechnol. Bioeng. 98, 611–615. Molna´ r, I., Jungmann, V., Stege, J., Trefzer, A., and Pachlatko, J.P. (2006). Biocatalytic conversion of avermectin into 400 -oxo-avermectin: discovery, characterization, heterologous expression and specificity improvement of the cytochrome P450 enzyme. Biochem. Soc. Trans. 34, 1236–1240. M€uller, J.E.N., Meyer, F., Litsanov, B., Kiefer, P., Potthoff, E., Heux, S., Quax, W.J., Wendisch, V.F., Brautaset, T., Portais, J.C., et al. (2015). Engineering Escherichia coli for methanol conversion. Metab. Eng. 28, 190–201. Nerem, R.M. (1991). Cellular engineering. Ann. Biomed. Eng. 19, 529–545. Nielsen, J. (1998). Metabolic engineering: techniques for analysis of targets for genetic manipulations. Biotechnol. Bioeng. 58, 125–132. Nielsen, J. (2014). Synthetic biology for engineering acetyl coenzyme a metabolism in yeast. mBio 5, e02153. Nielsen, M.T., Nielsen, J.B., Anyaogu, D.C., Holm, D.K., Nielsen, K.F., Larsen, T.O., and Mortensen, U.H. (2013). Heterologous reconstitution of the intact geodin gene cluster in Aspergillus nidulans through a simple and versatile PCR based approach. PLoS One 8, e72871. Nieuwenhuizen, P.J., and Lyon, D. (2011). Anticipating opportunities in industrial biotechnology: sizing the market and growth scenarios. J. Commer. Biotechnol. 17, 159–164. Nocon, J., Steiger, M.G., Pfeffer, M., Sohn, S.B., Kim, T.Y., Maurer, M., Rußmayer, H., Pfl€ugl, S., Ask, M., Haberhauer-Troyer, C., et al. (2014). Model based engineering of Pichia pastoris central metabolism enhances recombinant protein production. Metab. Eng. 24, 129–138. Ohya, T., Hayashi, T., Kiyama, E., Nishii, H., Miki, H., Kobayashi, K., Honda, K., Omasa, T., and Ohtake, H. (2008). Improved production of recombinant human antithrombin III in Chinese hamster ovary cells by {ATF}4 overexpression. Biotechnol. Bioeng. 100, 317–324. Ostergaard, S., Olsson, L., Johnston, M., and Nielsen, J. (2000). Increasing galactose consumption by Saccharomyces cerevisiae through metabolic engineering of the GAL gene regulatory network. Nat. Biotechnol. 18, 1283–1286. Park, J.H., Oh, J.E., Lee, K.H., Kim, J.Y., and Lee, S.Y. (2012). Rational design of Escherichia coli for L-isoleucine production. ACS Synth. Biol. 1, 532–540. Pfeifer, B.A., and Khosla, C. (2001). Biosynthesis of polyketides in heterologous hosts. Microbiol. Mol. Biol. Rev. 65, 106–118. Pfleger, B.F., Pitera, D.J., Smolke, C.D., and Keasling, J.D. (2006). Combinatorial engineering of intergenic regions in operons tunes expression of multiple genes. Nat. Biotechnol. 24, 1027–1032. Pfleger, B.F., Gossing, M., and Nielsen, J. (2015). Metabolic engineering strategies for microbial synthesis of oleochemicals. Metab. Eng. 29, 1–11. Pharkya, P., and Maranas, C.D. (2006). An optimization framework for identifying reaction activation/inhibition or elimination candidates for overproduction in microbial systems. Metab. Eng. 8, 1–13. Pickens, L.B., Tang, Y., and Chooi, Y.H. (2011). Metabolic engineering for the production of natural products. Annu. Rev. Chem. Biomol. Eng. 2, 211–236. Punt, P.J., Schuren, F.H., Lehmbeck, J., Christensen, T., Hjort, C., and van den Hondel, C.A. (2008). Characterization of the Aspergillus niger prtT, a unique regulator of extracellular protease encoding genes. Fungal Genet. Biol. 45, 1591–1599. Pybus, L.P., Dean, G., West, N.R., Smith, A., Daramola, O., Field, R., Wilkinson, S.J., and James, D.C. (2013). Model-directed engineering of ‘‘difficultto-express’’ monoclonal antibody production by Chinese hamster ovary cells. Biotechnol. Bioeng. 111, 372–385. Qi, K., Zhong, J.J., and Xia, X.X. (2014). Triggering respirofermentative metabolism in the crabtree-negative yeast Pichia guilliermondii by disrupting the CAT8 gene. Appl. Environ. Microbiol. 80, 3879–3887. Qian, Z.G., Xia, X.X., and Lee, S.Y. (2011). Metabolic engineering of Escherichia coli for the production of cadaverine: a five carbon diamine. Biotechnol. Bioeng. 108, 93–103. Ranganathan, S., Suthers, P.F., and Maranas, C.D. (2010). OptForce: an optimization procedure for identifying all genetic manipulations leading to targeted overproductions. PLoS Comput. Biol. 6, e1000744. Ratner, M. (2014). Genentech’s glyco-engineered antibody to succeed Rituxan. Nat. Biotechnol. 32, 6–7. Ribela, M.T., Bianco, A.C., and Bartolini, P. (1996). The use of recombinant human thyrotropin produced by Chinese hamster ovary cells for the preparation of immunoassay reagents. J. Clin. Endocrinol. Metab. 81, 249–256. Rodrigues, A.L., Trachtmann, N., Becker, J., Lohanatha, A.F., Blotenberg, J., Bolten, C.J., Korneli, C., de Souza Lima, A.O., Porto, L.M., Sprenger, G.A., et al. (2013). Systems metabolic engineering of Escherichia coli for production of the antitumor drugs violacein and deoxyviolacein. Metab. Eng. 20, 29–41. Rodrigues, A.L., Becker, J., de Souza Lima, A.O., Porto, L.M., and Wittmann, C. (2014). Systems metabolic engineering of Escherichia coli for gram scale production of the antitumor drug deoxyviolacein from glycerol. Biotechnol. Bioeng. 111, 2280–2289. Rosano, G.L., and Ceccarelli, E.A. (2014). Recombinant protein expression in Escherichia coli: advances and challenges. Front. Microbiol. 5, 172. Ruohonen, L., Toikkanen, J., Tieaho, V., Outola, M., Soderlund, H., and Keranen, S. (1997). Enhancement of protein secretion in Saccharomyces cerevisiae by overproduction of Sso protein, a late-acting component of the secretory machinery. Yeast 13, 337–351. Sathyamurthy, M., Lee, J.S., Park, J.H., Kim, Y.J., Jeong, J.Y., Jang, J.W., and Lee, G.M. (2012). Overexpression of PACEsol improves BMP-7 processing in recombinant CHO cells. J. Biotechnol. 164, 336–339. Schro¨ der, M. (2008). Engineering eukaryotic protein factories. Biotechnol. Lett. 30, 187–196. Siddiqui, M.S., Thodey, K., Trenchard, I., and Smolke, C.D. (2012). Advancing secondary metabolite biosynthesis in yeast with synthetic biology tools. FEMS Yeast Res. 12, 144–170. Stephanopoulos, G., and Vallino, J.J. (1991). Network rigidity and metabolic engineering in metabolite overproduction. Science 252, 1675–1681. Thodey, K., Galanie, S., and Smolke, C.D. (2014). A microbial biomanufacturing platform for natural and semisynthetic opioids. Nat. Chem. Biol. 10, 1–10. Tigges, M., and Fussenegger, M. (2006). Xbp1-based engineering of secretory capacity enhances the productivity of Chinese hamster ovary cells. Metab. Eng. 8, 264–272. Trosset, J.Y., and Carbonell, P. (2015). Synthetic biology for pharmaceutical drug discovery. Drug Des. Dev. Ther. 9, 6285–6302. Tyo, K.E., Liu, Z., Magnusson, Y., Petranovic, D., and Nielsen, J. (2014). Impact of protein uptake and degradation on recombinant protein secretion in yeast. Appl. Microbiol. Biotechnol. 98, 7149–7159. Valkonen, M., Ward, M., Wang, H., Penttil€a, M., and Saloheimo, M. (2003a). Improvement of foreign-protein production in Aspergillus niger var. awamori by constitutive induction of the unfolded-protein response. Appl. Environ. Microbiol. 69, 6979–6986. Valkonen, M., Penttil€a, M., and Saloheimo, M. (2003b). Effects of inactivation and constitutive expression of the unfolded- protein response pathway on protein production in the yeast Saccharomyces cerevisiae. Appl. Environ. Microbiol. 69, 2065–2072. van Dijl, J.M., de Jong, A., Smith, H., Bron, S., and Venema, G. (1991). Signal peptidase I overproduction results in increased efficiencies of export and maturation of hybrid secretory proteins in Escherichia coli. Mol. Gen. Genet. 227, 40–48. van Rossum, H.M., Kozak, B.U., Pronk, J.T., and van Maris, A.J. (2016). Engineering cytosolic acetyl-coenzyme A supply in Saccharomyces cerevisiae: pathway stoichiometry, free-energy conservation and redox-cofactor balancing. Metab. Eng. 36, 99–115. Vogl, T., Hartner, F.S., and Glieder, A. (2013). New opportunities by synthetic biology for biopharmaceutical production in Pichia pastoris. Curr. Opin. Biotechnol. 24, 1094–1101. 274 Cell Systems 4, March 22, 2017 Cell Systems Review Waegeman, H., and Soetaert, W. (2011). Increasing recombinant protein production in Escherichia coli through metabolic and genetic engineering. J. Ind. Microbiol. Biotechnol. 38, 1891–1910. Walsh, G. (2010). Biopharmaceutical benchmarks 2010. Nat. Biotechnol. 28, 917–924. Walsh, G. (2014). Biopharmaceutical benchmarks 2014. Nat. Biotechnol. 32, 992–1000. Wang, Y., and Yu, O. (2012). Synthetic scaffolds increased resveratrol biosynthesis in engineered yeast cells. J. Biotechnol. 157, 258–260. Wang, J., Liu, W., Ding, W., Zhang, G., and Liu, J. (2013). Increasing ethanol titer and yield in a gpd1D gpd2D strain by simultaneous overexpression of GLT1 and STL1 in Saccharomyces cerevisiae. Biotechnol. Lett. 35, 1859–1864. Ward, O.P. (2011). Production of recombinant proteins by filamentous fungi. Biotechnol. Adv. 30, 1119–1139. Ward, M., Wilson, L.J., Kodama, K.H., Rey, M.W., and Berka, R.M. (1990). Improved production of chymosin in Aspergillus by expression as a glucoamylase-chymosin fusion. Biotechnology (N Y) 8, 435–440. Weber, T., Blin, K., Duddela, S., Krug, D., Kim, H.U., Bruccoleri, R., Lee, S.Y., Fischbach, M.A., M€uller, R., Wohlleben, W., et al. (2015). antiSMASH 3.0-a comprehensive resource for the genome mining of biosynthetic gene clusters. Nucleic Acids Res. 43, W237–W243. Westwood, A.D., Rowe, D.A., and Clarke, H.R.G. (2010). Improved recombinant protein yield using a codon deoptimized DHFR selectable marker in a CHEF1 expression plasmid. Biotechnol. Prog. 26, 1558–1566. Wu, J., Du, G., Zhou, J., and Chen, J. (2014). Systems metabolic engineering of microorganisms to achieve large-scale production of flavonoid scaffolds. J. Biotechnol. 188, 72–80. Wurm, F.M. (2004). Production of recombinant protein therapeutics in cultivated mammalian cells. Nat. Biotechnol. 22, 1393–1398. Xiao, H., and Zhong, J.J. (2016). Production of useful terpenoids by higher-fungus cell factory and synthetic biology approaches. Trends Biotechnol. 34, 242–255. Xu, J.W., Xu, Y.N., and Zhong, J.J. (2012). Enhancement of ganoderic acid accumulation by overexpression of an n-terminally truncated 3-hydroxy-3methylglutaryl coenzyme a reductase gene in the basidiomycete Ganoderma lucidum. Appl. Environ. Microbiol. 78, 7968–7976. Yamauchi, Y., Hirasawa, T., Nishii, M., Furusawa, C., and Shimizu, H. (2014). Enhanced acetic acid and succinic acid production under microaerobic conditions by Corynebacterium glutamicum harboring Escherichia coli transhydrogenase gene pntAB. J. Gen. Appl. Microbiol. 60, 112–118. Yang, Z., Wang, S., Halim, A., Schulz, M.A., Frodin, M., Rahman, S.H., VesterChristensen, M.B., Behrens, C., Kristensen, C., Vakhrushev, S.Y., et al. (2015). Engineered CHO cells for production of diverse, homogeneous glycoproteins. Nat. Biotechnol. 33, 842–844. Yoon, J., Aishan, T., Maruyama, J., and Kitamoto, K. (2010). Enhanced production and secretion of heterologous proteins by the filamentous fungus Aspergillus oryzae via disruption of vacuolar protein sorting receptor gene Aovps10. Appl. Environ. Microbiol. 76, 5718–5727. Yu, J.L., Xia, X.X., Zhong, J.J., and Qian, Z.G. (2014). Direct biosynthesis of adipic acid from a synthetic pathway in recombinant Escherichia coli. Biotechnol. Bioeng. 111, 2580–2586. Zabala, D., Bran˜ a, A.F., Flo´ rez, A.B., Salas, J.A., and Me´ ndez, C. (2013). Engineering precursor metabolite pools for increasing production of antitumor mithramycins in Streptomyces argillaceus. Metab. Eng. 20, 187–197. Cell Systems 4, March 22, 2017 275 Cell Systems Review