Top Banner
Contents lists available at ScienceDirect Metabolic Engineering journal homepage: www.elsevier.com/locate/meteng Metabolic engineering of Escherichia coli for producing adipic acid through the reverse adipate-degradation pathway Mei Zhao a,b , Dixuan Huang a,b , Xiaojuan Zhang a , Mattheos A.G. Koas c,d , Jingwen Zhou a,b, , Yu Deng a,b, a National Engineering Laboratory for Cereal Fermentation Technology (NELCF), Jiangnan University, 1800 Lihu Road, Wuxi, Jiangsu 214122, China b School of Biotechnology, Jiangnan University, 1800 Lihu Rd, Wuxi, Jiangsu 214122, China c Department of Chemical and Biological Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180, USA d Department of Biological Sciences, Rensselaer Polytechnic Institute, Troy NY 12180, USA ARTICLE INFO Keywords: Adipic acid E. coli High titer The reverse adipate degradation pathway Metabolic engineering ABSTRACT Adipic acid is an important dicarboxylic acid mainly used for the production of nylon 66 bers and resins. Previous studies focused on the biological production of adipic acid directly from dierent substrates, resulting in low yields and titers. In this study, a ve-step reverse adipate-degradation pathway (RADP) identied in Thermobida fusca has been reconstructed in Escherichia coli BL21 (DE3). The resulting strain (Mad136) produced 0.3 g L 1 adipic acid with a 11.1% theoretical yield in shaken asks, and we conrmed that the step catalyzed by 5-Carboxy-2-pentenoyl-CoA reductase (Tfu_1647) as the rate-limiting step of the RADP. Overexpression of Tfu_1647 by pTrc99A carried by strain Mad146 produced with a 49.5% theoretical yield in shaken asks. We further eliminated pathways for major metabolites competing for carbon ux by CRISPR/Cas9 and deleted the succinate-CoA ligase gene to promote accumulation of succinyl-CoA, which is the precursor for adipic acid synthesis. The nal engineered strain Mad123146, which could achieve 93.1% of the theoretical yield in the shaken ask, was able to produce 68.0 g L 1 adipic acid by fed-batch fermentation. To the best of our knowl- edge, these results constitute the highest adipic acid titer reported in E. coli. 1. Introduction Adipic acid is a dicarboxylic acid that has extensive applications in the chemical industry, medicine, and lubricant manufacturing, and is mainly used for the production of nylon 66 bers and resins (Deng et al., 2016; Polen et al., 2013; Vardon et al., 2015; Yu et al., 2014). Additionally, adipic acid is one of 12 bio-based chemicals with the greatest market value from renewable substrates reported by the United States Department of Energy (Werpy and Petersen, 2004). The adipic acid market yields ~4676,850,000 US dollars annually, with 2850,000 US tons produced globally, which has increased ~4.1% annually (Polen et al., 2013). Adipic acid is currently produced from feedstocks derived from petroleum, specically by oxidation of a mixture of cyclohexanone and cycohexanol (KA oils) catalyzed by nitric acid (Niu et al., 2002; Polen et al., 2013; Sato et al., 1998). However, this chemical synthesis results in high levels of pollution and greenhouse gas emissions (U.S. Environmental Protection Agency, 2011). Therefore, there are great incentives to discover alternative approaches allowing for renewable and aordable adipic acid production. Recently, with the rapid devel- opment of biotechnology, the synthesis of adipic acid via metabolic engineering and synthetic biology approaches has attracted increased attention (Cheong et al., 2016; Deng et al., 2016; Polen et al., 2013; Vardon et al., 2015; Yu et al., 2014). A two-stage method consisting of biological synthesis and chemical reactions was established involving glucose fermentation to cis,cis-muconic acid (ccMA)(Jung et al., 2015; Weber et al., 2012; Wu et al., 2004; Wu et al., 2006) and ccMA hy- drogenation to adipic acid catalyzed by 10% Pt (Draths and Frost, 1994; Niu et al., 2002). However, the chemical conversion of ccMA to adipic acid is expensive, environmentally damaging, and results in low yield (0.17 g adipic acid per g glucose)(Vardon et al., 2015). Therefore, there is a need to discover alternative methods for the direct synthesis of adipic acid from renewable substrates, such as glucose and glycerol. Yu et al. constructed an articial adipic acid synthesis pathway (Yu et al., 2014) involving β-ketoadipyl-CoA thiolase (PaaJ) from Escher- ichia coli, 3-hydroxyacyl-CoA reductase (PaaH1) from Ralstonia eu- tropha, enoyl-CoA hydratase (Ech) from R. eutropha H16, trans-enoyl- CoA reductase (Ter) from Euglena gracilis, butyryl kinase (Buk1) from https://doi.org/10.1016/j.ymben.2018.04.002 Received 19 December 2017; Received in revised form 19 March 2018; Accepted 1 April 2018 Corresponding authors at: National Engineering Laboratory for Cereal Fermentation Technology (NELCF), Jiangnan University, 1800 Lihu Rd, Wuxi, Jiangsu 214122, China. E-mail addresses: [email protected] (J. Zhou), [email protected] (Y. Deng). Metabolic Engineering 47 (2018) 254–262 Available online 03 April 2018 1096-7176/ © 2018 International Metabolic Engineering Society. Published by Elsevier Inc. All rights reserved. T
9

Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

May 12, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

Contents lists available at ScienceDirect

Metabolic Engineering

journal homepage: www.elsevier.com/locate/meteng

Metabolic engineering of Escherichia coli for producing adipic acid throughthe reverse adipate-degradation pathway

Mei Zhaoa,b, Dixuan Huanga,b, Xiaojuan Zhanga, Mattheos A.G. Koffasc,d, Jingwen Zhoua,b,⁎,Yu Denga,b,⁎

aNational Engineering Laboratory for Cereal Fermentation Technology (NELCF), Jiangnan University, 1800 Lihu Road, Wuxi, Jiangsu 214122, Chinab School of Biotechnology, Jiangnan University, 1800 Lihu Rd, Wuxi, Jiangsu 214122, Chinac Department of Chemical and Biological Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180, USAd Department of Biological Sciences, Rensselaer Polytechnic Institute, Troy NY 12180, USA

A R T I C L E I N F O

Keywords:Adipic acidE. coliHigh titerThe reverse adipate degradation pathwayMetabolic engineering

A B S T R A C T

Adipic acid is an important dicarboxylic acid mainly used for the production of nylon 6–6 fibers and resins.Previous studies focused on the biological production of adipic acid directly from different substrates, resultingin low yields and titers. In this study, a five-step reverse adipate-degradation pathway (RADP) identified inThermobifida fusca has been reconstructed in Escherichia coli BL21 (DE3). The resulting strain (Mad136) produced0.3 g L−1 adipic acid with a 11.1% theoretical yield in shaken flasks, and we confirmed that the step catalyzed by5-Carboxy-2-pentenoyl-CoA reductase (Tfu_1647) as the rate-limiting step of the RADP. Overexpression ofTfu_1647 by pTrc99A carried by strain Mad146 produced with a 49.5% theoretical yield in shaken flasks. Wefurther eliminated pathways for major metabolites competing for carbon flux by CRISPR/Cas9 and deleted thesuccinate-CoA ligase gene to promote accumulation of succinyl-CoA, which is the precursor for adipic acidsynthesis. The final engineered strain Mad123146, which could achieve 93.1% of the theoretical yield in theshaken flask, was able to produce 68.0 g L−1 adipic acid by fed-batch fermentation. To the best of our knowl-edge, these results constitute the highest adipic acid titer reported in E. coli.

1. Introduction

Adipic acid is a dicarboxylic acid that has extensive applications inthe chemical industry, medicine, and lubricant manufacturing, and ismainly used for the production of nylon 6–6 fibers and resins (Denget al., 2016; Polen et al., 2013; Vardon et al., 2015; Yu et al., 2014).Additionally, adipic acid is one of 12 bio-based chemicals with thegreatest market value from renewable substrates reported by the UnitedStates Department of Energy (Werpy and Petersen, 2004). The adipicacid market yields ~4676,850,000 US dollars annually, with 2850,000US tons produced globally, which has increased ~4.1% annually (Polenet al., 2013).

Adipic acid is currently produced from feedstocks derived frompetroleum, specifically by oxidation of a mixture of cyclohexanone andcycohexanol (KA oils) catalyzed by nitric acid (Niu et al., 2002; Polenet al., 2013; Sato et al., 1998). However, this chemical synthesis resultsin high levels of pollution and greenhouse gas emissions (U.S.Environmental Protection Agency, 2011). Therefore, there are greatincentives to discover alternative approaches allowing for renewable

and affordable adipic acid production. Recently, with the rapid devel-opment of biotechnology, the synthesis of adipic acid via metabolicengineering and synthetic biology approaches has attracted increasedattention (Cheong et al., 2016; Deng et al., 2016; Polen et al., 2013;Vardon et al., 2015; Yu et al., 2014). A two-stage method consisting ofbiological synthesis and chemical reactions was established involvingglucose fermentation to cis,cis-muconic acid (ccMA)(Jung et al., 2015;Weber et al., 2012; Wu et al., 2004; Wu et al., 2006) and ccMA hy-drogenation to adipic acid catalyzed by 10% Pt (Draths and Frost, 1994;Niu et al., 2002). However, the chemical conversion of ccMA to adipicacid is expensive, environmentally damaging, and results in low yield(0.17 g adipic acid per g glucose)(Vardon et al., 2015). Therefore, thereis a need to discover alternative methods for the direct synthesis ofadipic acid from renewable substrates, such as glucose and glycerol.

Yu et al. constructed an artificial adipic acid synthesis pathway (Yuet al., 2014) involving β-ketoadipyl-CoA thiolase (PaaJ) from Escher-ichia coli, 3-hydroxyacyl-CoA reductase (PaaH1) from Ralstonia eu-tropha, enoyl-CoA hydratase (Ech) from R. eutropha H16, trans-enoyl-CoA reductase (Ter) from Euglena gracilis, butyryl kinase (Buk1) from

https://doi.org/10.1016/j.ymben.2018.04.002Received 19 December 2017; Received in revised form 19 March 2018; Accepted 1 April 2018

⁎ Corresponding authors at: National Engineering Laboratory for Cereal Fermentation Technology (NELCF), Jiangnan University, 1800 Lihu Rd, Wuxi, Jiangsu 214122, China.E-mail addresses: [email protected] (J. Zhou), [email protected] (Y. Deng).

Metabolic Engineering 47 (2018) 254–262

Available online 03 April 20181096-7176/ © 2018 International Metabolic Engineering Society. Published by Elsevier Inc. All rights reserved.

T

Page 2: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

Clostridium acetobutylicum, phosphate butyryltransferase (Ptb) from C.acetobutylicum in E. coli, producing ~0.0007 g L−1 adipic acid fromglucose. Based on this pathway, Cheong et al. engineered E. coliMG1655 (ΔldhA, ΔpoxB, Δpta, ΔadhE, and ΔsucD) to produce 2.5 g L−1

adipic acid from 50 g L−1 glycerol (9.6% theoretical yield)(Cheonget al., 2016). However, the adipic acid titers were far from those re-quired for industrial application.

Previously, there were no native adipic acid synthesis pathwaysdetermined in microorganisms. However, Deng et al. reported the ex-istence of a native adipic acid synthesis pathway in the thermophilicactinobacterium Thermobifida fusca (Deng and Mao, 2015) and identi-fied it as a reverse adipate-degradation pathway (RADP) that includedfive enzymes (Fig. 1): Tfu_0875 (β-ketothiolase), Tfu_2399 (3-hydro-xyacyl-CoA dehydrogenase), Tfu_0067 (3-hydroxyadipyl-CoA dehy-drogenase), Tfu_1647 (5-Carboxy-2-pentenoyl-CoA reductase) andTfu_2576-7 (adipyl-CoA synthetase). They reported that T. fusca B6produced 2.23 g L−1 adipic acid from 50 g L−1 glucose; however, due tothe lack of tools available for genetically engineering T. fusca, it wasdifficult to increase adipic acid titer and yield.

Here, we constructed the T. fusca RADP in E. coli for adipic acidproduction. After we confirmed the rate-limiting step (Tfu_1647) in theRADP, the corresponding enzyme was over-expressed for increasing theadipic acid production. Additionally, pathways for the metabolitescompeting for carbon flux were eliminated by CRISPR/Cas9, and theconcentration of the precursor of adipic acid, succinyl-CoA, was in-creased by deleting the succinyl-CoA synthetase alpha subunit (sucD)gene. Furthermore, we optimized fermentation conditions for the en-gineered strain to achieve a high titer of adipic acid. Our findings re-present the first example of expressing a native microorganism RADPpathway in E. coli, resulting in both high yield and titer of the targetproduct.

2. Materials and methods

2.1. Bacterial strains and cultivation

The strains used in this study are listed in Supplementary Table 1.The E. coli strains were grown on M9, Lysogeny broth (LB), super op-timal broth (SOB), super optimal broth with catabolite repression(SOC), 3-(N-morpholino) propanesulfonic acid broth (MOPS), and ter-rific broth (TB) (the media recipes are shown in Supplementarymethods) with addition of chemicals as desired in 250mL shake flaskswith 200 rpm agitation on 4 g L−1 glucose. Gene expression was in-duced by IPTG at various final concentrations. All genes from organismsother than E. coli were codon optimized for E. coli by Genewiz (Suzhou,

China). The production of adipic acid by E. coli in a 5-L bioreactor wasconducted at 37 °C for culturing cells and at 30 °C for inducing geneexpression with 1 vvm aeration and 400 rpm agitation with 1mM IPTG.

For the fed-batch fermentation, a 5-L bioreactor (Baoxing, Shanghai,China) was used to conduct the fed-batch fermentation for the finalstrain. The two-stage fermentation strategy was employed in this study.The initial culturing temperature was 37 °C. During the first stage, 2%of the working volume seed culture was inoculated to the bioreactorand fermentation was initiated with 20 g L−1 glucose. Isopropyl β-D-1-thiogalactopyranoside (IPTG) was added into the medium once OD(600 nm) reached 0.6 after inoculation and then the culturing tem-perature was decreased to 30 °C. During the second stage, desiredamount of 500 g L−1 glycerol added to the bioreactor. The pH wasmaintained at 7.0 by online measurement using a pH sensor and ad-dition of 2mol L−1 NaOH. The aeration and agitation rates varied ac-cording to different conditions.

2.2. Plasmid construction

Plasmids and primers used in this study are listed in SupplementaryTable 1 and Supplementary Table 2. The synthesized genes are shownin Supplementary Table 3. Tfu_0875 was synthesized and ligated intoplasmid pUC57-1 by Genewiz (Suzhou, China) and amplified by pri-mers pUC57-1F (EcoR I) and pUC57-1R (Hind III). The PCR product andpRSFDuet-1 plasmid were digested by EcoR I and Hind III, and Tfu_0875was ligated into pRSFDuet-1 using the T4 DNA ligase. The ligationproduct was introduced into E. coli JM109 for screening by colony PCRand Sanger sequencing. The final plasmid was named pRSF-Tfu_0875.Tfu_2399 was amplified by primers pUC57-2F (Bgl II) and pUC57-2R(Kpn I) and ligated into pRSF-Tfu_0875, which was digested with Bgl IIand Kpn I to form pAD1. Plasmids pAD3, pAD6, pAD9, pAD10, andpAD11 were constructed according to the same protocol describedpreviously, and the primers used are listed in Supplementary Table 2.Tfu_0067 and Tfu_1647 were amplified by the primers pUC57-3R andpUC57-3F and pUC57-4R and pUC57-4F from pUC57-3 and pUC57-4,respectively. PTrc99A was linearized by Nco I and Hind III restrictiondigest, and primers pAD4–0067F and pAD4–1647R contained 39-ntsequences homologous to the upstream and downstream regions ofpTrc99A at the Nco I and Hind III sites, with the two genes linked by anRBS (AAGAGGTATATATTA). Plasmid pAD4 was constructed usingGibson assembly methods (Gibson et al., 2009) and transformed into E.coli JM109 competent cells. The remaining plasmids developed in thisstudy were constructed following the same protocol described aboveand are described in Supplementary Table 1.

Fig. 1. Metabolic strategies for redirectingcarbon flux to adipic acid in Escherichia coli.The reverse adipate-degradation pathway in-cludes five steps in Thermobifida fusca: β-ke-tothiolase (Tfu_0875), 3-hydroxyacyl-CoA de-hydrogenase (Tfu_2399), 3-hydroxyadipyl-CoAdehydrogenase (Tfu_0067), 5-Carboxy-2-pen-tenoyl-CoA reductase (Tfu_1647), and adipyl-CoA synthetase (Tfu_2576-7). The ldhA, atoB,and sucD genes were deleted using theCRISPR/Cas9 system. Red font representsgenes or enzymes subject to overexpression,and blue font represents those deleted. Solidarrows indicate single steps, and dashed ar-rows indicate multiple steps. Red arrows re-present overexpressed genes or proteins, andblue arrows represent those deleted. ldhA, L-

lactate dehydrogenase; atoB, acetyl-CoA C-acetyltransferase; sucD, succinyl-CoA synthetase alpha subunit; TCA, tricarboxylic acid cycle; EMP, the glycolysis pathway;AA, adipic acid; AKG, a-ketoglutarate; CIT, citrate; ACO, aconitate; ICI, isocitrate; OAA, oxaloacetate; AcCoA, acetyl CoA; PYR, pyruvate; SucCoA, succinyl-CoA;SUCC, succinic acid; Glc, glucose; Gly, Glycerol; G3P, glycerol 3-phosphate; DHAP, dihydroxyacetone phosphate. (For interpretation of the references to color in thisfigure legend, the reader is referred to the web version of this article.)

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

255

Page 3: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

2.3. Gene deletions in E. coli BL21 (DE3) cells by CRISPR-Cas9

The general procedure of deleting genes in E. coli BL21 (DE3) wasdescribed previously by Jiang et al.(Jiang et al., 2015). The plasmidpCas harbored Cas9 and single-guide RNA (sgRNA) targeting plasmidpTargetF (Jiang et al., 2015). The pTargetF harboring sgRNA (con-stitutive expression) targeting the gene to be deleted and donor DNAwere transformed into cells with pCas, and 10mM arabinose was addedto induce λ-Red on pCas in order to promote homologous recombina-tion. The mutant lacking homologous recombination by the donor DNAdid not survive, whereas the one with the donor DNA replacing thetarget gene was able to grow on the agar plate. IPTG (0.5mM) was usedto induce the expression of sgRNA, which cut the targeting ‘pTargetF’using Cas9. The pCas was eliminated upon heating the culture to~42 °C. The sgRNAs targeting sucD, atoB, and ldhA genes are shown inSupplementary Table 3.

2.4. GC-MS identification of adipic acid

Adipic acid in the fermentation broth was extracted with ethylacetate. Samples were vacuum dried after re-dissolving sample withacetonitrile, and extracted using silylation reagent at a ratio of N,O-bis-(trimethylsilyl) trifluoroacetamide to trimethylchlorosilane of 99:1 andanalyzed by GC-MS (Cheong et al., 2016) using a GCMS-QP2010 ultramass selective detector (Shimadzu, Kyoto, Japan) at 70 eV electronimpact to record mass spectra. The scan ranged from 50m/z to 450m/z,and the detector gain was set to 1.67 kV. The interface, ion source, andquadrupole mass analyzer were maintained at 220 °C, 200 °C, and240 °C, respectively. The GC conditions were as follows: initial tem-perature was 50 °C, with a 1-min isothermal time; the heating programwas 8 °C per min up to 180 °C without isothermal time, followed by10 °C per min up to 240 °C, with an isothermal time of 5min; and thecarrier gas was helium with a constant flow rate of 1mLmin−1. GC-MSsoftware (Shimadzu Corporation Technologies, Kyoto Japan) was usedto determine all data. The volume of sample injection was 1 μL.

2.5. LCMS-Q-TOF identification of adipic acid

The purified adipic acid samples and the same amount of adipic acidstandard were dissolved in the water. The purified samples were ana-lyzed by Waters MALDI SYNAPT Q-TOF MS equipped with WatersAcquity UPLC BEH C18 column (2.1 mm x 50mm) and Waters AcquityPDA detector (200–400 nm)(Li et al., 2016). The mobile phase was: A:formic acid (0.1%), B: MeOH (100%) with 0.3 mLmin−1 of flow velo-city at 45 °C with gradient elution program (Supplementary Table 4).The isocratic elution is MS conditions: ESI+, Capillary: 3.5 kV, Cone: 30V, Source Block Temp: 100℃, Desolvation Temp: 400℃, DesolvationGas Flow: 700 L h−1, Cone Gas Flow: 50 L h−1, Collision Energy: 6 eV,mass range: 20–2000m/z,detector voltage: 1800 V.

2.6. NMR identification of adipic acid

Waters 1525 (Milford, MA, USA), a preparative UPLC was used toget the pure adipic acid by picking the peaks of it. Waters 1525 wasequipped with Waters Xbridge C18 (250mm L×10.0 mm i.d.) workingat 40 °C with UV–VIS detector (λ=210 nm). The mobile phase was80% methanol with isocratic elution. The adipic acid peaks were pickedand obtained manually around 10min of retention time. The adipicacid solution was transferred to a rotary evaporator to get rid of MeOHand then freeze-dried for NMR analysis.

1H NMR spectra were recorded on a Bruker Avance III (400MHz)spectrometer (Billerica, MA), using MeOD as a solvent and tetra-methylsilane as internal standard. Coupling constants J [Hz] were di-rectly taken from the spectra and were not averaged.

2.7. Metabolite quantification

Cell culture samples (1 mL) were collected by centrifugation at12,000g for 5min, and the supernatants were used for detection.Metabolites, such as glucose, glycerol, adipic acid, acetic acid, andbutyric acid, were measured by high-performance liquid chromato-graphy (HPLC; Rigol, Suzhou, China) equipped with a HPX-87H ion-exclusion column (Bio-Rad, Hercules, CA, USA) The mobile phaseconsisted of 5mM H2SO4-acetonitrile (97:3) at a rate of 0.3 mLmin−1,and the column temperature was maintained at 30 °C. The organic acidswere measured by UV–VIS detector and glucose and glycerol weremeasured by refractive index (RI) detector.

2.8. Acetyl-CoA detection

Cell culture samples (200 μL; OD (600 nm) ≈ 1) were collected andmixed with 600 μL − 80 °C aqueous 60% (v v−1) methanol solution tostop all biochemical reactions. Cell pellets were collected by cen-trifugation at 8000g for 10min at 4 °C, and cells were washed twicewith cold phosphate-buffered saline (PBS; pH 7.4). For cell lysis, 500 μLof 6% perchloric acid [2mmol L−1 dithiothreitol (DTT)] was added tocells, and following ultrasound fragmentation for 15min, cellulardebris and denatured proteins were removed by centrifugation at12,000g for 10min at 4 °C. K2CO3 was immediately added to the su-pernatant, and the pH was adjusted to 3.0 (Takamura and Nomura,1988). The supernatant was then filtered for HPLC measurement. A C18chromatography column (ODS2 Hypersil, 5 µm; Thermo Fisher Scien-tific, Waltham, MA, USA) connected to a UV detector (254 nm) wasused for detection using buffer solution A [0.2 mol L−1 sodium phos-phate (pH 5)], with buffer solution B [0.25mol L−1 800mL sodiumphosphate (pH 5) and 200mL acetonitrile mixture] used as the mobilephase. The flow velocity was 0.3 mLmin−1, and the column tempera-ture was maintained at 25 °C. The gradient elution is shown inSupplementary Table 5.

2.9. Measurements of the intracellular ATP and ADP concentrations

To measure the [ATP]/[ADP] ratio, 0.2mL of cell culture was mixedwith 0.2mL of 80 °C phenol (equilibrated with 10mM Tris–1mM EDTA[pH 8]) supplemented with 0.2-g glass beads (106m in diameter;Sigma-Aldrich, MO, USA) as previously described (Koebmann et al.,2002b). Treatment and measurement of the intracellular concentrationsof ATP and ADP were then performed by using ADP/ATP Ratio AssayKit (Sigma-Aldrich, MO, USA) as recommended by the manufacturer(Koebmann et al., 2002b).

2.10. Enzyme assays

Cell pellets were collected by centrifugation at 8000g for 10min at4 °C, and cells were washed twice with cold phosphate-buffered saline(PBS; pH 7.4). For cell lysis, 500 μL of 6% perchloric acid [2mmol L−1

dithiothreitol (DTT)] was added to cells, and following ultrasoundfragmentation for 5min, cellular debris were removed by centrifuga-tion at 12,000g 10min at 4 °C. The UV-2450 spectrometer (Shimadzu)was used in this study.

For reverse 5-Carboxy-2-pentenoyl-CoA reductase (Tfu_1647) ac-tivity: The enzymatic activity derived from pAD3 and pAD4 productsand harbored in E. coli BL21 (DE3) cells was measured by an increase inabsorbance at 340 nm caused by NAD+ reduction or at 600 nm by DCIPreduction.

For NAD-dependent activity: The solution consisted of 0.2 mMNAD+, 5mM MgCl2, 100 μL cell extract, 100mM Tris–HCl, and 5mMDTT. The initial reaction needed to add 500mM adipic acid. The de-finition of 1 U mg−1 was to obtain 1 μmol NADHmin−1 mg−1-protein(Deng et al., 2013).

For FAD-dependent activity: The solution consisted of 100 μM

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

256

Page 4: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

dichlorophenolindophenol (DCIP), 5 mM MgCl2, 100 μL cell extract andPBS (pH 7.0) was added to reach 0.9mL. The initial reaction was ac-tivated by adding 0.1mL 500mM adipic acid. The definition of1 Umg−1 was to obtain 1 μmol DCIPmin−1 mg−1-protein (Deng et al.,2013). The reductions of DCIP was measured at 600 nm with blue tocolorless.

For succinyl-CoA synthetase activity: Quantitative detection ofsuccinyl-CoA synthetase activity was obtained by using a commercialavailable kit according to manufacturer instructions (GenmedScientifics, Wilmington, DE, USA). Cells were cleaned with a cleaningsolution (reagent A), and after centrifugation at 4 °C for 5min at 300g,precipitated and dissolved in lysis buffer (reagent B). After vortexing for15 s, cells were placed on ice for 30min. After centrifugation under thesame conditions as described previously, the supernatant was used todetect the protein and enzyme activity. A Bradford protein assay kit(Beijing Solarbio, Beijing, China) was used to determine protein con-centration. Reagent C (130 μL), 20 μL enzymatic reaction buffer (re-agent D), 20 μL reaction buffer (reagent E), and 20 μL substrate solution(reagent F) were added to the 96-well plate to determine enzyme ac-tivity. After incubating at 30 °C for 3min, 10 μL of treated supernatantor 10 μL negative-control solution (reagent G) was added to the above96-well plates, respectively. Luminescence was detected on a micro-plate reader (BioTek Synergy 2, BioTek) at 340 nm every 1min for atotal of 5min. Enzyme activity was calculated using the followingformula (Deng et al., 2013):

=△ − △U μmol min A B( / ) ( )*D*0.2*C

0.01*0.622*0.5*5

A: sample readingB: blank readingC: protein concentration (mg mL−1)D: dilution

2.11. SDS-PAGE analysis

SDS-PAGE was performed using the SDS-PAGE gel preparation kit(Solarbio) according to manufacturer instructions. A culture sample(1mL) was taken 24 h after induction for SDS-PAGE analysis and pel-leted by centrifugation at 10,000g. Cell pellets were resuspended in50mM Tris–HCl (pH 8.0) and vortexed for 5min on ice. SDS-containingsample buffers (5 μL) were added to 15 μL of the samples and heated for10min, followed by 20 μL of each sample being loaded onto the gel.Gels were electrophoresed at 60 V until the samples fully entered intothe concentrate (Schagger, 2006). After 30min, the voltage waschanged to and maintained at 80 V. The gel was fixed for 30min andstained by Coomassie brilliant blue (Biotechnology company, Shanghai,China) for 3 h and destained twice in 10% acetic acid. Products werevisualized using an automatic image-analysis system (Beijing 61 gel-imaging analysis system, WD-9413B, Beijing, China).

2.12. Data availability

All experimental data were measured in triplicate, and error barsrepresent the standard deviation. The data that support the findings ofthis study are available from the authors on reasonable request; seeauthor contributions for specific data sets.

3. Results

3.1. Constructing the T. fusca RADP in E. coli

The stoichiometry associated with producing adipic acid from glu-cose and glycerol based on the RADP is shown in SupplementaryTable 6. This indicates that in the presence of 1mol glucose, 2/3moladipic acid would be produced by generating 2mol adenosine tripho-sphate (ATP) and 4mol NADH, resulting in a theoretical yield of 0.54 gadipic acid per g glucose. In the presence of 1mol glycerol, 1/3moladipic acid would be produced by generating 1mol ATP and 3molNADH, with a theoretical yield of 0.52 g-adipic acid per g-glycerol.However, the above theoretical yield calculation was based on the TCAcycle, a branched TCA cycle generating less excess NADH might in-crease the yield further (Tsuji et al., 2013).

Genes from T. fusca were codon optimized for E. coli and ligated intoDuet-1 plasmids, and their expression was induced by isopropyl β-D-1-thiogalactopyranoside (IPTG). E. coli BL21 (DE3) cells harboring pAD1(Tfu-0875 and Tfu-2399), pAD3 (Tfu-0067 and Tfu-1647), and pAD6(Tfu-2576 and Tfu-2577) were designated as E. coli Mad136, whichproduced 0.16 g L−1 adipic acid in M9 medium, with a yield of 0.04 gper g glucose (a 7.4% theoretical yield) (Fig. 2a). The product wasverified by gas chromatography-mass spectrometry (GC-MS)(Supplementary Figure 1) and Nuclear Magnetic Resonance (NMR)analysis. For NMR analysis, the adipic acid was purified by preparativeHPLC, which was then subject to freeze drying after getting rid ofmethanol. The NMR result showed that the pure product we extractedfrom the fermentation broth was adipic acid (Supplementary Figure 2).Besides, LCMS-Q-TOF analysis of the purified adipic acid showed thatthe [M+H]+ of was 81.0, 83.0, 101.0, 127.0, 145.0, 146.0, 244.9,269.9, 344.0, respectively, which were as the same as that of the adipicacid standard.

To determine the optimal medium for E. coli Mad136 production of

Fig. 2. Characterizations of Escherichia coli Mad136. (a) Production of adipicacid on M9, Lysogeny broth (LB), super optimal broth (SOB), super optimalbroth with catabolite repression (SOC), MOPS minimal medium (MOPS), andterrific broth (TB) media in shaken flasks. (b) [ATP]/[ADP] ratio from Mad146and BL21 (DE3) strains. E. coli Mad146: BL21 (DE3) harboring pAD1, pAD4,and pAD6. Error bars represent the s.d. from three independent assays.

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

257

Page 5: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

adipic acid, M9, Lysogeny broth (LB), super optimal broth (SOB), superoptimal broth with catabolite repression (SOC), 3-(N-morpholino)propanesulfonic acid broth(MOPS), and terrific broth (TB) media werestudied. Among these, SOB produced the highest titer from the Mad136strain at 0.3 g L−1 adipic acid, with a 0.06 g per g glucose yield (an11.1% theoretical yield) (Fig. 2a).

The Mad136 strain was then cultured in a 5-L bioreactor with 0, 1, 2and 3 vvm aeration and 400-rpm agitation in SOB medium(Supplementary Figure 3). There was very low adipic acid accumula-tion in the absence of aeration. However, once the aeration was in-creased to 1 vvm, the yield was increased to 18.5% theoretical yield.However, we increased the aeration to 2 vvm and 3 vvm, the yield ofadipic acid decreased slightly, indicating that high levels of aerationwere unfavorable for adipic acid synthesis.

3.2. Optimization of the adipic acid pathway

To determine the optimal combination of heterologous enzymescatalyzing the reactions necessary for adipic acid synthesis in E. coli,enzyme activities were studied in E. coli BL21 (DE3) cells grown in M9medium to eliminate any trace carbon other than glucose. The pro-duction of adipic acid varied significantly along with different combi-nations of enzymes (Fig. 3). BL21 (DE3) cells harboring pRSF-Tfu_0875-Tfu_2399, pET-Tfu_0067-Tfu_1647, pCD-Tfu_2576-Tfu_2577 and pRSF-PaaJ-PaaH1, pET-Tfu_0067-Tfu_1647, pCD-Tfu_2576-Tfu_2577 producedthe highest adipic acid yield (0.2 g L−1), whereas those harboring pRSF-Tfu_0875-Tfu_2399, pET-Ech-Ter, pCD-Buk1-Ptb, pRSF-PaaJ-PaaH1,pET-Ech-Ter, and pCD-Buk1-Ptb accumulated the lowest adipic acidyield (0.06 g L−1) from 4 g L−1 glucose. These results demonstratedthat strains expressing Ech and Ter genes produced significantly lessadipic acid than those harboring Tfu_0067 and Tfu_1647. Our resultsindicated that the presence of the Ter gene was homologous to Tfu_1647in regard to adipic acid production. Deng et al. (Deng and Mao, 2015)reported that Tfu_1647 was the rate-limiting step for adipic acidsynthesis in T. fusca. Therefore, we hypothesized that low levels of Tfu-1647 might hinder adipic acid accumulation. To test this hypothesis, anartificial operon including sequences for Tfu_0067 and Tfu_1647 wassynthesized and ligated into pTrc99A, a medium copy number plasmidcontaining a strong promoter (pTrc) to form pAD4. The new strain(Mad146) containing pAD1, pAD4, and pAD6 was grown in SOBmedium in the presence of 4 g L−1 glucose. Sodium dodecyl sulfatepolyacrylamide gel electrophoresis (SDS-PAGE) results showed thatTfu_1647 expression under the control of pTrc promoter exceeded thatobserved in other strains (Fig. 4a). We analyzed the ribosome-bindingsite (RBS) strength associated with Tfu_1647 translation using an RBScalculator (Borujeni et al., 2014) and found the existence of an addi-tional start-position in the RBS on pTrc-Tfu_1647 as compared with

pET-Tfu_1647 (Supplementary Table 7), indicating that the translationof Tfu_1647 on pTrc-Tfu_1647might be much stronger than that of pET-Tfu_1647. Based on the enzymatic assay, Tfu_1647 was confirmed to be

Fig. 3. Combinations of heterologous enzymescatalyzing the reactions necessary for adipicacid synthesis. PaaJ: β-ketoadipyl-CoA thio-lase, PaaH1: 3-hydroxyacyl-CoA reductase;Ech: enoyl-CoA hydratase; Ter: trans-enoyl-CoAreductase; Buk1: butyryl kinase; Ptb: phosphatebutyryltransferase. Tfu_0875: β-ketothiolase,Tfu_2399: 3-hydroxyacyl-CoA dehydrogenase,Tfu_0067: 3-hydroxyadipyl-CoA dehy-drogenase, Tfu_1647: 5-Carboxy-2-pentenoyl-CoA reductase, and Tfu_2576-7: adipyl-CoAsynthetase. Error bars represent the s.d. fromthree independent assays.

Fig. 4. Over-expression of Tfu_1647 by pTrc99A. (a) SDS-PAGE results. Lane 1,marker; lanes 2–6: protein lysates isolated from strains BL21 (wild-type; lane 2)and BL21 (pET-Tfu_0067-Tfu_1647; lane 3), BL21 (pTrc-Tfu_0067-Tfu_1647; lane4), BL21 (pRSF-Tfu_0875-Tfu_2399, pET-Tfu_0067-Tfu_1647, and pCDF-Tfu_2576-Tfu_2577; lane 5), and BL21 (pRSF-Tfu_0875-Tfu_2399, pTrc-Tfu_0067-Tfu_1647, and pCDF-Tfu_2576-Tfu_2577; lane 6). Strains were grownin SOB medium containing 0.4% glucose and induced with 1mM IPTG. Sampleswere separated by SDS-PAGE. Tfu_1647 bands at 42.2 kDa are identified byblack arrow and the suspected Tfu_0067 bands at 28.2 kDa are identified byblue arrow. Full gel for the portion of gel presented here was included inSupplementary Figure 10. (b) Specific reverse Tfu_1647 activity. Error barsrepresent the s.d. from three independent assays. (For interpretation of thereferences to color in this figure legend, the reader is referred to the web ver-sion of this article.)

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

258

Page 6: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

FADH2-dependent. Bioinformatics analysis indicates this enzyme be-longs to the Bcd-type family of enoyl-CoA reductases, which are re-versible (Deng and Mao, 2015). We measured the reverse activity ofTfu_1647: converting from adipyl-CoA to 2, 3-dehydroadipyl-CoA. Thereverse activities of Tfu_1647 following Tfu_1647 expression controlledby the pTrc promoter and T7 promoter were shown in Fig. 4b. Thespecific reverse Tfu_1647 activity observed after 24, 28 and 44 h offermentation was 0.72 Umg−1, 0.81 Umg−1, and 1.48 Umg−1, re-spectively, for the T7 promoter and 3.1 Umg−1, 1.83 Umg−1, and2.21 Umg−1, respectively, for the pTrc promoter. Thus, the expressionof Tfu_1647 controlled by pTrc promoter was much stronger than theone controlled by T7 promoter and it was also confirmed by the RBSresults above.

The final step of the RADP was catalyzed by Tfu_2576-7 subse-quently generating ATP as the product. Therefore, we measured ATPlevels in E. coli strains with and without the ability to produce adipicacid (Fig. 2b). Our results showed that during the log phase, ATP/ADPin strain Mad146 were 7.53, which was 4.8-fold greater than that ob-served in BL21 (DE3) cells (1.57). During the stationary phase, the ATPlevel measured in the Mad146 strain was 4.50, which was 4.4-foldgreater than that observed in BL21 (DE3) cells (1.02). These resultsindicated that overexpression of the heterologous enzymes allowedMad146 cells to accumulate significantly more ATP as compared withBL21 (DE3) cells.

The new strain (Mad146) containing pAD1, pAD4, and pAD6 wasgrown in SOB medium in the presence of 4 g L−1 glucose with 49.5%the theoretical yield in shake flasks. (Fig. 5a), which was 4.5-fold higherthan that observed from strain Mad136. These results confirmed thatthe production of adipic acid was correlated with Tfu_1647 expressionlevel.

3.3. Engineering the host strain

We deleted L-lactate dehydrogenase (ldhA), succinyl-CoA synthetasealpha subunit (sucD), and acetyl-CoA acetyltransferase (atoB) in-dividually to test the effects of their knockout on adipic acid synthesis.We employed CRISPR/Cas9 (Sander and Joung, 2014) to remove ldhAin E. coli and observed no lactic acid accumulation in the ldhA-deletionstrain, and thus adipic acid yield in the new strain (Mad1146) was in-creased from 49.5% to 61.7% theoretical yield in shaken flasks(Fig. 5a).

Butyric acid is synthesized from acetoacetyl-CoA and catalyzed byenzymes encoded by Hbd, Crt, Ter, and tesB (Saini et al., 2014; Yu et al.,2015), which are homologous to the Tfu_2399, Tfu_0067, Tfu_1647, andTfu_2576-7, respectively, used in this study. The fermentation resultsalso showed that butyric acid was another byproduct. To eliminatebutyric acid, we deleted atoB encoding acetyl-CoA acetyltransferaseinvolved in the conversion of acetyl-CoA to acetoacetyl-CoA in E. coli.The fermentation results using strain Mad2146 revealed no butyric acidproduced and an increase in adipic acid from a 49.5–68.5% theoreticalyield in shaken flasks (Fig. 5a).

To switch the carbon flux of succinyl-CoA from succinic acid toadipic acid, sucD encoding succinate-CoA ligase subunit alpha was de-leted in E. coli to test whether partial inhibition of succinate productionfrom succinyl-CoA would direct carbon flux from succinyl-CoA to adipicacid through the heterologous pathway. The specific succinyl-CoAsynthetase activity of the ΔsucD strain Mad3146 was0.0068 μmolmin−1, which was ~50% that of the parent strain Mad146(0.0103 μmolmin−1), indicating that the degradation of succinyl-CoAto succinate was largely blocked. The adipic acid yield of strainMad3146 (a 67.6% theoretical yield) increased significantly as a resultas compared with that observed in strain Mad146 (a 49.5% theoreticalyield) in shaken flasks.

Based on these results, we deleted the sucD, atoB, and ldhA genes inBL21 (DE3) cells harboring plasmids for adipic acid production, re-sulting in strain Mad123146. Our results revealed no byproducts other

than adipic acid produced by this strain (Fig. 5a), whereas adipic acidproduction reached a 93.1% theoretical yield in shaken flasks, whichwas 6.7-fold higher than that observed from strain Mad136.

In order to test the stability of the recombinant E. coli strain forindustrial application purpose, the strain Mad123146 was grown on4 g L−1 glucose in shake flasks and the production of adipic acid wasinduced by IPTG. Once the cells reached stationary phase, cells wereharvested and transferred to the fresh medium with IPTG in it to startproducing adipic acid. In this study, we transferred cells for 10 times. InSupplementary Figure 4, decreased production of adipic acid fromglucose was not observed throughout these transfers, confirming thestability of the recombinant E. coli strain.

Given that acetyl-CoA represents another adipic acid precursor, wefurther compared the intracellular concentration of acetyl-CoA in theparent and engineered strains. Acetyl-CoA levels in different strains areshown in Fig. 5b. Strains Mad1, Mad2, Mad3, and Mad23 exhibited 1.2-fold higher acetyl-CoA concentrations relative to the BL21 (DE3) strain,whereas strain Mad123 (ΔldhA, ΔatoB, and ΔsucD) exhibited 1.6-foldhigher acetyl-CoA concentrations as compared with the BL21 (DE3)strain. These results indicated that higher acetyl-CoA levels were ben-eficial for adipic acid production in the engineered E. coli strain.

Fig. 5. Strain engineering results. (a) Fermentation results for different en-gineered strains in shaken flasks on 4 g L−1 glucose. (b) Intracellular acetyl-CoAlevels of different strains. Error bars represent the s.d. from three independentassays.

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

259

Page 7: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

3.4. Production of adipic acid in 5-L bioreactors

To achieve high titers of adipic acid, fed-batch fermentation wasemployed using glycerol as the major substrate. Because the adipic acidproduction was sensitive to oxygen, we conducted fed-batch fermen-tation by not controlling oxygen (1 vvm aeration, 400 rpm agitation),and by controlling dissolved oxygen (DO) at 5% and 10%, respectively.In Supplementary Figure 5, although higher levels of DO helped accu-mulate additional biomass as compared with that accumulated at lowlevels of DO, the maximal titer of adipic acid production was observedat low oxygen levels (68.0 g L−1, 42.3 g L−1 and 36.1 g L−1 for notcontrol, DO 5% and DO 10%, respectively). The results of adipic acidproduction in the absence of DO control are shown in Fig. 6. We ob-served a maximal adipic acid titer of 68.0 g L−1 after an 88 h fermen-tation, with a maximal OD (600 nm) of 36.7 achieved after 64 h, re-presenting a 226 fold increase as compared with strain Mad136. TheDO under this condition was< 5% for most of the fermentation pro-cess.

The fermentation broth was then subject to the purification processto extract adipic acid (the general process is shown in SupplementaryMethods). After purification, adipic acid was subjected to the LCMS-Q-TOF analysis with gradient elution (Supplementary Table 4). The purityof adipic acid was 96.86% with a yield of 87.3%. The above data are

calculated according to the peak area according to the SupplementaryFigure 6.

4. Discussion

Adipic acid is a dicarboxylic acid that plays an important role inchemical industry, medicine, and lubricant manufacturing, with themost popular product derived from adipic acid being nylon-6, 6 (Denget al., 2016; Polen et al., 2013; Vardon et al., 2015; Yu et al., 2014).Previously, no native adipic acid synthesis pathway was found in mi-croorganisms until the discovery of a RADP by Deng et al. in the acti-nobacterium T. fusca capable of converting acetyl-CoA and succinyl-CoA to adipic acid (Deng and Mao, 2015). Here, we constructed the T.fusca RADP in E. coli BL21 (DE3) by expression of five enzymes(Tfu_0875, Tfu_2399, Tfu_0067, Tfu_1647, and Tfu_2576-7) for adipicacid production via Duet-1 plasmids. Our results indicated that theengineered E. coli strain accumulated only small amounts of adipic acid.Subsequent analysis revealed that low expression of Tfu_1647 con-stituted the rate-limiting enzyme involved in decreased adipic acidaccumulation, with overexpression of Tfu_1647 by pTrc99A resulting inhigh yields of adipic acid.

The strategies used to increase adipic acid yield were as follows: 1)partial inhibition of succinate production from succinyl-CoA to increasethe accumulation of succinyl-CoA, the precursor of adipic acid; 2) de-letion of atoB to eliminate butyric acid production; and 3) deletion ofldhA to eliminate lactic acid production. Due to the low efficiency ofhomologous recombination in E.coli BL21 (DE3) aided by the λ-Redsystem, the CRISPR/Cas9 system was employed in this study to enablegene deletion (Jiang et al., 2015).

The bioinformatics analysis indicated that Tfu_1647 belonged to theBcd-type family of enoyl-CoA reductases (Deng and Mao, 2015), whichwas confirmed by enzymatic assay. Thus, this enzyme was reversible,with its directionality dependent upon relative concentration and redoxstate. Highly aerobic conditions promote adipate degradation, giventhat adipic acid production was reduced at high aeration rates. Higheraeration rates enhanced respiration, which competed for the reducedpotential for producing adipic acid and increased cellular yield un-associated with adipic acid production. However, in the complete ab-sence of oxygen, the TCA cycle would be too weak to generate sufficientsuccinyl-CoA for adipic acid production.

During the late stationary phase, acetate might be assimilated bycells. Gupta et al. found the same phenomenon and they observed thatthe acetate could be reassimilated (Gupta et al., 2017). According to thestoichiometry of acetate production, acetate is generated from acetyl-CoA by phosphate acetyltransferase and acetate kinase while gen-erating ATP (Kim et al., 2015; Kumari et al., 2000). However, this re-action is reversible and converted acetate to acetyl-CoA. The reverseacetate-synthesis reaction might also provide higher levels of acetyl-CoA, which is the key precursor for adipic acid synthesis. Another ex-planation involves the acetyl-coenzyme A synthetase gene (acs)(Webster, 1966; Webster, 1967), which is silent during the early stagesof growth and can be activated upon acetate accumulation in the brothduring the late stationary phase in response to rising cyclic adenosinemonophosphate (cAMP) levels, low oxygen levels, and carbon fluxthrough pathways associated with acetate metabolism (Kumari et al.,2000).

Succinyl-CoA synthetase catalyzes the conversion of succinyl-CoA tosuccinate by generating ATP (Litsanov et al., 2012). However, succinyl-CoA is a precursor for adipic acid synthesis. A reasonable pool of suc-cinyl-CoA necessary to drive carbon flux toward adipic acid synthesis ishighly important. Here, rather than deleting all genes encoding suc-cinyl-CoA synthetase, the gene sucD encoding the alpha subunit wasdeleted in E. coli BL21 (DE3), resulting in the downregulation of suc-cinate synthesis (Cheong et al., 2016). We found that sucD deletionsignificantly increased adipic acid production. Although Tfu_2576-7 isannotated as a succinyl-CoA synthetase, the similarity between

Fig. 6. Fed-batch fermentation process associated with strain Mad123146.Substrates consumptions and dissolved oxygen in 5-L bioreactor; (b) Metaboliteexcretions and cell growth during fed-batch fermentation. The highest titer wasachieved in 5-L bioreactor with 1 vvm aeration rate and 400 rpm agitation ratein SOB medium. Error bars represent the s.d. from three independent assays.

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

260

Page 8: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

Tfu_2576-7 and sucC–D is ~47% (Supplementary Figure 7); therefore,the absence of sucD could not be fully compensated by Tfu_2577 instrain Mad123146 (ΔsucD).

For NADH generation, it is obvious that the strain Mad123146,which was deficient in both lactate dehydrogenase (encoded by ldhA)and acetyl-CoA acetyltransferase (encoded by atoB), showed sig-nificantly higher NADH concentration than that of Mad136, throughoutthe whole fermentation process (Supplementary Figure 8). The elevatedNADH could significantly drive the conversion of succinyl-CoA andacetyl-CoA to 3-hydroxyadipyl-CoA, which is the second step of thereverse adipate-degradation pathway (RADP).

The cells from mid-log phase during batch culture process wereharvested and tested for the intracellular succinyl-CoA concentrations(Bennett et al., 2009). The results are shown in Supplementary Figure 9.The engineered strains without sucD (Mad3146, Mad23146 andMad123146), a subunit of succinate-CoA ligase, showed much higherintracellular succinyl-CoA concentrations than those strains with sucD(Mad146 and Mad1146). Besides, the final strain Mad123146 had thehighest level of succinyl-CoA (0.0168mmol OD (600 nm). Thus, suc-cinyl-CoA, the precursor of adipic acid was accumulated in engineeredstrains to drive the production of adipic acid.

During the aerobic fermentation, glycerol is transported to E. coli bya glycerol transporter (glpF), which is then converted to Glycerol-3-phosphate by a glycerol kinase (glpK). Two respiratory glycerol-3-phosphate dehydrogenases (G3PDHs, encoded by the glpD and glpABC)then convert sn-Glycerol-3-phosphate to Glycerone phosphate (dihy-droxyacetone Phosphate, DHAP), which enters glycolysis pathway(Murarka et al., 2008).

Koebmann et al. showed that in the aerobic condition, at least 75%of the control over glycolysis of E. coli related to in the ATP-consumingreactions (Koebmann et al., 2002b). Based on above results, they over-expressed the genes encoding part of the F1 unit of the (F1F0) H+-AT-Pase, resulting in a 70% increase in the glycolytic flux. The samestrategy was successful to uncouple of glycolysis from biomass pro-duction in Lactococcus lactis(Koebmann et al., 2002a). Thus, the nextstep, we might introduce the ATP hydrolysis by over-expressing F1 unitof the (F1F0) H+-ATPase in E. coli and channel more flux to the targetproduct, adipic acid.

5. Conclusions

In this study, we optimized the adipic acid-synthesis pathway fromT. fusca in E. coli BL21 (DE3) by rational deletion of genes hinderingadipic acid production, resulting in a maximal adipic acid titer of68.0 g L−1 from glycerol as the major substrate. To the best of ourknowledge, this constitutes the highest titer reported in E. coli (Table 1).

Acknowledgements

This study was funded by National Natural Science Foundation ofChina (31500070, 31600044, 31670095), Natural Science Foundationof Jiangsu Province (BK20150136, BK20150151), the FundamentalResearch Funds for the Central Universities (JUSRP51701A,JUSRP51705A) and the Distinguished Professor Project of JiangsuProvince (5926010241150800). We thank Prof. Ping Xu from Shanghai

Jiaotong University for valuable discussion.

Conflicts of interest

None.

Appendix A. Supporting information

Supplementary data associated with this article can be found in theonline version at http://dx.doi.org/10.1016/j.ymben.2018.04.002.

References

Bennett, B.D., Kimball, E.H., Gao, M., Osterhout, R., Van Dien, S.J., Rabinowitz, J.D.,2009. Absolute metabolite concentrations and implied enzyme active site occupancyin Escherichia coli. Nat. Chem. Biol. 5, 593–599.

Borujeni, A.E., Channarasappa, A.S., Salis, H.M., 2014. Translation rate is controlled bycoupled trade-offs between site accessibility, selective RNA unfolding and sliding atupstream standby sites. Nucleic Acids Res. 42, 2646–2659.

Cheong, S., Clomburg, J.M., Gonzalez, R., 2016. Energy- and carbon-efficient synthesis offunctionalized small molecules in bacteria using non-decarboxylative Claisen con-densation reactions. Nat. Biotechnol. 34, 556–561.

Deng, Y., Ma, L.Z., Mao, Y., 2016. Biological production of adipic acid from renewablesubstrates: current and future methods. Biochem Eng. J. 105, 16–26.

Deng, Y., Mao, Y., 2015. Production of adipic acid by the native-occurring pathway inThermobifida fusca B6. J. Appl. Microbiol. 119, 1057–1063.

Deng, Y., Olson, D.G., Zhou, J.L., Herring, C.D., Shaw, A.J., Lynd, L.R., 2013. Redirectingcarbon flux through exogenous pyruvate kinase to achieve high ethanol yields inClostridium thermocellum. Metab. Eng. 15, 151–158.

Draths, K.M., Frost, J.W., 1994. Environmentally compatible synthesis of adipic acid fromD-glucose. J. Am. Chem. Soc. 116, 399–400.

Gibson, D.G., Young, L., Chuang, R.Y., Venter, J.C., Hutchison, C.A., Smith, H.O., 2009.Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods6, 343–345.

Gupta, A., Reizman, I.M.B., Reisch, C.R., Prather, K.L.J., 2017. Dynamic regulation ofmetabolic flux in engineered bacteria using a pathway-independent quorum-sensingcircuit. Nat. Biotechnol. 35, 273–279.

Jiang, Y., Chen, B., Duan, C., Sun, B., Yang, J., Yang, S., 2015. Multigene editing in theEscherichia coli genome via the CRISPR-Cas9 system. Appl. Environ. Microbiol. 81,2506–2514.

Jung, H.M., Jung, M.Y., Oh, M.K., 2015. Metabolic engineering of Klebsiella pneumoniaefor the production of cis,cis-muconic acid. Appl. Microbiol Biotechnol. 99,5217–5225.

Kim, J.N., Ahn, S.J., Burne, R.A., 2015. Genetics and physiology of acetate metabolism bythe Pta-Ack pathway of Streptococcus mutans. Appl. Environ. Microbiol. 81,5015–5025.

Koebmann, B.J., Solem, C., Pedersen, M.B., Nilsson, D., Jensen, P.R., 2002a. Expression ofgenes encoding F(1)-ATPase results in uncoupling of glycolysis from biomass pro-duction in Lactococcus lactis. Appl. Environ. Microbiol. 68, 4274–4282.

Koebmann, B.J., Westerhoff, H.V., Snoep, J.L., Nilsson, D., Jensen, P.R., 2002b. Theglycolytic flux in Escherichia coli is controlled by the demand for ATP. J. Bacteriol.184, 3909–3916.

Kumari, S., Beatty, C.M., Browning, D.F., Busby, S.J., Simel, E.J., Hovel-Miner, G., Wolfe,A.J., 2000. Regulation of acetyl coenzyme A synthetase in Escherichia coli. J.Bacteriol. 182, 4173–4179.

Li, L.L., Chang, M., Tao, G.J., Wang, X.S., Liu, Y., Liu, R.J., Jin, Q.Z., Wang, X.G., 2016.Analysis of phospholipids in Schizochytrium sp. S31 by using UPLC-Q-TOF-MS. Anal.Methods-Uk. 8, 763–770.

Litsanov, B., Brocker, M., Bott, M., 2012. Toward homosuccinate fermentation: metabolicengineering of Corynebacterium glutamicum for anaerobic production of succinatefrom glucose and formate. Appl. Environ. Microbiol. 78, 3325–3337.

Murarka, A., Dharmadi, Y., Yazdani, S.S., Gonzalez, R., 2008. Fermentative utilization ofglycerol by Escherichia coli and its implications for the production of fuels andchemicals. Appl. Environ. Microbiol. 74, 1124–1135.

Niu, W., Draths, K.M., Frost, J.W., 2002. Benzene-free synthesis of adipic acid.Biotechnol. Prog. 18, 201–211.

Polen, T., Spelberg, M., Bott, M., 2013. Toward biotechnological production of adipic acidand precursors from biorenewables. J. Biotechnol. 167, 75–84.

Table 1Comparison of cell factories for production of adipic acid.

Microorganism Media Cultivation mode Titer (g L− 1) productivity (g/l/h) Yield (g/g) Yield(% theoretical yield) Reference

E. coli R/2 Shake flask 0.000673 0.0000056 0.0000673 0.012 (Yu et al., 2014)E. coli MOPS Batch 2.5 0.017 0.05 9.47 (Cheong et al., 2016)E. coli N.A. Shake flask 2.03652 N.A. 0.202 37.41 (Draths and Frost, 1994)Thermobifida fusca B6 Hagerdahl Batch 2.23 0.034 0.0446 8.26 (Deng and Mao, 2015)E. coli SOB Fed-batch 68.0 0.810 0.378 72.7 This study

*N.A.: not available.

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

261

Page 9: Metabolic engineering of Escherichia coli for producing ...homepages.rpi.edu/~koffam/papers2/2018_Zhao.pdf · Metabolic engineering of Escherichia coli for producing adipic acid through

Saini, M., Wang, Z.W., Chiang, C.J., Chao, Y.P., 2014. Metabolic engineering ofEscherichia coli for production of butyric acid. J. Agric. Food Chem. 62, 4342–4348.

Sander, J.D., Joung, J.K., 2014. CRISPR-Cas systems for editing, regulating and targetinggenomes. Nat. Biotechnol. 32, 347–355.

Sato, K., Aoki, M., Noyori, R., 1998. A "Green" route to adipic acid: direct oxidation ofcyclohexenes with 30 percent hydrogen peroxide. Science 281, 1646–1647.

Schagger, H., 2006. Tricine-SDS-PAGE. Nat. Protoc. 1, 16–22.Takamura, Y., Nomura, G., 1988. Changes in the intracellular concentration of acetyl-CoA

and malonyl-CoA in relation to the carbon and energy metabolism of Escherichia coliK12. J. Gen. Microbiol. 134, 2249–2253.

Tsuji, A., Okada, S., Hols, P., Satoh, E., 2013. Metabolic engineering of Lactobacillusplantarum for succinic acid production through activation of the reductive branch ofthe tricarboxylic acid cycle. Enzym. Microb. Technol. 53, 97–103.

U.S. Environmental Protection Agency, 2011. O. o. A. P., Inventory of U.S. GreenhouseGas Emissions and Sinks: 1990–2009.

Vardon, D.R., Franden, M.A., Johnson, C.W., Karp, E.M., Guarnieri, M.T., Linger, J.G.,Salm, M.J., Strathmann, T.J., Beckham, G.T., 2015. Adipic acid production fromlignin. Energy Environ. Sci. 8, 617–628.

Weber, C., Brueckner, C., Weinreb, S., Lehr, C., Essl, C., Boles, E., 2012. Biosynthesis ofcis,cis-muconic acid and its aromatic precursors, catechol and protocatechuic acid,from renewable feedstocks by Saccharomyces cerevisiae. Appl. Environ. Microbiol. 78,

8421–8430.Webster Jr., L.T., 1966. Studies of the acetyl coenzyme A synthetase reaction. IV. The

requirement for monovalent cations. J. Biol. Chem. 241, 5504–5510.Webster Jr., L.T., 1967. Studies of the acetyl coenzyme A synthetase reaction. V. The

requirement for monovalent and divalent cations in partial reactions involving en-zyme-bound acetyl adenylate. J. Biol. Chem. 242, 1232–1240.

Werpy, T., Petersen, G., 2004. Top Value Added Chemicals from Biomass. US Departmentof Energy, Washington DC.

Wu, C.M., Lee, T.H., Lee, S.N., Lee, Y.A., Wu, J.Y., 2004. Microbial synthesis of cis,cis-muconic acid by Sphingobacterium sp GCG generated from effluent of a styrenemonomer (SM) production plant. Enzym. Microb. Technol. 35, 598–604.

Wu, C.M., Wu, C.C., Su, C.C., Lee, S.N., Lee, Y.A., Wu, J.Y., 2006. Microbial synthesis ofcis,cis-muconic acid from benzoate by Sphingobacterium sp mutants. Biochem Eng. J.29, 35–40.

Yu, J.L., Xia, X.X., Zhong, J.J., Qian, Z.G., 2014. Direct biosynthesis of adipic acid from asynthetic pathway in recombinant Escherichia coli. Biotechnol. Bioeng. 111,2580–2586.

Yu, L., Xu, M.M., Tang, I.C., Yang, S.T., 2015. Metabolic engineering of Clostridium tyr-obutyricum for n-butanol production through co-utilization of glucose and xylose.Biotechnol. Bioeng. 112, 2134–2141.

M. Zhao et al. Metabolic Engineering 47 (2018) 254–262

262