strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-bdo)...

31
Contents lists available at ScienceDirect Biotechnology Advances journal homepage: www.elsevier.com/locate/biotechadv Research review paper Strain engineering for microbial production of value-added chemicals and fuels from glycerol Adam W. Westbrook , Dragan Miscevic, Shane Kilpatrick, Mark R. Bruder, Murray Moo-Young, C. Perry Chou Department of Chemical Engineering, Waterloo, Ontario, Canada ARTICLE INFO Keyword: Glycerol Metabolic engineering Biorefinery Biofuels E. coli Clostridium Klebsiella Citrobacter Lactobacillus ABSTRACT While the widespread reliance on fossil fuels is driven by their low cost and relative abundance, this fossil-based economy has been deemed unsustainable and, therefore, the adoption of sustainable and environmentally compatible energy sources is on the horizon. Biorefinery is an emerging approach that integrates metabolic engineering, synthetic biology, and systems biology principles for the development of whole-cell catalytic platforms for biomanufacturing. Due to the high degree of reduction and low cost, glycerol, either refined or crude, has been recognized as an ideal feedstock for the production of value-added biologicals, though microbial dissimilation of glycerol sometimes can be difficult particularly under anaerobic conditions. While strain de- velopment for glycerol biorefinery is widely reported in the literature, few, if any, commercialized bioprocesses have been developed as a result, such that engineering of glycerol metabolism in microbial hosts remains an untapped opportunity in biomanufacturing. Here we review the recent progress made in engineering microbial hosts for the production of biofuels, diols, organic acids, biopolymers, and specialty chemicals from glycerol. We begin with a broad outline of the major pathways for fermentative and respiratory glycerol dissimilation and key end metabolites, and then focus our analysis on four key genera of bacteria known to naturally dissimilate glycerol, i.e. Klebsiella, Citrobacter, Clostridium, and Lactobacillus, in addition to Escherichia coli, and system- atically review the progress made toward engineering these microorganisms for glycerol biorefinery. We also identify the major biotechnological and bioprocessing advantages and disadvantages of each genus, and bot- tlenecks limiting the production of target metabolites from glycerol in engineered strains. Our analysis culmi- nates in the development of potential strategies to overcome the current technical limitations identified for commonly employed strains, with an outlook on the suitability of different hosts for the production of key metabolites and avenues for their future development into biomanufacturing platforms. https://doi.org/10.1016/j.biotechadv.2018.10.006 Received 16 December 2017; Received in revised form 3 October 2018; Accepted 10 October 2018 Abbreviations: 3-HPA, 3-hydroxypropionaldehyde; 1,2-PDO, 1,2-propanediol; 1,3-PDO, 1,3-propanediol; 1,3-PDOOR, 1,3-PDO oxidoreductase; 2,3-BDO, 2,3-bu- tanediol; 3-HH, 3-hydroxyhexanoate; 3-HP, 3-hydroxypropionic acid; 3-HV, 3-hydroxyvalerate; (R)-3-HV-CoA, (R)-3-hydroxyvaleryl-CoA; ALDH, aldehyde dehy- drogenase; ACE, Allele-Coupled Exchange; AOR, aldehyde oxidoreductase; asRNA, antisense RNA; ATP, adenosine Triphosphate; cAMP, cyclic adenosine mono- phosphate; CRE, catabolite repression element; CRISPR, Clustered Regularly Interspaced Palindromic Repeats; CRISPRi, CRISPR interference; Cas9, CRISPR- associated [protein] 9; CRP, cAMP receptor protein; dcw, dry cell weight; DHA, dihydroxyacetone; DHAK, DHA kinase; DHAP, dihydroxyacetone phosphate; DODHt, diol dehydratase; EDP, Entner-Doudoroff pathway; FAD, flavin adenine dinucleotide; FHL, formate hydrogen lyase; FDH, formate dehydrogenase; G3P, glycerol-3- phosphate; GDHt, glycerol dehydratase; GluDH, glucose dehydrogenase; GlyDH, glycerol dehydrogenase; GK, glycerol kinase; LDH, lactate dehydrogenase; MDH, malate dehydrogenase; MG, methylglyoxal; MGR, MG reductase; MGS, MG synthase; NADH, nicotinamide adenine dinucleotide; NOX, NADH oxidase; PHA, poly- hydroxyalkanoate; P(3HB-co-3HH), poly(3-hydroxyburyrate-co-3-hydroxyhexanoate); P(3HB-co-3HV), poly(3-hydroxybutyrate-co-3-hydroxyvalerate); P(3HP), poly (3-hydroxypropionate); PDU, propanediol utilization; PEP, phosphoenolpyruvate; PCK, PEP carboxykinase; PFL, pyruvate formate lyase; PDC, pyruvate decarbox- ylase; PDH, pyruvate dehydrogenase; PFOR, pyruvate:ferrodoxin oxidoreductase; PGL, phosphogluconolactonase; PPC, PEP carboxylase; PPP, pentose phosphate pathway; PTA, phosphotransacetylase; PYK, pyruvate kinase; PO, pyruvate oxidase; RBS, ribosome binding site; Sbm, sleeping beauty mutase; TCA, tricarboxylic acid; UDH, pyridine nucleotide transhydrogenase Corresponding authors. E-mail addresses: [email protected] (A.W. Westbrook), [email protected] (C.P. Chou). Biotechnology Advances 37 (2019) 538–568 Available online 17 October 2018 0734-9750/ © 2018 Elsevier Inc. All rights reserved. T

Upload: others

Post on 28-Jul-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Contents lists available at ScienceDirect

Biotechnology Advances

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

Research review paper

Strain engineering for microbial production of value-added chemicals andfuels from glycerolAdam W. Westbrook⁎, Dragan Miscevic, Shane Kilpatrick, Mark R. Bruder, Murray Moo-Young,C. Perry Chou⁎

Department of Chemical Engineering, Waterloo, Ontario, Canada

A R T I C L E I N F O

Keyword:GlycerolMetabolic engineeringBiorefineryBiofuelsE. coliClostridiumKlebsiellaCitrobacterLactobacillus

A B S T R A C T

While the widespread reliance on fossil fuels is driven by their low cost and relative abundance, this fossil-basedeconomy has been deemed unsustainable and, therefore, the adoption of sustainable and environmentallycompatible energy sources is on the horizon. Biorefinery is an emerging approach that integrates metabolicengineering, synthetic biology, and systems biology principles for the development of whole-cell catalyticplatforms for biomanufacturing. Due to the high degree of reduction and low cost, glycerol, either refined orcrude, has been recognized as an ideal feedstock for the production of value-added biologicals, though microbialdissimilation of glycerol sometimes can be difficult particularly under anaerobic conditions. While strain de-velopment for glycerol biorefinery is widely reported in the literature, few, if any, commercialized bioprocesseshave been developed as a result, such that engineering of glycerol metabolism in microbial hosts remains anuntapped opportunity in biomanufacturing. Here we review the recent progress made in engineering microbialhosts for the production of biofuels, diols, organic acids, biopolymers, and specialty chemicals from glycerol. Webegin with a broad outline of the major pathways for fermentative and respiratory glycerol dissimilation and keyend metabolites, and then focus our analysis on four key genera of bacteria known to naturally dissimilateglycerol, i.e. Klebsiella, Citrobacter, Clostridium, and Lactobacillus, in addition to Escherichia coli, and system-atically review the progress made toward engineering these microorganisms for glycerol biorefinery. We alsoidentify the major biotechnological and bioprocessing advantages and disadvantages of each genus, and bot-tlenecks limiting the production of target metabolites from glycerol in engineered strains. Our analysis culmi-nates in the development of potential strategies to overcome the current technical limitations identified forcommonly employed strains, with an outlook on the suitability of different hosts for the production of keymetabolites and avenues for their future development into biomanufacturing platforms.

https://doi.org/10.1016/j.biotechadv.2018.10.006Received 16 December 2017; Received in revised form 3 October 2018; Accepted 10 October 2018

Abbreviations: 3-HPA, 3-hydroxypropionaldehyde; 1,2-PDO, 1,2-propanediol; 1,3-PDO, 1,3-propanediol; 1,3-PDOOR, 1,3-PDO oxidoreductase; 2,3-BDO, 2,3-bu-tanediol; 3-HH, 3-hydroxyhexanoate; 3-HP, 3-hydroxypropionic acid; 3-HV, 3-hydroxyvalerate; (R)-3-HV-CoA, (R)-3-hydroxyvaleryl-CoA; ALDH, aldehyde dehy-drogenase; ACE, Allele-Coupled Exchange; AOR, aldehyde oxidoreductase; asRNA, antisense RNA; ATP, adenosine Triphosphate; cAMP, cyclic adenosine mono-phosphate; CRE, catabolite repression element; CRISPR, Clustered Regularly Interspaced Palindromic Repeats; CRISPRi, CRISPR interference; Cas9, CRISPR-associated [protein] 9; CRP, cAMP receptor protein; dcw, dry cell weight; DHA, dihydroxyacetone; DHAK, DHA kinase; DHAP, dihydroxyacetone phosphate; DODHt,diol dehydratase; EDP, Entner-Doudoroff pathway; FAD, flavin adenine dinucleotide; FHL, formate hydrogen lyase; FDH, formate dehydrogenase; G3P, glycerol-3-phosphate; GDHt, glycerol dehydratase; GluDH, glucose dehydrogenase; GlyDH, glycerol dehydrogenase; GK, glycerol kinase; LDH, lactate dehydrogenase; MDH,malate dehydrogenase; MG, methylglyoxal; MGR, MG reductase; MGS, MG synthase; NADH, nicotinamide adenine dinucleotide; NOX, NADH oxidase; PHA, poly-hydroxyalkanoate; P(3HB-co-3HH), poly(3-hydroxyburyrate-co-3-hydroxyhexanoate); P(3HB-co-3HV), poly(3-hydroxybutyrate-co-3-hydroxyvalerate); P(3HP), poly(3-hydroxypropionate); PDU, propanediol utilization; PEP, phosphoenolpyruvate; PCK, PEP carboxykinase; PFL, pyruvate formate lyase; PDC, pyruvate decarbox-ylase; PDH, pyruvate dehydrogenase; PFOR, pyruvate:ferrodoxin oxidoreductase; PGL, phosphogluconolactonase; PPC, PEP carboxylase; PPP, pentose phosphatepathway; PTA, phosphotransacetylase; PYK, pyruvate kinase; PO, pyruvate oxidase; RBS, ribosome binding site; Sbm, sleeping beauty mutase; TCA, tricarboxylicacid; UDH, pyridine nucleotide transhydrogenase

⁎ Corresponding authors.E-mail addresses: [email protected] (A.W. Westbrook), [email protected] (C.P. Chou).

Biotechnology Advances 37 (2019) 538–568

Available online 17 October 20180734-9750/ © 2018 Elsevier Inc. All rights reserved.

T

Page 2: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

1. Introduction

Currently, it is estimated that ~90% of the world’s energy re-quirements are met through the use of non-renewable fossil fuels suchas petroleum, natural gas, bitumens, oil shale, and coal (Parikka, 2004;Pöschl et al., 2010; Srirangan et al., 2012). These resources, particularlyoil and natural gas, are also the most important feedstock for the pro-duction of fine and commodity chemicals (Basiago, 1994; Liu et al.,2010). Nevertheless, the unrestricted use of fossil fuels is consideredunsustainable due to the finite supply and unequal distribution ofnatural reserves, coupled with climate change resulting from increasinggreenhouse gas emissions (Basiago, 1994; Börjesson, 2009; Sriranganet al., 2012). As a result, biorefinery becomes an emerging approachthrough integrating systems biology, genetic engineering, syntheticbiology, and metabolic engineering principles for the development ofwhole-cell biocatalytic platforms for manufacturing purposes (Menonand Rao, 2012; Octave and Thomas, 2009). While this appears to be apromising avenue, a major detriment to the use of biological platformsis the high cost of feedstock (Menon and Rao, 2012; Octave andThomas, 2009). Industrially, value-added products are derived fromagricultural crops (i.e. first-generation feedstock) or from lig-nocellulosic crops and agricultural wastes (i.e. second-generationfeedstock) (Srirangan et al., 2012). First-generation feedstocks (e.g.corn, starch, oilseed, and sugar) often have a high energy, oil, andcarbohydrate content and are currently used for the production ofbiodiesel (and other bio-esters), bioethanol (and other bioalcohols), andbiogas (Hein and Leemans, 2012; Srirangan et al., 2012). However,these bioprocess schemes are considered unsustainable due to theircompetition with the human food and animal feed markets, and therequirement for large arable lands (Schmidhuber, 2008). While notdirectly impacting the cost and availability of food supplies, second-generation feedstocks are not practical due to their inherent recalci-trance and the high cost of pre-treatment technologies required for theirvalorization (Luo et al., 2010; Srirangan et al., 2012).

Accordingly, several alternative biorefinery schemes have beenproposed to utilize biomass feedstock in an economically viable mannerto compete with existing petroleum-refinery technologies. One suchmodel is co-utilization of the starting feedstock and the byproductstream (Clomburg and Gonzalez, 2013; Yazdani and Gonzalez, 2007).For instance, glycerol is a major sugar alcohol byproduct associatedwith bioethanol (Raynaud, and xe, line, Sar, xe, abal, P., Meynial-Salles,I., Croux, C., Soucaille, P., 2003; Zeng and Biebl, 2002) and biodiesel(Balat and Balat, 2010) production. Transesterification of oils and fatswith an alcohol generates ~1 lb crude glycerol (with at least 80%purity) for every 10 lb biodiesel produced (López et al., 2009; Sabourin-Provost and Hallenbeck, 2009). Likewise, all first-generation-basedbioethanol programs generate significant amounts of glycerol as a fer-mentative end-product (Kim et al., 2008; Zeng and Biebl, 2002). Cur-rently, biodiesel and bioethanol represent the two largest biomass en-ergy programs in the world, and their significant growth led to a morethan 10-fold increase in global glycerol production between 2004 and2011, with ~700 million lb crude glycerol produced in the UnitedStates alone in 2011 (Clomburg and Gonzalez, 2013; López et al.,2009). Additionally, many other industries which derive products fromanimal fats and vegetable oils generate waste streams of high glycerolcontent, e.g. the oleochemical industry (Sabourin-Provost andHallenbeck, 2009; Yazdani and Gonzalez, 2007). This surplus in gly-cerol has plummeted its price to ~2.5 cents/lb in recent years, effec-tively rendering it a waste product with an associated disposal cost (Rajet al., 2008).

As a carbon source, glycerol offers several distinctive advantagesover traditional fermentable sugars. Due to the high degree of reduction(reductance κ = 4.67), glycolytic degradation of glycerol generatesapproximately twice the number of reducing equivalents (i.e. NADH)compared to xylose and glucose (κ = 4) (Murarka et al., 2008; Yazdaniand Gonzalez, 2007). Therefore, harnessing glycerol metabolism not

only results in higher yields but also expands the repertoire of chemi-cals and fuels generated from microbial systems. However, microbialdissimilation of glycerol is difficult under anaerobic conditions, as thecellular redox balance (i.e. the NAD+/NADH ratio) must be properlymaintained through terminal transfer of electrons to internally pro-duced organic compounds as opposed to molecular oxygen (Celińska,2010; Yazdani and Gonzalez, 2007). As a result, only select organisms(e.g. Clostridium pasteurianum and Klebsiella pneumoniae) are capable offermenting glycerol anaerobically. Glycerol metabolism generates in-termediates such as dihydroxyacetone (DHA) and 3-hydro-xypropionaldehyde (3-HPA), and reduced end-products such asethanol, 1,2-propanediol (1,2-PDO), 1,3-propanediol (1,3-PDO), and2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarkaet al., 2008; Yazdani and Gonzalez, 2007). Accordingly, glycerol fer-mentation has been extensively explored in both native and geneticallytractable hosts for the production of various value-added chemicals andfuels, including advanced alcohols (e.g. 1-propanol), organic acids (e.g.succinic, propionic, and 3-hydroxypropionic acids), natural products(e.g. terepenes), ketones (e.g. acetone and butanone) and bio(co)poly-mers (e.g. polyhydroxyalkanoate (PHA) and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) [P(3HB-co-3HV)]). This article reviews our cur-rent understanding of glycerol metabolism and recent approaches ofstrain engineering for glycerol biorefinery.

2. Overview of glycerol metabolism in microbes

Due to the high reductance of glycerol, there is an excess of reducingequivalents that remain unbalanced such that the overall redox balancewithin a cell is a vital principle that governs the nature and quantity ofmetabolites during glycerol dissimilation (Clomburg and Gonzalez,2013). The natural fermentative metabolism of glycerol has beenthoroughly studied in the Enterobacteriaceae family (Booth, 2005;Murarka et al., 2008). Under fermentative conditions, microorganismsmust be metabolically capable of consuming glycerol in the absence ofexternal electron acceptors. Specifically, for 1.1 mM of glycerol in-corporated into cell mass, 0.6 mM reducing equivalents (i.e. NADH) aregenerated. This is not an issue for other common carbohydrates, such asglucose and xylose, since their degree of reduction is lower than that ofthe cell biomass. Furthermore, excess electrons cannot be consumedthrough the production of reduced products such as ethanol or succi-nate since their production from glycerol is already redox-balanced.Thus, to achieve redox poise, there must be a metabolic pathway gen-erating more reduced products for consuming surplus reducingequivalents. The formation of 1,3-PDO (Homann et al., 1990; Saint-Amans et al., 1994; Schütz and Radler, 1984) and 1,2-PDO (Clomburgand Gonzalez, 2011) via ATP-neutral pathways results in successfulanaerobic dissimilation of glycerol due to the highly reduced nature ofthese metabolites (κ = 5.33).

Like other small and uncharged molecules, glycerol can cross thecytoplasmic membrane via two mechanisms: (i) passive diffusion underhigh concentrations; and (ii) facilitated diffusion with the aid of atransport protein (i.e. GlpF) under low concentrations (da Silva et al.,2009; Sun et al., 2008). Intracellular glycerol is subsequently metabo-lized either in the presence of electron acceptors (i.e. respiratorybranch) or in the absence of electrons acceptors (i.e. fermentativebranch) (Gonzalez et al., 2008). Fermentative metabolism of glycerolhas been previously reported in microbes within the genera of Klebsiella(Forage and Foster, 1982; Homann et al., 1990), Citrobacter (Danielet al., 1995; Seifert et al., 2001), Clostridium (Abbad-Andaloussi et al.,1995; Biebl, 2001; Macis et al., 1998), and Lactobacillus (Talarico et al.,1990), as well as in Escherichia coli (Dharmadi et al., 2006; Murarkaet al., 2008). The natural fermentative bioconversion of glycerol to amore reduced metabolite, such as 1,3-PDO, occurs both reductively andoxidatively in Klebsiella sp., Citrobacter sp., Clostridium sp., and certainstrains of Lactobacilli sp. (da Cunha and Foster, 1992; da Silva et al.,2009). In the reductive branch, a B12-dependent glycerol dehydratase

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

539

Page 3: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

(GDHt) dehydrates glycerol to form 3-HPA, which is subsequently re-duced to 1,3-PDO (Fig. 1) (Murarka et al., 2008). In the latter reaction,NAD+ is regenerated by NADH-dependent 1,3-PDO oxidoreductase(1,3-PDOOR), resulting in a balanced internal redox state (González-Pajuelo et al., 2006). Glycerol fermentation is the only known processfor anaerobic conversion for 1,3-PDO production (Homann et al.,1990). As 1,3-PDO is more reduced than glycerol, its formation is ac-companied by a more oxidized co-product (Clomburg and Gonzalez,2013). In Klebsiella sp. and Citrobacter sp., acetic acid is the major co-product during 1,3-PDO production under fermentative conditions(Homann et al., 1990), whereas mixed co-products are formed (i.e.butyrate, butanol, ethanol, and lactate) in Clostrodium sp. (Biebl, 2001;Clomburg and Gonzalez, 2013; Saint-Amans et al., 2001). Biosynthesisof 1,3-PDO relies on the availability of NADH, which can be naturallygenerated in Klebsiella sp. and Lactobacillus sp. via oxidation of 3-HPAinto 3-hydroxypropionic acid (3-HP) by NAD+-dependent aldehydedehydrogenase (ALDH) (Zhu et al., 2015). In the oxidative branch,NAD+-dependent glycerol dehydrogenase (GlyDH) catalyzes thetransformation of glycerol to dihydroxyacetone (DHA) (Fig. 1), which issubsequently phosphorylated by a DHA kinase (DHAK) to form dihy-droxyacetone phosphate (DHAP) (Shams Yazdani and Gonzalez, 2008).In Klebsiella sp. and Citrobacter sp., the genes encoding enzymes specific

to glycerol metabolism (i.e. DhaB/DhaBCE, DhaT, DhaD, and DhaK) arelocated in a DHA (dha) regulon (Zhu et al., 2002). In certain Clostridiumsp., such as Clostridium butyricum, the dha regulon is composed of threegenes encoding GDHt (i.e. DhaB1 encoded by dhaB1), its reactivatingprotein (i.e. DhaB2 encoded by dhaB2), and 1,3-PDOOR (i.e. DhaTencoded by dhaT) (Raynaud, and xe, line, Sar, xe, abal, P., Meynial-Salles, I., Croux, C., Soucaille, P., 2003), while the dha regulon can befound in two spatially different clusters on the chromosome of Clos-tridium perfringens (Ignatova et al., 2003) (Table 1).

E. coli can ferment glycerol anaerobically in a pH-dependent mannerwithout external electron acceptors (Booth, 2005; Dharmadi et al.,2006). The internal availability of CO2 generated by formate hydrogenlyase (FHL)-mediated oxidation of formate enables E. coli to anaerobi-cally metabolize glycerol. For E. coli to synthesize small molecules andfatty acids, it requires a steady supply of bicarbonate/CO2 substrates.Lowering the pH of the media as well as potassium and phosphateconcentrations under fermentative conditions activates the transcrip-tion of FHL and subsequently achieves the cellular requirement for CO2

upon formate accumulation (Murarka et al., 2008). Additionally, for-mate oxidation also produces H2 and, therefore, can contribute to re-ducing equivalent accumulation that other redox-balanced pathways,such as succinate and ethanol-generating pathways, cannot (Dharmadi

Fig. 1. Schematic representation of the major glycerol catabolism pathways in bacteria. Blue arrows represent reactions in the fermentative branch, displaying bothreductive and oxidative routes; green arrows represent reactions in the respiratory branch; red arrows represent reactions in the 1,2-PDO pathway; and gold arrowsrepresent reactions in the 2,3-BDO pathway. Solid line: direct bioconversion step; dashed line: multi-step bioconversion. Enzyme abbreviations: ACK, acetate kinase;ADH, bifunctional aldehyde/alcohol dehydrogenase; ae-G3PDH, aerobic glycerol-3-phosphate dehydrogenase; AKR, aldo-keto reductase; ALDH, aldehyde dehy-drogenase; ALDC, α-acetolactate decarboxylase; ALS, α-acetolactate synthase; an-G3PDH, anaerobic glycerol-3-phosphate dehydrogenase; AR, acetoin reductase;DODHt, diol dehydratase; DHAK, dihydroxyacetone kinase; FHL, formate hydrogen lyase; FRD, fumarate reductase; GDHt, glycerol dehydratase; GK, glycerol kinase;GlpF, glycerol diffusion facilitator; GlyDH, glycerol dehydrogenase; LDH, lactate dehydrogenase; MGR, methylglyoxal reductase; MGS, methylglyoxal synthase; PDH,pyruvate dehydrogenase; PduL, phosphate propanoyltransferase; PduP, propionaldehyde dehydrogenase; PduW, propionate kinase; PFL, pyruvate formate-lyase;PTA, phosphotransacetylase; PK, pyruvate kinase; 1,2-PDOOR; 1,2-propanediol oxidoreductase; 1,3-PDOOR, 1,3-propanediol oxidoreductase. Chemical inter-mediates and product abbreviations: α-AL, α-acetolactate; acetyl-P, acetyl phosphate; DHA, dihydroxyacetone; DHAP, dihydroxyacetone phosphate; G3P, glycerol-3-phosphate; MG, methylglyoxal; PEP, phosphoenolpyruvate; 1,2-PDO, 1,2-propanediol; 1,3-PDO, 1,3-propanediol; 3-HP, 3-hydroxypropionate; 3-HP-CoA, 3-hydro-xypropionyl-CoA; 3-HP-P, 3-hydroxypropionyl phosphate.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

540

Page 4: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Table1

Sum

mar

yof

enzy

mes

com

pris

ing

natu

ralp

athw

ays

for

glyc

erol

cata

bolis

min

Kleb

siella

sp.,Citrob

acter

sp.,Clostridium

sp.,La

ctob

acillus

sp.,

andE.

coli.

Locu

sta

gs,a

ndre

spec

tive

enzy

me

nam

esw

here

avai

labl

e,ar

epr

ovid

edfo

ra

repr

esen

tativ

est

rain

for

each

genu

sor

spec

ies.

Each

num

bere

dco

lum

nco

rres

pond

sto

one

ofth

enu

mbe

red

reac

tions

inFi

g.1.

Ferm

enta

tive

met

abol

ism

Resp

irat

ory

met

abol

ism

Org

anis

m1

23

45

67

89

Kleb

siella

sp.1,

2

(K.p

neum

oniae

MG

H78

578)

GD

Ht:

Dha

B⁎,a

(RS1

8900

,RS

1889

5,RS

1889

0)

DO

DH

t:Pd

uCD

E(R

S171

90⁎ ,

RS17

195,

RS17

200)

1,3-

PDO

OR:

Dha

T⁎

(RS1

8910

),Yq

hD⁎,

b

(RS1

8380

)

ALD

H:

YdcW

(RS1

0355

),Yn

eI⁎,

c

(RS0

8785

),Fe

aB⁎

(RS0

7805

),Pu

uC⁎

(RS0

5455

),Ba

dH⁎

(RS0

3160

),G

abD

(RS0

1360

)

PduP

(RS1

7245

)

PduL

(RS1

7225

)Pd

uW⁎

(RS1

7275

)G

lyD

H:

Dha

D⁎

(RS1

8930

),G

ldA

(RS2

2860

)

DH

AK:

Dha

KI⁎

(RS1

8955

),D

haKI

I⁎

(RS1

8940

,RS

1894

5,RS

1895

0)

GK:

Glp

K⁎,d

(RS2

1635

)

Citrob

acter

sp.3,

4

(C.f

reun

dii

CFN

IH1)

GD

Ht:

Dha

BCE⁎

(RS0

2570

,RS

0256

5,RS

0256

0)

DO

DH

t:Pd

uCD

E(R

S208

65,

RS20

870,

RS20

875)

1,3-

PDO

DH

:D

haT⁎,

b

(RS0

2580

)

NA

NA

NA

NA

Gly

DH

:D

haD

⁎,c

(RS0

2600

)G

ldA

(RS0

7260

)

DH

AK:

Dha

K⁎

(RS1

9390

)

GK:

Glp

Kd

(RS0

5005

)

Clostridium

sp.5,

6

(C.p

asteurianu

mA

TCC

6013

)

GD

Ht:

PduC

DE/

Dha

BCE

(RS1

1085

,RS

1108

0,RS

1107

5)D

haB1

(Gen

bank

AY1

1298

9.1)

j

1,3-

PDO

H:

Dha

T2(R

S110

50)

Dha

T1(R

S053

50)

NA

NA

NA

NA

Gly

DH

:G

ldA

(RS0

5875

,RS

1866

5),

RS13

630,

RS16

340

DH

AK:

Dha

K(R

S200

60)

GK:

Glp

K(C

A_C

1321

)k

Lactob

acillus

sp.7,

8

(L.r

euteri

DSM

2001

6)

DO

DH

t:Pd

uCD

E⁎

(RS0

9105

,RS

0910

0,RS

0909

5)

1,3-

PDO

OR:

RS00

160⁎ ,

RS09

040⁎

ALD

H:

PduP

⁎,e

(RS0

9045

)

PduP

(RS0

9045

)Pd

uL(R

S090

70)

PduW

(RS0

9035

)G

lyD

H:

RS09

575

NA

GK:

RS05

600

E.co

liN

AA

LDR:

YqhD

(b30

11)

ALD

H:

Ald

H⁎,

c

(b13

00)

NA

NA

NA

Gly

DH

:G

ldA

(b39

45)

DH

AK:

Dha

KLM

(b12

00,

b119

9,b1

198)

GK:

Glp

K⁎

(b39

26)

Resp

irat

ory

met

abol

ism

1,2-

PDO

Succ

inat

eLa

ctat

e2,

3-BD

O

Org

anis

m10

1112

1314

1516

1718

1920

Kleb

siella

sp.1,

2

(K.p

neum

oniae

MG

H78

578)

ae-G

3PD

H:

Glp

D(R

S204

85)

an-G

3PD

H:

Glp

ABC

(RS1

4215

,

NA

NA

NA

NA

NA

FRD

:Fr

dABC

D(R

S245

50,

PK:

PykA

(RS1

2780

),

LDH

:Ld

hA⁎

(RS0

7775

),

ALS

:Bu

dB⁎

(RS1

1100

)(con

tinue

don

next

page

)

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

541

Page 5: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Table1

(con

tinue

d)

Resp

irat

ory

met

abol

ism

1,2-

PDO

Succ

inat

eLa

ctat

e2,

3-BD

O

RS14

220,

RS14

225)

RS24

545,

RS24

540,

RS24

535)

PykF

(RS1

1500

)Ld

h1(R

S087

70),

Ldh2

(RS2

1315

)Citrob

acter

sp.3,

4

(C.f

reun

dii

CFN

IH1)

ae-G

3PD

H:

Glp

D(R

S047

15)

an-G

3PD

H:

Glp

ABC

(RS2

2435

,RS

2244

0,RS

2244

5)

NA

NA

NA

NA

NA

FRD

:Fr

dABC

D(R

S083

35,

RS08

330,

RS08

325,

RS08

320)

PK:

RS16

640,

RS20

115

LDH

:Ld

hA⁎

(RS1

8135

)

NA

Clostridium

sp.5,

6

(C.p

asteurianu

mA

TCC

6013

)

NA

an-G

3PD

H:

Glp

A(C

A_C

1322

)k

NA

NA

NA

NA

NA

NA

PK:

RS13

430

LDH

:RS

1342

0,RS

0850

5

NA

Lactob

acillus

sp.7,

8

(L.r

euteri

DSM

2001

6)

ae-G

3PD

H:

RS01

940

NA

NA

NA

NA

NA

NA

FRD

:RS

0804

0PK

:RS

0402

5LD

Hf :

RS00

995,

RS03

840,

RS03

845,

RS08

830,

RS09

895

NA

E.co

liae

-G3P

DH

:G

lpD

(b34

26)

an-G

3PD

HG

lpA

BC⁎

(b22

41,b

2242

,b2

243)

MG

S:M

gsA

(b09

63)

AKR

:Ye

aE⁎

(b17

81),

YghZ

(b30

01),

YafB

(b02

07),

YqhE

(b30

12),

YqhD

(b30

11)

glyD

H:

Gld

A⁎

(b39

45)

MG

RYd

jG(b

1771

)

1,2-

PDO

R:Fu

cO⁎

(b27

99)

FRD

:Fr

dABC

D⁎

(b41

54,

b415

3,b4

152,

b415

1)

PK:

PykA

(b18

54),

PykF

(b16

76)

LDH

:Ld

hA⁎

(b13

80)

NA

2,3-

BDO

Etha

nol

Ace

tate

Org

anis

m21

2223

2425

2627

2829

Kleb

siella

sp.1,

2

(K.p

neum

oniae

MG

H78

578)

AR:

BudC

(RS1

1105

)

ALD

C:Bu

dA⁎

(RS1

1095

)

AR:

BudC

(RS1

1105

)

PFL:

PflA

B(R

S050

10,

RS05

015)

PDC:

RS14

705

PDH

:A

ceE

(RS0

0635

),Po

xB⁎

(RS0

4870

)D

HLA

T:A

ceF

(RS0

0640

)

ALD

H:

Ald

A⁎

(RS0

8050

)RS

0289

0,RS

0542

5,A

stD

(RS0

6540

)

AD

H:

YigB

(RS2

5235

),RS

2268

0,Yq

hDb

(RS1

8380

),A

dhE⁎

(RS1

1855

),A

dhP

(RS0

9985

)

PTA

:Pt

a⁎

(RS1

4425

),Eu

tD(R

S149

85)

ACK

:A

ckA

(RS1

4420

)Td

cD(R

S123

55)

Citrob

acter

sp.3,

4

(C.f

reun

dii

CFN

IH1)

NA

NA

NA

PFL:

PflA

(RS1

4340

)

PDH

:A

ceEF

(RS1

0115

,RS

1012

0)

ALD

H:

RS16

255,

RS20

920

Ast

D(R

S162

55)

AD

H:

RS02

285,

RS04

980,

RS15

860,

RS16

775,

RS17

775

PTA

:Pt

a(R

S226

70)

ACK

:RS

2095

0,RS

2266

5Td

cD(R

S027

95)

Clostridium

sp.5,

6N

AN

AN

AN

AA

LDH

:A

dhE

AD

H:

Adh

EA

CK:

Ack

A(con

tinue

don

next

page

)

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

542

Page 6: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Table1

(con

tinue

d)

2,3-

BDO

Etha

nol

Ace

tate

(C.p

asteurianu

mA

TCC

6013

)PD

H:

NifJ

(RS0

2460

)

(RS0

7340

,RS

0275

0,RS

0758

0RS

1406

5,CA

_P01

62k ,

CA_P

0162

k )

(RS0

7340

,RS

0275

0,RS

0758

0RS

1406

5,CA

_P01

62k ,

CA_P

0035

k )

PTA

:Pt

a(R

S101

15)

(CLP

A_

RS10

110)

Lactob

acillus

sp.7,

8

(L.r

euteri

DSM

2001

6)

NA

NA

NA

NA

PDH

:RS

0341

0,RS

0341

5D

HLA

T:RS

0342

0

ALD

H:

RS00

180,

RS01

680

AD

Hg :

RS07

840,

RS08

065,

RS08

320,

RS09

040,

RS09

680

PTA

:RS

0207

5A

CK:

RS02

910,

RS09

035e

E.co

liN

AN

AN

APF

L:Pfl

B⁎

(b09

03),

TdcE

(b31

14)

PDH

(E1)

:A

ceE⁎

(b01

14)

DH

LAT

(E2)

:A

ceF⁎

(b01

15)

LAD

H(E

3):

Lpd⁎

(b01

16)

ALD

Hh :

Adh

E⁎(b

1241

)A

DH

i :A

dhE⁎

(b12

41),

Adh

P⁎(b

1478

)

PTA

:Pt

a⁎

(b22

97),

EutD

⁎(b

2458

)

ACK

:A

ckA

(b22

96)

1 K.pn

eumon

iae

andK.

oxytoc

a;2

locu

stag

prefi

xis

KPN

;34

locu

stag

prefi

xis

CFN

IH;5 C

.pasteurianu

man

dC.

acetob

utylicum

;6lo

cust

agpr

efix

isCL

PA;7 L

.reu

teri,

L.pa

nis,L.

collino

ides

,L.d

ioliv

oran

s,an

dL.

brevis;

8lo

cus

tag

prefi

xis

LREU

;ain

activ

ein

K.ox

ytoc

a;b

NA

D(P

)H-d

epen

dent

;cN

AD

(P)+

-dep

ende

nt;d

Pote

ntia

llyin

activ

epa

thw

ayin

som

est

rain

s;e

hom

olog

ueno

tpre

sent

inL.

panis;

fid

entifi

edpu

tativ

eLD

Hsw

ere

sele

cted

base

don

the

mos

tsig

nific

anth

omol

ogy

amon

gpu

tativ

eLD

Hsf

ound

inal

lspe

cies

incl

uded

(add

ition

alpu

tativ

eLD

Hsm

ayex

isti

nea

chsp

ecie

s);g

iden

tified

puta

tive

AD

Hsw

ere

sele

cted

base

don

the

mos

tsig

nific

anth

omol

ogy

amon

gpu

tativ

eA

DH

sfo

und

inal

lspe

cies

incl

uded

(add

ition

alpu

tativ

eA

DH

sm

ayex

isti

nea

chsp

ecie

s);h

addi

tiona

lAD

HE

may

exis

t;iad

ditio

nalA

DH

may

exis

t;jB 1

2-in

depe

nden

tGD

Htf

rom

C.bu

tyric

um;k

from

C.ac

etob

utylicum

ATC

C82

4;⁎

func

tion

ofen

zym

eex

peri

men

tally

dem

onst

rate

d.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

543

Page 7: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

et al., 2006). Also, under fermentative conditions, E. coli can recycle theexcess H2 formed via FHL by functional expression of native hydro-genase isoenzymes Hyd-1 and Hyd-2 (Murarka et al., 2008; Sawerset al., 1985; Sawers and Boxer, 1986). Compared to 1,3-PDO, theconversion of glycerol to 1,2-PDO is uncommon and less effective withsignificantly lower yields (Clomburg and Gonzalez, 2013; Zeng andSabra, 2011). In the oxidative branch, 1,2-PDO production starts withthe formation of DHAP via phosphorylation of DHA. Subsequently,DHAP enters the methylglyoxal (MG) pathway and is converted to theintermediate MG by a MG synthase (MGS; Fig. 1) (Hopper and Cooper,1972). The availability of DHAP is crucial since it represents a nodewhere carbon flux between glycolysis and 1,2-PDO formation diverges(Clomburg and Gonzalez, 2011). Conversion of MG to 1,2-PDO is pos-sible through either (i) the intermediate acetol via aldehyde oxidor-eductase (AOR), or (ii) the intermediate lactaldehyde via GlyDH or MGreductase (MGR; Fig. 1) (Clomburg and Gonzalez, 2011).

Under respiratory conditions, glycerol dissimilation can advancethrough two different two-step routes to DHAP (Fig. 1): (i) the aerobicglycerol kinase (GlpK)-glycerol-3-phosphate (G3P) dehydrogenase(GlpD) pathway; or (ii) the anaerobic GlpK-anaerobic G3P dehy-drogenase (GlpABC) pathway (Blankschien et al., 2010; Durnin et al.,2009). In either case, G3P is first synthesized via phosphorylation ofglycerol by GlpK at the expense of ATP (Fig. 1) (Murarka et al., 2008),or it can be directly imported into the cell via the G3P transporter, GlpT(encoded by glpT) (Jin et al., 1982). Through the aerobic pathway,GlpD, a plasma membrane-associated homodimer, generates DHAP viaoxidation of G3P. On the other hand, DHAP is synthesized from G3P byGlpABC (encoded by glpABC) anaerobically (Fig. 1). In addition tobeing a precursor to DHAP, G3P also serves as a precursor for lipidbiosynthesis or production of other metabolites (da Silva et al., 2009).Similar to the dha regulon in the fermentative pathway, the glp regulonin E. coli contains genes encoding enzymes found in the respiratorypathway (i.e. GlpF, GlpK, GlpD, and GlpABC). Similar glp regulons arealso present in Klebsiella sp. and Citrobacter sp. (da Silva et al., 2009),while the putative partial glp regulon of Lactobacillus sp. is inactive, due,in part, to the absence of DHAK (Morita et al., 2008).

3. Biological conversion of glycerol for the production of value-added chemicals in select microbial platforms

3.1. Klebsiella sp.

3.1.1. Natural metabolismKlebsiella sp. (i.e. K. pneumoniae) own natural pathways for flexible

and efficient metabolism of glycerol (Kumar and Park, 2017; Zenget al., 1993; Zhang et al., 2008; Zhang et al., 2009). Fermentativeglycerol dissimilation in Klebsiella sp. proceeds anaerobically throughcoupled oxidative and reductive branches (Fig. 1), providing outlets forenergy generation and redox balance. In the oxidative branch, glycerolis first converted to DHA by GlyDH (i.e. DhaD encoded by dhaD), andDHA is subsequently converted to DHAP by one of two DHAKs, i.e.DhaK I (encoded by dhaK) and DhaK II (encoded by dhaK123), withDhaK II playing the dominant role (Wang et al., 2003; Wei et al., 2014).Glycerol is simultaneously converted to 3-HPA by GDHt (i.e. DhaBencoded by dhaB123), followed by reduction of 3-HPA to 1,3-PDO,which is the major redox valve for glycerol dissimilation, via the NADH-dependent (DhaT encoded by dhaT) or putative NADPH-dependent(YqhD encoded by yqhD) 1,3-PDOORs (Ashok et al., 2011; Wang et al.,2003; Zeng et al., 1993). In Klebsiella sp., which synthesize B12 natu-rally, DhaB is B12-dependent and is activated by GdrAB (encoded bygdrAB), a reactivation factor that can reactivate glycerol- or oxygen-deactivated DhaB and protect it from substrate inhibition (Sun et al.,2003). Biosynthesis of 1,3-PDO can be limited by dhaB123 expression(Ahrens et al., 1998; Wang et al., 2003) given the detrimental effects of3-HPA accumulation (Wang et al., 2003). While a glp regulon is presentin K. pneumoniae, it is dormant in certain strains (Forage and Lin, 1982;

Wang et al., 2017b). Components of the oxidative and reductive bran-ches of the fermentative pathway of glycerol dissimilation (i.e. DhaD,DhaKII, DhaB, and DhaT) were expressed under aerobic conditions in aK. pneumoniae strain lacking a functional glp regulon, although theexpression levels were markedly lower compared to those observedunder anaerobic conditions (Forage and Lin, 1982). Conversely, gly-cerol kinase (i.e. GlpK encoded by glpK) and G3P dehydrogenase (GlpDencoded by glpD) were expressed under aerobic conditions in a strainpossessing a functional glp regulon, while the expression of DhaD,DhaKII, DhaB, and DhaT were highly repressed. Note that 1,3-PDOproduction under microaerobic conditions has been studied extensively(Chen et al., 2003a, 2003b; Chen et al., 2003a, 2003b; Liu et al., 2007;Mu et al., 2006; Xiu et al., 2007; Zhang and Xiu, 2009). Interestingly,the addition of exogenous cyclic adenosine monophosphate (cAMP),the inducer of the cAMP receptor protein (CRP) that relieves cataboliterepression in the absence of preferred carbon sources (Zheng et al.,2004), significantly increased the expression of DhaD, DhaKII, DhaB,and DhaT under aerobic conditions in either strain, although the foldincrease in expression was significantly higher in the strain possessing afunctional glp regulon (Forage and Lin, 1982). As the minor oxidizedproduct derived from the reductive branch of fermentative glyceroldissimilation, 3-HP is derived from 3-HPA by the NAD+-dependent(YdcW encoded by ydcW) and NADP+-dependent (YneI encoded byyneI) ALDHs (Fig. 1) (Luo et al., 2013; Luo et al., 2011a, 2011b). 3-HPproduction can provide compensatory NADH for 1,3-PDO synthesis(Zhu et al., 2015), and alleviate the accumulation of the toxic inter-mediate 3-HPA, which is also known to inhibit DhaT in the presence ofexcess glycerol (Hao et al., 2008a; Sun et al., 2008). Moreover, theexpression of other ALDHs, including PuuC (Ashok et al., 2013a; Ashoket al., 2011; Huang et al., 2012; Luo et al., 2011a, 2011b), FeaB (Huanget al., 2012), BadH (Huang et al., 2012), and GabD (Huang et al., 2012)(encoded by puuC, feaB, badH, and gabD, respectively), can potentiallyenhance 3-HP production in K. pneumoniae. Finally, alternate pathwaysderived from the propanediol utilization (PDU) system exist for theconversion of glycerol to 1,3-PDO and 3-HP in Klebsiella sp. (Bobiket al., 1999; Cho et al., 2015; Luo et al., 2012), as is the case in Sal-monella sp. (Bobik et al., 1999) and Lactobacillus sp. (Makarova et al.,2006; Morita et al., 2008; Sauvageot et al., 2002; Sriramulu et al.,2008). In this system, glycerol is first converted to 3-HPA by the B12-dependent diol dehydratase (DODHt) PduCDE (encoded by pduCDE)(Bobik et al., 1999; Cho et al., 2015; Morita et al., 2008; Sauvageotet al., 2002) with the reactivation factor PduGH (encoded by pduGH)(Bobik et al., 1999), followed by conversion of 3-HPA to 1,3-PDO via a1,3-PDOOR (e.g. DhaT). Alternatively, 3-HPA can be converted to 3-HP-CoA via PduP, a CoA-dependent propionaldehyde dehydrogenase en-coded by pduP, followed by phosphorylation of 3-HP-CoA to yield 3-HP-P via a phosphate propanoyltransferase PduL (encoded by pduL), andfinal conversion of 3-HP-P to 3-HP via a propionate kinase PduW (en-coded by pduW; Fig. 1) (Bobik et al., 1999; Honjo et al., 2015; Luo et al.,2012).

During anaerobic growth, pyruvate metabolism is mediated by boththe pyruvate formate lyase (PFL) and pyruvate dehydrogenase (PDH)complexes, with both enzyme systems presenting similar activities,while pyruvate:ferrodoxin oxidoreductase (PFOR) does not appear to beactive (Menzel et al., 1997) except under conditions of glycerol excess(Zeng et al., 1993). The activity of PFL was observed to increase anddecrease under glycerol limitation and excess, respectively (Menzelet al., 1997). Conversely, the activity of the PDH complex increaseswith glycerol concentration (Menzel et al., 1997). The phenomena ofoscillation of biomass concentration and glycerol consumption in con-tinuous anaerobic cultures of K. pneumoniae has been, in part, attrib-uted to the feedback control between the parallel pyruvate metabolismpathways for the production of acetate, CO2, 2,3-BDO, lactate, andethanol (Menzel et al., 1996). Interestingly, the activity of PFL, which isthe presumptive major player in fermentative metabolism of pyruvate,is strongly affected by oscillations in anaerobic continuous cultures,

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

544

Page 8: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

while the activities of pyruvate kinase (PYK) and the PDH complexwere marginally influenced (Menzel et al., 1998). Finally pyruvateoxidase (PO encoded by poxB) appears to play a significant role in bothaerobic and anaerobic pyruvate metabolism, as mutation of poxB re-sulted in excessive acetate accumulation and severely inhibited cellgrowth during aerobic cultivation, and abolished CO2 production underanaerobic conditions (Lin et al., 2016).

Other major fermentative products produced by Klebsiella sp. in-clude 2,3-BDO, acetate, lactate, succinate, and ethanol (Xu et al., 2009;Zeng et al., 1996). A substantial amount of NADH released through theactivities of DhaD and GlpD is oxidized and discharged as H2 throughpathways other than the 1,3-PDO, ethanol, lactate, and 2,3-BDO path-ways. When glycerol is in excess, the bulk of reducing equivalents areconsumed for 1,3-PDO production with a small amount being releasedas H2. Under glycerol limitation, ethanol production is more prevalentas it is energetically favorable, although ethanol is more toxic to Kleb-siella sp. than 1,3-PDO (Zeng et al., 1993).

3.1.2. Strain engineeringTable 2 summarizes various strategies for strain engineering of

Klebsiella sp. to convert glycerol to value-added products, while detailedtechnical discussion is provided in the following sections.

3.1.2.1. 1,3-PDO. To date, 1,3-PDO has been the most common targetmetabolite in engineered Klebsiella sp. The formation of 1,3-PDOsimultaneously generates 1 mol NADH per mol glycerol consumed asthe major redox valve for oxidative glycerol metabolism via theglycolytic pathway. Major strain engineering strategies to enhance1,3-PDO formation include: (1) disruption of the oxidative branch ofthe fermentative glycerol dissimilation pathway; (2) overexpression ofgenes in the reductive and/or oxidative branch of the fermentativeglycerol dissimilation pathway; and (3) inactivation of pathwaysleading to side metabolites, such as lactate, 2,3-BDO, ethanol,succinate, and acetate. Note that these strategies have been applied tobalance the NAD+/NADH ratio, increase carbon flux toward 1,3-PDO,and/or reduce toxic metabolite formation. The oxidative branch offermentative glycerol dissimilation produces DHAP, which issubsequently converted to the glycolytic intermediate glyceraldehyde3-phosphate. Glycolysis results in the synthesis of phosphoenolpyruvate(PEP), leading to the formation of succinate or pyruvate, the latter ofwhich is an intermediate in 2,3-BDO, lactate, acetate, and ethanolformation (Fig. 1). To inactivate oxidative glycerol metabolism in K.pneumoniae, dhaD and dhaK were mutated with the simultaneous use ofthe lacZ promoter from E. coli to regulate the expression of dhaB123, ordhaR, encoding the transcription factor that activates the expression ofthe dha regulon (DhaR), was mutated (Seo et al., 2009). In eitherapproach, the reductive branch of fermentative glycerol dissimilationwas also inactivated, necessitating the expression of native dhaT andgdrAB to restore 1,3-PDO production. Side metabolite formation wasmarkedly reduced, although 1,3-PDO production was not enhanced.Mutation of dhaD and dhaK with coexpression of native dhaB123 anddhaT resulted in abolished production of 2,3-BDO, lactate, and ethanol,but the 1,3-PDO titer and glycerol consumption were reduced relativeto the wild-type strain (Horng et al., 2010). In the absence of afunctional glp regulon for respiratory dissimilation of glycerol, theoxidative branch is the major pathway for fermentative glycerolmetabolism in K. pneumoniae, such that inactivation of the dharegulon potentially hinders glycerol dissimilation and reduces NADHlevels. Activating the glp regulon in K. pneumoniae via overexpression ofnative glpK significantly altered cellular metabolism with acetoinproduction being favored over reductive glycerol metabolism (Wanget al., 2017b), while reduced expression of glpK may mildly stimulateglycerol dissimilation and improve 1,3-PDO production under suitableaeration conditions. In parallel, expression of formate dehydrogenase(FDH) and/or pyridine nucleotide transhydrogenase (UDH) canpotentially regenerate NADH to restore the redox balance during 1,3-

PDO production. On the other hand, co-fermentation of glycerol andglucose is an alternate strategy to enhance 1,3-PDO production in theΔdha background, and can be achieved via mutation of crr, encodingglucose phosphotransferase protein (EIIAGlc) (Oh et al., 2013a; Wanget al., 2017c). Co-feeding limiting amounts of glucose with excessglycerol may stimulate biomass accumulation and NADH regenerationvia glucose metabolism, while glycerol can be converted to 1,3-PDO inthe absence of oxidative glycerol metabolism.

The expression of genes in the reductive branch of the fermentativeglycerol dissimilation pathway can increase carbon flux toward 1,3-PDO production with reduced 3-HPA accumulation, while expression ofgenes in the oxidative branch can improve glycerol dissimilation andNADH levels albeit at the expense of potential side metabolite forma-tion. To alleviate growth cessation associated with toxic 3-HPA, nativedhaT was overexpressed (Hao et al., 2008b). While cell growth con-tinued until glycerol was exhausted with reduced levels of all major sidemetabolites, 1,3-PDO production was not improved (Hao et al., 2008b),even in fedbatch cultures where a decrease in the growth rate wasobserved, eventually leading to growth stagnation (Zhao et al., 2009).Similarly, overexpression of native dhaD reduced ethanol and 2,3-BDOaccumulation without improving 1,3-PDO formation, suggesting thatthe redox balance in the cell is optimal for natural 1,3-PDO synthesis(Zhao et al., 2009). On the other hand, coexpression of native dhaD anddhaT significantly reduced 3-HPA accumulation, leading to a sub-stantial increase in 1,3-PDO production (Chen et al., 2009).

Coexpression of a putative 1,3-PDOOR bearing similarity to YqhDfrom E. coli and native gdrAB significantly increased the 1,3-PDO titer,relative to coexpression of native dhaT and gdrAB, in a strain with theoxidative branch of fermentative glycerol metabolism being disrupted(Seo et al., 2010). The activity of YqhD is NADPH-dependent, and, incontrast to DhaT, the overexpression of the putative 1,3-PDOOR couldpotentially mitigate the NAD+/NADH imbalance resulting from dis-ruption of the oxidative branch of fermentative glycerol metabolism.However, glycerol consumption and 1,3-PDO production rates weresignificantly lower relative to the parent strain coexpressing yqhD andgdrAB, in which the oxidative branch was intact, although the final 1,3-PDO titer was comparable (Seo et al., 2010). In addition, coexpressionof native dhaT and yqhD from E. coli reduced side metabolite formation,but provided only a marginal increase in the 1,3-PDO titer (Zhuge et al.,2010), while yqhD expression alone reduced the rate of glycerol dis-similation without affecting the 1,3-PDO titer (Oh et al., 2013b). Ex-pression of yqhD facilitated high 1,3-PDO titers in E. coli (Emptageet al., 2003a), although it seems to play a minor role in reductive gly-cerol metabolism in Klebsiella sp., and disruption of native dhaT nearlyabolishes 1,3-PDO production (Seo et al., 2009). Finally, basal expres-sion of native puuC from the leaky tac promoter increased and de-creased 1,3-PDO titers in cultures of lactate and lactate/2,3-BDO defi-cient mutants, respectively (Zhu et al., 2015). The conversion of 3-HPAto 3-HP (via PuuC) reduces 3-HPA levels, while producing NADH re-quired for 1,3-PDO production. The decrease in 1,3-PDO formation inthe lactate/2,3-BDO deficient strain was the result of pyruvate accu-mulation, which can interfere with glycerol dissimilation in 2,3-BDOdeficient mutants (Lee et al., 2014).

The major side metabolites produced during glycerol fermentationin Klebsiella sp. are lactate, 2,3-BDO, ethanol, succinate, and acetate,and inactivation of these pathways can potentially enhance 1,3-PDOproduction. Formation of acetate and 2,3-BDO during glycerol fer-mentation for 1,3-PDO production results in a surplus of 2 mol and 0.5mol NADH, respectively, per mol glycerol consumed, while formationof lactate and ethanol is redox balanced and formation of succinategenerates a deficit of 1 mol NADH per mol glycerol consumed. Whilethe formation of each metabolite, which is accompanied by ATP pro-duction for cell growth and maintenance, can relieve pyruvate accu-mulation resulting from the compromised tricarboxylic acid (TCA)cycle in Klebsiella sp. (Cabelli, 1955), reduced carbon flux toward 1,3-PDO formation may occur. Inactivating the ethanol pathway in K.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

545

Page 9: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Table 2Summary of literature for Klebsiella sp. strains engineered for value-added chemical production from glycerol.

Parent strain Product Genetic strategy Titer (g/L) Cultivationmode

Reference

Kp ZG25 1,3-PDO Mutate budC to reduce 2,3-BDO formation 67 fedbatch (Guo et al., 2013)Kp TUAC01 1,3-PDO Express dhaT 20 batch (Hao et al., 2008b)Kp KG2 (lacks ldhA) 1,3-PDO Mutate poxB and pta-ackA to reduce acetate formation 77 fedbatch (Lin et al., 2016)Kp J2B ΔldhA 1,3-PDO Mutate budA, budB, budC, or all three to reduce 2,3-BDO formation 31 fedbatch (Kumar et al., 2016)Kp KG1 1,3-PDO

Express dhaD and dhaT to elucidate their roles during microaerobiccultivation

45 fedbatch (Zhao et al., 2009)

Kp AK (lacks oxidativepathway)

1,3-PDOExpress dhaB123, dhaT, or yqhD, or dhaT and yqhD, or dhaB123 anddhaT (along with gdrAB)

26 fedbatch (Luo et al., 2013)

Kp Cu 1,3-PDO -Mutate dhaD and dhaK while inserting the lacZ promoter (E. coli)upstream of dhaB123 to interrupt oxidative metabolism withoutdisturbing the reductive branch-Mutate dhaR to interrupt oxidative metabolism-Coexpress dhaT and gdrAB

5 flask (Seo et al., 2009)

Kp ΔldhA 1,3-PDO-Mutate budB and pta to reduce 2,3-BDO and acetate formation,respectively-Coexpress pdc and aldB (Z. mobilis) to stimulate conversion ofpyruvate to ethanol to restore metabolism in ΔbudB mutant

68 fedbatch (Lee et al., 2014)

Kp CICIM B0057 1,3-PDO Express yqhD (E. coli) and/or dhaT 18 batch (Zhuge et al., 2010)Ko M5aI 1,3-PDO Mutate ldhA to reduce lactate formation 63 fedbatch (Yang et al., 2007)Kp AK and Cu 1,3-PDO Express yqhD to increase NADH levels 8 flask (Seo et al., 2010)Kp WT 1,3-PDO -Mutate budC to reduce 2,3-BDO formation

-Chromosomally express fdh (P. pastoris) to regenerate NADH72 fedbatch (Wu et al., 2013)

Kp KG2 and KG3 1,3-PDOExpress puuC via leakage from Ptac promoter to enhance 3-HP andNADH formation

69 fedbatch (Zhu et al., 2015)

Kp YMU2 1,3-PDO Mutate aldA to reduce ethanol formation 71 fedbatch (Zhang et al., 2006a,2006b)

Kp WT 1,3-PDO -Mutate dhaD and dhaK to interrupt oxidative metabolism-Coexpress dhaT and dhaB123

8 flask (Horng et al., 2010)

Ko M5aI 1,3-PDO Mutate budA to reduce 2,3-BDO formation 27 fedbatch (Zhang et al., 2012)Kp HR526 1,3-PDO Mutate ldhA to reduce lactate formation 102 fedbatch (Xu et al., 2009)Kp Cu 1,3-PDO -Mutate budB and ldhA to reduce 2,3-BDO and lactate formation,

respectively103 fedbatch (Oh et al., 2012b)

Kp KG1 1,3-PDO Coexpress dhaD and dhaT 17 flask (Zhao et al., 2009)Kp ACCC 10082 1,3-PDO Coexpress dhaD and dhaT 59 fedbatch (Chen et al., 2009)Kp 2-1 1,3-PDO -Mutate ldhA and aldH to reduce lactate and ethanol formation,

respectively88 fedbatch (Chen et al., 2016)

Kp CICIM B0057 1,3-PDOReduce 2,3-BDO formation by reducing expression of budA, budB,and budC via asRNA

22 flask (Lu et al., 2016)

Kp CGMCC 1.6366 1,3-PDO-Mutate pflB or acoABCD to reduce acetate formation

16 batch (Zhou et al., 2017a,2017b)

Kp KG2 1,3-PDO -Mutate pck or ppc to reduce succinate formation 72 fedbatch (Zhang et al., 2017)Kp KCTC 2242 1,3-PDO -Introduce point mutations in DhaT via error prone PCR to increase

activity-Mutate crr to relieve catabolite repression for glycerol/glucose co-fermentation to increase biomass and NADH levels-Overexpress glpF to enhance glycerol uptake-Coexpress fdh (P. pastoris), gdh (B. subtilis), and udh to regenerateNADH

86 fedbatch (Wang et al., 2017c)

Kp ATCC 200721ΔldhA

1,3-PDO-Mutate budA, budB, or budC to reduce 2,3-BDO formation

49 fedbatch (Oh et al., 2018)

Kp WT 1,3-PDO-Mutate arcA to increase expression of enzymes in the TCA cycle-Mutate ldhA to reduce lactate formation-Mutate crr to relieve catabolite repression for glycerol/glucose co-fermentation to increase biomass and NADH levels

78 fedbatch (Lu et al., 2018)

Kp KCTC 2242ΔwabG

1,3-PDO -Mutate ldhA, pflB, and budA to reduce lactate, acetate, and 2,3-BDOformation, respectively-Mutate dhaD and glpK to prevent glycerol assimilation into biomass-Modify the 5’-UTR of mtlA to improve the efficiency of mannitolusage-Mutate dhaK123 and overexpress dhaK2 to enhance expression ofdhaT

21 batch (Lee et al., 2018)

Ko PDL-0(WT) 1,3-PDO/lactate

-Mutate budA and budB, adhE, and ackA-pta to reduce 2,3-BDO,ethanol, and acetate formation, respectively-Replace ldhD with ldhL (L. casei) to facilitate L-lactate production

70 / 100 (D-or L-lactate)

fedbatch (Xin et al., 2017)

Kp KCTC12133BP 1,3-PDO/2,3-BDO -Mutate ldhA and mdh to reduce lactate and succinate formation,

respectively

125 fedbatch (Park et al., 2017)

(continued on next page)

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

546

Page 10: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Table 2 (continued)

Parent strain Product Genetic strategy Titer (g/L) Cultivationmode

Reference

Kp KCTC 2242 ΔldhA ΔptsG 1,3-PDO/2,3-BDO -Construct EDP from Z. mobilis and coexpress udh to regenerate

NADH-Coexpress dhaT

79 / 32 fedbatch (Wang et al., 2017d)

Kp DSMZ 2026 3-HP/1,3-PDO Express puuC and mutate dhaT to balance 3-HP and 1,3-PDO

production

16 / 17 fedbatch (Ashok et al., 2011)

Kp WM3 1,3-PDO/3-HP Express ALDHs from Zymomonas mobilis (adhB), Lactobacillus

collinoides (aldH and pduQ), K. pneumoniae (aldA, aldB, puuC, ydcW,etuE, feaB, gabD and badH), or E. coli (aldA, aldH and ycdW)

24 / 49 fedbatch (Huang et al., 2012)

Kp J2B 1,3-PDO/3-HP

Express KGSADH (A. brasilense) 16 / 12 fedbatch (Kumar et al., 2012)

Kp J2B 1,3-PDO/3-HP

-Express KGSADH (A. brasilense)-Mutate ldhA to reduce lactate formation

23 / 23 fedbatch (Kumar et al., 2013)

Kp DSMZ 2026 1,3-PDO/3-HP

Express dhaS (B. subtilis) 18 / 27 fedbatch (Su et al., 2015)

Kp AK 1,3-PDO/3-HP

Express aldH, dhaT amd gdrAB in mutant defective in oxidativeglycerol metabolism

23 / 7 fedbatch (Luo et al., 2011a,2011b)

Kp J2B 1,3-PDO/3-HP - Mutate ldhA, adhE, frdA, and pta-ackA to reduce lactate, ethanol,

succinate, and acetate formation, respectively-Mutate glpK or dhaD to reduce acetate without disrupting pta-ackAto improve glycerol consumption-Overexpress dhaT, or dhaB123 and gdrAB

21 / 43 fedbatch (Ko et al., 2017)

Kp J2B 3-HP-Mutate dhaT and yqhD to reduce 1,3-PDO formation-Express puuC, KGSADH (A. brasilense) or aldH (E. coli)

16 fedbatch (Ko et al., 2012)

Kp J2B ΔdhaT ΔyqhD 3-HP Mutate ahpF and adhE in a ΔdhaT ΔyqhD double mutant expressingKGSADH (A. brasilense) to reduce 1,3-PDO formation

9 fedbatch (Ko et al., 2015)

Kp DSMZ 2026 3-HP-Mutate glpK to force glycerol through the fermentative pathway inthe presence of nitrate,-Mutate dhaT to reduce 1,3-PDO formation-Overexpress puuC

23 fedbatch (Ashok et al., 2013a)

Kp DSMZ 2026 3-HP-Optimize puuC expression through promoter selection (Ptac)-Mutate ldh1 and ldh2, and pta to reduce lactate and acetateformation, respectively

84 fedbatch (Li, Y. et al., 2016)

Kp DSMZ 2026 3-HP-Coexpress dhaB123, gdrAB and puuC in a ΔdhaT ΔyqhD doublemutant

28 fedbatch (Ashok et al., 2013b)

Ko M1 2,3-BDO-Mutate pduC and ldhA to reduce 1,3-PDO and lactate formation,respectively

132 fedbatch (Cho et al., 2015)

Kp WT 1-butanol -Express ter (T. denticola), and bdhA and bdhB (C. acetobutylicum)-Express kivd (L. lactis) to facilitate the 2-keto acid pathway-Reduce 1,3-PDO formation by reducing expression of dhaB123,dhaT, and/or gdrAB via asRNA-Reduce 2,3-BDO formation by reducing expression of budB and/orbudC via asRNA

0.029 flask (Wang et al., 2014a,2014b, 2014c)

Kp WT 1-butanol-Express kivd (L. lactis), leuABCD, and adhe1 (C. acetobutylicum)-Enhance NADH regeneration via expression of fdh (P. pastoris), gdh,or udh-Reduce 1,3-PDO formation by reducing expression of dhaB123,dhaT, and/or gdrAB via asRNA

0.1 flask (Wang et al., 2015a,2015b)

Kp 13 1-butanol-Express kivd (L. lactis), leuABCD, and adhe1 (C. acetobutylicum)-Repress expression of metA, ilvI, ilvB, and alaA via CRISPRi toincrease carbon flux toward 1-butanol production

0.2 flask (Wang et al., 2017a)

Kp ATCC 200721 ΔldhA 2-butanol-Express ilvIH, ilvD, andkivd and adhA (L. lactis)-Mutate budA to reduce 2,3-BDO formation

0.3 batch (Oh et al., 2014)

GEM167 ethanol-Mutate ldhA to reduce lactate formation-Coexpress pdc and adhII (Z. mobilis)

31 fedbatch (Oh et al., 2012a)

ATCC 25955 D-lactate-Overexpress ldhA-Mutate dhaT and yqhD to reduce 1,3-PDO formation

142 fedbatch (Feng et al., 2014)

ATCC 25955 acetoin -Mutate budC, gldA and dhaD to prevent conversion of acetoin to 2,3-BDO

32 fedbatch (Wang et al., 2017b)

(continued on next page)

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

547

Page 11: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

pneumoniae via mutation of aldA (encoding ALDH; AldA) drasticallyreduced ethanol formation, resulting in significant increases in the yieldand titer of 1,3-PDO in a fedbatch fermentation, despite decreasedglycerol consumption and biomass accumulation. (Zhang et al., 2006a,2006b). On the other hand, mutation of aldH (encoding ALDH; AldH)and ldhA (encoding lactate dehydrogenase; LdhA) abolished ethanoland lactate formation, respectively, although the 1,3-PDO titer wasmarginally improved (Chen et al., 2016). Similarly, inactivation of ldhAsignificantly increased and decreased NADH and lactate levels, re-spectively, leading to substantial 2,3-BDO and ethanol formation, andthe 1,3-PDO titer increased slightly in a fedbatch fermentation (Xuet al., 2009). In contrast to K. pneumoniae, mutation of ldhA in Klebsiellaoxytoca slightly enhanced glycerol dissimilation, abolished lactateproduction, and significantly increased the 1,3-PDO titer, however, 2,3-BDO formation markedly increased (Yang et al., 2007).

Elevated 2,3-BDO levels can be particularly problematic since itsboiling point is similar to that of 1,3-PDO, making downstream pur-ification more difficult (Xiu and Zeng, 2008), and considerable efforthas been made to alleviate 2,3-BDO production. Mutation of ldhA andbudB, encoding α-acetolactate synthase (BudB), significantly decreasedlactate and 2,3-BDO formation, respectively, although the 1,3-PDO titerwas also reduced (Oh et al., 2012b). Inactivation of budC, encoding 2,3-BDO dehydrogenase/acetoin reductase (BudC), moderately decreasedand increased 2,3-BDO and 1,3-PDO production, respectively (Guoet al., 2013). The moderate reduction in 2,3-BDO production in culturesof the ΔbudC mutant is likely due to the promiscuity of DhaD, which cancatalyze the interconversion of acetoin and 2,3-BDO in response toelevated intracellular NADH levels during glycerol catabolism (Wanget al., 2014a, 2014b, 2014c), while an alternate/complementary ex-planation is the presence of the 2,3-BDO replenishment cycle in K.pneumoniae (Syu, 2001). Mutation of budA, encoding α-acetolactatedecarboxylase (BudA), in K. oxytoca abolished 2,3-BDO formation andmoderately increased the 1,3-PDO titer, although the acetate and lac-tate titers increased substantially (Zhang et al., 2012). Similarly, mu-tation of budA or budB in a K. pneumoniae ΔldhA mutant abolished 2,3-BDO formation, although the ΔldhA ΔbudA double mutant producedsignificantly less 1,3-PDO than the parent strain (Kumar et al., 2016).

Moreover, mutation of budC modestly increased and decreased the 1,3-PDO and 2,3-BDO titers, although mutation of the entire bud operon(i.e. budABC) resulted in the lowest 1,3-PDO titer. While the cultureperformance of the ΔbudABC mutant was significantly improved, i.e.production of the target metabolite (1,3-PDO) increased, by increasingthe buffering capacity of the medium in a shake flask, the 1,3-PDO titerdeclined significantly due to excessive pyruvate accumulation during afedbatch fermentation (Kumar et al., 2016), indicating inefficient op-eration of the TCA cycle. Similarly, reduced expression of budA andbudB via antisense RNA (asRNA)-mediated transcriptional interferencesignificantly decreased 2,3-BDO production without substantially af-fecting the 1,3-PDO titer, and repression of budC expression did notappreciably affect culture performance (Lu et al., 2016). In contrast,mutation of budA, budB, or budC in a ∆ldhA mutant reduced 2,3-BDOand 1,3-PDO production, with the ΔldhA ΔbudA double mutant pro-ducing significantly less 1,3-PDO than all other strains (Oh et al., 2018).To reduce pyruvate accumulation through increased ethanol synthesis,pdc and aldB, encoding pyruvate decarboxylase (PDC) and ALDH (AldB)of Zymomonas mobilis, respectively, were coexpressed in a ΔldhA ΔbudBdouble mutant (Oh et al., 2012b), resulting in restored glycerol dis-similation and 1,3-PDO production. These studies underscore the deli-cate balance between 1,3-PDO and 2,3-BDO production in Klebsiella sp.,and further suggest the unstable nature of the TCA cycle, which is acritical issue to be addressed.

Inactivating acetate pathways to enhance 1,3-PDO production hasalso been explored. Mutation of pta (encoding phosphotransacetylase;PTA) and ackA (encoding acetate kinase; AckA) restored cell growth,reduced acetate production, and modestly increased the 1,3-PDO titerunder the ΔldhA ΔpoxB background (Lin et al., 2016). Similarly, mu-tation of pflB, encoding PFL (PflB), decreased acetate production butdid not significantly improve the 1,3-PDO titer, while mutation ofacoABCD, encoding the PDH complex (AcoABCD), increased the pro-ductivity but decreased the yield of 1,3-PDO due to significantly in-creased formate and ethanol formation (Zhou et al., 2017a, 2017b).Finally, in an attempt to alleviate succinate formation, pck (encodingPEP carboxykinase; PCK) was inactivated, resulting in improved growthand reduced acetate and 2,3-BDO formation, albeit without significant

Table 2 (continued)

Parent strain Product Genetic strategy Titer (g/L) Cultivationmode

Reference

-Express glpK to activate silent glp operon-Express nox (L. brevis) to increase NAD+/NADH ratio

ATCC 25955 P(3HP) -Express dhaB123, gdrAB, and aldH (E. coli) to enhance 3-HPformation-Express prpE and phaC to convert 3-HP to P(3HP)-Mutate dhaT and yqhD to reduce 1,3-PDO formation

0.2 flask (Feng et al., 2015)

ackA, acetate kinase; acoABCD, pyruvate dehydrogenase complex; adhII, aldehyde/alcohol dehydrogenase; adhA, alcohol dehydrogenase; aldB, aldehyde dehy-drogenase; adhE, bifunctional acetaldehyde-CoA/alcohol dehydrogenase; alaA, alanine transaminase; ahpF, alkyl hydroperoxide oxidoreductase; aldH, aldehydedehydrogenase; bdhA, butanol dehydrogenase A; bdhB, butanol dehydrogenase B; budA, α-acetolactate decarboxylase; budB, α-acetolactate synthase; budC, 2,3-BDOdehydrogenase/acetoin reductase; dhaB123, glycerol dehydratase; crr, glucose phosphotransferase protein; dhaD, glycerol dehydrogenase; dhaK, dihydroxyacetonekinase; dhaR, transcriptional activator of the dha regulon; dhaS, aldehyde dehydrogenase; dhaT, NADH-dependent 1,3-PDO oxidoreductase; fdh, formate dehy-drogenase; frdA, succinate dehydrogenase; gdh, glucose dehydrogenase (Bacillus subtilis); gdrAB, glycerol dehydratase reactivation factor; gldA, glycerol dehy-drogenase; glpF, glycerol transporter; glpK, glycerol kinase; ilvB, acetolactate synthase isozyme 1 large subunit; ilvD, dihydroxyacid dehydratase; ilvI, acetolactatesynthase isozyme 3 large subunit; ilvIH, keto-acid reducto-isomerase; KGSADH, α-ketoglutaric semialdehyde dehydrogenase; kivD, α-ketoisovalerate decarboxylase;lacZ, β-galactosidase; ldh1, lactate dehydrogenase; ldh2, lactate dehydrogenase; ldhA, D-lactate dehydrogenase; ldhD, D-lactate dehydrogenase; ldhL, L-lactate de-hydrogenase; leuA, 2-isopropylmalate synthase; leuB, 3-isopropylmalate dehydrogenase; leuCD, 3-isopropylmalate dehydratase subunits C and D; mdh, malate de-hydrogenase; metA, pck, carboxykinase; nox, NADH oxidase; pdc, pyruvate decarboxylase; homoserine O-succinyltransferase; pduC, diol dehydratase large subunit;pflB, pyruvate formate lyase; phaC, polyhydroxyalkanoate synthase; poxB, pyruvate oxidase; ppc, phosphoenolpyruvate carboxylase; prpE, propionyl-CoA synthetase;pta, phosphotransacetylase; puuC, aldehyde dehydrogenase; ter, trans-2-enoyl-CoA reductase; udh, pyridine nucleotide transhydrogenase; yqhD, NADPH-dependent1,3-PDO oxidoreductase; other terms: ALDH, aldehyde dehydrogenase; 1,3-PDO, 1,3-propanediol; 2,3-BDO, 2,3-butanediol; 3-HP, 3-hydroxypropionic acid; asRNA,antisense RNA; CRISPRi, Clustered Regularly Interspaced Palindromic Repeats interference; EDP, Entner-Doudoroff pathway; Ko, Klebsiella oxytoca; Kp, Klebsiellapneumoniae; NADH, nicotinamide adenine dinucleotide; P(3HP), poly(3-hydroxypropionate); PCR, polymerase chain reaction.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

548

Page 12: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

changes in 1,3-PDO or succinate production, and mutation of ppc (en-coding PEP carboxylase; PPC) did not affect culture performance(Zhang et al., 2017). A considerable body of literature suggests thatcompletely abolishing side metabolite formation is not feasible inKlebsiella sp., potentially due to their incomplete TCA cycle that canlead to accumulation of inhibitory levels of pyruvate (Kumar et al.,2016). Prior work indicates that the flavin adenine dinucleotide (FAD)binding subunit of the succinate:quinone oxidoreductase complex, en-coded by sdhA (SdhA), and fumarase A, encoded by fumA (FumA),which successively convert succinate to fumarate and fumarate tomalate in the TCA cycle, are weakly expressed in K. pneumoniae(Cabelli, 1955; Kumar et al., 2016). Fumarase A and fumarase B (en-coded by fumB; FumB) contain [4Fe-4S] clusters that are sensitive tooxidative damage, while fumarase C (encoded by fumC; FumC) is in-sensitive to oxygen (Liochev and Fridovich, 1993). Accordingly, ex-pression of the appropriate fumarase isozyme from E. coli, selectedbased on the aeration conditions employed for glycerol fermentation,may activate the TCA cycle in Klebsiella sp., in turn, reducing sidemetabolite formation and promoting glycerol dissimilation. Similarly,expression of the succinate:quinone oxidoreductase complex, encodedby sdhABCDE (SdhABCDE), may also improve culture performance. Itmay be necessary to perform more extensive transcriptional analysis ofgenes encoding enzymes comprising the TCA cycle during glycerolfermentation to identify other gene targets in Klebsiella sp. for metabolicmanipulation.

Coproduction of 1,3-PDO with a second major metabolite is an al-ternate approach to completely eliminating side metabolite production.1,3-PDO/lactate co-production is a redox balanced process that canresult in high yields of both metabolites in K. oxytoca (Xin et al., 2017).2,3-BDO, ethanol, and acetate formation were either abolished or sig-nificantly reduced via mutation of budA and budB, adhE (encoding al-dehyde/alcohol dehydrogenase; AdhE), and ackA-pta, respectively,leading to a high 1,3-PDO/D-lactate yield of 0.95 mol/molglycerol.Moreover, optically pure L-lactate was co-produced with 1,3-PDO at asimilar yield by replacing the native D-LDH (encoded by ldhD) with L-LDH from Lactobacillus casei (encoded by ldhL) (Xin et al., 2017). High-level 1,3-PDO/2,3-BDO co-production from glycerol was recently de-monstrated in K. pneumoniae. 1,3-PDO/2,3-BDO co-production wassignificantly improved in K. pneumoniae via mutation of ldhA, whilemutation of mdh, encoding malate dehydrogenase (MDH), significantlyreduced succinate formation in the ΔldhA mutant with only a modestimprovement in the 1,3-PDO/2,3-BDO yield (Park et al., 2017). Con-struction of the Entner-Doudoroff pathway (EDP) from Z. mobilis in K.pneumoniae with coexpression of native dhaT and udh (encoding UDH)significantly improved 1,3-PDO/2,3-BDO co-production during gly-cerol/glucose co-fermentation (Wang et al., 2017d). This strategyaimed to increase NADPH levels via glucose metabolism through theEDP from Z. mobilis with subsequent regeneration of NADH fromNADPH via native UDH (Wang et al., 2017d). However, 1,3-PDO/2,3-BDO co-production may be practically infeasible due to the difficultyassociated with separating these compounds, and carbon loss is sig-nificant during 2,3-BDO production due to the successive decarbox-ylations from pyruvate to acetoin (Fig. 1).

Other strain engineering strategies to improve 1,3-PDO productionfrom glycerol include protein engineering of key enzymes, globallyupregulating the expression of enzymes in the TCA cycle, relievingcatabolite repression to enable co-fermentation of glycerol and glucose,the expression of glycerol transporters to enhance glycerol uptake, andthe expression of enzymes that can regenerate NADH. Transformationof integration cassettes generated with error prone PCR resulted inpoint mutations to DhaT (D41G and H57Y) that moderately increasedits activity and improved 1,3-PDO production (Wang et al., 2017c).Creating large mutant libraries via error prone PCR, or rational proteinengineering may be effective strategies to further enhance the activityof DhaT, or to relieve the inhibition of DhaT by 3-HPA. The ArcA-ArcBtwo component system negatively regulates the expression of genes

encoding enzymes of the TCA cycle in response to decreasing levels ofreduced quinones (Bekker et al., 2010; Shalel-Levanon et al., 2005).Mutation of arcA in K. pneumoniae resulted in a ≥9-fold increase in thetranscription of gltA (encoding citrate synthase; GltA), icd (encodingisocitrate dehydrogenase; ICD), sdhC, and sucC (encoding succinyl-CoAsynthetase β subunit; SucC), resulting in a significant decrease inacetate accumulation and the intracellular NAD+/NADH ratio, and anincrease in the specific 1,3-PDO titer (Lu et al., 2018). However, lactateproduction increased substantially, which was resolved by mutatingldhA in the ΔarcA mutant, resulting in complete cessation of lactateaccumulation and a slight increase in the 1,3-PDO titer. In the presenceof glucose or other preferred carbon sources, EIIAGlc is in the depho-sphorylated state and will bind to GlpK to inhibit glycerol uptake(Deutscher et al., 2006). Accordingly, mutation of crr in the ΔarcAΔldhA double mutant enabled co-fermentation of glucose and glycerol,resulting in a moderate increase in the 1,3-PDO titer, albeit with re-duced rates of cell growth and glycerol consumption, and increasedacetate and lactate production (Lu et al., 2018). Similarly, mutation ofcrr in K. pneumoniae moderately improved 1,3-PDO production, al-though, in this case, side metabolite formation was reduced (Wanget al., 2017c). Also, co-fermentation of glycerol and untreated molassessignificantly improved the 1,3-PDO titer in a ΔldhA Δcrr double mutant(Oh et al., 2013a). Moreover, overexpression of native glpF significantlyenhanced the glycerol consumption rate and modestly increased the1,3-PDO titer in a fedbatch fermentation (Wang et al., 2017c). Finally,expression of fdh, encoding FDH from Candida boidinii, in K. oxytoca,slightly increased the 1,3-PDO titer while ethanol production increasedsubstantially (Zhang et al., 2009). However, the intracellular NAD+/NADH ratio was not affected by the activity of the NADH regenerationsystem, as was the case in E. coli (Berrı,́ S.J., San, K.-Y., Bennett, G.,2002), suggesting that either K. oxytoca can efficiently recycle NADHthrough native pathways, expression levels of fdh were not sufficient toaffect NADH regeneration, or substantial conversion of formate to H2

and CO2 by the formate hydrogen lyase (FHL) complex occurred. Si-milarly, chromosomal insertion of fdh from Pichia pastoris in the budClocus of K. pneumoniae significantly increased biomass accumulationand decreased 2,3-BDO production, although the 1,3-PDO titer wasmodestly improved (Wu et al., 2013), while coexpression of fdh from P.pastoris, gdh (encoding glucose dehydrogenase; GluDH) from Bacillussubtilis, and native udh increased both 1,3-PDO and lactate production(Wang et al., 2017c).

Recently, multiple parallel strain engineering strategies led to a veryhigh 1,3-PDO yield of 0.76 mol/molglycerol in a K. pneumoniae strain inwhich wabG, encoding a protein involved in the transfer of galacturonicacid (WabG), was mutated to prevent lipopolysaccharide synthesis (Leeet al., 2018). Mutation of ldhA and pflB in the ∆wabG mutant abolishedlactate and acetate production, respectively, with modest increases inthe 1,3-PDO titer, while mutation of budA nearly eliminated 2,3-BDOproduction at the expense of increased acetate production and severelyreduced glycerol consumption and 1,3-PDO production. To preventglycerol assimilation into biomass, dhaD and glpK were mutated in the∆wabG ∆ldhA ∆pflB ∆budA quadruple mutant, and the resulting strainKMK-23 was cultured with glycerol and glucose as co-substrates toenable growth on glucose with glycerol being converted to 1,3-PDO at asignificantly higher yield than the parent strain (Lee et al., 2018).Subsequently, mannitol was selected as the most suitable co-substratewith glycerol, and the 5’- untranslated region (UTR) of mtlA, encodingthe mannitol-specific transporter (MtlA), was modified in the genome ofKMK-23 to improve the efficiency of mannitol utilization, which, inturn, improved cell growth and 1,3-PDO production in the resultingstrain KMK-23M. Finally, based on the observation that DhaL, i.e. ahomologue of the subunit of DhaKII encoded by dhaK2, positivelyregulates DhaR, in turn, increasing the expression of DhaT in E. coli(Bächler et al., 2005), dhaK123 was mutated in KMK-23M, resulting instrain KMK-46, followed by overexpression of dhaK2 in KMK-46, re-sulting in strain KMY, which produced significantly more 1,3-PDO than

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

549

Page 13: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

KMK-23M. Interestingly, 1,3-PDO production was abolished in KMK-46,and could not be rescued by simply overexpressing dhaR. Moreover, theoverexpression of dhaT in KMK-23M or KMK-46 resulted in significantlylower 1,3-PDO production, compared to KMY (Lee et al., 2018), sug-gesting the complex and poorly understood regulation of glycerol me-tabolism in K. pneumoniae.

3.1.2.2. 3-HP. Due to its use in the synthesis of many industriallyimportant chemicals, 3-HP has received increasing attention as a targetmetabolite in engineered Klebsiella sp. 3-HP production from glycerolgenerates 1 mol NADH per mol glycerol consumed, such that co-production with 1,3-PDO is favorable due to redox balance, or anexogenous electron acceptor can improve the 3-HP/1,3-PDO selectivity(Ashok et al., 2013a). Strain engineering strategies employed toenhance the marginal levels of naturally produced 3-HP primarilyfocused on enhancing 3-HPA conversion to 3-HP via expression ofALDHs, and inactivating 1,3-PDOORs to increase available 3-HPA.Balanced 3-HP/1,3-PDO co-production was achieved viaoverexpression of native puuC and mutation of dhaT and, relative tothe parent strain, the 3-HP titer was improved by more than 10-foldwith a similar 1,3-PDO titer (Ashok et al., 2011). The existence ofputative oxidoreductases which convert 3-HPA to 1,3-PDO is welldocumented in Klebsiella sp., making elimination of 1,3-PDOformation difficult (Ko et al., 2015; Ko et al., 2012). Similarly,overexpression of native puuC increased the 3-HP titer by 5.5-fold ina strain with an inactivated oxidative branch of fermentative glycerolmetabolism, however, no effect was observed in the wild-type strain,suggesting that puuC overexpression was merely generatingcompensatory NADH in light of disruption of oxidative glycerolmetabolism (Luo et al., 2011a, 2011b). The ydcW and aldH genesfrom E. coli and native ydcW were selected out of 14 ALDHs evaluatedto enhance 3-HP production (Huang et al., 2012). The 3-HP titer wasincreased by 10-fold in a fedbatch fermentation of the aldH-expressingstrain, while the 1,3-PDO titer declined somewhat (Huang et al., 2012).Similarly, expression of heterologous α-ketoglutaric semialdehydedehydrogenase from Azospirillum brasilense, encoded by KGSADH,significantly increased both the 3-HP and 1,3-PDO titers, relative tothe overexpression of native puuC, in a fedbatch fermentation of theΔdhaT mutant (Ko et al., 2012). The dhaT mutation respectivelyincreased and decreased 3-HP and 1,3-PDO production in strainsexpressing puuC or KGSADH, while mutation of yqhD had nosignificant effect on culture performance. 3-HP/1,3-PDO co-production was subsequently investigated in a wild-type strainexpressing KGSADH (Ko et al., 2017). Mutation of ldhA, adhE, andfrdA (encoding succinate dehydrogenase; FrdA) significantly reducedlactate and ethanol formation, while succinate and 3-HP/1,3-PDO co-production were only slightly affected. Due to a marked increase inacetate formation observed in the ΔldhA ΔadhE ΔfrdA triple mutant, pta-ackA were mutated resulting in significantly decreased acetateformation, albeit at the expense of glycerol consumption and 3-HPproduction (Ko et al., 2017). Accordingly, acetate formation wasreduced without disrupting pta-ackA by mutating glpK in the ΔldhAΔadhE ΔfrdA triple mutant, resulting in significantly increased glycerolconsumption and 3-HP/1,3-PDO co-production, while mutation of dhaDimpeded culture performance. Moreover, overexpression of dhaT, ordhaB123 and gdrAB, increased and decreased 3-HP/1,3-PDO co-production in the ΔldhA ΔadhE ΔfrdA triple mutant, and all geneticstrategies employed in this study either reduced or did not affect 3-HPproduction (Ko et al., 2017). High-level 3-HP/1,3-PDO co-productionwas also achieved via expression of KGSADH from A. brasilense in aresting cell system (Kumar et al., 2012), and mutation of ldhA couldsignificantly improve both 3-HP and 1,3-PDO titers (Ashok et al.,2013b). In contrast to the aforementioned ΔldhA ΔadhE ΔfrdA ΔglpKquadruple mutant expressing KGSADH, a ΔglpK single mutantoverexpressing native puuC produced less 3-HP relative to the wild-type strain expressing puuC. On the other hand, mutation of dhaT

increased the 3-HP titer significantly in the wild-type strainoverexpressing puuC, while 3-HP production was highest in culturesof a ΔglpK ΔdhaT double mutant supplemented with nitrate as anelectron acceptor (Ashok et al., 2013a). Subsequently, inactivation ofdhaT and yqhD neither eliminated 1,3-PDO production nor substantiallyincreased 3-HP formation in the wild-type strain, while coexpression ofnative puuC, dhaB123, and gdrAB significantly increased 3-HPproduction in the ΔdhaT ΔyqhD double mutant, relative to puuCoverexpression alone, during fedbatch operation (Ashok et al.,2013b). Further efforts to reduce 1,3-PDO formation for enhanced 3-HP production in a ΔdhaT ΔyqhD double mutant expressing KGSADHfrom A. brasilense via mutation of adhE and ahpF, encoding alkylhydroperoxide oxidoreductase (AhpF), were unsuccessful (Ko et al.,2015), suggesting that additional 1,3-PDOORs have yet to be identified.Finally, the 3-HP titer increased by 3.9-fold and the 1,3-PDO titerdecreased substantially upon expression of dhaS, encoding ALDH fromB. subtilis (DhaS), but with significantly reduced glycerol consumptionand biomass accumulation (Su et al., 2015).

For more effective 3-HP production, the strain engineering strate-gies described above can be integrated with bioprocessing strategies.For example, while overexpressing native puuC from the strong tacpromoter, the 3-HP titer increased by 6-fold relative to that obtainedunder the lac promoter, and an extremely high 3-HP titer of 84 g/L wasobtained with a high 3-HP/1,3-PDO selectivity by manipulating aera-tion conditions (Li et al., 2016a, 2016b). It has been demonstrated thatthe selectivity of 3-HP to 1,3-PDO is highly dependent on aerationconditions, with fully aerobic operation abolishing both metabolitesdue to insufficient B12 synthesis, fully anaerobic conditions favoring1,3-PDO formation, and microaerobic conditions being conducive to co-production (Ashok et al., 2011; Ashok et al., 2013b; Li et al., 2016a,2016b; Su et al., 2015). While an improved 3-HP/1,3-PDO selectivity isachievable without disrupting 1,3-PDOOR expression, 1,3-PDO pro-duction remains significant such that identification and mutation ofadditional 1,3-PDOORs may improve culture performance for 3-HPproduction. Advanced genome engineering technologies such as theCRISPR-Cas9 (Clustered Regularly Interspaced Palindromic Repeats-CRISPR-associated [protein] 9) system facilitate multiplexed gene mu-tations in various microbial hosts (Jiang et al., 2015; Westbrook et al.,2016), and CRISPR interference (CRISPRi) has been applied to meta-bolic engineering in K. pneumoniae (Wang et al., 2017a). Accordingly,multiplexed mutation of genes encoding putative 1,3-PDOORs via theCRISPR-Cas9 system may be a convenient and effective approach toabolish 1,3-PDO production. Moreover, expression of NADH oxidase(NOX) can regenerate NAD+ to maintain the redox balance in a 1,3-PDO-deficient mutant without sacrificing carbon flux to side metaboliteformation.

3.1.2.3. Other products. High level 2,3-BDO production from glycerolwas achieved in K. oxytoca through inactivation of pduC, encoding theDODHt large subunit, and ldhA (Cho et al., 2015). The 2,3-BDO titerreached 132 g/L in the ΔpduC ΔldhA double mutant, owing tosignificantly increased glycerol dissimilation and decreased sidemetabolite formation, relative to the parent strain (Cho et al., 2015).Interestingly, the ratio of (2S,3S)-2,3-BDO to meso-2,3-BDO was higherfor the ΔpduC ΔldhA double mutant, suggesting that it consumed moreNADH per mole of 2,3-BDO produced than the parent strain. Duringaerobic growth, a significantly lower intracellular NAD+/NADH ratiowas observed in a K. pneumoniae strain in which the NADH:quinoneoxidoreductase I complex (NDH-1) was inactivated, resulting inmarkedly improved 2,3-BDO production without any significantchange in 1,3-PDO levels (Zhang et al., 2018). Strain engineering for2,3-BDO production has recently been reviewed (Yang et al., 2017;Yang and Zhang, 2018).

Production of 1-butanol from glycerol was achieved in K. pneumo-niae although titers were in the mg/L range (Wang et al., 2014a, 2014b,2014c). The ter and bdhAB genes, encoding trans-2-enoyl-CoA reductase

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

550

Page 14: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

from Treponema denticola (Ter) and butanol dehydrogenases A (BdhA)and B (BdhB) from Clostridium acetobutylicum, respectively, were co-expressed to construct a modified CoA-dependent 1-butanol pathway,while coexpression of kivD, encoding α-ketoisovalerate decarboxylasefrom Lactococcus lactis (KivD), to facilitate the 2-keto acid pathwaysignificantly increased 1-butanol production. In addition, asRNAs weredesigned to target dhaB123, dhaT, gdrAB, budB, and budC to reduce theformation of 1,3-PDO and 2,3-BDO, although silencing gdrB expressionalone was most effective in enhancing 1-butanol production (Wanget al., 2014a, 2014b, 2014c). 1-Butanol production from glycerol in K.pneumoniae via the 2-keto acid pathway was significantly enhancedthrough overexpression of native leuABCD (encoding 2-isopropylmalatesynthase, 3-isopropylmalate dehydrogenase, and 3-isopropylmalatedehydratase subunits C and D), kivD from L. lactis, and adhE1 (encodingaldehyde/alcohol dehydrogenase; AdhE1) from C. acetobutylicum(Wang et al., 2015a, 2015b). Moreover, coexpression of fdh from P.Pastoris, and native udh and gdh to regenerate NADH, or simultaneoustranscription of asRNAs targeting the 1,3-PDO pathway markedly in-creased the 1-butanol titer in K. pneumoniae, and side metabolite for-mation was greatly reduced (Wang et al., 2015a, 2015b). In addition,repression of expression of metA (encoding homoserine O-succinyl-transferase; MetA), ilvI (encoding acetolactate synthase isozyme 3 largesubunit; IlvI), ilvB (encoding acetolactate synthase isozyme 1 largesubunit; IlvB), and alaA (encoding alanine transaminase; AlaA) viaCRISPRi significantly increased 1-butanol production in K. pneumoniae,with repression of ilvB expression alone increasing the titer by 2.5-fold(Wang et al., 2017a). A number of strain engineering strategies thathave been applied to enhance 1-butanol production in E. coli, for ex-ample (Atsumi et al., 2008; Dellomonaco et al., 2011; Dong et al., 2017;Ohtake et al., 2017), may also prove successful in Klebsiella sp. Mutationof pta can increase acetyl-CoA availability to drive acetoacetyl-CoAproduction and simultaneously reduce acetate accumulation (Ohtakeet al., 2017), and optimization of the ribosome binding site (RBS)translation initiation rate of AdhE2 (Dong et al., 2017; Ohtake et al.,2017), which is the most common bifunctional aldehyde/alcohol de-hydrogenase employed for 1-butanol production, may enhance 1-bu-tanol production in Klebsiella sp. Moreover, disruption of the FHLcomplex, in combination with overexpression of fdh, can mitigate theredox imbalance imposed by 1-butanol production in Klebsiella sp.(Dong et al., 2017).

Metabolic engineering of a ΔldhA K. pneumoniae mutant for 2-bu-tanol production from glycerol was explored through overexpression ofnative ilvIH (encoding keto-acid reducto-isomerase; IlvIH) and ilvD(encoding dihydroxyacid dehydratase; IlvD), and kivD and adhA (en-coding alcohol dehydrogenase; AdhA) from L. lactis, (Oh et al., 2014).The 2-butanol titer increased significantly upon mutation of budA, andthe ldhA mutation was essential for 2-butanol production which beganupon cessation of cell growth (Oh et al., 2014).

Enhanced ethanol production from glycerol was investigated in a K.pneumoniae mutant generated by γ-irradiation (Oh et al., 2011; Ohet al., 2012a). Coexpression of pdc and adhII (encoding aldehyde/al-cohol dehydrogenase; AdhII) from Z. mobilis moderately increased theethanol titer during fedbatch fermentation of crude glycerol (Oh et al.,2011), while mutation of ldhA significantly improved ethanol produc-tion in the strain coexpressing pdc and adhII (Oh et al., 2012a).

High-level production of optically pure D-lactate from glycerol wasachieved by expressing native ldhA in a ΔdhaT ΔyqhD double mutant ofK. pneumoniae (Feng et al., 2014). Overexpression of ldhA with andwithout mutation of dhaT and yqhD increased the lactate titer by 11.5-fold and 5.3-fold, respectively, and the lactate titer reached 142 g/L in afedbatch fermentation under suitable aeration conditions (Feng et al.,2014).

Acetoin production from glycerol has also been explored in en-gineered K. pneumoniae (Wang et al., 2017b). Mutation of budC, gldA(encoding GlyDH; GldA), and dhaD prevented conversion of acetoin to2,3-BDO, albeit with reduced glycerol consumption and no acetoin

production. On the other hand, activation of the silent glp regulon viaexpression of native glpK further hindered glycerol dissimilation, butenabled acetoin production. Moreover, expression of nox, encodingNOX from Lactobacillus brevis, significantly improved the acetoin titer,which reached 32 g/L in a fedbatch fermentation (Wang et al., 2017b).

Finally, biopolymer production from glycerol was achieved in en-gineered K. pneumoniae (Feng et al., 2015). Poly(3-hydroxypropionate)[P(3HP)], a polyhydroxyalkanoate (PHA) possessing high rigidity,ductility, and tensile strength in drawn films, is considered an attractivealternative to petroleum-based plastics (Andreeßen et al., 2010). P(3HP) production was achieved by first enhancing 3-HP biosynthesisthrough coexpression of native dhaB123 and gdrAB, and aldH from E.coli. 3-HP was converted to 3-hydroxypropionyl-CoA (3-HP-CoA) whichwas subsequently polymerized into P(3HP) upon expression of prpE(encoding propionyl-CoA synthetase; PrpE) from E. coli, and phaC(encoding PHA synthase; PhaC) from Ralstonia eutropha. The 3-HP titerincreased by 6.4-fold in a ΔdhaT ΔyqhD double mutant, though thespecific yield of P(3HP) was low, suggesting that 3-HP is not an idealsubstrate for PrpE, which natively produces propionyl-CoA from pro-pionate (Feng et al., 2015).

3.2. Citrobacter sp.

3.2.1. Natural metabolismCitrobacter sp. are capable of fermentative glycerol dissimilation

through coupled oxidative and reductive pathways (Fig. 1). Glycerol isoxidized to DHA by a soluble NAD+-dependent GlyDH (i.e. DhaD en-coded by dhaD) and is subsequently phosphorylated into DHAP byDhaK (encoded by dhaK) (Daniel and Gottschalk, 1992; Daniel et al.,1995). While ATP-dependent DhaK of Citrobacter freundii showsminimal homology to enzymes in other microorganisms (Daniel et al.,1995), it contains K- and L-domains displaying similar functions toDhaK and DhaL of E. coli, respectively (Siebold et al., 2003). Note thatenzymes encoded by the dha regulon (i.e. DhaB/DhaBCE, DhaD, DhaK,and DhaT) from Klebsiella sp. and Citrobacter sp. share 80-90% sequencesimilarity, compared to 30-80% between Citrobacter sp. and Clostridiumsp. (Sun et al., 2003). In parallel to fermentative glycerol oxidation,glycerol is converted to 3-HPA by GDHt (i.e. DhaBCE encoded bydhaBCE) (Seifert et al., 2001) and further reduced to 1,3-PDO by DhaT,encoded by dhaT (Ainala et al., 2013). Analogous to DhaB in Klebsiellasp., the activity of DhaBCE is dependent on B12, which is naturallysynthesized in Citrobacter sp. (Keuth and Bisping, 1994). Upon theconversion to 3-HPA, glycerol inactivates DhaBCE by disrupting thecobalt-carbon (Co-C) σ-bond of the attached B12 (Seifert et al., 2001),resulting in a closer binding of the modified coenzyme to the active site(Toraya, 2000). This altered dehydratase complex is reactivated byDhaF and DhaG, encoded by dhaF and dhaG, respectively (Seifert et al.,2001). In C. freundii, dhaF is located downstream of dhaBCE, whereasdhaG is located downstream of dhaT with an opposite transcriptionaldirection to dhaBCE (Seifert et al., 2001). These findings indicate thatdhaF is co-transcribed with dhaBCE, while dhaG is transcribed sepa-rately with dhaT to prevent accumulation of 3-HPA (Jiang et al., 2016;Seifert et al., 2001).

3.2.2. Strain engineeringDue to poor genetic tractability, Citrobacter sp. has been unpopular

as a host for bio-based production. However, some progress has beenmade toward strain engineering of Citrobacter sp. for 1,3-PDO produc-tion. Preliminary attempts to produce 1,3-PDO in C. freundii focused onfermentation strategies using purified (Boenigk et al., 1993) or crude(Metsoviti et al., 2013) glycerol. As the first approach for strain en-gineering, expression of dhaT from Shimwellia blattae in C. freundiisignificantly increased 1,3-PDO production compared to the wild-typestrain during a fedbatch fermentation, with significant lactate forma-tion observed as the cells entered the stationary phase (Celińska et al.,2015). Accordingly, focus should be placed on minimizing lactate

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

551

Page 15: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

synthesis in C. freundii to further enhance 1,3-PDO production, as wassuccessfully achieved in Citrobacter werkmanii (Maervoet et al., 2016).C. werkmanii is also a native host for high-level 1,3-PDO production(Maervoet et al., 2012a, 2012b) and preferentially utilizes glycerol forgrowth, even in the presence of multiple carbon sources (Maervoetet al., 2014). Accordingly, dhaD was mutated to force dissimilatedglycerol through the reductive pathway during co-fermentation with asecond carbon source that could also support cell growth. Of several co-substrates tested, glycerol/glucose co-fermentation was considered fa-vorable due to a moderate increase in the 1,3-PDO yield and the rela-tively low cost of glucose, although insufficient NADH levels in theabsence of glycerol oxidation resulted in 3-HPA accumulation(Maervoet et al., 2014). Moreover, ethanol, lactate, and acetate for-mation was significant during glycerol/glucose co-fermentation(Maervoet et al., 2014), leading to subsequent inactivation of ldhA andadhE to relieve the formation of these side metabolites (Maervoet et al.,2016). During glycerol/glucose co-fermentation, the ∆dhaD ∆adhEdouble mutant and ∆dhaD ∆ldhA ∆adhE triple mutant produced 2.2-foldand 2.7-fold more 1,3-PDO, respectively, compared to the ∆dhaD singlemutant. In addition, 3-HPA accumulation was significantly reduced inthe ∆dhaD ∆adhE double mutant, compared to the ∆dhaD single mutant,and was undetectable in the ∆dhaD ∆ldhA ∆adhE triple mutant(Maervoet et al., 2016). Lastly, mutation of arcA resulted in a sig-nificant increase in acetate, lactate, and succinate formation with 1,3-PDO production being only slightly improved, whereas the ∆dhaD∆ldhA ∆arcA triple mutant did not produce any 1,3-PDO (Maervoetet al., 2016). As previously noted for Klebsiella sp., another potentialstrategy to increase NADH levels for enhanced 1,3-PDO production inCitrobacter sp. is to simultaneously inactivate the FHL complex andoverexpress FDH. In addition, 2,3-BDO formation has not been reportedin Citrobacter sp. engineered for enhanced 1,3-PDO production and thisrepresents a major advantage over other natural 1,3-PDO producerssuch as Klebsiella sp. due to potentially reduced purification costs forseparating 1,3-PDO and 2,3-BDO.

Violacein, a blue-purple pigment possessing anti-bacterial, anti-viral, and anti-tumoral properties, is produced by diverse genera ofbacteria, including Chromobacterium, Collimonas, Duganella,Janthinobacterium, Microbulbifer, and Pseudomonas (Choi et al., 2015;Durán and Menck, 2001; Rettori and Durán, 1998). Biological pro-duction of violacein has been relatively ineffective, with the most stu-died strain Chromobacterium violaceum generating small amounts ofcrude violacein (Mendes et al., 2001; Sánchez et al., 2006). Expressionof the violacein biosynthesis operon, i.e. vioABCDE, encoding L-tryp-tophan oxidase (VioA), 2-imino-3-(indol-3-yl) propanoate dimerase(VioB), deoxyviolaceinate synthase (VioC), protodeoxyviolaceinatemonooxygenase (VioD), and a protein of unknown function (VioE),from C. violaceum, resulted in negligible violacein production in en-gineered E. coli and Streptomyces albus (Pemberton et al., 1991; Sánchezet al., 2006). On the other hand, expression of the violacein biosynth-esis operon from Duganella sp. in C. freundii resulted in a crude violaceintiter of 1.7 g/L from glycerol in shake flask cultures supplemented withthe precursor L-tryptophan (Jiang, P.-x., Wang, H.-s., Zhang, C., Lou, K.,Xing, X.-H., 2010), and the titer reached 4.1 g/L in a fedbatch culti-vation (Yang et al., 2011).

3.3. Clostridium sp.

3.3.1. Natural metabolismDespite the ability to utilize a wide variety of carbon sources for

Clostridium sp. (Tracy et al., 2012), natural catabolism of glycerol as thesole carbon source appears to be confined to a small subset, of which C.butyricum and C. pasteurianum are biotechnologically important. Ad-ditionally, strains of Clostridium beijerinckii, Clostridium diolis, Clos-tridium saccharobutylicum, and Clostridium propionicum were also de-rived for glycerol dissimilation (Dobson et al., 2012; Tracy et al., 2012;Wang et al., 2014a, 2014b, 2014c). Clostridium sp. normally consume

glycerol under anaerobic conditions via coupled oxidative and re-ductive branches of fermentative glycerol dissimilation, channelingglycerol into glycolysis to generate energy and reducing equivalentswhich are consumed through the production of reduced metabolitessuch as 1,3-PDO (Tracy et al., 2012; Wang et al., 2015a, 2015b), withother acids and alcohols, including acetate, propionate, butyrate, lac-tate, ethanol, and butanol being co-produced (Oh et al., 2014; Tracyet al., 2012). Similar to glycerol dissimilation in other anaerobic mi-croorganisms, certain Clostridium sp. predominantly channel glycerol toglycolysis through the intermediates DHA and DHAP. As such, Clos-tridium sp. encode GlyDH (i.e. GldA or DhaD) and DHAK (i.e. DhaK).Concurrently, glycerol is converted to 3-HPA and 1,3-PDO by GDHt(encoded by dhaB, pduD, or pduE) and 1,3-PDOOR (encoded by dhaT),respectively, to maintain redox poise (Kubiak et al., 2012; Sandovalet al., 2015; Sun et al., 2003). While GDHt is typically B12-dependent(Pyne et al., 2016a, 2016b; Saint-Amans et al., 2001; Siebold et al.,2003), GDHt from C. butyricum (i.e. DhaB1 encoded by dhaB1) is B12-independent, a phenomenon that appears to be uncommon in nature (Liet al., 2016a, 2016b). The completed genome sequence of C. diolis DSM15410 also identified a putatively B12-independent GDHt (Wang et al.,2013a, 2013b). Interestingly, putative genes encoding enzymes in-volved in glycerol dissimilation, i.e. glpK and glpD, have been identifiedin several members of Clostridium, including Clostridium kluyveri, C.propionicum, and C. acetobutylicum (Nölling et al., 2001; Poehlein et al.,2016; Seedorf et al., 2008). This is notable as these glycerol dissimila-tion pathways are predominantly found in aerobic microorganisms.Moreover, while various enzymes involved in glycerol dissimilationhave been annotated in other Clostridium sp., these pathways appear tobe incomplete or dormant (Biebl, 2001; Sun et al., 2003; Vasconceloset al., 1994). Despite having genes associated with glycerol dissimila-tion, most Clostridium sp. are unable to metabolize glycerol as the solecarbon source, ostensibly due to the absence or deficiency of a 1,3-PDOpathway (González-Pajuelo et al., 2005; Seedorf et al., 2008). C. acet-obutylicum will utilize glycerol in the presence of a more oxidizedcarbon source such as glucose or pyruvate, or an exogenous electronacceptor such as methyl viologen or neutral red, suggesting that itsatypical glycerol dissimilation pathway is still active (Girbal et al.,1995; Girbal and Soucaille, 1994; Peguin et al., 1994; Vasconceloset al., 1994).

3.3.2. Strain engineeringDue to limited genetic tools for Clostridium sp. and the fact that

glycerol is a recalcitrant carbon source under anaerobic conditions(Jang et al., 2012; Pyne et al., 2014), strategies directed toward im-proving glycerol utilization for the production of value-added meta-bolites in Clostridium sp. remain uncommon. Random mutagenesis hasbeen widely applied to improve glycerol dissimilation and, as such,mutants with improved phenotypes remain uncharacterized geneti-cally. A C. pasteurianum mutant was capable of consuming glycerol at arate almost double that of the parent strain, resulting in enhancedproduction of 1,3-PDO and 1-butanol (Jensen et al., 2012). Similarly, aC. pasteurianum mutant with improved glycerol tolerance and growthrate produced significantly more 1-butanol than the parent strain(Malaviya et al., 2012), while 1-butanol tolerance and production in-creased in another mutant, even though glycerol consumption was notaffected (Gallardo et al., 2017). Genome shuffling was applied to C.diolis to improve 1,3-PDO tolerance, resulting in a marked increase inthe 1,3-PDO yield with enhanced glycerol utilization compared to theparent strain (Otte et al., 2009). A C. pasteurianum mutant with im-proved tolerance to crude glycerol was derived through random mu-tagenesis followed by directed evolution under increasing concentra-tions of crude glycerol, resulting in substantially increased 1-butanolproduction and reduced 1,3-PDO formation, and the improved cultureperformance was found to be associated with a large deletion in spo0A,encoding a transcriptional regulator for sporulation (SpoOA) (Sandovalet al., 2015). In agreement with this work, mutation of spoOA

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

552

Page 16: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

significantly improved glycerol dissimilation with a concomitant in-crease in 1-butanol production (Schwarz et al., 2017). Mutation of dhaTin C. pasteurianum led to a significant decrease in 1,3-PDO production,however, glycerol dissimilation persisted with 1,2-PDO formation as analternative redox valve (Pyne, Michael E. et al., 2016). Although theidentification of genes encoding putative enzymes related to the 1,2-PDO pathway suggests that Clostridium sp. such as C. acetobutylicum,Clostridium difficile, and C. beijerinckii may synthesize 1,2-PDO or itsprecursors, this was the first experimental demonstration of 1,2-PDOproduction in Clostridium sp. (Huang, K.-x., Rudolph, F.B., Bennett,G.N., 1999; Liyanage et al., 2001; Pyne et al., 2016a, 2016b). 1,3-PDOproduction was recently abolished in C. pasteurianum by deleting theentire dhaBCE operon using Allele-Coupled Exchange (ACE), although1-butanol and side metabolite formation were not significantly affected(Schwarz et al., 2017). In contrast, mutation of rex, encoding the redox-responsive regulator (Rex) that represses expression of genes in the 1-butanol pathway in response to low NADH levels, significantly de-creased 1,3-PDO and butyrate formation with markedly improved 1-butanol production, in spite of reduced glycerol utilization. Finally,mutation of hydA, encoding ferredoxin dehydrogenase (HydA), sig-nificantly increased acetate, lactate, and ethanol formation, with a re-duction in butyrate and 1,3-PDO production, and a modest increase inthe 1-butanol titer (Schwarz et al., 2017). The results are in agreementwith an earlier study of asRNA-mediated repression of hydA expressionin C. pasteurianum grown in a complex medium with glucose (Pyneet al., 2015).

As DhaB1 of C. butyricum is B12-independent, 1,3-PDO productionusing this organism is of significant interest (Raynaud, and xe, line, Sar,xe, abal, P., Meynial-Salles, I., Croux, C., Soucaille, P., 2003; Saint-Amans et al., 2001). Due to the lack of genetic tools, improving 1,3-PDO production in C. butyricum has been limited to derivation of 1,3-PDO-tolerant mutants via random mutagenesis (Abbad-Andaloussiet al., 1995; Reimann et al., 1998; Reimann and Biebl, 1996). The 1,3-PDO pathway of C. butyricum was introduced into C. acetobutylicum(González-Pajuelo et al., 2005) which has more advanced genetic tools(Pyne et al., 2014). The resulting engineered strains metabolized gly-cerol anaerobically, resulting in a moderate increase in the 1,3-PDOtiter relative to the natural producer C. butyricum in fedbatch fermen-tations (González-Pajuelo et al., 2005). Improved 1,3-PDO productionfrom glycerol was achieved in C. beijerinckii via coexpression of gldAand dhaKLM from E. coli and, compared to the wild-type strain, theengineered strain also had a greatly increased specific growth rate(Wischral et al., 2016). While the tools available for genetic manip-ulation in Clostridium sp. are lagging behind other industrially im-portant bacteria, recent advances such as ACE and CRISPR can accel-erate the pace of strain development. Notably, 74% of Clostridium sp.harbor CRISPR-Cas machinery, and native CRISPR-Cas systems can beexploited for genome editing with significantly greater editing effi-ciency than heterologous platforms (Pyne et al., 2016a, 2016b). Abol-ishing 1,3-PDO production in C. pasteurianum did not significantly altermetabolite profiles, such that systems level in silico analysis of genemutations combined with an efficient toolkit for genome editing may benecessary to dramatically reduce the labor of strain optimization forhigh-level 1-butanol production. In the case of C. pasteurianum, high-efficiency transformation protocols have been developed (Pyne et al.,2013; Schwarz et al., 2017), however, low transformability still remainsan issue for many promising Clostridium sp. (González-Pajuelo et al.,2006; Li et al., 2016a, 2016b) and must be addressed prior to the ap-plication of advanced genome engineering technologies for metabolicengineering purposes.

3.4. Lactobacillus sp.

3.4.1. Natural metabolismCertain Lactobacillus sp. strains are industrially important for lactate

production through anaerobic dissimilation of glycerol with sugar co-

substrates (Schütz and Radler, 1984; Talarico et al., 1990). Glycerol isconverted to 1,3-PDO via the reductive glycerol dissimilation pathwayto regenerate NAD+ consumed through the pentose phosphate pathway(PPP) (da Cunha and Foster, 1992; Garai-Ibabe et al., 2008). However,oxidative glycerol dissimilation does not occur naturally under fer-mentative and respiratory conditions due to incomplete and inactivepathways, respectively, in Lactobacillus sp. such that they cannot con-sume glycerol in the absence of a more oxidized carbon source (Kanget al., 2014a, 2014b; Morita et al., 2008; Talarico et al., 1990). Putativeglycerol kinase (GK) and NADP+-dependent G3P dehydrogenase arecommon to many Lactobacillus sp., although they are presumably dor-mant or weakly expressed as is the case in certain Clostridium sp.(Nölling et al., 2001; Vasconcelos et al., 1994) and K. pneumoniaestrains (Forage and Lin, 1982; Wang et al., 2017b). Putative GlyDH wasidentified in the genomes of several Lactobacillus sp., although DHAK isnot present, preventing oxidative glycerol dissimilation under fermen-tative conditions (Morita et al., 2008). The reductive pathway of fer-mentative glycerol dissimilation begins with the conversion of glycerolto 3-HPA, a process initially thought to occur via either of two distinctB12-dependent enzymes, i.e. GDHt and DODHt (Talarico andDobrogosz, 1990). However, the three subunits of a putative dehy-dratase from Lactobacillus reuteri JCM1112 show sequence similarity toDODHt of Salmonella typhimurium (i.e. PduCDE encoded by pduCDE)(Morita et al., 2008). Moreover, a metabolosome-associated DODHt(i.e. PduCDE encoded by pduCDE) was identified in L. reuteri DSM20016 that required pre-incubation with 1,2-PDO to achieve high 3-HPA production from glycerol (Sriramulu et al., 2008). Similar pduoperons have also been found in Lactobacillus collinoides (Sauvageotet al., 2002) and L. brevis (Makarova et al., 2006), indicating thecommon absence of a dedicated GDHt for Lactobacilli. The naturalability to synthesize B12 is an attractive feature of many L. reuteri andcertain Lactobacillus panis strains (Kang et al., 2014a; Morita et al.,2008; Ricci et al., 2015; Santos et al., 2011; Taranto et al., 2003), al-though it is not common to all Lactobacillus sp. (Pflugl et al., 2014). Thereduction of 3-HPA to 1,3-PDO is catalyzed by one or more 1,3-PDOORs, of which the activity and expression are sensitive to growthconditions. L. reuteri DSM 20016 expresses two putative 1,3-PDOORs,encoded by lr_0030 and lr_1734, which are active during exponentialand stationary growth phases, respectively (Stevens et al., 2011). Inaddition, lr_0030 is down-regulated upon depletion of glucose, resultingin 3-HPA accumulation, a phenomenon also observed in L. panis (Kanget al., 2013) and L. collinoides (Sauvageot et al., 2000). Among lactate-producing bacteria, L. reuteri has the unique capability to produce andexcrete 3-HPA in large quantities. The 3-HPA/1,3-PDO ratio decreasesas the glucose/glycerol ratio increases in L. reuteri cultures (Luthi-Penget al., 2002), although 1,3-PDO formation is hindered when glycerol ispresent in excess (~30 g/L) (Bauer et al., 2010). The latter observationis not surprising given that expression of dhaT is repressed under highglycerol loading in L. panis (Kang et al., 2013), a process aggravated bythe 3-HPA-mediated inhibition of DhaT (Kang et al., 2014b). Con-versely, the 1,3-PDO titer increased as the glycerol concentration wasincreased up to 70 g/L during batch fermentations of Lactobacillusdiolivorans (Pflugl et al., 2012).

In general, cell growth can be maintained or enhanced upon theaddition of glycerol to anaerobic glucose cultures of Lactobacillus sp. (daCunha and Foster, 1992; El-Ziney et al., 1998; Pflugl et al., 2012) with ametabolic shift from ethanol and lactate to 1,3-PDO and acetate (daCunha and Foster, 1992; El-Ziney et al., 1998; Luthi-Peng et al., 2002;Pflugl et al., 2012). The formation of 1,3-PDO replaces ethanol and, to alesser extent, lactate for NAD+ replenishment, enabling the productionof additional ATP via acetate synthesis (da Cunha and Foster, 1992).Lactate can also be oxidized to generate acetate under low sugar con-centrations, with 1,3-PDO serving as the major redox outlet (da Cunhaand Foster, 1992). Production of 3-HP has also been observed in L.reuteri (Ramakrishnan et al., 2015), L. collinoides (Garai-Ibabe et al.,2008), and L. diolivorans (Garai-Ibabe et al., 2008), serving to restore

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

553

Page 17: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

the cellular redox balance upon depletion of fermentable sugars (Luoet al., 2011a, 2011b; Sriramulu et al., 2008). Expression of pduP from L.reuteri in K. pneumoniae markedly enhanced 3-HP production relative tothe wild-type strain (Luo et al., 2011a, 2011b), indicating that PduPmay be associated with the 3-HP pathway in Lactobacillus sp. as is thecase in Klebsiella sp. (Luo et al., 2012). Strains of L. diolivorans and L.collinoides isolated from bitter tasting ciders produced 3-HP as a majormetabolite during fermentation on fructose and glycerol (Garai-Ibabeet al., 2008). High sugar concentrations favored 1,3-PDO production,while equimolar quantities of 3-HP and 1,3-PDO were obtained underfructose limitation (Garai-Ibabe et al., 2008).

3.4.2. Strain engineeringEngineering Lactobacillus sp. to enhance glycerol dissimilation has

emerged recently for synthesis of 1,3-PDO and 3-HP. An attempt toimprove 1,3-PDO production in engineered L. reuteri was made throughexpression of yqhD from E. coli, resulting in a significant increase in thespecific 1,3-PDO productivity relative to the wild-type strain(Vaidyanathan et al., 2011). However, a shift from acetate and biomassto ethanol and lactate formation was observed, presumably due to thepreference of YqhD for NADPH, diminishing the activity of the nativeNADH-dependent 1,3-PDOOR (i.e. PduQ encoded by pduQ) and, inturn, disturbing the natural redox balance (Vaidyanathan et al., 2011).As part of an assessment of heterologous recombinases in L. reuteri,mutation of the catabolite repression element (CRE) located upstreamof the pdu operon resulted in a 3-fold increase in 3-HPA productionwhen cells grown overnight in a medium containing glucose weretransferred to a medium containing glycerol (van Pijkeren et al., 2012).The CRE mutant was subsequently evaluated for 3-HP/1,3-PDO co-production in a fedbatch resting cell fermentation and the specificproductivities of 3-HP and 1,3-PDO increased significantly, relative tothe wild-type strain, with essentially all glycerol being converted to thetarget metabolites (Dishisha et al., 2014).

L. panis has been engineered by expressing yqhD from E. coli toenhance 1,3-PDO production from thin stillage with a significant gly-cerol content (Kang et al., 2014b). The L. panis strain was further en-gineered to dissimilate glycerol as the sole carbon source through me-tabolic introduction of a respiratory pathway (Kang et al., 2014a,2014b). However, coexpression of glpK, glpF, glpD, and tpi, encodingtriosephosphate isomerase (Tpi), from E. coli resulted in poor cellgrowth and minimal glycerol consumption with lactate as the majorend metabolite, indicating that NADH generated through the syntheticrespiratory pathway was not effectively recycled. Accordingly, coex-pression of yqhD from E. coli was employed to restore redox poise (Kanget al., 2014a, 2014b). It was anticipated that YqhD could reduce acetateto acetaldehyde, which could be further reduced to ethanol via an en-dogenous alcohol dehydrogenase, while simultaneously enhanced 3-HPA conversion to 1,3-PDO would occur (Kang et al., 2014a). The celldensity and glycerol utilization increased dramatically upon coexpres-sion of yqhD, relative to the control strain expressing the synthetic re-spiratory pathway, and ethanol and 1,3-PDO were the major reducedmetabolites. Enhanced 1,3-PDO production was also demonstrated in L.diolivorans by overexpressing a native putative NADPH-dependent 1,3-PDOOR (i.e. PDO-DH (NADPH)) (Pflugl et al., 2013). The approachresulted in slightly improved cell growth and a moderate increase in1,3-PDO formation, indicating that the putative 1,3-PDOOR readilyuses NADH or NADPH as a cofactor to convert 3-HPA to 1,3-PDO.

While engineering of glycerol metabolism in Lactobacillus sp. has notbeen studied extensively, many of these microorganisms have beengranted the Generally Recognized as Safe (GRAS) status from the FDA,USA, making them ideal hosts for the conversion of glycerol to value-added products on an industrial scale. In particular, further explorationof 1,3-PDO production in Lactobacillus sp. is warranted due to theirnatural production potential and apparent minimal 2,3-BDO formationas a side metabolite. For example, to overcome the NADH accumulationissue upon the introduction of a respiratory pathway for dissimilation of

glycerol as the sole carbon source in L. panis (Kang et al., 2014a,2014b), cofactor engineering via expression of NOX could regenerateNAD+ without forming competing side products (i.e. ethanol), andexpression of DhaT from Klebsiella sp. or Clostridium sp. may improveconversion of 3-HPA to 1,3-PDO while consuming NADH. Moreover,expression of NOX may also alleviate NAD+ depletion in Lactobacillussp. engineered to overproduce 3-HP. To further enhance glycerol dis-similation, the native respiratory pathway could be activated and co-expression of glpF from a Gram-positive organism, e.g. a facultativeanaerobe such as B. subtilis, may enhance the functional expression ofthe membrane-bound permease, relative to GlpF from a Gram-negativebacterium. Similarly, an engineered fermentative pathway containingthe native putative GlyDH may facilitate glycerol dissimilation in theabsence of a more reduced carbon source. Finally, in light of the sub-stantial number of putative LDHs and aldehyde/alcohol dehy-drogenases in Lactobacillus sp., knocking out relevant genes by theCRISPR-Cas9 system could simplify the elimination of lactate andethanol production to enhance target metabolite formation.

3.5. Escherichia coli

3.5.1. Natural metabolismE. coli can naturally dissimilate glycerol aerobically and anaerobi-

cally via respiratory and fermentative pathways (Dharmadi et al.,2006). Glycerol is transported into E. coli through facilitated diffusionby GlpF (Richey and Lin, 1972), which can also transport polyols (e.g.erythritol), urea, glycine, and glyceraldehyde (Heller et al., 1980). Inthe cytosol, glycerol enters the respiratory or fermentative pathways,either of which produce DHAP as an intermediate of glycolysis. In therespiratory pathway, glycerol is phosphorylated to G3P by GK (i.e. GlpKencoded by glpK) (Hayashi and Lin, 1967), the presumed limiting stepfor aerobic dissimilation of glycerol (Lin, 1976a, 1976b; Zwaig et al.,1970). Subsequently, if oxygen acts as an electron acceptor, G3P isconverted to DHAP by aerobic-G3PDH (i.e. GlpD encoded by glpD), oranaerobic-G3PDH (i.e. GlpABC encoded by glpABC) in the presence ofelectron acceptors other than oxygen (i.e. nitrate and fumarate). In thefermentative pathway, glycerol is oxidized to DHA via NAD+-depen-dent GlyDH (i.e. GldA encoded by gldA) (Tang et al., 1979), followed byphosphorylation of DHA to DHAP via PEP-dependent DHAK (i.e.DhaKLM encoded by dhaKLM) (Gonzalez et al., 2008). The expressionof enzymes in the respiratory and fermentative pathways is under acollective control of the global regulators cAMP-CRP, Cra, ArcA, Fnr,and RpoS encoded by crp, cra, arcA, fnr, and rpoS, respectively (Iuchiand Lin, 1993; Martínez-Gómez et al., 2012).

In the absence an exogenous electron acceptor, E. coli cannotnaturally maintain redox poise through the production of reducedcompounds other than 1,2-PDO (Dharmadi et al., 2006). The net NADHproduced through the lactate, ethanol, succinate, and 1,2-PDO path-ways are 1, 0, 0, and -1 mol NADH per mol glycerol consumed, re-spectively, such that only 1,2-PDO formation can regenerate NAD+. Inaddition to redox balance, glycerol fermentation is dependent on pHand the availability of CO2 (Durnin et al., 2009). E. coli requires CO2 inthe form of bicarbonate (HCO3

-) for growth and biosynthesis of varioussmall molecules and cellular components (Dharmadi et al., 2006).Under fermentative conditions, CO2 production mainly relies on theactivity of the FHL complex (encoded by fdhF and hycBCDEFG) for theconversion of formate to CO2 and H2 (Iuchi and Lin, 1993). Tran-scription of the genes encoding the FHL complex requires formate andan acidic pH (Sawers et al., 2004). An alkaline environment causes anequilibrium shift toward HCO3

-, which cannot be transported across thecellular membrane of E. coli, thus reducing CO2 availability inside thecell (Dharmadi et al., 2006). H2 is co-produced with CO2 and can act asan electron donor, further contributing to the redox imbalance duringglycerol fermentation and potentially impairing growth (Dharmadiet al., 2006). Other glycerol fermentation products of E. coli includeethanol, succinate, and acetate (Gonzalez et al., 2008), and ethanol

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

554

Page 18: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Table3

Sum

mar

yof

liter

atur

efo

rE.

coli

stra

ins

engi

neer

edfo

ren

hanc

edna

tura

lmet

abol

itepr

oduc

tion

from

glyc

erol

.

Parent

strain

Product

Geneticstrategy

Titer(g/L)

Cultivationmode

Reference

ATC

C873

9et

hano

l-M

utat

efrdA

BCD,

ackA

-pta

andpo

xB,a

ndldhA

tore

duce

succ

inat

e,ac

etat

e,an

dla

ctat

efo

rmat

ion,

resp

ectiv

ely,

and

mut

ategldA

andglpK

for

cont

rolle

dex

pres

sion

-Ove

rexp

ress

dhaK

LM,g

ldA

,glpK,

andad

hEfo

rin

crea

sed

carb

onflu

xan

det

hano

lfor

mat

ion

47Fe

dbat

ch(W

ong

etal

.,20

14)

LY18

0et

hano

lFo

smid

cont

aini

ngm

etag

enom

icfr

agm

ent

isol

ated

from

anae

robi

cre

acto

r75

Batc

h(L

oace

set

al.,

2016

)M

G16

55et

hano

l-M

utat

efrdA

,pta

,and

fdhF

tore

duce

succ

inat

e,ac

etat

e,an

dCO

2,an

dH

2fo

rmat

ion,

resp

ectiv

ely

-Ove

rexp

ress

gldA

anddh

aKLM

for

incr

ease

dca

rbon

flux

thro

ugh

the

ferm

enta

tive

bran

ch10

Batc

h(Y

azda

nian

dG

onza

lez,

2008

)

BW25

113

H2

-Mut

atehy

cA,h

yaB,

hybC

,hyc

E,hy

fG,p

flB,f

hlA

,fdh

F,gldA

,orglpA

toaff

ect

H2

form

atio

nN

/ACr

imp-

top

vial

s(S

anch

ez-T

orre

set

al.,

2013

)BW

2511

3H

2M

utat

earoM

,gatZ,

ycgR

,oryfgI

toen

hanc

eH

2fo

rmat

ion

See

refe

renc

eCr

imp-

top

vial

s(T

ran

etal

.,20

15)

BW25

113

H2

Mut

atefrdC

,ldh

A,f

dnG

,ppc

,narG

,mgsA,a

ndhy

cAto

enha

nce

H2

form

atio

n15

8μm

ol/m

gpr

otei

nCr

imp-

top

vial

s(T

ran

etal

.,20

14)

ATC

C873

9su

ccin

ate

-Mut

atepfl

Bto

redu

cefo

rmat

e,et

hano

l,an

dac

etat

efo

rmat

ion

-Mut

ateptsI

tosh

iftPE

Pflu

xaw

ayfr

omD

HA

K-M

utat

ion

inpr

omot

orre

gion

ofpc

kge

neto

incr

ease

expr

essi

onle

vel

12Fl

ask

(Zha

nget

al.,

2010

)

MG

1655

succ

inat

e-M

utat

ead

hE,p

taan

dpo

xB,a

ndldhA

tore

duce

etha

nol,

acet

ate,

and

lact

ate

form

atio

n,re

spec

tivel

y-M

utat

epp

cto

rem

ove

sele

ctio

nre

quir

emen

t-E

xpre

sspy

c(L.lac

tis)

for

succ

inat

epr

oduc

tion

from

pyru

vate

node

14Fl

ask

(Bla

nksc

hien

etal

.,20

10)

BL21

(DE3

)su

ccin

ate

-Mut

atepp

can

dch

rom

osom

alex

pres

sion

ofgl

yoxy

late

shun

tgen

esac

eBAK

follo

wed

bydi

rect

edev

olut

ion

fori

ncre

ased

glyo

xyla

tesh

unt

activ

ityan

dre

activ

atio

nof

ppc

-Mut

atesdhC

DAB

tore

duce

prod

uctd

egra

datio

n-C

hrom

osom

alex

pres

sion

ofsucA

Bto

redu

ceα-

keto

glut

arat

efo

rmat

ion

43Fe

dbat

ch(L

iet

al.,

2013

)

MG

1655

D-la

ctat

e-M

utat

epta,

adhE

,and

frdA

tore

duce

acet

ate,

etha

nol,

and

succ

inat

efo

rmat

ion,

resp

ectiv

ely

-Mut

atedld

tore

duce

D-la

ctat

ede

grad

atio

n-O

vere

xpre

ssglpD

andglpK

toin

crea

seca

rbon

flux

thro

ugh

the

resp

irat

ory

path

way

45Fl

ask

(Maz

umda

ret

al.,

2010

)

BL21

(DE3

)D

-lact

ate

-Mut

atepta,

adhE

,frdA

,pflB

,and

pflCD

tore

duce

side

met

abol

itefo

rmat

ion

–Mut

atemgsA

anddld

tore

duce

MG

form

atio

nan

dD

-la

ctat

ede

grad

atio

n,re

spec

tivel

y-D

irec

ted

evol

utio

nfo

rin

crea

sed

crud

egl

ycer

olut

iliza

tion

-Chr

omos

omal

expr

essi

onof

glpD

andglpK

toin

crea

seca

rbon

flux

thro

ugh

the

resp

irat

ory

path

way

,and

chro

mos

omal

expr

essi

onof

focA

andldh

(L.h

elvetic

us)

toin

crea

sela

ctat

eex

port

and

form

atio

n,re

spec

tivel

y

105

Fedb

atch

(Wan

get

al.,

2015

a,20

15b)

MG

1655

L-la

ctat

e-M

utat

epfl

B,pta,

adhE

,and

frdA

tore

duce

side

met

abol

itefo

rmat

ion

-Mut

atemgsA

andldhA

topr

even

tD

-lact

ate

form

atio

n,an

dm

utat

elld

Dto

redu

ceL-

lact

ate

degr

adat

ion

-Ove

rexp

ress

glpK

andglpD

toin

crea

seca

rbon

flux

thro

ugh

the

resp

irat

ory

path

way

-Chr

omos

omal

expr

essi

onof

ldh

(S.b

ovis)

for

L-la

ctat

efo

rmat

ion

50Fl

ask

(Maz

umda

ret

al.,

2013

)

BW25

113

L-la

ctat

e-M

utat

eac

kA,p

ta,p

ps,p

flB,p

oxB,

adhE

,and

frdA

tore

duce

side

met

abol

itefo

rmat

ion

-Mut

ateldhA

andlld

topr

even

tD

-lact

ate

form

atio

nan

dpr

oduc

tdeg

rada

tion,

resp

ectiv

ely

-Chr

omos

omal

expr

essi

onof

ldh

(B.c

oagu

lans

)fo

rL-

lact

ate

form

atio

n

142

Fedb

atch

(Tia

net

al.,

2013

)

W31

10L-

phen

ylal

anin

e-M

utat

etyrA

tore

duce

L-tr

ypto

phan

form

atio

n-M

utat

epy

kAan

dpy

kFto

incr

ease

avai

labl

ePE

P,an

dm

utat

ephe

Aan

daroF

for

cont

rolle

dex

pres

sion

-Ove

rexp

ress

aroF

BLan

dfe

edba

ckre

sist

antp

heAfbr fo

rin

crea

sed

shik

imat

ean

dL-

phen

ylal

anin

efo

rmat

ion,

resp

ectiv

ely

-Fee

dla

ctat

ean

dam

mon

ia

13Fe

dbat

ch(W

eine

ret

al.,

2014

)

BL21

(DE3

)fu

mar

ate

-Pre

viou

sly

evol

ved

succ

inat

epr

oduc

ing

stra

inco

ntai

ning

ppc

mut

atio

nan

dex

pres

sing

chro

mos

omal

aceB

AK

-Mut

atefumABC

andaspA

tore

duce

fum

arat

ede

grad

atio

n-O

vere

xpre

sspp

cto

incr

ease

fum

arat

esy

nthe

sis

42Fe

dbat

ch(L

iet

al.,

2014

)

MG

1655

form

ate

-Mut

atefrdA

,pta

,and

fdhF

tore

duce

side

met

abol

itefo

rmat

ion

-Ove

rexp

ress

gldA

anddh

aKLM

for

incr

ease

dgl

ycer

olut

iliza

tion

7Ba

tch

(Yaz

dani

and

Gon

zale

z,20

08)

MG

1655

FFA

s-M

utat

efadD

andptsG

toen

hanc

eFF

Afo

rmat

ion,

and

mut

atefabR

tore

mov

ere

pres

sion

offa

tty

acid

synt

hesi

sge

nes

-Ove

rexp

ress

nadK

andpn

tAB

toin

crea

sein

trac

ellu

lar

redo

xeq

uiva

lent

s,an

dex

pres

san

acyl

-ACP

thio

este

rase

(R.c

ommun

is)fo

rin

crea

sed

FFA

form

atio

n

5Fl

ask

(Wu

etal

.,20

14)

BL21

(DE3

)D

HA

-Coe

xpre

ssno

x(E

.fae

calis

)an

dgldA

for

incr

ease

dN

AD

+re

gene

ratio

nan

dD

HA

form

atio

n,re

spec

tivel

y-E

xoge

nous

supp

lyof

NA

D+

2Te

sttu

be(Y

ang

etal

.,20

13)

Lin4

3ac

etol

-Mut

ategloA

tore

duce

MG

degr

adat

ion

-Exp

ress

yqhD

for

incr

ease

dac

etol

synt

hesi

s5

Flas

k(Z

huet

al.,

2013

)

(con

tinue

don

next

page

)

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

555

Page 19: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

formation is essential as it generates ATP and is redox-balanced(Murarka et al., 2008). Although succinate production is also redox-balanced, it contributes minimally to ATP generation (Unden andKleefeld, 2004), making it non-essential during glycerol fermentation(Murarka et al., 2008), whereas acetate production is often marginalduring glycerol fermentation.

The highly reduced nature of glycerol makes it an ideal feedstock forthe production of reduced compounds and derived value-added pro-ducts. Anaerobic conditions favor the production of reduced com-pounds as oxygen is not available as an electron acceptor. However,anaerobic glycerol fermentations require supplementation of richmedia components to counteract the production of reducing equiva-lents generated via biomass formation (κglycerol=4.7; κbiomass=4.3) andthe low activity of the 1,2-PDO pathway (Durnin et al., 2009). Incontrast to anaerobic fermentation, microaerobic conditions introducelimiting amounts of oxygen, effectively acting as an electron acceptorfor redox equivalents generated from biomass, such that rich mediacomponents are no longer required while the production of reducedcompounds is maintained (Durnin et al., 2009).

3.5.2. Strain engineeringTable 3 and Table 4 summarize strategies for strain engineering of E.

coli for enhanced production of natural and non-native metabolitesfrom glycerol, respectively. Selective target products are used fortechnical illustration in this section.

3.5.2.1. 1,3–PDO. Preliminary attempts to produce 1,3-PDO in E. coliby expressing the dha regulon from either K. pneumoniae (Sprengeret al., 1989; Tong et al., 1991) or C. freundii (Daniel and Gottschalk,1992) resulted in substantially lower 1,3-PDO titers than those obtainedwith natural production hosts (Cameron et al., 1998; Chotani et al.,2000). As a result, more advanced strain engineering strategies weredeveloped to enhance 1,3-PDO production in E. coli, primarilyincluding: (1) the improved expression of genes involved in thereductive branch of glycerol dissimilation from natural 1,3-PDOproducers; (2) the expression of native putative NADPH-dependent1,3-PDOOR, YqhD; and (3) reducing the formation of the toxicmetabolites MG, G3P, and 3-HPA. MG is an intermediate of 1,2-PDOproduction, and concentrations exceeding 0.5 mM can inhibit cellgrowth (Ackerman et al., 1974; Booth et al., 2003; MacLean et al.,1998a, 1998b). Anaerobic fermentation of glycerol by E. coli strainsexpressing the dha regulon from K. pneumoniae resulted in theaccumulation of MG to toxic levels with lowered 1,3-PDO production(Zhu et al., 2001). As MG can be converted to S- lactoylglutathinone byglyoxalase I, and subsequently to lactate by glyoxalase II (MacLeanet al., 1998a, 1998b), expression of glyoxalase I, i.e. Glo1 (encoded byglo1 from Pseudomonas putida), reduced intracellular MG levels andmoderately enhanced 1,3-PDO production (Zhu et al., 2001). Similarly,both cell growth and 1,3-PDO production were inhibited when theintracellular G3P concentration exceeded 10 mM in E. coli (Lin, 1976a,1976b; Zhu et al., 2002), and the addition of an exogenous electronacceptor such as fumarate could increase the activity of GlpABC and, inturn, curb G3P accumulation and dramatically improve 1,3-PDOproduction (Zhu et al., 2002).

The reductive branch of glycerol dissimilation from K. pneumoniae,C. freundii, and C. butyricum have been introduced in E. coli to facilitate1,3-PDO production. While dhaB123 and dhaT are divergently ex-pressed from individual promoters in K. pneumoniae, this arrangementresulted in a lower 1,3-PDO titer in E. coli, compared to the coexpres-sion of dhaB123 and dhaT from a single promoter (Skraly et al., 1998).Furthermore, native YqhD could convert 3-HPA to 1,3-PDO more effi-ciently than DhaT from K. pneumoniae (Emptage et al., 2003b), and thepreference of YqhD for NADPH may have contributed to higher 1,3-PDO titers by altering the reduced/oxidized cofactor ratios (Nakamuraand Whited, 2003). Coexpression of gdrAB and dhaB123 from K.pneumoniae with native yqhD significantly increased 1,3-PDOTa

ble3

(con

tinue

d)

Parent

strain

Product

Geneticstrategy

Titer(g/L)

Cultivationmode

Reference

MG

1655

1,2-

PDO

-Mut

ateac

kAan

dpta,

andldhA

tore

duce

acet

ate

and

lact

ate

form

atio

n,re

spec

tivel

y-M

utat

edh

aKan

dex

pres

sdh

aKL

(C.f

reun

dii)

for

ATP

depe

nden

tfu

nctio

nalit

y(c

ompa

red

tona

tive

PEP

depe

nden

ce)

-Exp

ress

msgA

,gldA

,and

yqhD

for

incr

ease

d1,

2-PD

Ofo

rmat

ion

6Ba

tch

(Clo

mbu

rgan

dG

onza

lez,

2011

)

Gen

epr

oduc

ts:a

ckA

,ace

tate

kina

se;a

ceBA

K,is

ocitr

ate

lyas

e,m

alat

esy

ntha

se,a

ndis

ocitr

ate

dehy

drog

enas

eki

nase

/pho

spho

ryla

se;a

dhE,

alde

hyde

-alc

ohol

dehy

drog

enas

e;aroF

BL,3

-deo

xy-D

-ara

bino

-hep

tulo

sona

te-7

-ph

osph

ate

synt

hase

isoe

nzym

e,3-

dehy

droq

uina

tesy

ntha

se,

and

shik

imat

eki

nase

;aroM

,pr

otei

nA

roM

;aspA

;as

part

ate

amm

onia

-lyas

e;dh

aKLM

,D

HA

kina

se,

PEP-

depe

nden

tdi

hydr

oxya

ceto

neki

nase

AD

P-bi

ndin

gsu

buni

t,an

dpr

otei

n-ly

sine

deac

etyl

ase;

dld,

quin

one-

depe

nden

tD-la

ctat

ede

hydr

ogen

ase;

fabR

,HTH

-type

tran

scri

ptio

nalr

epre

ssor

;fad

D,lo

ng-c

hain

fatt

yac

idCo

Alig

ase;

fdhf

,for

mat

ede

hydr

ogen

ase

H;fdn

G,f

orm

ate

dehy

drog

enas

eN

subu

nit

alph

a;fhlA

,fo

rmat

ehy

drog

enly

ase

tran

scri

ptio

nal

activ

ator

;focA

,fo

rmat

etr

ansp

orte

r1;

frdA

,fu

mar

ate

redu

ctas

efla

vopr

otei

nsu

buni

t(a

naer

obic

);frdA

BCD

,fum

arat

ere

duct

ase

enzy

me

com

plex

;frdC,

fum

arat

ere

duct

ase

subu

nit

C;fumABC

,fum

aras

esu

buni

tsA

,B,a

ndC;

gatZ

,D-ta

gato

se-1

,6-b

isph

osph

ate

aldo

lase

subu

nit;

gldA

,gly

cero

ldeh

ydro

gena

se;g

loA

,lac

toyl

glut

athi

one

lyas

e;glpA

,ana

erob

icgl

ycer

ol-3

-pho

spha

tede

hydr

ogen

ase

subu

nitA

;glpK,

glyc

erol

kina

se;h

yaB,

hydr

ogen

ase-

1la

rge

chai

n;hy

bC,h

ydro

gena

se-2

larg

ech

ain;

hycA

,for

mat

ehy

drog

enly

ase

regu

lato

rypr

otei

n;hy

cE,f

orm

ate

hydr

ogen

lyas

esu

buni

t5;

hyfG

,hy

drog

enas

e-4

com

pone

ntG

;ldh,

lact

ate

dehy

drog

enas

e;ldhA

,la

ctat

ede

hydr

ogen

ase

A;lld

D,L-

lact

ate

dehy

drog

enas

e;mgsA

,m

ethy

lgly

oxal

synt

hase

;na

dK,

NA

Dki

nase

;na

rG,

resp

irat

ory

nitr

ate

redu

ctas

e1

alph

ach

ain;

nox,

NA

DH

oxid

ase;

pck,

PEP

carb

oxyk

inas

e;pfl

B,fo

rmat

eac

etyl

tran

sfer

ase

1;pfl

CD,p

yruv

ate

form

ate

lyas

esu

buni

tsC

and

D;p

heA

,P-p

rote

in;p

ntAB,

prot

ein

PntA

B;po

xB,p

yruv

ate

dehy

-dr

ogen

ase;

ppc,

PEP

carb

oxyl

ase;

pta,

phos

phot

rans

acet

ylas

e;ptsG

,PTS

syst

emgl

ucos

e-sp

ecifi

cEI

ICB

com

pone

nt;p

tsI,P

EPph

osph

otra

nsfe

rase

;pyc

,pyr

uvat

eca

rbox

ylas

e;py

kA,p

yruv

ate

kina

seII;

pykF

,pyr

uvat

eki

nase

I;sdhC

DAB,

succ

inat

ede

hydr

ogen

ase

com

plex

;suc

AB,

alph

a-ke

togl

utar

ate

dehy

drog

enas

e;tyrA

,T-p

rote

in;y

cgR,

flage

llar

brak

epr

otei

n;yfgI

,unc

hara

cter

ized

prot

ein;

yqhD

,NA

D(P

)H-d

epen

dent

1,3-

prop

aned

iold

ehy-

drog

enas

e.A

bbre

viat

ions

:1,2

-PD

O,1

,2-p

ropa

nedi

ol;a

cyl-A

CP,a

cylc

arri

erpr

otei

n;A

TP,a

deno

sine

trip

hosp

hate

;CO

2,ca

rbon

diox

ide;

DH

A,d

ihyd

roxy

acet

one;

DH

AK,

dihy

drox

yace

tone

kina

se,F

FA,f

ree

fatt

yac

ids;

H2,

hydr

ogen

gas;

MG

,met

hylg

lyox

al;N

AD

+,n

icot

inam

ide

aden

ine

dinu

cleo

tide;

PEP,

phos

phoe

nolp

yruv

ate.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

556

Page 20: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

production relative to the strain expressing only dhaB123 and yqhD,(Hong et al., 2015), indicating that glycerol- and/or oxygen-mediatedinactivation of DhaB was a critical issue in these E. coli strains. Ad-ditionally, supplementation of the TCA cycle intermediates succinate,fumarate, and malate significantly reduced the NAD+/NADH ratio andfurther improved 1,3-PDO production (Hong et al., 2015). Subse-quently, gapN, encoding NADP+-dependent glyceraldehyde-3-phos-phate dehydrogenase (GAPDN) from C. acetobutylicum, was coexpressedwith native yqhD, and dhaB123 and gdrAB from K. pneumoniae, to re-generate NADPH, resulting in significantly improved 1,3-PDO produc-tion relative to the strain coexpressing yqhD, dhaB123, and gdrAB (Yanget al., 2018). Fine-tuning the expression of gapN via modification of its5’-UTR led to significant improvements to culture performance, whichwas further enhanced by mutating ptsG, encoding the phospho-transferase system (PTS) glucose-specific EIICB protein, and over-expressing galP (encoding galactose permease; GalP) and glk (encodingglucokinase; GLK) to relieve carbon catabolite repression when glyceroland glucose were provided as co-substrates (Yang et al., 2018). Relativeto the expression of genes from the reductive branch of glycerol dis-similation from K. pneumoniae, coexpression of native yqhD and dhaBCEfrom C. freundii resulted in significantly improved 1,3-PDO production

(Zhang et al., 2006a, 2006b). On the other hand, a considerable amountof acetate was generated by engineered E. coli expressing dhaBCE anddhaFG from C. freundii and dhaT from K. pneumoniae, leading to cellgrowth inhibition and decreased 1,3-PDO production (Przystałowskaet al., 2015). Finally, coexpression of native yqhD with dhaB1 anddhaB2 from C. butyricum resulted in high-level 1,3-PDO production withthe titer reaching 104 g/L in an anaerobic fedbatch fermentation, al-though acetate accumulation eventually limited culture performance(Tang et al., 2009). This strain engineering approach is particularlyattractive, as the expression of DhaB1 from C. butyricum does not re-quire B12.

Although large-scale 1,3-PDO production using engineered E. colihas been demonstrated (Celińska, 2010), strategies to eliminate bio-processing limitations and further enhance culture performance requireto be developed. For example, complete growth inhibition of E. colioccurred for 1,3-PDO concentrations exceeding 120 g/L (Cameronet al., 1998), such that improving the 1,3-PDO tolerance of E. colithrough directed evolution or rational strain engineering may be aneffective strategy to improve 1,3-PDO production. Cell membrane en-gineering was applied to improve rates of biocatalysis in E. coli (Chen,2007), protein secretion (Cao et al., 2017) and biopolymer production

Table 4Summary of literature for E. coli strains engineered to produce 1,3-PDO and 3-HP from glycerol.

Parent strain Product Genetic strategy Titer (g/L)

Cultivation mode Reference

AG1 1,3-PDO Express dha regulon (K. pneumoniae) N/A Flask (Tong et al., 1991)ELC707 1,3-PDO Express dha regulon (C. freundii) 5 Flask (Daniel and Gottschalk, 1992)AG1 1,3-PDO Coexpress dhaB and dhaT (K. pneumoniae) from trc promoter 6 Fedbatch (Skraly et al., 1998)AG1 1,3-PDO -Express dha regulon (K. pneumoniae)

-Express glo1 (P. putida) to reduce MG formation2 Flask (Zhu et al., 2001)

AG1 1,3-PDO -Express dha regulon (K. pneumoniae)-Mutate glpK and feed fumarate to reduce G3P formation

4 Flask (Zhu et al., 2002)

JM109 1,3-PDO -Coexpress dhaB and gdrAB (K. pneumoniae), and yqhD-Feed succinate to decrease NAD+/NADH ratio

13 Fedbatch (Hong et al., 2015)

JM109 1,3-PDO Coexpress dhaBCE (C. freundii) and yqhD 41 Fedbatch (Zhang et al., 2006a, 2006b)BL21 1,3-PDO Coexpress dhaBCEFG (C. freundii) and dhaT (K. pneumoniae) 11 Fedbatch (Przystałowska et al., 2015)K-12 ER2925 1,3-PDO Coexpress dhaB12 (C. butyricum) and yqhD 104 Fedbatch (Tang et al., 2009)JM109 1,3-PDO -Coexpress yqhD, and dhaB and gdrAB (K. pneumoniae)

-Coexpress gapN (C. acetobutylicum) to generate NADPH, and fine-tuneexpression via 5’-UTR engineering-mutate ptsG and overexpress galP and glk to relieve catabolite repressionduring glycerol/glucose co-feeding

14 Flask (Yang et al., 2018)

BL21 3-HP Coexpress dhaB123 (K. pneumoniae) and aldH 0.6 Fedbatch (Raj et al., 2008)BL21 3-HP Coexpress dhaB123 (K. pneumoniae) and aldH from low-copy and high-copy

plasmids, respectively31 (Raj et al., 2009)

BL21 3-HP Coexpress dhaB and dhaR (L. brevis) and aldH 14 Flask (Kwak et al., 2013)W3110 3-HP -Coexpress dhaB123 and gdrAB from K. pneumoniae, and aldH

-Overexpress glpF to increase glycerol uptake-Mutate ackA-pta, yqhD, and glpR to reduce acetate, 3-HP, and G3P formation,respectively

42 Fedbatch (Jung et al., 2014)

BL21 3-HP -Coexpress dhaB and dhaR (L. brevis), and PSALDH (P. aeruginosa)-Mutate glpK and yqhD to reduce G3P and 3-HP formation, respectively

57 Fedbatch (Kim et al., 2014)

W 3-HP Coexpress dhaB and gdrAB (K. pneumoniae), and KGSADH (A. brasilense) 42 Fedbatch (Sankaranarayanan et al., 2014)W3110 3-HP -Coexpress dhaB and gdrAB (K. pneumoniae), and engineered gabD4 (C. necator)

with point mutations E209Q and E269Q-Mutate pta-ackA and yqhD to reduce acetate and 3-HP formation, respectively

72 Fedbatch (Chu et al., 2015)

BW25113 3-HP -Implement metabolic toggle switch to conditionally repress the expression ofgapA in ΔgapA-Coexpress dhaB123 and gdrAB (K. pneumoniae), and araE (A. brasilense)-Mutate yqhD and to reduce 1,3-PDO formation

6 Fedbatch (Tsuruno et al., 2015)

B (ATCC11303) 3-HP Coexpress dhaB123 and gdrAB (K. pneumoniae), KGSADH (A. brasilense), andpduPLW (K. pneumoniae)

5 Fedbatch (Honjo et al., 2015)

Gene products: ackA, acetate kinase; aldH, aldehyde dehydrogenase; araE, α-ketoglutaric semialdehyde dehydrogenase; dhaB123, glycerol dehydratase; dhaBCDEFG,glycerol dehydratase and reactivation factors; dhaBT, glycerol dehydratase and 1,3-propanediol oxidoreductase; dhaR, dihydroxyacetone reductase; dhaT, 1,3-propanediol oxidoreductase; gabD4, aldehyde dehydrogenase; gapA, glyceraldehyde 3-phosphate dehydrogenase; gdrAB, glycerol dehydratase reactivase; glo1,glyoxylase 1; glpF, glycerol uptake facilitator protein; glpK, glycerol kinase; glpR, glycerol-3-phosphate regulon repressor; KGSADH, alpha-ketoglutarate semialdehydedehydrogenase; pduPLW, aldehyde dehydrogenase, phosphate propanoyltransferase, and propionate kinase; PSALDH, semialdehyde dehydrogenase complex; pta,phosphotransacetylase; yqhD, NAD(P)H-dependent 1,3-propanediol dehydrogenase. Abbreviations: 3-HP, 3-hydroxypropionate; G3P, glycerol-3-phosphate; MG,methylglyoxal; NAD+, nicotinamide adenine dinucleotide.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

557

Page 21: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

(Westbrook et al., 2018) in B. subtilis, and solvent and acid tolerance inC. acetobutylicum (Zhao et al., 2003) and E. coli (Tan et al., 2017), and,hence, may be an effective strategy to enhance the 1,3-PDO tolerance ofE. coli. To restore the potential cofactor imbalance arising from theconversion of 3-HPA to 1,3-PDO by YqhD, native zwf, encodingNADP+-dependent glucose 6-phosphate-1-dehydrogenase (Zwf), can beoverexpressed to regenerate NADPH with low-level glucose supple-mentation. Moreover, the accumulation of acetate to inhibitory levelsobserved during high-level 1,3-PDO production (Tang et al., 2009) maybe relieved by mutation of ptsH, encoding the histidine protein of thesugar PTS that participates in the phosphorylation of DHA by DhaKLM,which significantly reduced acetate formation during anaerobic succi-nate production from glycerol (Zhang et al., 2010).

3.5.2.2. 3-HP. Strain engineering strategies to enhance 3-HPproduction in E. coli have primarily focused on balancing theactivities of GDHt and ALDH in the synthetic 3-HP pathway (Fig. 1),and limiting side metabolite formation. Low-level 3-HP production wasfirst demonstrated in E. coli via coexpression of DhaB from K.pneumoniae with one of four ALDHs, of which AldH4 from S.cerevisiae, encoded by aldH4, was most effective for converting 3-HPAto 3-HP (Suthers and Cameron, 2005). Excessive 3-HPA accumulationwas observed to hinder 3-HP production in a strain coexpressingdhaB123 from K. pneumoniae and native aldH (encoding AldH) fromindividual low-copy plasmids, due to an imbalance between theactivities of DhaB and AldH (Raj et al., 2008). Subsequent workrevealed that the activity of GdrAB was critical to prevent theinactivation of DhaB, and the coexpression of native aldH, anddhaB123 and gdrAB from K. pneumoniae improved the 3-HP titer by6-fold (Rathnasingh et al., 2009) relative to the corresponding strainwithout gdrAB coexpression (Raj et al., 2008). Moreover, the expressionof KGSADH from A. brasilense, in place of native aldH, further increasedthe 3-HP titer by 2.5-fold (Rathnasingh et al., 2009). Fine-tuning theexpression of dhaB123 from K. pneumonia by engineering its 5’- UTRalso served to manage 3-HPA accumulation (Lim et al., 2016). KGSADHfrom A. brasilense was expressed from a relatively strong promoter in ahighly 3-HPA-tolerant E. coli strain that also expressed dhaB123 withengineered 5’-UTRs from a separate promoter, resulting in a substantialincrease in the 3-HP yield (Lim et al., 2016). Lastly, parallel expressionof the partial PDU pathway (i.e. pduPLW) from K. pneumoniae and araE,encoding α-ketoglutaric semialdehyde dehydrogenase from A.brasilense (AraE), in a strain also expressing dhaB123 and gdrAB fromK. pneumoniae resulted in a significantly increased 3-HP titer, comparedto a strain that did not express pduPLW, suggesting that conversion of 3-HPA to 1,3-PDO may have decreased due to increased 3-HPA fluxthrough the partial PDU pathway (Honjo et al., 2015).

Several strategies were explored to minimize the formation of majorside metabolites, including acetate, lactate, succinate, and 1,3-PDO, forenhanced 3-HP production in engineered E. coli. Inactivating ackA-ptaand yqhD in a strain coexpressing dhaB123 and gdrAB from K. pneu-moniae and native aldH, significantly decreased acetate and 1,3-PDOproduction (Jung et al., 2014). 3-HP production increased by 2-fold inthe ΔackA Δpta ΔyqhD triple mutant compared to the parent straincoexpressing dhaB123, gdrAB, and aldH, while mutation of glpR, en-coding the repressor of the glp regulon (GlpR), and overexpression ofglpF modestly enhanced 3-HP production (Jung et al., 2014). On theother hand, inactivating respiratory glycerol dissimilation via mutationof glpK significantly improved 3-HP production in a strain coexpressingdhaB and dhaR, encoding the reactivation factor of DhaB (DhaR), fromL. brevis, and native aldH (Kim et al., 2014; Kwak et al., 2013). More-over, mutation of yqhD significantly improved 3-HP production in theΔglpK mutant, and the expression of PSALDH, encoding semialdehydedehydrogenase from Pseudomonas aeruginosa, in place of aldH, furtherenhanced culture performance (Kim et al., 2014). In silico analysis ofgene mutations based on a genome-scale metabolic model identifiedtpiA, encoding triosephosphate isomerase (TpiA), and zwf as targets for

inactivation to enhance 3-HP production from glycerol (Tokuyamaet al., 2014). Mutation of tpiA in a strain coexpressing dhaB123 andgdrAB from K. pneumoniae and native aldH increased the 3-HP titer by2-fold, while mutation of zwf had only a minor effect on culture per-formance. Moreover, mutation of yqhD significantly reduced 1,3-PDOformation and, in turn, 3-HP production was markedly increased,compared to the ΔtpiA Δzwf double mutant coexpressing dhaB123,gdrAB, and aldH (Tokuyama et al., 2014). Similarly, mutation of ackA-pta and yqhD significantly reduced acetate and 1,3-PDO formation, witha concomitant increase in 3-HP production in a strain coexpressingdhaB123 and gdrAB from K. pneumoniae and native aldH (Chu et al.,2015). Moreover, 3-HP production was further enhanced by expressinggabD4, encoding ALDH from C. necator (GabD4), in place of native aldH,while expression of engineered GabD4, with point mutations E209Qand E269Q, significantly improved 3-HP production during a fedbatchco-cultivation using glycerol and glucose (Chu et al., 2015). Finally,repressing the expression of genes in central metabolic pathways canredirect carbon flux toward target metabolite production withoutcompromising host cell viability. A previously engineered metabolictoggle switch (Gardner et al., 2000) was used to conditionally repressthe expression of gapA, encoding glyceraldehyde 3-phosphate dehy-drogenase (GapA), in a ΔgapA mutant coexpressing dhaB123 and gdrABfrom K. pneumoniae and araE from A. brasilense to redirect carbon fluxfrom glycolysis to the synthetic 3-HP pathway, while minimizinggrowth inhibition observed in ΔgapA mutants (Tsuruno et al., 2015).The conditional repression of gapA expression resulted in a significantincrease in 3-HP production compared to the strain coexpressingdhaB123, gdrAB, and araE, whereas mutation of yqhD dramatically re-duced 1,3-PDO formation and, in turn, improved the 3-HP titer andyield (Tsuruno et al., 2015).

Other technical issues to be addressed for enhanced 3-HP produc-tion from glycerol in E. coli include recycling of NADH and cir-cumventing the requirement of B12-dependent GDHt. Excess NADH mayaccumulate during 3-HP production, potentially resulting in the for-mation of reduced side metabolites. While NADH can be oxidized byNOX using molecular oxygen without diverting carbon from 3-HPproduction, B12-independent DhaB1 from C. butyricum may not functionin the presence of oxygen (Andreeßen et al., 2010; Tang et al., 2009),such that a NAD+ regeneration system functional under anaerobicconditions would be required in this context. NADH:quinone oxidor-eductase II, encoded by ndh in E. coli, oxidizes NADH during nitraterespiration (Tran et al., 1997), such that coexpression of native ndh, anddhaB1 and dhaB2 from C. butyricum (in addition to ALDH) may facil-itate anaerobic production of 3-HP without B12 if exogenous nitrate isprovided. Alternatively, protein engineering of DhaB1 to reduce itssensitivity to oxygen may enable B12-independent 3-HP productionunder microaerobic conditions. A similar strategy was employed foraerobic 1-propanol production through the L-threonine degradationpathway, whereby engineered aero-tolerant AdhE successively con-verted propionyl-CoA to propionaldehyde and 1-propanol (Choi et al.,2012). In the absence of B12 dependency, E. coli ΔyqhD mutants aresuperior hosts for 3-HP production relative to other microbial hostsdiscussed herein due to the minimal formation of 1,3-PDO.

3.5.2.3. Bio(co)polymers. PHAs are a family of biologically producedpolyesters composed of various (R)-hydroxycarboxylates as themonomers (Mozzi et al., 2006; Teeka et al., 2012) and have emergedas substitutes for conventional plastics due to their thermoplastic andelastomer properties, biodegradability, non-toxicity, and structuraldiversity (Leong et al., 2014; Reddy et al., 2003; Teeka et al., 2012).Unlike most PHAs, P(3HP) is not naturally produced by bacteria(Linares-Pastén et al., 2015), leading to the exploration of syntheticpathways in engineered hosts such as E. coli. Marginal P(3HP)production was achieved in E. coli coexpressing dhaB1 from C.butyricum, pduP from Salmonella enterica, and phaC from C. necator,from either pure or crude glycerol (Andreeßen et al., 2010). On the

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

558

Page 22: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

other hand, coexpressing dhaB123 and gdrAB from K. pneumoniae withpduP from S. typhimurium and phaC from C. necator resulted in a highspecific P(3HP) yield using glucose and glycerol as co-substrates in afedbatch cultivation (Wang et al., 2013a, 2013b). In spite of the highspecific P(3HP) yield, plasmid instability was an issue that wasaddressed using a plasmid addiction system for the coexpression ofpduP from S. typhimurium and phaC from C. necator, and chromosomalexpression of dhaB123 and gdrAB from K. pneumoniae, resulting in ahigh specific P(3HP) yield of 68% of dry cell weight (dcw) in a fedbatchcultivation (Gao et al., 2014). To circumvent the requirement of B12-dependent GDHt, P(3HP) was produced via the β-alanine pathway in E.coli (Wang et al., 2014a, 2014b, 2014c). In this engineered redox-neutral pathway, native L-aspartate decarboxylase (i.e. PanD encodedby panD) generated β-alanine that was converted to P(3HP) throughsubsequent actions of a β-alanine-pyruvate transaminase from P. putida(i.e. GabT encoded by gabT), native 3-hydroxyacid dehydrogenase (i.e.YdfG encoded by ydfG) and PrpE, and PhaC from C. necator. Althoughexogenous B12 was not required, P(3HP) production was low (Wanget al., 2014a, 2014b, 2014c), with only a modest improvement obtainedby expressing panD from Corynebacterium glutamicum in place of nativepanD (Lacmata et al., 2017). Lastly, engineered E. coli was used in atwo-stage cultivation with resting cells of L. reuteri for P(3HP)production from glycerol (Linares-Pastén et al., 2015). Due to a high3-HPA tolerance, L. reuteri was used in the first stage to convert glycerolto 3-HPA, which was, in turn, used as a substrate by E. coli coexpressingpduP from L. reuteri (Sabet-Azad et al., 2013) and phaC fromChromobacterium sp. (Bhubalan et al., 2011), leading to a highspecific P(3HP) yield (Linares-Pastén et al., 2015). Although notdirectly related to engineering of glycerol metabolism, other studieshave shown the potential of using pure or crude glycerol for theproduction of poly(3-hydroxybutyrate) (de Almeida et al., 2007;Ganesh et al., 2015; Shah et al., 2014; Shalel Levanon et al., 2005).

Due to high crystallinity, homopolymer PHAs are too brittle and stifffor many industrial applications (Leong et al., 2014), and such technicaldrawbacks can be overcome by the use of copolymers (Singh andMohanty, 2007). P(3HB-co-3HV) was produced in engineered E. coliwith a high specific yield of 80% of dcw via coexpression of bktB, en-coding β-ketothiolase (BktB), phaB, encoding acetoacetyl-CoA re-ductase (PhaB), and phaC, from R. eutropha, in addition to native sucCD,using glycerol as the primary carbon source with supplemental succi-nate and propionate (Bhatia et al., 2015). Previous studies have re-ported high C3 metabolite production from glycerol due to moreNADPH and ATP generation relative to glucose, favoring threoninebiosynthesis and directing carbon flux toward propionyl-CoA, a pre-cursor to (R)-3-hydroxyvaleryl-CoA [(R)-3-HV-CoA] (Tseng et al.,2010). High-level propionyl-CoA and, in turn, 1-propanol productionwas achieved from glycerol without supplemental succinate and pro-pionate by activating the native genomic sleeping beauty mutase (Sbm)operon (Srirangan et al., 2013; Srirangan et al., 2014). This approachwas extended to P(3HB-co-3HV) production by mutating dhaK andcoexpressing bktB and phaCAB from C. necator in the propanogenic E.coli strain, resulting in a high specific P(3HB-co-3HV) yield of 66% ofdcw (Srirangan et al., 2016b). Moreover, inactivating respiratory gly-cerol dissimilation via mutation of glpD in the strain coexpressing bktBand phaCAB increased the 3-hydroxyvalerate (3-HV) content by 2.6-fold to 18.5 mol%, which is the highest reported level for E. coli-based P(3HB-co-3HV) biosynthesis using an unrelated carbon source (Sriranganet al., 2016b). Lastly, poly(3-hydroxyburyrate-co-3-hydroxyhexanoate)[P(3HB-co-3HH)] was synthesized in E. coli coexpressing phaAB from C.necator, and phaJ (encoding (R)-specific enoyl-CoA hydratase; PhaJ)and phaC from Aeromonas hydrophila, although P(3HB-co-3HH) pro-duction was low on pure glycerol (Phithakrotchanakoon et al., 2013)and was negligible when crude glycerol was used(Phithakrotchanakoon et al., 2015).

As is the case for 3-HP production, the regeneration of reducingequivalents and elimination of the B12-reliance for functional

expression of GDHt are technical challenges to overcome for enhancedbiopolymer production in E. coli. Excess NADH can accumulate during P(3HP) production due to the conversion of 3-HPA to 3-HP-CoA viaPduP, potentially hindering culture performance (Andreeßen et al.,2010), while NADPH is consumed through the conversion of β-keto-valeryl-CoA to 3-HV-CoA by PhaB during P(3HB-co-3HV) production.Accordingly, the same strategies proposed to recycle reducing equiva-lents during B12-independent 3-HP and 1,3-PDO production may beeffective for P(3HP) and P(3HB-co-3HV) production, respectively.Protein engineering is also a potential strategy to alter bio(co)polymercomposition. Expression of engineered PhaJ from A. hydrophila con-taining a V130A point mutation substantially increased the 3-hydro-xyhexanoate (3-HH) fraction in P(3HB-co-3HH) produced in E. coli,without compromising the specific P(3HB-co-3HH) yield (Hu et al.,2007). Moreover, bio(co)polymer content can be modified by manip-ulating glycerol dissimilation through either the fermentative or re-spiratory pathways as previously noted (Srirangan et al., 2016b).

3.5.2.4. Other products. Production of 1-propanol from glycerol in E.coli was demonstrated using two distinct pathways. In the firstapproach, the natural 1,2-PDO pathway of E. coli was modified usingan ATP-dependent DHAK, encoded by dhaK from S. blattae, andextended using components of the pdu operon from K. pneumoniae(Matsubara et al., 2016). 1,2-PDO was converted to propionaldehydevia PduCDE with the reactivation factor PduGH, and then to 1-propanolvia PduQ. In contrast to native PEP-dependent DhaKLM, ATP-dependent DHAK decreased reliance on pyruvate production and, inturn, increased flux into the 1-propanol pathway (Matsubara et al.,2016). The second strategy to achieve 1-propanol production fromglycerol involved the activation of the dormant Sbm operon of E. coli(Srirangan et al., 2013; Srirangan et al., 2014). This pathway beginswith the conversion of succinyl-CoA to L-methylmalonyl-CoA via amethylmalonyl-CoA mutase (Sbm encoded by sbm), followed by theformation of propionyl-CoA from succinyl-CoA by a methylmalonyl-CoA decarboxylase (YqfG encoded by yqfG), with a propionyl-CoA::succinate transferase (YgfH encoded by ygfH) facilitating theinterconversion between succinyl-CoA and propionyl-CoA. Theactivation of the Sbm operon introduced an intracellular carbon fluxcompetition between the C2-fermentative pathway (with ethanol andacetate as the major metabolites) and C3-fermentative pathway (with 1-propanol and propionate as the major metabolites), and glycerol wasidentified as a suitable carbon source favoring direction of thedissimilated carbon flux toward the C3-fermentative pathway(Srirangan et al., 2014). The propionyl-CoA derived from the Sbmoperon and acetyl-CoA were fused by either PhaA or BktB from C.necator to form 3-ketovaleryl-CoA, which was further converted tobutanone via a ketone-formation pathway containing acetoacetyl-CoA:acetate/butyrate:CoA transferase (i.e. CtfAB encoded by ctfAB)and acetoacetate decarboxylase (i.e. Adc encoded by adc) from C.acetobutylicum (Akawi et al., 2015; Srirangan et al., 2016a).Manipulating various genes involved in the glycerol dissimilationpathway potentially enhanced the production of propionate (Akawiet al., 2015) and butanone (Srirangan et al., 2016a) from glycerol undervarious fermentation conditions.

Small amounts (up to 154 mg/L) of 1-butanol were produced fromglycerol by reconstructing the 1-butanol pathway from C. acet-obutylicum in E. coli (Zhou et al., 2014). To use glycerol as a carbonsource, expression of native fdh1, encoding FDH, and inactivation ofadhE, ldhA, and frdBC associated with side metabolite production aimedto overcome the NADH deficit resulting from 1-butanol synthesis (Zhouet al., 2014). Significantly higher 1-butanol titers of up to 6.9 g/L wereobtained from crude glycerol in a ΔadhE ΔldhA ΔfrdA Δpta quadruple E.coli mutant expressing the 1-butanol pathway from C. acetobutylicum,with phaA from C. necator and ter from T. denticola expressed in place ofthil and bcd-etfAB, respectively, by engineering parallel NADH re-generation systems (Saini et al., 2017). Coexpression of the engineered

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

559

Page 23: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

native NADH-insensitive PDH complex (encoded by aceEF-lpdA⁎), andnative pgl (encoding 6-phosphogluconolactonase; PGL) and zwf re-spectively stimulated the decarboxylation of pyruvate to acetyl-CoA,and increased flux through the PPP to regenerate NADPH, which wasconverted to NADH via overexpressed native UDH (encoded by udhA),resulting in a significant increase in the 1-butanol titer. Similarly, co-expression of native gldA and dhaKLM presumably increased NADHlevels and, in turn, 1-butanol production, which could be further en-hanced by restricting carbon flux into the TCA cycle via promoter en-gineering to reduce the expression of gltA to conserve acetyl-CoA (Sainiet al., 2017).

E. coli was also engineered to produce 2,3-BDO isomers from gly-cerol. For the production of (R,R)-2,3-BDO, alsS, encoding α-acet-olactate synthase (AlsS) that converts pyruvate to (S)-2-acetolactate,and alsD, encoding α-acetolactate decarboxylase (AlsD) that converts(S)-2-acetolactate to (R)-acetoin, from B. subtilis were coexpressed withadh, encoding secondary alcohol dehydrogenase that converts (R)-acetoin to (R,R)-2,3-BDO from C. beijerinckii, resulting in a (R,R)-2,3-BDO titer of 9.5 g/L (Shen et al., 2012). Similarly, coexpression of budA,encoding α-acetolactate decarboxylase (BudA) that converts (S)-2-acetolactate to (R)-acetoin, and budC, encoding 2,3-BDO dehy-drogenase/acetoin reductase that converts (R)-acetoin to meso-2,3-BDO, from K. pneumoniae resulted in a meso-2,3-BDO titer of 6.9 g/Lfrom crude glycerol (Lee et al., 2012).

Production of up to 6.2 g/L 2-hydroxyisovalerate (an industriallyimportant surfactant, emulsifier, and intermediate for the synthesis ofpolymers, solvents, and fertilizers) from glycerol was achieved in E. coliby exploiting the native L-valine biosynthetic pathway (Cheong et al.,2018). First, two pyruvate molecules were condensed to generate 2-acetolactate by AlsS from B. subtilis, followed by its sequential con-version to 2,3-dihydroxyisovalerate, 2-ketoisovalerate, and finally 2-hydroxyisovalerate by native acetohydroxy-acid isomeroreductase (i.e.,IlvC encoded by ilvC), native IlvD, and 2-hydroxyacid dehydrogenase(i.e., PanE endoded by panE) from L. lactis, respectively. Inactivation ofcompeting mixed-acid pathways resulted in a significant decrease in 2-hydroxyisovalerate production, highlighting the importance of thesepathways for maintaining the overall intracellular redox balance(Cheong et al., 2018).

β-carotene production from glycerol was achieved in E. coli byconstructing an alternative glycerol utilization pathway in a previouslydeveloped strain (Ye et al., 2016) (Guo et al., 2018). Glycerol was se-quentially converted to D-glyceraldehyde and D-glycerate via aldehydereductase (i.e., Alrd encoded by alrd) and ALDH (i.e., AldH encoded byaldH) from C. necator, respectively, and 2-phospho-D-glycerate wasderived from D-glycerate by a native glycerol kinase (i.e., GarK encodedby garK). The synthetic glycerol utilization pathway simultaneouslyincreased the rate of glycerol consumption and the supply of earlyisoprenoid precursors, i.e., G3P and pyruvate, enabling the productionof 75 mg/L of β-carotene (Guo et al., 2018).

Production of 5-hydroxytryptophan (5-HTP), a precursor of ser-otonin used to treat depression, insomnia, and headaches, from glycerolwas recently demonstrated in E. coli (Wang et al., 2018). As a precursorof 5-HTP biosynthesis, tetrahydrobiopterin (BH4) was produced in athree-step pathway via GTP cyclohydrolase (GCHI), 6-pyruvate-tetra-hydropterin synthase (PTPS), and sepiapterin reductase (SPR), encodedby mtrA from B. subtilis, human PTPS, and human SPR, respectively. 5-HTP was synthesized from BH4 and L-tryptophan by tryptophan hy-droxylase I or II, i.e., TPH1/2 encoded by human TPH1/2, and BH4 wasregenerated via a two-step pathway comprised of pterin-4α-carbinola-mine dehydratase (PCD) and dihydropteridine reductase (DHPR), en-coded by human PCD and DHPR, respectively (Wang et al., 2018). Toavoid the requirement for exogenous L-tryptophan supplementation,the native L-tryptophan biosynthetic pathway, including engineeredfeedback resistant trpE⁎ (encoding tryptophan-insensitive anthranilatesynthase subunit TrpE⁎) and aroH⁎ (encoding tryptophan-insensitive 3-deoxy-7-phosphoheptulonate synthase AroH⁎), were coexpressed in the

5-HTP-producing strain. 5-HTP production was further enhanced byexpressing engineered TPH2, i.e. TPH145 containing deletions of 145N-terminal and 24 C-terminal amino acids, and by reducing the copynumber of the plasmid containing L-tryptophan biosynthesis genes andselecting different promoters to drive the synthesis and regeneration ofBH4. These strain engineering strategies resulted in the highest 5-HTPtiter ever reported for bio-based production (5.1 g/L) (Wang et al.,2018).

L-3,4-dihydroxyphenylalanine (L-DOPA), a precursor of dopamineused in the treatment of Parkinson’s disease and dopamine-responsivedystonia, was synthesized from glycerol with an E. coli strain expressinga modified Shikimate pathway (Das et al., 2018). L-DOPA is derivedfrom L-tyrosine, which is the major effector of TyrR (L-tyrosine re-pressor encoded by tyrR)-based repression of aromatic amino acidbiosynthesis (Keseler et al., 2016), such that tyrR was mutated to in-crease L-tyrosine levels. To further boost L-tyrosine overproduction,feedback- resistant native 3-deoxy-D-arabino-heptulosonate synthase(i.e., DAHP⁎ encoded by aroG⁎), shikimate kinase (i.e., AroL encoded byaroL), chorismate synthase (i.e., AroC encoded by aroC), 3-phos-phoshikimate 1-carboxyvinyltransferase (i.e., AroA encoded by aroA),and L-tyrosine aminotransferase (i.e., TyrB encoded by tyrB) were co-expressed in a tyrR mutant also containing mutations to pykF (encodingpyruvate kinase I; PykF) and serA (encoding phosphoglycerate dehy-drogenase; SerA) to block competing amino acid biosynthetic path-ways. The direct conversion of L-tyrosine to L-DOPA was achieved bycoexpressing native 4-hydroxyphenylacetate-3-hydrolase (i.e., HpaBCencoded by hpaBC) in the L-tyrosine overproducing mutant, resulting inthe production of up to 12.5 g/L L-DOPA, with a marked reduction inacetate accumulation compared to previous studies in which L-DOPAwas derived from glucose (Das et al., 2018).

Itaconate is a versatile biological used in the production of resins,plastics, rubbers, paints, surfactants, and lubricants (Dwiarti et al.,2007). Itaconate production from glycerol was demonstrated in E. coliexpressing engineered cis-aconitate decarboxylase from Aspergillus ter-reus (CadA encoded by cadA), which converts cis-aconitate (a TCA cycleintermediate) to itaconate (Jeon et al., 2016). A library of synonymouscodon variants was constructed by introducing random point mutationsin the first 10 codons of CadA, and inclusion body formation was nearlyabolished when certain engineered CadA variants were expressed, re-sulting in up to 4-fold increases in the itaconate titer, which reached 7.2g/L in a fedbatch cultivation (Jeon et al., 2016). Moreover, inactivationof icd and coexpression of codon optimized cadA from A. terreus, pyc(encoding pyruvate carboxylase; Pyc) from C. glutamicum, and nativeacnB (encoding aconitase; AcnB) and gltA resulted in a high itaconatetiter of 43 g/L in a fedbatch cultivation (Chang et al., 2017).

Deoxyviolacein, a drug with antagonistic activity against tumors,Gram-positive bacteria, and fungal plant pathogens (Rodrigues et al.,2014), was first produced from glucose in engineered E. coli expressinga truncated deoxyviolacein gene cluster, i.e. vioABCE with vioD omitted,from C. violaceum (Rodrigues et al., 2013). The deoxyviolacein-produ-cing strain was further modified via inactivation of araBAD, eliminatingL-arabinose metabolism, and, among various carbon sources, glycerolwas determined to be most effective for deoxyviolacein production upto 1.6 g/L (Rodrigues et al., 2014).

Valerenadiene is an anxiolytic compound derived from the medic-inal herb Valeriana officinalis, and has been produced in small quantitiesin engineered E. coli (Nybo et al., 2017). Coexpression of codon opti-mized vds, encoding valerenadiene synthase from V. officinalis (VDS),and genes comprising the mevalonic acid pathway from S. cerevisiaeresulted in a valerenadiene titer of 62 mg/L from glycerol (Nybo et al.,2017).

Small amounts (up to 38 mg/L) of toxic acrylic acid was synthesizedfrom glycerol using a previously constructed P(3HP)-producing mutant,in which 3-HP-CoA was derived via coexpression of dhaB123 and gdrABfrom K. pneumoniae and pduP from S. typhimurium (Tong et al., 2016).Acrylyl-CoA was derived from 3-HP-CoA via an enoyl-CoA hydratase

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

560

Page 24: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

from the propionyl-CoA synthase complex of Chloroflexus aurantiacusand the CoA moiety was subsequently removed by a putative en-dogenous CoA transferase (Tong et al., 2016).

Serinol production (up to 3.3 g/L) from glycerol was achieved viaexpression of a bifunctional DHAP aminotransferase/dihy-drorhizobitoxine synthase from Bradyrhizobium elkanii, which mediatesDHAP conversion to serinol phosphate and subsequent depho-sphorylation (Andreeßen and Steinbüchel, 2012).

4. 4.0 Future perspectives

4.1. Selecting the appropriate production host for glycerol biorefinery

Factors associated with the selection of a proper microbial platformfor target metabolite production from glycerol include growth re-quirements, product tolerance, pathogenicity, and genetic tractability.While K. pneumoniae is genetically amenable and has a high capacity forthe conversion of glycerol into diols and organic acids with minimalnutritional requirements, the pathogenicity limits its application tobiomanufacturing, particularly for human consumption products. Onthe other hand, E. coli possesses all of the same positive attributes, witheven greater ease of genetic manipulation and a proven track record inbiomanufacturing, and has been engineered to dissimilate glycerol intomany non-native metabolites with similar efficiency to natural produ-cers. For example, using E. coli, 1,3-PDO is produced commercially fromglucose (Celińska, 2010), and B12-independent production of 1,3-PDOfrom glycerol has been achieved on a similar scale (Tang et al., 2009).Moreover, side metabolite formation appears less problematic in E. coli,potentially due to its relatively stable TCA cycle, and the insignificant2,3-BDO production during glycerol fermentation can simplify pur-ification of 1,3-PDO. E. coli is also an ideal candidate for 3-HP pro-duction due to its high production capacity and lack of redundant 1,3-PDOORs, although the engineered strains often express B12-dependentGDHt which is a technical challenge difficult to overcome.

Certain Clostridium sp. are particularly well suited to the productionof specific metabolites from glycerol, e.g. C. pasteurianum and C. bu-tyricum as 1-butanol and 1,3-PDO producers, respectively. However,they are obligate anaerobes that are recalcitrant to genetic manipula-tion and, thus, may be restricted to the production of native productswith limited strain improvement. While Citrobacter sp. have not beenwell characterized and, therefore, lack advanced genetic tools, certainstrains are attractive for glycerol biorefinery. In particular, C. werkmaniireadily utilizes co-substrates with a preference for glycerol, such thatblocking oxidative pathways for glycerol dissimilation to channel thecarbon flux into the reductive branch as well as engineering of cofactorregeneration systems that exploit glucose metabolism may be effectivestrategies in this host for target metabolite formation. Moreover, ex-ceptionally high 1,3-PDO yields from co-substrates have been reportedin cultures of C. werkmanii relative to common 1,3-PDO producer suchas Klebsiella sp. (Maervoet et al., 2012a, 2012b). Finally, Lactobacillussp. are highly alcohol and acid tolerant and many have been grantedGRAS status, making them suitable hosts for the production of biofuelsand organic acids (Bosma et al., 2017). Moreover, the genomes ofLactobacillus sp. are substantially smaller than other bacteria discussedherein (Bosma et al., 2017). For example, K. pneumoniae MGH 78578has a 5.32 Mb genome encoding 4,996 proteins, whereas L. reuteri DSM20016 has a 2 Mb genome encoding 1,917 proteins. Accordingly, it isanticipated that minimizing side metabolite formation or overcomingcomplex regulatory mechanisms that control flux distribution would besignificantly easier in L. reuteri. Note that certain species of this genusare fastidious in their nutritional requirements, although other speciessuch as L. reuteri can proliferate in less complex media (Bosma et al.,2017). Overall, while E. coli is the most preferred host for bio-basedproduction from glycerol, certain organisms are inherently specializedfor the production of specific metabolites. Hence, the development ofmore sophisticated genetic tools and comprehensive understanding of

global metabolism may facilitate the derivation of superior productionhosts.

4.2. Strategies to enhance target metabolite production from glycerol

General strategies to enhance target metabolite production fromglycerol are universal among the hosts discussed herein, although thespecific approach and relative importance depend on the strain andproduct. For example, further improvements to 1,3-PDO production inE. coli or K. pneumoniae will likely hinge on improving the producttolerance of these strains and reducing inhibitory metabolite formation.In general, improving product tolerance may be achieved via cellmembrane engineering (Tan et al., 2017; Zhao et al., 2003), directedevolution (Sandoval et al., 2015), or modulating the host stress re-sponse via overexpression of heat shock proteins (Zingaro andPapoutsakis, 2012). Reducing side metabolite formation in E. coliduring 1,3-PDO production can be effectively achieved by inactivatingkey genes (e.g. ptsH), whereas engineering of the TCA cycle in K.pneumoniae may be required to alleviate pyruvate accumulation in 2,3-BDO-deficient mutants. On the other hand, engineering cofactor re-generation systems that rely on co-substrate utilization should be a keyapproach to improve 1,3-PDO production in Citrobacter sp. and Lacto-bacillus sp. In the case of Lactobacillus sp., partially activating dormantfermentative or respiratory pathways for oxidative glycerol dissimila-tion may also serve to restore redox poise.

A major disadvantage in using E. coli for the production of 3-HP andP(3HP), or other oxidized metabolites derived from 3-HPA, is the re-quirement for exogenous B12. Biosynthesis of B12 in E. coli required thecoexpression of 22 genes contained in six operons and three plasmids(Ko et al., 2014), which, if implemented in a strain engineered toproduce a second major metabolite (e.g. 3-HP), could pose a largeburden on the host cell and compromise culture performance. As pre-viously outlined, NAD+ can be regenerated anaerobically via NADH:-quinone oxidoreductase II with supplemental nitrate, which may fa-cilitate 3-HP and P(3HP) production in strains expressing B12-independent DhaB1 from C. butyricum. Alternatively, protein en-gineering of DhaB1 to reduce or abolish its sensitivity to oxygen maycircumvent the requirement of anaerobic operation, and this was suc-cessfully achieved with AdhE from E. coli for aerobic 1-propanol pro-duction (Choi et al., 2012). Yet, another possibility is the selection of aproduction host that can naturally synthesize B12 such as Citrobacter sp.,certain Lactobacillus sp., and K. pneumoniae. While K. pneumoniae is apathogen, attenuated strains lacking endogenous endotoxins have beendeveloped through rational strain design (Huynh et al., 2015; Junget al., 2013), which is an approach that led to the development of non-hemolytic Streptococcus equisimilis for commercial hyaluronic acidproduction (Kim et al., 1996; Liu et al., 2011), and endotoxin-free E. colifor recombinant protein production (Mamat et al., 2015). Finally, theapplication of mixed microbial cultures is another potential strategy toenhance the bioconversion of glycerol to value-added products, and hasbeen studied extensively for the production of 1,3-PDO (Jiang et al.,2018; Jiang et al., 2017; Sun et al., 2018; Zhou et al., 2017a, 2017b;Zhou et al., 2018).

While the strategies proposed herein may resolve certain limitationsidentified through various strain engineering approaches, a systemsbiology approach that incorporates metabolomics, proteomics, andtranscriptomics may serve to elucidate the multiple layers of regulatoryarchitecture that orchestrate glycerol metabolism. Evidence suggeststhat regulatory mechanisms such as post-transcriptional regulation (i.e.via non-coding RNAs), post-translational regulation (e.g. phosphoryla-tion), and allosteric regulation significantly affect metabolic flux dis-tribution, which is another important topic that has recently been re-viewed (Liu et al., 2017). A systems level approach combined with next-generation genome editing tools, such as the CRISPR-Cas9 systemwhich has been implemented in various microorganisms (Choi and Lee,2016), is a logical step toward engineering microbial platforms for

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

561

Page 25: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

industrial-scale glycerol biorefinery. However, the functionality of theCRISPR-Cas9 system in a given host might not necessarily result indramatic improvements to the frequency with which desired genomeediting events occur. Moreover, it has been our experience that higherorder multiplexing of gene mutations may be limited by the efficiencyof the host recombination system, such that down-regulating the ex-pression of gene targets via CRISPRi (Qi et al., 2013), which reliesprimarily on the binding affinity of dCas9 to its assigned target, may bea more effective approach to large-scale pathway modulation.

Acknowledgements

Natural Sciences and Engineering Research Council, Canada.

References

Abbad-Andaloussi, S., Manginot-Durr, C., Amine, J., Petitdemange, E., Petitdemange, H.,1995. Isolation and characterization of Clostridium butyricum DSM 5431 mutants withincreased resistance to 1,3-Propanediol and altered production of acids. Applied andEnvironmental Microbiology 61 (12), 4413–4417.

Ackerman, R.S., Cozzarelli, N.R., Epstein, W., 1974. Accumulation of toxic concentrationsof methylglyoxal by wild-type Escherichia coli K-12. Journal of Bacteriology 119 (2),357–362.

Ahrens, K., Menzel, K., Zeng, A.P., Deckwer, W.D., 1998. Kinetic, dynamic, and pathwaystudies of glycerol metabolism by Klebsiella pneumoniae in anaerobic continuousculture: III. Enzymes and fluxes of glycerol dissimilation and 1,3-propanediol for-mation. Biotechnology and Bioengineering 59 (5), 544–552.

Ainala, S.K., Ashok, S., Ko, Y., Park, S., 2013. Glycerol assimilation and production of 1,3-propanediol by Citrobacter amalonaticus Y19. Applied Microbiology andBiotechnology 97 (11), 5001–5011.

Akawi, L., Srirangan, K., Liu, X., Moo-Young, M., Perry Chou, C., 2015. EngineeringEscherichia coli for high-level production of propionate. Journal of IndustrialMicrobiology & Biotechnology 42 (7), 1057–1072.

Andreeßen, B., Steinbüchel, A., 2012. Biotechnological conversion of glycerol to 2-amino-1, 3-propanediol (serinol) in recombinant Escherichia coli. Applied microbiology andbiotechnology 93 (1), 357–365.

Andreeßen, B., Lange, A.B., Robenek, H., Steinbüchel, A., 2010. Conversion of glycerol topoly(3-hydroxypropionate) in recombinant Escherichia coli. Applied andEnvironmental Microbiology 76 (2), 622–626.

Ashok, S., Raj, S.M., Rathnasingh, C., Park, S., 2011. Development of recombinantKlebsiella pneumoniae ∆dhaT strain for the co-production of 3-hydroxypropionicacid and 1,3-propanediol from glycerol. Applied Microbiology and Biotechnology 90(4), 1253–1265.

Ashok, S., Mohan Raj, S., Ko, Y., Sankaranarayanan, M., Zhou, S., Kumar, V., Park, S.,2013a. Effect of puuC overexpression and nitrate addition on glycerol metabolismand anaerobic 3-hydroxypropionic acid production in recombinant Klebsiella pneu-moniae ΔglpKΔdhaT. Metabolic Engineering 15, 10–24.

Ashok, S., Sankaranarayanan, M., Ko, Y., Jae, K.-E., Ainala, S.K., Kumar, V., Park, S.,2013b. Production of 3-hydroxypropionic acid from glycerol by recombinantKlebsiella pneumoniae ΔdhaTΔyqhD which can produce vitamin B12 naturally.Biotechnology and Bioengineering 110 (2), 511–524.

Atsumi, S., Cann, A.F., Connor, M.R., Shen, C.R., Smith, K.M., Brynildsen, M.P., Chou,K.J., Hanai, T., Liao, J.C., 2008. Metabolic engineering of Escherichia coli for 1-bu-tanol production. Metabolic engineering 10 (6), 305–311.

Bächler, C., Schneider, P., Bähler, P., Lustig, A., Erni, B., 2005. Escherichia coli dihy-droxyacetone kinase controls gene expression by binding to transcription factorDhaR. The EMBO journal 24 (2), 283–293.

Balat, M., Balat, H., 2010. Progress in biodiesel processing. Applied Energy 87 (6),1815–1835.

Basiago, A.D., 1994. The limits of technological optimism. Environmentalist 14 (1),17–22.

Bauer, R., du Toit, M., Kossmann, J., 2010. Influence of environmental parameters onproduction of the acrolein precursor 3-hydroxypropionaldehyde by Lactobacillus re-uteri DSMZ 20016 and its accumulation by wine lactobacilli. International journal offood microbiology 137 (1), 28–31.

Bekker, M., Alexeeva, S., Laan, W., Sawers, G., De Mattos, J.T., Hellingwerf, K., 2010. TheArcBA two-component system of Escherichia coli is regulated by the redox state ofboth the ubiquinone and the menaquinone pool. Journal of bacteriology 192 (3),746–754.

Berrı,́ S.J., San, K.-Y., Bennett, G., 2002. The effect of NAPRTase overexpression on thetotal levels of NAD, the NADH/NAD+ ratio, and the distribution of metabolites inEscherichia coli. Metabolic engineering 4 (3), 238–247.

Bhatia, S.K., Yi, D.H., Kim, H.J., Jeon, J.M., Kim, Y.H., Sathiyanarayanan, G., Seo, H.M.,Lee, J.H., Kim, J.H., Park, K., Brigham, C.J., Yang, Y.H., 2015. Overexpression ofsuccinyl-CoA synthase for poly (3-hydroxybutyrate-co-3-hydroxyvalerate) productionin engineered Escherichia coli BL21(DE3). Journal of Applied Microbiology 119 (3),724–735.

Bhubalan, K., Chuah, J.-A., Shozui, F., Brigham, C.J., Taguchi, S., Sinskey, A.J., Rha, C.,Sudesh, K., 2011. Characterization of the Highly Active PolyhydroxyalkanoateSynthase of Chromobacterium sp. Strain USM2. Applied and EnvironmentalMicrobiology 77 (9), 2926–2933.

Biebl, H., 2001. Fermentation of glycerol by Clostridium pasteurianum — batch andcontinuous culture studies. Journal of Industrial Microbiology and Biotechnology 27(1), 18–26.

Blankschien, M.D., Clomburg, J.M., Gonzalez, R., 2010. Metabolic engineering ofEscherichia coli for the production of succinate from glycerol. Metabolic Engineering12 (5), 409–419.

Bobik, T.A., Havemann, G.D., Busch, R.J., Williams, D.S., Aldrich, H.C., 1999. ThePropanediol Utilization (pdu) Operon ofSalmonella enterica Serovar TyphimuriumLT2 Includes Genes Necessary for Formation of Polyhedral Organelles Involved inCoenzyme B12-Dependent 1,2-Propanediol Degradation. Journal of Bacteriology 181(19), 5967–5975.

Boenigk, R., Bowien, S., Gottschalk, G., 1993. Fermentation of glycerol to 1,3-propanediolin continuous cultures of Citrobacter freundii. Applied Microbiology andBiotechnology 38 (4), 453–457.

Booth, I., 2005. Glycerol and Methylglyoxal Metabolism. EcoSal Plus.Booth, I.R., Ferguson, G.P., Miller, S., Li, C., Gunasekera, B., Kinghorn, S., 2003. Bacterial

production of methylglyoxal: a survival strategy or death by misadventure?Biochemical Society Transactions 31 (6), 1406–1408.

Börjesson, P., 2009. Good or bad bioethanol from a greenhouse gas perspective–Whatdetermines this? Applied Energy 86 (5), 589–594.

Bosma, E.F., Forster, J., Nielsen, A.T., 2017. Lactobacilli and pediococci as versatile cellfactories–Evaluation of strain properties and genetic tools. Biotechnology Advances.35 (4), 419–442.

Cabelli, V.J., 1955. The tricarboxylic acid cycle in the oxidative and synthetic metabolismof Klebsiella pneumoniae. Journal of bacteriology 70 (1), 23.

Cameron, D., Altaras, N., Hoffman, M., Shaw, A., 1998. Metabolic engineering of pro-panediol pathways. Biotechnology Progress 14 (1), 116–125.

Cao, H., Heel, A.J., Ahmed, H., Mols, M., Kuipers, O.P., 2017. Cell surface engineering ofBacillus subtilis improves production yields of heterologously expressed α-amylases.Microbial Cell Factories 16 (1), 56.

Celińska, E., 2010. Debottlenecking the 1, 3-propanediol pathway by metabolic en-gineering. Biotechnology advances 28 (4), 519–530.

Celińska, E., Drożdżyńska, A., Jankowska, M., Białas, W., Czaczyk, K., Grajek, W., 2015.Genetic engineering to improve 1,3-propanediol production in an isolated Citrobacterfreundii strain. Process Biochemistry 50 (1), 48–60.

Chang, P., Chen, G.S., Chu, H.-Y., Lu, K.W., Shen, C.R., 2017. Engineering efficient pro-duction of itaconic acid from diverse substrates in Escherichia coli. Journal ofBiotechnology 249, 73–81.

Chen, R., 2007. Permeability issues in whole-cell bioprocesses and cellular membraneengineering. Applied Microbiology and Biotechnology 74 (4), 730–738.

Chen, X., Xiu, Z., Wang, J., Zhang, D., Xu, P., 2003a. Stoichiometric analysis and ex-perimental investigation of glycerol bioconversion to 1, 3-propanediol by Klebsiellapneumoniae under microaerobic conditions. Enzyme and Microbial Technology 33(4), 386–394.

Chen, X., Zhang, D.-J., Qi, W.-T., Gao, S.-J., Xiu, Z.-L., Xu, P., 2003b. Microbial fed-batchproduction of 1, 3-propanediol by Klebsiella pneumoniae under micro-aerobic con-ditions. Applied Microbiology and Biotechnology 63 (2), 143–146.

Chen, Z., Liu, H., Liu, D., 2009. Regulation of 3-hydroxypropionaldehyde accumulation inKlebsiella pneumoniae by overexpression of dhaT and dhaD genes. Enzyme andMicrobial Technology 45 (4), 305–309.

Chen, L., Ma, C., Wang, R., Yang, J., Zheng, H., 2016. Deletion of ldhA and aldH genes inKlebsiella pneumoniae to enhance 1,3-propanediol production. Biotechnology Letters38 (10), 1769–1774.

Cheong, S., Clomburg, J.M., Gonzalez, R., 2018. A synthetic pathway for the productionof 2-hydroxyisovaleric acid in Escherichia coli. Journal of industrial microbiology &biotechnology 1–10.

Cho, S., Kim, T., Woo, H.M., Kim, Y., Lee, J., Um, Y., 2015. High production of 2,3-butanediol from biodiesel-derived crude glycerol by metabolically engineeredKlebsiella oxytoca M1. Biotechnology for Biofuels 8 (1), 1–12.

Choi, K.R., Lee, S.Y., 2016. CRISPR technologies for bacterial systems: current achieve-ments and future directions. Biotechnology advances 34 (7), 1180–1209.

Choi, Y.J., Park, J.H., Kim, T.Y., Lee, S.Y., 2012. Metabolic engineering of Escherichia colifor the production of 1-propanol. Metabolic engineering 14 (5), 477–486.

Choi, S.Y., Yoon, K.H., Lee, J.I., Mitchell, R.J., 2015. Violacein: Properties and Productionof a Versatile Bacterial Pigment. BioMed Research International 2015, 8.

Chotani, G., Dodge, T., Hsu, A., Kumar, M., LaDuca, R., Trimbur, D., Weyler, W., Sanford,K., 2000. The commercial production of chemicals using pathway engineering.Biochimica et Biophysica Acta (BBA)-Protein Structure and Molecular Enzymology1543 (2), 434–455.

Chu, H.S., Kim, Y.S., Lee, C.M., Lee, J.H., Jung, W.S., Ahn, J.H., Song, S.H., Choi, I.S.,Cho, K.M., 2015. Metabolic engineering of 3-hydroxypropionic acid biosynthesis inEscherichia coli. Biotechnology and bioengineering 112 (2), 356–364.

Clomburg, J.M., Gonzalez, R., 2011. Metabolic engineering of Escherichia coli for theproduction of 1,2-propanediol from glycerol. Biotechnology and Bioengineering 108(4), 867–879.

Clomburg, J.M., Gonzalez, R., 2013. Anaerobic fermentation of glycerol: a platform forrenewable fuels and chemicals. Trends in biotechnology 31 (1), 20–28.

da Cunha, M.V., Foster, M.A., 1992. Sugar-glycerol cofermentations in lactobacilli: thefate of lactate. Journal of bacteriology 174 (3), 1013–1019.

da Silva, G.P., Mack, M., Contiero, J., 2009. Glycerol: A promising and abundant carbonsource for industrial microbiology. Biotechnology Advances 27 (1), 30–39.

Daniel, R., Gottschalk, G., 1992. Growth temperature-dependent activity of glycerol de-hydratase in Escherichia coli expressing the Citrobacter freundii dha regulon. FEMSmicrobiology letters 100 (1-3), 281–285.

Daniel, R., Stuertz, K., Gottschalk, G., 1995. Biochemical and molecular characterizationof the oxidative branch of glycerol utilization by Citrobacter freundii. Journal of

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

562

Page 26: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Bacteriology 177 (15), 4392–4401.Das, A., Tyagi, N., Verma, A., Akhtar, S., Mukherjee, K.J., 2018. Metabolic engineering of

Escherichia coli W3110 strain by incorporating genome-level modifications andsynthetic plasmid modules to enhance L-Dopa production from glycerol. PreparativeBiochemistry and Biotechnology 1–12.

de Almeida, A., Nikel, P.I., Giordano, A.M., Pettinari, M.J., 2007. Effects of Granule-Associated Protein PhaP on Glycerol-Dependent Growth and Polymer Production inPoly(3-Hydroxybutyrate)-Producing Escherichia coli. Applied and EnvironmentalMicrobiology 73 (24), 7912–7916.

Dellomonaco, C., Clomburg, J.M., Miller, E.N., Gonzalez, R., 2011. Engineered reversal ofthe [bgr]-oxidation cycle for the synthesis of fuels and chemicals. Nature 476 (7360),355–359.

Deutscher, J., Francke, C., Postma, P.W., 2006. How phosphotransferase system-relatedprotein phosphorylation regulates carbohydrate metabolism in bacteria.Microbiology and Molecular Biology Reviews 70 (4), 939–1031.

Dharmadi, Y., Murarka, A., Gonzalez, R., 2006. Anaerobic fermentation of glycerol byEscherichia coli: A new platform for metabolic engineering. Biotechnology andBioengineering 94 (5), 821–829.

Dishisha, T., Pereyra, L.P., Pyo, S.-H., Britton, R.A., Hatti-Kaul, R., 2014. Flux analysis ofthe Lactobacillus reuteri propanediol-utilization pathway for production of 3-hy-droxypropionaldehyde, 3-hydroxypropionic acid and 1, 3-propanediol from glycerol.Microbial cell factories 13 (1), 1.

Dobson, R., Gray, V., Rumbold, K., 2012. Microbial utilization of crude glycerol for theproduction of value-added products. Journal of Industrial Microbiology &Biotechnology 39 (2), 217–226.

Dong, H., Zhao, C., Zhang, T., Zhu, H., Lin, Z., Tao, W., Zhang, Y., Li, Y., 2017. A sys-tematically chromosomally engineered Escherichia coli efficiently produces butanol.Metabolic Engineering. 44, 284–292.

Durán, N., Menck, C.F.M., 2001. Chromobacterium violaceum: A Review ofPharmacological and Industiral Perspectives. Critical Reviews in Microbiology 27 (3),201–222.

Durnin, G., Clomburg, J., Yeates, Z., Alvarez, P.J.J., Zygourakis, K., Campbell, P.,Gonzalez, R., 2009. Understanding and harnessing the microaerobic metabolism ofglycerol in Escherichia coli. Biotechnology and Bioengineering 103 (1), 148–161.

Dwiarti, L., Otsuka, M., Miura, S., Yaguchi, M., Okabe, M., 2007. Itaconic acid productionusing sago starch hydrolysate by Aspergillus terreus TN484-M1. Bioresource tech-nology 98 (17), 3329–3337.

El-Ziney, M., Arneborg, N., Uyttendaele, M., Debevere, J., Jakobsen, M., 1998.Characterization of growth and metabolite production of Lactobacillus reuteri duringglucose/glycerol cofermentation in batch and continuous cultures. BiotechnologyLetters 20 (10), 913–916.

Emptage, M., Haynie, S.L., Laffend, L.A., Pucci, J.P., Whited, G., 2003a. Process for thebiological production of 1,3-propanediol with high titer. Google Patents.

Emptage, M., Haynie, S.L., Laffend, L.A., Pucci, J.P., Whited, G., 2003b. Process for thebiological production with high titer. United States.

Feng, X., Ding, Y., Xian, M., Xu, X., Zhang, R., Zhao, G., 2014. Production of opticallypure d-lactate from glycerol by engineered Klebsiella pneumoniae strain. BioresourceTechnology 172, 269–275.

Feng, X., Xian, M., Liu, W., Xu, C., Zhang, H., Zhao, G., 2015. Biosynthesis of poly(3-hydroxypropionate) from glycerol using engineered Klebsiella pneumoniae strainwithout vitamin B12. Bioengineered 6 (2), 77–81.

Forage, R.G., Foster, M.A., 1982. Glycerol fermentation in Klebsiella pneumoniae: func-tions of the coenzyme B12-dependent glycerol and diol dehydratases. Journal ofBacteriology 149 (2), 413–419.

Forage, R.G., Lin, E.C., 1982. DHA system mediating aerobic and anaerobic dissimilationof glycerol in Klebsiella pneumoniae NCIB 418. Journal of Bacteriology 151 (2),591–599.

Gallardo, R., Alves, M., Rodrigues, L., 2017. Influence of nutritional and operationalparameters on the production of butanol or 1, 3-propanediol from glycerol by amutant Clostridium pasteurianum. New biotechnology 34, 59–67.

Ganesh, M., Senthamarai, A., Shanmughapriya, S., Natarajaseenivasan, K., 2015. Effectiveproduction of low crystallinity Poly(3-hydroxybutyrate) by recombinant E. coli strainJM109 using crude glycerol as sole carbon source. Bioresource Technology 192,677–681.

Gao, Y., Liu, C., Ding, Y., Sun, C., Zhang, R., Xian, M., Zhao, G., 2014. Development ofGenetically Stable Escherichia coli Strains for Poly(3-Hydroxypropionate)Production. PLoS ONE 9 (5), e97845.

Garai-Ibabe, G., Ibarburu, I., Berregi, I., Claisse, O., Lonvaud-Funel, A., Irastorza, A.,Duenas, M.T., 2008. Glycerol metabolism and bitterness producing lactic acid bac-teria in cidermaking. International journal of food microbiology 121 (3), 253–261.

Gardner, T.S., Cantor, C.R., Collins, J.J., 2000. Construction of a genetic toggle switch inEscherichia coli. Nature 403, 339.

Girbal, L., Soucaille, P., 1994. Regulation of Clostridium acetobutylicum metabolism asrevealed by mixed-substrate steady-state continuous cultures: role of NADH/NADratio and ATP pool. Journal of Bacteriology 176 (21), 6433–6438.

Girbal, L., Croux, C., Vasconcelos, I., Soucaille, P., 1995. Regulation of metabolic shifts inClostridium acetobutylicum ATCC 824. FEMS Microbiology Reviews 17 (3),287–297.

Gonzalez, R., Murarka, A., Dharmadi, Y., Yazdani, S.S., 2008. A new model for theanaerobic fermentation of glycerol in enteric bacteria: Trunk and auxiliary pathwaysin Escherichia coli. Metabolic Engineering 10 (5), 234–245.

González-Pajuelo, M., Meynial-Salles, I., Mendes, F., Andrade, J.C., Vasconcelos, I.,Soucaille, P., 2005. Metabolic engineering of Clostridium acetobutylicum for theindustrial production of 1,3-propanediol from glycerol. Metabolic Engineering 7(5–6), 329–336.

González-Pajuelo, M., Meynial-Salles, I., Mendes, F., Soucaille, P., Vasconcelos, I., 2006.

Microbial Conversion of Glycerol to 1,3-Propanediol: Physiological Comparison of aNatural Producer, Clostridium butyricum VPI 3266, and an Engineered Strain,Clostridium acetobutylicum DG1(pSPD5). Applied and Environmental Microbiology72 (1), 96–101.

Guo, X., Fang, H., Zhuge, B., Zong, H., Song, J., Zhuge, J., Du, X., 2013. budC knockout inKlebsiella pneumoniae for bioconversion from glycerol to 1,3-propanediol.Biotechnology and Applied Biochemistry 60 (6), 557–563.

Guo, J.-Y., Hu, K.-L., Bi, C.-H., Li, Q.-Y., Zhang, X.-L., 2018. Construction of an alternativeglycerol-utilization pathway for improved β-carotene production in Escherichia coli.Journal of industrial microbiology & biotechnology, 1-9.

Hao, J., Lin, R., Zheng, Z., Sun, Y., Liu, D., 2008a. 3-Hydroxypropionaldehyde guidedglycerol feeding strategy in aerobic 1, 3-propanediol production by Klebsiellapneumoniae. Journal of industrial microbiology & biotechnology 35 (12),1615–1624.

Hao, J., Wang, W., Tian, J., Li, J., Liu, D., 2008b. Decrease of 3-hydroxypropionaldehydeaccumulation in 1,3-propanediol production by over-expressing dhaT gene inKlebsiella pneumoniae TUAC01. Journal of Industrial Microbiology & Biotechnology35 (7), 735–741.

Hayashi, S.-I., Lin, E., 1967. Purification and properties of glycerol kinase fromEscherichia coli. Journal of Biological Chemistry 242 (5), 1030–1035.

Hein, L., Leemans, R., 2012. The impact of first-generation biofuels on the depletion ofthe global phosphorus reserve. Ambio 41 (4), 341–349.

Heller, K.B., Lin, E., Wilson, T.H., 1980. Substrate specificity and transport properties ofthe glycerol facilitator of Escherichia coli. Journal of Bacteriology 144 (1), 274–278.

Homann, T., Tag, C., Biebl, H., Deckwer, W.-D., Schink, B., 1990. Fermentation of gly-cerol to 1,3-propanediol by Klebsiella and Citrobacter strains. Applied Microbiologyand Biotechnology 33 (2), 121–126.

Hong, E., Kim, J., Ha, S.-j., Ryu, Y., 2015. Improved 1, 3-propanediol production byEscherichia coli from glycerol due to Co-expression of glycerol dehydratase re-activation factors and succinate addition. Biotechnology and Bioprocess Engineering20 (5), 849–855.

Honjo, H., Tsuruno, K., Tatsuke, T., Sato, M., Hanai, T., 2015. Dual synthetic pathway for3-hydroxypropionic acid production in engineered Escherichia coli. Journal ofbioscience and bioengineering 120 (2), 199–204.

Hopper, D.J., Cooper, R.A., 1972. The purification and properties of Escherichia colimethylglyoxal synthase. Biochemical Journal 128 (2), 321–329.

Horng, Y.-T., Chang, K.-C., Chou, T.-C., Yu, C.-J., Chien, C.-C., Wei, Y.-H., Soo, P.-C.,2010. Inactivation of dhaD and dhaK abolishes by-product accumulation during 1,3-propanediol production in Klebsiella pneumoniae. Journal of Industrial Microbiology& Biotechnology 37 (7), 707–716.

Hu, F., Cao, Y., Xiao, F., Zhang, J., Li, H., 2007. Site-directed mutagenesis of Aeromonashydrophila enoyl coenzyme A hydratase enhancing 3-hydroxyhexanoate fractions ofpoly (3-hydroxybutyrate-co-3-hydroxyhexanoate). Current microbiology 55 (1),20–24.

Huang, K.-x., Rudolph, F.B., Bennett, G.N., 1999. Characterization of methylglyoxalsynthase from Clostridium acetobutylicum ATCC 824 and its use in the formation of1,2-propanediol. Applied and Environmental Microbiology 65 (7), 3244–3247.

Huang, Y., Li, Z., Shimizu, K., Ye, Q., 2012. Simultaneous production of 3-hydro-xypropionic acid and 1,3-propanediol from glycerol by a recombinant strain ofKlebsiella pneumoniae. Bioresource Technology 103 (1), 351–359.

Huynh, D.T.N., Kim, A.-Y., Seol, I.-H., Jung, S., Lim, M.-C., Lee, J.-A., Jo, M.-R., Choi, S.-J., Kim, B., Lee, J., 2015. Inactivation of the virulence factors from 2, 3-butanediol-producing Klebsiella pneumoniae. Applied microbiology and biotechnology 99 (22),9427–9438.

Ignatova, Z., Mahsunah, A., Georgieva, M., Kasche, V., 2003. Improvement of post-translational bottlenecks in the production of penicillin amidase in recombinantEscherichia coli strains. Applied and Environmental Microbiology 69 (2), 1237–1245.

Iuchi, S., Lin, E., 1993. Adaptation of Escherichia coli to redox environments by geneexpression. Molecular microbiology 9 (1), 9–15.

Jang, Y.-S., Lee, J., Malaviya, A., Seung, D.Y., Cho, J.H., Lee, S.Y., 2012. Butanol pro-duction from renewable biomass: Rediscovery of metabolic pathways and metabolicengineering. Biotechnology Journal 7 (2), 186–198.

Jensen, T.Ø., Kvist, T., Mikkelsen, M., Christensen, P., Westermann, P., 2012.Fermentation of crude glycerol from biodiesel production by Clostridium pasteur-ianum. Journal of Industrial Microbiology & Biotechnology 39.

Jeon, H.G., Cheong, D.E., Han, Y., Song, J.J., Choi, J.H., 2016. Itaconic acid productionfrom glycerol using Escherichia coli harboring a random synonymous codon-sub-stituted 5′-coding region variant of the cadA gene. Biotechnology and bioengineering113 (7), 1504–1510.

Jiang, P.-x., Wang, H.-s., Zhang, C., Lou, K., Xing, X.-H., 2010. Reconstruction of theviolacein biosynthetic pathway from Duganella sp. B2 in different heterologous hosts.Applied Microbiology and Biotechnology 86 (4), 1077–1088.

Jiang, Y., Chen, B., Duan, C., Sun, B., Yang, J., Yang, S., 2015. Multigene editing in theEscherichia coli genome via the CRISPR-Cas9 system. Applied and EnvironmentalMicrobiology 81 (7), 2506–2514.

Jiang, W., Wang, S., Wang, Y., Fang, B., 2016. Key enzymes catalyzing glycerol to 1,3-propanediol. Biotechnology for Biofuels 9, 57.

Jiang, L., Liu, H., Mu, Y., Sun, Y., Xiu, Z., 2017. High tolerance to glycerol and highproduction of 1, 3-propanediol in batch fermentations by microbial consortium frommarine sludge. Engineering in Life Sciences 17 (6), 635–644.

Jiang, L., Dai, J., Sun, Y., Xiu, Z., 2018. The effects of ionic liquid 1-ethyl-3-methylimi-dazolium trifluoromethanesulfonate on the production of 1,3-propanediol from crudeglycerol by microbial consortium. Bioprocess and Biosystems Engineering 41 (8),1079–1088.

Jin, R.Z., Forage, R.G., Lin, E.C., 1982. Glycerol kinase as a substitute for dihydrox-yacetone kinase in a mutant of Klebsiella pneumoniae. Journal of Bacteriology 152

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

563

Page 27: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

(3), 1303–1307.Jung, S.-G., Jang, J.-H., Kim, A.-Y., Lim, M.-C., Kim, B., Lee, J., Kim, Y.-R., 2013. Removal

of pathogenic factors from 2, 3-butanediol-producing Klebsiella species by in-activating virulence-related wabG gene. Applied microbiology and biotechnology 97(5), 1997–2007.

Jung, W.S., Kang, J.H., Chu, H.S., Choi, I.S., Cho, K.M., 2014. Elevated production of 3-hydroxypropionic acid by metabolic engineering of the glycerol metabolism inEscherichia coli. Metabolic engineering 23, 116–122.

Kang, T.S., Korber, D.R., Tanaka, T., 2013. Glycerol and environmental factors: effects on1,3-propanediol production and NAD+ regeneration in Lactobacillus panis PM1.Journal of Applied Microbiology, n/a-n/a.

Kang, T.S., Korber, D.R., Tanaka, T., 2014a. Bioconversion of glycerol to 1,3-propanediolin thin stillage-based media by engineered Lactobacillus panis PM1. Journal ofIndustrial Microbiology & Biotechnology 41 (4), 629–635.

Kang, T.S., Korber, D.R., Tanaka, T., 2014b. Transcriptional repressor role of PocR on the1,3-propanediol biosynthetic pathway by Lactobacillus panis PM1. BiotechnologyLetters 36 (6), 1263–1269.

Keseler, I.M., Mackie, A., Santos-Zavaleta, A., Billington, R., Bonavides-Martínez, C.,Caspi, R., Fulcher, C., Gama-Castro, S., Kothari, A., Krummenacker, M., 2016. TheEcoCyc database: reflecting new knowledge about Escherichia coli K-12. Nucleicacids research 45 (D1), D543–D550.

Keuth, S., Bisping, B., 1994. Vitamin B12 production by Citrobacter freundii or Klebsiellapneumoniae during tempeh fermentation and proof of enterotoxin absence by PCR.Applied and Environmental Microbiology 60 (5), 1495–1499.

Kim, J.-H., Yoo, S.-J., Oh, D.-K., Kweon, Y.-G., Park, D.-W., Lee, C.-H., Gil, G.-H., 1996.Selection of a Streptococcus equi. mutant and optimization of culture conditions forthe production of high molecular weight hyaluronic acid. Enzyme and MicrobialTechnology 19 (6), 440–445.

Kim, Y., Mosier, N.S., Hendrickson, R., Ezeji, T., Blaschek, H., Dien, B., Cotta, M., Dale, B.,Ladisch, M.R., 2008. Composition of corn dry-grind ethanol by-products: DDGS, wetcake, and thin stillage. Bioresource Technology 99 (12), 5165–5176.

Kim, K., Kim, S.-K., Park, Y.-C., Seo, J.-H., 2014. Enhanced production of 3-hydro-xypropionic acid from glycerol by modulation of glycerol metabolism in recombinantEscherichia coli. Bioresource technology 156, 170–175.

Ko, Y., Ashok, S., Zhou, S., Kumar, V., Park, S., 2012. Aldehyde dehydrogenase activity isimportant to the production of 3-hydroxypropionic acid from glycerol by re-combinant Klebsiella pneumoniae. Process Biochemistry 47 (7), 1135–1143.

Ko, Y., Ashok, S., Ainala, S.K., Sankaranarayanan, M., Chun, A.Y., Jung, G.Y., Park, S.,2014. Coenzyme B12 can be produced by engineered Escherichia coli under bothanaerobic and aerobic conditions. Biotechnology journal 9 (12), 1526–1535.

Ko, Y., Ashok, S., Seol, E., Ainala, S.K., Park, S., 2015. Deletion of putative oxidor-eductases from Klebsiella pneumoniae J2B could reduce 1,3-propanediol during theproduction of 3-hydroxypropionic acid from glycerol. Biotechnology and BioprocessEngineering 20 (5), 834–843.

Ko, Y., Seol, E., Sekar, B.S., Kwon, S., Lee, J., Park, S., 2017. Metabolic engineering ofKlebsiella pneumoniae J2B for co-production of 3-hydroxypropionic acid and 1, 3-propanediol from glycerol: Reduction of acetate and other by-products. Bioresourcetechnology 244, 1096–1103.

Kubiak, P., Leja, K., Myszka, K., Celińska, E., Spychała, M., Szymanowska-Powałowska,D., Czaczyk, K., Grajek, W., 2012. Physiological predisposition of various Clostridiumspecies to synthetize 1,3-propanediol from glycerol. Process Biochemistry 47 (9),1308–1319.

Kumar, V., Park, S., 2017. Potential and limitations of Klebsiella pneumoniae as a mi-crobial cell factory utilizing glycerol as the carbon source. Biotechnology advances.36 (1), 150–167.

Kumar, V., Sankaranarayanan, M., Jae, K.-e., Durgapal, M., Ashok, S., Ko, Y., Sarkar, R.,Park, S., 2012. Co-production of 3-hydroxypropionic acid and 1,3-propanediol fromglycerol using resting cells of recombinant Klebsiella pneumoniae J2B strain over-expressing aldehyde dehydrogenase. Applied Microbiology and Biotechnology 96 (2),373–383.

Kumar, V., Sankaranarayanan, M., Durgapal, M., Zhou, S., Ko, Y., Ashok, S., Sarkar, R.,Park, S., 2013. Simultaneous production of 3-hydroxypropionic acid and 1, 3-pro-panediol from glycerol using resting cells of the lactate dehydrogenase-deficient re-combinant Klebsiella pneumoniae overexpressing an aldehyde dehydrogenase.Bioresource technology 135, 555–563.

Kumar, V., Durgapal, M., Sankaranarayanan, M., Somasundar, A., Rathnasingh, C., Song,H., Seung, D., Park, S., 2016. Effects of mutation of 2,3-butanediol formationpathway on glycerol metabolism and 1,3-propanediol production by Klebsiellapneumoniae J2B. Bioresource Technology 214, 432–440.

Kwak, S., Park, Y.-C., Seo, J.-H., 2013. Biosynthesis of 3-hydroxypropionic acid fromglycerol in recombinant Escherichia coli expressing Lactobacillus brevis dhaB anddhaR gene clusters and E. coli K-12 aldH. Bioresource technology 135, 432–439.

Lacmata, S.T., Kuiate, J.-R., Ding, Y., Xian, M., Liu, H., Boudjeko, T., Feng, X., Zhao, G.,2017. Enhanced poly(3-hydroxypropionate) production via β-alanine pathway inrecombinant Escherichia coli. PLOS ONE 12 (3), e0173150.

Lee, S., Kim, B., Park, K., Um, Y., Lee, J., 2012. Synthesis of pure meso-2, 3-butanediolfrom crude glycerol using an engineered metabolic pathway in Escherichia coli.Applied biochemistry and biotechnology 166 (7), 1801–1813.

Lee, S.-M., Hong, W.-K., Heo, S.-Y., Park, J.M., Jung, Y.R., Oh, B.-R., Joe, M.-H., Seo, J.-W., Kim, C.H., 2014. Enhancement of 1,3-propanediol production by expression ofpyruvate decarboxylase and aldehyde dehydrogenase from Zymomonas mobilis inthe acetolactate-synthase-deficient mutant of Klebsiella pneumoniae. Journal ofIndustrial Microbiology & Biotechnology 41 (8), 1259–1266.

Lee, J.H., Jung, M.-Y., Oh, M.-K., 2018. High-yield production of 1, 3-propanediol fromglycerol by metabolically engineered Klebsiella pneumoniae. Biotechnology for bio-fuels 11 (1), 104.

Leong, Y.K., Show, P.L., Ooi, C.W., Ling, T.C., Lan, J.C.-W., 2014. Current trends inpolyhydroxyalkanoates (PHAs) biosynthesis: Insights from the recombinantEscherichia coli. Journal of Biotechnology 180, 52–65.

Li, N., Zhang, B., Chen, T., Wang, Z., Tang, Y.-j., Zhao, X., 2013. Directed pathwayevolution of the glyoxylate shunt in Escherichia coli for improved aerobic succinateproduction from glycerol. Journal of Industrial Microbiology & Biotechnology 40(12), 1461–1475.

Li, N., Zhang, B., Wang, Z., Tang, Y.-J., Chen, T., Zhao, X., 2014. Engineering Escherichiacoli for fumaric acid production from glycerol. Bioresource Technology 174, 81–87.

Li, Q., Chen, J., Minton, N.P., Zhang, Y., Wen, Z., Liu, J., Yang, H., Zeng, Z., Ren, X., Yang,J., 2016a. CRISPR-based genome editing and expression control systems inClostridium acetobutylicum and Clostridium beijerinckii. Biotechnology journal 11(7), 961–972.

Li, Y., Wang, X., Ge, X., Tian, P., 2016b. High Production of 3-Hydroxypropionic Acid inKlebsiella pneumoniae by Systematic Optimization of Glycerol Metabolism. ScientificReports 6, 26932.

Lim, H.G., Noh, M.H., Jeong, J.H., Park, S., Jung, G.Y., 2016. Optimum Rebalancing ofthe 3-Hydroxypropionic Acid Production Pathway from Glycerol in Escherichia coli.ACS Synthetic Biology 5 (11), 1247–1255.

Lin, E., 1976a. Glycerol dissimilation and its regulation in bacteria. Annual Reviews inMicrobiology 30 (1), 535–578.

Lin, E.C.C., 1976b. Glycerol Dissimilation and its Regulation in Bacteria. Annual Reviewsin Microbiology 30 (1), 535–578.

Lin, J., Zhang, Y., Xu, D., Xiang, G., Jia, Z., Fu, S., Gong, H., 2016. Deletion of poxB, pta,and ackA improves 1,3-propanediol production by Klebsiella pneumoniae. AppliedMicrobiology and Biotechnology 100 (6), 2775–2784.

Linares-Pastén, J.A., Sabet-Azad, R., Pessina, L., Sardari, R.R.R., Ibrahim, M.H.A., Hatti-Kaul, R., 2015. Efficient poly(3-hydroxypropionate) production from glycerol usingLactobacillus reuteri and recombinant Escherichia coli harboring L. reuteri propio-naldehyde dehydrogenase and Chromobacterium sp. PHA synthase genes.Bioresource Technology 180, 172–176.

Liochev, S.I., Fridovich, I., 1993. Modulation of the fumarases of Escherichia coli in re-sponse to oxidative stress. Archives of biochemistry and biophysics 301 (2), 379–384.

Liu, H.-J., Zhang, D.-J., Xu, Y.-H., Mu, Y., Sun, Y.-Q., Xiu, Z.-L., 2007. Microbial pro-duction of 1, 3-propanediol from glycerol by Klebsiella pneumoniae under micro-aerobic conditions up to a pilot scale. Biotechnology letters 29 (8), 1281–1285.

Liu, C., Li, F., Ma, L.P., Cheng, H.M., 2010. Advanced materials for energy storage.Advanced materials 22 (8).

Liu, L., Liu, Y., Li, J., Du, G., Chen, J., 2011. Microbial production of hyaluronic acid:current state, challenges, and perspectives. Microbial Cell Factories 10 (1), 99.

Liu, Y., Li, J., Du, G., Chen, J., Liu, L., 2017. Metabolic engineering of Bacillus subtilisfueled by systems biology: Recent advances and future directions. Biotechnologyadvances 35 (1), 20–30.

Liyanage, H., Kashket, S., Young, M., Kashket, E.R., 2001. Clostridium beijerinckii andClostridium difficile detoxify methylglyoxal by a novel mechanism involving glyceroldehydrogenase. Applied and Environmental Microbiology 67 (5), 2004–2010.

Loaces, I., Rodríguez, C., Amarelle, V., Fabiano, E., Noya, F., 2016. Improved glycerol toethanol conversion by E. coli using a metagenomic fragment isolated from an anae-robic reactor. Journal of Industrial Microbiology & Biotechnology, 1-12.

López, J.Á.S., Santos, M.d.l.Á.M., Pérez, A.F.C., Martín, A.M., 2009. Anaerobic digestionof glycerol derived from biodiesel manufacturing. Bioresource technology 100 (23),5609–5615.

Lu, X., Ji, G., Zong, H., Zhuge, B., 2016. The role of bud ABC on 1, 3-propanediol and 2, 3-butanediol production from glycerol in Klebsiella pneumoniae CICIM B0057.Bioengineered 7 (6), 439–444.

Lu, X., Ren, S., Lu, J., Zong, H., Song, J., Zhuge, B., 2018. Enhanced 1, 3-propanediolproduction in Klebsiella pneumoniae by a combined strategy of strengthening theTCA cycle and weakening the glucose effect. Journal of applied microbiology 124 (3),682–690.

Luo, L., van der Voet, E., Huppes, G., 2010. Biorefining of lignocellulosicfeedstock–Technical, economic and environmental considerations. Bioresourcetechnology 101 (13), 5023–5032.

Luo, L.H., Seo, J.-W., Oh, B.-R., Seo, P.-S., Heo, S.-Y., Hong, W.-K., Kim, D.-H., Kim, C.H.,2011a. Stimulation of reductive glycerol metabolism by overexpression of an alde-hyde dehydrogenase in a recombinant Klebsiella pneumoniae strain defective in theoxidative pathway. Journal of Industrial Microbiology & Biotechnology 38 (8),991–999.

Luo, L.H., Seo, J.W., Baek, J.O., Oh, B.R., Heo, S.Y., Hong, W.K., Kim, D.H., Kim, C.H.,2011b. Identification and characterization of the propanediol utilization proteinPduP of Lactobacillus reuteri for 3-hydroxypropionic acid production from glycerol.Applied Microbiology and Biotechnology 89 (3), 697–703.

Luo, L.H., Kim, C.H., Heo, S.-Y., Oh, B.-R., Hong, W.-K., Kim, S., Kim, D.-H., Seo, J.-W.,2012. Production of 3-hydroxypropionic acid through propionaldehyde dehy-drogenase PduP mediated biosynthetic pathway in Klebsiella pneumoniae.Bioresource Technology 103 (1), 1–6.

Luo, L.H., Seo, J.-W., Heo, S.-Y., Oh, B.-R., Kim, D.-H., Kim, C.H., 2013. Identification andcharacterization of Klebsiella pneumoniae aldehyde dehydrogenases increasing pro-duction of 3-hydroxypropionic acid from glycerol. Bioprocess and BiosystemsEngineering 36 (9), 1319–1326.

Luthi-Peng, Q., Dileme, F.B., Puhan, Z., 2002. Effect of glucose on glycerol bioconversionby Lactobacillus reuteri. Applied Microbiology and Biotechnology 59 (2-3), 289–296.

Macis, L., Daniel, R., Gottschalk, G., 1998. Properties and sequence of the coenzyme B12-dependent glycerol dehydratase of Clostridium pasteurianum. FEMS MicrobiologyLetters 164 (1), 21–28.

MacLean, M., Ness, L., Ferguson, G., Booth, I., 1998a. The role of glyoxalase I in thedetoxification of methylglyoxal and in the activation of the KefB K+ efflux system in

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

564

Page 28: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Escherichia coli. Molecular microbiology 27 (3), 563–571.MacLean, M.J., Ness, L.S., Ferguson, G.P., Booth, I.R., 1998b. The role of glyoxalase I in

the detoxification of methylglyoxal and in the activation of the KefB K+ effluxsystem in Escherichia coli. Molecular Microbiology 27 (3), 563–571.

Maervoet, V.E., Beauprez, J., De Maeseneire, S.L., Soetaert, W.K., De Mey, M., 2012a.Citrobacter werkmanii, a new candidate for the production of 1, 3-propanediol: strainselection and carbon source optimization. Green Chemistry 14 (8), 2168–2178.

Maervoet, V.E.T., Beauprez, J., De Maeseneire, S.L., Soetaert, W.K., De Mey, M., 2012b.Citrobacter werkmanii, a new candidate for the production of 1,3-propanediol: strainselection and carbon source optimization. Green Chemistry 14 (8), 2168–2178.

Maervoet, V.E.T., De Maeseneire, S.L., Avci, F.G., Beauprez, J., Soetaert, W.K., De Mey,M., 2014. 1,3-propanediol production with Citrobacter werkmanii DSM17579: effectof a dhaD knock-out. Microbial Cell Factories 13, 70.

Maervoet, V.E.T., De Maeseneire, S.L., Avci, F.G., Beauprez, J., Soetaert, W.K., De Mey,M., 2016. High yield 1,3-propanediol production by rational engineering of the 3-hydroxypropionaldehyde bottleneck in Citrobacter werkmanii. Microbial CellFactories 15, 23.

Makarova, K., Slesarev, A., Wolf, Y., Sorokin, A., Mirkin, B., Koonin, E., Pavlov, A.,Pavlova, N., Karamychev, V., Polouchine, N., Shakhova, V., Grigoriev, I., Lou, Y.,Rohksar, D., Lucas, S., Huang, K., Goodstein, D.M., Hawkins, T., Plengvidhya, V.,Welker, D., Hughes, J., Goh, Y., Benson, A., Baldwin, K., Lee, J.H., Diaz-Muniz, I.,Dosti, B., Smeianov, V., Wechter, W., Barabote, R., Lorca, G., Altermann, E.,Barrangou, R., Ganesan, B., Xie, Y., Rawsthorne, H., Tamir, D., Parker, C., Breidt, F.,Broadbent, J., Hutkins, R., O'Sullivan, D., Steele, J., Unlu, G., Saier, M.,Klaenhammer, T., Richardson, P., Kozyavkin, S., Weimer, B., Mills, D., 2006.Comparative genomics of the lactic acid bacteria. Proc Natl Acad Sci U S A 103 (42),15611–15616.

Malaviya, A., Jang, Y.-S., Lee, S.Y., 2012. Continuous butanol production with reducedbyproducts formation from glycerol by a hyper producing mutant of Clostridiumpasteurianum. Applied Microbiology and Biotechnology 93 (4), 1485–1494.

Mamat, U., Wilke, K., Bramhill, D., Schromm, A.B., Lindner, B., Kohl, T.A., Corchero, J.L.,Villaverde, A., Schaffer, L., Head, S.R., 2015. Detoxifying Escherichia coli for en-dotoxin-free production of recombinant proteins. Microbial cell factories 14 (1), 57.

Martínez-Gómez, K., Flores, N., Castañeda, H.M., Martínez-Batallar, G., Hernández-Chávez, G., Ramírez, O.T., Gosset, G., Encarnación, S., Bolivar, F., 2012. New insightsinto Escherichia coli metabolism: carbon scavenging, acetate metabolism and carbonrecycling responses during growth on glycerol. Microbial cell factories 11 (1), 1.

Matsubara, M., Urano, N., Yamada, S., Narutaki, A., Fujii, M., Kataoka, M., 2016.Fermentative production of 1-propanol from d-glucose, l-rhamnose and glycerolusing recombinant Escherichia coli. Journal of bioscience and bioengineering. 122(4), 421–426.

Mazumdar, S., Clomburg, J.M., Gonzalez, R., 2010. Escherichia coli strains engineered forhomofermentative production of D-lactic acid from glycerol. Applied andEnvironmental Microbiology 76 (13), 4327–4336.

Mazumdar, S., Blankschien, M.D., Clomburg, J.M., Gonzalez, R., 2013. Efficient synthesisof L-lactic acid from glycerol by metabolically engineered Escherichia coli. MicrobialCell Factories 12 (1), 1.

Mendes, A.S., de Carvalho, J.E., Duarte, M.C.T., Durán, N., Bruns, R.E., 2001. Factorialdesign and response surface optimization of crude violacein for Chromobacteriumviolaceum production. Biotechnology Letters 23 (23), 1963–1969.

Menon, V., Rao, M., 2012. Trends in bioconversion of lignocellulose: biofuels, platformchemicals & biorefinery concept. Progress in Energy and Combustion Science 38 (4),522–550.

Menzel, K., Zeng, A.-P., Biebl, H., Deckwer, W.-D., 1996. Kinetic, dynamic, and pathwaystudies of glycerol metabolism by Klebsiella pneumoniae in anaerobic continuousculture: I. The phenomena and characterization of oscillation and hysteresis.Biotechnology and Bioengineering 52 (5), 549–560.

Menzel, K., Zeng, A.-P., Deckwer, W.-D., 1997. Enzymatic evidence for an involvement ofpyruvate dehydrogenase in the anaerobic glycerol metabolism of Klebsiella pneu-moniae. Journal of Biotechnology 56 (2), 135–142.

Menzel, K., Ahrens, K., Zeng, A.P., Deckwer, W.D., 1998. Kinetic, dynamic, and pathwaystudies of glycerol metabolism by Klebsiella pneumoniae in anaerobic continuousculture: IV. Enzymes and fluxes of pyruvate metabolism. Biotechnology andBioengineering 60 (5), 617–626.

Metsoviti, M., Zeng, A.-P., Koutinas, A.A., Papanikolaou, S., 2013. Enhanced 1,3-propa-nediol production by a newly isolated Citrobacter freundii strain cultivated on bio-diesel-derived waste glycerol through sterile and non-sterile bioprocesses. Journal ofBiotechnology 163 (4), 408–418.

Morita, H., Toh, H., Fukuda, S., Horikawa, H., Oshima, K., Suzuki, T., Murakami, M.,Hisamatsu, S., Kato, Y., Takizawa, T., Fukuoka, H., Yoshimura, T., Itoh, K.,O'Sullivan, D.J., McKay, L.L., Ohno, H., Kikuchi, J., Masaoka, T., Hattori, M., 2008.Comparative genome analysis of Lactobacillus reuteri and Lactobacillus fermentumreveal a genomic island for reuterin and cobalamin production. DNA research : aninternational journal for rapid publication of reports on genes and genomes 15 (3),151–161.

Mozzi, F., Vaningelgem, F., Hébert, E.M., Van der Meulen, R., Foulquié Moreno, M.R.,Font de Valdez, G., De Vuyst, L., 2006. Diversity of Heteropolysaccharide-ProducingLactic Acid Bacterium Strains and Their Biopolymers. Applied and EnvironmentalMicrobiology 72 (6), 4431–4435.

Mu, Y., Teng, H., Zhang, D.-J., Wang, W., Xiu, Z.-L., 2006. Microbial production of 1, 3-propanediol by Klebsiella pneumoniae using crude glycerol from biodiesel prepara-tions. Biotechnology letters 28 (21), 1755–1759.

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. Applied and environmental microbiology 74 (4), 1124–1135.

Nakamura, C.E., Whited, G.M., 2003. Metabolic engineering for the microbial production

of 1,3-propanediol. Current Opinion in Biotechnology 14 (5), 454–459.Nölling, J., Breton, G., Omelchenko, M.V., Makarova, K.S., Zeng, Q., Gibson, R., Lee,

H.M., Dubois, J., Qiu, D., Hitti, J., Production, G.T.C.S.C., Finishing, BioinformaticsTeams Gtc Sequencing Center, G.T.C.W.M., Department of Pathology, U.S.U.o.t.H.S.,The National Center for Biotechnology Information, T.N.I.o.H.B.M., Insa,D.d.G.B.T.c.F., Department of, B., Cell Biology, R.U.H.T, Wolf, Y.I., Tatusov, R.L.,Sabathe, F., Doucette-Stamm, L., Soucaille, P., Daly, M.J., Bennett, G.N., Koonin,E.V., Smith, D.R., 2001. Genome Sequence and Comparative Analysis of the Solvent-Producing Bacterium Clostridium acetobutylicum. Journal of Bacteriology 183 (16),4823–4838.

Nybo, S.E., Saunders, J., McCormick, S.P., 2017. Metabolic engineering of Escherichiacoli for production of valerenadiene. Journal of Biotechnology 262, 60–66.

Octave, S., Thomas, D., 2009. Biorefinery: toward an industrial metabolism. Biochimie 91(6), 659–664.

Oh, B.-R., Seo, J.-W., Heo, S.-Y., Hong, W.-K., Luo, L.H., Joe, M.-h., Park, D.-H., Kim, C.H.,2011. Efficient production of ethanol from crude glycerol by a Klebsiella pneumoniaemutant strain. Bioresource technology 102 (4), 3918–3922.

Oh, B.-R., Seo, J.-W., Heo, S.-Y., Hong, W.-K., Luo, L.H., Kim, S., Kwon, O., Sohn, J.-H.,Joe, M.-h., Park, D.-H., Kim, C.H., 2012a. Enhancement of ethanol production fromglycerol in a Klebsiella pneumoniae mutant strain by the inactivation of lactate de-hydrogenase. Process Biochemistry 47 (1), 156–159.

Oh, B.-R., Seo, J.-W., Heo, S.-Y., Hong, W.-K., Luo, L.H., Kim, S., Park, D.-H., Kim, C.H.,2012b. Optimization of Culture Conditions for 1,3-Propanediol Production fromGlycerol Using a Mutant Strain of Klebsiella pneumoniae. Applied Biochemistry andBiotechnology 166 (1), 127–137.

Oh, B.-R., Hong, W.-K., Heo, S.-Y., Luo, L.H., Kondo, A., Seo, J.-W., Kim, C.H., 2013a. Theproduction of 1, 3-propanediol from mixtures of glycerol and glucose by a Klebsiellapneumoniae mutant deficient in carbon catabolite repression. Bioresource technology130, 719–724.

Oh, B.-R., Seo, J.-W., Heo, S.-Y., Luo, L.H., Hong, W.-K., Park, D.-H., Kim, C.-H., 2013b.Efficient production of 1,3-propanediol from glycerol upon constitutive expression ofthe 1,3-propanediol oxidoreductase gene in engineered Klebsiella pneumoniae withelimination of by-product formation. Bioprocess and Biosystems Engineering 36 (6),757–763.

Oh, B.-R., Heo, S.-Y., Lee, S.-M., Hong, W.-K., Park, J.M., Jung, Y.R., Kim, D.-H., Sohn, J.-H., Seo, J.-W., Kim, C.H., 2014. Production of 2-butanol from crude glycerol by agenetically-engineered Klebsiella pneumoniae strain. Biotechnology Letters 36 (1),57–62.

Oh, B.-R., Lee, S.-M., Heo, S.-Y., Seo, J.-W., Kim, C.H., 2018. Efficient production of 1, 3-propanediol from crude glycerol by repeated fed-batch fermentation strategy of alactate and 2, 3-butanediol deficient mutant of Klebsiella pneumoniae. Microbial cellfactories 17 (1), 92.

Ohtake, T., Pontrelli, S., Laviña, W.A., Liao, J.C., Putri, S.P., Fukusaki, E., 2017.Metabolomics-driven approach to solving a CoA imbalance for improved 1-butanolproduction in Escherichia coli. Metabolic Engineering 41, 135–143.

Otte, B., Grunwaldt, E., Mahmoud, O., Jennewein, S., 2009. Genome Shuffling inClostridium diolis DSM 15410 for Improved 1,3-Propanediol Production. Applied andEnvironmental Microbiology 75 (24), 7610–7616.

Parikka, M., 2004. Global biomass fuel resources. Biomass and Bioenergy 27 (6),613–620.

Park, J.M., Rathnasingh, C., Song, H., 2017. Metabolic engineering of Klebsiella pneu-moniae based on in silico analysis and its pilot-scale application for 1, 3-propanedioland 2, 3-butanediol co-production. Journal of industrial microbiology & bio-technology 44 (3), 431–441.

Peguin, S., Goma, G., Delorme, P., Soucaille, P., 1994. Metabolic flexibility of Clostridiumacetobutylicum in response to methyl viologen addition. Applied Microbiology andBiotechnology 42 (4), 611–616.

Pemberton, J., Vincent, K., Penfold, R., 1991. Cloning and heterologous expression of theviolacein biosynthesis gene cluster from Chromobacterium violaceum. CurrentMicrobiology 22 (6), 355–358.

Pflugl, S., Marx, H., Mattanovich, D., Sauer, M., 2012. 1,3-Propanediol production fromglycerol with Lactobacillus diolivorans. Bioresource Technology 119, 133–140.

Pflugl, S., Marx, H., Mattanovich, D., Sauer, M., 2013. Genetic engineering of Lactobacillusdiolivorans. FEMS Microbiology Letters 344 (2), 152–158.

Pflugl, S., Marx, H., Mattanovich, D., Sauer, M., 2014. Heading for an economic industrialupgrading of crude glycerol from biodiesel production to 1,3-propanediol byLactobacillus diolivorans. Bioresource Technology 152, 499–504.

Phithakrotchanakoon, C., Champreda, V., Aiba, S.-i., Pootanakit, K., Tanapongpipat, S.,2013. Engineered Escherichia coli for Short-Chain-Length Medium-Chain-LengthPolyhydroxyalkanoate Copolymer Biosynthesis from Glycerol and Dodecanoate.Bioscience, Biotechnology, and Biochemistry 77 (6), 1262–1268.

Phithakrotchanakoon, C., Champreda, V., Aiba, S.-i., Pootanakit, K., Tanapongpipat, S.,2015. Production of Polyhydroxyalkanoates from Crude Glycerol Using RecombinantEscherichia coli. Journal of Polymers and the Environment 23 (1), 38–44.

Poehlein, A., Schlien, K., Chowdhury, N.P., Gottschalk, G., Buckel, W., Daniel, R., 2016.Complete Genome Sequence of the Amino Acid-Fermenting Clostridium propionicumX2 (DSM 1682). Genome Announcements 4 (2) (e00294-00216).

Pöschl, M., Ward, S., Owende, P., 2010. Evaluation of energy efficiency of various biogasproduction and utilization pathways. Applied Energy 87 (11), 3305–3321.

Przystałowska, H., Zeyland, J., Szymanowska-Powałowska, D., Szalata, M., Słomski, R.,Lipiński, D., 2015. 1, 3-Propanediol production by new recombinant Escherichia colicontaining genes from pathogenic bacteria. Microbiological research 171, 1–7.

Pyne, M.E., Moo-Young, M., Chung, D.A., Chou, C.P., 2013. Development of an electro-transformation protocol for genetic manipulation of Clostridium pasteurianum.Biotechnology for biofuels 6 (1), 50.

Pyne, M.E., Bruder, M., Moo-Young, M., Chung, D.A., Chou, C.P., 2014. Technical guide

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

565

Page 29: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

for genetic advancement of underdeveloped and intractable Clostridium.Biotechnology Advances 32 (3), 623–641.

Pyne, M.E., Moo-Young, M., Chung, D.A., Chou, C.P., 2015. Antisense-RNA-mediatedgene downregulation in Clostridium pasteurianum. Fermentation 1 (1), 113–126.

Pyne, M.E., Bruder, M.R., Moo-Young, M., Chung, D.A., Chou, C.P., 2016a. Harnessingheterologous and endogenous CRISPR-Cas machineries for efficient markerlessgenome editing in Clostridium. Scientific reports 6.

Pyne, M.E., Sokolenko, S., Liu, X., Srirangan, K., Bruder, M.R., Aucoin, M.G., Moo-Young,M., Chung, D.A., Chou, C.P., 2016b. Disruption of the reductive 1,3-propanediolpathway triggers production of 1,2-propanediol for sustained glycerol fermentationby Clostridium pasteurianum. Applied and Environmental Microbiology. 82 (17),5375–5388.

Qi, Lei S., Larson, Matthew H., Gilbert, Luke A., Doudna, Jennifer A., Weissman, JonathanS., Arkin, Adam P., Lim, Wendell A., 2013. Repurposing CRISPR as an RNA-guidedplatform for sequence-specific control of gene expression. Cell 152 (5), 1173–1183.

Raj, S.M., Rathnasingh, C., Jo, J.-E., Park, S., 2008. Production of 3-hydroxypropionicacid from glycerol by a novel recombinant Escherichia coli BL21 strain. ProcessBiochemistry 43 (12), 1440–1446.

Raj, S.M., Rathnasingh, C., Jung, W.-C., Park, S., 2009. Effect of process parameters on 3-hydroxypropionic acid production from glycerol using a recombinant Escherichiacoli. Applied microbiology and biotechnology 84 (4), 649–657.

Ramakrishnan, G.G., Nehru, G., Suppuram, P., Balasubramaniyam, S., Gulab, B.R.,Subramanian, R., 2015. Bio-transformation of glycerol to 3-hydroxypropionic acidusing resting cells of Lactobacillus reuteri. Current Microbiology 71 (4), 517–523.

Rathnasingh, C., Raj, S.M., Jo, J.-E., Park, S., 2009. Development and evaluation of ef-ficient recombinant Escherichia coli strains for the production of 3-hydroxypropionicacid from glycerol. Biotechnology and Bioengineering 104 (4), 729–739.

Raynaud, C., Sarçabal, P., Meynial-Salles, I., Croux, C., Soucaille, P., 2003. MolecularCharacterization of the 1,3-Propanediol (1,3-PD) Operon of Clostridium butyricum.Proceedings of the National Academy of Sciences of the United States of America 100(9), 5010–5015.

Reddy, C.S.K., Ghai, R., Rashmi, Kalia, V.C., 2003. Polyhydroxyalkanoates: an overview.Bioresource Technology 87 (2), 137–146.

Reimann, A., Biebl, H., 1996. Production of 1,3-propanediol by Clostridium butyricumDSM 5431 and product tolerant mutants in fedbatch culture: Feeding strategy forglycerol and ammonium. Biotechnology Letters 18 (7), 827–832.

Rettori, D., Durán, N., 1998. Production, extraction and purificationof violacein: an an-tibiotic pigment producedby Chromobacterium violaceum. World Journal ofMicrobiology and Biotechnology 14 (5), 685–688.

Ricci, M.A., Russo, A., Pisano, I., Palmieri, L., de Angelis, M., Agrimi, G., 2015. Improved1,3-Propanediol Synthesis from Glycerol by the Robust Lactobacillus reuteri StrainDSM 20016. Journal of microbiology and biotechnology 25 (6), 893–902.

Richey, D., Lin, E., 1972. Importance of facilitated diffusion for effective utilization ofglycerol by Escherichia coli. Journal of bacteriology 112 (2), 784–790.

Rodrigues, A.L., Trachtmann, N., Becker, J., Lohanatha, A.F., Blotenberg, J., Bolten, C.J.,Korneli, C., de Souza Lima, A.O., Porto, L.M., Sprenger, G.A., 2013. Systems meta-bolic engineering of Escherichia coli for production of the antitumor drugs violaceinand deoxyviolacein. Metabolic engineering 20, 29–41.

Rodrigues, A.L., Becker, J., de Souza Lima, A.O., Porto, L.M., Wittmann, C., 2014. Systemsmetabolic engineering of Escherichia coli for gram scale production of the antitumordrug deoxyviolacein from glycerol. Biotechnology and bioengineering 111 (11),2280–2289.

Sabet-Azad, R., Linares-Pastén, J.A., Torkelson, L., Sardari, R.R.R., Hatti-Kaul, R., 2013.Coenzyme A-acylating propionaldehyde dehydrogenase (PduP) from Lactobacillusreuteri: Kinetic characterization and molecular modeling. Enzyme and MicrobialTechnology 53 (4), 235–242.

Sabourin-Provost, G., Hallenbeck, P.C., 2009. High yield conversion of a crude glycerolfraction from biodiesel production to hydrogen by photofermentation. Bioresourcetechnology 100 (14), 3513–3517.

Saini, M., Wang, Z.W., Chiang, C.-J., Chao, Y.-P., 2017. Metabolic engineering ofEscherichia coli for production of n-butanol from crude glycerol. Biotechnology forbiofuels 10 (1), 173.

Saint-Amans, S., Perlot, P., Goma, G., Soucaille, P., 1994. High production of 1,3-pro-panediol from glycerol by Clostridium butyricum VPI 3266 in a simply controlled fed-batch system. Biotechnology Letters 16 (8), 831–836.

Saint-Amans, S., Girbal, L., Andrade, J., Ahrens, K., Soucaille, P., 2001. Regulation ofCarbon and Electron Flow inClostridium butyricum VPI 3266 Grown on Glucose-Glycerol Mixtures. Journal of Bacteriology 183 (5), 1748–1754.

Sánchez, C., Braña, A.F., Méndez, C., Salas, J.A., 2006. Reevaluation of the ViolaceinBiosynthetic Pathway and its Relationship to Indolocarbazole Biosynthesis.ChemBioChem 7 (8), 1231–1240.

Sanchez-Torres, V., Yusoff, M.Z.M., Nakano, C., Maeda, T., Ogawa, H.I., Wood, T.K.,2013. Influence of Escherichia coli hydrogenases on hydrogen fermentation fromglycerol. International Journal of Hydrogen Energy 38 (10), 3905–3912.

Sandoval, N.R., Venkataramanan, K.P., Groth, T.S., Papoutsakis, E.T., 2015. Whole-genome sequence of an evolved Clostridium pasteurianum strain reveals Spo0A de-ficiency responsible for increased butanol production and superior growth.Biotechnology for Biofuels 8 (1), 1–18.

Sankaranarayanan, M., Ashok, S., Park, S., 2014. Production of 3-hydroxypropionic acidfrom glycerol by acid tolerant Escherichia coli. Journal of industrial microbiology &biotechnology 41 (7), 1039–1050.

Santos, F., Spinler, J.K., Saulnier, D.M., Molenaar, D., Teusink, B., de Vos, W.M.,Versalovic, J., Hugenholtz, J., 2011. Functional identification in Lactobacillus reuteriof a PocR-like transcription factor regulating glycerol utilization and vitamin B12synthesis. Microbial Cell Factories 10 (1), 1.

Sauvageot, N., Gouffi, K., Laplace, J.-M., Auffray, Y., 2000. Glycerol metabolism in

Lactobacillus collinoides: production of 3-hydroxypropionaldehyde, a precursor ofacrolein. International journal of food microbiology 55 (1), 167–170.

Sauvageot, N., Muller, C., Hartke, A., Auffray, Y., Laplace, J.-M., 2002. Characterisationof the diol dehydratase pdu operon of Lactobacillus collinoides. FEMS MicrobiologyLetters 209 (1), 69–74.

Sawers, R.G., Boxer, D.H., 1986. Purification and properties of membrane-bound hy-drogenase isoenzyme 1 from anaerobically grown Escherichia coli K12. EuropeanJournal of Biochemistry 156 (2), 265–275.

Sawers, R.G., Ballantine, S.P., Boxer, D.H., 1985. Differential expression of hydrogenaseisoenzymes in Escherichia coli K-12: evidence for a third isoenzyme. Journal ofBacteriology 164 (3), 1324–1331.

Sawers, R.G., Blokesch, M., Böck, A., 2004. Anaerobic formate and hydrogen metabolism.EcoSal Plus 1 (1).

Schmidhuber, J., 2008. Impact of an increased biomass use on agricultural markets,prices and food security: A longer-term perspective, Energy Security in Europe:Proceedings from the Conference ‘Energy Security in Europe. pp. 133-170.

Schütz, H., Radler, F., 1984. Anaerobic reduction of glycerol to propanediol-1.3 byLactobacillus brevis and Lactobacillus buchneri. Systematic and AppliedMicrobiology 5 (2), 169–178.

Schwarz, K.M., Grosse-Honebrink, A., Derecka, K., Rotta, C., Zhang, Y., Minton, N.P.,2017. Towards improved butanol production through targeted genetic modificationof Clostridium pasteurianum. Metabolic engineering 40, 124–137.

Seedorf, H., Fricke, W.F., Veith, B., Brüggemann, H., Liesegang, H., Strittmatter, A.,Miethke, M., Buckel, W., Hinderberger, J., Li, F., Hagemeier, C., Thauer, R.K.,Gottschalk, G., 2008. The genome of Clostridium kluyveri, a strict anaerobe withunique metabolic features. Proceedings of the National Academy of Sciences of theUnited States of America 105 (6), 2128–2133.

Seifert, C., Bowien, S., Gottschalk, G., Daniel, R., 2001. Identification and expression ofthe genes and purification and characterization of the gene products involved inreactivation of coenzyme B12-dependent glycerol dehydratase of Citrobacterfreundii. European Journal of Biochemistry 268 (8), 2369–2378.

Seo, M.-Y., Seo, J.-W., Heo, S.-Y., Baek, J.-O., Rairakhwada, D., Oh, B.-R., Seo, P.-S., Choi,M.H., Kim, C.H., 2009. Elimination of by-product formation during production of 1,3-propanediol in Klebsiella pneumoniae by inactivation of glycerol oxidative pathway.Applied Microbiology and Biotechnology 84 (3), 527–534.

Seo, J.-W., Seo, M.-Y., Oh, B.-R., Heo, S.-Y., Baek, J.-O., Rairakhwada, D., Luo, L.H.,Hong, W.-K., Kim, C.H., 2010. Identification and utilization of a 1,3-propanedioloxidoreductase isoenzyme for production of 1,3-propanediol from glycerol inKlebsiella pneumoniae. Applied Microbiology and Biotechnology 85 (3), 659–666.

Shah, P., Chiu, F.-S., Lan, J.C.-W., 2014. Aerobic utilization of crude glycerol by re-combinant Escherichia coli for simultaneous production of poly 3-hydroxybutyrateand bioethanol. Journal of Bioscience and Bioengineering 117 (3), 343–350.

Shalel Levanon, S., San, K.-Y., Bennett, G.N., 2005. Effect of oxygen on the Escherichiacoli ArcA and FNR regulation systems and metabolic responses. Biotechnology andBioengineering 89 (5), 556–564.

Shalel-Levanon, S., San, K.-Y., Bennett, G.N., 2005. Effect of oxygen, and ArcA and FNRregulators on the expression of genes related to the electron transfer chain and theTCA cycle in Escherichia coli. Metabolic engineering 7 (5-6), 364–374.

Shams Yazdani, S., Gonzalez, R., 2008. Engineering Escherichia coli for the efficientconversion of glycerol to ethanol and co-products. Metab. Eng. 10 (6), 340–351.

Shen, X., Lin, Y., Jain, R., Yuan, Q., Yan, Y., 2012. Inhibition of acetate accumulationleads to enhanced production of (R,R)-2,3-butanediol from glycerol in Escherichiacoli. Journal of Industrial Microbiology & Biotechnology 39(11), 1725-1729.

Siebold, C., Arnold, I., Garcia-Alles, L.F., Baumann, U., Erni, B., 2003. Crystal Structure ofthe Citrobacter freundii Dihydroxyacetone Kinase Reveals an Eight-stranded α-Helical Barrel ATP-binding Domain. Journal of Biological Chemistry 278 (48),48236–48244.

Singh, S., Mohanty, A.K., 2007. Wood fiber reinforced bacterial bioplastic composites:Fabrication and performance evaluation. Composites Science and Technology 67 (9),1753–1763.

Skraly, F.A., Lytle, B.L., Cameron, D.C., 1998. Construction and characterization of a 1, 3-propanediol operon. Applied and environmental microbiology 64 (1), 98–105.

Sprenger, G.A., Hammer, B.A., Johnson, E.A., Lin, E.C., 1989. Anaerobic Growth ofEscheuichia coli on Glycerol by Importing Genes of the dha Regulon from Klebsiellapneumoniae. Journal of General Microbiology 135, 1255–1262.

Sriramulu, D.D., Liang, M., Hernandez-Romero, D., Raux-Deery, E., Lünsdorf, H., Parsons,J.B., Warren, M.J., Prentice, M.B., 2008. Lactobacillus reuteri DSM 20016 producescobalamin-dependent diol dehydratase in metabolosomes and metabolizes 1, 2-pro-panediol by disproportionation. Journal of bacteriology 190 (13), 4559–4567.

Srirangan, K., Akawi, L., Moo-Young, M., Chou, C.P., 2012. Towards sustainable pro-duction of clean energy carriers from biomass resources. Applied Energy 100,172–186.

Srirangan, K., Akawi, L., Liu, X., Westbrook, A., Blondeel, E.J., Aucoin, M.G., Moo-Young,M., Chou, C.P., 2013. Manipulating the sleeping beauty mutase operon for the pro-duction of 1-propanol in engineered Escherichia coli. Biotechnology for biofuels 6(1), 1.

Srirangan, K., Liu, X., Westbrook, A., Akawi, L., Pyne, M.E., Moo-Young, M., Chou, C.P.,2014. Biochemical, genetic, and metabolic engineering strategies to enhance copro-duction of 1-propanol and ethanol in engineered Escherichia coli. Applied micro-biology and biotechnology 98 (22), 9499–9515.

Srirangan, K., Liu, X., Akawi, L., Bruder, M., Moo-Young, M., Chou, C.P., 2016a.Engineering Escherichia coli for microbial production of butanone. Applied and en-vironmental microbiology 82 (9), 2574–2584.

Srirangan, K., Liu, X., Tran, T.T., Charles, T.C., Moo-Young, M., Chou, C.P., 2016b.Engineering of Escherichia coli for direct and modulated biosynthesis of poly(3-hy-droxybutyrate-co-3-hydroxyvalerate) copolymer using unrelated carbon sources.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

566

Page 30: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Scientific Reports 6, 36470.Stevens, M.J., Vollenweider, S., Meile, L., Lacroix, C., 2011. 1, 3-Propanediol dehy-

drogenases in Lactobacillus reuteri: impact on central metabolism and 3-hydro-xypropionaldehyde production. Microbial cell factories 10 (1), 1.

Su, M., Li, Y., Ge, X., Tian, P., 2015. 3-Hydroxypropionaldehyde-specific aldehyde de-hydrogenase from Bacillus subtilis catalyzes 3-hydroxypropionic acid production inKlebsiella pneumoniae. Biotechnology Letters 37 (3), 717–724.

Sun, J., van den Heuvel, J., Soucaille, P., Qu, Y., Zeng, A.-P., 2003. Comparative GenomicAnalysis of dha Regulon and Related Genes for Anaerobic Glycerol Metabolism inBacteria. Biotechnology Progress 19 (2), 263–272.

Sun, Y.Q., Qi, W.T., Teng, H., Xiu, Z.L., Zeng, A.P., 2008. Mathematical modeling ofglycerol fermentation by Klebsiella pneumoniae: Concerning enzyme-catalytic re-ductive pathway and transport of glycerol and 1,3-propanediol across cell membrane.Biochemical Engineering Journal 38 (1), 22–32.

Sun, Y.-Q., Shen, J.-T., Yan, L., Zhou, J.-J., Jiang, L.-L., Chen, Y., Yuan, J.-L., Feng, E.-M.,Xiu, Z.-L., 2018. Advances in bioconversion of glycerol to 1, 3-propanediol: prospectsand challenges. Process Biochemistry. 71, 134–146.

Suthers, P.F., Cameron, D.C., 2005. Production of 3-hydroxypropionic acid in re-combinant organisms United States.

Syu, M.-J., 2001. Biological production of 2,3-butanediol. Applied Microbiology andBiotechnology 55 (1), 10–18.

Talarico, T.L., Dobrogosz, W.J., 1990. Purification and characterization of glycerol de-hydratase from Lactobacillus reuteri. Applied and environmental microbiology 56(4), 1195–1197.

Talarico, T.L., Axelsson, L.T., Novotny, J., Fiuzat, M., Dobrogosz, W.J., 1990. Utilizationof glycerol as a hydrogen acceptor by Lactobacillus reuteri: purification of 1, 3-pro-panediol: NAD+ oxidoreductase. Applied and Environmental Microbiology 56 (4),943–948.

Tan, Z., Khakbaz, P., Chen, Y., Lombardo, J., Yoon, J.M., Shanks, J.V., Klauda, J.B.,Jarboe, L.R., 2017. Engineering Escherichia coli membrane phospholipid head dis-tribution improves tolerance and production of biorenewables. MetabolicEngineering 44, 1–12.

Tang, C., Ruch, F., Lin, C., 1979. Purification and properties of a nicotinamide adeninedinucleotide-linked dehydrogenase that serves an Escherichia coli mutant for glycerolcatabolism. Journal of bacteriology 140 (1), 182–187.

Tang, X., Tan, Y., Zhu, H., Zhao, K., Shen, W., 2009. Microbial Conversion of Glycerol to1,3-Propanediol by an Engineered Strain of Escherichia coli. Applied andEnvironmental Microbiology 75 (6), 1628–1634.

Taranto, M.P., Vera, J.L., Hugenholtz, J., De Valdez, G.F., Sesma, F., 2003. Lactobacillusreuteri CRL1098 Produces Cobalamin. Journal of Bacteriology 185 (18), 5643–5647.

Teeka, J., Imai, T., Reungsang, A., Cheng, X., Yuliani, E., Thiantanankul, J., Poomipuk,N., Yamaguchi, J., Jeenanong, A., Higuchi, T., Yamamoto, K., Sekine, M., 2012.Characterization of polyhydroxyalkanoates (PHAs) biosynthesis by isolatedNovosphingobium sp. THA_AIK7 using crude glycerol. Journal of IndustrialMicrobiology & Biotechnology 39 (5), 749–758.

Tian, K., Shi, G., Lu, F., Singh, S., Wang, Z., 2013. High-efficiency L-lactate productionfrom glycerol by metabolically engineered Escherichia coli. Chinese Journal ofBiotechnology 29 (9), 1268–1277.

Tokuyama, K., Ohno, S., Yoshikawa, K., Hirasawa, T., Tanaka, S., Furusawa, C., Shimizu,H., 2014. Increased 3-hydroxypropionic acid production from glycerol, by mod-ification of central metabolism in Escherichia coli. Microbial Cell Factories 13, 64.

Tong, I.-T., Liao, H.H., Cameron, D.C., 1991. 1, 3-Propanediol production by Escherichiacoli expressing genes from the Klebsiella pneumoniae dha regulon. Applied and en-vironmental microbiology 57 (12), 3541–3546.

Tong, W., Xu, Y., Xian, M., Niu, W., Guo, J., Liu, H., Zhao, G., 2016. Biosynthetic pathwayfor acrylic acid from glycerol in recombinant Escherichia coli. Applied microbiologyand biotechnology 100 (11), 4901–4907.

Toraya, T., 2000. Radical catalysis of B12 enzymes: structure, mechanism, inactivation,and reactivation of diol and glycerol dehydratases. Cellular and Molecular LifeSciences CMLS 57 (1), 106–127.

Tracy, B.P., Jones, S.W., Fast, A.G., Indurthi, D.C., Papoutsakis, E.T., 2012. Clostridia: theimportance of their exceptional substrate and metabolite diversity for biofuel andbiorefinery applications. Current Opinion in Biotechnology 23 (3), 364–381.

Tran, Q.H., Bongaerts, J., Vlad, D., Unden, G., 1997. Requirement for the proton-pumpingNADH dehydrogenase i of Escherichia coli in respiration of NADH to fumarate and itsbioenergetic implications. The FEBS Journal 244 (1), 155–160.

Tran, K.T., Maeda, T., Wood, T.K., 2014. Metabolic engineering of Escherichia coli toenhance hydrogen production from glycerol. Applied Microbiology andBiotechnology 98 (10), 4757–4770.

Tran, K.T., Maeda, T., Sanchez-Torres, V., Wood, T.K., 2015. Beneficial knockouts inEscherichia coli for producing hydrogen from glycerol. Applied Microbiology andBiotechnology 99 (6), 2573–2581.

Tseng, H.-C., Harwell, C.L., Martin, C.H., Prather, K.L., 2010. Biosynthesis of chiral 3-hydroxyvalerate from single propionate-unrelated carbon sources in metabolicallyengineered E. coli. Microbial Cell Factories 9 (1), 1–12.

Tsuruno, K., Honjo, H., Hanai, T., 2015. Enhancement of 3-hydroxypropionic acid pro-duction from glycerol by using a metabolic toggle switch. Microbial cell factories 14(1), 1.

Unden, G., Kleefeld, A., 2004. C4-dicarboxylate degradation in aerobic and anaerobicgrowth. EcoSal Plus 1 (1).

Vaidyanathan, H., Kandasamy, V., Ramakrishnan, G.G., Ramachandran, K., Jayaraman,G., Ramalingam, S., 2011. Glycerol conversion to 1, 3-Propanediol is enhanced by theexpression of a heterologous alcohol dehydrogenase gene in Lactobacillus reuteri.AMB express 1 (1), 1.

van Pijkeren, J.-P., Neoh, K.M., Sirias, D., Findley, A.S., Britton, R.A., 2012. Exploringoptimization parameters to increase ssDNA recombineering in Lactococcus lactis and

Lactobacillus reuteri. Bioengineered 3 (4), 209–217.Vasconcelos, I., Girbal, L., Soucaille, P., 1994. Regulation of carbon and electron flow in

Clostridium acetobutylicum grown in chemostat culture at neutral pH on mixtures ofglucose and glycerol. Journal of Bacteriology 176 (5), 1443–1450.

Wang, W., Sun, J., Hartlep, M., Deckwer, W.-D., Zeng, A.-P., 2003. Combined use ofproteomic analysis and enzyme activity assays for metabolic pathway analysis ofglycerol fermentation by Klebsiella pneumoniae. Biotechnology and Bioengineering83 (5), 525–536.

Wang, Q., Yang, P., Liu, C., Xue, Y., Xian, M., Zhao, G., 2013a. Biosynthesis of poly(3-hydroxypropionate) from glycerol by recombinant Escherichia coli. BioresourceTechnology 131, 548–551.

Wang, Y., Tao, F., Tang, H., Xu, P., 2013b. Genome Sequence of Clostridium diolis StrainDSM 15410, a Promising Natural Producer of 1,3-Propanediol. GenomeAnnouncements 1 (4).

Wang, M., Fan, L., Tan, T., 2014a. 1-Butanol production from glycerol by engineeredKlebsiella pneumoniae. RSC Advances 4 (101), 57791–57798.

Wang, Q., Yang, P., Xian, M., Feng, L., Wang, J., Zhao, G., 2014b. Metabolic engineeringof Escherichia coli for poly(3-hydroxypropionate) production from glycerol andglucose. Biotechnology Letters 36 (11), 2257–2262.

Wang, Y., Tao, F., Xu, P., 2014c. Glycerol dehydrogenase plays a dual role in glycerolmetabolism and 2, 3-butanediol formation in Klebsiella pneumoniae. Journal ofBiological Chemistry, jbc. M113. 525535.

Wang, M., Hu, L., Fan, L., Tan, T., 2015a. Enhanced 1-Butanol Production in EngineeredKlebsiella pneumoniae by NADH Regeneration. Energy & Fuels 29 (3), 1823–1829.

Wang, Z.W., Saini, M., Lin, L.-J., Chiang, C.-J., Chao, Y.-P., 2015b. Systematic engineeringof Escherichia coli for D-lactate production from crude glycerol. Journal ofAgricultural and Food Chemistry 63 (43), 9583–9589.

Wang, M., Liu, L., Fan, L., Tan, T., 2017a. CRISPRi based system for enhancing 1-butanolproduction in engineered Klebsiella pneumoniae. Process Biochemistry 56, 139–146.

Wang, Y., Tao, F., Xin, B., Liu, H., Gao, Y., Zhou, N.-Y., Xu, P., 2017b. Switch of metabolicstatus: redirecting metabolic flux for acetoin production from glycerol by activating asilent glycerol catabolism pathwayc. Metabolic engineering 39, 90–101.

Wang, M., Wang, G., Zhang, T., Fan, L., Tan, T., 2017c. Multi-modular engineering of 1, 3-propanediol biosynthesis system in Klebsiella pneumoniae from co-substrate. Appliedmicrobiology and biotechnology 101 (2), 647–657.

Wang, M., Zhou, Y., Tan, T., 2017d. Cofactor Engineering for Enhanced Production ofDiols by Klebsiella pneumoniae From Co-Substrate. Biotechnology Journal. 12 (11),1700176 (6 pages).

Wang, H., Liu, W., Shi, F., Huang, L., Lian, J., Qu, L., Cai, J., Xu, Z., 2018. MetabolicPathway Engineering for High-level Production of 5-Hydroxytryptophan inEscherichia coli. Metabolic Engineering. 48, 279–287.

Wei, D., Wang, M., Jiang, B., Shi, J., Hao, J., 2014. Role of dihydroxyacetone kinases Iand II in the dha regulon of Klebsiella pneumoniae. Journal of Biotechnology 177,13–19.

Weiner, M., Albermann, C., Gottlieb, K., Sprenger, G.A., Weuster-Botz, D., 2014. Fed-batch production of L-phenylalanine from glycerol and ammonia with recombinantEscherichia coli. Biochemical Engineering Journal 83, 62–69.

Westbrook, A.W., Moo-Young, M., Chou, C.P., 2016. Development of a CRISPR-Cas9 toolkit for comprehensive engineering of Bacillus subtilis. Applied and EnvironmentalMicrobiology 82 (16), 4876–4895.

Westbrook, A.W., Ren, X., Moo-Young, M., Chou, C.P., 2018. Engineering of cell mem-brane to enhance heterologous production of hyaluronic acid in Bacillus subtilis.Biotechnology and Bioengineering. 115, 216–231.

Wischral, D., Zhang, J., Cheng, C., Lin, M., De Souza, L.M.G., Pessoa, F.L.P., Pereira Jr.,N., Yang, S.-T., 2016. Production of 1,3-propanediol by Clostridium beijerinckii DSM791 from crude glycerol and corn steep liquor: Process optimization and metabolicengineering. Bioresource Technology 212, 100–110.

Wong, M.S., Li, M., Black, R.W., Le, T.Q., Puthli, S., Campbell, P., Monticello, D.J., 2014.Microaerobic conversion of glycerol to ethanol in Escherichia coli. Applied andEnvironmental Microbiology 80 (10), 3276–3282.

Wu, Z., Wang, Z., Wang, G., Tan, T., 2013. Improved 1,3-propanediol production byengineering the 2,3-butanediol and formic acid pathways in integrative recombinantKlebsiella pneumoniae. Journal of Biotechnology 168 (2), 194–200.

Wu, H., Karanjikar, M., San, K.-Y., 2014. Metabolic engineering of Escherichia coli forefficient free fatty acid production from glycerol. Metabolic Engineering 25, 82–91.

Xin, B., Tao, F., Wang, Y., Liu, H., Ma, C., Xu, P., 2017. Coordination of metabolicpathways: Enhanced carbon conservation in 1, 3-propanediol production by couplingwith optically pure lactate biosynthesis. Metabolic Engineering 41, 102–114.

Xiu, Z.-L., Zeng, A.-P., 2008. Present state and perspective of downstream processing ofbiologically produced 1,3-propanediol and 2,3-butanediol. Applied Microbiology andBiotechnology 78 (6), 917–926.

Xiu, Z.-L., Chen, X., Sun, Y.-Q., Zhang, D.-J., 2007. Stoichiometric analysis and experi-mental investigation of glycerol–glucose co-fermentation in Klebsiella pneumoniaeunder microaerobic conditions. Biochemical Engineering Journal 33 (1), 42–52.

Xu, Y.-Z., Guo, N.-N., Zheng, Z.-M., Ou, X.-J., Liu, H.-J., Liu, D.-H., 2009. Metabolism in1,3-propanediol fed-batch fermentation by a D-lactate deficient mutant of Klebsiellapneumoniae. Biotechnology and Bioengineering 104 (5), 965–972.

Yang, Z., Zhang, Z., 2018. Recent advances on production of 2, 3-butanediol using en-gineered microbes. Biotechnology advances. https://doi.org/10.1016/j.biotechadv.2018.03.019.

Yang, G., Tian, J., Li, J., 2007. Fermentation of 1,3-propanediol by a lactate deficientmutant of Klebsiella oxytoca under microaerobic conditions. Applied Microbiologyand Biotechnology 73 (5), 1017–1024.

Yang, C., Jiang, P., Xiao, S., Zhang, C., Lou, K., Xing, X.-H., 2011. Fed-batch fermentationof recombinant Citrobacter freundii with expression of a violacein-synthesizing genecluster for efficient violacein production from glycerol. Biochemical Engineering

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

567

Page 31: Strain engineering for microbial production of value-added ... · 2,3-butanediol (2,3-BDO) (Clomburg and Gonzalez, 2013; Murarka et al., 2008; Yazdani and Gonzalez, 2007). Accordingly,

Journal 57, 55–62.Yang, W., Zhou, Y., Zhao, Z.K., 2013. Production of dihydroxyacetone from glycerol by

engineered Escherichia coli cells co-expressing gldA and nox genes. African Journal ofBiotechnology 12 (27).

Yang, T., Rao, Z., Zhang, X., Xu, M., Xu, Z., Yang, S.-T., 2017. Metabolic engineeringstrategies for acetoin and 2, 3-butanediol production: advances and prospects.Critical reviews in biotechnology 37 (8), 990–1005.

Yang, B., Liang, S., Liu, H., Liu, J., Cui, Z., Wen, J., 2018. Metabolic engineering ofEscherichia coli for 1, 3-propanediol biosynthesis from glycerol. Bioresource tech-nology 267, 599–607.

Yazdani, S.S., Gonzalez, R., 2007. Anaerobic fermentation of glycerol: a path to economicviability for the biofuels industry. Current opinion in biotechnology 18 (3), 213–219.

Yazdani, S.S., Gonzalez, R., 2008. Engineering Escherichia coli for the efficient conversionof glycerol to ethanol and co-products. Metabolic Engineering 10 (6), 340–351.

Li, Y., Wang, X., Ge, X., Tian, P., 2016. High Production of 3-Hydroxypropionic Acid inKlebsiella pneumoniae by Systematic Optimization of Glycerol Metabolism. ScientificReports 6, 26932.

Ye, L., Zhang, C., Bi, C., Li, Q., Zhang, X., 2016. Combinatory optimization of chromo-somal integrated mevalonate pathway for β-carotene production in Escherichia coli.Microbial cell factories 15 (1), 202.

Zeng, A.P., Biebl, H., 2002. Bulk Chemicals from Biotechnology: The Case of 1,3-Propanediol Production and the New Trends. In: Schügerl, K., Zeng, A.P., Aunins,J.G., Bader, A., Bell, W., Biebl, H., Biselli, M., Carrondo, M.J.T., Castilho, L.R., Chang,H.N., Cruz, P.E., Fuchs, C., Han, S.J., Han, M.R., Heinzle, E., Hitzmann, B., Köster, D.,Jasmund, I., Jelinek, N., Lang, S., Laatsch, H., Lee, J., Miirkl, H., Maranga, L.,Medronho, R.A., Meiners, M., Nath, S., Noll, T., Scheper, T., Schmidt, S., Schüigerl,K., Stäirk, E., Tholey, A., Wagner-Döbler, I., Wandrey, C., Wittmann, C., Yim, S.C.,Zeng, A.P. (Eds.), Tools and Applications of Biochemical Engineering Science.Springer Berlin Heidelberg, pp. 239–259 Berlin, Heidelberg.

Zeng, A.-P., Sabra, W., 2011. Microbial production of diols as platform chemicals: Recentprogresses. Current Opinion in Biotechnology 22 (6), 749–757.

Zeng, A.P., Biebl, H., Schlieker, H., Deckwer, W.D., 1993. Pathway analysis of glycerolfermentation by Klebsiella pneumoniae: Regulation of reducing equivalent balanceand product formation. Enzyme and Microbial Technology 15 (9), 770–779.

Zeng, A.-P., Menzel, K., Deckwer, W.-D., 1996. Kinetic, dynamic, and pathway studies ofglycerol metabolism by Klebsiella pneumoniae in anaerobic continuous culture: II.Analysis of metabolic rates and pathways under oscillation and steady-state condi-tions. Biotechnology and Bioengineering 52 (5), 561–571.

Zhang, Q., Xiu, Z., 2009. Metabolic pathway analysis of glycerol metabolism in Klebsiellapneumoniae incorporating oxygen regulatory system. Biotechnology progress 25 (1),103–115.

Zhang, X., Li, Y., Zhuge, B., Tang, X., Shen, W., Rao, Z., Fang, H., Zhuge, J., 2006a.Construction of a novel recombinant Escherichia coli strain capable of producing 1,3–propanediol and optimization of fermentation parameters by statistical design.World Journal of Microbiology and Biotechnology 22 (9), 945–952.

Zhang, Y., Li, Y., Du, C., Liu, M., Cao, Z.a., 2006b. Inactivation of aldehyde dehy-drogenase: A key factor for engineering 1,3-propanediol production by Klebsiellapneumoniae. Metabolic Engineering 8 (6), 578–586.

Zhang, Q., Teng, H., Sun, Y., Xiu, Z., Zeng, A., 2008. Metabolic flux and robustnessanalysis of glycerol metabolism in Klebsiella pneumoniae. Bioprocess and BiosystemsEngineering 31 (2), 127–135.

Zhang, Y., Huang, Z., Du, C., Li, Y., Cao, Z.a., 2009. Introduction of an NADH re-generation system into Klebsiella oxytoca leads to an enhanced oxidative and re-ductive metabolism of glycerol. Metabolic Engineering 11 (2), 101–106.

Zhang, X., Shanmugam, K., Ingram, L.O., 2010. Fermentation of glycerol to succinate by

metabolically engineered strains of Escherichia coli. Applied and environmentalmicrobiology 76 (8), 2397–2401.

Zhang, G., Yang, G., Wang, X., Guo, Q., Li, Y., Li, J., 2012. Influence of Blocking of 2,3-Butanediol Pathway on Glycerol Metabolism for 1,3-Propanediol Production byKlebsiella oxytoca. Applied Biochemistry and Biotechnology 168 (1), 116–128.

Zhang, Y., Ji, Z., Lin, J., Xu, D., Fu, S., Gong, H., 2017. Deleting pck improves growth andsuppresses by-product formation during 1, 3-propanediol fermentation by Klebsiellapneumoniae. Journal of Applied Microbiology. 123 (3), 678–687.

Zhang, L., Bao, W., Wei, R., Fu, S., Gong, H., 2018. Inactivating NADH: quinone oxi-doreductases affects the growth and metabolism of Klebsiella pneumoniae.Biotechnology and applied biochemistry. https://doi.org/10.1002/bab.1684.

Zhao, Y., Hindorff, L.A., Chuang, A., Monroe-Augustus, M., Lyristis, M., Harrison, M.L.,Rudolph, F.B., Bennett, G.N., 2003. Expression of a cloned cyclopropane fatty acidsynthase gene reduces solvent formation in Clostridium acetobutylicum ATCC 824.Applied and Environmental Microbiology 69 (5), 2831–2841.

Zhao, L., Ma, X., Zheng, Y., Zhang, J., Wei, G., Wei, D., 2009. Over-expression of glyceroldehydrogenase and 1,3-propanediol oxidoreductase in Klebsiella pneumoniae andtheir effects on conversion of glycerol into 1,3-propanediol in resting cell system.Journal of Chemical Technology & Biotechnology 84 (4), 626–632.

Zheng, D., Constantinidou, C., Hobman, J.L., Minchin, S.D., 2004. Identification of theCRP regulon using in vitro and in vivo transcriptional profiling. Nucleic acids re-search 32 (19), 5874–5893.

Zhou, P., Zhang, Y., Wang, P., Xie, J., Ye, Q., 2014. Butanol production from glycerol byrecombinant Escherichia coli. Annals of microbiology 64 (1), 219–227.

Zhou, J.-J., Shen, J.-T., Jiang, L.-L., Sun, Y.-Q., Mu, Y., Xiu, Z.-L., 2017a. Selection andcharacterization of an anaerobic microbial consortium with high adaptation to crudeglycerol for 1, 3-propanediol production. Applied microbiology and biotechnology101 (15), 5985–5996.

Zhou, J., Wang, D., Wang, C., Gu, J., Kim, C.H., Shi, J., Jiang, B., Wang, M., Hao, J.,2017b. The Role of the Pyruvate Acetyl-CoA Switch in the Production of 1, 3-Propanediol by Klebsiella pneumoniae. Applied biochemistry and biotechnology 181(3), 1199–1210.

Zhou, J.-J., Shen, J.-T., Wang, X.-L., Sun, Y.-Q., Xiu, Z.-L., 2018. Stability and oscillatorybehavior of microbial consortium in continuous conversion of crude glycerol to 1, 3-propanediol. Applied microbiology and biotechnology, 1-15.

Zhu, M.M., Skraly, F.A., Cameron, D.C., 2001. Accumulation of methylglyoxal in anae-robically grown Escherichia coli and its detoxification by expression of thePseudomonas putida glyoxalase I gene. Metabolic engineering 3 (3), 218–225.

Zhu, M.M., Lawman, P.D., Cameron, D.C., 2002. Improving 1, 3-Propanediol Productionfrom Glycerol in a Metabolically Engineered Escherichia coliby ReducingAccumulation of sn-Glycerol-3-phosphate. Biotechnology progress 18 (4), 694–699.

Zhu, H., Yi, X., Liu, Y., Hu, H., Wood, T.K., Zhang, X., 2013. Production of acetol fromglycerol using engineered Escherichia coli. Bioresource Technology 149, 238–243.

Zhu, C., Jiang, X., Zhang, Y., Lin, J., Fu, S., Gong, H., 2015. Improvement of 1,3-propa-nediol production in Klebsiella pneumoniae by moderate expression of puuC (en-coding an aldehyde dehydrogenase). Biotechnology Letters 37 (9), 1783–1790.

Zhuge, B., Zhang, C., Fang, H., Zhuge, J., Permaul, K., 2010. Expression of 1,3-propa-nediol oxidoreductase and its isoenzyme in Klebsiella pneumoniae for bioconversionof glycerol into 1,3-propanediol. Applied Microbiology and Biotechnology 87 (6),2177–2184.

Zingaro, K.A., Papoutsakis, E.T., 2012. Toward a semisynthetic stress response system toengineer microbial solvent tolerance. Mbio 3 (5), e00308–e00312.

Zwaig, N., Kistler, W., Lin, E., 1970. Glycerol kinase, the pacemaker for the dissimilationof glycerol in Escherichia coli. Journal of bacteriology 102 (3), 753–759.

A.W. Westbrook et al. Biotechnology Advances 37 (2019) 538–568

568