Available online at www.sciencedirect.com
ScienceDirect Recent advances in systems metabolic engineering tools and strategies Tong Un Chae1, So Young Choi1, Je Woong Kim, Yoo-Sung Ko and Sang Yup Lee Metabolic engineering has been playing increasingly important roles in developing microbial cell factories for the production of various chemicals and materials to achieve sustainable chemical industry. Nowadays, many tools and strategies are available for performing systems metabolic engineering that allows systems-level metabolic engineering in more sophisticated and diverse ways by adopting rapidly advancing methodologies and tools of systems biology, synthetic biology and evolutionary engineering. As an outcome, development of more efficient microbial cell factories has become possible. Here, we review recent advances in systems metabolic engineering tools and strategies together with accompanying application examples. In addition, we describe how these tools and strategies work together in simultaneous and synergistic ways to develop novel microbial cell factories. Address Metabolic and Biomolecular Engineering National Research Laboratory, Department of Chemical and Biomolecular Engineering (BK21 Plus Program), BioProcess Engineering Research Center, Institute for the BioCentury, KAIST, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of Korea Corresponding author: Lee, Sang Yup (
[email protected]) Both authors contributed equally to this work.
1
Current Opinion in Biotechnology 2017, 47:67–82 This review comes from a themed issue on Tissue, cell & pathway engineering Edited by Kobi Benenson and Matthias Lutolf
http://dx.doi.org/10.1016/j.copbio.2017.06.007 0958-1669/# 2017 Elsevier Ltd. All rights reserved.
occurring microorganisms are rather low. Metabolic engineering has made it possible to improve three major performance indices (titer, yield and productivity) in bio-based production of chemicals and materials of interest. Some of the early satisfactory results on performance improvement encouraged us to put more effort on advancing the strategies of metabolic engineering towards development of biorefinery processes capable of producing diverse chemicals and materials in cost effective manner. To meet the need to develop industrially competitive microorganisms, systems metabolic engineering that integrates metabolic engineering with tools and strategies of systems biology, synthetic biology and evolutionary engineering has been established [1–3]. Various systems metabolic engineering tools (Figure 1) have allowed us to perform advanced strain engineering tasks including construction of novel metabolic pathways, genome-wide identification of metabolic engineering targets, fine-tuning and control of gene expression, multiplex genome engineering, creation of synthetic circuits, and increasing tolerance to target chemicals or intermediates when needed. Such systems metabolic engineering tools and strategies have been employed not only for strain development but also for optimizing bioprocess variables such as medium composition, pH, aeration, cultivation mode, and nutrient feeding strategies. In this paper, we review recent advances in systems metabolic engineering tools and strategies together with accompanying application examples (Table 1). The tools and strategies are classified into three categories for convenience: systems biology, synthetic biology and evolutionary engineering. We also discuss how these tools and strategies have played synergistic roles in developing microbial cell factories.
Introduction Climate change is one of the most important global risks requiring attention with utmost priority. One of the major contributors for climate change is our heavy use of fossil oil and gas. Thus, there has been much interest in producing chemicals and materials by microbial cell factories from renewable non-food biomass, instead of relying on petrochemical routes. Microorganisms can theoretically produce all the metabolites present in their metabolic network, but the efficiencies of producing most, if not all, of these chemicals by employing naturally www.sciencedirect.com
Systems biology tools and strategies for metabolic engineering Systems biology aims at interpreting cellular phenomena at systems level by employing a wide range of data and tools, most notably omics data analysis and computational simulations. It plays important roles in system metabolic engineering by suggesting more systematic and highthroughput engineering approaches compared to conventional metabolic engineering [4]. To accomplish the desired phenotype such as cell growth and target chemical Current Opinion in Biotechnology 2017, 47:67–82
68 Tissue, cell & pathway engineering
Figure 1
(a)
(b)
Transcriptomics
(c) :to be deleted
Fluxomics
Proteomics
(d)
Multi-omics
Module 1
Module 2
Genome models
:up-regulated :down-regulated
MOMA ROOM OptKnock FSEOF FVSEOF with GR
Metabolomics
(e)
(f)
V1
GIMME GIMMEp GIM3E iMAT, E-Flux
Reduced flux space V2
V3
(g)
Ribosome
DNA
:Heterologous gene :Endogenous gene
BNICE GEMPath PathTracer
(h) Gene
sRNA DNA SDM SSM
Hfq mRNA Promoter
Structure prediction
Active site
(i)
RBS
Synthetic sRNA RNA interference Antisense RNA
Flux
Gene expression components
(j)
TIR
(k)
A
C
B Enz2 Enz1
DNA
Enz1 B Enz2 Enz1
Enz2
Coculture
(o)
Mg2+ Mn2+ ↑ dGTP dTTP↑ DNAP dCTP dATP↓ Ep-PCR
Promoter
System biology Synthetic biology
Output
Sensor
UV
Product
C
Microcompartment
Synthetic scaffold
(n)
Species2
B
C
3′ 5′
B
A
dCas9
(m)
A
Species1
Enz3 A
CRISPR, CRISPRi
STAR
D
PAM
RNAP
STAR Antisense
(l)
Cas9 sgRNA
RNAP
Recombineering
MNNG NTG
Evolutionary engineering
Input Promoter
DNA Shuffling
Sensor
Increasing diversity
Output
Biosensor
Oligo pool
MAGE YOGE RAGE Current Opinion in Biotechnology
Tools and strategies for systems metabolic engineering. Systems biology tools and strategies (represented in purple box) include (a) omics, (b) in silico-based gene target identification, (c) omics integrated genome-scale metabolic model, (d) pathway prediction and (e) enzyme design. Synthetic biology tools and strategies (represented by blue box) include (f) gene expression component based pathway optimization, (g) RNA based gene downregulation, (h) RNA based gene overexpression, (i) CRISPR based gene over/downregulation, (j) synthetic scaffold, (k) microcompartment and (l) co-culture. Evolutionary engineering tools and strategies include (m) increasing genetic diversity, (n) biosensor and (o) oligo-mediated evolution.
production, various tools and strategies in system metabolic engineering have been developed and used in constructing high performance production strains.
Omics Analyses of omics including genomics, transcriptomics, proteomics, metabolomics, and fluxomics can provide Current Opinion in Biotechnology 2017, 47:67–82
invaluable system-wide information on cellular and metabolic characteristics under various genotypic and environmental conditions. By analyzing omics data, target genes to be manipulated can be identified for enhancing chemical production capability (Figure 1a). Recently interest in fluxomics is keep increasing since ultimate purpose of metabolic engineering is to optimize flux www.sciencedirect.com
www.sciencedirect.com
Table 1 Recent applications of systems metabolic engineering tools and strategies Category Systems biology
Purpose Understanding cellular metabolic status, bottleneck or potential
Genome-wide knockout target identification
Pathway prediction
Current Opinion in Biotechnology 2017, 47:67–82
Enzyme design
Synthetic biology
Pathway optimization
Strain
Product
S. cerevisiae
Ethanol
M. succiniciproducens
Succinic acid
C. tyrobutyricum
N/A a
Genomics/ Phenomics/ Transcriptomics/ Modeling approach MOMA
E. coli
40 chemicals
S. cerevisiae
Amorphadiene
OptKnock
E. coli
1,4-Butanediol
Flux response analysis FSEOF
E. coli
Fumaric acid
E. coli
Lycopene
FVSEOF with GR
E. coli
Putrescine
tSOT
S. coelicolor
Actinorhodin
System based on semisupervised Gaussian process GEM-Path
N/A
Pinocembrin
E. coli
20 Chemicals
MapMaker/ PathTracer
E. coli
1777 Chemicals
Homology modeling
E. coli
3-HP
de novo computational design Promoters
N/A
N/A
E. coli
Violacein
RBS library calculator
E. coli
Cartenoid
Effect of tool/strategy utilization
References
Identify reason for inefficient ethanol production from xylose to be low glycolysis flux Reveal underlying mechanism allowing simultaneously usage sucrose and glycerol by fruA deleted mutant Whole genome sequence and unique metabolic characteristics were elucidated allowing start of metabolic engineering Potential of seven E. coli strains as host for 40 chemicals were evaluated in both aerobic and anaerobic condition
[7]
Identification of novel 10 gene knockout target and increase production yield by 8 to 10-fold Production enhanced by knocking out suggested knock out targets (ldhA, pflB, adhE, and mdhA) with a modified lpdA in microaerobic condition Identification of ppc as overexpression target and result in enhancement of production titer by 2.8-fold Prediction of idi and mdh as overexpression target and result in enhanced production by 2.7-fold Successful prediction of 5 overexpression targets (glk, acnA, acnB, ackA and ppc) which resulted in increased production by 20.5% on average Overexpression of ribulose 5-phosphate 3-epimerase and NADP dependent malic enzyme increased production by 2 and 1.8-fold, respectively Successfully predicted most efficient pathway for production
[17]
[8] [9]
[10]
[18]
[22] [23] [24]
[26]
[32]
Identified 245 unique heterologous pathways for 20 commodity compounds Demonstrated total 1777 non-natural products can be potentially synthesized by E. coli by adding known heterologous reactions Homology modeling based engineering of GabD4 enzyme resulted in increased activity to 1.4-fold and also produced 71.9 g/L of 3-HP Developed new energy-efficient formate assimilation pathway
[33]
Balancing each biosynthesis gene (vioABEDC) expressions using 5 mutants T4 promoters resulted in 63fold improvement in titer Generated RBS sets to efficient explore expression landscape of crtEBI gene helping to find optimal expression levels for optimal production
[52]
[35]
[39]
[40]
[56]
Recent systems metabolic engineering tools and strategies Chae et al. 69
Gene-wide overexpression target identification
Tool/Strategy Metabolomics/ Fluxomics Transcriptomics/ Proteomics Genomics/ Proteomics
Category
Purpose
Tool/Strategy
Strain
Product
sRNA
E. coli
1,3-DAP
RNAi
S. cerevisiae
Itaconic acid
asRNA
E. coli
Malonyl-CoA derivatives
CRISPR
S. cerevisiae
Mevalonate
E. coli
b-Carotene
C. glutamicum
L-Lysine/L-
CRISPRi
Glutamate
Synthetic scaffold
Microcompartment
Coculture
E. coli
PHAs
E. coli
Mevalonate
B. subtilis
GlcNAs
S. cerevisiae
Isobutanol
E. coli
Ethanol
T. reesei/E. coli
Isobutanol
E. coli/S. cerevisiae
UV and NTG
C. glutamicum
Oxygenated taxanes cis,cis-Muconic acid L-arginine
NTG/genome shuffling
C. beijerinckii
Isopropanol
RAGE
S. cerevisiae
Acetic acid
Ep-PCR
S. cerevisiae
Lycopene
Biosensor RBS libraries Biosensor MAGE
C. glutamicum E. coli C. glutamicum E. coli
L-Lysine PHAs L-Valine Naringenin/ Glucaric acid
E. coli/E. coli Evolutionary engineering
Improving tolerance
www.sciencedirect.com
Improving efficiency of pathway
Global optimization of pathway a
N/A, not applicable.
Effect of tool/strategy utilization Enable fast screening of knockout target ( pfkA) required and led to 169.5% increase in titer Identified ade3 as engineering target for enhanced production Malonyl-CoA pool was increased by repression of fatty acid biosynthesis genes ( fabD, fabH, fabB and fabF) resulting in enhanced production of its derivatives Five gene related to mevalonate biosynthesis were deleted in combinatorial way and found best combination More than 100 genetic variants related of MEP pathway were generated and best mutant harboring 15 mutations successfully produced 2.0 g/L of b-carotene Enhanced production in C. glutamicum was achieved through separate downregulation of pyk, pck and pgi gene Multiple knockdown of sad, sucD, sucC, sdhA and sdhB genes increased 4-hydroxybutyrate fraction in PHAs Intermediate toxicity problem was solved by using protein scaffold resulting in 77-fold increase in production titer GlcNAc production was optimized by spatial organization of GlmS and GNA1 in 1:2 ratio resulting in 20.58 g/L production Isobutanol production pathway was constructed in mitochondria resulting in 260% increase in production titer Ethanol production pathway was constructed in BMC and production has enhanced Demonstrated direct fermentation of lignocellulosic biomass into isobutanol Demonstrated direct fermentative production oxygenated taxanes from xylose Succeeded high yield production from mixed xylose and glucose Increased tolerance against L-arginine analogues and production of L-arginine by 2-fold Increased tolerance to 35 g/L by NTG and further enhancement to 50 g/L by genome shuffling resulting in increased IBE production Three synergistic knockdown targets which confers improved tolerance were identified Create CrtYB with solely phytoene synthase function and increase activity of CrtE resulting in 2-fold increased production Increased production by 89% Increased PHB accumulation contents up to 92% Increased production by 100% Increased production of naringenin and glucaric acid by 36-fold and 22-fold
References [47] [63] [65,66]
[73] [74]
[76]
[77] [78] [79]
[80]
[81] [83] [84] [85] [86] [88]
[97] [87]
[92] [93] [91] [95]
70 Tissue, cell & pathway engineering
Current Opinion in Biotechnology 2017, 47:67–82
Table 1 (Continued )
Recent systems metabolic engineering tools and strategies Chae et al. 71
distribution towards product formation. To quantitate the flux distribution, 13C metabolic flux analysis (13C-MFA) is usually applied using isotope labeling patterns of metabolites as direct experimental parameters [5]. The flux calculation using 13C-MFA is becoming more accurate as various methodologies are in development to increase the precision. As one example, study on selecting optimal tracer for parallel labeling experiment has been reported [6]. Nowadays, multi-omics approach, which analyzes various omics data together, rather than just a single type of omics, is usually applied. The multi-omics approach can offer comprehensive understanding of the host strain by compensating each omics’ drawbacks. The multiomics approach can be used to elucidate various phenomena in a metabolically engineered strain and to identify further engineering targets. For example, by analyzing metabolomic and fluxomic data, low glycolysis flux was revealed to be one of the potential reasons for inefficient ethanol production by Saccharomyces cerevisiae when using xylose as carbon source [7]. More recently, transcriptomics and proteomics were applied simultaneously to reveal underlying mechanism for sucrose and glycerol dual carbon source utilization of homo-succinic acid producing Mannheimia succiniciproducens PALFK strain. The analysis showed that deletion of fruA gene encoding fructose phosphotransferase system (PTS) system allowed simultaneous uptake of sucrose and glycerol by increasing cAMP level, which activates glycerol utilization pathway [8]. Using the PALFK strain, succinic acid could be produced with titer, yield and productivity of 78.4 g/L, 1.64 mol/mol glucose equivalent, and 6.02 g/L/ h, respectively. When a membrane cell recycle bioreactor system was used, this engineered strain was able to produce succinic acid with an impressively high productivity of 38.6 g/L/h [8]. Multi-omics approaches are not only applied to genetically/metabolically engineered strains, but also to wildtype host strains to select potentially the best host strain to work with for further metabolic engineering. Clostridium tyrobutyricum is a native butyric acid producer and has a potential for further enhanced production of butyric acid and its derivative products through metabolic engineering. However, metabolic engineering studies have not been extensively performed on C. tyrobutyricum due to the limited information on its genetic and metabolic characteristics. To address this problem, genomic and proteomic studies were performed, which revealed interesting metabolic characteristics very different from other clostridial species; butyric acid formation does not rely on the well-known phosphotransbutyrylase and butyrate kinase pathway, but rather uses CoA transferase system [9]. In another study, strain specific differences of seven widely used Escherichia coli strain (BL21, C, Crooks, DH5a, K-12 MG1655, K-12 W3110 and W) were quantified using www.sciencedirect.com
multi-omics (genomics, phenomics and transcriptomics) together with in silico approaches. Through integrated modeling, the potentials and capabilities of these seven strains for producing 40 selected chemicals were quantitatively evaluated in both aerobic and anaerobic conditions [10].
Genome-scale metabolic model (GEM) As the first whole genome sequence was reported in the mid-1990s, researches on the GEM opened new era in metabolic engineering. Nowadays, GEM and various simulation tools have become a fundamental framework for the system-wide analysis and prediction of cellular metabolisms and functions, and also for providing system metabolic engineering strategies. Accordingly, various GEMs have been developed for various metabolic engineering platform strains such as E. coli, Corynebacterium glutamicum, Ralstonia eutropha, C. acetobutyricum, M. succiniciproducens, S. cerevisiae, microalgae, and many others [11]. To utilize GEMs for genome-scale metabolic analyses, various algorithms based on constraint-based flux analysis have been developed to predict genetic alteration targets such as gene knockout or amplification targets (Figure 1b). As detailed algorithms applied to each tools have already been extensively reviewed in several papers [11,12,13], here we will focus more on applications examples. Several computational tools for identifying gene knockout targets for increased production of target compound have been developed [11]. MOMA is one of the early algorithms developed and widely used for identifying gene knockout targets to enhance production of various chemicals such as lycopene [14], L-valine [15], poly(lactic acid) [16] and terpenoid [17]. As a recent example, ten novel knockout targets for terpenoid production in S. cerevisiae were identified and once deleted from the chromosome, most resulted in enhanced terpenoid yield (8 to 10-fold) compared to the wild-type [17]. OptKnock is another popular algorithm for identifying gene knockout targets and has been used in metabolic engineering for the enhanced production of various chemicals. One representative example is 1,4-butanediol production [18]. In detail, OptKnock simulation suggested ldhA, pflB, adhE and mdhA as knockout targets in anaerobic condition. Three target genes (ldhA, pflB and adhE) were deleted to implement the simulation result but the mutant resulted in no cell growth. The problem was solved by introducing anaerobically functioning pyruvate dehydrogenase E3 subunit (lpdA gene from Klebsiella pneumonia) and changing culture condition to micro-aerobic. In the modified condition, deletion of all suggested knockout targets (ldhA, pflB, adhE and mdhA) finally led to increased production of 1,4-butandiol [18]. RobustKnock is an algorithm derived from OptKnock to overcome the potential optima problem caused by competing pathways [19]. OptSwap is another algorithm based on RobustKnock with additional function Current Opinion in Biotechnology 2017, 47:67–82
72 Tissue, cell & pathway engineering
of optimizing cofactor specificities of oxidoreductases. OptSwap has been successfully applied to an E. coli model to theoretically design efficient producers of L-alanine, succinate, acetate and lactate [20]. Cofactor modification analysis (CMA) allows finding targets for cofactor specificity by monitoring flux distribution changes [21]. Several algorithms have also been used for identifying gene amplification targets. Several notable ones include flux response analysis, flux scanning based on enforced objective flux (FSEOF), and flux variability scanning based on enforced objective flux (FVSEOF) with grouping reaction (GR) [11]. By using flux response analysis, ppc gene was selected as an overexpression target to enhance fumaric acid production in E. coli. Experimental result was consistent with the simulation prediction, resulting in 2.8-fold higher production of fumaric acid [22]. Similarly, FSEOF was used to identify gene amplification targets for enhanced lycopene production. Indeed, overexpression of idi and mdh genes identified by FSEOF in E. coli increased lycopene production by 2.7fold [23]. FVSEOF with GR constraints was developed to overcome limitations of FSEOF such as alternative optima and lack of thorough consideration of physiological state. It was applied to metabolically engineered E. coli strain producing putrescine, and successfully predicted glk, acnA, acnB, ackA, and ppc genes as amplification targets. All of these target genes enhanced putrescine production by 20.5% upon their amplification on average [24]. Although various algorithms have been developed to reduce the solution space of GEMs, it is still not enough for accurate prediction of cellular metabolic phenotypes. To address this problem, GEMs integrated with omics data have been developed to build more accurate condition-specific models (Figure 1c). Various tools for integrating transcriptomic data with GEMs have been extensively developed such as gene inactivity moderated by metabolism and expression (GIMME), integrative metabolic analysis tool (iMAT), E-Flux, E-Flux2, probabilistic regulation of metabolism (PROM) and transcriptomics-based strain optimization tool (tSOT) [11,25,26]. Furthermore, proteomic and metabolomic data have also been integrated with GEMs by gene inactivity moderated by metabolism and expression by proteome (GIMMEp) and gene inactivation moderated by metabolism, metabolomics and expression (GIM3E), respectively [11]. Recently, tSOT was applied for secondary metabolite overproduction in Streptomyces (actinorhodin) coelicolor. Overexpression of tSOT targets, ribulose 5phosphate 3-epimerase and NADP-dependent malic enzyme, resulted in 2 and 1.8-fold increased production, respectively [26]. Also, GEMs have expanded the capabilities to integrate protein structure as in GEM-PRO [27], metabolism and protein expression as in metabolism with gene expression (ME) model [28], and many Current Opinion in Biotechnology 2017, 47:67–82
cellular processes by modular simulations at separate time points to give a whole-cell model [29], the details of which described in two recent minireviews [30,31].
Pathway prediction Computer algorithms have also been developed to predict the synthetic pathways for biosynthesis of various natural and non-natural chemicals (Figure 1d); these algorithms include biochemical network integrated computational explorer (BNICE), RetroPath, GEM-Path, OptStrain and DESHARKY [11]. Recently, a systematic approach to select enzyme for designed biosynthetic pathway was developed based on semisupervised Gaussian process. The system has successfully predicted a promiscuous enzyme responsible for generating N-acetyl-L-leucine and the most efficient metabolic pathway for producing pinocembrin. Furthermore, quantitative prediction of Michaelis constant (KM) value for given enzyme-substrate pair was possible [32]. GEM-Path is another great tool that has successfully been used to identify 245 unique heterologous pathways for 20 commodity chemicals based on reaction promiscuity [33]. Computational metabolic pathway search with analysis tools including MapMaker and PathTracer [34] were used to suggest the great potential of E. coli to produce 1777 non-native chemicals, of which 279 have commercial usage [35]. As synthetic pathway design is no longer constrained for any target molecule, identification of useful and industrially important chemicals to produce is becoming increasingly important. One of the promising and less explored fields is traditional oriental medicine (TOM). Recently, systematic analysis of TOM-derived compounds with natural human metabolites revealed potentially active components. Although some more time will be needed to fully characterize active compounds and their mechanisms of action, these molecules will be promising targets to produce by metabolic engineering [36].
Enzyme design As more and more advanced metabolic engineering strategies are being developed for producing increasing number of chemicals, the need for developing or creating enzymes with new characteristics such as novel activity, enhanced activity or enhanced stability are keep increasing. Such new or improved enzymes can be created with computational protein design tools (extensively reviewed in [37,38]), which can identify core parts of protein structure, provide target sites for engineering, and even allow de novo protein design from the scratch (Figure 1e). These protein design tools are now widely used in metabolic engineering fields. Recently, a novel aldehyde dehydrogenase (GabD4 from R. eutropha) which can efficiently convert 3-hydroxypropionaldehyde (3-HPA) to 3-hydroxypropionic acid (3-HP) was identified. To further enhance the activity, homology modeling was used to predict crystal structure and select engineering target site. When the target site was engineered by www.sciencedirect.com
Recent systems metabolic engineering tools and strategies Chae et al. 73
site-directed mutagenesis (SDM) and site-saturation mutagenesis (SSM), a mutant GabD4 having 1.4-fold higher activity compared to the wild-type enzyme was obtained. The use of this mutant GabD4 resulted in production of 71.9 g/L of 3-HP [39]. In another example, Siegel and coworkers created artificial and highly efficient one-carbon assimilation pathway by de novo computational design of a new enzyme, named as formolase [40].
Synthetic biology tools and strategies for metabolic engineering Synthetic biology that covers DNA synthesis, synthetic protein scaffolds design, construction and modulation of synthetic expression components, design and construction of novel metabolic pathways, complex cellular network and even multicellular systems helps us to develop designable and reprogrammable organism with the objectives [1,41]. Synthetic biology can complement traditional metabolic engineering by providing various strategies including those described above for producing non-native and non-natural chemicals and materials and optimizing cellular fluxes. Some of the key contributions of synthetic biology in metabolic engineering are described below.
Pathway construction Enzymes found from nature can be used for establishing novel metabolic pathways, and often the enzyme promiscuity further enlarges the space of biological reactions that can be realized. For example, succinic semialdehyde dehydrogenase (encoded by the yneI gene) in E. coli was recently found to have promiscuous activity on malonic semialdehyde. The finding has expanded E. coli metabolism and allowed microbial production of malonic acid for the first time [42]. Furthermore, new metabolic pathway can be designed rationally or through the use of in silico computational tools, followed by experimental development of corresponding enzymes and metabolic pathways. Development of synthetic pathways to produce increasing number of currently petroleum-based chemicals showcases great potential of microbial cell factories towards moving into bio-based chemical industry. Various chemicals have been produced by engineered microorganisms harboring diverse synthetic pathways. Recent representative examples include fuels (short chain alkane [43], branched alcohols [44]), polymers (poly(lactate-co-glycolate) [45]), nylon precursors (adipic acid [46], 1,3-diaminopropane [47], 6-aminocaproic acid [48]), and natural products (opoids [49], protoilludene [50], artemisinic acid [51]). In one example, the nondecarboxylative Claisen condensation (mediated by thiolase) and subsequent b-reduction reaction were used to synthesize 18 chemicals of 10 classes including v-hydroxy, v-1-oxo, 2-methyl acids and v-1-methyl alcohols [46]. The selected starting chemicals termed as primers (phenylacetyl, succinyl, glutaryl, isobutyryl, acetyl, propionyl and glycolyl-CoAs) and extenders (acetyl, propionyl and www.sciencedirect.com
glycolyl-CoAs) were combinatorially fed into E. coli expressing thiolase and reductase, which yielded various important chemicals such as adipic acid [46]. In addition to bulk chemicals, many natural products that can be used as flavors, fragrances, nutraceuticals, and pharmaceuticals normally derived from plants are receiving much attention. This is because the chemical synthesis of such natural products is rather difficult due to the complex molecular structures with regio-selectivities and enantioselectivities, while current plant extraction-based process typically results in extremely low yield. One of the representative examples is production of artemisinic acid which can be chemically converted to potent antimalarial drug, artemisinin. Systematic metabolic engineering of S. cerevisiae allowed production of artemisinic acid to more than 25 g/L [51], and the upgraded process of which has been commercialized.
Pathway optimization Once desired synthetic pathway is constructed, the next step is optimizing the metabolic pathway fluxes to improve product titer, productivity and yield in order to achieve an economically viable bioprocess. Traditional metabolic engineering approaches allowed pathway flux optimization to good extents, but more precise and predictable tools for metabolic flux control are required to develop higher performance strains with less time and effort. An intuitive way to fine-tune metabolic flux is to control gene expression level by modulating gene expression components such as promoter, ribosomal binding site (RBS), terminator, 30 or 50 untranslated region (UTR) and transcription factor (Figure 1f). These components can be designed to accomplish desired expression levels of target genes. Optimizing promoter sets are one of the most widely used strategy and recently applied successfully to violacein production in E. coli [52]. Violacein is synthesized from L-tryptophan through five enzymatic reactions catalyzed by VioA, B, E, D and C followed by a non-enzymatic reaction. Five mutant T7 promoters having different strengths were designed and used to express the five violacein biosynthetic genes in various combinations; 3125 variants were constructed. The best violacein producer showing 63-fold improvement over the control strain could be obtained [52]. The 50 UTR, especially RBS, are also widely used to regulate gene expression level. Various computational tools to design RBS were developed such as RBS calculator [53], RBS designer [54] and UTR designer [55]. Recently, a tool called RBS library calculator [56] was developed for designing RBS’s to control expression of multiple genes. This tool can generate a set of RBS’s to cover broad expression spaces of multiple genes (proteins) across 10,000-fold range. Using this RBS library calculator, 73 RBS variants for three genes (crtEBI) of Current Opinion in Biotechnology 2017, 47:67–82
74 Tissue, cell & pathway engineering
carotenoid biosynthesis were designed and tested. As a result, optimal expression levels of crtEBI genes for carotenoid production could be found. As an alternative approach different from computational tools, the empirical model and oligos for protein expression changes (EMOPEC), the free tool for designing RBS based on empirical data was developed [57]. This tool allowed modulating expression level of any gene in E. coli strain within 2-fold of desired target expression level by changing a few base pairs. The design tools described above enable better forward metabolic engineering through more easily designable and controllable expression of genes. Another powerful and well-studied tool for controlling gene expression employs regulatory RNAs. Regulatory RNAs can control gene expression by interacting with DNA, RNA, protein and metabolite at different stages such as mRNA degradation, translation and transcription elongation [58]. Among various regulatory RNAs, transacting RNAs, which work based on recognition of target DNA or RNA sequences by the base-pairing rule, are modular, tunable and designable. Due to such fascinating characteristics, trans-acting RNAs have been employed extensively for genome-wide engineering and fine control of gene expression in various forms such as synthetic small regulatory RNA (sRNA), RNA interference (RNAi), antisense RNA (asRNA), small transcription activating RNAs (STARs) and CRISPRi. The sRNA system (Figure 1g) was developed for efficient attenuation of target gene expression in genome-wide scale, and has been demonstrated for efficient metabolic engineering of E. coli [59,60]. The system consists of a MicC scaffold recruiting Hfq protein and a target-specific sequence binding to the translation initiation region of the target gene. Once Hfq protein is bound to MicC scaffold, Hfq protein facilitates binding of sRNA with target mRNA as well as degradation of mRNA. The target binding sequence of sRNA can be forward designed to achieve desirable level of knockdown. Using the sRNA system, gene knockdown can be achieved in more rapid, easy, modular, portable, multiplex, fine-tunable and undoable manner, offering unique advantages over conventional genome engineering. The sRNA knockdown system was successfully applied to metabolic engineering of E. coli for production of various chemicals such as cadaverine, L-tyrosine [59], phenol [61] and 1,3-diaminopropane (1,3-DAP) [47]. In the case of 1,3-DAP production, 128 synthetic sRNAs targeting the genes encoding glycolytic enzymes, TCA cycle enzymes, transcription factors, transporters and cell division proteins were applied and 11 positive targets leading to enhanced 1,3-DAP production were identified [47]. Among these 11 knockdown targets screened, the three most effective and nonessential targets ( pykF, ptsI and pfkA) were knocked out; knockout of the pfkA gene led to 165.9% increase in Current Opinion in Biotechnology 2017, 47:67–82
1,3-DAP titer. The sRNA system was also successfully applied to C. acetobutylicum [62]. Since C. acetobutyricum is one of the most difficult bacteria to genetically engineer, the sRNA technology has a great potential as a new easy-to-use and rapid knockdown tool in metabolic engineering of clostridial species. As demonstrated for Clostridia, the sRNA tool can be further expanded to facilitate systems metabolic engineering of various microorganisms. RNAi, a gene silencing system in eukaryotes, can also be used to control and optimize metabolic fluxes (Figure 1g). The key component of RNAi system is an RNA-induced silencing complex (RISC) protein. The RISC protein is guided to the target mRNA by a small interfering RNA (siRNA), which is generated by degradation of foreign double stranded RNA by a protein called dicer. Then, the target mRNA is cleaved by argonaute, which is an essential catalytic component of RISC protein. As a metabolic engineering tool, RNAi system has been introduced to S. cerevisiae by expressing hairpin structure RNA (degraded into siRNA by dicer), dicer and argonaute using plasmids [63]. The system was applied to itaconic acid production, and suggested the potential of ade3 gene as an engineering target and Sigma 10560-4A as a promising base strain [63]. Another example is asRNA, a single-stranded RNA complementary to the target mRNA (Figure 1g). Once asRNA binds to the target mRNA, ribosome cannot bind to the mRNA thereby blocking translation. The asRNA system has been successfully applied to knocking down gene expression in C. acetobutylicum [64], a strain that is extremely difficult to genetically manipulate. As mentioned earlier, the sRNA system is now added for gene expression knockdown control in C. acetobutyricum [62]. Recently, the asRNA system has also been applied to E. coli. Using the loop-stem structure synthetic asRNA having complementary sequences to the target at the loop position, the fatty acid biosynthesis genes ( fabD, fabH, fabB and fabF) were effectively repressed to enhance intracellular malonyl-CoA concentration. This resulted in enhanced production of resveratrol, naringenin and 4-hydroxycoumarin [65,66]. While above mentioned RNA systems have been used to down-regulate gene expression, recently developed STARs (Figure 1h) demonstrate that the small RNAs can also be used to up-regulate gene expression [67]. The STARs are designed starting from the natural pT181 attenuator to disrupt the formation of a target transcriptional terminator hairpin placed upstream of a gene. This plasmid-based system was highly orthogonal and showed up to 94-fold activation. However, application of the STARs system is limited for metabolic engineering due to the inability to increase the expression level of chromosomal genes. It can be used as a new tool in constructing logic gates in synthetic circuits [67]. www.sciencedirect.com
Recent systems metabolic engineering tools and strategies Chae et al. 75
CRISPR-Cas system is a prokaryotic adaptive immune system that has been receiving much interest for genome engineering of various hosts, especially eukaryotes, since it enables rapid, simple and robust engineering compared with the conventional tools. Among various types of CRISPR-Cas system, the type II system is mainly used for genome engineering application due to its simplicity. In the type II system, trans-activating CRISPR RNA (tracrRNA) together with CRISPR RNA (crRNA) are used to direct Cas9 to the target DNA sequence. Cas9, an endonuclease, introduces double strand break. The tracrRNA and crRNA can be combined into a single guide RNA (sgRNA), which makes use of the type II system more convenient. Differently from the RNA-based systems described above, CRISPR-Cas system can be used to directly engineer DNA, thereby enabling gene knockout as well as insertion [Figure 1i; [68]]. CRISPR-Cas based system has been used to engineer various host strains including Bacillus subtilis, Clostridium sp., E. coli, Lactobacillus reuteri, S. cerevisiae, Streptococcus sp. and Streptomyces sp. [68]. One of the attractive advantages of the CRISPR-Cas system for metabolic engineering is its multiplex capability. Multiplex genome engineering has been demonstrated in E. coli [69], B. subtilis [70], Streptomyces sp. [71] and S. cerevisiae [72]. In the case of S. cerevisiae, all possible combinatorial knockout of 5 genes related to mevalonate production were rapidly constructed to find the best combination [73]. In another study, CRISPR-Cas system was used in multiplex mode to generate various types of genomic modifications in E. coli to produce b-carotene [74]. The system could integrate the b-carotene synthetic pathways into E. coli genome to combinatorially optimize methylerythritolphosphate (MEP) pathway together with central metabolic pathway for improved b-carotene production. Among more than 100 genetic variants comprising 33 genomic modifications, the strain (XF237T) having 15 mutations showed the highest b-carotene production (2.0 g/L titer) by fed-batch fermentation [74]. The CRISPR-Cas system can be also used for gene downregulation by using a deactivated Cas9 (dCas9) protein. The dCas9 protein has the target DNA binding capability without the endonuclease activity, and thus can block transcriptional initiation or elongation [Figure 1i; [75]]. This phenomenon called CRISPR interference (CRISPRi) has been applied in several metabolic engineering hosts including B. subtilis, Clostridium sp., C. glutamicum, E. coli and Streptomyces sp. [68]. For example, the pyk, pck and pgi genes in C. glutamicum were separately downregulated using the CRISPRi system, which resulted in the increase of L-lysine and L-glutamate titers comparable to those obtained by corresponding gene deletions [76]. This CRISPRi based system was also applied for increasing the 4-hydroxybutyrate monomer fraction in poly(3-hydroxybutyrate-co-4-hydroxybutyrate) www.sciencedirect.com
in E. coli through knocking down multiple genes (sad, sucD, sucC, sdhA and sdhB) simultaneously [77]. Another way to control the metabolic fluxes is spatial modulation of enzymes using synthetic protein or DNA scaffold (Figure 1j). When the enzymes comprising the metabolic pathway are located closely, the product of the first enzyme more efficiently reaches the next enzyme with low probability of side reactions by other enzymes. Thus, metabolite channeling through this approach is particularly advantageous when the biosynthetic steps include unstable or toxic intermediates. This strategy was first adopted using protein scaffold for the production of mevalonate, resulting in 77-fold increase of titer by resolving intermediate toxicity problem [78]. In another recent example, B. subtilis producing N-acetylglucosamine (GlcNAc) was developed by spatial organization of the two key biosynthetic enzymes (GlmS and GNA1) in various ratios using a DNA scaffold system; the GlmS and GNA1 scaffolds at 1:2 ratio showed the highest GlcNAc production up to 20.58 g/L by fed-batch fermentation [79]. Oppositely from the protein scaffold concept, spatial separation of enzymes from other enzymes or cell components can also be beneficial for metabolic flux optimization in some cases (Figure 1k). This can be done by using natural organelles in eukaryotic cells. Although prokaryotic cells are mostly lacking in specific organelles, spatial separation of enzymes can also be achieved by using bacterial microcompartment (BMC). Encapsulation of enzymes within natural organelles or BMCs can increase the metabolic intermediate concentration, reduce the effect of toxic intermediates to general cell metabolism, and offer beneficial environment that is different from the cytosol. As an example, this strategy was applied to isobutanol production in S. cerevisiae [80]. When an isobutanol biosynthetic pathway was constructed using an a-ketoacid decarboxylase and alcohol dehydrogenase in mitochondria, isobutanol production increased up to 260% mainly due to greater local enzyme concentration. Also, BMCs were successfully used to enhance ethanol production in E. coli by encapsulating two enzymes, pyruvate decarboxylase and alcohol dehydrogenase. In addition, it was demonstrated that the purified BMCs containing two enzymes could produce ethanol from pyruvate [81]. Another approach to optimize metabolic fluxes is employing a co-culture system that involves a consortia of multiple microorganisms (Figure 1l). Coculture system can provide several advantages such as efficient utilization of various carbon sources and no need to construct a long and complex metabolic pathway in one host strain, thereby avoiding potential metabolic burden [82]. In one example, Trichoderma reesei, which hydrolyzes lignocellulosic biomass into simple sugars, and engineered E. coli, Current Opinion in Biotechnology 2017, 47:67–82
76 Tissue, cell & pathway engineering
which utilizes the produced sugars to biosynthesize isobutanol, were cocultured to directly produce isobutanol from cellulosic biomass [83]. Also, an engineered E. coli strain capable of producing taxadiene and S. cerevisiae capable of converting taxadiene to oxygenated taxanes were cocultured to directly produce oxygenated taxanes from xylose [84]. The E. coli–E. coli consortia was also employed for the efficient production of various chemicals such as cis,cis-muconic acid, 4-hydroxybenzoic acid, 3aminobenzoic acid, butanol, and flavan-3-ols [82,85]. Although coculture system is an attractive strategy for producing chemicals and materials through synergistically using the advantages of different host strains, there are also various challenges to be solved as systems gets more complex compared to monoculture system. Maintaining the stability and population ratio of cocultured species as originally designed is a key challenge. This might become even more difficult in industrial scale.
Evolutionary engineering tools and strategies for metabolic engineering In developing an industrial microbial strain, several interlinked complex factors should be considered, such as product titer, yield and productivity, byproducts formation, product toxicity and tolerance, and cell growth rate in the context of fermentation (culture medium, culture condition, and feeding mode) and downstream processes. In some cases, it is difficult, if not impossible, to rationally optimize the cellular performance considering such complex factors, especially when links between genotype and phenotype are unknown due to the lack of sufficient prior knowledge. Evolutionary engineering is a powerful method to address this problem. Evolutionary engineering, which mimics natural evolution process but performed in much more rapid, rational, and efficient ways, allows improvement of cellular phenotype through applying smart evolution pressure. For convenience, evolutionary engineering strategies are divided into two types, directed and semi-rational evolution, in this paper.
Directed evolution In directed evolution, increasing genetic diversity is performed in random way by using numerous mutagenesis methods (Figure 1m). Traditional random mutagenesis methods using error prone-PCR (ep-PCR) and/or applying the chemical and physical mutagens are still widely used. As an example, a C. glutamicum mutant strain having increased tolerance against L-arginine analogues could be obtained by N-methyl-N-nitro-N0 -nitroguanidine (NTG) and ultraviolet treatment. In addition to tolerance enhancement, this mutant strain could produce 34.2 g/L of L-arginine, which was 2-fold higher than that obtained with the parent strain. Further systems metabolic engineering of this mutant strain allowed development of an engineered C. glutamicum strain capable of producing 92.5 and 81.2 g/L of L-arginine by laboratory and industrial scale fermentation, respectively [86]. EpCurrent Opinion in Biotechnology 2017, 47:67–82
PCR was recently used to optimize production of lycopene in S. cerevisiae by creating CrtYB only possessing phytoene synthase function and by enhancing the catalytic activity of CrtE. Additional metabolic engineering of this strain produced 1.61 g/L of lycopene [87]. Increasing genetic diversity has been further accelerated through various techniques such as genome shuffling, staggered extension process (StEP), incremental truncation for the creation of hybrid enzymes (ITCHY) and nonhomologous random recombination (NPR). Genome shuffling was applied to C. beijerinckii DSM6423 strain to enhance its tolerance to isopropanol. As a first step, the wild-type strain was chemically mutated with NTG to obtain a mutant strain which can tolerate more than 35 g/L of isobutanol. Then, the selected mutant was subjected to genome shuffling. As a result, mutant tolerating up to 50 g/L of isopropanol was obtained. The resulting strain also showed enhanced isopropanol/butanol/ethanol (IBE) titer [88]. When a large library of evolved strains is used, a proper screening method is required to select mutants with desired characteristics. If the desired characteristic can be linked with distinguishable features such as color and fast growth, the desired mutant can be easily screened using instruments such as fluorescence-activated cell sorting (FACS), calorimetric assays, spectrophotometer or microfluidic devices. One of the easy characteristics to be evolved is tolerance towards the target chemical since it can be directly accessed by measuring growth rate in presence of the target chemical or its derivatives. Also, enhanced production of a target chemical or reduced production of a byproduct can be easily screened if the product or byproduct has its own color. As an example, carotenoid hyper-producing S. cerevisiae strain could be screened using oxidative stress as a selection pressure. The resulting mutant strain accumulated carotenoid up to 18 mg/g dry cell weight, which was 3-fold higher than that obtained with the parent strain [89]. When there are no easy intuitive ways of screening the desired phenotype, biosensors can be developed and used to link production capability with distinguishable characteristics such as color (e.g., by triggering expression of fluorescence protein) or antibiotic resistance (e.g., by triggering expression of antibiotic resistance gene) (Figure 1n). Numerous biosensors for screening metabolites such as butanol, fatty acid, isopentenyl pyrophosphate, L-lysine, L-tryptophan, L-valine, malonyl-CoA, NADPH, naringenin and succinic acid have been developed [90]. In addition, transcriptional regulators are also widely used as biosensor. For example, an L-valine responsive sensor based on Lrp transcriptional regulator of C. glutamicum was developed and used for adaptive laboratory evolution for enhanced production of L-valine in C. glutamicum by 100% [91]. Riboswitches are also used to construct biosensors. Riboswitches regulate gene expreswww.sciencedirect.com
Recent systems metabolic engineering tools and strategies Chae et al. 77
sion at the translational and transcriptional level by changing the conformations of secondary RNA structures when target metabolites are bound. For example, an Llysine riboswitch was constructed for the dynamic control of L-lysine exporter lysE, which successfully enhanced Llysine production in C. glutamicum [92].
Semi-rational evolution Generating genetic diversity can proceed at single genelevel, pathway-level and even at whole genome-levels. Since the random mutation space becomes explosively larger when the system to be evolved goes from single gene-level to the whole genome-level, it become extremely difficult to screen the right mutant/variant if the library is generated in a completely random way; it becomes nearly impossible to cover all random mutants when it becomes genome-level. To address this issue, various methods which can introduce mutations in semirational manner have been developed. By using these methods, focused libraries which contain fewer mutants but are richer in beneficial mutations can be generated, thereby making the desired screening easier. A semi-rational evolution approach of using rationally designed RBS libraries was taken to optimize metabolic fluxes for poly(3-hydroxybutyrate) (PHB) production in E. coli [93]. Expression of phaC, phaA and phaB genes involved in PHB biosynthesis was regulated by the RBS library, which was constructed by oligo-linker mediated assembly (OLMA) method. Then, the library was screened by visual selection and high-throughput screening (FACS), which allowed identification of strain variants accumulating PHB to the contents of 0–92%. This kind of semi-rational approach, which combines directed evolution and rational design, is likely to be more widely applied to developing strains capable of producing other products. In order to make focused libraries at genomic level, oligo mediated targeted mutation generation methodologies can be used (Figure 1o). One of the popular methods is multiplex automated genome engineering (MAGE), which can increase the diversity of the targeted sites by using recursive rounds of single-stranded DNA (ssDNA) recombination-based engineering in E. coli strain [94]. Recently, MAGE was applied to enrich rare high-producing strains and enhance production of naringenin and glucaric acid [95]. More recently, a modified method of MAGE called yeast oligo-mediated genome engineering (YOGE) was developed for genome engineering of S. cerevisiae [96]. As another approach for genome-scale evolution, RNAi-assisted genome evolution (RAGE) was developed and used in engineering S. cerevisiae. Application of RAGE has identified three synergistic knockdown targets conferring improved tolerance against acetic acid [97]. www.sciencedirect.com
Synergistic applications of the tools and strategies of systems biology, synthetic biology and evolutionary engineering: systems metabolic engineering All the tools and strategies of systems biology, synthetic biology and evolutionary engineering describe above can be simultaneously used in synergistic manner in order to construct high performance microbial cell factories. Two representative examples are shown below to demonstrate how various tools and strategies can be synergistically employed to perform systems metabolic engineering (Figure 2). Poly(lactate-co-glycolate) (PLGA) is US Food and Drug Administration (FDA) approved commercial polyester widely used for medical applications due to its biodegradability and biocompatibility. Recently, a metabolically engineered E. coli strain capable of producing PLGA was developed [Figure 2a; [45]]. Microbial production of this non-natural polyester PLGA could not be achieved until recently due to the lack of known natural metabolic pathway. Thus, a synthetic metabolic pathway for the biosynthesis of PLGA was designed by mimicking natural polyhydroxyalkanoates (PHAs) biosynthesis. The pathway is composed of two steps: activation of lactate and glycolate into lactyl-CoA and glycolyl-CoA, respectively, and copolymerization of lactyl-CoA and glycolyl-CoA. Since there is no natural enzyme catalyzing the polymerization of glycolyl-CoA, evolutionary engineering approach was taken to develop a PHA synthase variant [98]. However, introduction of the synthetic pathway to the wildtype E. coli XL1-Blue strain did not result in PLGA production due to the limited intracellular lactate and glycolate pools. This problem was solved by designing a new pathway through the application of systems biology tools. More specifically, the heterologous Dahms pathway from Caulobacter crescentus was introduced into E. coli to increase glycolate pool using xylose as a sole carbon source. However, the growth of the engineered strain was significantly retarded. To diagnose the cause for growth retardation, in silico genome-scale metabolic simulation was performed, which identified reduced ATP availability and poor glycolysis flux as the reasons. This problem was solved by simultaneous utilization of glucose through ptsG gene deletion for removal of catabolite repression. Furthermore, fermentative byproducts were removed (deletion of adhE, pflB, frdB and poxB genes) and lactate pool was enhanced (overexpression of ldhA gene and deletion of dld gene) through system-wide metabolic analyses. Finally, PLGA could be successfully produced from glucose and xylose. This example demonstrates the efficiency of systems metabolic engineering that integrates systems biology, synthetic biology and evolutionary engineering tools and strategies in strain development. Another great example of systems metabolic engineering is recent demonstration of opioids (specifically, thebaine Current Opinion in Biotechnology 2017, 47:67–82
78 Tissue, cell & pathway engineering
Figure 2
(a)
(b) Xylose
Glucose
In silico simulation
Glucose
Xylose Caulobacter crescentus
pTac
xylB
Aro4pQ166K
E4P
PTS
Synthetic pathway Glucose
Aro7pT266I
DAHP
F6P
4-HPP
TyrHWR
xylB NCS
Metabolic engineering
xylC
G6P
Xylonolactone
Metabolic engineering Chromosomal deletion Overexpression
6OMT CNMT NMCH 4’OMT
2-Keto-3-deoxy-xylonate
Bioinformatic analysis Search COR-like enzyme
Candidates selection
Verification
PEP Formate
IdhA
dId
oA
pct540
LA-C
PhaC1
Enzyme engineering
Isocitrate glcB aceB
Malate
Heterologous genes Mammalian Bacterial Plant
(R)-Reticuline Chimeric enzyme
Correctly Incorrectly processed & processed abnormal glycosylation (CFS) (SaISyn)
poxB
LA
GA
oA
LA-C
Chromosomal Plasmid gene
DRS -DRR
Ethanol
Acetate
Enzyme engineering
L-DOPA
Synthetic pathway
Retention time (h)
adhE
Ac-CoA
DODC
(S)-Reticuline
Domain Analysis (S)-reticuline CYP 82Y1-like domain COR-like domain (DRS-DRR)
Pyruvate
DOP
LC-MS (R)-reticuline
pflB
Glycolaldehyde
4-HPAA
NCCL
Elimination feedback inhibition Overexpression
xylC
Xylonate
L-TYR
PEP
C
N
Cytosol
C
fusion protein
ER matrix N
N
C
CFS1-83 SalSyn92-504
N C
Glyoxylate
SaISyn
Salutaridine
GA-CoA LA-CoA glcDEFG
PhaC1437 Pseudomonas sp. MBEL 6-19
PLGA
SaIR
Glycolate
phaC1437
Fumarate
frdB
Succinate
SaIAT
morB T6ODM
Hydrocodone
Thebaine
7-ASTDO
Current Opinion in Biotechnology
Two representative examples showing synergistic applications of system biology, synthetic biology and evolutionary engineering tools and strategies. (a) Production of poly(lactate-co-glycolate) from glucose and xylose by metabolically engineered E. coli. (b) Production of opoids (thebaine and hydrocodone) from glucose by metabolically engineering S. cerevisiae. Application of systems biology, synthetic biology, evolutionary engineering tools and strategies are represented by purple, blue and gray box, respectively. Genes and proteins are (in alphabetical order): aceB, malate synthase A; adhE, aldehyde-alcohol dehydrogenase; Aro4p, 3-deoxy-D-arabino-2-heptulosonic acid 7-phosphate synthase; Aro7p, chorismate mutase; CNMT, coclaurine N-methyltransferase; DRS-DRR, 1,2-dehydroreticuline synthase-1,2-dehydroreticuline reductase; dld, lacate dehydrogenase; DODC, 3,4-dihydroxyphenylalanine decarboxylase; frdD, fumarate reductase; glcB, malate synthase G; glcDEFG, glycolate oxidase; ldhA, lactate dehydrogenase; morB, morphinone reductase; NCS, (S)-norcoclaurine synthase; NMCH, N-methylcoclaurine hydroxylase; 40 OMT, 30 -hydroxy-N-methylcoclurine 40 -O-methyltransferase; 6OMT, norcoclaurine 6-O-methyltransferase; pct, propionyl-CoA transferase; phaC, polyhydroxyalkanoate synthase; pflB, pyruvate formate lyase; poxB, pyruvate dehydrogenase; SalAT, salutaridinol 7-Oacetyltransferase; SalR, salutaridine reductase; SalSyn, salutaridine synthase; T6ODM, thebaine 6-O-demethylase; TyrH, L-tyrosine hydroxylase; xylB, xylose dehydrogenase; xylC, xylonolactonase. Metabolites are (in alphabetical order): Ac-CoA, acetyl-CoA; 7-ASTDO, 7-O-acetylsalutaridinol; DAHP, 3-deoxy-D-arabino-2-heptulosonic acid 7-phosphate; DOP, dopamine; L-DOPA, 3,4-dihydroxyphenylalanine; E4P, erythrose 4-phophate; G6P, glucose 6-phosphate; GA, glycolate; GA-CoA, glycolyl-CoA; 4-HPAA, 4-hydroxyphenylacetaldehyde; 4-HPP, 4-hydroxyphenylpyruvate; LA, lactate; LA-CoA, lactyl-CoA; NCCL, (S)-norcoclaurine; PEP, phosphoenolpyruvate; L-TYR, L-tyrosine.
and hydrocodone) production in an engineered yeast strain [49] (Figure 2b). The synthetic biosynthetic pathways for thebaine and hydrocodone comprising 21 and 23 genes, respectively, were constructed by carefully selecting genes from plants, mammals, bacteria and yeast. As expected, it was not trivial to construct such complex whole pathway and make it functional in one strain. The biosynthesis of opioids requires isomerization of (S)-reticuline to (R)-reticuline. In Papaver species, (S)reticuline is converted to (R)-reticuline via its oxidation to 1,2-dehydroeticuline followed by stereospecific reduction. However, the genes encoding the required enzymes, 1,2-dehydroreticuline synthase (DRS) and 1,2-dehydroreticuline reductase (DRR), were unknown. Based on the previous finding that gene-silencing of codeinon reductase (COR) resulted in (S)-reticuline Current Opinion in Biotechnology 2017, 47:67–82
accumulation, bioinformatic searches were performed to find 38 COR-like enzymes. Interestingly, four of the identified enzymes possessed both COR-like domain and cytochrome P450 oxidase 82Y1-like domain in one open reading frame. Since epimerization of (S)-reticuline required both oxidation and reduction, it was considered that this natural fusion protein could catalyze the isomerization reaction. One of the fusion protein from P. bracteatum (Pbr.89405) was confirmed to have (S)-reticuline isomerization activity. Loss of the protein activity due to improper protein folding or processing is one of the most frequent challenges often encountered when genes from heterologous sources are used to construct a synthetic metabolic pathway. The same problem happened with SalSyn, which converts (R)-reticuline to salutaridine. The properly processed SalSyn anchors its N-terminus www.sciencedirect.com
Recent systems metabolic engineering tools and strategies Chae et al. 79
at the outer endoplasmic reticulum (ER) domain and the catalytic domain in the cytosol. However, SalSyn was incorrectly processed when expressed in yeast, resulting in catalytic domain at the rumen of ER and abnormal glycosylation. Simple modification of glycosylation site to change glycosylation pattern was unsuccessful. As an alternative, a chimeric protein was designed by fusion of SalSyn from P. somniferum and P. bracteatum with Cheilanthifoline synthase (CFS) from plant, which is 61–68% identical to SalSyn but properly processed without glycosylation. Several fusion proteins were found to be properly processed and showed efficient conversion of (R)-reticuline to salutaridine. This great example also demonstrates the importance of taking systems approach in efficient strain development through the integrated use of synthetic biology and evolutionary engineering.
Conflict of interest Authors declare that they do not have any conflict of interest.
Acknowledgements We would like to thank Won Jun Kim for helpful discussion. This work was supported by the Technology Development Program to Solve Climate Changes on Systems Metabolic Engineering for Biorefineries (NRF2012M1A2A2026556 and NRF-2012M1A2A2026557) and also by the Intelligent Synthetic Biology Center through the Global Frontier Project (2011-0031963) from the Ministry of Science, ICT and Future Planning through the National Research Foundation of Korea.
References and recommended reading Papers of particular interest, published within the period of review, have been highlighted as: of special interest of outstanding interest 1.
Lee JW, Na D, Park JM, Lee J, Choi S, Lee SY: Systems metabolic engineering of microorganisms for natural and nonnatural chemicals. Nat Chem Biol 2012, 8:536-546.
2.
King ZA, Lloyd CJ, Feist AM, Palsson BO: Next-generation genome-scale models for metabolic engineering. Curr Opin Biotechnol 2015, 35:23-29.
3.
Jensen MK, Keasling JD: Recent application of synthetic biology tools for yeast metabolic engineering. FEMS Yeast Res 2015, 15:1-10.
Conclusions and future perspectives Systems metabolic engineering has become a core technology for constructing efficient microbial cell factories. Recently, the power of systems metabolic engineering has been further enhanced by the development of advanced tools and strategies as discussed throughout the paper. Successful examples on developing microbial strains by systems metabolic engineering for the enhanced production of desired chemicals and materials are increasingly available. However, there are still more rooms for new tools and strategies that can further facilitate microbial cell factory development. Although genome-scale metabolic models have been successfully employed in flux optimization and strain design, more precise genome-scale metabolic model needs to be developed by incorporating missing metabolic information and integrating regulatory and signaling networks in simulation-competent mode. More accurate enzyme and pathway design tools for the production of nonnatural chemicals are also needed. In particular, a current bottleneck in systems metabolic engineering is to engineer or even create an enzyme for a desired reaction. Protein structure prediction tools are still not perfect, and forward design of proteins having altered reaction characteristics based on the structure is even more difficult, if not impossible. Additionally, determination of the structures of proteins and their mutant/engineered variants requires much time and effort. In this sense, a recent paper [99] reporting successful generation of structural models for hundreds of protein families of previously unknown structures is of great interest. Furthermore, more efficient, convenient, multiplex, robust, and large-scale genome manipulation tools still need to be developed. As such tools and strategies of systems metabolic engineering are being continuously developed and used in synergistic way, more efficient and cost-effective development of microbial cell factories suitable for the bio-based industrial production of chemicals and materials will become a reality. www.sciencedirect.com
4. Lee SY, Kim HU: Systems strategies for developing industrial microbial strains. Nat Biotechnol 2015, 33:1061-1072. The paper nicely described how systems metabolic engineering could develop normal microbes into industrial microbial strains. They represented general ten strategies in great details from host selection to optimization of bioprocesses for the industrial scale production. L-Arginine, 1,4-butanediol, and L-lysine producing microbial strains were exemplified as representatives to guide the journeys for the biocommercialization of chemicals. 5.
Gopalakrishnan S, Maranas CD: 13C metabolic flux analysis at a genome-scale. Metab Eng 2015, 32:12-22.
6.
Crown SB, Long CP, Antoniewicz MR: Optimal tracers for parallel labeling experiments and 13C metabolic flux analysis: a new precision and synergy scoring system. Metab Eng 2016, 38:10-18.
7.
Wasylenko TM, Stephanopoulos G: Metabolomic and metabolic flux analysis of a xylose-consuming Saccharomyces cerevisiae strain expressing xylose isomerase. Biotechnol Bioeng 2015, 112:470-483.
13
C-
8.
Lee JW, Yi J, Kim TY, Choi S, Ahn JH, Song H, Lee MH, Lee SY: Homo-succinic acid production by metabolically engineered Mannheimia succiniciproducens. Metab Eng 2016, 38:409-417. High productivity and yield of homo-succinic acid were achieved by M. succiniciproducens by integrated approaches of in silico genome scale metabolic flux analysis, omics, and reconstruction of central carbon metabolism. Their systematic metabolic approaches including dual carbon source usage, deregulation of carbon catabolite repression, and membrane cell recycle bioreactor system led to succeed in the yield and productivity for low-density and medium-density PALFK strain, which were 1.56 and 1.64 mol/mol glucose equivalent and 2.50 g/L/h and 6.02 g/L/h, respectively.
9.
Lee J, Jang Y, Han M, Kim JY, Lee SY: Deciphering Clostridium tyrobutyricum metabolism based on the whole-genome sequence and proteome analyses. mBio 2016 http://dx.doi.org/ 10.1128/mBio.00743-16.
10. Monk JM, Koza A, Campodonico MA, Machado D, Seoane JM, Palsson BO, Herrgard MJ, Feist AM: Multi-omics quantification of species variation of Escherichia coli links molecular features with strain phenotypes. Cell Syst 2016, 3:238-251. Seven commonly used E. coli strains were characterized by multi-omics data to give a framework for host selection based on the strain-specific features. The dataset from each strain were gathered in both aerobic and Current Opinion in Biotechnology 2017, 47:67–82
80 Tissue, cell & pathway engineering
anaerobic culture and showed that metabolic physiologies and expression level of encoding genes were widely different among seven strains. Their comprehensive data types from genomics, phenomics, transcriptomics, and genome-scale modeling were highly reliable, resulting in the validation of the link between phenotypes and molecular features. 11. Kim BJ, Kim WJ, Kim DI, Lee SY: Applications of genome-scale metabolic network model in metabolic engineering. J Ind Microbiol Biotechnol 2015, 42:339-348. 12. O’Brien EJ, Monk JM, Palsson BO: Using genome-scale models to predict biological capabilities. Cell 2015, 161:971-987. This paper described the entire processes how genome-scale models have been developed and utilized for understanding the cellular metabolism and predicting the phenotype under certain perturbations. Not only just its effect on system biology was introduced, but also related tools and methods to interpret the cellular behavior and the direction that has been made for the development were very nicely described. The six grand challenges from cell to systems biology were pointed out by using GEMs and this Primer paper summarized those related contents very nicely. 13. Mala P, Rocha M, Rocha I: In silico constraint-based strain optimization methods: the quest for optimal cell factories. Microbiol Mol Biol Rev 2016, 80:45-67. 14. Alper H, Jin Y, Moxley JF, Stephanopoulos G: Identifying gene targets for the metabolic engineering of lycopene biosynthesis in Escherichia coli. Metab Eng 2005, 7:155-164. 15. Park JH, Lee KH, Kim TY, Lee SY: Metabolic engineering of Escherichia coli for the production of L-valine based on transcriptome analysis and in silico gene knockout simulation. Proc Natl Acad Sci U S A 2007, 104:7782-7797. 16. Jung YK, Kim TY, Park SJ, Lee SY: Metabolic engineering of Escherichia coli for the production of polylactic acid and its copolymers. Biotechnol Bioeng 2010, 105:161-171. 17. Sun Z, Meng H, Li J, Wang J, Li Q, Wang Y, Zhang Y: Identification of novel knockout targets for improving terpenoids biosynthesis in Saccharomyces cerecisiae. PLoS ONE 2014 http://dx.doi.org/10.1371/journal.pone.0112615. The production of terpenoid, specifically amorphadiene, was enhanced from S. cerevisiae single knockout mutants. To improve the precursor isopentenyl diphosphate, several target genes were predicted by MOMA as deletion targets and those were predicted as targets that would not inhibit the yeast growth. Among them, hxk2 (encoding hexokinase) deleted mutant produced the highest titer of amorphadiene (54.55 mg/ L) which is 12-fold increase compared to wild type.
from transcriptomic data. PLoS ONE 2016 http://dx.doi.org/ 10.1371/journal.pone.0157101. 26. Kim M, Yi JS, Lakshmanan M, Lee DY, Kim BG: Transcriptomicsbased strain optimization tool for designing secondary metabolite overproducing strains of Streptomyces coelicolor. Biotechnol Bioeng 2016, 113:651-660. 27. Chang RL, Andrews K, Kim D, Li Z, Godzik A, Palsson BO: Structural systems biology evaluation of metabolic thermotolerance in Escherichia coli. Science 2013, 340:12201223. 28. Liu J, O’Brien E, Lerman J, Zengler K, Palsson BO, Feist AM: Reconstruction and modeling protein translocation and compartmentalization in Escherichia coli at the genomescale. BMC Syst Biol 2014 http://dx.doi.org/10.1186/s12918-0140110-6. The ME model of E. coli was reconstructed by including protein translocation and cell membrane composition. This reconstructed ME model computed proteome composition, abundance and activity in four types of membrane (cytoplasm, periplasm, inner and outer membrane) and predicted the cellular functions in the limited space of membranes. This model could be able to give the relationship between the energy efficiency in the pathway and the spatial need. Metabolic engineering strategies relating transport or membrane-bound enzymes could be greatly advanced through this model. 29. Karr JR, Sanghvi JC, Macklin DN, Gutschow MV, Jacobs JM, Bolival B, Assad-Garcia N, Glass JI, Covert MW: A whole-cell computational model predicts phenotype from genotype. Cell 2012, 2:389-401. 30. Karr JR, Takahashi K, Funahashi A: The principles of whole-cell modeling. Curr Opin Biotechnol 2015, 27:18-24. 31. O’Brien EJ, Palsson BO: Computing the function proteome: recent progress and future prospects for genome-scale models. Curr Opin Biotechnol 2015, 34:125-134. 32. Mellor J, Grigoras I, Carbonell P, Faulon JL: Semisupervised Gaussian process for automated enzyme search. ACS Synth Biol 2016, 5:518-528.
18. Yim H, Haselbeck R, Niu W, Pujol-Baxley C, Burgard A, Boldt J, Khandurina J, Trawick JD, Osterhout RE, Stephen R et al.: Metabolic engineering of Escherichia coli for direct production of 1,4-butanediol. Nat Chem Biol 2011, 21:488-494.
33. Campodonico MA, Andrews BA, Asenjo JA, Palsson BO, Feist AM: Generation of an atlas for commodity chemical production in Escherichia coli and a novel pathway prediction algorithm. GEM-Path. Metab Eng 2014, 25:140-158. Non-native compounds were targeted for their biosynthesis and those heterologous biosynthetic pathways were proposed through GEM-Path’s integration of GEM and reaction promiscuity analysis. For 20 commodity chemicals, they predicted 245 unique biosynthetic pathways and further analyzed their pathways by coupling with feasible growth-related production. GEM-Path can give a big outline for pathway design in chemical productions.
19. Tepper N, Shlomi T: Predicting metabolic engineering knockout strategies for chemical production: accounting for competing pathways. Bioinformatics 2010, 26:536-543.
34. Tervo CJ, Reed JL: MapMaker and PathTracer for tracking carbon in genome-scale metabolic networks. Biotechnol J 2016, 11:648-661.
20. Zachary AK, Adam MF: Optimizing cofactor specificity of oxidoreductase enzymes for the generation of microbial production strains-OptSwap. Ind Biotechnol 2013, 9:236-246.
35. Zhang X, Tervo CJ, Reed JL: Metabolic assessment of E. coli as a biofactory for commercial products. Metab Eng 2016, 35:6474. Total 1777 non-native chemicals were validated for their production in E. coli under glucose minimal medium conditions using MapMaker and PathTracer. They computed the E. coli’s potential abilities by using integrated modeling approaches with adding heterologous reactions. Among 1777 non-native chemicals, 279 showed potentials in the commercial applications. They also identified precursors and pathways for predicted chemicals proposing the possibility for further metabolic engineering to be developed.
21. Lakshmanan M, Chung BKS, Liu CC, Kim SW, Lee DY: Cofactor modification analysis: a computational framework to identify cofactor specificity engineering targets for strain improvement. J Bioinf Comput Biol 2013 http://dx.doi.org/ 10.1142/S0219720013430063. 22. Song CW, Kim DI, Choi S, Jang JW, Lee SY: Metabolic engineering of Escherichia coli for the production of fumaric acid. Biotechnol Bioeng 2013, 110:2025-2034. 23. Choi HS, Lee SY, Kim TY, Woo HM: In silico identification of gene amplification targets for improvement of lycopene. Appl Environ Microbiol 2010, 76:3097-3105. 24. Park JM, Park HM, Kim WJ, Kim HU, Kim TY, Lee SY: Flux variability scanning based on enforced objective flux for identifying gene amplification targets. BMC Syst Biol 2012, 6:106. 25. Kim MK, Lane A, Kelley JJ, Lun DS: E-Flux2 and SPOT: validated methods for inferring intracellular metabolic flux distributions Current Opinion in Biotechnology 2017, 47:67–82
36. Kim HU, Ryu JY, Lee JO, Lee SY: A systems approach to traditional oriental medicine. Nat Biotechnol 2015, 33:264-268. 37. Davey JA, Chica RA: Multistate approaches in computational protein design. Protein Sci 2012, 21:1241-1252. 38. Hilvert D: Design of protein catalysts. Annu Rev Biochem 2013, 82:447-470. 39. Chu HS, Kim YS, Lee CM, Lee JH, Jung WS, Ahn JH, Song SH, Choi IS, Cho KM: Metabolic engineering of 3-hydroxypropionic acid biosynthesis in Escherichia coli. Biotechnol Bioeng 2015, 112:356-364. www.sciencedirect.com
Recent systems metabolic engineering tools and strategies Chae et al. 81
40. Siegel JB, Smith AL, Poust S, Wargacki AJ, Bar-Even A, Louw C, Shen BW, Eiben CB, Tran HM, Noor E et al.: Computational protein design enables a novel one-carbon assimilation pathway. Proc Natl Acad Sci U S A 2015, 112:3704-3709. To solve difficulties for introducing heterologous carbon fixation cycles into the host, formolase (FLS) was designed computationally. Synthetic metabolic pathway for one-carbon assimilation with FLS was successfully integrated into E. coli. Compared to the natural formate assimilation pathways, the newly designed pathway possessed several advantages such as high energy efficiency and less backward flux. 41. Cameron DE, Bashor CJ, Collins JJ: A brief history of synthetic biology. Nat Rev Microbiol 2014, 12:381-390. 42. Song CW, Kim JW, Cho IJ, Lee SY: Metabolic engineering of Escherichia coli for the production of 3-hydroxypropionic acid and malonic acid through b-alanine route. ACS Synth Biol 2016, 5:1256-1263. 43. Choi YJ, Lee SY: Microbial production of short-chain alkanes. Nature 2013, 502:571-574.
55. Seo SW, Yang JS, Kim I, Yang J, Min BE, Kim S, Jung GY: Predictive design of mRNA translation initiation region to control prokaryotic translation efficiency. Metab Eng 2013, 15:67-74. 56. Farasat I, Kushwaha M, Collens J, Easterbrook M, Guido M, Salis HM: Efficient search, mapping, and optimization of multi protein genetic systems in diverse bacteria. Mol Syst Biol 2014 http://dx.doi.org/10.15252/msb.20134955. RBS Library Calculator was developed which can suggest sets of RBSs to efficiently explore multi-protein expression space across a 10,000-fold range. Furthermore, search algorithm was combined with system-level kinetic modeling making sequence-expression-activity map (SEAMAP). The creation of SEAMAP will accelerate optimization of multi-protein system. 57. Bonde MT, Pederson M, Klausen MS, Jensen SI, Wulff T, Harrison S, Nielsen AT, Herrgard MJ, Sommer MO: Predictable tunning of protein expression in bacteria. Nat Methods 2016 http://dx.doi.org/10.1038/nmeth.3727.
44. Atsumi S, Taizo H, Liao JC: Non-fermentative pathways for synthesis of branched-chain higher alcohols as biofuels. Nature 2008, 451:86-89.
58. Si T, HamediRad M, Zhao H: Regulatory RNA-assisted genome engineering in microorganisms. Curr Opin Biotechnol 2015, 36:85-90.
45. Choi SY, Park SJ, Kim WJ, Yang JE, Lee H, Shin J, Lee SY: One step fermentative production of poly(lactate-co-glycolate) from carbohydrates in Escherichia coli. Nat Biotechnol 2016, 34:435-440. The first microbial production of poly(lactate-co-glycolate) was demonstrated. Synthetic pathway was designed and construction with help of evolutionary engineering approach. Metabolic fluxes were further optimized using in silico simulation, thereby making direct production possible from cabohydrates (glucose and xylose). This is a good example of how systems metabolic engineering are used to construct advanced microbial cell factory.
59. Na D, Yoo SM, Chung H, Park H, Park JH, Lee SY: Metabolic engineering of Escherichia coli using synthetic small regulatory RNAs. Nat Biotechnol 2013, 31:170-174.
46. Cheong S, Clomburg JM, Gonzalez R: Energy-efficient and carbon-efficient synthesis of functionalized small molecules in bacteria using non-decarboxylative Claisen condensation reactions. Nat Biotechnol 2016, 34:556-561. Non-decarboxylative Claisen condensation reaction followed by subsequent b-reduction reaction was used with various starting chemicals. As a result, 18 chemicals from 10 different classes were biosynthesized including v-phenylalkanoic, a,v-dicarboxylic, v-hydroxy, v-1-oxo, v1-methyl, 2-methyl, 2-methyl-2-enolic and 2,3-dihydroxy acids, bhydroxy-v-lactones, and v-1-methyl alcohols. 47. Chae TU, Kim WJ, Choi S, Park SJ, Lee SY: Metabolic engineering of Escherichia coli for the production of 1,3diaminopropane, a three carbon diamine. Sci Rep 2015 http:// dx.doi.org/10.1038/srep13040. 48. Turk SC, Kloosterman WP, Ninaber DK, Kolen KP, Knutova J, Suir E, Schurmann M, Raemaker-Franken PC, Muller M, Wildeman SM et al.: Metabolic engineering toward sustainable production of nylon-6. ACS Synth Biol 2015, 5:65-73. 49. Galanie S, Thodey K, Trenchard IJ, Interrante MF, Smolke CD: Complete biosynthesis of opioids in yeast. Science 2015, 349:1095-1100. 50. Yang L, Wang C, Zhou J, Kim SW: Combinatorial engineering of hydrid mevalonate pathways in Escherichia coli for protoiludene production. Microb Cell Fact 2016 http:// dx.doi.org/10.1186/s12934-016-0409-7. 51. Paddon CJ, Westfall PJ, Benjamin K, Fisher K, McPhee D, Leavell MD, Tai A, Main A, Eng D, Polichuk DR et al.: High-level semi-synthetic production of the potent antimalarial artemisinin. Nature 2013, 496:528-532. 52. Jones JA, Vernacchio VR, Lachance DM, Lebovich M, Fu L, Shirke AN, Schultz VL, Cress B, Linhardt RJ, Koffas MA: ePathOptimize: a combinatorial approach for transcriptional balancing of metabolic pathways. Sci Rep 2015 http:// dx.doi.org/10.1038/srep11301. 53. Salis HM, Mirsky EA, Voigt CA: Automated design of synthetic ribosome binding sites to control protein expression. Nat Biotechnol 2009, 27:946-950. 54. Na D, Lee D: RBSDesigner: software for designing synthetic ribosome binding sites that yields a desired level of protein expression. Bioinformatics 2010, 26:2633-2634. www.sciencedirect.com
60. Yoo SM, Na D, Lee SY: Design and use of synthetic regulatory small RNAs to control gene expression in Escherichia coli. Nat Protoc 2013, 8:1694-1707. 61. Kim B, Park H, Na D, Lee SY: Metabolic engineering of Escherichia coli for the production of phenol from glucose. Biotechnol J 2014, 9:621-629. 62. Cho C, Lee SY: Efficient gene knockdown in Clostridium acetobutylicum by synthetic small regulatory RNAs. Biotechnol Bioeng 2016, 114:374-383. Synthetic small regulatory RNAs (sRNA) tool have been widely used for metabolic engineering in E. coli strains. The tool is expanded to other strain, Clostridium acetobutylicum, for the first time. As C. acetobutylicum is a hard strain to engineer, the newly developed sRNA tool will become a key tool for efficient downregulation of target genes. 63. Crook NC, Schmitz AC, Alper HS: Optimization of a yeast RNA interference system for controlling gene expression and enabling rapid metabolic engineering. ACS Synth Biol 2013, 3:307-313. 64. Desau RP, Papoutsakis ET: Antisense RNA strategies for metabolic engineering of Clostridium acetobutylicum. Appl Environ Microbiol 1999, 65:936-945. 65. Wu J, Yu O, Du G, Zhou J, Chen J: Fine-tuning of the fatty acid pathway by synthetic antisense RNA for enhanced (2S)naringenin production from L-tyrosine in Escherichia coli. Appl Environ Microbiol 2014, 23:7283-7292. 66. Yang Y, Lin Y, Li L, Linhardt RJ, Yan Y: Regulating malonyl-CoA metabolism via synthetic antisense RNAs for enhanced biosynthesis of natural products. Metab Eng 2015, 29:217-226. 67. Chappell J, Takahashi MK, Lucks JB: Creating small transcription activating RNAs. Nat Chem Biol 2015, 11:214-220. The capability of small RNAs was expanded for upregulation of target genes. The tool is named as small transcriptional activiating RNAs (STARs). The systems was constructed from natural pT181 attenuator and further engineered. Finally, four orthogonal STARs could be generated which showed up to 94-fold activation. The tools can be utilized as new component in transcriptional logic gates. 68. Choi KR, Lee SY: CRISPR technologies for bacterial systems: current achievements and future directions. Biotechnol Adv 2016, 34:1180-1209. 69. Jiang Y, Chen B, Duan C, Sun B, Yang J, Yang S: Multigene editing in the Escherichia coli genome via CRISPR-Cas9 system. Appl Environ Microbiol 2015, 81:2506-2514. 70. Zhang K, Duan X, Wu J: Multiple disruption in undomesticated Bacillus subtilis ATCC 6051a using CRISPR/Cas9 system. Sci Rep 2016 http://dx.doi.org/10.1038/srep27943. Current Opinion in Biotechnology 2017, 47:67–82
82 Tissue, cell & pathway engineering
71. Cobb RE, Wang Y, Zhao H: High-efficiency multiplex genome editing of Streptomyces species using an engineered CRISPR/ Cas system. ACS Synth Biol 2014, 4:723-728. 72. Bao Z, Xiao H, Liang J, Zhang L, Xiong X, Sun N, Si T, Zhao H: Homology-integrated CRISPR-Cas (HI-CRISPR) system for one-step multigene disruption in Saccharomyces cerevisiae. ACS Synth Biol 2014, 4:585-594.
and inefficient utilization of mixed sugar. Furthermore, authors have successfully demonstrated coculture strategy using same species (E. coli) for the first time achieving enhanced cis,cis-muconic acid and 4hydroxybenzoic acid production. 86. Park SH, Kim HU, Kim TY, Park JS, Kim SS, Lee SY: Metabolic engineering of Corynebacterium glutamicum for L-arginine production. Nat Commun 2014, 5:4618.
73. Jakociunas T, Bonde I, Herrgard M, Harrison SJ, Kristensen M, Pedersen LE, Jensen MK, Keasling JD: Multiplex metabolic pathway engineering using CRISPR/Cas9 in Saccharomyces cerevisiae. Metab Eng 2015, 28:213-222.
87. Xie W, Lv X, Ye L, Zhou P, Yu H: Construction of lycopeneoverproducing Saccharomyces cerevisiae by combining directed evolution and metabolic engineering. Metab Eng 2015, 30:69-78.
74. Li Y, Lin Z, Huang C, Zhang Y, Wang Z, Tang Y, Chen T, Zhao X: Metabolic engineering of Escherichia coli using CRISPR-Cas9 meditated genome editing. Metab Eng 2015, 31:13-21.
88. Gerando HM, Fayolle-Guichard F, Rudant L, Millah SK, Monot F, Ferreira NL, Lopez-Contreras AM: Improving isopropanol tolerance and production of Clostridium beijerinckii DSM 6423 by random mutagenesis and genome shuffling. Appl Microbiol Biotechnol 2016, 100:5427-5436.
75. Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, Lim WA: Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 2013, 152:1173-1183. 76. Cleto S, Jensen JV, Wendisch VF, Lu TK: Corynebacterium glutamicum metabolic engineering with CRISPR interference (CRISPRi). ACS Synth Biol 2016, 5:375-385. 77. Lv L, Ren YL, Chen JC, Wu Q, Chen GQ: Application of CRISPRi for prokaryotic metabolic engineering involving multiple genes, a case study: controllable P (3HB-co-4HB) biosynthesis. Metab Eng 2015, 29:160-168. 78. Dueber JE, Wu GC, Malmirchegini GR, Moon TS, Petzold CJ, Ullal AV, Prather KL, Kealing JD: Synthetic protein scaffolds provide modular control over metabolic flux. Nat Biotechnol 2009, 27:753-759.
89. Reyes LH, Gomez JM, Kao KC: Improving carotenoids production in yeast via adaptive laboratory evolution. Metab Eng 2014, 21:26-33. 90. Williams TC, Pretorius IS, Paulsen IT: Synthetic evolution of metabolic productivity using biosensors. Trends Biotechnol 2016, 34:371-381. 91. Mahr R, Gatgens C, Gatgens J, Polen T, Kalinowski J, Frunzke J: Biosensor-driven adaptive laboratory evolution of L-valine production in Corynebacterium glutamicum. Metab Eng 2015, 32:184-194. 92. Zhou LB, Zeng AP: Engineering a lysine-on riboswitch for metabolic control of lysine production in Corynebacterium glutamicum. ACS Synth Biol 2015, 4:1335-1340.
79. Liu Y, Zhu Y, Ma W, Shin HD, Li J, Liu L, Du G, Chen J: Spatial modulation of key pathway enzymes by DNA-guided scaffold system and respiration chain engineering for improved Nacetylglucosamine production in Bacillus subtilis. Metab Eng 2014, 24:61-69.
93. Li T, Ye J, Shen R, Zong Y, Zhao X, Lou C, Chen G-Q: Semirational approach for ultrahigh poly(3-hydroxybutyrate) accumulation in Escherichia coli by combining one-step library construction and high-throughput screening. ACS Synth Biol 2016, 5:13081317.
80. Avalos JL, Fink GR, Stephanopoulos G: Compartmentalization of metabolic pathways in yeast mitochondria improves the production of branched-chain alcohols. Nat Biotechnol 2013, 31:335-341.
94. Wang HH, Isaacs FJ, Carr PA, Sun ZZ, Xu G, Forest CR, Church GM: Programming cells by multiplex genome engineering and accelerated evolution. Nature 2009, 460:894898.
81. Lawrence AD, Frank S, Newnhan S, Lee MJ, Brown IR, Xue WF, Rowe ML, Mulvhill DP, Prentice MB, Howard MJ et al.: Solution structure of a bacterial microcompartment targeting peptide and its application in the construction of an ethanol bioreactor. ACS Synth Biol 2014, 3:454-465.
95. Raman S, Rogers JK, Taylor ND, Church GM: Evolution-guided optimization of biosynthetic pathways. Proc Natl Acad Sci U S A 2014, 111:17803-17808. Identifying optimal cells among billions is the key for engineering to succeed. MAGE was used to increase genetic diversity. To reduce unproductive variants, metabolite-responsive transcription factors were used to control TolC. Strains with enhanced production of naringenin and glucaric acid were developed.
82. Zhang H, Wang X: Modular co-culture engineering, a new approach for metabolic engineering. Metab Eng 2016, 37:114121. 83. Minty JJ, Singer ME, Scholz SA, Bae CH, Ahn JH, Foster CE, Liao JC, Lin XN: Design and characterization of synthetic fungal-bacterial consortia for direct production of isobutanol from cellulosic biomass. Proc Natl Acad Sci U S A 2013, 110:14592-14597. 84. Zhou K, Qiao K, Edgar S, Stephanopoulos G: Distributing a metabolic pathway among a microbial consortium enhances production of natural products. Nat Biotechnol 2015, 33:337383. 85. Zhang H, Pereira B, Li Z, Stephanopoulos G: Engineering Escherichia coli coculture system for the production of biochemical products. Proc Natl Acad Sci U S A 2015, 112:82668271. Coculture system was suggested as general strategy to overcome universal challenges in metabolic engineering such as intermediate secretion
Current Opinion in Biotechnology 2017, 47:67–82
96. DiCarlo JE, Conley AJ, Penttila M, Jantti J, Wang HH, Church GM: Yeast oligo-mediated genome engineering (YOGE). ACS Synth Biol 2013, 2:741-749. 97. Si T, Luo Y, Bao Z, Zhao H: RNAi-assisted genome evolution in Saccharomyces cerevisiae for complex phenotype engineering. ACS Synth Biol 2015, 4:283-291. 98. Yang TH, Jung YK, Kang HO, Kim TY, Park SJ, Lee SY: Tailormade type II Pseudomonas PHA synthases and their use for the biosynthesis of polylactic acid and its copolymer in recombinant Escherichia coli. Appl Microbiol Biotechnol 2011, 90:603-614. 99. Ovchinnikov S, Park H, Varghese N, Huang P-S, Pavlopoulos GA, Kim DE, Kamisetty H, Kyrpides NC, Baker D: Protein structure determination using metagenome sequence data. Science 2017, 355:294-298.
www.sciencedirect.com