2,3-butanediol production using acetogenic bacteria

187
2,3-Butanediol production using acetogenic bacteria Dissertation Submitted for the fulfillment of the requirements for the doctoral degree Dr. rer. nat. at the Faculty of Natural Sciences, Ulm University Catarina Erz from Stuttgart – Bad Cannstatt 2017

Upload: others

Post on 26-Apr-2022

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 2,3-Butanediol production using acetogenic bacteria

2,3-Butanediol production using

acetogenic bacteria

Dissertation

Submitted for the fulfillment of the requirements for the doctoral degree Dr. rer. nat. at the Faculty of Natural Sciences,

Ulm University

Catarina Erz

from

Stuttgart – Bad Cannstatt

2017

Page 2: 2,3-Butanediol production using acetogenic bacteria
Page 3: 2,3-Butanediol production using acetogenic bacteria

The present study was performed at the Institute of Microbiology and Biotechnology,

Ulm University, under the direction of Prof. Dr. Peter Dürre.

Faculty dean in office: Prof. Dr. Peter Dürre

First reviewer: Prof. Dr. Peter Dürre

Second reviewer: Prof. Dr. Bernhard J. Eikmanns

Date of the doctoral examination: October 11th, 2017

Page 4: 2,3-Butanediol production using acetogenic bacteria
Page 5: 2,3-Butanediol production using acetogenic bacteria

Content

Content

Abbreviations ____________________________________________________ I

1. Introduction ___________________________________________________ 1

2. Material and Methods __________________________________________ 10

2.1 Microorganisms, plasmids, and primers ____________________________10

2.1.1 Bacterial strains _________________________________________________ 10

2.1.2 Vectors and plasmids ____________________________________________ 11

2.1.3 Primers ________________________________________________________ 14

2.2 Chemicals, enzymes, and gases ___________________________________17

2.3 Bacterial cultivation ____________________________________________19

2.3.1 Cultivation media _______________________________________________ 19

2.3.1.1 Modified 1.0 ATCC 1612 medium ________________________________ 19

2.3.1.2 Modified ATCC 1612 medium ___________________________________ 22

2.3.1.3 Lyophilization medium ________________________________________ 22

2.3.1.4 Lysogeny broth medium -modified- ______________________________ 23

2.3.1.5 Rajagopalan medium -modified- _________________________________ 23

2.3.1.6 Super optimal broth medium -modified- __________________________ 24

2.3.1.7 Tanner medium -modified- _____________________________________ 25

2.3.1.8 Yeast tryptone fructose medium -modified-________________________ 25

2.3.2 Antibiotic selection ______________________________________________ 26

2.4 Growth conditions _____________________________________________27

2.4.1 Aerobic cultivation ______________________________________________ 27

2.4.2 Anaerobic cultivation ____________________________________________ 27

2.4.3 Strain conservation ______________________________________________ 28

Page 6: 2,3-Butanediol production using acetogenic bacteria

Content

2.4.3.1 Glycerol ____________________________________________________ 28

2.4.3.2 SMP buffer/DMSO ___________________________________________ 28

2.4.3.3 Lyophilization _______________________________________________ 29

2.4.4 Determination of growth and metabolic production parameters _________ 29

2.4.4.1 Determination of optical density ________________________________ 29

2.4.4.2 Quantification of substrate and microbial metabolic products by high

performance liquid chromatography ____________________________ 30

2.4.4.3 Quantification of microbial metabolic products by gas chromatography _ 31

2.5 Nucleic acid methods___________________________________________ 32

2.5.1 Nucleic acid isolation ____________________________________________ 32

2.5.1.1 Plasmid isolation from E. coli using the ZyppyTM Plasmid Miniprep Kit ___ 32

2.5.1.2 Plasmid isolation from E. coli using QIAGEN Plasmid Midi Kit __________ 32

2.5.1.3 Genomic DNA isolation from Gram positive bacteria using MasterPureTM

Gram Positive DNA Purification Kit ______________________________ 33

2.5.1.4 RNA isolation from clostridia ___________________________________ 33

2.5.2 Polymerase chain reaction ________________________________________ 35

2.5.2.1 Composition of PCR reaction mixtures with different DNA polymerases _ 35

2.5.2.2 Standard PCR program ________________________________________ 36

2.5.2.3 Insertion of restriction recognition sequences via PCR _______________ 37

2.5.3 DNA modification _______________________________________________ 37

2.5.3.1 Restriction enzyme digestion of DNA _____________________________ 37

2.5.3.2 Dephosphorylation of digested plasmids __________________________ 38

2.5.3.3 DNA ligation ________________________________________________ 38

2.5.3.4 In-Fusion® HD Cloning Kit ______________________________________ 39

2.5.3.5 Radioactive labelling of DNA fragments ___________________________ 39

Page 7: 2,3-Butanediol production using acetogenic bacteria

Content

2.5.4 Purification of nucleic acids _______________________________________ 40

2.5.4.1 Purification of amplified and digested DNA ________________________ 40

2.5.4.2 Purification of radioactively labelled DNA _________________________ 41

2.5.4.3 Purification of ligated DNA by microdialysis ________________________ 41

2.5.5 Gel electrophoresis ______________________________________________ 41

2.5.5.1 Non-denaturating gel electrophoresis ____________________________ 41

2.5.5.2 Denaturating gel electrophoresis ________________________________ 42

2.5.6 Determination of nucleic acid concentration __________________________ 43

2.6 DNA sequencing and data analysis ________________________________43

2.6.1 Sequencing of vectors and DNA fragments ___________________________ 43

2.6.2 Genome sequencing _____________________________________________ 44

2.6.3 Mapping for single nucleotide polymorphism (SNP) analysis _____________ 44

2.6.4 Comparison of locus tags of different acetogens _______________________ 44

2.7 DNA transfer in bacteria _________________________________________45

2.7.1 DNA transfer in Escherichia coli ____________________________________ 45

2.7.1.1 Preparation of electrocompetent E. coli cells _______________________ 45

2.7.1.2 Preparation of fast competent E. coli cells _________________________ 45

2.7.1.3 Transformation of electrocompetent E. coli cells ____________________ 45

2.7.1.4 Preparation of chemically competent E. coli cells ___________________ 46

2.7.1.5 Transformation of chemically competent E. coli cells ________________ 47

2.7.2 DNA transfer in A. woodii and Clostridium by electroporation____________ 47

2.7.2.1 Preparation of electrocompetent A. woodii and Clostridium cells _______ 47

2.7.2.2 Transformation of electrocompetent A. woodii and Clostridium ________ 47

2.7.3 DNA transfer in Clostridium by conjugation ___________________________ 48

2.7.4 Northern blot ___________________________________________________ 49

Page 8: 2,3-Butanediol production using acetogenic bacteria

Content

2.7.4.1 Transfer of RNA onto a nylon membrane __________________________ 49

2.7.4.2 Hybridization of radioactively labelled probe with RNA ______________ 50

3. Results ______________________________________________________ 53

3.1 Improvement of 2,3-BD production in acetogenic bacteria ____________ 53

3.1.1 Natural 2,3-BD production in acetogenic bacteria _____________________ 53

3.1.2 Construction of plasmids for enhancement of natural 2,3-BD production __ 54

3.1.3 Homologous 2,3-BD overproduction in C. ljungdahlii ___________________ 56

3.1.4 Homologous 2,3-BD overproduction in C. coskatii _____________________ 60

3.1.5 Modifications of 2,3-BD production plasmid pJIR750_23BD _____________ 63

3.1.5.1 Removal of a potential terminator structure _______________________ 64

3.1.5.2 Exchange of bdh with adh (primary-secondary alcohol dehydrogenase) _ 66

3.1.5.3 BudABC operon from Raoultella terrigena _________________________ 68

3.1.5.4 Growth experiments of different C. ljungdahlii strains harboring the modified

2,3-BD plasmids ______________________________________________ 70

3.1.6 Sequencing of C. ljungdahlii [pMTL82151] and C. ljungdahlii WT with SNP

analysis _______________________________________________________ 75

3.1.7 Detection and quantification of transcripts responsible for 2,3-BD production

__________________________________________________________________ 77

3.1.8 Heterologous 2,3-BD production in A. woodii _________________________ 79

3.1.9 Enhancement of 2,3-BD production by overexpressing nifJ ______________ 82

3.2 Heterologous butanol production ________________________________ 90

4. Discussion ____________________________________________________ 95

4.1 Natural microbial 2,3-BD production ______________________________ 95

4.2 Enhancement of 2,3-BD production using acetogens ________________ 103

Page 9: 2,3-Butanediol production using acetogenic bacteria

Content

4.2.1 Enhancement of 2,3-BD production via overexpression of alsS, budA, and bdh

__________________________________________________________________ 103

4.2.2 Modifications of 2,3-BD production plasmid to optimize 2,3-BD production 107

4.2.3 PFOR enzyme as potential bottle-neck in 2,3-BD production? ___________ 112

4.3 Detection of transcripts from synthetic 2,3-BD operons in C. ljungdahlii 116

4.4 Investigation of mutated C. ljungdahlii [pMTL82151] showing increased flux

towards ethanol and 2,3-BD ___________________________________ 118

4.5 A. woodii as suitable host for 2,3-BD production from H2 + CO2? ______ 122

4.6 Heterologous butanol production in C. ljungdahlii - the conflict of large

plasmid transformation _______________________________________ 124

5. Summary ___________________________________________________ 127

6. Zusammenfassung ____________________________________________ 129

7. References __________________________________________________ 132

8. Supplement _________________________________________________ 159

Contribution to this work ________________________________________ 165

Danksagung ___________________________________________________ 167

Curriculum vitae ________________________________________________ 171

Statement_____________________________________________________ 173

Page 10: 2,3-Butanediol production using acetogenic bacteria
Page 11: 2,3-Butanediol production using acetogenic bacteria

Abbreviations I

Abbreviations

A adenosine

A. Acetobacterium; Aeromonas; Aspergillus

AA amino acid

abbr. abbreviation

ad. addiere (fill up to)

ADP adenosine diphosphate

AG Aktiengesellschaft (joint-stock company)

AG & Co. Aktiengesellschaft (joint-stock company) & Compagnie (trading company)

ATCC American Type Culture Collection

ATP adenosine triphosphate

B. Bacillus; Butyribacterium

BLAST Basic Local Alignement Search Tool

Bp base pair(s)

2,3-BD 2,3-butanediol

C cytosine

c centi (10-2)

C. Clostridium, Corynebacterium

°C degree(s) Celsius

C1 compound with one carbon

C4 compound with four carbons

CA California

cal calory

CGMCC China General Microbiological Culture Collection Center

CICC China Center of Industrial Culture Collection

CO carbon monoxide

CO2 carbon dioxide

D D-isomer

Δ delta; gene deletion

DAD diode array detector

[α-32P]-dATP deoxyadenosine triphosphate, labeled on the alpha phosphate group with 32P

Page 12: 2,3-Butanediol production using acetogenic bacteria

II Abbreviations

dATP deoxyadenosine triphosphate

DMSO dimethylsulfoxide

DSMZ Deutsche Sammlung von MIkroorganismen und Zellkulturen (German

Collection of Microorganisms and Cell Culture)

DNA deoxyribonucleic acid

dNTP deoxyribonucleotide triphosphate

E. Escherichia; Enterobacter

EDTA ethylenediaminetetraacetic acid

et al. et alii (and others)

E-value expect value

Fd ferredoxin

Fd2- reduced ferredoxin

FID flame ionization detector

fwd forward

G guanine

g gram

GC gas chromatography

GmbH Gesellschaft mit beschränkter Haftung (limited liability company)

h hour(s)

H+ proton(s)

H2 hydrogen

H2O water

HPLC high-performance liquid chromatography

ID inner diameter

Inc. Incorporated

IPTG isopropyl β-D-1-thiogalactopyranoside

K kilo (103)

K. Klebsiella

L L-isomer

l liter

λ lambda; wavelength

LB Luria-Bertani; lysogeny broth

Page 13: 2,3-Butanediol production using acetogenic bacteria

Abbreviations III

Ltd. Limited

M molar

m milli (10-3); meter

µ micro (10-6)

MA Massachusetts

max. maximal

MES 2-(N-morpholino)ethanesulfonic acid

min minute

mod. modified

n nano (10-9)

N2 nitrogen

Na+ sodium ion

NAD+ oxidized nicotinamide adenine dinucleotide

NADH reduced nicotinamide adenine dinucleotide

NADP+ oxidized nicotinamide adenine dinucleotidephosphate

NADPH reduced nicotinamide adenine dinucleotidephosphate

NC North Carolina

NH New Hampshire

NJ New Jersey

O2 oxygen

OD600nm optical density at 600 nm

P promoter

P. Paenibacillus

PCR polymerase chain reaction

pH negative decade logarithm of the proton concentration (potentia hydrogenii)

Pi inorganic phosphate

PIPES 1,4-piperazinediethanesulfonic acid

R. Raoultella

rev reverse

RID refraction index detector

rpm rounds per minute

s second

Page 14: 2,3-Butanediol production using acetogenic bacteria

IV Abbreviations

S. Saccharomyces; Serratia; Synechococcus; Synechocystis

SMP sucrose-magnesium-phosphate

SNP single nucleotide polymorphism

SOB super optimal broth

sp. Species

SSF simultaneous saccharification and fermentation

subsp. subspecies

Syngas synthesis gas

T terminator; thymine

TE tris-EDTA

TM trademark

TSE tris-sucrose-EDTA

U uracile

UV ultraviolet

V volume

w weight

WI Wisconsin

WT wild type

x g times gravity (unit of relative centrifugal force)

Page 15: 2,3-Butanediol production using acetogenic bacteria

1. Introduction 1

1. Introduction

There are four stable isomers of butanediol (BD): 1,2-BD, 1,3-BD, 2,3-BD, and 1,4-BD. Two of

these four different isomers are important for commercial use. On the one hand, 1,4-BD is a

significant bulk chemical, with production of 1.3 million tons per year from fossil resources

(Zeng and Sabra, 2011). On the other hand, 2,3-BD is considered as an important fine and

potential platform chemical, with impact on the specialty chemical market. It can be easily

converted to butanes, butenes, and butadienes, which are building blocks used for production

of specialty chemicals. The key downstream products of 2,3-BD have a potential global market

of approximately 32 million tons per year with a value of around 43 billion dollars on the sales

market (Köpke et al., 2011b). Transparancy market research company reports that the global

market size of 2,3-BD was estimated at over 61.8 kilo tons in 2012 and is predicted to expend

to 74 kilo tons by 2018. There are three stereoisomers of 2,3-BD: the optically active forms

2S,3S-BD (dextrorotatory/L(+)-form) and 2R,3R-BD (levorotatory/D(-)-form), as well as the

optically inactive form 2R,3S-BD (meso-form) (Figure 1). Since the levorotatory form has a low

freezing point of -60 °C, it is considered to be used commercially as an anti-freeze agent (Kuhz

et al., 2017). Furthermore, 2R,3R-BD as well as 2S,3S-BD are excellent chiral compounds for

asymmetric synthesis of liquid crystals (Xiao et al., 2010; Liu et al., 2011; Zeng and Sabra,

2011). Apart from applications of the optical forms, 2,3-BD can be converted to convenient

derivatives via certain chemical reactions (Ji et al., 2011a; Nie et al., 2011; Gubbels et al.,

2013). The dehydration of 2,3-BD leads to the excellent organic solvent methyl ethyl ketone

(MEK), which is used for resins, fuels, paints, and laquers (Tran and Chambers, 1987; Ji et al.,

2012; Zhang et al., 2012b). Further dehydrating leads to 1,3-butadiene, finding applications in

manufacturing synthetic rubbers, polyesters, and polyurethanes (Haveren et al., 2008).

Figure 1: Stereoisomers of 2,3-butanediol

Page 16: 2,3-Butanediol production using acetogenic bacteria

2 1. Introduction

Moreover, dehydrogenation of 2,3-BD can form acetoin and diacetyl, which are used for

flavoring dairy products and margarines, giving a buttery taste (Bartowsky and Henschke,

2004; Faveri et al., 2003; Ji et al., 2013). Additionally, diacetyl can be applied as a bacteriostatic

food additive, since it inhibits growth of some microorganisms (Celińska and Grajek, 2009). As

a further chemical reaction, esterification of 2,3-BD can form a polyimide precursor, which

finds application in lotions, drugs, and cosmetics. Other products formed by esterification of

2,3-BD with maleic acid are polyurethane-melamides (PUMAs), which are useful in

cardiovascular applications (Petrini et al., 1999). Not only dehydration and esterification are

possible, but also polymerization of 2,3-BD. The polymerization leads to polyesters with high

potential for industry, if they are bio-based and even bio-degradable (Aeschelmann and Carus,

2015; Hu et al., 2016; Debuissy et al., 2016). Furthermore, due to its high octane rating 2,3-BD

might be a potential aviation fuel (Celińska and Grajek, 2009; Ji et al., 2012) and it is useful as

raw material in the manufacture of printing inks, pesticides, plasticizers, moisturizing and

softening agents, explosives, fumigants, etc. (Magee and Kosaric, 1987; Garg and Jain, 1995;

Syu, 2001; Kuhz et al., 2017).

The chemical production of 2,3-BD is performed by removal of butadiene and isobutene from

crack gases (product from steam cracking petroleum refining), whereby a C4-hydrocarbon

fraction is obtained, consisting of approximately 77 % butenes and 23 % of a mixture

containing butane and isobutane (Gräfje et al., 2000). The chlorohydrination of this fraction

with a chlorine/water solution and subsequent cyclization of the chlorohydrins with sodium

hydroxide leads to a butene oxide mixture. This mixture contains 55 % trans-2,3-butene oxide,

30 % cis-2,3-butene oxide, and 15 % 1,2-butene oxide (Gräfje et al., 2000). By subsequent

hydrolysis of the butene oxides, a mixture of butanediols is obtained and separation is

performed by vacuum fractionation. By using an excess of water during hydrolysis, formation

of polyethers is avoided. Thereby, the fractionation of butanediols is easier than of the butene

oxide mixture (Gräfje et al., 2000). The optically inactive meso-2,3-BD is obtained from trans-

2-butene via trans-2,3-butene oxide, whereas the racemic mixture of D(-)-2,3-BD and

L(+)-2,3-BD is formed in an analogous manner from cis-2-butene via cis-2,3-butene oxide

(Gräfje et al., 2000). Furthermore, MEK is also formed as a by-product (Myszkowski and

Zielinski, 1965).

Page 17: 2,3-Butanediol production using acetogenic bacteria

1. Introduction 3

The chemical chiral synthesis of 2,3-BD as well as its separation represent expensive steps.

Thus, application of bacteria for biotechnological production of 2,3-BD with a high

enantiomeric purity reveals an alternative, sustainable, and competitive approach. Moreover,

bio-based 2,3-BD production would be independent from oil supply. This economic aspect has

boosted the overall interest in the biotechnological process recently. The microbial 2,3-BD

production has a history of more than 100 years. In 1906, Harden and Walpole reported for

the first time microbial 2,3-BD production in Klebsiella pneumoniae, formerly known as

Aerobacter aerogenes and Klebsiella aerogenes (Harden and Walpole, 1906). At that time,

their method was very cost-effective and less expensive than chemical synthesis. In 1926, 2,3-

BD accumulation was also detected in Paenibacillus polymyxa (formerly Bacillus polymyxa;

reclassified by Ash et al., 1993) by Garg and Jain (Garg and Jain, 1995). The first microbial

industrial-scale production of 2,3-BD was proposed in 1933 (Fulmer et al., 1933). During World

War II there was a lack of 1,3-butadiene (pre-curser for synthetic rubber), which highly

promoted interest in 2,3-BD fermentation. Thus, the peak of pilot-scale development for

manufacturing 2,3-BD and its conversion to 1,3-butadiene was reached by this time (Ji et al.,

2011a). However, it suddenly ended when less expensive ways for 1,3-butadiene production

by chemical synthesis with petroleum as feed-stock were available (Ji et al., 2011a; Białkowska

et al., 2016). The advantage of the chemical method over biotechnological production lasted

until the mid-1970s, since crude oil prices highly increased due to gradual depletion of that

feedstock. This again promoted interest in 2,3-BD production from biomass (Ji et al., 2011a;

Białkowska et al., 2016). Nowadays, fossil fuel prices remain very unstable, chemical

compounds catalyzing the unique diol structure during chemical synthesis become more and

more expensive, and the petrochemical synthesis requires high energy input. This shows that

the focus needs to shift towards a sustainable microbial 2,3-BD production. Until today, the

best 2,3-BD production strains are sugar- or citrate-fermenting bacteria with the majority

belonging to risk group 2 organisms, e. g. Klebsiella sp. and Serratia marcescens with 150 g/l

and 152 g/l, respectively (Ma et al., 2009; Zhang et al., 2010a). However, application of these

organisms in industrial processes is unfavorable due to safety requirements of

risk group 2 organisms (Celińska and Grajek, 2009; Ji et al., 2011a). One sustainable alternative

is represented by acetogenic bacteria (acetogens), which are obligate anaerobe organisms

capable of using CO or/and CO2 + H2 as sole carbon and energy source to produce the central

intermediate acetyl-CoA via the reductive acetyl-CoA pathway, which is also known as Wood-

Page 18: 2,3-Butanediol production using acetogenic bacteria

4 1. Introduction

Ljungdahl pathway (Wood, 1991; Ragsdale and Pierce, 2008; Figure 2). This pathway is

considered to be one of the oldest biochemical pathways in life and was first described in

Moorella thermoacetica (formerly known as Clostridium thermoaceticum) (Fontaine et al.,

1942; Daniel et al., 1990) by Harland Goff Wood and Lars Gerhard Ljungdahl (Ljungdahl, 1986;

Wood, 1991; Drake et al., 2008). The Wood-Ljungdahl pathway is one of six pathways capable

Figure 2: Overview of the Wood-Ljungdahl pathway of acetogens with enzymes and natural products starting from the central intermediate acetyl-CoA with focus on 2,3-BD production. Rnf, Rhodobacter

nitrogen fixation; ATPase, adenosine triphosphatase; CODH, carbon monoxide dehydrogenase; Acs/CODH, complex of acetyl-CoA synthase with carbon monoxide dehydrogenase; THF, tetrahydrofolate; [CO], enzyme-bound CO; Fdh, formate dehydrogenase; Fhs, formyl-THF synthetase; Mtc, methenyl-THF cyclohydrolase; Mtd, methylene-THF deyhdrogenase; Mtr, methylene-THF reductase; Met, methyl transferase; AlsS, acetolactate synthase; BudA, acetolactate decarboxylase; Bdh, 2,3-BD dehydrogenase; sAdh, primary-secondary alcohol dehydrogenase; Co-FeS-P, corrinoid iron-sulfur protein; HS-CoA, coenzyme A; [H], reducing equivalent.

Page 19: 2,3-Butanediol production using acetogenic bacteria

1. Introduction 5

of fixing CO2, but instead of being circular it follows a linear manner (Fuchs, 2011). It consists

of the methyl and carbonyl branch (Figure 2), leading to the central intermediate acetyl-CoA.

In the methyl branch, CO2 is reduced stepwise to methyl-tetrahydrofolate (THF) via the

enzymes formate dehydrogenase (Fdh), formyl-THF synthetase (Fhs), methenyl-THF

cyclohydrolase (Mtc), methylene-THF dehydrogenase (Mtd), and methylene-THF reductase

(Mtr) (Ljungdahl, 1986; Ragsdale and Pierce, 2008). Activation of formate to formate-THF by

the enzyme formyl-THF synthetase requires one mol of ATP. After binding the methyl-group

of methyl-THF to a corrinoid iron-sulfur protein (CoFeS-P) in the last step of the methyl branch,

the enzyme complex Acs/CODH (acetyl-CoA synthase/carbon monoxide dehydrogenase)

merges the enzyme-bound methyl group (methyl branch), CO (carbonyl branch), and

coenzyme A (CoA) to the central intermediate acetyl-CoA. Afterwards, it is either used for

anabolism (production of biomass) or converted to other products (Ragsdale, 2007). The Rnf

complex (Rhodobacter nitrogen fixation; ferredoxin:NAD+ oxidoreductase) plays an important

role for energy conservation in acetogens during autotrophic growth. It couples the transfer

of electrons from reduced ferredoxin to NAD+ with simultaneous translocation of cations

across the cell membrane, leading to a cation gradient (Imkamp et al., 2007; Müller et al.,

2008; Biegel and Müller, 2010; Tremblay et al., 2012; Hess et al., 2016). Afterwards, this

gradient is used by an ATPase for generation of additional ATP (Reidlinger and Müller, 1994).

Thereby, this system is also coupled to a flavin-based electron-bifurcating hydrogenase

providing cations from oxidation of hydrogen with concomitant reduction of oxidized

ferredoxin (Schuchmann and Müller 2012, Wang et al., 2013a, Buckel and Thauer, 2013).

Other acetogens such as Moorella thermoacetica harbour other systems involved in energy

conservation, including a so-called energy-converting hydrogenase (Ech) complex,

cytochromes and quinones (Schuchmann and Müller, 2014). Over 100 acetogenic species

were identified to date, of which 90 % produce acetate as sole end product (Köpke et al.,

2011a). In addition, ethanol, butanol, butyrate, lactate, hexanol, and hexanoate can be by-

products depending on the strain (Bruant et al., 2010; Schiel-Bengelsdorf and Dürre, 2012;

Phillips et al., 2015; Ramió-Pujol et al., 2015). Recently, 2,3-BD production was shown for

C. ljungdahlii (Figure 3A), Clostridium autoethanogenum (Figure 3B), and Clostridium ragsdalei

(1.4-2 mM) from steel mill waste gas (Köpke et al., 2011b). These organisms mainly produce

Page 20: 2,3-Butanediol production using acetogenic bacteria

6 1. Introduction

the D(-)-form with 94 %, while only 6 % of meso-2,3-BD is formed. The production of 2,3-BD

in C. autoethanogenum from pyruvate is catalyzed by the enzymes acetolactate synthase

(AlsS), acetolactate decarboxylase (BudA), and 2,3-BD dehydrogenase (Bdh) or primary-

secondary alcohol dehydrogenase (Adh) (Köpke et al., 2011b; Köpke et al., 2014). Acetogens

being typically used for commercial syngas fermentation are C. ljungdahlii (Figure 3A) and

C. autoethanogenum (Figure 3B), C. ragsdalei, Clostridium coskatii, Clostridium

carboxidivorans, Clostridium aceticum, M. thermoacetica (formerly known as Clostridium

thermoaceticum), Acetobacterium woodii, and Butyribacterium methylotrophicum. They can

use pure CO, or syngas, or other waste gases originating from e. g. steel mills or chemical

production lines (Daniell et al., 2016; Dürre, 2016a). Gas fermentation is a process showing

several advantages such as the use of waste gases diminishing environmental pollution by

reduction of industrial gas emissions. Reducing CO2-emissions is a major topic of today’s

society, due to increasing global warming. Furthermore, gas fermentation represents a non-

cellulosic process not only saving food resources, but also agricultural land. Aerobic as well as

anaerobic gas fermentation have both reached commercial level. Especially acetogens have

attracted attention in gas fermentation, including industrial waste gases and synthesis gas

(syngas; a mixture of mainly CO and H2), since these organisms are less sensitive to variations

and contaminants in the composition of the gas as well as leading to a high product specifity

(Köpke et al., 2011a; Schiel-Bengelsdorf and Dürre, 2012; Dürre and Eikmanns, 2015).

In addition to microbial 2,3-BD production, biotechnologically produced 1-butanol (butanol)

is a valuable product for industry. Due to its high heating value, low corrosiveness compared

to ethanol, low volatility, and energy density similar to gasoline, it is a more interesting biofuel

Figure 3: Scanning electron microscope image of C. ljungdahlii (A) and C. autoethanogenum (B) growing with fructose as energy and carbon source. The white bar represents the scale.

2 µm 2 µm

A B

Page 21: 2,3-Butanediol production using acetogenic bacteria

1. Introduction 7

than ethanol. The four-carbon alcohol is commonly used as a chemical and solvent in paints,

coatings, sealants, textiles, printing inks, glues, and plastics (Hahn et al., 2000). In 2013, a study

presented by the company Informa Economics revealed that the annual global market of

butanol exceeded 3.6 million tons/year and was valued over 6 billion dollars at that time.

Butanol was produced petrochemically in the past decades. In this process called oxo-

synthesis, propylene is converted to butyraldehyde by using a homogenous catalyst via

hydroformylation with CO. Subsequently, butyraldehyde is hydrogenated over a

heterogenous catalyst to butanol. The Guerbet reaction represents a second pathway for

chemical synthesis of butanol (O’Lenick, 2001; Kozlowski and Davis, 2013). This process

consists of three steps: dehydrogenation of ethanol to acetaldehyde, acetaldehyde aldol-

condensation to form crotonaldehyde, and hydrogenation of crotonaldehyde to form 1-

butanol (Kozlowski and Davis, 2013; Sun and Wang, 2014). In order to achieve high activity

and selectivity of the Guerbet reaction, the catalyst used in this process needs to have certain

characteristics. On the one hand, it can be a basic agent like an alkali metal hydroxide, a salt

dissolved in the reaction medium (homogenous catalyst), or a solid base (heterogenous

catalyst) (Kuhz et al., 2017). On the other hand, it needs to facilitate the dehydrogenation of

ethanol at the respective reaction temperature. Examples for typical dehydrogenating agents

are metals including platinum, nickel, and copper (Veibel and Nielsen, 1967; Kozlowski and

Davis, 2013). If a homogenous catalyst is used in the formation of butanol, a precious metal

catalyst such as an organometallic complex needs to be applied in order to perform the

de/hydrogenation steps and an inorganic base aids in the aldol coupling step (Koda et al.,

2009; Chakraborty et al., 2015). A different alternative to this three-step mechanism

represents a coupling step without obtaining two molecules of acetaldehyde (Faba et al.,

2016). This process involves a direct surface coupling between the α-carbon of an aldehyde

and one alcohol (Yang and Meng, 1993; Ndou et al., 2003).

Instead of chemical synthesis, 66 % of butanol used worldwide was produced by acetone-

butanol-ethanol (ABE) fermentation until 1950 (Dürre, 2008), which is one of the oldest

bioprocesses in history. The first discovery of biological butanol production was found in

Vibrion butyrique (Pasteur, 1862), but it propably was not a pure culture. Most likely it

contained Clostridium butyricum, which is also able to produce butanol under certain

conditions (Dürre, 2005). The most commonly used bacteria for biotechnological butanol

Page 22: 2,3-Butanediol production using acetogenic bacteria

8 1. Introduction

production are clostridia such as Clostridium acetobutylicum, Clostridium beijerinckii,

Clostridium pasteurianum, Clostridium sporogenes, Clostridium saccharobutylicum, and

Clostridium saccharoperbutylacetonicum (Visioli et al., 2014). The metabolism of these strains

shows an acidogenic phase in the exponential growth phase converting the substrate to acids

(acetate and butyrate) followed by the solventogenic phase, in which the substrate and

produced acids are transformed into solvents (acetone, butanol, and ethanol) (Dürre, 2005).

It is characterized by a typical ABE ratio of 3:6:1 (Jones and Woods, 1986; Formanek et al.,

1997). Due to rising fossil oil production after the World War II, butanol production by ABE

fermentation was too expensive compared to oxo-synthesis and the last ABE-plant closed in

the late 1980s (Li et al., 2014a). Recently, interest in biotechnological butanol production

dramatically increased again, since acetogens were reported to produce butanol from the

renewable and sustainable feedstock syngas. The organism Butyribacterium

methylotrophicum was the first acetogen reported to naturally produce butanol

autotrophically (Grethlein et al., 1991). For the formation of butanol two mol of acetyl-CoA

are condensed to acetoacetyl-CoA, which is stepwise reduced to butanol (Fast and

Papoutsakis, 2012). Recently, C. ljungdahlii was genetically modified to produce butanol

growing on syngas by introduction of the plasmid pSOBptb (Köpke et al., 2010). By

overexpression of the six genes thlA (thiolase), hbd (3-hydroxybutyryl-CoA dehydrogenase),

crt (crotonase), bcd (butyryl-CoA dehydrogenase), adhE (bifunctional butyraldehyde/alcohol

dehydrogenase), and bdhA (butanol dehydrogenase) production of 2 mM butanol with

C. ljungdahlii was detected (Köpke et al., 2010). This work demonstrated that production of

butanol by syngas fermentation is possible.

The main focus of this study was to enhance 2,3-BD production of different acetogens by

overexpressing the genes alsS (acetolactate synthase), budA (acetolactate decarboxylase),

and bdh (2,3-BD dehydrogenase) responsible for 2,3-BD formation starting from pyruvate in

C. ljungdahlii. For this purpose, the respective genes were cloned into two different shuttle

vectors containing different replicons from Gram-positive bacteria (Clostridium perfringens,

Clostridium botulinum) enabling investigation of their influence on 2,3-BD formation.

Furthermore, examination of additional overexpression of nifJ (pyruvate:ferredoxin

oxidoreductase) was performed to elucidate, if this modification can direct the metabolic flux

towards 2,3-BD production.

Page 23: 2,3-Butanediol production using acetogenic bacteria

1. Introduction 9

The second aim of this study was the modification of the above-mentioned butanol plasmid

pSOBptb (Köpke et al., 2010). The main disadvantage of this plasmid is that it does not contain

the genes etfA and etfB (electron-transferring flavoproteins), which were shown to be

essential for reduction of crotonyl-CoA to butyryl-CoA e. g. in C. kluyveri (Li et al., 2008). These

proteins are also considered to be needed for butanol formation in C. acetobutylicum (Lütke-

Eversloh and Bahl, 2011). In this study, the plasmid pSB3C5-UUMKS 3 containing the genes

etfA and etfB from C. acetobutylicum (Schuster, 2011) should be equipped with two different

replicons from Gram-positive bacteria (C. perfringens, Clostridium butyricum). This aims at

construction of a recombinant acetogenic strain capable of efficient butanol formation.

Page 24: 2,3-Butanediol production using acetogenic bacteria

10 2. Material and Methods

2. Material and Methods

2.1 Microorganisms, plasmids, and primers

2.1.1 Bacterial strains

Bacterial strains used in this work are shown in Table 1.

Table 1: Bacterial strains

Strain Relevant geno- or phenotype* Origin/Reference

Acetobacterium woodii

DSM 1030

Type strain DSMZ** GmbH,

Brunswick, Germany

Clostridium aceticum

DSM 1496

Type strain DSMZ GmbH,

Brunswick, Germany

Clostridium autoethanogenum

DSM 10061

Type strain DSMZ GmbH,

Brunswick, Germany

Clostridium carboxidivorans

DSM 15243

Type strain DSMZ GmbH,

Brunswick, Germany

Clostridium coskatii

PTA-10522

Type strain ATCC***, Manassas,

VA, USA

Clostridium ljungdahlii

DSM 13528

Type strain DSMZ GmbH,

Brunswick, Germany

Clostridium ragsdalei

DSM 15248

Type strain DSMZ GmbH,

Brunswick, Germany

Escherichia coli

CA434

hsdS20 (r-B, m-B), supE44, thi-1,

recAB, ara-14, leuB5proA2, lacY1,

galK, rpsL20 (StrR), xyl-5, mtl-1

including the conjugative plasmid

R702 (TetR, SmR, SuR, HgR, Tra+,

Mob+)

Purdy et al., 2002

*Standard genotype abbreviations for E. coli (Berlyn, 1998)

**Deutsche Sammlung von Mikroorganismen und Zellkulturen (German Collection of Microorganisms

and Cell Culture)

***American Type Culture Collection

Page 25: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 11

Table 1: Bacterial strains (continued)

Strain Relevant geno- or phenotype* Origin/Reference

ER2275

trp31 his1 tonA2 rpsL104

supE44 xyl-7 mtl-2 metB1

e14- Δ(lac)U169 endA1 recA1

R(zgbZ10::Tn10) Tcs

Δ(mcr-hsd-mrr) 114::1510

(F´ proAB traD36 lacIq Δ M15

zzf::mini Tn10 (KmR))

Mermelstein and

Papoutsakis, 1993

XL1-Blue MRF'

Δ(mcrA)183 Δ(mcrCB-

hsdSMR-mrr)173 endA1

supE44 thi-1 recA1 gyrA96

relA1 lac (F’proAB lacIq

ZΔM15 Tn10 (TetR))

Agilent

Technologies, Santa

Clara, CA, USA

*Standard genotype abbreviations for E. coli (Berlyn, 1998)

2.1.2 Vectors and plasmids

Vectors and plasmids constructed and used in this work are listed in Table 2.

Table 2: Vectors and plasmids

Vector/Plasmid Size [bp] Characteristics Origin/Reference

pACYC184 4,245 p15A ori (rep), Cmr (catP), Tcr

(tetR)

DSMZ GmbH,

Brunswick, Germany

pACYC184_MCljI 7,414 pACYC184 carrying

CLJU_c03310 and

CLJU_c03320 from

C. ljungdahlii under control

of Pptb-buk from

C. acetobutylicum

Matthias Beck,

unpublished

pACYC184_MCljI_pSC101ori 8,327 pACYC184_MCljI carrying

pSC101 ori (rep101)

This study

Page 26: 2,3-Butanediol production using acetogenic bacteria

12 2. Material and Methods

Table 2: Vectors and plasmids (continued)

Vector/Plasmid Size [bp] Characteristics Origin/Reference

pCE2 14,692 pSB3C5-UUMKS 2 carrying

pIP404 ori (rep)

This study

pCE3 14,686 pSB3C5-UUMKS 3 carrying

pCB102 ori (repH)

This study

pCE3_traJ 15,452 pCE3 carrying traJ of

pMTL82151

This study

pJIR750 6,568 pMB1 ori (rep), pIP404 ori

(rep), Cmr (catP), lacZ

Bannam and Rood,

1993

pJIR750_23BD

10,960 pJIR750 carrying alsS, budA,

and bdh from C. ljungdahlii

under control of

Ppta-ack from C. ljungdahlii

This study

pJIR750_23BD_budAshort

10,867 pJIR750_23BD carrying a 3'-

end shortened sequence of

budA from C. ljungdahlii

This study

pJIR750_23BD_budAshort_adh 10,798 pJIR750_23BD_budAshort

carrying CLJU_c24860 from

C. ljungdahlii instead of bdh

This study

pJIR750_budABC 9,771 pJIR750 carrying budA, budB,

and budC from Raoultella

terrigena

This study

pJIR750_budABC_Ppta-ack 10,153 pJIR750_budABC carrying

Ppta-ack from C. ljungdahlii

This study

pJIR750_budABCoperon 10,342 pJIR750_budABC_Ppta-ack

carrying Tsol-adc from

C. acetobutylicum

This study

pJIR750_23BD_PFOR 14,510 pJIR750_23BD carrying nifJ

from C. ljungdahlii

This study

pJIR750_23BD_PFOR_traJ 15,274 pJIR750_23BD_PFOR carrying

traJ of pMTL82151

This study

Page 27: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 13

Table 2: Vectors and plasmids (continued)

Vector/Plasmid Size [bp] Characteristics Origin/Reference

pJIR750_oBDH_PFOR 13,212 pJIR750_23BD without bdh

carrying nifJ from

C. ljungdahlii

This study

pJIR750_oBDH_PFOR_traJ 13,976 pJIR750_oBDH_PFOR carrying

traJ of pMTL82151

This study

pJR3 14,199 pSB3C5-UUMKS 3 carrying

pCB102 ori (repH)

This study

pJR3_traJ 14,965 pJR3 carrying traJ of

pMTL82151

This study

pJIR750_oBDH_PFOR_traJ 13,976 pJIR750_oBDH_PFOR carrying

traJ of pMTL82151

This study

pMTL82151 5,254 colE1 ori (rep), pBP1 ori (repA,

orf2), Cmr (catP), traJ

Heap et al., 2009

pMTL82151_23BD 9,646 pMTL82151 carrying alsS,

budA, and bdh from

C. ljungdahlii under control of

Ppta-ack from C. ljungdahlii

This study

pMTL82151_23BD_PFOR 13,196 pMTL82151_23BD carrying

nifJ from C. ljungdahlii

This study

pMTL82151_oBDH_PFOR 11,898 pMTL82151_23BD without

bdh carrying nifJ from

C. ljungdahlii

This study

pMTL83151 4,476 colE1 ori (rep), pCB102 ori

(repH), Cmr (catP), traJ

Heap et al., 2009

pMK 2,393 colE1 ori (rep), Kmr (kanR) Thermo Fisher

Scientific Inc.,

Waltham, MA, USA

pMK_budABCoperon 5,537 colE1 ori (rep), budA, budB,

budC from R. terrigena, Kmr

(kan)

Biopolis S. L.,

Valencia, Spain

Page 28: 2,3-Butanediol production using acetogenic bacteria

14 2. Material and Methods

Table 2: Vectors and plasmids (continued)

Vector/Plasmid Size [bp] Characteristics Origin/Reference

pSB3C5 3,828 p15A ori (rep), Cmr (catP),

ccdB

ATG:biosynthetics

GmbH, Freiburg,

Germany

pSB3C5-UUMKS 3 12,515 pSB3C5 carrying crt, bcd,

etfA, etfB, hbd, thlA, adhE2

from C. acetobutylicum under

control of Pptb-buk from

C. acetobutylicum

ATG:biosynthetics

GmbH, Freiburg,

Germany

pSC101 9,263 pSC101 ori (rep101), Tcr (tetR) DSMZ GmbH,

Brunswick,

Germany/Cohen

and Chang, 1977

pUC57 2,710 ColE1 ori (rep), Apr (bla), lacZ GenScript USA Inc.,

Piscataway, NJ, USA

pUC57_23BD

10,960 pUC carrying carrying alsS,

budA and bdh from

C. ljungdahlii under control of

Ppta-ack

GenScript USA Inc.,

Piscataway, NJ, USA

2.1.3 Primers

Primers used in this study are presented in Table 3. They were synthesized by biomers.net

GmbH (Ulm, Germany) and dissolved in sterile, ultrapure water to get a concentration of

100 pmol/µl. Primers were used for either amplification of DNA fragments or sequencing. For

storage purposes, primers were kept at -20°C. When required, restriction sites added at the

5'-end are underlined in bold and primers that contain overlapping regions to the utilized

cloning vector for “In-Fusion® HD Cloning” are italicized.

Page 29: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 15

Table 3: Primers

Primer Sequence (5' → 3') Restriction enzyme

fD1* AGAGTTTGATCCTGGCTCAG

rP2* ACGGCTACCTTGTTACGACTT

pMTL82151catP_fwd CATACCGGGAATATGTAG

pMTL82151catP_rev GTTTGCAAGCAGCAGATTAC

pMTL82151_23BDSeq1 GCAGAACGAGCACTTTCAATATG

pMTL82151_23BDSeq2 TGCCCGTAGGTGCTTTAG

pMTL82151_23BDSeq3 GTTGGTGATGGCGGTTTC

pMTL82151_23BDSeq4 AGCCTATGGCTGAAGTTG

pMTL82151_23BDSeq5 CAAGAGTACTGCACCAGTAG

pJIR750catP_fwd CTTTCGGCTCGATTTCAC

pJIR750catP_rev CGGCAAGTGTCCAAGAAG

pIP404PstI_fwd ACACTGCAGATCCCGCTTTAATCCCAC PstI

pIP404SalI_rev ACAGTCGACGGTTTACTTCAACGGC SalI

CljbudAKpnI_fwd ACAGGTACCGAGGTGAATGTAATATGG KpnI

Clj-budABamHI_rev ACAGGATCCCTATCAATACTTTATTTTTC BamHI

CljADHBamHI_fwd ACAGGATCCGGAGGTTGTATTATGAAAGG BamHI

CljADHSalI_rev ACAGTCGACGTTCTTGGCATCAGGTTGAG SalI

budABC_SEQ CTGAATGCACCTGCCAGGAG

BudABC_Seq1 CATATGCAGGGCCTTAAC

BudABC_Seq2 GCGCCTGCACGCCAATATTC

BudABC_Seq3 TGGTCTCCAACGCCTTCCG

BudABC_Seq4 GGCCGCTGTATCCATATGAC

pIP404Seq1 CTGTCGACGGTTTACTTC

pIP404Seq2 GGAGCAGATAGACAAGCCTTAG

pIP404Seq3 TCTAATGTTAGAAGTACC

CPECPpta-ackEcoRI_fwd CACACAGGAAACAGCTATGACCATGATTAC

GAATTCCCATGTCAAGAACTCTGTTTATTTC

EcoRI

CPECPpta-ackEcoRI_rev CCTGGCAGGTGCATTCAGGATAATGATTCAC

GAATTCGGTGGTGGCTTTAAATTTAACACAAA

ATTAC

EcoRI

*(Weisburg et al., 1991)

Page 30: 2,3-Butanediol production using acetogenic bacteria

16 2. Material and Methods

Table 3: Primers (continued)

Primer Sequence (5' → 3') Restriction enzyme

TsoladcXbaI_fwd ACATCTAGACTAGCAAAGAAATACTATG XbaI

TsoladcHindIII_rev ACAAAGCTTCCTTAGAATCCATTAC HindIII

pJIRSEQ_fwd GATAACCGTATTACCGCCTTTG

pJIR_budABC_Pptaack

SEQ_rev

CAATCAGTTCGCCATCGAGTTC

pCB102SalI_fwd ACAGTCGACAGCCTGAATGGCGAATGG SalI

pCB102PstI_rev ACACTGCAGCCGGCCGCTTATAATCCATAAC PstI

pJR3Seq1 TCGACAGCCTGAATGGCGAATG

pJR3Seq2 GGGTGAGCAAAGTGACAGAG

pJR3Seq3 AATATTGAGAGTGCCGACACAG

pJR3Seq4 CAAAGCCGTTTCCAAATC

pSC101SacII_fwd ACACCGCGGAGCGCAGCGAACTGAATGTC SacII

pSC101ClaI_rev ACAATCGATGAACAGCTTTAAATGCAC ClaI

CljalsS-sonde_fwd GTTTAGAAGCCGAAGGAG

CljalsS-sonde_rev TTATAGCACCGCTTCCAG

CljBudA-sonde_fwd GAGGTGAAAGTCCCAAAC

CljBudA-sonde_rev CCAGTTCAGTGACTACAG

Clj23bdh-sonde_fwd GTGGTTCTGACTTGCATGAG

Clj23bdh-sonde_rev GAGAATGAAGGGCAACTG

PFORInfusionBamHI_ fwd TAAGTAAGTCCACCTGTCCGGATCCTGAAAAG

GAGAGGAATTTTTATG

BamHI

PFORInfusionBamHI_ rev CTATGTGCCATTATGACACGGATCCAATTTACT

TGATTACTGTTCATTATCAAG

BamHI

PFORohneBDHBamHI_ fwd TAAGTAAGTCCACCTGTCCGGATCCTGAAAAG

GAGAGGAATTTTTATG

BamHI

PFORohneBDHSalI_ rev CAAGCTTGCATGCCTGCAGGTCGACAATTAAA

AAAGTACCAGACAGC

SalI

PFOR-Seq1 TCCAGTAACTCGTGGTACAG PFOR-Seq1

PFOR-Seq2 AACGCTGCTATAGACAAGGG PFOR-Seq2

PFOR-Seq3 GCTGAAGGTGAAGGAACAAGAG PFOR-Seq3

PFOR-Seq4 CACCAGAAGGAACTACAG PFOR-Seq4

Page 31: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 17

Table 3: Primers (continued)

Primer Sequence (5' → 3') Restriction enzyme

PFOR-Seq5 CTACAAGCAGGAGCATTGAC

traJInfusionPstI_fwd AACCCTTAGTGACTCCTGCAGCGATCGGTCTTG

CCTTGC

PstI

traJInfusionPstI_rev AACCCTTAGTGACTCCTGCAGCGATCGGTCTTG

CCTTGC

PstI

traJseq1 GATTAAAGCGGGATCTGC

traJSeq_rev CGCCTCCATCCAGTCTATTC

InfusionTraJEheI_fwd CAGCCTGAATGGCGAATGGCGCCTGCTTGCGG

GTCATTATAG

EheI

InfusionTraJEheI_rev GAGAAAATACCGCATCAGGCGCCGATCGGTCT

TGCCTTGC

EheI

pJR3-traJ Seq CCTGGCGTTACCCAACTTAATC

M13-FP* TGTAAAACGACGGCCAGT

M13-RP* CAGGAAACAGCTATGACC

*GATC Biotech AG, Constance, Germany

2.2 Chemicals, enzymes, and gases

The chemicals and enzymes used in this work were all purchased from the following

companies:

AppliChem GmbH, Darmstadt, Germany

Biozym Scientific GmbH, Oldenburg, Germany

Carl Roth GmbH & Co. KG, Karlsruhe, Germany

Difco Laboratories, Augsburg, Germany

Epicentre Technologies Corp., an Illumina company, Madison, WI, USA

GE Healthcare EUROPE GmbH, Munich, Germany

Genaxxon Bioscience GmbH, Ulm, Germany

Invitrogen GmbH, Karlsruhe, Germany

MACHEREY-NAGEL GmbH & Co. KG, Dueren, Germany

Merck KGaA, Darmstadt, Germany

Qiagen GmbH, Hilden, Germany

SERVA Electrophoresis GmbH, Heidelberg, Germany

Page 32: 2,3-Butanediol production using acetogenic bacteria

18 2. Material and Methods

Sigma-Aldrich Chemie GmbH, Steinheim, Germany

Takara Bio USA Inc., Mountain View, CA, USA

Thermo Fisher Scientific, Waltham, MA, USA

VWR International GmbH, Darmstadt, Germany

ZYMO Research Europe GmbH, Freiburg, Germany

All gases and gas mixtures used during this study are listed in Table 4. Gases, excluding syngas

and H2 + CO2, were purchased from the company MTI IndustrieGase AG (Neu-Ulm, Germany).

Syngas and H2 + CO2 were ordered at Westfalen AG (Münster, Germany).

Table 4: Utilized gases and gas mixtures

Gas Gas composition Purity Application

Forming gas 095.0 % N2

005.0 % H2

3.0

3.0

Anaerobic chamber

N2 100.0 % N2 5.0 Anaerobic media

preparation, gas

chromatography (carrier gas)

Synthetic air 079.5 % N2

020.5 % O2

5.0

5.0

Gas chromatography

(detector gas)

H2 + CO2 067.0 % H2

033.0 % CO2

5.0

3.0

Autotrophic growth

experiments

Synthesis gas (syngas) 040.0 % CO

040.0 % H2

010.0 % CO2

010.0 % N2

2.0

5.0

3.0

5.0

Autotrophic growth

experiments

N2 + CO2 080.0 % N2

020.0 % CO2

5.0

3.0

Anaerobic media preparation

H2 100.0 % H2 5.0 Gas chromatography

(detector gas)

Page 33: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 19

2.3 Bacterial cultivation

2.3.1 Cultivation media

In general, for preparing culture media (see chapter 2.3.1.1 - 2.3.1.8), chemicals were weighed

to corresponding composition into beakers and dissolved in water of the ultrapure water

system “18.2 MΩ-cm, Type I water” (PURELAB Classic, ELGA LabWater, Celle). After adjusting

pH and end volume, media were autoclaved at 121 °C and 1.2 bar for 15 to 20 min. Heat-liable

components such as antibiotics were filter-sterilized (“Filtropur S 2.0”, pore size 0.2 µm,

Sarstedt AG & Co, Nümbrecht, Germany) and were added to the media directly prior to use.

In order to prepare solid media, 1.5 % agar were added to liquid culture media before

autoclaving. Afterwards, media were cooled down to approximately 55 °C and antibiotics were

supplemented. The prepared media were poured into sterile petri dishes and solidified

overnight. Solid agar plates were sealed in sterile plastic bags and stored at 4 °C.

Growth of clostridial strains requires anaerobic culture media. Resazurin was supplemented

to anaerobic media as an indicator for oxygen. Under anaerobic conditions, resazurin is

reduced to the colourless dihydroresorufin, while under aerobic conditions the reoxidized

resorufin is formed appearing as a pink colour. Media were prepared aerobically, filled in

Hungate tubes (Ochs GmbH, Bovenden, Germany) or glass flasks (Müller Krempel AG, Bülach,

Switzerland, or Thermo Fisher Scientific Inc., Waltham, MA, USA), closed with butyl rubber

stoppers (Ochs GmbH, Bovenden or Maag Technic GmbH, Göppingen) and screw caps (Ochs

GmbH, Bovenden, Germany or Thermo Fisher Scientific Inc., Waltham, MA, USA). Finally,

vacuum was applied which was followed by purging with N2 + CO2 to get an anaerobic

atmosphere. These last two steps were repeated 5 times. Media which contained agar

(1.5 % [w/v]) were poured into sterile petri dishes in the anaerobic chamber after cooling

down to 55 °C. Agar plates were enabled to solidify overnight, sealed in sterile plastic bags,

and were stored in the anaerobic chamber at room temperature.

2.3.1.1 Modified 1.0 ATCC 1612 medium (Hoffmeister, unpublished)

Cultivation of C. aceticum was performed in modified 1.0 ATCC 1612 medium.

Modified 1.0 ATCC 16121 medium

NH4Cl 00.20 g 03.7 mM

KH2PO4 00.33 g 02.0 mM

Page 34: 2,3-Butanediol production using acetogenic bacteria

20 2. Material and Methods

K2HPO4 00.45 g 03.0 mM

MgSO4 x 7 H2O* 00.10 g 00.4 mM

Vitamin solution 02.00 ml 00.2 % [v/v]

SL-9 Trace element solution 02.00 ml 00.2 % [v/v]

Bacto® yeast extract 03.00 g 00.3 % [w/v]

NaHCO3 10.00 g 00.1 M000000000

Cysteine HCl x H2O 00.30 g 01.7 mM

Na2S x 9 H2O 00.30 g 01.2 mM

Resazurin 01.00 mg 04.4 µM

H2O ad 1,000 ml0000

*Sterile, anaerobic stock solution (750 mM) was supplemented after autoclaving.

As carbon source, a sterile and anaerobic fructose stock solution (1.11 M) was added after

autoclaving. The pH was adjusted to 8.5 by adding Na2CO3 (sterile, anaerobic stock solution;

5 % [w/v]). For cultivation of C. aceticum on syngas, additional trace elements (sterile,

anaerobic stock solutions) were supplemented to 50 ml medium prior to use (0.2 ml FeCl2

stock solution, 0.1 ml Na2SeO3 + Na2WO4 stock solution, 0.4 ml NiCl2 stock solution, and 0.4 ml

of ZnSO4 stock solution).

FeCl2 Stock solution

FeCl2 x 4 H2O 0.12 g 8.10 mM

H2O ad 50 ml0000

Na2SeO3 + Na2WO4 Stock solution

NaOH 0.5 g 15.5 mM

Na2SeO3 x 5 H2O 3.0 mg 11.4 µM

Na2WO4 x 2 H2O 4.0 mg 12.1 µM

H2O ad 1000 ml0000

NiCl2 Stock solution

NiCl2 x 6 H2O 0.03 g 0.36 mM

H2O ad 50 ml0000

ZnSO4 Stock solution

ZnSO4 x 7 H2O 0.06 g 3.91 mM

H2O ad 50 ml0000

Page 35: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 21

Vitamin solution -modified- (Wolin et al., 1963)

Pyridoxine HCl 50 mg 205 µM

Thiamine HCl 50 mg 148 µM

Riboflavin 50 mg 130 µM

Calcium pantothenate 50 mg 105 µM

α-Lipoic acid 25 mg 120 µM

4-Aminobenzoic acid 50 mg 365 µM

Nicotinic acid 50 mg 405 µM

Vitamin B12 25 mg 020 µM

D-Biotin 25 mg 100 µM

Folic acid 25 mg 055 µM

H2O ad 1,000 ml0000

SL-9 Trace element solution (Tschech and Pfennig, 1984)

Nitrilotriacetic acid 12.8 g 67.0 mM

MnCl2 x 4 H2O 00.1 g 00.5 mM

FeCl2 x 4 H2O 02.0 g 10.1 mM

CoCl2 x 6 H2O 00.2 g 00.8 mM

ZnCl2 70.0 mg 00.5 mM

CuCl2 x 2 H2O 02.0 mg 11.7 µM

NaMoO4 x 2 H2O 36.0 mg 00.2 mM

NiCl2 x 6 H2O 24.0 mg 00.1 mM

H3BO3 06.0 mg 97.0 µM

NaCl 05.0 g 85.6 mM

H2O ad 1,000 ml0000

For preparation of trace element stock solution, nitrilotriacetic acid was dissolved in water by

adjusting pH to 6 using KOH. Afterwards, the other components were added and the final pH

was adjusted to 7.0 using KOH.

Page 36: 2,3-Butanediol production using acetogenic bacteria

22 2. Material and Methods

2.3.1.2 Modified ATCC 1612 medium (Hoffmeister et al., 2016)

Cultivation of A. woodii was performed in modified ATCC 1612 medium.

Modified ATCC 1612 medium

NH4Cl 00.20 g 03.7 mM

KH2PO4 01.76 g 12.9 mM

K2HPO4 08.44 g 48.5 mM

MgSO4 x 7 H2O* 00.33 g 01.3 mM

Vitamin stock solution (see above) 02.00 ml 00.2 % [v/v]

SL-9 trace element stock solution (see above) 02.00 ml 00.2 % [v/v]

Bacto® yeast extract 02.00 g 00.2 % [w/v]

NaHCO3 10.00 g 00.1 M000000000

Cysteine HCl x H2O 00.30 g 01.7 mM

Na2S x 9 H2O 00.30 g 01.2 mM

Resazurin 01.00 mg 04.4 µM

H2O ad 1,000 ml0000

*Sterile, anaerobic solution was supplemented after autoclaving.

As carbon source, a sterile and anaerobic stock solution of fructose (1.11 M) was added after

autoclaving.

2.3.1.3 Lyophilization medium (Flüchter, unpublished)

All anaerobic strains were stored by lyophilization, using lyophilization medium.

Lyophilization medium

Meat extract 002 g 000.2 % [w/v]

Sucrose 120 g 350.6 mM

Resazurin 001 mg 004.4 µM

H2O (anaerobic) ad 1,000 ml

Prior to use, 0.3 ml of anaerobic ferrous (II) sulfide stock solution (20 mg/ml) were added to

5 ml medium.

Page 37: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 23

2.3.1.4 Lysogeny broth medium -modified- (Bertani, 1951)

Cultivation of E. coli cells was done in lysogeny broth (LB) medium.

LB medium -modified-

Bacto® tryptone 10 g 001.0 % [v/v]

NaCl 10 g 171.1 mM

Bacto® yeast extract 05 g 000.5 % [v/v]

H2O ad 1,000 ml0000

2.3.1.5 Rajagopalan medium -modified- (Rajagopalan et al., 2002)

Cultivation of C. ragsdalei was performed in Rajagopalan medium.

Rajagopalan medium -modified-

MES 20.00 g 102.0 mM

Bacto® yeast extract 02.00 g 000.2 % [v/v]

NH4Cl 01.00 g 019.0 mM

NaCl 01.00 g 017.0 mM

KCl 00.10 g 001.3 mM

KH2PO4 00.10 g 000.7 mM

MgSO4 x 7 H2O 00.20 g 000.8 mM

CaCl2 00.02 g 000.2 mM

L-Cysteine-HCl x H2O 01.00 g 005.0 mM

Vitamin solution 10.00 ml 001.0 % [v/v]

PETC 1754 trace element solution 10.00 ml 001.0 % [v/v]

Resazurin 01.00 mg 004.4 µM

H2O ad 1,000 ml0000

All components were mixed and pH was adjusted to 6.1 using NaOH. As carbon source, a

sterile and anaerobic stock solution of fructose (1.11 M) was added after autoclaving.

Vitamin solution (Tanner, 2007)

Pyridoxine HCl 10 mg 49.0 µM

Thiamine HCl 05 mg 15.0 µM

Riboflavin 05 mg 13.0 µM

Calcium pantothenate 05 mg 21.0 µM

α-Lipoic acid 05 mg 24.0 µM

Page 38: 2,3-Butanediol production using acetogenic bacteria

24 2. Material and Methods

4-Aminobenzoic acid 05 mg 36.0 µM

Nicotinic acid 05 mg 41.0 µM

Vitamin B12 05 mg 03.7 µM

D-Biotin 02 mg 08.2 µM

Folic acid 02 mg 04.5 µM

2-Mercaptoethanesulfonic acid (MESNA) 02 mg 14.0 µM

H2O ad 1,000 ml0000

PETC 1754 trace element solution

Nitrilotriacetic acid 2.00 g 105 µM

MnSO4 x H2O 1.00 g 059 µM

Fe(NH4)2 (SO4)2 x 6 H2O 0.80 g 020 µM

CoCl2 x 6 H2O 0.20 g 712 µM

ZnSO4 x 7 H2O 1.00 mg 003 µM

CuCl2 x 2 H2O 0.02 g 117 µM

NaMoO4 x 2 H2O 0.02 g 097 µM

Na2SeO3 x 5 H2O 0.02 g 102 µM

NiCl2 x 6 H2O 0.02 g 084 µM

Na2WO4 x 2 H2O 0.02 g 064 µM

H2O ad 1,000 ml0000

For preparation of trace element solution, nitrilotriacetic acid was dissolved in water by

adjusting pH to 6 using KOH. Afterwards, the other components were added.

2.3.1.6 Super optimal broth medium -modified- (Sun et al., 2009b)

Chemically competent cells of E. coli were prepared in super optimal broth (SOB).

SOB medium -modified-

Bacto® tryptone 020.0 g 02.0 % [v/v]

NaCl 000.5 g 08.6 mM

Bacto® yeast extract 005.0 g 00.5 % [v/v]

KCl 190.0 g 02.6 M

MgCl2 (2M) 005.0 ml 10.0 mM

H2O ad 1,000 ml0000

*Sterile solution was added after autoclaving.

Page 39: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 25

2.3.1.7 Tanner medium -modified- (Tanner, 2007)

Cultivation of C. ljungdahlii, C. autoethanogenum, and C. coskatii was achieved in Tanner

medium. For cultivation of C. carboxidivorans, 3 g of yeast extract and for C. coskatii

2 g of yeast extract was used.

Tanner medium -modified-

MES 20.0 g 102.00 mM

Bacto® yeast extract 00.5 g 000.05 % [v/v]

Mineral solution 25.0 ml 000.50 % [v/v]

PETC 1754 Trace element solution (see above) 10.0 ml 001.00 % [v/v]

Vitamin solution (see above) 10.0 ml 001.00 % [v/v]

L-Cysteine HCl x H2O 01.0 g 005.00 mM

Resazurin 01.0 mg 004.40 µM

H2O ad 1,000 ml0000

All components were mixed and pH was adjusted to 5.9 using NaOH. As carbon source, a

sterile and anaerobic fructose stock solution (1.11 M) was added after autoclaving.

Mineral solution (Tanner, 2007)

NaCl 080 g 001.4 M

NH4Cl 100 g 001.9 M

KCl 010 g 134.0 mM

KH2PO4 010 g 073.0 mM

MgSO4 x 7 H2O 020 g 081.0 mM

CaCl2 x 2 H2O 004 g 027.0 mM

H2O ad 1,000 ml0000

2.3.1.8 Yeast tryptone fructose medium -modified- (Leang et al., 2013)

Cultivation of C. ljungdahlii after transformation via electroporation (see chapter 2.7.2) or

conjugation (see chapter 2.7.3) was performed on solid yeast tryptone fructose (YTF) medium.

YTF medium -modified-

Bacto® yeast extract 10 g 1.0 % [w/v]

Bacto® tryptone 16 g 1.6 % [w/v]

Fructose 05 g 27.8 mM

NaCl 04 g 68.4 mM

L-Cysteine-HCl x H2O 01 g 05.7 mM

Page 40: 2,3-Butanediol production using acetogenic bacteria

26 2. Material and Methods

Resazurin 01 mg 004.4 µM

H2O ad 1,000 ml0000

2.3.2 Antibiotic selection

The selection of recombinant strains was achieved by using antibiotics which are listed in

Table 5. Antibiotic stock solutions were 1000-fold concentrated, sterile-filtrated (sterile filter

“Filtropur S 2.0”, pore size 0.2 µm, Sarstedt AG & Co, Nümbrecht, Germany) and stored at

-20 °C. Addition of antibiotics to the media was done prior to use.

Table 5: Antibiotics used in this work

Antibiotic Stock solution

[mg/ml] (solvent)

Final

concentration

[μg/ml]

Organism Supplier

Ampicillin 100 (in H2O) 100 E. coli Carl Roth GmbH

& Co.KG,

Karlsruhe,

Germany

Chloramphenicol 30 (in 96 % [v/v] ethanol) 30 E. coli

Tetracycline HCl 10 (in 50 % [v/v] ethanol) 10 E. coli Fluka Chemie

GmbH, Buchs,

Switzerland

Kanamycine 50 (in H2O) 50 E. coli Carl Roth GmbH

& Co.KG,

Karlsruhe,

Germany

Colistin sodium

methanesulfonate

10 (in H2O) 10 E. coli Sigma-Aldrich

Chemie GmbH,

Steinheim,

Germany

Thiamphenicol* 15 (in

dimethylformamide or

acetonitrile)

7.5 A. woodii,

C. ljungdahlii,

C. coskatii

*light-sensitive solution

Page 41: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 27

2.4 Growth conditions

2.4.1 Aerobic cultivation

In general, E. coli was grown at a temperature of 37 °C. For preparation of chemically

competent cells (see chapter 2.7.1.4), cells were grown at 18 °C. In order to cultivate E. coli

for e.g. plasmid isolation (see chapter 2.5.1.1), 5 ml LB medium (see chapter 2.3.1.4) in test

tubes, closed with loose aluminium caps were used. Inoculation was performed with 10 µl of

E. coli stock (see chapter 2.4.3.1). For isolating higher amounts of plasmid DNA (see chapter

2.5.1.2), 500 ml medium in baffled 2-l Erlenmeyer flasks were inoculated with 5 ml of growing

culture. Optimal growth of E. coli strains was maintained by continuous oxygenation on a

rotary shaker at 120 to 180 rpm at 37 °C. Agar plates with E. coli strains were incubated upside

down at 37 °C.

2.4.2 Anaerobic cultivation

For 5 ml culture volume, Acetobacterium and Clostridium strains were grown anaerobically in

Hungate tubes (Ochs GmbH, Bovenden, Germany), which were closed with airtight synthetic

rubber stoppers (Ochs GmbH, Bovenden, Germany) and plastic screw caps (Ochs GmbH,

Bovenden, Germany). For larger volumes (50 ml and 100 ml), glass flasks (Müller & Krempel

AG, Bülach, Switzerland or Thermo Fisher Scientific Inc., Waltham, MA, USA), sealed with

natural rubber stoppers (Maag Technic GmbH, Göppingen, Germany) and stainless-steel

screw caps (Müller & Krempel AG, Bülach, Switzerland or Thermo Fisher Scientific Inc.,

Waltham, MA, USA) were used. Cultures of A. woodii, C. aceticum, and C. ragsdalei were

grown at 30 °C, while C. ljungdahlii, C. carboxidivorans, C. autoethanogenum, and C. coskatii

were cultivated at 37 °C. Incubation of agar plates took place upside down in an incubator

within the anaerobic chamber at 37 °C. Heterotrophic growth experiments were performed

in 50 ml medium in 125-ml anaerobic glass flasks (see above) and for autotrophic growth

experiments either 50 ml medium in 500-ml glass flasks or 100 ml in 1-l glass flasks were used.

The respective medium was inoculated with the corresponding strain to an optical density at

600 nm (OD600nm) of 0.1 for heterotrophic growth experiments and to OD600nm of 0.05 for

autotrophic growth experiments, respectively. Thereafter, OD600nm was monitored and

samples for product analysis were withdrawn. After each growth experiment, genomic DNA

was isolated and the identity of the strain was verified by amplification and sequencing of the

16S rDNA using the primers fD1 and rP2 (Table 3). For plasmid verification either primers

Page 42: 2,3-Butanediol production using acetogenic bacteria

28 2. Material and Methods

pMTL82151catP_fwd and pMTL82151catP_rev for pMTL82151 backbone or pJIR750catP_fwd

and pJIR750catP_rev for pJIR750 based vectors were used. Reactivation of sucrose-

magnesium-phosphate (SMP) buffer/dimethylsulfoxide (DMSO) cultures (see chapter 2.4.3.2)

was achieved by inoculation of 5 ml medium with 10 µl of culture. Subsequent passages of the

strains into medium were performed with approximately 10 % [v/v] of the final culture

volume. Lyophilized cells (see chapter 2.4.3.3) were reactivated by adding 2 ml of respective

medium to the cells, the pellet was allowed to dissolve and afterwards returned to the

Hungate tube. Three days of incubation at the respective temperature were sufficient for

reactivation.

2.4.3 Strain conservation

2.4.3.1 Glycerol

For conservation of E. coli strains, cells were grown over night at 37 °C in 5 ml LB medium (see

chapter 2.3.1.4) with respective antibiotics added (Table 5), if needed. 500 µl of this culture

were mixed with 500 µl of 50 % [v/v] glycerol and afterwards stored in cryogenic vials at

-80 °C.

2.4.3.2 SMP buffer/DMSO

For short-time conservation of anaerobic strains, the (sucrose-magnesium-phosphate) SMP

buffer/DMSO method was used (Hoffmeister, 2017). 5 ml medium in a Hungate tube were

inoculated with 500 µl of growing cells. When late exponential growth phase was reached,

cells were harvested in a “Hungate tube centrifuge EBA20” (Andreas Hettich GmbH & Co. KG,

Tuttlingen, Germany) at 2,400 x g for 10 min at 4 °C. Afterwards, supernatant was discarded

with a 5-ml syringe and cell pellet was dissolved in 400 µl SMP buffer and 200 µl anti-freezing

buffer. The cell suspension was stored at -80 °C.

SMP buffer

Sucrose 92.40 g 270 mM

MgCl2 x 6 H2O 00.20 g 001 mM

NaH2PO4 00.84 g 007 mM

H2O ad 1,000 ml0000

Page 43: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 29

The pH was adjusted to 6 by adding HCl, buffer was boiled for 20 min and cooled on ice for 20

min, while sparging with N2. Buffer was transferred in an anaerobic chamber and filled in

airtight glass flasks (see chapter 2.4.2).

Anti-freezing buffer

SMP buffer 20 ml 40 % [v/v]

DMSO 30 ml 60 % [v/v]

Anti-freezing buffer was prepared with anaerobic SMP buffer and DMSO in an anaerobic

chamber and filled in airtight glass flasks (see chapter 2.4.2).

2.4.3.3 Lyophilization

In order to preserve anaerobic strains for a long time, lyophilization was performed.

Acetobacterium or Clostridium were grown in a Hungate tube to late exponential growth

phase. Thereupon, cells were sedimented in a “Hungate tube centrifuge EBA20” (Andreas

Hettich GmbH & Co. KG, Tuttlingen, Germany) at 2,400 x g for 10 min at 4 °C, supernatant was

removed with a 5-ml syringe and resuspended in 1.5 ml of anaerobic lyophilization medium

(see chapter 2.3.1.3). Aliquots of 0.4 ml were transferred under sterile conditions into sterile

8-ml glass vials, closed instantly with sterile cotton wool plugs (Paul Hartmann AG,

Heidenheim, Germany) and flash-freezed in liquid N2. For keeping cells in this frozen state, a

cooling block (-80 °C) was used to take the cells to the pre-cooled lyophilisator (-54 °C) (Christ

Alpha 1-4 freeze-dryer; Martin Christ Gefriertrocknungsanlagen GmbH, Osterode am Harz,

Germany). Then, vacuum was applied for 19 h, which represents the first drying step.

Afterwards, cotton wool plugs were exchanged under sterile conditions with butyl rubber

stoppers and plastic screw caps (Ochs GmbH, Bovenden, Germany). Thereafter, a second

vacuum step took place by piercing each butyl rubber stopper with a sterile cannula which

was linked to the lyophilisator by a sterilised manifold. 24 hours later, cannulas were removed

and the freeze-dried cells were stored at 4 °C.

2.4.4 Determination of growth and metabolic production parameters

2.4.4.1 Determination of optical density

For determination of bacterial growth, optical density at 600 nm (OD600nm) of 1 ml sample was

determined in 1.6-ml semi-micro cuvettes with 10 mm optical pathway (Sarstedt AG & Co.,

Nümbrecht, Germany) by using “Ultrospec® 3100 pro” (Amersham Bioscience Europe GmbH,

Page 44: 2,3-Butanediol production using acetogenic bacteria

30 2. Material and Methods

Freiburg, Germany). If OD600nm was exceeding a value of 0.3, samples were diluted by with

water to obtain values in the linear range of the photometer. In this case, the sterile medium

being used as blank was also diluted in the same ratio with water. For calculation of OD600nm

the respective dilution factor was taken into account.

2.4.4.2 Quantification of substrate and microbial metabolic products by high performance

liquid chromatography

The method and parameters for high performance liquid chromatography (HPLC) analysis

were described previously (Nielsen et al., 2010). The following method was used to determine

fructose, acetate, lactate, and stereoisomers of 2,3-BD in 300 µl cell culture supernatant

(centrifugation 15,000 x g, 30 min, 4 °C), which were transferred in 2-ml vials (CS-

Chromatographie Service GmbH, Langerwehe, Germany) and sealed with aluminium caps (CS-

Chromatographie Service GmbH, Langerwehe, Germany). Analysis was performed with a

“Infinity 1260” HPLC (Agilent Technologies, Waldbronn, Germany), equipped with an

autosampler and pre-column followed by a CS-Organic acid column (CS-Chromatographie

Service GmbH, Langerwehe, Germany). For detection of organic acids (e.g. acetate, lactate) a

diode array UV detector (DAD) was used. Analysis of fructose and solvents (e.g. 2R,3R-BD) was

performed with a refraction index detector (RID). Calibration was achieved by using 1-ml

samples with different concentrations of the investigated compounds (1 mM-200 mM) which

served as external standards. All data analysis was done with “Agilent OPEN-Lab CDS Version

A.01.03.” (Agilent Technologies, Waldbronn, Germany).

The following parameter setup was used for HPLC analysis:

Column: CS-Organic acid (150 mm x 8 mm)

Pre-column: CS-Organic acid (40 mm x 8 mm)

Packing material: Poly(styrene-divinylbenzene)

Mobile phase: 5 mM H2SO4 (0.7 ml/min)

Injection temperature: 25 °C

Column temperature: 40 °C

Detector temperature: 35 °C

Detection wavelength (DAD): 210 nm

Running time: 24 min

Sample volume: 20 µl at 15 °C autosampler

Page 45: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 31

2.4.4.3 Quantification of microbial metabolic products by gas chromatography

The method used for quantification of metabolic products by gas chromatography (GC) was

modified from a previously described protocol (Playne, 1985). For this purpose, 1 ml of culture

supernatant (centrifugation at 15,000 x g, 30 min, 4 °C) in 2-ml vials (CS-Chromatographie

Service GmbH, Langerwehe, Germany), sealed with aluminium caps (CS-Chromatographie

Service GmbH, Langerwehe, Germany) was needed to measure ethanol as well as 2,3-BD in

Clostridium and A. woodii by using a “Clarus 600” gas chromatograph (PerkinElmer LAS GmbH,

Rodgau, Germany). It was equipped with a flame ionization detector (FID), an autosampler,

and a column packed with Porapak P. Isobutanol solution (110 mM isobutanol in 2 M HCl)

served as internal standard (100 µl for 1 ml sample). For calibration, one vial containing

internal standard and 5 mM of each analyzed compound was used. All data analysis was

executed with “TotalChrom Version 6.3.1” (PerkinElmer, LAS GmbH, Rodgau, Germany).

The GC analysis was performed with the following conditions for ethanol analysis:

Column: Metal (ID 2 mm x 2,000 mm)

Packing material: Porapak P (80/100 mesh)

Carrier gas: N2 (2.17 bar)

Injection temperature: 200 °C

Detector temperature: 300 °C

Detector type: Flame ionization detector (FID)

Detector gases: H2 (45 ml/min flow rate),

synthetic air (450 ml/min flow rate)

Temperature setup: 130 °C for 1 min,

130°C to 140 °C at 5 °C/min,

140 °C for 10 min

140 °C to 170 °C at 25 °C/min,

170 °C for 1 min

Sample volume: 1 µl

For 2,3-BD analysis along with acetonitrile (solvent of thiamphenicol for 2,3-BD analysis in

A. woodii) the following parameters were used:

Column: Metal (ID 2 mm x 2,000 mm)

Packing material: Porapak P (80/100 mesh)

Carrier gas: N2 (1.5 bar)

Page 46: 2,3-Butanediol production using acetogenic bacteria

32 2. Material and Methods

Injection temperature: 200 °C

Detector temperature: 300 °C

Detector type: Flame ionization detector (FID)

Detector gases:

H2 (45 ml/min flow rate),

synthetic air (450 ml/min flow rate)

Temperature setup: 160 °C for 1 min,

160°C to 170 °C at 1 °C/min,

Sample volume: 1 µl

2.5 Nucleic acid methods

2.5.1 Nucleic acid isolation

2.5.1.1 Plasmid isolation from E. coli using the ZyppyTM Plasmid Miniprep Kit

For isolating small amounts of plasmid DNA, the “ZyppyTM Plasmid Miniprep Kit” (ZYMO

Research Europe GmbH, Freiburg, Germany) was used. This kit contained all necessary

reagents and columns. The instructions were slightly modified in the following way. 3 ml of an

overnight grown E. coli culture were harvested at 13,000 x g for 1 min in the same 1.5-ml

reaction tube. All steps were performed at room temperature. The cell pellet was dissolved in

600 µl sterile water. Then, a washing step with 400 µl Zymo-SpinTM wash buffer was conducted

(13,000 x g for 30 s). After the last washing step, drying of the column was performed at

10,000 x g for 3 min. Finally, plasmid DNA was eluted with 35 µl sterile water at 13,000 x g for

15 s, instead of tris-EDTA (TE) buffer. Plasmid DNA was stored at -20 °C.

2.5.1.2 Plasmid isolation from E. coli using QIAGEN Plasmid Midi Kit

For isolating larger volumes of plasmid DNA with a high concentration, which is needed for

transformation of plasmid DNA in anaerobic bacteria, the “QIAGEN Plasmid Midi Kit”

(QIAGEN-tip 100; Qiagen GmbH, Hilden, Germany) was used. The protocol which was applied

is for isolating very low-copy plasmids and cosmids with slight modifications. E. coli was grown

in 500 ml LB medium in baffled 2-l Erlenmeyer flasks with shaking at 300 rpm overnight. Cells

were harvested in a centrifugation vessel at 6,000 x g for 15 min at 4 °C. Then, the bacterial

pellet was resuspended in 20 ml buffer P1 containing RNase A. Afterwards, 20 ml of buffer P2

were added to the sample and mixed by vigorously inverting the sealed vessel 4-6 times. After

incubation at room temperature for 5 min, 20 ml of chilled buffer P3 were added and the

solution was mixed immediately and thoroughly by inverting 4-6 times. Thereafter, incubation

Page 47: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 33

for 30 min on ice followed. The sample was centrifuged at 20,000 x g for 30 min at 4 °C to

sediment the cell debris. Before centrifugation the sample was mixed again. After adding

buffer P1, P2, and P3 and centrifugation at 20,000 x g for 30 min at 4 °C, the supernatant

containing plasmid DNA was directly filtered over a pre-wetted, folded filter paper (ø 125 mm;

Schleicher & Schuell Bioscience GmbH, Dassel, Germany) to get rid of cell debris. In the end,

DNA pellet was air-dried for 15 min at 60 °C and resuspended in 150 µl Tris-EDTA (TE) buffer

(pH 8). Plasmid DNA was stored at -20 °C.

TE buffer

Tris 1.20 g 10 mM

EDTA, disodium salt 0.37 g 01 mM

H2O ad 1,000 ml0000

The pH value was adjusted to 8, using 1 M HCl.

2.5.1.3 Genomic DNA isolation from Gram positive bacteria using MasterPureTM Gram

Positive DNA Purification Kit

In order to isolate genomic DNA from Gram-positive bacteria for 16S rDNA PCR or

retransformation into E. coli cells, the “MasterPureTM Gram Positive DNA Purification Kit”

(Epicentre Technologies Corp., an Illumina company, Madison, WI, USA) was used. Harvesting

of cells was done with 4 ml of a stationary growth phase culture at 13,000 x g for 3 min at

room temperature and isolation was handled according to the kit instructions with slight

modifications. Incubation of cells with lysozyme was performed overnight at 37 °C. After the

washing step with 70 % [v/v] ethanol and centrifugation at 10,000 x g for 2 min, DNA was

allowed to air-dry and resuspended in 35 µl TE buffer. Genomic DNA was stored at -20 °C.

2.5.1.4 RNA isolation from clostridia

For isolation of RNA, all material and reagents were autoclaved twice to eliminate RNases.

While working with RNA, wearing nitrile gloves was obligatory to minimize the possibility of

RNA degradation by RNases of human skin. Furthermore, all work surfaces were cleaned with

“RNase-Exitus PlusTM “(AppliChem GmbH, Darmstadt, Germany), which destroys RNases. All

reagents were treated with RNase-free SafeSeal-Tips® (Biozym Scientific GmbH, Oldenburg).

Clostridium culture for RNA isolation was grown in 50 ml of the respective medium for 65 h

and harvested in 50-ml falcon tubes at 6,000 x g for 10 min at 4 °C. Cell pellet was stored at

-80 °C overnight. After thawing the pellet on ice, 600 µl lysis buffer (20 mg/ml lysozyme in tris-

Page 48: 2,3-Butanediol production using acetogenic bacteria

34 2. Material and Methods

sucrose-EDTA (TSE) buffer) were added to the sample and vortexed. The following steps were

performed with 700 µl of this sample in 2-ml reaction tubes. Cells were incubated at 37 °C for

5 min. After adding 1 ml chilled TRI Reagent® (Sigma-Aldrich Chemie GmbH, Steinheim), the

sample was pipetted up and down. Thereafter, 0.3 ml of chloroform/isoamyl alcohol (24:1 v/v)

were applied and the solution was mixed vigorously. After an incubation at room temperature

for 3 min, sample was centrifuged at 12,000 x g for 15 min at 4 °C. The upper aqueous phase

was transferred into a 1.5-ml reaction tube and 0.5 ml cold isopropanol were added. After

inverting the tube several times, sample was centrifuged at 12,000 x g for 10 min at 4 °C.

Supernatant was removed and pellet was washed with 1 ml of chilled 70 % ethanol [v/v]. Then,

sample was centrifuged at 12,000 x g for 4 min at 4 °C and supernatant was discarded. Finally,

RNA pellet was air-dried for 20 min and resuspended in 44 µl of RNase-free water. If the pellet

could not be completely resuspended, sample was incubated at 55 °C for 15 min.

The RNA isolation protocol was completed with a DNA digest method by

Invitrogen™ Ambion™ TURBO DNA-free Kit (Thermo Fisher Scientific Inc., Waltham, MA, USA).

For standard DNA digestion, the following reaction mixture was used:

RNA 44 µl ~ 20 µg/µl

10 x TURBO DNase buffer 05 µl 10 % [v/v]

TURBO DNase 01 µl 02 % [v/v]

H2O ad 50 µl

After 30 min incubation at 37 °C, another 1 µl of TURBO DNase was added and sample was

incubated for 30 min at 37 °C. Thereafter, 5 µl DNase inactivation reagent were applied and

sample was vortexed. After incubation at room temperature for 5 min, RNA was centrifuged

at 10,000 x g for 90 s. DNA-free RNA was transferred into a new 1.5-ml reaction tube.

Finally, RNA concentration was determined by using the spectral photometer

NanoDrop 2000c (Thermo Fisher Scientific Inc., Waltham, MA, USA) (see chapter 2.5.6).

TSE buffer

Sucrose 62.50 g 730 mM

Tris-HCl 01.50 g 050 mM

EDTA, disodium salt 04.65 g 050 mM

H2O ad 250 ml0000

The pH value was adjusted to 7.6, using 1 M NaOH.

Page 49: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 35

2.5.2 Polymerase chain reaction

Polymerase chain reaction (PCR) is a method, which was used in this study for amplification

of specific DNA fragments, e.g. for cloning procedures or strain verification (Mullis et al.,

1986). PCR requires a three-step cycling process. In the first step, denaturation at 95 °C

separates double-stranded DNA. In the second step, single-stranded primers attach to the

dissociated DNA, while each primer is complementary to either 5' end or 3' end of the

sequence of interest. In the third step, DNA polymerase synthesizes the new DNA strands

starting from the respective primer (Schochetman et al., 1988). Primer design was performed

with the software “Clone Manager 7.11“ (Scientific & Educational Software, Cary, NC USA) or

NEBuilder® Assembly Tool (New England Biolabs® Inc., Ipswich, MA, USA). For different DNA

polymerases, the polymerase chain reaction mixtures changed slightly (see chapter 2.5.2.1).

2.5.2.1 Composition of PCR reaction mixtures with different DNA polymerases

In this study, the amplification of DNA fragments for proof of plasmid presence in recombinant

strains was done with Taq DNA polymerase (Genaxxon Bioscience GmbH, Ulm, Germany). In

contrast, amplification of genes, promoters, or terminators for cloning procedures, for which

proof reading is obligatory, was performed with ReproFast DNA polymerase (Genaxxon

Bioscience GmbH, Ulm, Germany). If problems occurred with DNA amplification (e.g. no PCR

product), the ReproFast DNA polymerase was used with a slightly modified method including

“FailSafeTM PCR Premix E” (Epicentre Technologies Corp., an Illumina company, Madison, WI,

USA). This premix contains a PCR enhancer with betaine which is helpful to amplify DNA

fragments. DNA fragments to be used in the cloning procedure “In-Fusion® HD Cloning Kit”

(Takara Bio USA Inc., Mountain View, CA, USA), were amplified with the “CloneAmpTM HiFi

polymerase” (Takara Bio USA Inc., Mountain View, CA, USA).

PCR with Taq DNA polymerase

DNA template 3.00 µl ~ 10.00 ng/µl

dNTPs (10 mM each) 1.00 µl ~ 10.20 mM each

10 x Buffer with MgCl2 5.00 µl ~ 10.00 % [v/v]

Primers (100 µM each) 0.50 µl each ~ 11.00 µM each

Taq polymerase 0.25 µl ~ 10.04 U/µl

H2O ad 50 µl

Page 50: 2,3-Butanediol production using acetogenic bacteria

36 2. Material and Methods

PCR with ReproFast polymerase

DNA template 3.00 µl ~ 10.00 ng/µl

dNTPs (10 mM each) 1.00 µl ~ 10.20 mM each

10 x ReproFast buffer with MgSO4 5.00 µl ~ 10.00 % [v/v]

Primers (100 µM each) 0.50 µl each ~ 11.00 µM each

ReproFast polymerase (5 U/µl) 0.25 µl ~ 10.03 U/µl

H2O ad 50 µl

PCR with ReproFast polymerase and FailSafeTM PCR Premix E

DNA template 03.00 µl ~ 10.00 ng/µl

FailSafeTM PCR Premix E 25.00 µl ~ 50.00 % [v/v]

Primers (100 µM each) 00.50 µl each ~ 11.00 µM each

ReproFast polymerase (5 U/µl) 00.5 µl ~ 10.05 U/µl

H2O ad 50 µl

PCR with CloneAmpTM HiFi PCR Premix

DNA template 01.00 µl <100.0 ng

2 x CloneAmpTM HiFi PCR Premix 12.50 µl <050.0 % [v/v]

Primers (10 µM each) 01.00 µl each <000.2 µM

H2O ad 25 µl

2.5.2.2 Standard PCR program

The standard PCR program was performed with the following conditions:

1. Initial denaturation 95 °C 300 s

2. Denaturation 95 °C 045 s

3. Annealing Variable* 045 s

4. Elongation 72 °C Variable**

5. Final elongation 72 °C 600 s

6. Cooling 12 °C Forever

*The annealing temperature was either calculated with the software Tm calculator (Thermo Fisher

Scientific Inc., Waltham, MA, USA) or NEBuilder® Assembly Tool (New England Biolabs® Inc., Ipswich,

MA,USA)

**Elongation time was calculated by size of the amplified DNA fragment and speed of the utilized DNA

polymerase: Taq polymerase (1,000 bps/60s) and ReproFast polymerase (600 bps/60s)

MA, USA)

Page 51: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 37

Steps 2-4, which represent one complete PCR cycle, were repeated 32 times before going to

step 5. For PCRs with “CloneAmpTM HiFi PCR Premix” (Takara Bio USA Inc., Mountain View, CA,

USA) a slightly modified PCR program was used:

1. Denaturation 95 °C 010 s

2. Annealing 55 °C 015 s

3. Elongation 72 °C Variable*

4. Cooling 04 °C Forever

*Elongation time was calculated by size of the amplified DNA fragment and speed of the CloneAmpTM

HiFi polymerase (1,000 bps/30-60 s)

Steps 1-3, which are one complete PCR cycle, were repeated 30 times before going to step 4.

2.5.2.3 Insertion of restriction recognition sequences via PCR

By means of PCR, restriction recognition sequences can be attached to the DNA fragment of

interest. The desired restriction recognition sequence was linked to the 5'-end of the utilized

primer. Additionally, the bases ACA were set upstream of the restriction recognition

sequence, since restriction enzymes require several bases to bind and cut DNA.

2.5.3 DNA modification

2.5.3.1 Restriction enzyme digestion of DNA

Restriction enzyme digestion was used for analytical and preparative digest of plasmids and

DNA fragments for cloning procedures and recombinant strain analysis, respectively.

Composition of the reaction mixture, when using one restriction enzyme, was prepared

following the instructions of the manufacturer, whereas the setting for using two restriction

enzymes was planned with the “DoubleDigestTM tool” (Thermo Fisher Scientific Inc., Waltham,

MA, USA). If vector was digested with only one enzyme for linearization, dephosphorylation

(see chapter 2.5.3.2) was mandatory after inactivation of restriction enzyme.

Analytical restriction enzyme digestion

DNA (plasmid) 3 µl ~ 150.0 ng/µl

10 x FastDigest green buffer 2 µl ~ 010.0 % [v/v]

FastDigest restriction enzyme 1 µl 0~ 00.5 U/µl

H2O ad 20 µl

The analytical digestion was performed in 1.5-ml reaction tubes at 37 °C for 30 min.

Page 52: 2,3-Butanediol production using acetogenic bacteria

38 2. Material and Methods

Preparative restriction enzyme digestion

DNA (plasmid or DNA fragment) 5 µl ~ 150.0 ng/µl

10 x FastDigest green buffer 5 µl ~ 010.0 % [v/v]

FastDigest restriction enzyme 2 µl ~ 000.5 U/µl

H2O ad 50 µl

The preparative digestion was performed in 1.5-ml reaction tubes at 37 °C for 30 min.

2.5.3.2 Dephosphorylation of digested plasmids

To avoid religation of linearized plasmids, which were constructed by cutting with one

restriction enzyme, the 5'-phosphate groups of the sticky ends must be removed prior to

ligation (see chapter 2.5.3.3). For this purpose, the “FastAP Thermosensitive Alkaline

Phosphatase” (Thermo Fisher Scientific Inc., Waltham, MA, USA) was the enzyme of choice.

After inactivation of restriction enzymes used in the digestion reaction, 1 µl of FastAP

Thermosensitive Alkaline Phosphatase (1 U/l) were added to the preparative digestion,

vortexed and incubated at 37 °C for 30 min. After an inactivation step at 65 °C for 15 min, the

dephosphorylated DNA was purified via “DNA Clean & ConcentratorTM-5 Kit” (ZYMO Research

Europe GmbH, Freiburg, Germany) (see chapter 2.5.4.1). This purified vector was ready to be

used in DNA ligation.

2.5.3.3 DNA ligation

In order to construct plasmids of choice, purified DNA fragments (e.g. digested PCR product

or restriction enzyme digested DNA fragment; see chapter 2.5.3.1) and purified, digested

vectors were linked by DNA ligation (Weiss et al., 1968). The enzyme needed for DNA ligations

is “T4 DNA ligase” (Thermo Fisher Scientific Inc., Waltham, MA, USA), which is capable to form

phosphodiester bonds of exposed 5'-phosphate and adjacent 3'-hydroxyl terminal groups in

double-stranded DNA molecules. This reaction requires ATP as cofactor, which is present in

the DNA ligase buffer. ATP is a molecule, which is very unstable, since it contains three

negatively charged phosphate groups. For this reason, ligase buffer is stored in 10-µl aliquots

for not thawing and freezing too often. The ratio between insert DNA fragment and vector

backbone was set in between 1:1 or 7:1. There was always a control for each ligation

experiment, which did not contain any insert DNA. This reaction served as control for

recircularization of vector backbone. Ligations were carried out by using the following reaction

composition:

Page 53: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 39

Insert DNA 1-7 µl ~ 10.00-70.00 ng/µl

Linearised vector 0-1 µl 00000~ 10.00 ng/µl

10 x T4 DNA ligase buffer 0-2 µl 00000~ 10.00 % [v/v]

T4 DNA ligase (5 U/µl) 0-1 µl 000000~ 0.25 U/µl

H2O ad 20 µl

Ligation reaction mixture was pipetted into a 0.2-ml microtube and was incubated on a

thermal cycler for 1.5 h at 22 °C. After inactivation of T4 DNA ligase, ligation reaction mixture

was transformed into chemically competent E. coli cells (see chapter 2.7.1.4). If ligation

mixture was transformed into electrocompetent E. coli cells, ligation mixture was purified by

dialysis (see chapter 2.5.4.3).

2.5.3.4 In-Fusion® HD Cloning Kit

By using “In-Fusion HD Cloning Kit” (Takara Bio USA Inc., Mountain View, CA, USA) one or more

DNA fragments can be cloned directly into any vector. The “In-Fusion Enzyme” fuses DNA

fragments (e.g. PCR products and linearized vectors) by recognizing a 15-bp overlap at their

ends. This overlap is engineered by primer design for amplification of desired DNA fragments.

Designing of required primers (see chapter 2.1.3) was accomplished by using the web-based

“NEBuilder® Assembly Tool” (New England Biolabs® Inc., Ipswich, MA, USA). “In-Fusion HD

Cloning” belongs to the next generation cloning techniques like circular polymerase extension

cloning (CPEC) or Gibson Assembly (Quan and Tian, 2011; Gibson et al., 2009). DNA fragment

and linearized vector were cut from agarose gel and purified (see chapter 2.5.5.1). The “In-

Fusion cloning” reaction was performed by using the following set-up:

Purified PCR fragment 0.5 µl 50-100 ng

Purified linearized vector 4.0 µl 50-200 ng

5 x In-Fusion HD enzyme premix 2.0 µl 20 % [v/v]

H2O ad 10 µl

The reaction mixture was incubated at 50 °C for 15 min in 0.25-ml microtubes. Subsequently,

sample was placed on ice and 5 µl of “In-Fusion cloning” reaction mixture was transformed

into chemically competent E. coli cells (see chapter 2.7.1.5).

2.5.3.5 Radioactive labelling of DNA fragments

In order to prepare [α-32P]-labelled DNA probes (10mCi/ml) for Northern blot experiments

“Random Primers DNA Labelling System Kit” (Thermo Fisher Scientific Inc., Waltham, MA,

Page 54: 2,3-Butanediol production using acetogenic bacteria

40 2. Material and Methods

USA) was performed according to manufacturer’s instructions. For radioactive labelling, 25 ng

(dissolved in 20 µl sterile water) of the desired DNA fragment were used. After addition of

5 µl stop buffer, non-incorporated radioactive bases were removed by using “illustra™

MicroSpin™ G-25 Columns” (GE Healthcare, Buckinghamshire, GB) (see chapter 2.5.4.2). For

measurement of total radioactivity, 3 µl sample (1:1000 dilution) were transferred to a

scintillation vessel and 4 ml of scintillation solution were added. The radioactivity of the

sample was measured in a “Tri-Carb® 2800TR Liquid Scintillation Analyzer” (PerkinElmer LAS

GmbH, Rodgau, Germany).

2.5.4 Purification of nucleic acids

2.5.4.1 Purification of amplified and digested DNA

In order to purify DNA fragments after PCR (see chapter 2.5.2) or after DNA digest (see chapter

2.5.3.1) either “ZymocleanTM Gel DNA Recovery Kit” or “DNA Clean & ConcentratorTM-5 Kit”

(ZYMO Research Europe GmbH, Freiburg, Germany) were used. If no unspecific PCR fragments

were present or for purification of digested DNA fragments, the “Clean & ConcentratorTM-5

Kit” (ZYMO Research Europe GmbH, Freiburg) was used to remove all undesired reaction

components. The kit was used according to the manufacturer’s instructions with slight

modifications. After washing the column, it was dried by centrifugation at 10,000 x g for 3 min.

Finally, it was placed into a sterile 1.5-ml reaction tube and DNA elution was achieved by

adding 10 µl sterile water to the column, and centrifugation at 10,000 x g for 30 s. If PCR

reactions contained DNA fragments, caused by unspecific binding of primers or for purification

of preparative digests (see chapter 2.5.3.1) the “ZymocleanTM Gel DNA Recovery Kit” (ZYMO

Research Europe GmbH, Freiburg, Germany) was applied. At first, the DNA fragment of

interest was cut with a scalpel from stained agarose gel (see chapter 2.5.5.1) by transferring

the gel on a “UST-30L-8E” UV-transilluminator (λ = 365 nm; biostep GmbH, Burkhardtsdorf,

Germany)). The gel slice was placed into a 1.5-ml reaction tube and weighed. Subsequently,

the kit was used according to manufacturer’s instructions with the following modifications. As

mentioned above, a drying step was performed at 10,000 x g for 3 min after the washing step

and DNA was eluted with 10 µl of sterile water (10,000 x g, 30 s).

Page 55: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 41

2.5.4.2 Purification of radioactively labelled DNA

For the purpose of separating radioactive labelled DNA from non-incorporated [α-32P]-dATP,

the “illustra™ MicroSpin™ G-25 Columns” (GE Healthcare, Buckinghamshire, GB) were used

according to the manufacturer’s instructions.

2.5.4.3 Purification of ligated DNA by microdialysis

If plasmid DNA was transformed after ligation into electrocompetent E. coli cells (see chapter

2.7.1.3), purification via microdialysis was mandatory. This step is required, since T4 DNA

ligase buffer (Thermo Fisher Scientific Inc., Waltham, MA, USA) contains a high concentration

of MgCl2 (100 mM). This high salt concentration in ligation reaction mixture would cause arcing

during electroporation. Microdialysis was performed with the complete ligation reaction

mixture by using “MFTM-Membrane Filters” (0.025 µm; Merck KGaA, Darmstadt, Germany).

The sample was pipetted onto the filter, which was placed in an ultra-pure water containing

petri dish. After 15 min incubation, the scaled-down volume of ligation sample was ready to

be used in transformation of electrocompetent E. coli cells.

2.5.5 Gel electrophoresis

In order to separate nucleic acids according to their size, the method gel electrophoresis was

applied. The phosphate groups of nucleic acids are loaded negatively and so they move to the

anode in an electric field. Due to the molecular structure of agarose gels, nucleic acids move

at different speeds through the gel. Thus, smaller fragments move faster than larger nucleic

acids (Green and Sambrook, 2012).

2.5.5.1 Non-denaturating gel electrophoresis

Agarose gel electrophoresis can be used to separate DNA fragments according to their size. In

this manner, restriction digests and PCR DNA fragments were analyzed, using the size indicator

“GeneRulerTM DNA ladder Mix” (Thermo Fisher Scientific Inc., Waltham, MA, USA). DNA with

a size over 1000 bp was separated by 0.8 % agarose [w/v] and for smaller fragments 2.0 %

agarose [w/v] was used. Agarose was dissolved in 1 x TAE buffer by heating in a microwave

and then poured in an electrophoresis chamber equipped with a plastic comb. After

polymerization, the gel was overlaid with 1 x tris-acetic acid-EDTA buffer (TAE) buffer and DNA

sample was mixed with “6 x DNA loading dye” (Thermo Fisher Scientific Inc., Waltham, MA,

USA) in a ratio of 5:1. This dye contains tracking components to monitor electrophoresis by

migration of xylene cyanol FF (4160 bp) and bromphenol blue (370 bp). Electrophoresis was

Page 56: 2,3-Butanediol production using acetogenic bacteria

42 2. Material and Methods

performed at 120 V covering a distance of 7 cm. For visualizing the DNA at a wavelength of

312 nm with a “TFP-M/WL” UV-transilluminator (biostep GmbH, Burkhardtsdorf, Germany),

the gel was stained with 0.5 % [w/v] ethidium bromide. In order to save the results of DNA

staining, the gel was photographed by using a photo documentation facility (Decon Science

Tec GmbH, Hohengandern, Germany).

50 x TAE buffer (Green and Sambrook, 2012)

Tris 242.3 g 02 M

Glacial acetic acid 057.1 g 01 M

EDTA 014.6 g 50 mM

H2O ad 1,000 ml

The pH was adjusted to 7.5 with HCl.

2.5.5.2 Denaturating gel electrophoresis

In order to separate RNA, denaturating gel electrophoresis was the method of choice (Lehrach

et al., 1977). For this purpose, 2.4 g agarose was boiled in 180 ml water (autoclaved twice)

and afterwards 20 ml 10 x MOPS-EDTA-sodium acetate (MESA) buffer, 3.6 ml formaldehyde,

and 2 µl ethidium bromide were added. After cooling down to 50 °C, the solution was poured

in an electrophoresis chamber equipped with a plastic comb. Then, the gel was allowed to

polymerize for 1 h. Afterwards, it was overlaid with 1 x MAE running buffer and pre-

electrophoresis was performed at 100 V for 30 min. RNA samples were diluted 1:2 with “2 x

RNA loading dye” (Thermo Fisher Scientific Inc., Waltham, MA, USA) and pipetted into the gel

pockets. 3 µl of “RiboRulerTM High Range DNA Ladder” (Thermo Fisher Scientific Inc., Waltham,

MA, USA) served as size indicator. The electrophoresis of RNA molecules was performed at

100 V for 3 h and 45 min. RNA was made visible by an “TFP-M/WL” UV transilluminator (λ =

312 nm; biostep GmbH, Burkhardtsdorf, Germany) and gel was photographed by using a

photo documentation facility (Decon Science Tec GmbH, Hohengandern, Germany). Bands of

the size standard were marked with a scalpel in the gel for later labelling by probe

hybridization.

Page 57: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 43

10 x MESA buffer

MOPS 41.85 g 200 mM

Sodium acetate x 3 H2O 06.80 g 050 mM

EDTA, disodium salt 01.86 g 005 mM

H2O ad 1,000 ml

The pH was adjusted to 7 with HCl.

1 x MESA running buffer

10 x MESA buffer 150 ml 10 % [v/v]

Formaldehyde 030 ml 03 % [v/v]

H2O ad 1,500 ml

2.5.6 Determination of nucleic acid concentration

The absorption of DNA was measured at wavelengths of 260 nm and 280 nm with a spectral

photometer “NanoDrop 2000c” (Thermo Fisher Scientific Inc., Waltham, MA, USA). Ratio of

absorptions (A260/A280) informs about contamination of DNA with proteins. The ratio for

DNA should be in between 1.7 and 2.0 and for RNA between 2.0 and 2.2. If the absorption

accounts for 1, the concentration of double stranded DNA is 50 µg/ml and for RNA 40 µg/ml,

respectively (Green and Sambrook, 2012). The solvent of DNA and RNA served as blank value.

2.6 DNA sequencing and data analysis

2.6.1 Sequencing of vectors and DNA fragments

Sequencing of vectors or DNA fragments was done by GATC Biotech AG (Constance, Germany)

using custom primers or universal primers. The method “SUPREMErun” is based on Sanger

sequencing. It requires purified DNA (30-100 ng/µl for plasmids and 10-50 ng/µl for DNA

fragments in a volume of 20 µl) in 1.5-ml reaction tubes. Resulting sequences were analyzed

using “Clone Manager 7.11“ (Scientific & Educational Software, Cary, NC, USA) for plasmid

validation after cloning and transformation processes. Validation of bacterial strains via 16S

rDNA sequencing was achieved with Basic Local Alignment Search Tool (BLAST) by using the

method “Nucleotide BLAST” (BLASTN). This browser-based tool of the National Center for

Biotechnological Information (NCBI) (http://blast.ncbi.nlm.nih.gov) aligns different nucleotide

databases using a nucleotide query (Altschul et al., 1990). If any mismatches of bases

appeared, chromatograms of the original sequencing data were further analyzed.

Page 58: 2,3-Butanediol production using acetogenic bacteria

44 2. Material and Methods

2.6.2 Genome sequencing

Chromosomal DNA of C. ljungdahlii DSM 13528 and C. ljungdahlii [pMTL82151] was isolated

using the “MasterPureTM Gram Positive DNA Purification Kit” (Epicentre, Madison, WI, USA)

(see chapter 2.5.1.3). Sequencing of respective strains was performed by the Göttingen

Genomics Laboratory (Göttingen, Germany). Resequencing of C. ljungdahlii wildtype was done

to search for possible sequence errors in the original sequence in order to exclude false

positive single nucleotide polymorphisms (SNPs). Illumina shotgun libraries were generated

from the extracted DNA according to the manufacturer’s protocol. Sequencing was performed

using a MiSeq instrument and the MiSeq reagent kit version 3, as recommended by the

manufacturer (Illumina Inc., San Diego, CA, USA) resulting in 3,480,880 paired end reads for

C. ljungdahlii [pMTL82151] and 3,092,112 paired end reads for DSM 13528 with a size of

301 bp. Quality filtering using Trimmomatic version 0.36 (Bolger et al., 2014) resulted in

3,231,581 reads for the recombinant and 3,016,568 reads for the wild type strain,

respectively.

2.6.3 Mapping for single nucleotide polymorphism (SNP) analysis

Bowtie2 (Langmead and Salzberg, 2012) and SAMtools (Li, 2011) were used to map the

trimmed reads of C. ljungdahlii [pMTL82151] on the reference genome (CP001666.1). Variant

calling was performed using VarScan v2.3.9 (Koboldt et al., 2012) with the pileup2snp option.

Finally, changes in genome sequence and resulting changes in amino acid translation were

edited with an unpublished in-house tool of Göttingen Genomics Laboratory (Göttingen,

Germany).

2.6.4 Comparison of locus tags of different acetogens

The orthology detection tool “Proteinortho” (Lechner et al., 2011) version 4.26 (default

specification with following cut off settings: blast = blastp v2.2.24, E-value = 1-6, algorithm-

connection = 0.1, coverage = 0.3, percent_identity = 25, adaptive_similarity = 0.95,

incorporated_pairs = 1, incorporated_singles = 1, selfblast = 1, unambiguous = 0;) was used by

the Göttingen Genomics Laboratory (Göttingen, Germany) to find orthologous genes within

genome sequences of different acetogenic species.

Page 59: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 45

2.7 DNA transfer in bacteria

2.7.1 DNA transfer in Escherichia coli

2.7.1.1 Preparation of electrocompetent E. coli cells

The modified method of choice for preparing electrocompetent E. coli cells was adapted from

a published protocol (Dower et al., 1988). Respective E. coli strain was grown overnight in

5 ml LB medium (see chapter 2.3.1.4). This sample was used to inoculate 250 ml LB medium

in a baffled 2-l Erlenmeyer flask to an OD600nm of 0.1. Cells were incubated at 37 °C and

150 rpm to an OD600nm of 0.5-0.8. Afterwards, culture was transferred to a centrifugation

vessel and incubated on ice for 20 min. Cells were harvested at 4,000 x g for 10 min at 4 °C.

After two washing steps with 250 ml chilled and sterile water (4,000 x g, 10 min, 4°C), cells

were resuspended in 50 ml cold 10 % [v/v] glycerol. Thereupon, centrifugation at 4,000 x g for

10 min at 4 °C followed and culture was suspended in 30 ml cold 10 % [v/v] glycerol. After a

final centrifugation step at 4,000 x g for 10 min at 4 °C, cells were resuspended in 1 ml cold

10 % [v/v] glycerol and 50-µl aliquots were first flash-freezed in liquid N2 and then stored in

1.5-ml reaction tubes at -80 °C or directly used for electroporation procedure.

2.7.1.2 Preparation of fast competent E. coli cells

For preparing smaller amounts of electrocompetent E. coli cells, 5 ml of an overnight grown

culture were transferred into two 2-ml reaction tubes. All centrifugation steps were

performed at room temperature. After centrifugation at 5,040 x g for 1 min, cells were washed

twice with 1 ml chilled, sterile water (5,040 x g, 1 min). Subsequently, culture was resuspended

in cold 10 % glycerol [v/v] and after a final centrifugation at 5,040 x g for 3 min, suspended in

100 µl cold 10 % glycerol [v/v]. For electroporation 50-µl aliquots were used. Short-term

storage of fast competent cells was achieved at -80 °C.

2.7.1.3 Transformation of electrocompetent E. coli cells

According to the described protocol for transformation of electrocompetent or fast

competent E. coli cells (Dower et al., 1988), 50 µl of cells (see chapter 2.7.1.1 and 2.7.1.2) were

thawed on ice and mixed with approx. 500 ng of purified plasmid DNA (see chapter 2.5.4) or

dialyzed ligation (see chapter 2.5.4.3). The sample was transferred to pre-cooled

electroporation cuvettes with a gap width of 2 mm (Biozym Scientific GmbH, Hessisch

Oldendorf, Germany). The electric pulse (2500 V, 25 µF, and 200 Ω; retention time 4.6-5.0 ms)

was performed with “Gene Pulser XcellTM pulse generator” (Bio-Rad Laboratories GmbH,

Page 60: 2,3-Butanediol production using acetogenic bacteria

46 2. Material and Methods

Munich, Germany). Afterwards, 800 µl sterile LB medium (see chapter 2.3.1.4) were added to

the pulsed cells and transferred to a sterile 1.5-ml reaction tube. After incubation at 37 °C and

1350 rpm, the sample was centrifuged at 6,000 x g for 1 min at room temperature, 800 µl of

supernatant were discarded, cells were suspended in the remaining suspension, and plated

on selective LB agar plates (see chapter 2.3.1.4). Plates were incubated upside down at 37 °C

until colonies appeared.

2.7.1.4 Preparation of chemically competent E. coli cells

Chemically competent E. coli cells were prepared according to a previously described protocol

(Inoue et al., 1990). 5 ml of an overnight grown E. coli culture were used to inoculate 250 ml

SOB medium (see chapter 2.3.1.6) in a baffled 2-l Erlenmeyer flask to an OD600nm of 0.1. Cells

were incubated at 18 °C, until they reached an OD600nm of 0.6-0.8. Thereafter, the culture was

transferred to centrifugation vessels and incubated on ice for 30 min. After cells were

harvested at 3,900 x g for 10 min at 4 °C, they were suspended in 40 ml tris-borate (TB) buffer

and incubated on ice for another 10 min. Subsequently, cells were centrifuged again at

3,900 x g for 10 min at 4 °C, the sediment was suspended in 10 ml TB buffer and 1.5 ml of

sterile DMSO were added. Finally, 200-µl aliquots were flash-freezed in liquid N2 and stored

at -80 °C or used directly for transformation.

TB buffer

Solution I

PIPES 756 mg 10 mM

CaCl2 420 mg 15 mM

H2O ad 125 ml

The pH was adjusted to 6.7 with KOH.

Solution II

KCl 4.66 g 250.0 mM

MnCl2 x 4 H2O 1.72 g 034.7 mM

H2O ad 125 ml

Both solutions were autoclaved separately and combined afterwards.

Page 61: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 47

2.7.1.5 Transformation of chemically competent E. coli cells

In order to transform chemically competent E. coli cells, the following modified protocol was

used (Inoue et al., 1990). Cells were thawed on ice, then 20 µl of ligation mixture (see chapter

2.5.3.3), 5 µl of “In-Fusion® HD Cloning Kit” sample (see chapter 2.5.3.4), or 500 ng plasmid

DNA were added, and incubated on ice for 10 min. Heat shock was performed at 42 °C for

1 min. Afterwards, cells were incubated on ice for another 10 min, 800 µl pre-warmed (37 °C)

LB medium (see chapter 2.3.1.4) were added, and further incubated at 1350 rpm and 37 °C

for 1 h. Thereafter, cells were centrifuged at 4,000 x g for 1 min and 800 µl of supernatant

were discarded. Cell pellet was resuspended in the remaining LB medium and plated on

selective agar plates, which were incubated upside down at 37 °C, until colonies appeared.

2.7.2 DNA transfer in A. woodii and Clostridium by electroporation

2.7.2.1 Preparation of electrocompetent A. woodii and Clostridium cells

The protocol used for preparing electrocompetent A. woodii and Clostridium cells was slightly

modified from a previously described method (Leang et al., 2013). All plastic material

(e.g. syringes, 5 µg plasmid DNA in 1.5-ml reaction tubes, electroporation cuvettes) was placed

in an anaerobic chamber the day before transformation. An early stationary growth phase

culture of the respective strain was used to inoculate 100 ml modified ATCC 1612 medium

(see chapter 2.3.1.2) or Tanner medium (see chapter 2.3.1.7), which contained

DL-threonine (150 mM) and fructose (40 mM), to an OD600nm of 0.09. DL-Threonine is used to

permeabilize the cell wall, but the exact mechanism is not known (Zhang et al., 2011).

A. woodii and Clostridium were incubated over night at 30 °C and 37 °C, respectively, until they

reached an OD600nm of 0.3-0.7. Cells were harvested in 50-ml falcon tubes in an anaerobic

chamber by centrifugation at 6,000 x g for 10 min at 4 °C. Afterwards, cells were washed twice

with 30 ml chilled SMP buffer (see chapter 2.4.3.2) (6,000 x g, 10 min, 4°C) and suspended in

600 µl SMP buffer and 120 µl anti-freezing buffer (see chapter 2.4.3.2). Cells were used directly

for transformation or stored as 500-µl aliquots in cryogenic vials at -80 °C for further use.

2.7.2.2 Transformation of electrocompetent A. woodii and Clostridium

In an anaerobic chamber, electrocompetent cells were first transferred to a 1.5-ml reaction

tube containing 5 µg of plasmid DNA and incubated on ice for 5 min. Then, sample was placed

to a pre-cooled electroporation cuvette with a gap width of 1 mm (Biozym Scientific GmbH,

Hessisch Oldendorf, Germany) and electroporation was performed with “Gene Pulser XcellTM

Page 62: 2,3-Butanediol production using acetogenic bacteria

48 2. Material and Methods

pulse generator” (Bio-Rad Laboratories GmbH, Munich, Germany) at 0.625 kV, 25 µF, and

600 Ω with a retention time of 9-11 ms. Afterwards, 800 µl from a Hungate tube with

respective medium were transferred to the cuvette and the mixture was placed back into the

Hungate tube. This tube was incubated at 37 °C, until an increase in OD600nm was observed. In

case of Clostridium, 600 µl cells were plated in an anaerobic chamber on selective YTF agar

plates (see chapter 2.3.1.8) until colonies appeared. Single colonies were selected in the

anaerobic chamber and transferred to Hungate tubes using a syringe. In contrast, A. woodii

was mixed with 5 µl of the respective antibiotic directly in the Hungate tube. When cells were

growing in presence of antibiotic, they were inoculated into fresh medium containing the

respective antibiotic, genomic DNA was isolated (see chapter 2.5.3.1) and recombinant strain

was verified via PCR targeting 16S rDNA (fD1 and rP2) and vector backbone (see chapter 2.4.2),

respectively.

2.7.3 DNA transfer in Clostridium by conjugation

Conjugation of plasmid DNA in Clostridium was performed with a previously described, slightly

modified protocol (Mock et al., 2015). This protocol is based on methods for conjugation of

Clostridium acetobutylicum and Clostridium difficile (Purdy et al., 2002). As conjugation donor

strain, E. coli CA434 was used to transfer large plasmid DNA in Clostridium. By conjugation,

restriction modification (R-M) systems of bacteria can be partially eluded. At first, the plasmid

to be conjugated into Clostridium was transformed in fast competent E. coli CA434 cells (see

chapter 2.7.1.3). The resulting recombinant E. coli CA434 strain was inoculated the day before

conjugation into 5 ml LB medium (see chapter 2.3.1.4), containing tetracycline and the

respective antibiotic depending on the plasmid used, and was incubated overnight at

37 °C. In contrast, Clostridium was inoculated in Tanner medium (40 mM fructose) 9 h earlier

than E. coli CA434, since it grows more slowly. On the day of conjugation, 2 ml LB medium

containing respective antibiotics were inoculated with 100 µl of the overnight grown E. coli

CA434. When both organisms reached an OD600nm of about 0.5, both cultures were transferred

into an anaerobic chamber. E. coli CA434 was transferred to a sterile 2-ml reaction tube, and

centrifuged at 3,000 x g for 3 min at room temperature. Supernatant was discarded, cell pellet

was carefully washed with sterile phosphate-buffered saline (PBS) and centrifuged again as

mentioned above. Thereafter, cells were carefully mixed with 200 µl of Clostridium culture

and spotted onto YTF agar plates (see chapter 2.3.1.8) without any antiobiotic selection. After

Page 63: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 49

incubation of plates at 37 °C for 24 h without inverting, cells were harvested by flooding the

agar plates with 500 µl PBS. Before flooding the plates, a small piece of agar at the lower part

of the plate was removed by using a sterile inoculation loop. This way, the PBS/cell mixture

aggregated in the gap, remaining cells were scraped off with the same inoculation loop and

pipetted on selective YTF agar plates supplemented with 10 μg/ml colistin (to counterselect

the E. coli CA434 donor strain) and 7.5 μg/ml thiamphenicol (to select plasmid uptake). Plates

were incubated upside down for 3-4 days at 37 °C, until colonies appeared and conjugants

were re-streaked on a new selective YTF agar plate, incubated again under the same

conditions to further eliminate E. coli CA434 cells and to get single colonies which were

inoculated in 5 ml modified Tanner medium. When cells were growing in liquid medium,

genomic DNA was isolated (see chapter 2.5.1.3) and the recombinant strain was verified via

PCR of 16S rDNA (fD1 and rP2) (see chapter 2.4.2) and PCR of vector backbone (see chapter

2.4.2), respectively.

PBS

NaCl 8.0 g 0.14 M

Na2HPO4 x 2 H2O 1.4 g 7.86 mm

KCl 0.2 g 2.68 mm

KH2PO4 0.2 g 1.47 mM

H2O ad 1,000 ml

2.7.4 Northern blot

RNA of Clostridium ljungdahlii was investigated via Northern blot analysis. By using this

method, RNA transcripts were analyzed according to their size and amount (Alwine et al.,

1977).

2.7.4.1 Transfer of RNA onto a nylon membrane

Separation of 10 µg RNA was achieved in a denaturating agarose gel (see chapter 2.5.5.1)

according to its size and amount. Afterwards, RNA was transferred to a nitrocellulose

membrane using capillary blot. The setup of the Northern blot is depicted in Figure 4. One

piece of Whatman paper (20 x 12 cm; Whatman International Ltd., Maidstone, GB) was placed

on a blotting device (Wissenschaftliche Werkstatt Feinwerktechnik, Ulm University, Germany),

which was filled with 10 x saline-sodium citrate (SSC) buffer. The ends of this Whatman paper

were dipped in 10 x SSC buffer to ensure capillary force. Three more layers of Whatman paper

Page 64: 2,3-Butanediol production using acetogenic bacteria

50 2. Material and Methods

(14.5 x 20 cm), the agarose gel (see chapter 2.5.5.2), one piece of nitrocellulose membrane

(MACHEREY-NAGEL GmbH & Co. KG, Düren, Germany), one more Whatman paper, a paper

towel stack, and a weight (~500 g) were placed onto the block (Figure 4). The capillary force

originates from the paper towel stack and 10 x SSC buffer. SSC buffer migrates to the paper

towel stack and transfers negatively loaded RNA from agarose gel to the positively loaded

nitrocellulose membrane. Blotting was performed for 15 h at room temperature and

thereafter the nitrocellulose membrane was dried at 37 °C. Finally, RNA was covalently fixed

to the membrane by using “Ultraviolet Crosslinker” (GE Healthcare EUROPE GmbH, Munich,

Germany) at a wavelength of 254 nm for 90 s and 120,000 µJ/cm2.

10 x SSC buffer

NaCl 87.7 g 3.0 M

Trisodium citrate x H2O 44.1 g 0.3 M

H2O ad 1,000 ml

2.7.4.2 Hybridization of radioactively labelled probe with RNA

In order to hybridize the radioactively labelled probe with RNA, nitrocellulose membrane was

washed initially with RNase-free water and then overlaid with hybridization buffer.

Afterwards, the membrane was transferred without any bubbles and overlapping to a

hybridization tube and 20 ml hybridization buffer were added. Then, pre-hybridization was

performed in a “rotary oven OV2” (Biometra GmbH, Göttingen, Germany) at 65 °C for 1 h.

Denaturation of the radioactively labelled DNA probe was achieved by boiling at 95 °C for

5 min and afterwards it was added to the pre-hybridized membrane. Hybridization was

Figure 4: Scheme of Northern blot setup

Page 65: 2,3-Butanediol production using acetogenic bacteria

2. Material and Methods 51

finished after incubation for 18 h at 65 °C in the rotary oven. After discarding hybridization

buffer, the membrane was washed twice with 50 ml wash buffer I. At first, buffer was

discarded immediately after washing, while the second washing step was performed for

15 min at 65 °C in the rotary oven. Thereupon, two further washing steps with 50 ml wash

buffer II and wash buffer III followed under the same conditions as mentioned above. Before

transferring the nitrocellulose membrane to a film cassette, it was wrapped in cling film, a

phosphor imaging plate was placed on the wrapped membrane for 5 h and 24 h, and finally

analyzed by a “Fujifilm BAS-1000 Bio Imaging Analyzer IPR 1000” (Fuji Photo Film Co., Ltd.,

Tokyo, Japan) with the software “Image Reader V 1.2” (Fuji Photo Film Co., Ltd., Tokyo. Japan).

Phosphor imaging plates allow computed radiography and were always cleaned with an eraser

(raytest Isotopenmessgeräte GmbH, Straubenhardt, Germany) before and after use.

Hybridization buffer

5 x Hybridization buffer 10.0 ml 20 % [v/v]

50 x Denhardt solution 10.0 ml 20 % [v/v]

10 % SDS 02.5 ml 05 % [v/v]

H2O ad 50 ml

Incubation at 65 °C for 15 min, then addition of:

NaCl 02.9 g 990 mM

Herring sperm DNA 00.5 ml 010 mg/ml [w/v]

5 x Hybridization solution

Tris 3.03 g 250.0 mM

Sodium pyrophosphate 0.50 g 018.8 mM

H2O ad 100 ml

The pH was adjusted to 7.5 with HCl

50 x Denhardt solution

Bovine serum albumin

(BSA; albumin fraction V)

1 g 1 % [w/v]

Polyvinylpyrrolidone - K 30 1 g 1 % [w/v]

Ficoll 400 1 g 1 % [w/v]

H2O ad 100 ml

Page 66: 2,3-Butanediol production using acetogenic bacteria

52 2. Material and Methods

Wash buffer I

10 x SSC buffer 60 ml 20 % [v/v]

10 % SDS [w/v] 03 ml 01 % [v/v]

H2O ad 300 ml

Wash buffer II

10 x SSC buffer 10 ml 10 % [v/v]

10 % SDS [w/v] 01 ml 01 % [v/v]

H2O ad 100 ml

Wash buffer III

10 x SSC buffer 1 ml 1 % [v/v]

10 % SDS [w/v] 1 ml 1 % [v/v]

H2O ad 100 ml

Page 67: 2,3-Butanediol production using acetogenic bacteria

3. Results 53

3. Results

3.1 Improvement of 2,3-BD production in acetogenic bacteria

3.1.1 Natural 2,3-BD production in acetogenic bacteria

At first, a suitable host for improvement of natural 2,3-BD production using syngas as energy

and carbon source was characterized. It was already reported that C. autoethanogenum,

0 50 100 150 200 250 300 350 4000.1

1

Gro

wth

[O

D6

00

nm

]

Time [h]

0

20

40

60

80

100

120

140

160

Ace

tate

, Eth

ano

l [m

M]

0

1

2

3

4

5

2,3

-BD

[m

M]

0 50 100 150 200 250 3000.01

0.1

1

Gro

wth

[O

D6

00

nm

]

Time [h]

0

20

40

60

80

100

120

140

Ace

tate

, Eth

ano

l [m

M]

0

2

4

6

2,3

-BD

[m

M]

0 100 200 300 400 500 600 7000.01

0.1

1

Gro

wth

[O

D6

00

nm

]

Time [h]

0

10

20

30

40

50

60

Ace

tate

, Eth

ano

l [m

M]

0

1

2

3

4

2,3

-BD

[m

M]

0 50 100 150 200 250 3000.01

0.1

1

Gro

wth

[O

D6

00

nm

]

0

10

20

30

40

50

60

70

80

Time [h]A

ceta

te [

mM

]

0 50 100 150 200 250 300 3500.01

0.1

1

Gro

wth

[O

D6

00

nm

]

Time [h]

0

10

20

30

40

50

Ace

tate

, Eth

an

ol [

mM

]

0

1

2

3

Bu

tyra

te [

mM

]

A B

C D

E

Figure 5: Growth and production profile of C. autoethanogenum (A), C. ljungdahlii (B) in 50 ml Tanner

medium, C. ragsdalei (C) in 50 ml Rajagopalan medium, C. aceticum (D) in 50 ml modified 1.0 ATCC

1612 medium, and C. carboxidivorans (E) in 50 ml Tanner medium (3 g yeast extract), growing with

syngas (1.8 bar overpressure) as energy and carbon source. Growth ( ), growth control ( ),

acetate production ( ) , ethanol production ( ), butyrate production ( ), and 2,3-BD

production production ( ).

Page 68: 2,3-Butanediol production using acetogenic bacteria

54 3. Results

C. ragsdalei, and C. ljungdahlii naturally produce 1.4-2 mM 2,3-BD from steel mill waste gas

(Köpke et al., 2011b). In order to determine the best acetogenic natural 2,3-BD producer, the

above-mentioned organisms as well as C. aceticum and C. carboxidivorans were examined in

autotrophic growth experiments. All organisms were cultivated in 500-ml glass flasks

containing 50 ml of the respective medium using syngas as energy and carbon source

(overpressure 1.8 bar). All growth experiments were performed as biological duplicates.

Growth and production profile of C. autoethanogenum, C. ljungdahlii, C. ragsdalei,

C. aceticum, and C. carboxidivorans with syngas is depicted in Figure 5. The organisms

C. autoethanogenum, C. ljungdahlii, and C. ragsdalei produced mainly acetate

(52.6 mM-138.7 mM) and ethanol (21.8 mM-49.1 mM). Only small amounts of 2,3-BD were

detected with a maximal concentration of 3.7 mM, 4.8 mM, and 3.1 mM, respectively

(Figure 5A and B and C). In contrast, C. aceticum produced only acetate (71.4 mM) without

any by-products, and C. carboxidivorans produced acetate (42.6 mM) and ethanol (26.7 mM)

in addition to small amounts of butyrate (2.4 mM) (Figure 5D and E). This experiment

demonstrated that C. ljungdahlii showed the highest natural 2,3-BD production with a

concentration of 4.8 mM. For further genetic engineering experiments, C. ljungdahlii was the

organism of choice in order to enhance 2,3-BD production. Furthermore, there is an efficient

and reproducible transformation protocol available (Leang et al., 2013), which facilitates

genetic engineering of this organism. Moreover, the acetogen C. coskatii represents a further

interesting organism for enhancement of natural 2,3-BD overproduction showing 98.3 %

average nucleotide identity (ANI) compared to C. ljungdahlii (Bengelsdorf et al., 2016).

Analysis with “Proteinortho” detection tool revealed that C. coskatii harbors the homologous

genes alsS (CCOS_39110), budA (CCOS_38400), and bdh (CCOS_06940) for 2,3-BD production.

So, C. coskatii was also taken into account to overproduce 2,3-BD.

3.1.2 Construction of plasmids for enhancement of natural 2,3-BD production

For homologous overproduction of 2,3-BD, the three genes alsS (acetolactate synthase;

CLJU_c38920), budA (acetolactate decarboxylase; CLJU_c08380), and bdh (2,3-BD

dehydrogenase; CLJU_c23220) from C. ljungdahlii (Köpke et al., 2011b; 2014) were

synthesized in form of an operon (GenScript USA Inc., Piscataway, NJ, USA) under control of

promoter Ppta-ack from C. ljungdahlii. This native promoter is constitutive and originates from

pta-ack operon of C. ljungdahlii (Hoffmeister et al., 2016; Köpke and Mueller, 2016). The

Page 69: 2,3-Butanediol production using acetogenic bacteria

3. Results 55

enzymes acetolactate synthase (AlsS), acetolactate decarboxylase (BudA), and 2,3-BD

dehydrogenase (Bdh) promote 2,3-BD production starting from two molecules of pyruvate

(Figure 6). The synthetic 2,3-BD operon was received in form of the plasmid pUC57_23BD

(GenScript USA Inc., Piscataway, NJ, USA (Figure 7)). In order to clone the synthetic 2,3-BD

operon into vectors pMTL82151 and pJIR750, the respective plasmid and vectors were

digested using restriction enzymes SacI and SalI. Subsequently, ligation of the purified 2,3-BD

Figure 7: Cloning strategy for construction of plasmids pMTL82151_23BD and pJIR750_23BD. Plasmid pUC57_23BD, as well as vectors pMTL82151 and pJIR750 were digested using restriction enzymes SacI and SalI. 2,3-BD operon (Ppta-ack, alsS, budA, and bdh) was ligated with linearized pMTL82151 and pJIR750, resulting in plasmids pMTL82151_23BD and pJIR750_23BD. Ppta-ack, (promoter region upstream of pta-ack operon from C. ljungdahlii); alsS, acetolactate synthase from C. ljungdahlii; budA, acetolactate decarboxylase from C. ljungdahlii; bdh, 2,3-BD dehydrogenase from C. ljungdahlii; bla, ampicillin resistance gene; rep, ColE1, replicon for Gram-negative bacteria; repA/orf2, pBP1-replicon for Gram-positive bacteria from C. botulinum; rep, pMB1-replicon for Gram-negative bacteria; rep, pIP404-replicon for Gram-positive bacteria from C. perfringens; catP, chloramphenicol resistance gene; traJ, gene for conjugal transfer function.

Figure 6: 2,3-BD production in C. ljungdahlii and C. coskatii originating from pyruvate via enzymes AlsS (acetolactate synthase), BudA (acetolactate decarboxylase), Bdh (2,3-BD dehydrogenase), and LdhA (lactate dehydrogenase).

Page 70: 2,3-Butanediol production using acetogenic bacteria

56 3. Results

operon into the linearized vectors was performed. The resulting plasmids were named

pMTL82151_23BD and pJIR750_23BD (Figure 7). Successful cloning was confirmed by

restriction enzyme digestion of these plasmids using SacI and SalI as well as sequencing (M13-

FP, M13-RP, pMTL82151_23BDSeq1, pMTL82151_23BDSeq2, pMTL82151_23BDSeq3,

pMTL82151_23BDSeq4, pMTL82151_23BDSeq5) by GATC Biotech AG (Constance, Germany).

3.1.3 Homologous 2,3-BD overproduction in C. ljungdahlii

The empty vectors pMTL82151 and pJIR750, as well as the two different 2,3-BD

overproduction plasmids pMTL82151_23BD and pJIR750_23BD (Figure 7) were transformed

into C. ljungdahlii. The transformed strains growing in presence of thiamphenicol were verified

using two different methods. On the one hand, isolation of genomic DNA from transformed

C. ljungdahlii strains was performed and catP of pMTL82151 backbone (expected DNA

fragment 1,121 bp) or pJIR750 backbone (expected DNA fragment 1,075 bp) was amplified via

PCR (pMTL82151catP_fwd, pMTL82151catP_rev, pJIR750catP_fwd, pJIR750catP_rev)

(Figure 8A). On the other hand, isolated genomic DNA from recombinant C. ljungdahlii strains

was transformed into chemically competent E. coli XL-1 Blue MRF' cells and single colonies

growing on agar plates containing chloramphenicol were picked for plasmid DNA isolation.

1 M 2 3 4 5 7 6 M bp

5 4 3 2 M 1 bp

4 3 2 1 bp

M B CA

bp

Figure 8: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion of retransformed plasmids (B), and amplification of 16S rDNA gene of genomic DNA from recombinant C. ljungdahlii strains (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,121-bp (pMTL82151 backbone) and 1,075-bp (pJIR750 backbone) DNA fragments resulting from PCR of catP from C. ljungdahlii [pMTL82151] (1), C. ljungdahlii [pMTL82151_23BD] (2), C. ljungdahlii [pJIR750] (3), and C. ljungdahlii [pJIR750_23BD] (4). (B) Analytical restriction enzyme digestion using ApaI of retransformed pMTL82151 with expected DNA fragments 4,488 bp and 766 bp (1), using BamHI and Bsu15I of retransformed pMTL82151_23BD with expected DNA fragments 7,983 bp and 1,663 bp (2-4), using BamHI and SacI of retransformed pJIR750_23BD with expected DNA fragments 7,890 bp and 3,070 bp (5), using NdeI of retransformed pJIR750 with expected DNA fragments 4,773 bp and 1,825 bp (7), and undigested control (6). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of C. ljungdahlii WT as control (1), C. ljungdahlii [pMTL82151] (2), C. ljungdahlii [pMTL82151_23BD] (3), C. ljungdahlii [pJIR750] (4), and C. ljungdahlii

[pJIR750_23BD] (5).

--1,500

--1,000 1,500--- 2,000---

5,000--

1,000--

700--

1,500---

3,000----

6,000----- 10,000-----

1,000---

Page 71: 2,3-Butanediol production using acetogenic bacteria

3. Results 57

This procedure is called retransformation and plasmid DNA isolated from recombinant E. coli

XL-1 Blue MRF' is called retransformed plasmid. Figure 8B and C show the analytical restriction

enzyme digestions using ApaI of retransformed pMTL82151 (expected DNA fragments

4,488 bp, 766 bp), NdeI of retransformed pJIR750 (expected DNA fragments 4,773 bp,

1,825 bp), BamHI and Bsu15I of retransformed pMTL82151_23BD (expected DNA fragments

7,983 bp, 1,663 bp), and BamHI and SacI of retransformed pJIR750_23BD (expected DNA

fragments 7,890 bp, 3,070 bp). Additionally, amplification of 16S rDNA gene (1,500 bp; fD1,

rP2) and concomitant sequencing of the resulting DNA fragment by GATC Biotech AG

(Constance, Germany) supported correct identity of the recombinant C. ljungdahlii strains

(Figure 8D). Using these methods, recombinant strains C. ljungdahlii [pMTL82151],

C. ljungdahlii [pJIR750], C. ljungdahlii [pMTL82151_23BD], and C. ljungdahlii [pJIR750_23BD]

were verified. The recombinant C. ljungdahlii strains were analyzed under both, heterotrophic

and autotrophic growth conditions.

Figure 9 presents growth, fructose consumption, and production pattern of C. ljungdahlii

wildtype (WT), C. ljungdahlii [pMTL82151], C. ljungdahlii [pJIR750], C. ljungdahlii

[pMTL82151_23BD] and C. ljungdahlii [pJIR750_23BD] under heterotrophic conditions.

Growth experiments were performed as biological triplicates in 125-ml glass flasks containing

50 ml Tanner medium with 40 mM fructose as energy and carbon source. All strains showed

a similar growth behaviour. C. ljungdahlii WT and C. ljungdahlii [pMTL82151] reached a max.

OD600nm of 2.2 after 72 h while the max. OD600nm of C. ljungdahlii [pJIR750] and C. ljungdahlii

[pJIR750_23BD] was 2.0 after 72 h. The strain C. ljungdahlii [pMTL82151] displayed the lowest

max. OD600nm of 1.7 after 48 h. In the medium of C. ljungdahlii WT and C. ljungdahlii

[pJIR750_23BD] fructose concentration in the beginning was a little lower with 33.0 mM and

30.2 mM, respectively. Furthermore, the highest acetate concentration was detected for

C. ljungdahlii WT with 102.0 mM, while for the other strains acetate concentration reached

values between 87.9 mM and 96.8 mM. Regarding ethanol production, C. ljungdahlii

[pMTL82151] showed a slightly higher ethanol production with 11.2 mM compared to the

other strains ranging from 4.9 mM to 8.3 mM. In contrast, C. ljungdahlii [pJIR750_23BD] and

C. ljungdahlii [pMTL82151_23BD] showed slightly higher 2,3-BD concentrations with 1.2 mM

and 1.1 mM, respectively, in comparison to the strains with the empty vectors and the WT

strain (0.8 mM-1.0 mM). This shows that C. ljungdahlii [pJIR750_23BD] and C. ljungdahlii

Page 72: 2,3-Butanediol production using acetogenic bacteria

58 3. Results

[pMTL82151_23BD] produced slightly higher amounts of 2,3-BD with fructose as carbon

source compared to C. ljungdahlii WT, C. ljungdahlii [pMTL82151], and C. ljungdahlii [pJIR750].

Afterwards, the 2,3-BD production with syngas was examined. The growth experiment was

performed in 1-l glass flasks containing 100 ml Tanner medium and syngas as energy and

carbon source with 1.8 bar overpressure. Figure 10 shows growth and production pattern of

C. ljungdahlii WT, C. ljungdahlii [pMTL82151], C. ljungdahlii [pJIR750], C. ljungdahlii

[pMTL82151_23BD], and C. ljungdahlii [pJIR750_23BD] under autotrophic growth conditions.

Regarding growth, it is noticeable that C. ljungdahlii [pJIR750] showed a more slowly and

decreased growth with a max. OD600nm of 0.7 after 1,344 h in comparison to the other strains.

0

2

4

6

8

10

12

0.01

0.1

1

Gro

wth

[O

D6

00

nm

]

0

10

20

30

40

50

Fru

cto

se [

mM

]

0

20

40

60

80

100

120

Ace

tate

[m

M]

Eth

ano

l [m

M]

0 25 50 75 100 125 150 175 200 2250.00.20.40.60.81.01.21.4

Time [h]

2,3

-BD

[m

M]

Figure 9: Growth (solid line), fructose consumption (dashed line), and production pattern (solid lines) of C. ljungdahlii WT ( ), C. ljungdahlii [pMTL82151] ( ), C. ljungdahlii [pJIR750] ( ), C. ljungdahlii [pMTL82151_23BD] ( ), and C. ljungdahlii [pJIR750_23BD] ( ) growing in 50 ml Tanner medium containing 40 mM fructose as carbon source.

Page 73: 2,3-Butanediol production using acetogenic bacteria

3. Results 59

The highest OD600nm of 1.8 after 1,008 h was detected for C. ljungdahlii WT compared to

C. ljungdahlii [pMTL82151] with 1.5 after 840 h, C. ljungdahlii [pJIR750_23BD] with 1.4 after

432 h, and C. ljungdahlii [pMTL82151_23BD] with 1.2 after 840 h. With respect to acetate

production, all recombinant C. ljungdahlii strains showed decreased acetate concentrations

(260.4 mM-241.6 mM) compared to C. ljungdahlii WT (287.8 mM). The stain C. ljungdahlii

[pJIR750_23BD] produced least acetate with a concentration of 219.6 mM. In contrast, all

recombinant C. ljungdahlii strains except of C. ljungdahlii [pJIR750] (60.4 mM) showed an

increased ethanol production compared to C. ljungdahlii WT (71.9 mM) with concentrations

of 82.0 mM to 83.3 mM. The highest ethanol concentration of 92.8 mM showed

C. ljungdahlii [pMTL82151]. Regarding 2,3-BD production, only C. ljungdahlii [pJIR750_23BD]

Figure 10: Growth and production pattern of C. ljungdahlii WT ( ), C. ljungdahlii [pMTL82151] ( ),C. ljungdahlii [pJIR750] ( ), C. ljungdahlii [pMTL82151_23BD] ( ), and C. ljungdahlii

[pJIR750_23BD] ( ) growing in 100 ml Tanner medium with syngas (1.8 bar overpressure).

0

20

40

60

80

100

120

0.01

0.1

1

Gro

wth

[O

D6

00

nm

]

050

100150200250300350

Ace

tate

[m

M]

Eth

an

ol [

mM

]

0 500 1000 1500 2000 25000

2

4

6

8

Time [h]

2,3

-BD

[m

M]

Page 74: 2,3-Butanediol production using acetogenic bacteria

60 3. Results

produced more 2,3-BD (6.4 mM) than C. ljungdahlii WT (4.4 mM). In contrast, C. ljungdahlii

[pMTL82151_23BD] did not produce more 2,3-BD (3.8 mM) than C. ljungdahlii WT.

In summary, heterotrophic (Figure 9) as well as autotrophic (Figure 10) growth experiments

with C. ljungdahlii WT and the recombinant C. ljungdahlii strains revealed that C. ljungdahlii

[pJIR750_23BD] produced higher amounts of 2,3-BD compared to C. ljungdahlii WT. In

contrast, C. ljungdahlii [pMTL82151_23BD] did only produce higher amounts of 2,3-BD

compared to C. ljungdahlii WT under heterotrophic growth conditions.

3.1.4 Homologous 2,3-BD overproduction in C. coskatii

For investigation of 2,3-BD overproduction in C. coskatii, the vectors pMTL82151 and pJIR750

as well as the two 2,3-BD production plasmids pMTL82151_23BD and pJIR750_23BD

(Figure 7) were transformed into C. coskatii. The respective recombinant C. coskatii strains

were verified either by isolation of genomic DNA and amplification (pMTL82151catP_fwd,

pMTL82151catP_rev, pJIR750catP_fwd, pJIR750catP_rev) of catP originating from

pMTL82151 backbone (expected DNA fragment 1,121 bp; Figure 11A) and pJIR750 backbone

(expected DNA fragment 1,075 bp; Figure 11A), respectively or via retransformation of

genomic DNA from recombinant strains into E. coli XL1-Blue MRF' and subsequent plasmid

isolation and analytical restriction enzyme digestion. Figure 11B depicts the digestion of

Figure 11: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion (B), and amplification of 16S rDNA gene of genomic DNA from recombinant C. coskatii strains (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,121-bp (pMTL82151 backbone) and 1,075-bp (pJIR750 backbone) DNA fragments resulting from PCR of catP

from C. coskatii [pMTL82151] (3), C. coskatii [pMTL82151_23BD] (4), C. coskatii [pJIR750] (7), C. coskatii [pJIR750_23BD] (8), positive control with empty vector pMTL82151 (2) and pJIR750 (6) as template, and negative control with sterile water as template (1, 5). (B) Analytical restriction enzyme digestion using SalI and SacI of retransformed pJIR750_23BD with expected DNA fragments 6,545 bp and 4,415 bp (1), using NdeI of retransformed pJIR750 with expected DNA fragments 4,773 bp and 1,825 bp (2), using EcoRI of retransformed pMTL82151 with expected DNA fragment 5,254 bp (3) and of retransformed pMTL82151_23BD with expected DNA fragment 5,284 bp, 4,338 bp, and 22 bp (4). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of C. coskatii

[pJIR750_23BD] (3), C. coskatii [pMTL82151] (6), C. coskatii [pMTL82151_23BD] (7), C. coskatii

[pJIR750] (4), and C. coskatii [pJIR750] (8), negative control with sterile water as template (1, 4), and positive control with genomic DNA from C. coskatii WT (2, 5).

Page 75: 2,3-Butanediol production using acetogenic bacteria

3. Results 61

retransformed pJIR750_23BD using SalI and SacI (expected DNA fragments 6,545 bp,

4,415 bp), retransformed pJIR750 using NdeI (expected DNA fragments 4,773 bp, 1,825 bp),

retransformed pMTL82151 using EcoRI (expected DNA fragment 5,254 bp), and retransformed

pMTL82151_23BD using EcoRI (expected DNA fragments 5,284 bp, 4,338 bp, 22 bp).

Furthermore, amplification of 16S rDNA gene (1,500 bp; fD1, rP2), sequencing of DNA

fragment by GATC Biotech AG (Constance, Germany) with subsequent BLAST analysis

supported correct identity of the recombinant strains (Figure 11C). This way, recombinant

strains C. coskatii [pMTL82151], C. coskatii [pJIR750], C. coskatii [pMTL82151_23BD], and

C. coskatii [pJIR750_23BD] were verified and subsequently growth and production patterns

were investigated under heterotrophic, as well as autotrophic growth conditions.

Heterotrophic growth experiments were performed as biological triplicates in 125-ml glass

flasks containing 50 ml Tanner medium with 30 mM fructose as energy and carbon source.

Figure 12 presents growth, fructose consumption, and production pattern of C. coskatii WT,

C. coskatii [pMTL82151], C. coskatii [pJIR750], C. coskatii [pMTL82151_23BD], and C. coskatii

[pJIR750_23BD]. In terms of growth behaviour, no significant difference between C. coskatii

WT and the recombinant strains occurred. Although this heterotrophic growth experiment

was conducted with less fructose (30 mM) compared to C. ljungdahlii (Figure 9), none of the

C. coskatii strains were able to consume fructose completely. The C. coskatii strains reached

max. OD600nm values of 2.2-1.9 after 219 h. Furthermore, C. coskatii [pJIR750_23BD] and

C. coskatii [pMTL82151] produced more acetate (92.0 mM and 78.0 mM) than C. coskatii WT

(67.8 mM). Moreover, it is noticeable that C. coskatii [pMTL82151_23BD] produced more

ethanol (15.7 mM) compared to the other tested strains (6.1 mM-7.7 mM), but it did not

produce more 2,3-BD (0.3 mM) than C. coskatii WT (1.1 mM). Also, C. coskatii [pJIR750_23BD]

produced a similar 2,3-BD concentration (1.0 mM) compared to C. coskatii WT.

All C. coskatii strains were also investigated under autotrophic growth conditions with syngas

(Figure 13). The growth experiment was performed in 500-ml glass flasks containing 50 ml

Tanner medium and syngas as energy and carbon source with 1.0 bar overpressure.

Overpressure of syngas in anaerobic glass flasks was reduced from 1.8 bar to 1 bar, since the

problem of bursting glass flasks occurred, when overpressure was too high and moreover

500-ml glass flasks were more stable than 1-l glass flasks. Regarding growth of the different

Page 76: 2,3-Butanediol production using acetogenic bacteria

62 3. Results

C. coskatii strains under autotrophic growth conditions (Figure 13), it was striking that all

strains grew more slowly and to decreased OD600nm values of 0.8. The strain C. coskatii

[pJIR750] showed lowest growth (OD600nm 0.5) compared to C. ljungdahlii (OD600nm 1.8;

Figure 10). In terms of acetate production, all C. coskatii strains showed very similar acetate

concentrations (113.7 mM-117.3 mM) with C. coskatii [pJIR750] producing least acetate

(101.4 mM), which can be ascribed to decreased growth. According to the results of the

growth experiment using fructose as carbon source (Figure 12), C. coskatii [pMTL82151_23BD]

produced more ethanol (7.6 mM) than the other C. coskatii strains (2.3-3.0 mM) and regarding

2,3-BD concentration, it is noticeable that C. coskatii WT produced less 2,3-BD than

Figure 12: Growth (solid line), fructose consumption (dashed line), and production pattern (solid lines) of C. coskatii WT ( ), C. coskatii [pMTL82151] ( ), C. coskatii [pJIR750] ( ), C. coskatii [pMTL82151_23BD]( ), and C. coskatii [pJIR750_23BD] ( ) growing in 50 ml Tanner medium containing 30 mM fructose.

0

5

10

15

20

0.01

0.1

1G

row

th [

OD

60

0n

m]

0

10

20

30

Fru

cto

se [

mM

]

0

20

40

60

80

100

Ace

tate

[m

M]

Eth

ano

l [m

M]

0 25 50 75 100 125 150 175 200 2250.0

0.2

0.4

0.6

0.8

1.0

1.2

Time [h]

2,3

-BD

[m

M]

Page 77: 2,3-Butanediol production using acetogenic bacteria

3. Results 63

C. ljungdahlii WT. Moreover, the recombinant C. coskatii strains did not show enhancement

in 2,3-BD production compared to C. coskatii WT.

3.1.5 Modifications of 2,3-BD production plasmid pJIR750_23BD

An optimization of the 2,3-BD production plasmid pJIR750_23BD was obligatory to further

enhance the homologous 2,3-BD production in C. ljungdahlii. In autotrophic growth

experiments with syngas, C. ljungdahlii [pJIR750_23BD] produced higher amounts of 2,3-BD

(6.4 mM) than C. ljungdahlii [pMTL82151_23BD] (3.8 mM) (Figure 10). For this reason, further

modification steps of 2,3-BD plasmids were only performed with pJIR750_23BD.

Figure 13: Growth and production pattern of C. coskatii WT ( ), C. coskatii [pMTL82151] ( ), C. coskatii [pJIR750] ( ), C. coskatii [pMTL82151_23BD] ( ), and C. coskatii [pJIR750_23BD]( ) growing in 50 ml Tanner medium with syngas (1.0 bar overpressure).

0

2

4

6

8

0.01

0.1

1

Gro

wth

[O

D6

00

nm

]

0

25

50

75

100

125

Ace

tate

[m

M]

Eth

ano

l [m

M]

0 250 500 750 1000 1250 1500 1750 2000 22500.0

0.5

1.0

1.5

2.0

Time [h]

2,3

-BD

[m

M]

Page 78: 2,3-Butanediol production using acetogenic bacteria

64 3. Results

3.1.5.1 Removal of a potential terminator structure

The DNA analytical program “Clone Manager 7.11“ (Scientific & Educational Software, Cary,

NC, USA) was applied to analyze the DNA sequence of the synthetic 2,3-BD operon. Due to

this analysis, a hairpin loop structure was identified downstream of budA gene (Figure 14).

Although a poly-A tail is missing, it could not be excluded that transcription of bdh from

plasmid pJIR750_23BD is aborted at this terminator structure. For removal of the potential

terminator structure downstream of budA, a shortened sequence of budA was amplified

(CljbudAKpnI_fwd, Clj-budABamHI_rev) using genomic DNA from C. ljungdahlii WT. The

amplified DNA fragment, as well as pJIR750_23BD were digested with restriction enzymes

KpnI and BamHI and afterwards ligated. Successful cloning was confirmed by analytical

restriction digestion with KpnI and SalI, as well as by sequencing (pMTL82151_23BDSeq3,

pMTL82151_23BDSeq4) by GATC Biotech AG (Constance, Germany). The resulting plasmid

was named pJIR750_23BD_budAshort and has a size of 10,867 bp, which is 93 bp shorter than

pJIR750_23BD with 10,960 bp (Figure 15). The modified plasmid pJIR750_23BD_budAshort

was transformed into C. ljungdahlii WT. Transformation was only successful after in vivo

methylation using E. coli ER2275 [pACYC184_MCljI] and isolation of plasmid DNA using

“QIAGEN Plasmid Midi Kit” (QIAGEN-tip 100; Qiagen GmbH, Hilden, Germany). Afterwards,the

transformed strain growing in presence of thiamphenicol was verified by PCR of 16S rDNA

gene (fD1, rP2; expected DNA fragment 1,500 bp; Figure 16C) with genomic DNA as template

followed by sequencing (fD1, rP2) by GATC Biotech AG (Constance, Germany). The presence

Figure 14: 2,3-BD operon containing Ppta-ack (promoter region upstream of pta-ack operon from C. ljungdahlii), alsS (acetolactate synthase from C. ljungdahlii), budA (acetolactate decarboxylase from C. ljungdahlii), bdh (2,3-BD dehydrogenase from C. ljungdahlii), and Tbdh (terminator of bdh from C. ljungdahlii) of plasmid pJIR750_23BD showing hairpin loop structure after budA gene with respective sequence and energy content.

[bp]

free energy: -9.3 kcal/mol

Tbdh

Page 79: 2,3-Butanediol production using acetogenic bacteria

3. Results 65

of plasmid was confirmed by PCR of catP (pJIR750catP_fwd, pJIR750catP_rev; expected DNA

fragment 1,075 bp; Figure 16A) with genomic DNA as template as well as analytical restriction

enzyme digestion with KpnI and BamHI (expected DNA fragments 10,120 bp, 747 bp;

Figure 16B) after retransformation of genomic DNA from the recombinant C. ljungdahlii strain

into E. coli XL1-Blue MRF' and subsequent plasmid isolation.

Figure 15: Cloning strategy for construction of pJIR750_23BD_budAshort. Plasmid pJIR750_23BD as well as DNA fragment budA (short) were digested with restriction enzymes KpnI and BamHI. DNA fragment budAshort was ligated with linearized pJIR750_23BD, resulting in plasmid pJIR750_23BD_budAshort. Ppta-ack, promoter region upstream of pta-ack operon from C. ljungdahlii; alsS, acetolactate synthase from C. ljungdahlii; budA, acetolactate decarboxylase from C. ljungdahlii; bdh, 2,3-BD dehydrogenase from C. ljungdahlii; rep, pMB1-replicon for Gram-negative bacteria; rep, pIP404-replicon for Gram-positive bacteria from C. perfringens; catP, chloramphenicol resistance gene.

Figure 16: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion (B), and amplification of 16S rDNA gene of genomic DNA from recombinant C. ljungdahlii strain (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,075-bp DNA fragments resulting from PCR of catP from C. ljungdahlii [pJIR750_23BD_budAshort] (3), positive control with empty vector pJIR750 as template (2), and negative control with sterile water as template (1). (B) Analytical restriction enzyme digestion using KpnI and BamHI of retransformed pJIR750_23BD_budAshort with expected DNA fragments 10,120 bp and 747 bp (1). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of C. ljungdahlii [pJIR750_23BD_budAshort] (2), and water as template for negative control (1).

Page 80: 2,3-Butanediol production using acetogenic bacteria

66 3. Results

3.1.5.2 Exchange of bdh with adh (primary-secondary alcohol dehydrogenase)

As a second approach to improve the pJIR750_23BD production plasmid, the gene bdh, which

encodes 2,3-BD dehydrogenase, was exchanged with the adh gene (primary-secondary

alcohol dehydrogenase). Recently, the primary-secondary alcohol dehydrogenase (sAdh) from

C. autoethanogenum (CAETHG_0553) was identified, which is able to convert acetoin to 2,3-

BD and also acetone to 1,3-propanediol (Köpke et al., 2014). The identical adh gene

(CLJU_c24860) is also present in the genome of C. ljungdahlii. Instead of using NADH as

reducing equivalent for reduction of acetoin to 2,3-BD, sAdh requires NADPH (Figure 17). To

test whether the primary-secondary alcohol dehydrogenase shows a higher performance for

2,3-BD production in C. ljungdahlii, plasmid pJIR750_23BD_budAshort_adh was constructed

(Figure 18). At first, adh was amplified (CljADHBamHI_fwd, CljADHSalI_rev) using genomic

DNA from C. ljungdahlii WT. Subsequently, purified adh gene and plasmid

pJIR750_23BD_budAshort were linearized using restriction enzymes BamHI and SalI and

ligated with DNA fragment adh. Confirmation of successful cloning was achieved by analytical

restriction enzyme digestion using BamHI and SalI and sequencing (M13-FP) by GATC Biotech

AG (Constance, Germany). The resulting plasmid was named pJIR750_23BD_budAshort_adh

(Figure 18). It was also necessary to methylate the plasmid in vivo, using E. coli ER2275

[pACYC184_MCljI] prior to transformation into C. ljungdahlii and isolation of plasmid DNA was

performed with “QIAGEN Plasmid Midi Kit” (QIAGEN-tip 100; Qiagen GmbH, Hilden, Germany)

to receive high plasmid DNA concentrations. Verification of the C. ljungdahlii

Figure 17: 2,3-BD production in C. ljungdahlii originating from pyruvate via enzymes AlsS (acetolactate synthase), BudA (acetolactate decarboxylase), and sAdh (primary-secondary alcohol dehydrogenase).

Page 81: 2,3-Butanediol production using acetogenic bacteria

3. Results 67

[pJIR750_23BD_budAshort_adh] strain was done by PCR of catP (pJIR750catP_fwd,

pJIR750catP_rev; expected DNA fragment 1,075 bp; Figure 19A) using genomic DNA as

template as well as analytical restriction enzyme digestion using BamHI and SalI (expected

DNA fragments 9,615 bp, 1,183 bp; Figure 19B) after retransformation of genomic DNA from

recombinant C. ljungdahlii strain into E. coli XL1-Blue MRF' and subsequent plasmid isolation.

Figure 18: Cloning strategy for construction of pJIR750_23BD_budAshort_adh. Plasmid pJIR750_23BD_budAshort, as well as DNA fragment adh were digested with restriction enzymes BamHI and SalI. DNA fragment adh was ligated with linearized pJIR750_23BD_budAshort, resulting in plasmid pJIR750_23BD_budAshort_adh. Ppta-ack, promoter region upstream of pta-ack operon from C. ljungdahlii; alsS, acetolactate synthase from C. ljungdahlii; budA, acetolactate decarboxylase from C. ljungdahlii; bdh, 2,3-BD dehydrogenase from C. ljungdahlii; adh, primary-secondary alcohol dehydrogenase from C. ljungdahlii; rep, pMB1-replicon for Gram-negative bacteria; rep, pIP404-replicon for Gram-positive bacteria from C. perfringens; catP, chloramphenicol resistance gene.

Figure 19: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion (B), and amplification of 16S rDNA gene of genomic DNA from recombinant C. ljungdahlii strain (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,075-bp DNA fragments resulting from PCR of catP from C. ljungdahlii [pJIR750_23BD_budAshort_adh] (3), positive control with empty vector pJIR750 as template (2), and negative control with sterile water as template (1). (B) Analytical restriction enzyme digestion using BamHI and SalI of retransformed pJIR750_23BD_budAshort_adh with expected DNA fragments 9,615 bp and 1,183 bp (1). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of C. ljungdahlii

[pJIR750_23BD_budAshort_adh] (2), and water as template for negative control (1).

Page 82: 2,3-Butanediol production using acetogenic bacteria

68 3. Results

Furthermore, 16S rDNA gene was amplified (fD1, rP2; expected DNA fragment 1,500 bp;

Figure 19C) using genomic DNA from C. ljungdahlii [pJIR750_23BD_budAshort_adh] as

template and sequenced (fD1, rP2) by GATC Biotech AG (Constance, Germany) to verify the

recombinant strain.

3.1.5.3 BudABC operon from Raoultella terrigena

As a third possibility for modification of pJIR750_23BD to enhance 2,3-BD production, the 2,3-

BD operon budABC from Raoultella terrigena (formerly known as Klebsiella terrigena;

Blomqvist et al., 1993) was cloned into vector pJIR750 under control of a promoter from

C. ljungdahlii (Ppta-ack) and a terminator from C. acetobutylicum (Tsol-adc). Tsol-adc is a strong rho-

independent terminator, which is functional in sense as well as antisense direction. Figure 20

shows the 2,3-BD production pathway of R. terrigena, in which acetolactate decarboxylase

(BudA), which converts S-acetolactate to acetoin, is encoded by budA, and the gene budB

encodes acetolactate synthase (AlsS), which metabolizes pyruvate to S-acetolactate. The final

reduction of R-acetoin or S-acetoin to meso-2,3-BD and 2S,3S-BD, respectively, is catalyzed by

acetoin reductase (Acr), encoded by budC, which can also operate as 2,3-BD dehydrogenase

and oxidizes 2,3-BD to acetoin with concomitant reduction of NAD+. In order to test

heterologous 2,3-BD production with budABC operon from R. terrigena in C. ljungdahlii,

plasmid pJIR750_budABCoperon was constructed. Firstly, plasmid pMK_budABCoperon was

digested with EcoRI and BamHI to obtain the budABC operon. The operon was ligated with

linearized pJIR750 resulting in plasmid pJIR750_budABC (Figure 21). Afterwards, Ppta-ack was

Figure 20: 2,3-BD production in R. terrigena originating from pyruvate via enzymes AlsS (acetolactate synthase), BudA (acetolactate decarboxylase), and Acr (acetoin reductase).

Page 83: 2,3-Butanediol production using acetogenic bacteria

3. Results 69

amplified (CPECPpta-ackEcoRI_fwd, CPECPpta-ackEcoRI_rev) and cloned into linearized (with

EcoRI digested) plasmid pJIR750_budABC using “In-Fusion HD Cloning” resulting in plasmid

pJIR750_ budABC_Ppta-ack (Figure 21). To finish plasmid construction, Tsol-adc was amplified

(TsoladcXbaI_fwd, TsoladcHindIII_rev), DNA fragment was digested with restriction enzymes

XbaI and HindIII, and ligated with linearized pJIR750_budABC_Ppta-ack. The resulting plasmid

was named pJIR750_budABCoperon (Figure 21). Successful cloning of pJIR750_budABCoperon

was confirmed by analytical restriction enzyme digestion using NdeI (expected fragments

5,425 bp, 3,092 bp, 1,825 bp) and sequencing (M13-RP, budABC_SEQ, BudABC_Seq1,

BudABC_Seq2, BudABC_Seq3, BudABC_Seq4, M13_FP) by GATC Biotech AG (Constance,

Germany). Plasmid pJIR750_budABCoperon was transformed into C. ljungdahlii, but again in

vivo methylation and plasmid isolation with “QIAGEN Plasmid Midi Kit” (QIAGEN-tip 100;

Qiagen GmbH, Hilden, Germany) was mandatory prior to the transformation process.

Recombinant strain C. ljungdahlii [pJIR750_budABCoperon] was verified by both PCR of catP

(pJIR750catP_fwd, pJIR750catP_rev; expected DNA fragment 1,075 bp; Figure 22A) with

genomic DNA as template and analytical restriction enzyme digestion using EcoRI (expected

Figure 21: Cloning strategy for construction of pJIR750_budABCoperon. Vector pJIR750 as well as DNA fragment budABC were digested with restriction enzymes EcoRI and BamHI. DNA fragment budABC was ligated with linearized pJIR750, resulting in plasmid pJIR750_budABC. Ppta-ack was cloned into linearized plasmid pJIR750_budABC via “In-Fusion HD cloning”, resulting in plasmid pJIR750_budABC_Ppta-ack. Tsol-adc was digested with XbaI and HindIII and ligated with linearized plasmid pJIR750_budABC_Ppta-ack, resulting in plasmid pJIR750_budABCoperon. rep (pMB1), pMB1-replicon for Gram-negative bacteria; rep (pIP404), pIP404-replicon for Gram-positive bacteria from C. perfringens; catP, chloramphenicol resistance gene; lacZ, β-galactosidase gene of lac-operon from E. coli; budABC, budABC operon from R. terrigena; Ppta-ack, promoter region upstream of pta-

ack operon from C. ljungdahlii; Tsol-adc, terminator region downstream of sol-adc operon from C. acetobutylicum.

Page 84: 2,3-Butanediol production using acetogenic bacteria

70 3. Results

DNA fragments 9,960 bp, 382 bp; Figure 22B) after retransformation of genomic DNA into

E. coli XL1-Blue MRF' and subsequent plasmid isolation. Afterwards, identity of recombinant

C. ljungdahlii strain growing in the presence of thiamphenicol was verified by PCR of 16S rDNA

gene (fD1, rP2; expected DNA fragment 1,500 bp; Figure 22C) with genomic DNA as template

followed by sequencing (fD1, rP2) by GATC Biotech AG (Constance, Germany).

3.1.5.4 Growth experiments of different C. ljungdahlii strains harboring the modified 2,3-BD

plasmids

The recombinant strains C. ljungdahlii [pJIR750_23BD_budAshort], C. ljungdahlii

[pJIR750_23BD_budAshort_adh], and C. ljungdahlii [pJIR750_budABCoperon] were

investigated under both heterotrophic and autotrophic growth conditions. Heterotrophic

growth experiments were performed as biological triplicates in 125-ml glass flasks containing

50 ml Tanner medium as well as 40 mM fructose as energy and carbon source.

In Figure 23, growth, pH, fructose consumption, and production pattern of C. ljungdahlii WT,

C. ljungdahlii [pMTL82151], C. ljungdahlii [pJIR750], C. ljungdahlii [pMTL82151_23BD],

C. ljungdahlii [pJIR750_23BD], C. ljungdahlii [pJIR750_23BD_budAshort], C. ljungdahlii

[pJIR750_23BD_budAshort_adh], and C. ljungdahlii [pJIR750_budABCoperon] under

heterotrophic conditions is presented. In terms of growth behaviour, it was noticeable that

C. ljungdahlii [pJIR750_23BD], C. ljungdahlii [pMTL82151_23BD], as well as C. ljungdahlii

Figure 22: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion (B), and amplification of 16S rDNA gene of genomic DNA from recombinant C. ljungdahlii strain (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,075-bp DNA fragments resulting from PCR of catP from C. ljungdahlii [pJIR750_budABCoperon] (2), positive control with empty vector pJIR750 as template (1). (B) Analytical restriction enzyme digestion using EcoRI of retransformed pJIR750_budABCoperon with expected DNA fragments 9,960 bp and 382 bp (1). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of C. ljungdahlii [pJIR750_budABCoperon] (2), and water as template for negative control (1).

Page 85: 2,3-Butanediol production using acetogenic bacteria

3. Results 71

[pJIR750] showed decreased max. OD600nm values compared to the other strains of 0.9, 1.7,

and 1.6, respectively (Figure 23). Moreover, C. ljungdahlii [pMTL82151_23BD] and

C. ljungdahlii [pJIR750] did not consume fructose completely. All other strains reached max.

OD600nm values above 2. The pH decreased concomitantly with increasing acetate

concentrations, but it is striking that C. ljungdahlii [pMTL82151] produced with 52.1 mM much

less acetate than the other C. ljungdahlii strains (96.1 mM to 110.4 mM). Due to less acetate

Figure 23: Growth, pH, fructose consumption, and production pattern of C. ljungdahlii WT ( ), C. ljungdahlii [pMTL82151] ( ), C. ljungdahlii [pJIR750] ( ), C. ljungdahlii [pMTL82151_23BD] ( ), C. ljungdahlii [pJIR750_23BD] ( ), C. ljungdahlii [pJIR750_23BD_budAshort] ( ),C. ljungdahlii [pJIR750_23BD_budAshort_adh] ( ), and C. ljungdahlii [pJIR750_budABCoperon] ( ) growing in in 50 ml Tanner medium containing 40 mM fructose.

0.1

1

3.5

4.0

4.5

5.0

5.5

6.0

6.5

0

5

10

15

20

25

30

35

40

45

0

20

40

60

80

100

120

0 50 100 150 200 250 300 350 4000

5

10

15

20

25

30

35

40

0 50 100 150 200 250 300 350 4000

1

2

3

4

5

6

7

8

Gro

wth

[O

D6

00

nm

]

pH

Fru

cto

se [

mM

]

Ace

tate

[m

M]

Eth

an

ol [

mM

]

Time [h]

2,3

-BD

[m

M]

Time [h]

Page 86: 2,3-Butanediol production using acetogenic bacteria

72 3. Results

production, pH of C. ljungdahlii [pMTL82151] decreased only to a value of 4.5, while all other

strains showed pH values of 4.25 after 356 h. Regarding ethanol and 2,3-BD production,

C. ljungdahlii [pMTL82151] showed even more differences compared to the other strains.

After 66 h C. ljungdahlii [pMTL82151] produced 31.5 mM ethanol, but after 74 h ethanol

concentration decreased to 19.3 mM with a subsequent increase to 24.1 mM. Also,

C. ljungdahlii [pJIR750] produced higher ethanol amounts of 29.5 mM after 212 h. In terms of

2,3-BD production, C. ljungdahlii [pMTL82151] produced the highest 2,3-BD concentration of

6.5 mM, which is 3.6 times higher than C. ljungdahlii WT (1.8 mM). From the newly

constructed strains C. ljungdahlii [pJIR750_23BD_budAshort], C. ljungdahlii

[pJIR750_23BD_budAshort_adh], and C. ljungdahlii [pJIR750_budABCoperon], only

C. ljungdahlii [pJIR750_23BD_budAshort_adh] produced higher amounts of 2,3-BD with

2.1 mM compared to C. ljungdahlii WT, but still lower concentrations than C. ljungdahlii

[pJIR750_23BD] with 2.7 mM.

Figure 24 presents growth, pH, and production pattern of C. ljungdahlii WT, C. ljungdahlii

[pMTL82151], C. ljungdahlii [pJIR750], C. ljungdahlii [pMTL82151_23BD], C. ljungdahlii

[pJIR750_23BD], C. ljungdahlii [pJIR750_23BD_budAshort], C. ljungdahlii

[pJIR750_23BD_budAshort_adh], and C. ljungdahlii [pJIR750_budABCoperon] under

autotrophic conditions. Growth experiments were performed as biological triplicates in

500-ml glass flasks containing 50 ml Tanner medium with syngas (1 bar overpressure) as

energy and carbon source. All strains showed a very similar growth behaviour with a max.

OD600nm from 1.7 for C. ljungdahlii [pJIR750_23BD_budAshort] to 2.2 for C. ljungdahlii

[pJIR750_23BD_budAshort_adh]. The pH dropped to 4 within 384 h for all C. ljungdahlii strains

except for C. ljungdahlii [pJIR750], which decreased the pH within 816 h (Figure 24).

Concerning max. acetate concentrations, C. ljungdahlii [pMTL82151] produced 120.1 mM

followed by C. ljungdahlii WT and C. ljungdahlii [pJIR750_23BD_budAshort_adh] with

119.8 mM and 110.9 mM, respectively. In contrast, C. ljungdahlii [pJIR750_23BD],

C. ljungdahlii [pJIR750_23BD_budAshort], and C. ljungdahlii [pJIR750_budABCoperon]

produced less than 100 mM acetate (98 mM, 99 mM, and 98.6 mM, respectively). Regarding

ethanol production, C. ljungdahlii [pMTL82151] produced higher ethanol concentrations

(68.1 mM) compared to the other C. ljungdahlii strains (57.1 mM-31.6 mM), which was in

accordance with the growth experiment under heterotrophic conditions (Figure 23). However,

Page 87: 2,3-Butanediol production using acetogenic bacteria

3. Results 73

in terms of 2,3-BD production, C. ljungdahlii [pMTL82151] did not produce higher 2,3-BD

concentrations (1.4 mM) than the other C. ljungdahlii strains (Figure 24), which does not

reflect the results of the heterotrophic growth experiment (Figure 23). The highest 2,3-BD

production is shown for C. ljungdahlii [pJIR750_23BD_budAshort_adh] (3.3 mM), which is very

similar compared to C. ljungdahlii [pMTL82151_23BD], C. ljungdahlii [pJIR750_23BD], and

Figure 24: Growth, pH, and production pattern of C. ljungdahlii WT ( ), C. ljungdahlii [pMTL82151]( ) C. ljungdahlii [pJIR750] ( ), C. ljungdahlii [pMTL82151_23BD] ( ), C. ljungdahlii [pJIR750_23BD] ( ), C. ljungdahlii [pJIR750_23BD_budAshort] ( ),C. ljungdahlii [pJIR750_23BD_budAshort_adh] ( ), and C. ljungdahlii [pJIR750_budABCoperon] ( ) growing in 50 ml Tanner medium with syngas (1 bar overpressure).

0.1

1

4

5

6

0

20

40

60

80

100

120

0 500 1000 1500 2000 2500 30000

10

20

30

40

50

60

70

80

0 500 1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

Gro

wth

[O

D6

00

nm

]

pH

Ace

tate

[m

M]

Eth

ano

l [m

M]

Time [h]

2,3

-BD

[m

M]

Time [h]

Page 88: 2,3-Butanediol production using acetogenic bacteria

74 3. Results

C. ljungdahlii [pJIR750_23BD_budAshort] (3.1 mM, 2.9 mM, and 2.8 mM, respectively) and

about two times higher than C. ljungdahlii WT (1.7 mM).

To sum it up, modifications of plasmid pJIR750_23BD did not lead to much higher 2,3-BD

concentrations with 3.3 mM for C. ljungdahlii [pJIR750_23BD_budAshort_adh] compared to

2.9 mM for C. ljungdahlii [pJIR750_23BD] (Figure 24). Furthermore, the strains C. ljungdahlii

[pJIR750_23BD_budAshort] and C. ljungdahlii [pJIR750_budABCoperon] did not produce more

2,3-BD than C. ljungdahlii [pJIR750_23BD] under autotrophic growth conditions (Figure 24).

Nevertheless, a higher ethanol and 2,3-BD production was shown for C. ljungdahlii

[pMTL82151] under heterotrophic growth conditions (Figure 23). Since this strain contains the

vector pMTL82151 without any genes for 2,3-BD production, it was interesting to find the

reason for higher ethanol and 2,3-BD concentrations in C. ljungdahlii [pMTL82151]. Table 6

and Table 7 present the summarized growth and production pattern of the investigated

C. ljungdahlii strains under heterotrophic and autotrophic growth conditions.

Table 6: Summary of max. OD600nm, max. acetate production, max. ethanol production, and max. 2,3-BD production of recombinant C. ljungdahlii strains in comparison to C. ljungdahlii WT under heterotrophic growth conditions.

Strain Max.

OD600nm

Max. acetate

production [mM]

Max. ethanol

production

[mM]

Max. 2,3-BD

production

[mM]/[mg/l]

Heterotrophic growth conditions

C. ljungdahlii WT 2.7 105.0 8.5 1.8/162.2

C. ljungdahlii [pMTL82151] 2.0 52.1 31.5 6.5/585.8

C. ljungdahlii [pJIR750] 1.6 96.1 7.2 1.0/90.1

C. ljungdahlii [pMTL82151_23BD] 1.7 119.5 7.5 1.3/117.2

C. ljungdahlii [pJIR750_23BD] 0.9 98.4 29.5 2.7/243.3

C. ljungdahlii

[pJIR750_23BD_budAshort]

2.3 95.1 6.2 1.5/135.1

C. ljungdahlii

[pJIR750_23BD_budAshort_adh]

2.8 103.1 7.1 2.1/189.3

C. ljungdahlii

[pJIR750_budABCoperon]

2.6 110.4 9.0 1.9/171.2

Page 89: 2,3-Butanediol production using acetogenic bacteria

3. Results 75

Table 7: Summary of max. OD600nm, max. acetate production, max. ethanol production, and max. 2,3-BD production of recombinant C. ljungdahlii strains in comparison to C. ljungdahlii WT under autotrophic growth conditions.

Strain Max.

OD600nm

Max. acetate

production [mM]

Max. ethanol

production

[mM]

Max. 2,3-BD

production

[mM]/[mg/l]

Autotrophic growth conditions

C. ljungdahlii WT 2.0 119.8 41.9 1.7/153.2

C. ljungdahlii [pMTL82151] 1.9 120.1 68.1 1.5/135.2

C. ljungdahlii [pJIR750] 1.5 105.4 40.6 1.1/099.1

C. ljungdahlii [pMTL82151_23BD] 1.8 109.0 52.5 3.1/279.4

C. ljungdahlii [pJIR750_23BD] 1.9 98.0 57.1 2.9/261.3

C. ljungdahlii

[pJIR750_23BD_budAshort]

1.7 99.0 42.2 2.8/252.3

C. ljungdahlii

[pJIR750_23BD_budAshort_adh]

2.2 110.9 50.3 3.3/297.4

C. ljungdahlii

[pJIR750_budABCoperon]

1.6 98.6 31.6 1.8/162.2

3.1.6 Sequencing of C. ljungdahlii [pMTL82151] and C. ljungdahlii WT with SNP

analysis

Growth experiments using fructose or syngas as energy and carbon source (Figure 23 and

Figure 24) demonstrated that C. ljungdahlii [pMTL82151] produced decreased acetate

concentrations accompanied by increased ethanol and 2,3-BD production. Thus, the question

evoked why this production pattern occurred. Since C. ljungdahlii [pMTL82151] harbors the

vector pMTL82151 without any genes for 2,3-BD production, there might be one or more

mutations in genes on the chromosome that affect ethanol and 2,3-BD production. For this

reason, sequencing of C. ljungdahlii [pMTL82151] and C. ljungdahlii WT was performed in

cooperation with the Göttingen Genomics Laboratory (Göttingen, Germany). SNP analysis of

C. ljungdahlii [pMTL82151] compared to C. ljungdahlii WT revealed many SNPs in different

locus tags, which are presented in Table 12 (chapter 7. Supplement). Table 8 summarizes locus

tags CLJU_c07590, CLJU_c07600, CLJU_c16510, CLJU_c16520, CLJU_c24850, and

Page 90: 2,3-Butanediol production using acetogenic bacteria

76 3. Results

Table 8: Results of SNP analysis of C. ljungdahlii [pMTL82151] compared to C. ljungdahlii WT

Locus tag* Annotation AA position changed Original AA Changed AA

CLJU_c07590 putative secretion

protein HlyD

1123, 1126, 1144, 1146,

1148, 1171

V, Y, H, H, A, G D, F, R, R, T, D

CLJU_c07600 putative transporter

protein

2794, 2797, 2821, 2836,

2848, 2854

D, G, M, Q, P, I G, V, K, P, H, R

CLJU_c16510 bifunctional

aldehyde/alcohol

dehydrogenase

544, 568, 607 M, A, S R, V, no AA

CLJU_c16520 bifunctional

aldehyde/alcohol

dehydrogenase

520, 532, 1768, 1769 D, I, E, E A, S, D, D

CLJU_c24850 signal-transduction

and transcriptional-

control protein

802, 1342 E, R V, T

CLJU_c24870 signal-transduction

and transcriptional-

control protein

1817, 1818 M, M I, I

*according to IMG /MER (https://img.jgi.doe.gov/cgi-bin/mer/main.cgi), 25th of april, 2017.

CLJU_c24870 showing one or more changed amino acids. The locus tag CLJU_c07590 (putative

secretion protein HylD) showed six changed AA and CLJU_c07600 (putative transporter

protein) displayed five changed AA. The locus tags CLJU_c16510 (gene adh1) and CLJU_c16520

(gene adh2), which are both annotated as bifunctional aldehyde/alcohol dehydrogenases,

showed three and four changed AA, respectively. Moreover, locus tag CLJU_c16510 contained

a stop codon at AA position 607 (length of total gene 2613 bp). This means that a shortened

protein with 668 less amino acids was translated in C. ljungdahlii [pMTL82151] compared to

C. ljungdahlii WT. Furthermore, locus tags CLJU_c24850 (gene stc1) and CLJU_c24870 (gene

stc2), which are both annotated as signal-transduction and transcriptional-control proteins

showed two changed AA. Therefore, strain C. ljungdahlii [pMTL82151] harboring these SNPs

is designated as mutated C. ljungdahlii [pMTL82151] in further experiments.

Page 91: 2,3-Butanediol production using acetogenic bacteria

3. Results 77

3.1.7 Detection and quantification of transcripts responsible for 2,3-BD production

In a further experiment, mRNAs of the genes alsS, budA, and bdh were detected by Northern

blot analysis to show that the transcript alsS-budA-bdh is completely expressed in

recombinant strains C. ljungdahlii [pMTL82151_23BD], C. ljungdahlii [pJIR750_23BD],

C. ljungdahlii [pJIR750_23BD_budAshort], and C. ljungdahlii [pJIR750_23BD_adh]. All

C. ljungdahlii strains for Northern blot analysis were grown with 40 mM fructose and RNA was

isolated after 65 h during early stationary growth phase, when 2,3-BD production starts, which

was checked by HPLC analysis (data not shown). Figure 25 depicts the gene organisation of

the synthetic 2,3-BD operon (alsS-budA-bdh) as well as the genes alsS, budA, and bdh present

on the chromosome of C. ljungdahlii with respective sizes of mRNA. For Northern blot

experiment, mRNAs originating from the chromosome were detected by using RNA of

C. ljungdahlii WT, C. ljungdahlii [pMTL82151], and C. ljungdahlii [pJIR750] and as negative

control, RNA from E. coli XL1-Blue MRF' was applied. Figure 26 shows Northern blot analysis

with probes alsS (Figure 26A), budA (Figure 26B), and bdh (Figure 26C), which were amplified

using CljalsS-sonde_fwd, CljalsS-sonde_rev, CljBudA-sonde_fwd, CljBudA-sonde_rev,

Clj23bdh-sonde_fwd, and Clj23bdh-sonde_rev, respectively. The probes were hybridized with

RNA from C. ljungdahlii WT, C. ljungdahlii [pMTL82151], C. ljungdahlii [pJIR750], C. ljungdahlii

Figure 25: Schematic organisation of 2,3-BD genes on chromosome of C. ljungdahlii with size of transcript alsS (A), budA (B), and bdh (C) as well as sizes of synthetic 2,3-BD operons of overproduction plasmids pMTL81151_23BD as well as pJIR750_23BD (D), pJIR750_23BD_budAshort (E), and pJIR750_23BD_budAshort_adh (F).

Page 92: 2,3-Butanediol production using acetogenic bacteria

78 3. Results

[pMTL82151_23BD], C. ljungdahlii [pJIR750_23BD], C. ljungdahlii [pJIR750_23BD_budAshort],

C. ljungdahlii [pJIR750_23BD_budAshort_adh], and E. coli XL1-Blue MRF' as negative control.

The transcript alsS from C. ljungdahlii chromosome was successfully detected in C. ljungdahlii

WT at a size of 1,670 bases (Figure 26A, lane 1) and the transcript alsS-budA-bdh could be

identified in C. ljungdahlii [pJIR750_23BD] at a size of 3,802 bases (Figure 26A, lane 5). In

contrast, no signal at all was detected for C. ljungdahlii [pMTL82151] (Figure 26A, lane2),

C. ljungdahlii [pMTL82151_23BD] (Figure 26A, lane 4), and C. ljungdahlii

[pJIR750_23BD_budAshort_adh] (Figure 26A, lane 7). Moreover, a signal with a smaller size of

about 1,000 bases was detected in the strain C. ljungdahlii [pJIR750] (Figure 26A, lane 3) and

in C. ljungdahlii [pJIR750_23BD_budAshort] an additional signal at 1,500 bases occurred

(Figure 26A, lane 6). Analysis of transcript budA of C. ljungdahlii WT (Figure 26B, lane 1)

revealed a signal at size of about 1,250 bases, but it actually should show a length of 720 bases.

Moreover, this transcript was also present in C. ljungdahlii [pMTL82151] and C. ljungdahlii

[pJIR750_23BD_budAshort_adh] (Figure 26B, lane 2,7), but not in the other recombinant

strains. A shorter transcript at a size of 750 bases was again existing in

C. ljungdahlii [pMTL82151], C. ljungdahlii [pJIR750], C. ljungdahlii [pJIR750_23BD], and

C. ljungdahlii [pJIR750_23BD_budAshort_adh]. Furthermore, the transcript alsS-budA-bdh is

again detected solely in C. ljungdahlii [pJIR750_23BD] (Figure 26B, lane 5), but in C. ljungdahlii

[pMTL82151_23BD] a very weak signal occured at a size of about 2,500 bases, which is shorter

than the transcript alsS-budA-bdh. Figure 26C shows Northern blot analysis with bdh probe

and transcript bdh with a size of 1,074 bases is present in C. ljungdahlii WT as well as

C. ljungdahlii [pMTL82151] (Figure 26C, lane 1 and 2). The recombinant strain C. ljungdahlii

[pJIR750_23BD_budAshort_adh] was not analyzed, since the plasmid contains the adh gene

Figure 26: Northern blot analysis with alsS probe (A), budA probe (B), and bdh probe (C) hybridized with RNA of C. ljungdahlii WT (1), C. ljungdahlii [pMTL82151] (2), C. ljungdahlii [pJIR750] (3), C. ljungdahlii [pMTL82151_23BD] (4), C. ljungdahlii [pJIR750_23BD] (5), C. ljungdahlii [pJIR750_23BD_budAshort] (6), C. ljungdahlii [pJIR750_23BD_budAshort_adh] (7), and E. coli XL1-Blue MRF' as negative control (8); RiboRuler High Range RNA Ladder (M).

Page 93: 2,3-Butanediol production using acetogenic bacteria

3. Results 79

instead of the bdh gene and so detection of the transcript alsS-budAshort-adh is not possible

by bdh probe in this strain. Besides, in accordance with results of alsS probe and budA probe,

transcript alsS-budA-bdh was only detected in C. ljungdahlii [pJIR750_23BD] (Figure 26C, lane

5). Moreover, no transcript at all was detected in C. ljungdahlii [pMTL82151_23BD] and

C. ljungdahlii [pJIR750_23BD_budAshort] and as mentioned above, a shorter signal at a size

of 500 bases was present.

Summarizing the results of Northern blot analysis with probes alsS, budA, and bdh, it can be

stated that transcript originating from C. ljungdahlii chromosome was always present in

C. ljungdahlii WT but not in each recombinant strain. Furthermore, transcript for 2,3-BD

overproduction was only detected in C. ljungdahlii [pJIR750_23BD].

3.1.8 Heterologous 2,3-BD production in A. woodii

The acetogenic organism A. woodii is a promising organism for heterologous 2,3-BD

production, since it was used previously for heterologous production of acetone with H2 + CO2

(Hoffmeister et al., 2016). For that reason, A. woodii was also taken into account to

overproduce 2,3-BD with the above mentioned 2,3-BD production plasmids (Figure 7,

Figure 15, Figure 18, Figure 21). For investigation of 2,3-BD production in A. woodii, the five

different unmethylated 2,3-BD production plasmids pMTL82151_23BD, pJIR750_23BD,

pJIR750_23BD_budAshort, pJIR750_23BD_budAshort_adh, and pJIR750_budABCoperon

were transformed into A. woodii. The recombinant strains A. woodii [pMTL82151] and

A. woodii [pJIR750] already existed (Hoffmeister, 2017), but were verified after reactivation of

lyophilized cultures by amplification (pMTL82151catP_fwd, pMTL82151catP_rev,

pJIR750catP_fwd, pJIR750catP_rev) of catP from pMTL82151 backbone (expected DNA

fragment 1,121 bp; Figure 27A) or pJIR750 backbone (expected DNA fragment 1,075 bp;

Figure 27A). Furthermore, 16S rDNA was amplified (fD1, rP2; expected DNA fragment

1,500 bp; Figure 27C) with genomic DNA as template followed by sequencing (fD1, rP2) by

GATC Biotech AG (Constance, Germany). New recombinant strains were verified either by

isolation of genomic DNA and amplification of catP (expected DNA fragment 1,500 bp;

Figure 27A) or via retransformation of genomic DNA from recombinant strains into E. coli XL1-

Blue MRF', subsequent plasmid isolation, and analytical restriction enzyme digestion (Figure

27B).

Page 94: 2,3-Butanediol production using acetogenic bacteria

80 3. Results

Figure 27B shows the analytical restriction enzyme digestions using Eco32I of retransformed

plasmid pJIR750_23BD_budAshort (expected DNA fragments 6,233 bp, 3,875 bp, 759 bp),

using BamHI and SacI of retransformed pJIR750_23BD (expected DNA fragments 7,890 bp,

3,070 bp), using BamHI and Bsu15I of retransformed pMTL82151_23BD (expected DNA

fragments 7,983 bp, 1,663 bp), using BamHI and SalI of retransformed

pJIR750_23BD_budAshort_adh (expected DNA fragments 9,522 bp, 1,218 bp), and using NdeI

of retransformed pJIR750_budABCoperon (expected DNA fragments 5,425 bp, 3,092 bp,

1,825 bp). Additionally, amplification of 16S rDNA gene (1,500 bp; fD1, rP2) and subsequent

sequencing of the resulting DNA fragment supported correct identity of the recombinant

A. woodii strains (Figure 27C). Using these methods, recombinant strains A. woodii

[pMTL82151_23BD], A. woodii [pJIR750_23BD], A. woodii [pJIR750_23BD_budAshort],

A. woodii [pJIR750_23BD_budAshort_adh], and A. woodii [pJIR750_budABCoperon] were

verified.

Figure 27: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion (B), and amplification of 16S rDNA gene of genomic DNA from recombinant A. woodii strains (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,121-bp (pMTL82151 backbone) and 1,075-bp (pJIR750 backbone) DNA fragments resulting from PCR of catP

from A. woodii [pMTL82151] (3), A. woodii [pMTL82151_23BD] (4), A. woodii [pJIR750] (7), A. woodii

[pJIR750_23BD] (8), A. woodii [pJIR750_23BD_budAshort] (9), A. woodii

[pJIR750_23BD_budAshort_adh] (10), A. woodii [pJIR750_budABCoperon] (11), positive control with empty vector pMTL82151 (2) and pJIR750 (6) as template, and negative control with sterile water astemplate (1, 5). (B) Analytical restriction enzyme digestion using Eco32I of retransformed pJIR750_23BD_budAshort with expected DNA fragments 6,233 bp, 3,875 bp and 759 bp (1), using BamHI and SacI of retransformed pJIR750_23BD with expected DNA fragments 7,890 bp and 3,070 bp (2), using BamHI and Bsu15I of retransformed pMTL82151_23BD with expected DNA fragments 7,983 bp and 1,663 bp (3), using BamHI and SalI of retransformed pJIR750_23BD_budAshort_adh with expected DNA fragments 9,522 bp and 1,218 bp (4), and using NdeI of retransformed pJIR750_budABCoperon with expected DNA fragment 5,425 bp, 3,092 bp and 1,825 bp (4). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of A. woodii [pMTL82151] (2), A. woodii [pJIR750] (3), A. woodii [pMTL82151_23BD] (4), A. woodii [pJIR750_23BD] (5), A. woodii [pJIR750_23BD_budAshort] (6), A. woodii [pJIR750_23BD_budAshort_adh] (7), A. woodii

[pJIR750_budABCoperon] (8), and negative control with sterile water as template (1).

Page 95: 2,3-Butanediol production using acetogenic bacteria

3. Results 81

Figure 28 presents growth, pH, and production pattern of A. woodii WT, A. woodii

[pMTL82151], A. woodii [pJIR750], A. woodii [pMTL82151_23BD], A. woodii [pJIR750_23BD],

A. woodii [pJIR750_23BD_budAshort], A. woodii [pJIR750_23BD_budAshort_adh], and

A. woodii [pJIR750_budABCoperon] under autotrophic conditions with H2 + CO2 as energy and

carbon source. Growth experiments were performed as biological triplicates in 500-ml glass

flasks containing 50 ml modified ATCC 1612 medium and H2 + CO2 as energy and carbon source

(1 bar overpressure). In terms of growth behaviour, A. woodii WT and A. woodii [pJIR750]

showed higher max. OD600nm values with 1.0 and 1.1, respectively, after 120 h compared to

the other strains (0.3-0.5). After this maximum OD600nm (120 h), growth of all strains

decreased. Furthermore, the pH started at 7.5 and decreased to 5.5 after 984 h in accordance

Figure 28: Growth, pH, and production pattern of A. woodii WT ( ), A. woodii [pMTL82151]( ), A. woodii [pJIR750] ( ), A. woodii [pMTL82151_23BD] ( ), A. woodii

[pJIR750_23BD] ( ), A. woodii [pJIR750_23BD_budAshort] ( ), A. woodii

[pJIR750_23BD_budAshort_adh] ( ), and A. woodii [pJIR750_budABCoperon] ( ) growing in 50 ml modified ATCC 1612 medium with H2 + CO2 (1 bar overpressure).

0.01

0.1

1

5

6

7

8

0 100 200 300 400 500 600 700 800 900 10000

20

40

60

80

100

120

140

160

180

Gro

wth

[O

D6

00

nm

]p

HA

ceta

te [

mM

]

Time [h]

Page 96: 2,3-Butanediol production using acetogenic bacteria

82 3. Results

with increasing acetate concentration. Acetate was the only detectable product in all strains

with A. woodii [pJIR750_23BD_budAshort] producing most (166.0 mM). Also,

A. woodii [pJIR750_23BD] and A. woodii [pJIR750_23BD_budAshort_adh] produced slightly

higher amounts of acetate with 163.9 mM and 156.3 mM, respectively, compared to A. woodii

WT (148.5 mM). All in all, none of the recombinant A. woodii strains was able to produce 2,3-

BD heterologously.

3.1.9 Enhancement of 2,3-BD production by overexpressing nifJ

The enzyme pyruvate:ferredoxin oxidoreductase (PFOR) converts acetyl-CoA to pyruvate by

oxidizing reduced ferredoxin. Additionally, two mol of pyruvate are needed to build one mol

of 2,3-BD (Figure 29). The company LanzaTech Inc. (Skokie, IL, USA) recently published a

patent (Köpke and Mueller, 2016) in which they describe C. autoethanogenum strains that

overexpress the gene nifJ, which encodes PFOR enzyme, in addition to the genes alsS and

budA. The enzymes capable of reducing acetoin to 2,3-BD were not overproduced, since

investigation of enzyme activities revealed that 2,3-BD dehydrogenase (bdh) and primary-

secondary alcohol dehydrogenase (adh) from C. autoethanogenum WT showed a high enzyme

activity with 1.2 U/mg (0.8 U/mg with NADH and 0.4 U/mg with NADPH) compared to PFOR

with 0.11 U/mg (Köpke and Mueller, 2016). According to the results of LanzaTech Inc. (Skokie,

IL, USA), it seems not necessary to overproduce 2,3-BD dehydrogenase or primary-secondary

Figure 29: 2,3-BD production in C. ljungdahlii originating from acetyl-CoA via enzymes PFOR (pyruvate:ferredoxin oxidoreductase), AlsS (acetolactate synthase), BudA (acetolactate decarboxylase), BDH (2,3-BD dehydrogenase).

Page 97: 2,3-Butanediol production using acetogenic bacteria

3. Results 83

alcohol dehydrogenase for 2,3-BD overproduction. Therefore, new plasmids were constructed

starting from the plasmids pMTL82151_23BD and pJIR750_23BD. In one attempt, the nifJ gene

was cloned in addition to alsS, budA, and bdh into pMTL82151_23BD and pJIR750_23BD and

in another attempt, the gene bdh was exchanged against nifJ. The sequence of nifJ was

amplified (PFORInfusionBamHI_fwd, PFORInfusionBamHI_rev, PFORohneBDHBamHI_fwd,

PFORohneBDHBamHI_rev) using genomic DNA from C. ljungdahlii WT as template, and

plasmids pMTL82151_23BD as well as pJIR750_23BD were digested either using BamHI for

linearization or BamHI and SalI for cutting out bdh (Figure 30 and Figure 31). Subsequently,

amplified DNA fragments were cloned into the linearized plasmids by “In-Fusion HD Cloning”.

Successful cloning was confirmed by analytical restriction digestion using restriction enzymes

BamHI or BamHI and SalI, as well as sequencing (PFOR-Seq1, PFOR-Seq2, PFOR-Seq3, PFOR-

Figure 30: Cloning strategy for construction of plasmids pMTL82151_23BD_PFOR (A) and pMTL82151_23BD_oBDH_PFOR (B). Plasmid pMTL82151_23BD was digested either with restriction enzymes BamHI for linearization or BamHI and SalI for cutting out bdh. Gene nifJ was coned into linearized pMTL82151_23BD by “In-Fusion HD cloning”, resulting in plasmids pMTL82151_23BD_PFOR and pMTL82151_23BD_oBDH_PFOR. Ppta-ack, promoter region upstream of pta-ack operon from C. ljungdahlii; alsS, acetolactate synthase from C. ljungdahlii; budA, acetolactate decarboxylase from C. ljungdahlii; bdh, 2,3-BD dehydrogenase from C. ljungdahlii; nifJ, PFOR from C. ljungdahlii; rep (ColE1), ColE1-replicon for Gram-negative bacteria; repA (pBP1)/orf2, pBP1-replicon for Gram-positive bacteria from C. botulinum; rep (pMB1, replicon for Gram-negative bacteria); rep(pIP404), pIP404-replicon for Gram-positive bacteria from C. perfringens; catP, chloramphenicol resistance gene; traJ, gene for conjugal transfer function.

Page 98: 2,3-Butanediol production using acetogenic bacteria

84 3. Results

Seq4, PFOR-Seq5) by GATC Biotech AG (Constance, Germany). The resulting plasmids were

named pMTL82151_23BD_PFOR, pMTL82151_23BD_oBDH_PFOR, pJIR750_23BD_PFOR, and

pJIR750_23BD_oBDH_PFOR (Figure 30A and B as well as Figure 31A and C). All constructed

plasmids were in vivo methylated using E. coli ER2275 [pACYC184_MCljI] and subsequently

isolated using “QIAGEN Plasmid Midi Kit” (QIAGEN-tip 100; Qiagen GmbH, Hilden, Germany).

Attempts of transforming the methylated plasmids pMTL82151_23BD_PFOR,

pMTL82151_23BD_oBDH_PFOR, pJIR750_23BD_PFOR, and pJIR750_23BD_oBDH_PFOR into

C. ljungdahlii by electroporation were not successful and since pMTL82151 vector harbors

already genes for conjugal transfer (oriT and traJ), conjugation experiments were conducted

in order to transform the newly constructed plasmids into C. ljungdahlii. Since vector pJIR750

does not contain any genes for conjugation, the oriT/traJ region of pMTL82151 was amplified

Figure 31: Cloning strategy for construction of plasmids pJIR750_23BD_PFOR (A), pJIR750_23BD_oBDH_PFOR (C), pJIR750_23BD_PFOR_traJ (B), and pJIR750_23BD_oBDH_PFOR_traJ (D). Plasmid pJIR750_23BD was digested either using restriction enzymes BamHI (A) for linearization or BamHI and SalI for cutting out bdh (C). Gene nifJ was cloned into linearized pJIR750_23BD by In-Fusion HD cloning, resulting in plasmids pJIR750_23BD_PFOR (A) and pJIR750_23BD_oBDH_PFOR (C). For construction of plasmids pJIR750_23BD_PFOR_traJ (B) and pJIR750_23BD_oBDH_PFOR_traJ (D), digestion of pJIR750_23BD_PFOR and pJIR750_23BD_oBDH_PFOR with EheI was performed. DNA fragment oriT/traJ was cloned into linearized pJIR750_23BD_PFOR and pJIR750_23BD_oBDH_PFOR by In-Fusion HD cloning. Ppta-ack, promoter region upstream of pta-ack

operon from C. ljungdahlii; alsS, acetolactate synthase from C. ljungdahlii; budA, acetolactate decarboxylase from C. ljungdahlii; bdh, 2,3-BD dehydrogenase from C. ljungdahlii; nifJ, PFOR from C. ljungdahlii; rep (ColE1), replicon for Gram-negative bacteria; repA/orf2 (pBP1), pBP1-replicon for Gram-positive bacteria from C. botulinum; rep (pMB1), pMB1-replicon for Gram-negative bacteria; rep (pIP404), pIP404-replicon for Gram-positive bacteria from C. perfringens; catP, chloramphenicol resistance gene; traJ, gene for conjugal transfer function; oriT, origin of conjugal transfer function.

Page 99: 2,3-Butanediol production using acetogenic bacteria

3. Results 85

(InfusionTraJEheI_fwd, InfusionTraJEheI_rev). Afterwards, plasmids pJIR750_23BD_PFOR and

pJIR750_23BD_oBDH_PFOR were linearized by restriction enzyme digestion using EheI and

subsequently oriT/traJ was inserted using “In-Fusion HD cloning” (Figure 31D). Successful

cloning was confirmed by analytical restriction digestion using EheI, as well as sequencing

(traJseq1, traJSeq_rev) by GATC Biotech AG (Constance, Germany). The resulting plasmids

were named pJIR750_23BD_PFOR_traJ and pJIR750_23BD_oBDH_PFOR_traJ (Figure 31B

and D). Subsequent conjugation experiments were only successful for plasmids

pMTL82151_23BD_PFOR and pMTL82151_23BD_oBDH_PFOR. The transformed strains

growing in presence of thiamphenicol and colistin were verified by two different methods. On

the one hand, isolation of genomic DNA from recombinant C. ljungdahlii strains was

performed and catP (expected DNA fragment 1,121 bp) was amplified via PCR

(pMTL82151catP_fwd, pMTL82151catP_rev) to verify presence of the plasmid (Figure 32A).

On the other hand, isolated genomic DNA was retransformed into chemically competent

E. coli XL-1 Blue MRF' cells and single colonies growing on agar plates containing

chloramphenicol were picked for plasmid DNA isolation. Figure 32B and C show the analytical

restriction enzyme digestions using NdeI of retransformed plasmids pMTL82151_23BD_PFOR

(expected DNA fragments 8,719 bp, 4,477 bp) and pMTL82151_23BD_oBDH_PFOR (expected

Figure 32: Analysis of DNA fragments resulting from amplification of catP (A), analytical restriction enzyme digestion (B), and amplification of 16S rDNA gene of genomic DNA from recombinant C. ljungdahlii strains (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) 1,121-bp (pMTL82151 backbone) DNA fragment resulting from PCR of catP from C. ljungdahlii

[pMTL82151_23BD_PFOR] (3), C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] (4), positive control with empty vector pMTL82151 (2) as template, and negative control with sterile water as template(1). (B) Analytical restriction enzyme digestion using NdeI of retransformed pMTL82151_23BD_oBDH_PFOR with expected DNA fragments 7,421 bp and 4,477 bp (2), and undigested control (1), using NdeI of retransformed pMTL82151_23BD_PFOR with expected DNA fragments 8,719 bp and 4,477 bp (4), and undigested control (3). (C) 1,500-bp DNA fragment resulting from amplification of 16S rDNA gene of C. ljungdahlii [pMTL82151_23BD_PFOR] (3), C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] (4), positive control C. ljungdahlii WT (2), and negative control sterile water (1).

Page 100: 2,3-Butanediol production using acetogenic bacteria

86 3. Results

DNA fragments 7,421 bp, 4,477 bp). Additionally, amplification of 16S rDNA gene (1,500 bp;

fD1, rP2) and concomitant sequencing of the resulting DNA fragment by GATC Biotech AG

(Constance, Germany) supported correct identity of the recombinant strains (Figure 32C).

Using these methods, recombinant strains C. ljungdahlii [pMTL82151_23BD_PFOR] and

C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] were verified. These recombinant

C. ljungdahlii strains were analyzed under both heterotrophic and autotrophic growth

conditions.

Figure 33 presents growth, fructose consumption, and production pattern of C. ljungdahlii WT,

mutated C. ljungdahlii [pMTL82151], C. ljungdahlii [pMTL82151_23BD_PFOR], and

C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] under heterotrophic conditions. Growth

experiments were performed as biological triplicates in 125-ml glass flasks containing 50 ml

Tanner medium with 30 mM fructose as energy and carbon source. In terms of growth,

C. ljungdahlii WT reached a higher max. OD600nm (2.4) after 50 h compared to the other strains

with 1.9-1.8 after 50 h. Regarding fructose concentration it was striking that C. ljungdahlii

[pMTL82151_23BD_PFOR] was the only strain, which was not completely consuming fructose.

The pH decreased to 4.0 after 72 h for all strains and mutated C. ljungdahlii [pMTL82151]

showed a decreased acetate production of 46.9 mM after 218 h compared to

C. ljungdahlii WT and C. ljungdahlii [pMTL82151_23BD_PFOR] of 62.9 mM and 63.0 mM,

respectively. In contrast, C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] produced the highest

acetate concentration with 71.4 mM. Comparing ethanol production of the different

C. ljungdahlii strains, it was noticeable that in accordance with previous results (Figure 23)

mutated C. ljungdahlii [pMTL82151] produced higher ethanol concentrations (23.1 mM) in

contrast to the other strains (5.8 mM-7.1 mM). In terms of 2,3-BD production, C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] exposed the highest 2,3-BD concentration after 218 h with

2.2 mM, which is very similar to mutated C. ljungdahlii [pMTL82151] with 1.9 mM. However,

C. ljungdahlii [pMTL82151_23BD_PFOR] produced comparable concentrations of 2,3-BD as

C. ljungdahlii WT with 0.7 mM after 218 h.

Figure 34 depicts growth, pH, and production pattern of C. ljungdahlii wildtype (WT), mutated

C. ljungdahlii [pMTL82151], C. ljungdahlii [pMTL82151_23BD_PFOR], and C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] under autotrophic conditions. Growth experiments were

performed as biological triplicates in 500-ml glass flasks containing 50 ml Tanner medium with

Page 101: 2,3-Butanediol production using acetogenic bacteria

3. Results 87

syngas (1 bar overpressure) as energy and carbon source. All recombinant C. ljungdahlii strains

reached a lower max. OD600nm (1.2-1.3) compared to C. ljungdahlii WT (1.7). The strain

C. ljungdahlii [PMTL82151_23BD_oBDH_PFOR] exhibited the lowest OD600nm of 0.8. After

2,207 h, pH decreased to 4 for all C. ljungdahlii strains. Regarding acetate production, all

strains produced very similar concentrations after 2,207 h from 103.8 mM to 110.2 mM. In

contrast, ethanol and 2,3-BD production showed a completely different picture. It was striking

Figure 33: Growth, pH, fructose consumption, and production pattern of C. ljungdahlii WT ( ), mutated C. ljungdahlii [pMTL82151] ( ), C. ljungdahlii [pMTL82151_23BD_PFOR] ( ), and C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] ( ) growing in 50 ml Tanner medium containing 30 mM fructose.

0.1

1

3.5

4.0

4.5

5.0

5.5

6.0

0

5

10

15

20

25

30

35

0

20

40

60

80

100

0 50 100 150 200 2500

5

10

15

20

25

0 50 100 150 200 2500.0

0.5

1.0

1.5

2.0

2.5

3.0

Gro

wth

[O

D6

00

nm

]

pH

Fru

cto

se [

mM

]

Ace

tate

[m

M]

Eth

ano

l [m

M]

Time [h]

2,3

-BD

[m

M]

Time [h]

Page 102: 2,3-Butanediol production using acetogenic bacteria

88 3. Results

that C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] produced much less ethanol (14.1 mM)

compared to the other strains (43.5 mM-50.9 mM) and most importantly, comparing 2,3-BD

production C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] produced the highest 2,3-BD

concentration (4.7 mM) besides mutated C. ljungdahlii [pMTL82151] (5.6 mM), which is more

than twice as much compared to C. ljungdahlii WT (2.2 mM). However, recombinant strain

Figure 34: Growth, pH, and production pattern of C. ljungdahlii WT ( ), mutated C. ljungdahlii [pMTL82151] ( ),C. ljungdahlii [pMTL82151_23BD_PFOR] ( ), and C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] ( ) growing in 50 ml Tanner medium with syngas (1 bar overpressure).

0.01

0.1

1

4.0

4.5

5.0

5.5

6.0

0

20

40

60

80

100

120

0 500 1000 1500 20000

10

20

30

40

50

60

0 500 1000 1500 20000

1

2

3

4

5

6

Gro

wth

[O

D6

00

nm

]

pH

Ace

tate

[m

M]

Eth

ano

l [m

M]

Time [h]

2,3

-BD

[m

M]

Time [h]

Page 103: 2,3-Butanediol production using acetogenic bacteria

3. Results 89

C. ljungdahlii [pMTL82151_23BD_PFOR] produced similar amounts of 2,3-BD (2.5 mM)

compared to C. ljungdahlii WT.

In summary, only C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] led to an improvement of

2,3-BD production, but it was still lower compared to mutated C. ljungdahlii [pMTL82151] with

syngas as energy and carbon source (Figure 34). Table 9 presents the summarized growth and

production pattern of the investigated C. ljungdahlii strains under heterotrophic and

autotrophic growth conditions.

Table 9: Summary of max. OD600nm, max. acetate production, max. ethanol production, and max. 2,3-BD production of recombinant C. ljungdahlii strains in comparison to C. ljungdahlii WT under heterotrophic and autotrophic growth conditions.

Strain Max.

OD600nm

Max. acetate

production [mM]

Max. ethanol

production

[mM]

Max. 2,3-BD

production

[mM]/[mg/l]

Heterotrophic growth conditions

C. ljungdahlii WT 2.4 77.7 6.6 1.0/090.1

Mutated C. ljungdahlii

[pMTL82151]

1.8 55.7 23.1 2.4/216.3

C. ljungdahlii

[pMTL82151_23BD_PFOR]

1.9 67.0 5.8 0.8/072.1

C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR]

1.9 71.4 7.1 2.2/198.3

Autotrophic growth conditions

C. ljungdahlii WT 2.0 108.5 50.9 2.2/198.3

Mutated C. ljungdahlii

[pMTL82151]

1.3 110.2 44.2 5.6/504.7

C. ljungdahlii

[pMTL82151_23BD_PFOR]

1.5 103.8 43.5 2.5/225.3

C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR]

0.9 109.0 14.1 4.7/423.6

Page 104: 2,3-Butanediol production using acetogenic bacteria

90 3. Results

3.2 Heterologous butanol production

Another aim of this study was the heterologous production of butanol in C. ljungdahlii. In a

previous study, the heterologous production of 2 mM butanol in C. ljungdahlii with the genes

originating from C. acetobutylicum was reported (Köpke et al., 2010). The genes for butanol

production from C. acetobutylicum, which are thlA (thiolase), hbd (3-hydroxybutyryl-CoA

dehydrogenase), crt (crotonase), bcd (butyryl-CoA dehydrogenase), bdhA (butanol

dehydrogenase), and adhE (bifunctional aldehyde/alcohol dehydrogenase) were located on

the plasmid pSOBptb under control of the ptb promoter. The plasmid pSOBptb does not contain

the genes etfA and etfB, which were shown to be essential for reduction of crotonyl-CoA to

butyryl-CoA in C. kluyveri (Li et al., 2008) as well as C. acetobutylicum (Lütke-Eversloh and Bahl,

2011). Therefore, plasmid pSB3C5-UUMKS 3 was synthesized by the company

ATG:biosynthetics GmbH (Freiburg, Germany) in a further study, containing the genes for

butanol production (thlA, hbd, crt, bcd, and adhE2) as well as etfA and etfB from

C. acetobutylicum (Schuster, 2011). Thus, pSB3C5-UUMKS 3 harbors all genes encoding

enzymes necessary for heterologous butanol production in C. ljungdahlii (Figure 35). However,

it only contains an origin of replication for Gram-negative bacteria (ori p15A) and butanol

production was only shown in E. coli (Schuster, 2011). For enabling the butanol plasmid to

Figure 35: Butanol production in C. ljungdahlii with enzymes originating from C. acetobutylicum starting from acetyl-CoA. Thl (thiolase), Hbd (3-hydroxybutyryl-CoA dehydrogenase), Crt (crotonase), Bcd (butyryl-CoA dehydrogenase), EtfA/B (electron transfer flavoproteins), AdhE2 (bifunctional aldehyde/alcohol dehydrogenase).

Page 105: 2,3-Butanediol production using acetogenic bacteria

3. Results 91

replicate in C. ljungdahlii, the origins of replication from C. perfringens (ori pIP404) and

C. butyricum (ori pCB102) were cloned into plasmid pSB3C5-UUMKS 3. For this purpose, ori

pIP404 and ori pCB102 were amplified (pIP404PstI_fwd, pIP404SalI_rev, pCB102SalI_fwd

pCB102PstI_rev) using vector pJIR750 and pMTL83151 as template, respectively. Afterwards,

plasmid pSB3C5-UUMKS 3 and amplified DNA fragments were digested using restriction

enzymes PstI and SalI. Subsequently, ligation of the digested and purified ori pIP404 and ori

pCB102 into the linearized plasmid pSB3C5-UUMKS 3 was performed. The resulting plasmids

were named pCE3 and pJR3 (Figure 36). Restriction enzyme digestion using PstI and SalI of

pCE3 (Figure 37A; expected DNA fragments 12,495 bp, 2,191 bp) and pJR3 (Figure 37A;

expected DNA fragments 12,495 bp, 1,704 bp) as well as sequencing (pIP404Seq1,

pIP404Seq2, pIP404Seq3, pJR3Seq1, pJR3Seq2, pJR3Seq3, pJR3Seq4) by GATC Biotech AG

(Constance, Germany) confirmed successful cloning. Prior to transformation of the two

butanol plasmids pCE3 and pJR3 into C. ljungdahlii, in vivo methylation with E. coli ER2275

[pACYC184_MCljI] was performed. However, for this purpose, the origin of replication for

Gram-negative bacteria p15A of the respective plasmid had to be exchanged, since pCE3 as

well as pJR3 harbor the identical ori. The pSC101-replicon also shows a very low copy number

just like the p15A-replicon. Therefore, p15A-replicon of in vivo methylation plasmid

Figure 36: Cloning strategy for construction of plasmids pCE3 and pJR3. Plasmid pSB3C5-UUMKS 3 as well as repH (pCB102) and rep (pIP404) were digested using restriction enzymes PstI and SalI. Rep

(pIP404) and repH (pCB102) were ligated into linearized pSB3C5-UUMKS 3, resulting in plasmids pCE3 and pJR3, respectively. Pptb , promoter region upstream of ptb-buk operon from C. acetobutylicum; thlA, thiolase from C. acetobutylicum; hbd, 3-hydroxybutyryl-CoA dehydrogenase from C. acetobutylicum; crt, crotonase from C. acetobutylicum; bcd, butyryl-CoA dehydrogenase from C. acetobutylicum; etfA, electron transfer flavoprotein from C. acetobutylicum; etfB, electron transfer flavoprotein from C. acetobutylicum; adhE2, bifunctional aldehyde/alcohol dehydrogenase from C. acetobutylicum; rep (p15A), p15A-replicon for Gram-negative bacteria; rep (pIP404), pIP404-replicon for Gram-positive bacteria from C. perfringens; repH (pCB102), pCB102-replicon for Gram-positive bacteria from C. butyricum; catP, chloramphenicol resistance gene.

Page 106: 2,3-Butanediol production using acetogenic bacteria

92 3. Results

pACYC184_MCljI was exchanged with pSC101-replicon. For this purpose, rep101/ori (pSC101)

was amplified (pSC101SacII_fwd, pSC101ClaI_rev) using vector pSC101 as template.

Subsequently, plasmid pACYC184_MCljI and amplified DNA fragment were digested using

restriction enzymes Bsu15I and SacII. Afterwards, ligation of the digested and purified ori

pSC101 into the linearized plasmid pACYC184_MCljI was performed and the resulting plasmid

was named pACYC184_MCljI_pSC101 (Figure 38). Restriction enzyme digestion using Bsu15I

and SacII of pACYC184_MCljI_pSC101 (Figure 37B; expected DNA fragments 6,730 bp,

1,597 bp) as well as sequencing (pSC101SacII_fwd) by GATC Biotech AG (Constance, Germany)

Figure 37: Analysis of DNA fragments resulting from analytical restriction enzyme digestion of plasmids pCE3 and pJR3 (A), analytical restriction enzyme digestion of plasmid pACYC184_MCljI_pSC101 (B), and analytical restriction enzyme digestion of plasmids pCE3_traJ and pJR3_traJ (C) using 0.8 % agarose gels. GeneRuler DNA Ladder Mix (M). (A) Analytical restriction enzyme digestion using PstI and SalI of pCE3 with expected DNA fragments 12,495 bp and 2,191 bp (1) as well as pJR3 with expected DNA fragments 12,495 bp and 1,704 bp (2). (B) Analytical restriction enzyme digestion using Bsu15I and SacII of pACYC184_MCljI_pSC101 with expected DNA fragments 6,730 bp and 1,597 bp (1). (C) Analytical restriction enzyme digestion using PstI of pCE3_traJ with expected DNA fragments 14,686 bp and 766 bp (2) as well as pJR3_traJ with expected DNA fragments 14,199 bp and 7,66 bp (1).

Figure 38: Cloning strategy for construction of plasmid pACYC184_MCljI_pSC101. Plasmid pACYC184_MCljI and rep101/ori (pSC101) were digested using restriction enzymes Bsu15I and SacII. Rep101/ori (pSC101) was ligated into linearized pACYC184_MCljI, resulting in plasmid pACYC184_MCljI_pSC101. Pptb, promoter region upstream of ptb-buk operon from C. acetobutylicum; CLJU_c03310, locus tag encoding methylase subunit; CLJU_c03320, locus tag encoding specifity subunit; rep (p15A), p15A-replicon for Gram-negative bacteria; tetR, tetracycline resistance gene; rep101/ori (pSC101), pSC101-replicon for Gram-negative bacteria.

Page 107: 2,3-Butanediol production using acetogenic bacteria

3. Results 93

confirmed successful cloning. Thus, in vivo methylation of pCE3 and pJR3 was achieved using

E. coli ER2275 [pACYC184_MCljI_pSC101] and for plasmid isolation “QIAGEN Plasmid Midi Kit”

(QIAGEN-tip 100; Qiagen GmbH, Hilden, Germany) was applied in order to get high yields of

plasmid concentration. Nevertheless, no successful transformation was achieved by using

electroporation method with competent C. ljungdahlii cells. As a further attempt, conjugation

experiments were performed to transform the butanol plasmids into C. ljungdahlii. Since pCE3

and pJR3 do not contain any genes for conjugal transfer, the oriT/traJ region of vector

pMTL82151 was amplified (traJInfusionPstI_fwd, traJInfusionPstI_rev) and cloned into the

respective plasmids. Plasmids pCE3 and pJR3 were linearized by restriction enzyme digestion

using PstI and oriT/traJ was inserted by “In-Fusion HD cloning”. On the one hand, successful

cloning of pCE3_traJ (Figure 37C, expected DNA fragments 14,199 bp, 766 bp) and pJR3_traJ

(Figure 37C, expected DNA fragments 14,686 bp, 766 bp) was confirmed by analytical

restriction digestion with PstI, and on the other hand by sequencing (pJR3-traJ Seq, traJseq1,

traJSeq_rev) of plasmids by GATC Biotech AG (Constance, Germany). The resulting plasmids

Figure 39: Cloning strategy for construction of plasmids pCE3_traJ and pJR3_traJ. Plasmids pCE3 and pJR3 and DNA fragments were digested with restriction enzyme PstI. OriT/traJ was cloned into linearized pCE3 and pJR3 by “In-Fusion HD Cloning”, resulting in plasmids pCE3_traJ and pJR3_traJ, respectively. Pptb , promoter region upstream of ptb-buk operon from C. acetobutylicum; thlA, thiolase; hbd, 3-hydroxybutyryl-CoA dehydrogenase; crt, crotonase; bcd, butyryl-CoA dehydrogenase; etfA, electron transfer flavoprotein; etfB, (lectron transfer flavoprotein; adhE2, bifunctional aldehyde/alcohol dehydrogenase; rep (p15A), p15A-replicon for Gram-negative bacteria; rep (pIP404), pIP404-replicon for Gram-positive bacteria from C. perfringens; repH

(pCB101), pCB102-replicon for Gram-positive bacteria from C. butyricum; catP, chloramphenicol resistance gene; traJ, gene for conjugal transfer function; oriT, origin of conjugal transfer function.

Page 108: 2,3-Butanediol production using acetogenic bacteria

94 3. Results

were named pCE3_traJ and pJR3_traJ (Figure 39). However, conjugation experiments using

pCE3_traJ and pJR3_traJ for transformation of C. ljungdahlii failed to produce recombinant

strains.

Page 109: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 95

4. Discussion

4.1 Natural microbial 2,3-BD production

Microbial 2,3-BD production is often accompanied by mixed-acid fermentation, a metabolism

found in most members of the family Enterobacteriaceae and some other microorganisms

including both aerobic and anaerobic types. Acidic products of mixed acid fermentation are

acetate, lactate, formate, and succinate, which lower the intracellular pH. Thus, production of

2,3-BD is assumed to prevent excessive acidification by switching the metabolism towards this

neutral compound (Vivijs et al., 2014a; b). Moreover, external supplementation of acids

induces 2,3-BD biosynthesis (Nakashimada et al., 2000; Lee et al., 2016). The reversible

reaction in-between acetoin and 2,3-BD with concomitant NAD(P)H/ NAD(P)+ formation might

also play a role in regulation of the NAD(P)H/NAD(P)+ ratio (Celińska and Grajek, 2009).

Bacteria can reuse 2,3-BD very often in case other carbon and energy sources are scarce, so

2,3-BD production is also considered to be a carbon and energy-storing strategy (Xiao and Xu,

2007). In the past years, several microorganisms including yeasts and bacteria from different

genera were reported to produce 2,3-BD (Ji et al., 2011a). Table 10 shows an overview of

microorganisms capable of 2,3-BD production, utilized substrate, culture method, and titer

since 2009. Since the best natural 2,3-BD producers belong to the risk group 2, research is also

focussing in metabolic engineering of risk group 1 organisms to avoid high safety

requirements, which are unfavorable for industrial-scale processes (Celińska and Grajek, 2009;

Ji et al., 2011a), but in most cases 2,3-BD concentrations do not reach yields of risk group 2

organisms. So far, Serratia marcescens and Klebsiella pneumoniae are the most efficient 2,3-

BD production strains with 152 g/l and 150 g/l, respectively (Table 10). However recently,

some bacteria generally regarded as safe (GRAS) such as Paenibacillus polymyxa, Bacillus

amyloliquefaciens, and Bacillus licheniformis were shown to produce comparable amounts of

2,3-BD as risk group 2 organisms (Table 10). Many 2,3-BD production strains use sugars like

hexoses and pentoses as substrate, which is the reason why more sustainable substrates such

as glycerol, hydrolysates of cellulose, corn stover hydrolysates, and starch have been recently

used for 2,3-BD production (Table 10). In 2011, the publication reporting that acetogens are

capable of 2,3-BD production using steel mill waste gas attracted great attention, although

only producing low amounts of 1.4 mM-2 mM (Köpke et al., 2011b). Furthermore, it was

Page 110: 2,3-Butanediol production using acetogenic bacteria

96 4. Discussion

Table 10: Microbial 2,3-BD production (titer) of different bacterial species and applied substrate and culture condition (since 2009)

Strain Substrate Culture method Titer

(g/l)

Reference

Engineered natural 2,3-BD production strains

Bacillus amyloliquefaciens B10-

127

Glucose Fed-batch 092.3 Yang et al., 2011

B. amyloliquefaciens B10-127 Glucose Fed-batch 061.4 Yang et al., 2012

B. amyloliquefaciens pBG Glucose Fed-batch 132.9 Yang et al., 2013

B. amyloliquefaciens GAR Crude glycerol Fed-batch 102.3 Yang et al.,

2015b

B. amyloliquefaciens TUL-308 Sugarcane

molasses

Fed-batch 060.0 Sikora et al.,

2016

B. atrophaeus NRS-213 Glucose Fed-batch 029.9 Kallbach et al.,

2017

B. licheniformis BL8 Xylose Batch 013.8 Wang et al.,

2012b

B. licheniformis 10-1-A Glucose Fed-batch 115.7 Li et al., 2013

B. licheniformis DSM8785 Glucose Fed-batch 144.7 Jurchescu et al.,

2013

B. licheniformis X-10 Corn stover

hydrolysate

Fed-batch 074.0 Li et al., 2014b

B. licheniformis ATCC

14580

Inulin

hydrolysate

SSF* batch 103.0 Li et al., 2014c

B. licheniformis WX-02 ΔbudC Glucose Shake flask 030.8 Vivijs et al.,

2014b

B. licheniformis NCIMB**

8059

Apple pomace

hydrolysates

Fed-batch 113.0 Białkowska et al.,

2015

B. mojavensis B-14698 Glucose Batch 037.8 Kallbach et al.,

2017

B. subtilis ATCC*** 23857 Glucose Batch 006.1 Biswas et al.,

2012

*simoultaneous saccharification and fermentation

**National Collections of Industrial, Marine and Food Bacteria

***American Type Culture Collection

Page 111: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 97

Table 10: Microbial 2,3-BD production (titer) of different bacterial species and applied substrate and culture condition (continued)

Strain Substrate Culture method Titer

(g/l)

Reference

Engineered natural 2,3-BD production strains

B. subtilis BSF20 Glucose Shake flask 049.3 Fu et al., 2014

B. subtilis AFYL Glucose Shake flask 044.2 Yang et al., 2015c

B. subtilis WN1394 Glucose Batch, IPTG 002.4 De Oliveira and

Nicholson, 2016

B. subtilis TUL 322 Sugarcane

molasses

Fed-batch 075.0 Białkowska et al.,

2016

B. vallismortis B-14891 Glucose Batch 060.4 Kallbach et al.,

2017

Clostridium autoethanogenum

DSM23693

Steel mill gas Batch 000.4 Köpke and Chen,

2013

C. autoethanogenum LZ1561 Synthetic steel

mill gas

Batch 001.1 Köpke and

Mueller, 2016

C. autoethanogenum LZ1561 Synthetic steel

mill gas

Continous

fermentation

009.0 Köpke and

Mueller, 2016

Enterobacter cloacea subsp.

dissolvens SDM

Cassava

powder

SSF* batch 093.9 Wang et al.,

2012a

E. cloacea SDM 09 Glucose Fed-batch 119.4 Li et al., 2015a

E. cloacae CGMCC** 605 Sugarcane

molasses

Fed-batch 099.5 Dai et al., 2015

E. aerogenes EMY 01 Glucose Fed-batch 118.1 Jung et al., 2012

E. aerogenes EMY-68 Sugarcane

molasses

Fed-batch 098.7 Jung et al., 2015

E. aerogenes EMY-70S Sugarcane

molasses

Fed-batch 140.0 Jung et al., 2015

Klebsiella pneumoniae SDM Glucose Fed-batch 150.0 Ma et al., 2009

*simoultaneous saccharification and fermentation

**China General Microbiological Culture Collection Center

Page 112: 2,3-Butanediol production using acetogenic bacteria

98 4. Discussion

Table 10: Microbial 2,3-BD production (titer) of different bacterial species and applied substrate and culture condition (since 2009) (continued)

Strain Substrate Culture

method

Titer

(g/l)

Reference

Engineered natural 2,3-BD production strains

K. pneumoniae G31 Glycerol Fed-batch 049.2 Petrov and

Petrova, 2009

K. pneumoniae CICC* 10011 Jerusalem

artichoke tuber

**SSF batch 084.0 Sun et al., 2009a

K. pneumoniae CICC* 10011 Jerusalem

artichoke stalk

and tuber

**SSF batch 067.4 Li et al., 2010a

K. pneumoniae G31 Glycerol Fed-batch 070.0 Petrov and

Petrova, 2010

K. pneumoniae SDM Corncob

molasses

Fed-batch 078.9 Wang et al., 2010

K. pneumoniae SGJSB04 Glucose Fed-batch 101.5 Kim et al., 2012

K. pneumoniae KG-rs Glucose Shake flask 024.5 Guo et al., 2014b

K. pneumoniae ΔadhEΔldhA Glucose Fed-batch 115.0 Guo et al., 2014a

K. pneumoniae KMK-05 Glucose Shake flask 031.1 Jung et al., 2014

K. pneumoniae SGSB105 Glucose Fed-batch 090.0 Kim et al., 2014a

K. pneumoniae G31-A Starch **SSF batch 053.8 Tsvetanova et al.,

2014

K. pneumoniae M3 Crude glycerol Fed-batch 131.5 Cho et al., 2015a

K. pneumoniae SGSB112 Glucose Fed-batch 061.0 Lee et al., 2015

K. oxytoca ME-UD-3 Glucose Batch 095.5 Ji et al., 2009

K. oxytoca ACCC 10370 Corncob acid

hydrolysate

Fed-batch 035.7 Cheng et al., 2010

K. oxytoca ME-XJ-8 Glucose Fed-batch 130.0 Ji et al., 2010

K. oxytoca ME-CRPin Glucose, Xylose Shake flask 023.9 Ji et al., 2011b

K. oxytoca ME-UD-3 Glucose Fed-batch 127.9 Nie et al., 2011

K. oxytoca GSC12206 (ΔldhA) Glucose Fed-batch 115.0 Kim et al., 2013a

*China Center of Industrial Culture Collection

**simoultaneous saccharification and fermentation

Page 113: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 99

Table 10: Microbial 2,3-BD production (titer) of different bacterial species and applied substrate and culture condition (since 2009) (continued)

Strain Substrate Culture method Titer

(g/l)

Reference

Engineered natural 2,3-BD production strains

K. oxytoca ΔldhAΔpflB Glucose Fed-batch 113.0 Park et al., 2013

K. oxytoca M1 Glucose Fed-batch 142.5 Cho et al., 2015b

K. oxytoca KMS005-73T Glucose Fed-batch 117.4 Jantama et al.,

2015

K. oxytoca

ΔldhAΔpflBΔbudC::PBDH

Glucose Fed-batch 106.7 Park et al., 2015

Paenibacillus polymyxa ZJ-9 Jerusalem

artichoke

tuber

Batch 036.9 Gao et al., 2010

P. polymyxa DSM365 Sucrose Fed-batch 111.0 Häßler et al.,

2012

Serratia marcescens H30 Sucrose Fed-batch 139.9 Zhang et al.,

2010a

S. marcescens H30 Sucrose Fed-batch 152.0 Zhang et al.,

2010b

S. marcescens ΔslaC-bdhA Sucrose Fed-batch 089.8 Bai et al., 2015

Natural 2,3-BD production strains

Raoultella planticola CECT* 843 Pure glycerol Batch 030.5 Ripoll et al., 2016

R. terrigena CECT* 4519 Raw glycerol Batch 033.7 Ripoll et al., 2016

Construction of pathway for 2,3-BD production

Clostridium acetobutylicum

pWUR460

Glucose Batch 002.0 Siemerink et al.,

2011

Clostridium sp. MBD136 Syngas Continous

fermentation

009.2 Tyurin and

Kiriukhin, 2013

Corynebacterium glutamicum

ATCC**13032

Glucose Two-stage

fermentation

006.3 Radoš et al., 2015

*The Spanish type culture collection of microorganisms

**American Type Culture Collection

Page 114: 2,3-Butanediol production using acetogenic bacteria

100 4. Discussion

Table 10: Microbial 2,3-BD production (titer) of different bacterial species and applied substrate and culture condition (since 2009) (continued)

Strain Substrate Culture

method

Titer

(g/l)

Reference

Construction of pathway for 2,3-BD production

C. glutamicum SGSC102 Glucose/

cassava powder

Fed-batch 018.9/

012.0

Yang et al., 2015a

Escherichia coli JCL260 Glucose Shake flask 006.1 Yan et al., 2009

E. coli JM109LPAP Glucose Batch 014.5 Li et al., 2010b

E. coli YYC202ΔldhAΔilvC Glucose Shake flask 001.1 Nielsen et al.,

2010

E. coli SGSB03 Glucose Batch 015.7 Lee et al., 2012

E. coli SGSB03 Crude glycerol Batch 006.9 Lee et al., 2012

E. coli BW25113/ΔackA Glycerol Shake flask 009.6 Shen et al., 2012

E. coli Cellodextrin SSF* batch 005.5 Shin et al., 2012

E. coli DSM02 Algal

hydrolysate

Fed-batch 014.1 Mazumdar et al.,

2013

E. coli Diacetyl Fed-batch 026.8 Wang et al.,

2013b

E. coli (pETDuet-bdhfdh) Diacetyl Fed-batch 031.7 Wang et al.,

2013b

E. coli mlc-XT7-LAFC-YSDXL Glucose + xylose Shake flask 054.0 Nakashima et al.,

2014

E. coli BL21/pET-RABC Glucose Fed-batch 073.8 Xu et al., 2014

E. coli budBbudC Glucose Shake flask 002.2 Chu et al., 2015

E. coli MQ1 Glucose Fed-batch 115.0 Ji et al., 2015

E. coli S30 Glucose Lysate with

NAD+, ATP

082.0 Kay and Jewett,

2015

E. coli YJ2 Glucose Shake flask 030.5 Tong et al., 2016

Sacharomyces cerevisiae B2C-

a1a3a5

Glucose Shake flask 002.3 Ng et al., 2012

*simoultaneous saccharification and fermentation

Page 115: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 101

Table 10: Microbial 2,3-BD production (titer) of different bacterial species and applied substrate and culture condition (since 2009) (continued)

Strain Substrate Culture method Titer

(g/l)

Reference

Construction of pathway for 2,3-BD production

S. cerevisiae BD4 Glucose Fed-batch 096.2 Kim et al., 2013b

S. cerevisiae BD4X Xylose Fed-batch 043.6 Kim et al., 2014b

S. cerevisiae JL0432 Glucose +

galactose

Fed-batch 100.0 Lian et al., 2014

S. cerevisiae BD5_t2nox Glucose Shake flask ~31.0 Kim et al., 2015

Synechococcus elongatus CO2 shake flask 002.4 Oliver et al., 2013

Synechocystis sp. PCC6830 CO2 shake flask 000.4 Savakis et al.,

2013

reported that LanzaTech uses already 2,3-BD produced by C. autoethanogenum from steel mill

waste gas as a precursor for the production of 1,3-butadiene in pilot/demo scale (Bengelsdorf

et al., 2013; Köpke and Havill, 2014). 2,3-BD formation from CO and water is coupled to CO2

formation (1).

11 CO + 5 H2O → C4O2H10 + 7 CO2 (1)

Acetogens are unique organisms since they can fix CO2 and/or CO via the non-circular Wood-

Ljungdahl pathway (Figure 2), which is considered to be one of the oldest metabolic pathways

in life (Russell and Martin, 2004). Moreover, it is assumed to be the most efficient carbon

fixation pathway since it acts in a linear manner (Fast and Papoutsakis, 2012). There are over

100 different acetogens, including at least 25 different genera (Daniell et al., 2016) while the

best characterised species either belong to the group of Gram-positive bacteria with a low GC

content (e.g. C. ljungdahlii, C. autoethanogenum, C. ragsdalei, C. coskatii, A. woodii,

Clostridium aceticum, and M. thermoacetica) (Schiel-Bengelsdorf and Dürre, 2012) or to

anaerobic spirochaetes, which live as symbionts in the gut of termites (Fuchs, 2007).

Acetogens can use a variety of different substrates including C1-compounds such as formate,

CO, CO2, and methanol (Küsel and Drake, 2005), sugars (hexoses and pentoses), glycerol and

cellulose (Olson et al., 2012), carboxylic acids (pyruvate, lactate), aldehydes (benzaldehyde,

glyoxylate), and many more (Drake et al., 2008). Moreover, acetogenic clostridia were shown

Page 116: 2,3-Butanediol production using acetogenic bacteria

102 4. Discussion

to utilize electricity as energy source in order to fix CO2 (Nevin et al., 2010; Lovley, 2011; Nevin

et al., 2011; Lovley and Nevin, 2013). Besides the main product acetate of the Wood-Ljungdahl

pathway, acetogens can also produce ethanol, lactate, butyrate, butanol, hexanol, hexanoate

and 2,3-BD (Liou et al., 2005; Drake et al., 2008; Köpke et al., 2011b; Phillips et al., 2015; Li et

al., 2015b). The acetogens A. woodii, M. thermoacetica, and C. aceticum mainly are considered

to produce acetate as sole end product, while B. methylotrophicum and C. carboxidivorans are

capable of producing butanol as by-product, and C. ljungdahlii, C. autoethanogenum,

C. ragsdalei, and C. coskatii are mainly considered for ethanol and 2,3-BD production besides

acetate (Daniell et al., 2016). In this study, C. ljungdahlii, C. ragsdalei, and C. autoethanogenum

were first investigated for natural 2,3-BD production with synthetic syngas as energy and

carbon source and furthermore C. aceticum and C. carboxidivorans were also tested for

natural 2,3-BD production (Figure 5). Since these organisms produce 94 % of 2R,3R-BD, while

only 6 % of meso-2,3-BD is formed (Köpke et al., 2011b), only the D(-)-form was investigated

in analytics.

The growth experiment was conducted to find the best natural 2,3-BD production strain in

order to work on genetic engineering leading to enhanced 2,3-BD production. However, 2,3-

BD production was only detected in the organsims C. ljungdahlii, C. autoethanogenum, and

C. ragsdalei, while C. ljungdahlii produced the highest 2,3-BD concentration of 4.8 mM

(Figure 5B). Furthermore, Köpke et al. (2011) reported that the genes, which are involved in

2,3-BD formation in C. autoethanogenum are upregulated massively during stationary growth

phase (15-fold increased to normalized mRNA level). This is also the growth phase, when the

majority of 2,3-BD is produced. Furthermore, only gene bdh was shown to be expressed

constitutively at high normalized mRNA levels over the whole period of growth (Köpke et al.,

2011b). In this study, formation of 2,3-BD in the stationary growth phase was also shown for

the organisms C. ljungdahlii, C. autoethanogenum, and C. ragsdalei. Thus, it might be

concluded that formation of the neutral compound 2,3-BD in acetogens also represents a

protection mechanism of both excessive acidification and excessive reducing equivalents, like

speculated for other 2,3-BD producing bacteria (Ji et al., 2011a). Therefore, 2,3-BD production

starts at the transition to the stationary growth phase, when pH decreases due to acetate

formation and a potential excess of reducing equivalents is present.

Page 117: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 103

4.2 Enhancement of 2,3-BD production using acetogens

4.2.1 Enhancement of 2,3-BD production via overexpression of alsS, budA, and bdh

Since testing of natural 2,3-BD production in some acetogens resulted in C. ljungdahlii showing

the highest 2,3-BD concentration (4.8 mM), this organism was chosen for further

enhancement of 2,3-BD production. C. ljungdahlii (Figure 3A) was isolated from chicken yard

waste in 1988 (Barik et al., 1988). It is a Gram-positive, rod-shaped, spore-forming, motile,

mesophilic, and obligate anaerobic bacterium with a GC content of 31 mol % (Tanner et al.,

1993). Moreover, it can grow autotrophically using H2 + CO2 and CO, but it also uses substrates

such as formate, ethanol, pyruvate, fumarate as well as sugars such as fructose and xylose for

heterotrophic growth (Tanner et al., 1993). A few years later, the description of a very close

relative (genome similarity of 99.3 %; Bengelsdorf et al., 2016) named C. autoethanogenum

was published, which was isolated from rabbit feces (Abrini et al., 1994). Furthermore, the

organism C. coskatii, which was was isolated from estuary sediment collected from Coskata-

Coatue Wildlife Refuge (Zahn and Saxena, 2011), was shown to be a very close relative of

C. ljungdahlii, C. autoethanogenum, and C. ragsdalei (Bengelsdorf et al., 2016). Therefore,

C. coskatii was considered to be a natural 2,3-BD producer. In 2011, C. ljungdahlii as well as

C. autoethanogenum were reported to produce 2,3-BD from steel mill waste gas (Köpke et al.,

2011b). Furthermore, toxicity tests using external addition of 2,3-BD with

C. autoethanogenum showed that growth and acetate production ceased with addition of

40 to 50 g/l 2,3-BD (Köpke et al., 2011b). This shows that C. ljungdahlii might also be

considered as a potential commercial 2,3-BD production strain displaying a high 2,3-BD

tolerance. Another advantage of choosing this organism for enhancement of 2,3-BD

production is the efficient and reproducible transformation protocol, enabling genetic

engineering of C. ljungdahlii (Leang et al., 2013). Furthermore, the acetogen C. coskatii, which

was not tested for natural 2,3-BD production by this time, harbors the same genes alsS

(CCOS_39110), budA (CCOS_38400), and bdh (CCOS_06940) for 2,3-BD formation as

C. ljungdahlii (CLJU_c38920, CLJU_c08380, CLJU_c23220) and C. autoethanogenum

(CAETHG_1740, CAETHG_2932, CAETHG_0385). This organism produces mainly acetate but

also ethanol is formed as side product from syngas (Zahn and Saxena, 2011). Overexpression

of the genes alsS, budA, and bdh encoding the acetolactate synthase (AlsS), acetolactate

decarboxylase (BudA), and 2,3-BD dehydrogenase (Bdh) might enhance natural 2,3-BD

Page 118: 2,3-Butanediol production using acetogenic bacteria

104 4. Discussion

production of C. ljungdahlii and C. coskatii using glycolysis and/or Wood-Ljungdahl pathway.

The enzyme AlsS converts pyruvate to S-acetolactate, which is further decarboxylated to R-

acetoin by BudA (Figure 6), while acetoin is a molecule, which is extremely prone to

spontaneous racemization via an enolate intermediate (Köpke et al., 2014). Thus, it is possible

that small amounts of R-acetoin are in vivo spontaneously racemized by enzyme BudA. Finally,

the enzyme Bdh reduces R-acetoin to 2R,3R-BD and S-actoin to meso-2,3-BD, respectively

(Figure 6). The three genes alsS, budA, and bdh were synthesized in form of an operon under

control of the constitutive promoter Ppta-ack and cloned into two E. coli/Clostridium shuttle-

vectors namely pMTL82151 and pJIR750. These vectors harbor different replicons for Gram-

positive bacteria (pMTL82151, pBP1 from C. botulinum; pJIR750, pIP404 from C. perfringens),

which were shown to yield good transformation efficiencies for C. ljungdahlii (Leang et al.,

2013). Since C. coskatii is a very close relative of C. ljungdahlii (Bengelsdorf et al., 2016), the

same transformation protocol was successfully applied for transformation of this organism.

Furthermore, the pBP1-replicon and pIP404-replicon might influence the plasmid copy

number differently inC. ljungdahlii as well as C. coskatii and thus, lead to changes in 2,3-BD

production. After transformation of the vectors pMTL82151 and pJIR750 as well as the 2,3-BD

production plasmids pMTL82151_23BD and pJIR750_23BD into C. ljungdahlii and C. coskatii,

growth experiments with fructose as well as syngas as energy and carbon source were

conducted to provide insights into growth and production pattern in comparison to the

respective wildtype strains.

Growth experiments using fructose as carbon source showed only slight differences in terms

of growth and production pattern of the recombinant C. ljungdahlii strains compared to

C. ljungdahlii WT (Figure 9). The strain C. ljungdahlii [pJIR750_23BD] produced the highest

2,3-BD concentration of 1.2 mM, which is 1.5-fold higher compared to C. ljungdahlii WT

(0.8 mM). Moreover, C. ljungdahlii [pMTL82151_23BD] showed a slight increase in 2,3-BD

production (1.1 mM) compared to C. ljungdahlii WT. The growth experiment using syngas as

energy and carbon source (Figure 10) confirmed the results of the heterotrophic growth

experiment (Figure 9) in C. ljungdahlii [pJIR750_23BD] being the best 2,3-BD production strain.

This strain produced the highest concentration of 2,3-BD (6.4 mM), which is a 1.5-fold increase

compared to C. ljungdahlii WT (4.3 mM). In contrast, C. ljungdahlii [pMTL82151_23BD] did not

produce higher amounts of 2,3-BD (3.8 mM) compared to C. ljungdahlii WT growing with

Page 119: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 105

syngas. This is contrary to the results of another autotrophic growth experiment with

C. ljungdahlii [pMTL82151_23BD] (Figure 24, Table 7), showing a 1.8-fold increase in 2,3-BD

production (3.1 mM) compared to C. ljungdahlii WT (1.7 mM). Furthermore, it was striking

that all C. ljungdahlii strains except C. ljungdahlii [pMTL82151_23BD] showed higher product

formation and thus higher 2,3-BD concentrations in the autotrophic growth experiment

depicted in Figure 10, compared to the other autotrophic growth experiment (Figure 24). This

might be explained due to higher overpressure of syngas (1.8 bar) conducted in this

experiment (Figure 10) compared to 1.0 bar (Figure 24). It was shown in various fermentation

experiments that increasing pressure enhances gas solubility and improves mass transfer and

therefore production (Ungerman and Heindel, 2007). The gas-to-liquid mass transfer of

substrate into the culture medium represents the rate-limiting step in the process of gas

fermentation due to the low solubility of CO and H2 in water (Daniell et al., 2016). This is in

accordance to one of the gas laws formulated by William Henry in 1803, which states that the

solubility of a gas in a liquid is directly proportional to the pressure of the gas above the liquid.

Furthermore, LanzaTech considered that providing sufficient or elevated levels of CO during

fermentation process, leads to production of higher energy products, such as 2,3-BD (Simpson

et al., 2011). Moreover, Table 10 shows that the same C. autoethanogenum LZ1561 strain

produced 1.1 g/l 2,3-BD during batch fermentation, while 9.0 g/l 2,3-BD (eight-fold increase)

were formed during continuous fermentation. This shows that the 2,3-BD production of

C. ljungdahlii [pJIR750_23BD] might be further enhanced via continuous fermentation

experiments. Another important observation in the growth experiments using C. ljungdahlii is

the fact that although overexpression of the genes for 2,3-BD formation was realized using a

constitutive promoter from C. ljungdahlii, 2,3-BD production did not start earlier during

growth compared to C. ljungdahlii WT. 2,3-BD production was still formed during late

stationary growth phase, which would support the assumption that 2,3-BD production is

heavily dependent on either acetate formation or excess of reducing equivalents.

Regarding heterotrophic growth experiments of C. coskatii strains using fructose as carbon

source, it was striking that none of the recombinant C. coskatii strains produced higher

2,3-BD concentrations compared to C. coskatii WT (Figure 12). Furthermore, C. coskatii WT

(1.1 mM) and C. coskatii [pJIR750_23BD] (1.0 mM) were the only strains with 2,3-BD

concentrations above the detection limit. Moreover, all C. coskatii strains showed lower

Page 120: 2,3-Butanediol production using acetogenic bacteria

106 4. Discussion

fructose consumption compared to C. ljungdahlii strains. The recombinant C. coskatii strains

did only consume 20.7-26.4 mM fructose compared to recombinant C. ljungdahlii strains

(40-30 mM) growing in Tanner medium. Additionally, C. coskatii required higher yeast extract

concentrations (2 g instead of 0.5 g) to reach a comparable OD600nm in Tanner medium as

C. ljungdahlii. These heterotrophic growth experiments already indicated that C. coskatii is not

an efficient 2,3-BD production strain, nevertheless, C. coskatii WT and the recombinant

C. coskatii strains were also investigated under autotrophic growth conditions using syngas as

energy and carbon source. All C. coskatii strains showed decreased OD600nm values of 0.5-0.8

(Figure 13) compared to C. ljungdahlii strains (0.7-1.8) (Figure 10) but this did not lead to a

decreased acetate production. In contrast, ethanol production was significantly decreased in

all C. coskatii strains compared to C. ljungdahlii strains growing with syngas. In

C. autoethanogenum and other acetogens it was shown that the enzyme aldehyde:ferredoxin

oxidoreductase (AOR) plays an important role in ethanol formation (Fast and Papoutsakis,

2012; Bertsch and Müller, 2015; Mock et al., 2015; Richter et al., 2016; Liew et al., 2017). The

enzyme AOR catalyzes the reversible reduction of an acid to its corresponding aldehyde

(White et al., 1989). In several acetogens, undissociated acetic acid is reduced to acetaldehyde

(Napora-Wijata et al., 2014), which is further metabolised to ethanol via a monofunctional

alcohol dehydrogenase (Adh) or a bifunctional aldehyde/alcohol dehydrogenase (AdhE)

(Figure 41). This pathway was postulated recently (Bertsch and Müller, 2015) alongside the

classic pathway of ethanol formation via the bifunctional AdhE (Figure 41). One great

advantage of the ethanol pathway via AOR is that one mol ATP is generated by the enzyme

acetate kinase (Ack) via substrate-level phosphorylation during acetate formation. Moreover,

proteome studies of C. ljungdahlii revealed that both AORs (aor1 CLJU_c20110, aor2

CLJU_c20210) and Adh (adh2, CLJU_c39950) were highly abundant during syngas

fermentation under acidogenesis as well as solventogenesis (Richter et al., 2016), which is in

accordance to the assumption that ethanol formation via AOR is more advantageous due to

ATP generation. Thus, ethanol production can immediately take place, as soon as overflow

reactants such as undissociated acetic acid and reducing equivalents reach a critical

concentration (Richter et al., 2016). This is the case when an excess of acetyl-CoA as well as

reducing equivalents cannot be further utilized in biomass formation (Richter et al., 2016).

Furthermore, Richter et al. (2016) postulated that ethanol production is exclusively achieved

via reduction of undissociated acetic acid during fermentation of syngas. The genomes of

Page 121: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 107

C. autoethanogenum (aor1 CAETHG_0092, aor2 CAETHG_0102) as well as C. ljungdahlii (aor1

CLJU_c20110, aor2 CLJU_c20210) contain two aor genes and it turned out that these genes

are missing in C. coskatii (Bengelsdorf et al., 2016). This would explain why all C. coskatii strains

produced significantly less ethanol under autotrophic growth conditions. Furthermore, the

recombinant C. coskatii [pMTL82151_23BD] strain produced higher ethanol concentrations

(15.7 mM/7.6 mM) compared to C. coskatii WT (6.1 mM/2.3 mM). In contrast, no higher

2,3-BD amounts were detected growing with fructose as carbon source as well as with syngas,

respectively. It might be assumed that flux towards 2,3-BD originating from pyruvate via

enzymes AlsS, BudA, and Bdh is not sufficient and in this case Bdh might rather be converting

acetaldehyde to ethanol than acetoin to 2,3-BD. The enzyme Bdh from C. ljungdahlii was

characterized recently but acetaldehyde was not tested as possible substrate of this enzyme

(Tan et al., 2015). Moreover, the primary-secondary alcohol dehydrogenase (sAdh) from

C. autoethanogenum (CAETHG_0553), which is also present in C. ljungdahlii (CLJU_c24860), is

not only capable of reducing acetoin to 2,3-BD, but it can also utilize acetaldehyde as substrate

(Köpke et al., 2014). Thus, it cannot be excluded that Bdh from C. ljungdahlii is also capable of

reducing acetaldehyde to ethanol. In this case, Bdh would play the same role as ethanol

formation via AOR in acetogens: getting rid of excess of reducing equivalents when growth is

limited during stationary growth phase and biomass can no longer be produced. However,

C. coskatii [pJIR750_23BD] harboring the same synthetic 2,3-BD operon, but a different

replicon (pIP404 from C. perfringens) for Gram-postive bacteria did not show the effect of

higher ethanol production. This might indicate that the pBP1-replicon from C. botulinum

(pMTL82151) leads to higher plasmid copy numbers in C. coskatii compared to pIP404-replicon

(pJIR750). Nevertheless, this assumption needs to be proven by experimental data such as

investigation of transcript alsS-budA-bdh from C. coskatii by Northern blot analysis or

determination of plasmid copy numbers of pMTL82151_23BD and pJIR750_23BD in C. coskatii

via qRT-PCR. In summary, heterotrophic as well as autotrophic growth experiments with

C. coskatii strains showed that this acetogen is not a suitable 2,3-BD production strain.

4.2.2 Modifications of 2,3-BD production plasmid to optimize 2,3-BD production

For further enhancement of 2,3-BD production using C. ljungdahlii, several modifications of

the 2,3-BD production plasmid pJIR750_23BD were conducted. Firstly, the sequence of budA

from the synthetic 2,3-BD operon alsS-budA-bdh was shortened at the 3'-end of budA and

Page 122: 2,3-Butanediol production using acetogenic bacteria

108 4. Discussion

cloned into the plasmid pJIR750_23BD to avoid a potential rho-independent terminator

structure (Figure 14). Although the sequence of the terminator structure does not show a

poly-A tail and the Gibbs free energy is rather low (-9.3 kcal/mol; Figure 14) compared to most

rho-independent terminator structures, for example of the Gram-positive bacterium B. subtilis

(~15 kcal/mol; De Hoon et al., 2005), abortion of transcription due to this hairpin loop cannot

be excluded. As a second modification of pJIR750_23BD, the gene bdh of the alsS-budA-bdh

operon was exchanged with adh (CLJU_c24860), encoding the primary-secondary alcohol

dehydrogenase sAdh from C. ljungdahlii and the identical gene is also present in

C. autoethanogenum. This enzyme was shown to be capable of reducing S-acetoin to 2R,3R-

BD as well as acetone to isopropanol in C. autoethanogenum (Köpke et al., 2014). If enzyme

sAdh shows a better substrate specifity or a higher enzymatic activity compared to Bdh this

would also lead to a higher 2,3-BD production in C. ljungdahlii. As a third option for

modification of pJIR750_23BD, the budABC operon from Raoultella terrigena was cloned into

the vector pJIR750 under control of Ppta-ack from C. ljungdahlii and Tsol-adc from

C. acetobutylicum. This organism was first described as K. terrigena in 1981 and it was

reported as a Gram-negative bacterium, which can be found in water and soil (Izard et al.,

1981). In 2001, it was renamed R. terrigena together with Raoultella planticola and Raoultella

ornithinolytica due to 16S rDNA and rpoB gene sequencing studies (Drancourt et al., 2001).

Via overexpression of the 2,3-BD operon budABC from R. terrigena with a constitutive

promoter from C. ljungdahlii and a terminator from C. acetobutylicum, 2,3-BD overproduction

using heterologous genes might be achieved in C. ljungdahlii (Figure 40). Investigation of the

acetoin reductase (Acr) encoded by budC from R. terrigena via BLAST analysis on protein level

revealed high similarity (95 %) to Acr from K. pneumoniae. This enzyme installs a S

stereocenter during reduction of R-acetoin, which leads to production of meso-2,3-BD (Yan et

al., 2009; Zhang et al., 2012a). Thus, it is very likely that Acr from R. terrigena also installs a

S-stereocenter, which would lead to formation of meso-2,3-BD in C. ljungdahlii, if correct

transcription and translation of the heterologous genes takes place. Transformation of the

modified pJIR750_23BD plasmids into C. ljungdahlii via electroporation was only successful

after in vivo methylation using E. coli ER2275 [pACYC184_MCljI]. This plasmid contains the

modification (methylase) subunit (CLJU_c03310) and the specifity subunit (CLJU_c03320) from

the C. ljungdahlii restriction-modification (R-M) system (Matthias Beck, unpublished).

R-M systems degrade exogenous DNA and thus previous in vivo methylation of plasmids might

Page 123: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 109

be beneficial to overcome restriction barriers of the organism of choice by stabilizing

transformed DNA and therefore boosting transformation efficiency (Chen et al., 2008; Suzuki,

2012). Although the modified plasmids are smaller in size (10,867 bp-10,342 bp) compared to

pJIR750_23BD (10,960 bp), it might be possible that the batch of electrocompetent

C. ljundahlii cells was not as efficient as the one used in transformation of pJIR750_23BD and

therefore in vivo methylation enhanced transformation of the modified pJIR750_23BD

plasmids into C. ljungdahlii. After successful transformation, the recombinant C. ljungdahlii

strains harboring the modified 2,3-BD production plasmid were investigated in terms of

growth behavior and production pattern under heterotrophic and autotrophic growth

conditions.

In growth experiments using fructose as carbon source C. ljungdahlii [pJIR750_23BD] showed

decreased growth with an OD600nm of 0.9 compared to C. ljungdahlii WT (OD600nm 2.7), but also

a higher ethanol concentration (29.5 mM) compared to C. ljungdahlii WT (8.5 mM) (Figure 23).

This might be explained by analytical problems either in acetate detection via HPLC or ethanol

analysis via GC, since calculation of carbon recovery shows that carbon concentration of

product formation (266.6 mM) exceeds carbon concentration of substrate consumption

(237.6 mM). The strain C. ljungdahlii [pJIR750_23BD] did not produce higher ethanol

concentrations in the other growth experiment with fructose (Figure 9), which suggests that

problems in ethanol analytics occurred. Regarding the recombinant strains harboring the

Figure 40: 2,3-BD production in C. ljungdahlii with genes originating from R. terrigena starting from pyruvate with budB (acetolactate synthase), budA (acetolactate decarboxylase), and budC (acetoin reductase).

Page 124: 2,3-Butanediol production using acetogenic bacteria

110 4. Discussion

modified 2,3-BD production plasmids, only C. ljundahlii [pJIR750_23BD_budAshort_adh] as

well as C. ljungdahlii [pJIR750_budABCoperon] achieved slightly increased 2,3-BD

concentrations of 2.1 mM and 1.9 mM compared to C. ljungdahlii WT (1.8 mM). The growth

experiment using syngas as energy and carbon source (Figure 24) showed a slightly different

picture, especially regarding 2,3-BD production. Under these conditions, C. ljungdahlii

[pJIR750_23BD_budAshort_adh] showed the highest 2,3-BD production of 3.3 mM, which is

1.9-fold higher compared to C. ljungdahlii WT (1.7 mM). Recently, it was shown that sAdh of

C. autoethanogenum displays a lower enzyme activity (0.4 U/mg) compared to Bdh (0.8 U/mg)

(Köpke et al., 2016). However, overexpression of adh in C. ljungdahlii

[pJIR750_23BD_budAshort_adh] led to similar 2,3-BD concentrations (3.3 mM) compared to

C. ljungdahlii [pJIR750_23BD] (2.9 mM) as well as C. ljungdahlii [pMTL82151_23BD] (3.1 mM),

in which bdh is overexpressed. This shows that lower enzyme activity of sAdh is either not

present in C. ljungdahlii or it does not have a massive effect on 2,3-BD production. The strain

C. ljungdahlii [pJIR750_23BD_budAshort] did not produce higher 2,3-BD amounts compared

to C. ljungdahlii [pJIR750_23BD] neither using fructose (1.5 mM vs. 2.7 mM) nor syngas

(2.8 mM vs. 2.9 mM) as energy and carbon source. This shows that the potential rho-

independent terminator structure downstream of budA does not seem to affect expression of

the synthetic 2,3-BD operon. Moreover, the recombinant strain C. ljungdahlii

[pJIR750_budABCoperon], harboring the heterologous genes for 2,3-BD production from

R. terrigena produced similar 2,3-BD concentrations (1.8 mM) compared to C. ljungdahlii WT

(1.7 mM). Thus, neither using fructose as carbon source nor syngas as energy and carbon

source led to higher meso-2,3-BD production in C. ljungdahlii [pJIR750_budABCoperon]

compared to the natural 2R,3R-BD production of C. ljungdahlii WT. The budABC operon from

R. terrigena was cloned directly into the vector pJIR750, but comparison of the GC content of

R. terrigena (56.7 %) with C. ljungdahlii (31 %) shows that there is a major difference. Also, the

codon usage table of R. terrigena compared to C. ljungdahlii (Table 11) displays differences in

the abundance of triplets coding for the different amino acids. The complete codon usage

table is presented in Table 13 (chapter 7. Supplement). For example, Table 11 shows that

C. ljungdahlii prefers GCA triplet to encode alanine, whereas R. terrigena uses GCC triplet.

Further differences are present for amino acids glycine, isoleucine, leucine, asparagine,

proline, glutamine, arginine, serine, threonine, and valine. Thus, it is possible that codon usage

Page 125: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 111

optimization of the budABC operon from R. terrigena would lead to higher 2,3-BD

concentrations in C. ljungdahlii, due to faster translation rates or higher translation efficiency.

Table 11: Comparison of codon usage in R. terrigena and C. ljungdahlii

Amino acid (abbr.) Triplet Abundance [%] R. terrigena* Abundance [%] C. ljungdahlii**

Alanine (A) GCA 18 41

GCC 30 15

GCG 29 6

GCT 23 38

Cysteine (C) UGC 62 44

UGU 38 56

Glycine (G) GGA 12 42

GGC 39 16

GGG 18 15

GGU 30 27

Isoleucine (I) AUA 12 43

AUC 35 17

AUU 53 40

Leucine (L) CUA 10 14

CUC 8 7

CUG 36 10

CUU 14 20

UUA 17 32

UUG 14 17

Asparagine (N) AAC 50 27

AAT 50 73

Proline (P) CCA 21 38

CCC 15 15

CCG 37 8

CCU 27 39

*according to Codon usage database

(http://www.kazusa.or.jp/codon/cgi-bin/showcodon.cgi?species=577&aa=1&style=N, 14th of June,

2017.)

**according to Anja Poehlein, Göttingen Genomics Laboratory (Göttingen, Germany)

Page 126: 2,3-Butanediol production using acetogenic bacteria

112 4. Discussion

Table 11: Comparison of codon usage in R. terrigena and C. ljungdahlii (continued)

Amino acid (abbr.) Triplet Abundance [%] R. terrigena* Abundance [%] C. ljungdahlii**

Glutamine (Q) CAA 37 61

CAG 63 39

Arginine (R) AGA 13 48

AGG 8 30

CGA 11 6

CGC 30 4

CGG 10 5

CGU 29 7

Serine (S) AGC 24 14

AGU 15 21

UCA 14 21

UCC 14 16

UCG 17 3

UCU 16 25

Threonine (T) ACA 17 36

ACC 38 19

ACG 29 7

ACU 15 38

Valine (V) GUA 14 38

GUC 22 12

GUG 32 17

GUU 32 33

*according to Codon usage database

(http://www.kazusa.or.jp/codon/cgi-bin/showcodon.cgi?species=577&aa=1&style=N, 14th of June,

2017.)

**according to Anja Poehlein, Göttingen Genomics Laboratory (Göttingen, Germany)

4.2.3 PFOR enzyme as potential bottle-neck in 2,3-BD production?

In order to further enhance natural 2,3-BD production in C. ljungdahlii, the gene nifJ encoding

the pyruvate:ferredoxin oxidoreductase (PFOR) from C. ljungdahlii was overexpressed. This

organism harbors two nifJ genes (CLJU_c29340 and CLJU_c09340), which are also present in

C. autoethanogenum. Köpke et al. (2011) reported that in C. autoethanogenum only one nifJ

gene (CAETHG_0928) is constituvely expressed at a high level, while the other (CAETHG_3029)

Page 127: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 113

is only upregulated at the end of the growth in a batch. This was the reason for choosing the

homologous gene from C. ljungdahlii (CLJU_c09340) for overexpression in this organism.

Moreover, it was shown that PFOR exhibits a lower enzyme activity (0.11 U/mg) in

C. autoethanogenum compared to Bdh and sAdh (1.2 U/mg) and therefore conversion of

acetyl-CoA to pyruvate seems to be the rate limiting step in 2,3-BD formation (Köpke et al.,

2016). They concluded that overexpression of nifJ together with alsS and budA is sufficient to

remove the bottleneck in 2,3-BD production (Köpke et al., 2016). Hence, overexpressing nifJ

can be useful in order to increase the amount of PFOR and thus enhance flux towards 2,3-BD

in C. ljungdahlii. This was tested by cloning either nifJ additionally into pMTL82151_23BD as

well as pJIR750_23BD or exchange of bdh with nifJ in both plasmids. Transformation of these

plasmids via electroporation did not lead to any success and via conjugation only

pMTL82151_23BD_PFOR and pMTL82151_23BD_oBDH_PFOR were transformed successfully

into C. ljungdahlii. The newly constructed plasmids harboring nifJ were the largest plasmids

for 2,3-BD production in this study with sizes of 11,898 bp to 15,274 bp. A correlation between

plasmid size and transformation effiency was reported in many bacteria, e. g. E. coli (Hanahan,

1983; Sheng et al., 1995; Chan et al., 2002) and B. subtilis (Ohse et al., 1995) only to mention

a few of them. A more detailed description why conjugation was used for transformation of

large plasmids is given in chapter 4.6.

Growth experiments using fructose (Figure 33) as well as syngas as energy and carbon source

(Figure 34) were conducted to give insights into growth behavior and production pattern of

C. ljungdahlii overexpressing nifJ. Under both growth conditions only C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] showed increased 2,3-BD concentrations of 2.2 mM and

4.7 mM, respectively, compared to C. ljungdahlii WT (1 mM and 2.2 mM). This correlates to a

2.1 and 2.2-fold, respectively, higher 2,3-BD production of C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] compared to C. ljungdahlii WT. These results demonstrate

that overexpression of nifJ together with alsS and budA is leading to an increased flux towards

2,3-BD. Nevertheless, the production is 2.6-fold lower compared to results of the company

Lanzatech (1.1 g/L = 12.2 mM), which were achieved with C. autoethanogenum

overexpressing alsS and budA in batch culture using synthetic syngas es energy and carbon

source (Table 10; Köpke et al., 2016). There are several possible reasons why the recombinant

C. ljungdahlii strain in this study produced lower 2,3-BD concentrations. Firstly, LanzaTech

Page 128: 2,3-Butanediol production using acetogenic bacteria

114 4. Discussion

used the plasmid pMTL83195 harboring the pCB102-replicon from C. butyricum compared to

the pBP1-replicon from C. botulinum used in this study. It is possible that pMTL83195 exhibits

a higher plasmid copy number in C. autoethanogenum compared to pMTL82151 in

C. ljungdahlii. Secondly, Lanzatech uses the ferredoxin gene promoter (Pfdx) from

C. perfringens instead of Ppta-ack, which is reported as one of the strongest promoters in the

genus Clostridum (Takamizawa et al., 2004). Thirdly, Lanzatech uses higher overpressure

(2.0 bar) in batch fermentations combined with utilization of a shaking incubator, which both

lead to a better gas-to-liquid mass transfer of CO. These are all factors, which might be starting

points to further increase 2,3-BD production in C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR]. The best 2,3-BD production of the recombinant

C. ljungdahlii strains constructed in this study (0.4-0.6 g/l), are much below 2,3-BD

concentrations of recombinant S. marcescens (152 g/l) and K. pneumoniae (150 g/l) strains

(Table 10). However, these strains use glucose as carbon source and are unattractive for

industrial fermentation processes, due to belonging to the risk group 2. Moreover, Lanzatech

already showed that 2,3-BD production from syngas has great potential for industry by

opening a pilot plant at BlueScope Steel (Glenbrook, New Zealand) producing CO-based 2,3-

BD with 15,000 gal/year (Regalbuto et al., 2017).

The autotrophic growth experiment (Figure 34) revealed that C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] showed a decreased max. OD600nm (0.9) compared to the

other recombinant C. ljungdahlii strains (1.3 and 1.5) as well as compared to C. ljungdahlii WT

(2.0). Moreover, this strain produced less ethanol (14.1 mM) compared to the other

C. ljungdahlii strains (43.5 mM-50.9 mM). A possible explanation might be presented, if

consumption and generation of reducing equivalents are taken into account (Figure 41). This

figure was obtained taking the data of a recently published study on C. autoethanogenum into

account (Richter et al., 2016). Overexpression in C. ljungdahlii

[pMTL82151_23BD_oBDH_PFOR] leads to higher conversion of acetyl-CoA to pyruvate and

thereby reduced ferredoxin (Fd2-) is consumed. This explains higher 2,3-BD production

compared to C. ljungdahlii WT. Moreover, if more reduced Fd2- is consumed via the PFOR

reaction, less Fd2- is available for ethanol production via the AOR pathway. It was already

Page 129: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 115

mentioned above that ethanol production is expected to be exclusively achieved via reduction

of undissociated acetic acid in C. ljungdahlii during fermentation on syngas (Richter et al.,

2016), which would explain lower ethanol production in C. ljungdahlii

Figure 41: Enzymes of the Wood-Ljungdahl pathway with corresponding reducing equivalents in C. ljungdahlii. Fdh and electron bifurcating hydrogenase HytCBDE1AE2 can form a complex to directly reduce CO2 to formate with H2. Rnf, Rhodobacter nitrogen fixation; ATPase, adenosine triphosphatase; Nfn, NADH-dependent reduced ferredoxin:NADP+ oxidoreductase; CODH, carbon monoxide dehydrogenase; FdhA, formate dehydrogenase; Acs/CODH, complex of acetyl-CoA synthase with carbon monoxide dehydrogenase; THF, tetrahydrofolate; [CO], enzyme-bound CO; Fdh, formate dehydrogenase; Fhs, formyl-THF synthetase; Mtc, methenyl-THF cyclohydrolase; Mtd, methylene-THF deyhdrogenae; Mtr, methylene-THF reductase; Met, methyl transferase; Pfor, pyruvate:ferredoxin oxidoreductase; AlsS, acetolactate synthase; BudA, acetolactate decarboxylase; Bdh, 2,3-BD dehydrogenase; sAdh, primary-secondary alcohol dehydrogenase; Pta, phosphotransacetylase; Ack, acetate kinase; AdhE, bifunctional aldehyde/alcohol dehydrogenase; Adh, monofunctional alcohol dehydrogenase; Aor, aldehyde:ferredoxin oxidoreductase; Co-FeS-P, corrinoid iron-sulfur protein; HS-CoA, coenzyme A. Based on the model of C. ljungdahlii published by Richter et al. (2016).

Page 130: 2,3-Butanediol production using acetogenic bacteria

116 4. Discussion

[pMTL82151_23BD_oBDH_PFOR]. Furthermore, reduction of acetyl-CoA to ethanol via

acetate by AOR pathway yields one ATP due to substrate level phosphorylation, which is

lacking if AOR reaction cannot take place, due to lower amounts of Fd2-. Additionally, the

decreased growth of C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR] might be explained by

the minimized Fd2- pool for the Rnf (Rhodobacter nitrogen fixation; ferredoxin:NAD+

oxidoreductase) complex. This complex needs Fd2- to reduce oxidized nicotinamide adenine

dinucleotide (NAD+) with concomitant translocation of cations (H+ in C. ljungdahlii, Na+ in

A. woodii) over the cell membrane (Imkamp et al., 2007; Müller et al., 2008; Biegel and Müller,

2010; Tremblay et al., 2012; Hess et al., 2016). Thereby, a proton gradient or sodium gradient

is generated, which is further used by an ATPase to provide additional ATP, since the Wood-

Ljungdahl pathway alone leads to no net ATP gain (Reidlinger and Müller, 1994). The ATP,

which is generated during acetate production via substrate level phosphorylation is needed

for activation of formate to formyl-THF. A detailed discussion of the phenotype of mutated

C. ljungdahlii [pMTL82151] is given in section 4.4.

The recombinant strain C. ljungdahlii [pMTL82151_23BD_PFOR], which harbors the plasmid

for overexpression of alsS, budA, bdh, and nifJ did not produce higher 2,3-BD concentrations

(0.8 mM and 2.5 mM) compared to C. ljungdahlii WT (1.0 mM and 2.2 mM) neither using

fructose nor syngas as energy and carbon source. The first three genes of the plasmid

pMTL82151_23BD_PFOR are identical to pMTL82151_23BD_oBDH_PFOR, so C. ljungdahlii

[pMTL82151_23BD_PFOR] was expected to produce similar or even higher amounts of

2,3-BD, due to additionally harboring the bdh gene. The plasmid pMTL82151_23BD_PFOR was

verified successfully in C. ljungdahlii [pMTL82151_23BD_PFOR] after performing the growth

experiments, but it cannot be excluded that mutations were present in this plasmid which

might affect 2,3-BD production of this strain. Thus, the sequence of pMTL82151_23BD_PFOR

from recombinant strain C. ljungdahlii [pMTL82151_23BD_PFOR] needs to be further

elucidated.

4.3 Detection of transcripts from synthetic 2,3-BD operons in C. ljungdahlii

Northern blot experiments were conducted to investigate, if the transcripts of the synthetic

2,3-BD operons alsS-budA-bdh, alsS-budAshort-bdh, and alsS-budAshort-adh (Figure 25) are

completely expressed in C. ljungdahlii. Moreover, it was investigated if utilization of one

promoter is sufficient for transcription of three genes. RNA was isolated in the early stationary

Page 131: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 117

growth phase (after 65 h), when 2,3-BD formation starts and transcripts for 2,3-BD production

were expected to be present. This was also shown in C. autoethanogenum via quantitative

PCR, where a slight up-regulation of genes involved in 2,3-BD formation was shown after 50 h

in the late exponential growth phase (Köpke et al., 2011b). Detection of transcripts alsS, budA,

and bdh from C. ljungdahlii chromosome was successful in C. ljungdahlii WT with sizes of

1,670 bases, 1,240 bases, and 1,074 bases, respectively (Figure 26). The transcript budA was

expected to have a size of 720 bases, but Figure 25B shows that budA is located downstream

of a coding sequence (CLJU_c08370), which is annotated as a hypothetical protein. Thus, it

seems budA and CLJU_c08370 form an operon, leading to a polycistronic mRNA with a size of

1,240 bases compared to 720 bases, in case of budA being a monocistronic mRNA. Since genes

for natural 2,3-BD production are located on the chromosome of C. ljungdahlii, they are

expected to be present in all investigated C. ljungdahlii strains. However, transcript alsS was

only detected in C. ljungdahlii WT, transcript budA in C. ljungdahlii WT, C. ljungdahlii

[pMTL82151], as well as C. ljungdahlii [pJIR750_23BD_budAshort_adh], and transcript bdh in

C. ljungdahlii WT as well as C. ljungdahlii [pMTL82151] (Figure 26). A possible explanation

might be that the respective mRNAs were not formed yet after 65 h, when RNA was isolated

from the different C. ljungdahlii strains. Furthermore, alsS-budA-bdh trancript of the synthetic

2,3-BD operon is only present in C. ljungdahlii [pJIR750_23BD], which is in accordance with

results of growth experiments using fructose as well as syngas as energy and carbon source

(Figure 9 and Figure 10) showing C. ljungdahlii [pJIR750_23BD] as best 2,3-BD producer

compared to C. ljungdahlii WT. Moreover, the transcript alsS-budA-bdh was not detected in

C. ljungdahlii [pMTL82151_23BD], although it produced a higher 2,3-BD concentration

compared to C. ljungdahlii WT in a growth experiment using syngas as carbon source (Figure

24). There might be much less transcript present in C. ljungdahlii [pMTL82151_23BD] cells

compared to C. ljungdahlii [pJIR750_23BD], which would show that plasmid replication of

vector pMTL82151 is lower compared to pJIR750 in C. ljungdahlii. Furthermore, the transcripts

of the other 2,3-BD operons alsS-budAshort-bdh and alsS-budAshort-adh could not be

detected via Northern blot analysis. This was unexpected, since all synthetic 2,3-BD operons

share the same constitutive promoter from C. ljungdahlii (Ppta-ack) with the only difference in

exchange of one gene of the operon. Either the transcripts were not formed yet after 65 h,

which is otherwise unlikely since a constitutive promoter was used for expression, or the

transcript was absent, due to a so far unknown reason. Moreover, Northern blot analysis

Page 132: 2,3-Butanediol production using acetogenic bacteria

118 4. Discussion

revealed several shorter signals at sizes of around 500 bases, 750 bases, 1,000 bases, and

1,500 bases. In contrast, these shortened signals were never present in C. ljungdahlii WT, so

they might be the result of unspecific binding of the probes either to the vector or to the

plasmid. However, alignment of the probes with the respective plasmids via in silico analysis

did not reveal any high homologies, so there might be another unknown reason for these

truncated signals.

4.4 Investigation of mutated C. ljungdahlii [pMTL82151] showing increased flux

towards ethanol and 2,3-BD

The mutated C. ljungdahlii [pMTL82151] strain produced slightly higher 2,3-BD concentrations

of 2.4 mM with fructose as carbon source (Figure 33) and 5.6 mM with syngas as energy and

carbon source (Figure 34) compared to C. ljungdahlii [pMTL82151_23BD_oBDH_PFOR]

(2.2 mM and 4.7 mM). This correlates to a 2.4-fold and 2.5-fold higher 2,3-BD production

compared to C. ljungdahlii WT, which produced 1.0 mM and 2.2 mM 2,3-BD, respectively.

Furthermore, this strain did not produce higher 2,3-BD concentrations (1.5 mM) in another

growth experiment using syngas as energy and carbon source (Figure 24) compared to

C. ljungdahlii WT (41.9 mM and 1.7 mM, respectively). This did not reflect the results from the

growth experiment using fructose as carbon source (Figure 23) and the other growth

experiment using syngas as energy and carbon source (Figure 34). It might be assumed that

mutated C. ljungdahlii [pMTL82151] did not show the phenotype of higher ethanol production

as well as higher 2,3-BD production in this growth experiment (Figure 24), due to worse CO

supply. As mentioned above, sufficient CO supply is an important feature for 2,3-BD

production (Simpson et al., 2011). Since C. ljungdahlii [pMTL82151] did still contain the vector

pMTL82151, which was confirmed by isolation of genomic DNA with concomitant

retransformation into E. coli XL-1 Blue MRF' as well as sequencing by GATC Biotech AG

(Constance, Germany), it was assumed that one or more mutations located on the

chromosome were responsible for the phenotype of higher ethanol as well as 2,3-BD

production. Recently, a spontaneous C. ljungdahlii mutant showing higher ethanol production

was reported (Whitham et al., 2017). The strain designated as C. ljungdahlii OTA-1 was

obtained after repeated subculturing of C. ljungdahlii WT in a rich mixotrophic medium and

storage for 2 weeks at room temperature (Whitham et al., 2017). Thus, it seems like

C. ljungdahlii has a potential to develop mutations during cultivation over a long period. The

Page 133: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 119

assumption that mutation of C. ljungdahlii also appeared in this study was confirmed by

sequencing genomes of C. ljungdahlii WT and C. ljungdahlii [pMTL82151] with subsequent SNP

analysis. Several SNPs leading to a change in the translated AA were found in mutated

C. ljungdahlii [pMTL82151] (Table 8). The spontaneous mutant strain showed this phenotype

after lyophilisation and reactivation. Since most of the SNPS were present in genes, which

were not assumed to be involved in ethanol or 2,3-BD production, only SNPs in gene adhE1

were examined more closely via comparison of protein domains (Figure 43) and protein

structure modelling (Figure 42). Two of the not investigated locus tags are CLJU_C07590 as

well as CLJU_c07600, which are annotated as putative secretion protein HlyD and putative

transporter protein, respectively (Table 8). Ethanol and 2,3-BD are small uncharged polar

molecules, which can pass the cell membrane via passive diffusion (Lodish et al., 2004). Hence,

it is unlikely that transporter or secrection proteins are involved in higher ethanol or higher

2,3-BD production. Moreover, locus tags CLJU_c24850 and CLJU_c24870 are designated as

genes stc1 and stc2, which encode signal-transduction and transcriptional regulator proteins.

They are also present in the stc cluster of the fungus Aspergillus nidulans, encoding enzymes

needed for synthesis of a toxic and carcinogenic secondary metabolite called sterigmatocystin

(precurser of the fungal toxin aflatoxin) (Hicks et al., 1997). The acetogen C. ljungdahlii does

not contain such a stc cluster, thus it might be concluded that genes stc1 and stc2 were

obtained during evolution via horizontal gene transfer, but do not fulfill a specific function in

this organism. The SNP introducing a stop codon in adhE1 (CLJU_c16510) (Table 8) seems to

Figure 42: Structure of enzyme models for AdhE1 from C. ljungdahlii WT (A) and mutated C. ljungdahlii [pMTL82151] (B). Modelling of protein sequences was performed using I-TASSER: Protein Structure & Function Predictions (http://zhanglab.ccmb.med.umich.edu/I-TASSER/).

Page 134: 2,3-Butanediol production using acetogenic bacteria

120 4. Discussion

be the most important SNP of this study, since this gene encodes a bifunctional

aldehyde/alcohol dehydrogenase, which catalyzes the reduction of acetyl-CoA to

acetaldehyde as well as the subsequent reduction of acetaldehyde to ethanol. This enzyme is

typically composed of a N-terminal acetylating aldehyde dehydrogenase (Ald) domain and a

C-terminal Fe-dependent Adh domain (Membrillo-Hernandez et al., 2000; Extance et al.,

2013). Figure 42 and Figure 43 show the impact of the stop codon on the enzyme structure of

AdhE1 and its protein domains in comparison to C. ljungdahlii WT. The stop codon leads to a

truncated version of AdhE1 with 202 AA in mutated C. ljungdahlii [pMTL82151] (Figure 42B

and Figure 43B) compared to AdhE1 of C. ljungdahlii WT with 870 AA (Figure 42A and Figure

43A). There is another gene (adhE2) encoding a bifunctional aldehyde/alcohol

dehydrogenase, which is located directly downstream of adhE1 on the chromosome of

C. ljungdahlii (Köpke et al., 2010; Leang et al., 2013). The two genes are also present in

C. autoethanogenum and supposed to be a potential result of gene duplication (Humphreys

et al., 2015). Moreover, it was reported that deletion of adhE2 showed no effect on ethanol

Figure 43: Protein domains of AdhE1 from C. ljungdahlii WT (A) and truncated AdhE1 from mutated C. ljungdahlii [pMTL82151] (B). Analysis of protein sequences were performed using InterPro: protein sequence analysis & classification (http://www.ebi.ac.uk/interpro/search/sequence-search).

Page 135: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 121

production in C. ljungdahlii using fructose as carbon source (Leang et al., 2013) and

overexpression of adhE2 in an adhE1/adhE2-deficient C. ljungdahlii strain did not result in

restoring ethanol production (Banerjee et al., 2014). These findings mean either that adhE2

does not encode a functional bifunctional aldehyde/alcohol dehydrogenase or

posttranscriptional factors limit expression and/or enzyme activity of AdhE2 (Banerjee et al.,

2014). In contrast, Liew et al. (2017) reported that adhE2 inactivation resulted in 63 % lower

ethanol production in C. autoethanogenum compared to C. autoethanogenum WT. Hence,

although being very similar at the genetic level (Bengelsdorf et al., 2016), both organisms show

differences in the phenotype concerning ethanol production (Cotter et al., 2009; Köpke et al.,

2010; Brown et al., 2014; Liew et al., 2016; Martin et al., 2016; Marcellin et al., 2016). The

truncated AdhE1 of mutated C. ljungdahlii [pMTL82151] in this study was assumed to be equal

to adhE1 knouckout due to the missing alcohol dehydrogenase domain and the truncated

aldehyde dehydrogenase N-terminal domain (Figure 43B). The inactivated adhE1 phenotype

of this study, showing higher ethanol and 2,3-BD production with fructose as well as syngas

as energy and carbon source, reflects more the results observed in C. autoethanogenum (Liew

et al., 2017) than in C. ljungdahlii (Banerjee et al., 2014). The deletion of adhE1 in C. ljungdahlii

was reported to result in a strain, which produced 6-fold less ethanol compared to

C. ljungdahlii WT (Banerjee et al., 2014). This is contrary to the results presented in this study.

However, adhE1 inactivation in C. autoethanogenum was shown to yield higher ethanol

production (171-183 %) with CO as carbon source but decreased 2,3-BD production compared

to C. autoethanogenum WT (Liew et al., 2017). Furthermore, the increased ethanol production

only occured, when strains were growing with CO but not during heterotrophic growth with

fructose (Liew et al., 2017). This shows that mutated C. ljungdahlii [pMTL82151] represents a

new phenotype compared to adhE1 inactivation strains described in literature. The increase

in ethanol production and 2,3-BD production of mutated C. ljungdahlii [pMTL82151] might be

explained considering the pool of reducing equivalents (Figure 41). Higher ethanol

concentrations are in agreement with the assumption that ethanol formation via AOR

pathway is more favourable for heterotrophic as well as autotrophic ethanol biosynthesis, due

to generation of ATP (Fast and Papoutsakis, 2012; Bertsch and Müller, 2015; Mock et al., 2015;

Richter et al., 2016; Valgepea et al., 2017a). Furthermore, Valgepea et al. (2017a)

demonstrated that ethanol production is strongly correlated with biomass concentrations

during steady-state culturing of C. autoethanogenum. The organism channels more carbon

Page 136: 2,3-Butanediol production using acetogenic bacteria

122 4. Discussion

into reduced by-products such as ethanol and 2,3-BD at high biomass concentrations.

Additionally, they showed that several transcripts were down-regulated with higher ethanol

production in C. autoethanogenum, for example the transcript of bifunctional

aldehyde/alcohol dehydrogenase AdhE (CAETHG_3747) (Valgepea et al., 2017a). This

observation fits to the results being observed during this study that a truncated AdhE1 enzyme

leads to higher ethanol production in mutated C. ljungdahlii [pMTL82151]. The higher 2,3-BD

production might also result from inactivation of adhE1 leading to an excess of reducing

equivalents, which can be consumed during 2,3-BD formation. This shows that mutated

C. ljungdahlii [pMTL82151] might be a target for further enhancement of 2,3-BD production

in C. ljungdahlii. Several passages of the strain without addition of antibiotic would result in

loss of the vector pMTL82151, due to missing selection pressure. Subsequently, allelic

exchange (Al-Hinai et al., 2012) or ClosTron mutagenesis (Heap et al., 2007) might be applied

to inactivate genes encoding AOR. These methods of genetic manipulation to inactivate genes

in clostridia were successfully applied for example in C. autoethanogenum recently (Mock et

al., 2015; Minton et al., 2016; Liew et al., 2017). Inactivation of aor would result in knock down

of the AOR pathway. Afterwards, pMTL82151_23BD_oBDH_PFOR plasmid could be

transformed into this strain to create flux towards 2,3-BD and not towards ethanol, since

excess of reducing equivalents would be consumed solely during 2,3-BD formation. A further

method to enhance 2,3-BD production in mutated C. ljungdahlii [pMTL82151] might be

represented by the arginine deiminase pathway, which was recently reported to provide

C. autoethanogenum with additional ATP, if arginine is used as nitrogen resource (Valgepea et

al., 2017b). This pathway would also lead to higher biomass concentrations and as mentioned

above thereby producing higher 2,3-BD concentrations. However, RNA sequencing showed

that the Wood-Ljungdahl pathway was down-regulated in strains growing with arginine

(Valgepea et al., 2017b), which might be unfavourable for 2,3-BD production using syngas as

energy and carbon source.

4.5 A. woodii as suitable host for 2,3-BD production from H2 + CO2?

The plasmids for 2,3-BD production were transformed into A. woodii to investigate, if

A. woodii can form 2,3-BD heterologously while growing on H2 + CO2. In order to test if 2,3-BD

formation from H2 + CO2 as sole energy and carbon source is energetically profitable in

A. woodii, stoichiometric energy balancing was performed. The calculation was done with

Page 137: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 123

incorporation of relevant energy mechanisms and reducing equivalents consuming-reactions

depicted in Figure 44 (Poehlein et al., 2012). Furthermore, the postulation that Mtr is not

reducing oxidized ferredoxin and therefore is not electron-bifurcating was incorporated into

the following equations (Bertsch et al., 2015). The formation of 2,3-BD as sole end-product

would be energetically unfavorable in A. woodii resulting in a negative ATP gain of -1.7 ATP

(Bertsch et al., 2015) (2).

4 CO2 + 11 H2 → C4O2H10 + 4 H2O + -1.7 ATP (2)

Figure 44: Stoichiometrical heterologous 2,3-BD production in A. woodii with H2 + CO2 as sole energy and carbon source as well as relevant enzymes. Rnf, Rhodobacter nitrogen fixation; ATPase, adenosine triphosphatase; Acs/CODH, complex of acetyl-CoA synthase with carbon monoxide dehydrogenase; THF, tetrahydrofolate; [CO], enzyme-bound CO; Fdh, formate dehydrogenase; Fhs, formyl-THF synthetase; Mtc, methenyl-THF cyclohydrolase; Mtd, methylene-THF deyhdrogenase; Mtr, methylene-THF reductase; Met, methyl transferase; PFOR, pyruvate:ferredoxin oxidoreductase; AlsS, acetolactate synthase; BudA, acetolactate decarboxylase; Bdh, 2,3-BD dehydrogenase; sAdh, primary-secondary alcohol dehydrogenase; Co-FeS-P, corrinoid iron-sulfur protein; HS-CoA, coenzyme A.

Page 138: 2,3-Butanediol production using acetogenic bacteria

124 4. Discussion

In conclusion, production of 6 mol acetate/mol 2,3-BD is needed to get a positive ATP balance

of 0.1 ATP in A. woodii using H2 + CO2 as sole energy and carbon source (3).

16 CO2 + 35 H2 → C4O2H10 + 6 C2H3OO- + 16 H2O + 0.1 ATP (3)

Nevertheless, no recombinant A. woodii strain was able to produce 2,3-BD heterologously

using H2 + CO2 as sole energy and carbon source (Figure 28). By this time, it was unknown that

A. woodii can use divalent alcohols such as 2,3-BD as well as 1,2-propanediol via oxidation as

carbon source (Hess et al., 2015; Schuchmann and Müller, 2016). In this reaction 2,3-BD is

oxidized to acetate with CO2 as electron acceptor, which is reduced to an additional acetate

(Hess et al., 2015) (4).

C4O2H10+ 1.5 CO2 + 1.3 ADP + 1.3 Pi→ 2.75 C2H3OO- + 1.3 ATP (4)

This finding explains why 2,3-BD was not detected in the conducted growth experiment.

However, A. woodii [pJIR750_23BD], A. woodii [pJIR750_23BD_budAshort], and A. woodii

[pJIR750_23BD_budAshort_adh] displayed higher acetate concentrations of 163.9 mM,

166.0 mM, and 156.3 mM, respectively, compared to A. woodii WT (148.5 mM). These results

would support the assumption that 2,3-BD was produced in A. woodii, but subsequently used

as carbon source. Nevertheless, it was shown in a previous study that vector pJIR750 leads to

higher acetate concentrations in A. woodii (Straub, 2012). Hence, it cannot be excluded that

increased acetate concentrations were plasmid-based. This might have been elucidated, if

samples for HPLC analysis would have been taken more often to see if 2,3-BD was formed and

subsequently consumed in case of A. woodii producing 2,3-BD concentrations above the

detection limit of HPLC analysis.

4.6 Heterologous butanol production in C. ljungdahlii - the conflict of large

plasmid transformation

In the past, heterologous butanol production (2 mM) from CO using C. ljungdahlii with the

genes for butanol formation from C. acetobutylicum was reported (Köpke et al., 2010).

Butanol production from CO as sole carbon and energy source is coupled to the production of

CO2 (5).

12 CO + 5 H2O→ C4OH10 + 8 CO2 (5)

Page 139: 2,3-Butanediol production using acetogenic bacteria

4. Discussion 125

However, glycerol cultures of C. ljungdahlii [pSOBptb] could not be reactivated successfully.

Thus, a new recombinant C. ljungdahlii strain producing butanol heterologously had to be

constructed. For this purpose the plasmid pSB3C5-UUMKS3 (Schuster, 2011) was used

containing two important differences compared to pSOBptb. Firstly, this plasmid contains etfA

and eftB, which were shown to be essential for reduction of crotonyl-CoA to butyryl-CoA in

C. kluyveri (Li et al., 2008) and are assumed to be also needed in C. acetobutylicum for this

reaction (Lütke-Eversloh and Bahl, 2011). Furthemore, pSB3C5-UUMKS 3 harbors the gene

adhE2 instead of bdhA and adhE1, since recently published data showed that AdhE2 is the key

enzyme responsible for butanol formation under acidogenesis and alcoholgenesis in

C. acetobutylicum (Yoo et al., 2016). Prior to transformation of the butanol plasmids pCE3 and

pJR3 into C. ljungdahlii in vivo methylattion using E. coli ER2275 [pACYC184_MCljI] was

performed. Therefore, the p15A-replicon of pACYC184_MCljI was exchanged, since pCE3 as

well as pJR3 harbor the identical p15A-replicon. It is known from literature that two identical

origins of replication cannot coexist in one organism (Novick, 1987). The pSC101-replicon was

chosen to maintain the low-copy number of pACYC184_MCljI and furthermore p15A-replicon

was described to coexist in one cell with pSC101-replicon (Peterson and Phillips, 2008).

However, transformation of the large butanol plasmids (14,199 bp-15,425 bp; Figure 36

and Figure 39) was not successful via electroporation. The effect of decreasing transformation

efficiency with increasing plasmid size was also shown for the acetogen A. woodii (Jonathan

Baker, University of Nottingham, unpublished). Although electrotransformation is a method

generally simple and efficient, conjugation is still a commonly used and frequently more

efficient method, most notably for transformation of large plasmids (Schweizer, 2008).

Bacterial conjugation requires direct contact between donor and recipient cell and was first

described between different E. coli strains (Lederberg and Tatum, 1946). Conjugation shows

at least two important advantages over electroporation. As mentioned above, electroporation

efficiencies decrease with increasing plasmid sizes (Szostková and Horáková, 1998), while

conjugation is lacking this limitation since even entire genomes have been successfully

transformed into E. coli using this method (Isaacs et al., 2011). The second advantage is that

methylation of plasmid DNA prior to transformation is obsolete, since during conjugation the

plasmid is transferred from the donor cell to recipient cell in form of a single strand, which is

afterwards methylated in second strand synthesis (Dominguez and O’Sullivan, 2013). Thus,

overcoming R-M system barriers of the organism can be achieved by using conjugation.

Page 140: 2,3-Butanediol production using acetogenic bacteria

126 4. Discussion

However, also conjugation experiments with the butanol plasmids into C. ljungdahlii were not

successful in this study. It might be concluded that the limit of transformable plasmid size was

reached. A patent published by the company LanzaTech reported recombinant

C. autoethanogenum DSM23693 and C. ljungdahlii DSM13528 strains harboring the butanol

production plasmid pMTL85245-thlA-crt-hbd (Köpke and Liew, 2012). This plasmid contains

the pIM13 replicon (Bacillus subtilis) and the genes thlA (thiolase), hbd (3-hydroxybutyryl-CoA

dehydrogenase), crt (crotonase), bcd (butyryl-CoA dehydrogenase), etfA (electrontransfer

flavoprotein), and etfB (electrontransfer flavoprotein) under control of the promoter Ppta-ack

for butanol synthesis. Furthermore, genes encoding phosphotransbutyrylase, butyrate kinase,

aldehyde:ferredoxin oxidoreductase, and butanol dehydrogenase were expressed. These

genes allow C. autoethanogenum and C. ljungdahlii to produce two molecules of acetyl-CoA,

which are afterwards combined to acetoacetyl-CoA and eventually converted via 3-

hydroxybutyryl-CoA and crotonyl-CoA to butyryl-CoA. Thereafter, butyryl-CoA is transformed

to butyryl phosphate catalyzed by phosphotransbutyrylase and this compound is further

converted to butyrate via the enzyme butyrate kinase (Dürre, 2016b). Recombinant

C. ljungdahlii strains produced 6 mM butanol under static conditions (without shaking), while

C. autoethanogenum produced higher 1-butanol concentrations of 25.7 mM and butyrate

production occurred (3.7 mM) (Köpke and Liew, 2012). Hence, this approach is beneficial due

to improved energetics of the pathway, as the butyrate kinase reaction yields an additional

ATP (Dürre, 2016b) and should be considered for further investigations on butanol production

in acetogens.

Page 141: 2,3-Butanediol production using acetogenic bacteria

5. Summary 127

5. Summary

1) Natural 2,3-BD production was detected in the acetogens C. ljungdahlii,

C. autoethanogenum, and C. ragsdalei using syngas as energy and carbon source. Acetate

was the main product and ethanol was a side product of these organisms. 2,3-BD was only

a minor side product, while C. ljungdahlii produced the highest 2,3-BD concentration of

4.8 mM.

2) 2,3-BD production plasmids were constructed using the genes encoding acetolactate

synthase (alsS), acetolactate decarboxlyase (budA), and 2,3-BD dehydrogenase (bdh) from

C. ljungdahlii under control of the constitutive Ppta-ack promoter from the same organism.

The synthetic 2,3-BD operon alsS-budA-bdh was cloned successfully into the E. coli-

Clostridium shuttle vectors pMTL82151 and pJIR750 containing different replicons for

Gram-positive bacteria.

3) Overexpression of the synthetic 2,3-BD operon alsS-budA-bdh using pJIR750_23BD

resulted in a 1.5-fold higher 2,3-BD production (1.2 and 6.4 mM 2,3-BD, respectively)

compared to C. ljungdahlii WT (0.8 and 4.3 mM 2,3-BD, respectively) with fructose as well

as syngas as energy and carbon source. Overexpression of the synthetic 2,3-BD operon

alsS-budA-bdh using pMTL82151_23BD resulted in a 1.8-fold increase of 2,3-BD

production (3.1 mM) compared to C. ljungdahlii WT (17 mM) using syngas as energy and

carbon source.

4) Overexpression of the synthetic 2,3-BD operon alsS-budA-bdh did not lead to an

improvement of 2,3-BD production in C. coskatii.

5) Three modified 2,3-BD production plasmids were constructed using a shortened sequence

of budA, exchange of bdh with primary-secondary alcohol dehydrogenase (adh) from

C. ljungdahlii, and the budABC operon from R. terrigena under control of Ppta-ack promoter

from C. ljungdahlii and Tsol-adc terminator from C. acetobutylicum.

6) Heterotrophic growth experiments with modified 2,3-BD production plasmids showed

nearly no improvement of 2,3-BD production in C. ljungdahlii, whereas autotrophically

grown C. ljungdahlii strains overexpressing adh produced 2.9-fold higher 2,3-BD

concentrations (3.3 mM) compared to C. ljungdahlii WT (1.7 mM). Prevention of the

potential rho-independent terminator structure as well as heterologous 2,3-BD

Page 142: 2,3-Butanediol production using acetogenic bacteria

128 5. Summary

production using budABC operon from R. terrigena did not enhance 2,3-BD production

compared to overexpression of the alsS-budA-bdh operon.

7) Recombinant A. woodii strains harboring 2,3-BD production plasmids did not produce 2,3-

BD heterologously using H2 + CO2 as energy and carbon source.

8) Genes encoding pyruvate:ferredoxin oxidoreductase (nifJ) and conjugal transfer region

(traJ/oriT) were cloned successfully into pMTL82151_23BD and pJIR750_23BD.

Conjugation into C. ljungdahlii was only successful for pMTL82151-based plasmids.

9) Overexpression of nifJ via alsS-budA-nifJ operon resulted in 2.2-fold (2.2 mM 2,3-BD) and

2.1-fold (4.7 mM 2,3-BD) higher 2,3-BD concentrations compared to C. ljungdahlii WT (1.0

mM and 2.2 mM 2,3BD) using fructose and syngas as energy and carbon source,

respectively. Overexpression of nifJ via alsS-budA-nifJ-bdh operon did not improve 2,3-BD

production compared to C. ljungdahlii WT.

10) Recombinant strain C. ljungdahlii [pMTL82151] showed the phenotype of decreased

acetate production with concomitant increased ethanol and 2,3-BD production compared

to C. ljungdahlii WT. 2,3-BD production was 2.4-fold increased (2.4 mM 2,3-BD) using

fructose and 2.5-fold increased (5.6 mM 2,3-BD) using syngas as energy and carbon source

compared to C. ljungdahlii WT (1.0 mM and 2.2 mM 2,3-BD, respectively). Genome

sequencing with subsequent SNP analysis revealed several SNPs located in the

chromosome of the spontaneously mutated C. ljungdahlii [pMTL82151] strain. The

respective strain was designated mutated C. ljungdahlii [pMTL82151]. Most important,

one SNP leading to a stop codon and therefore to a truncated AdhE1 enzyme seems to be

the reason for the observed phenotype.

11) Transcripts alsS, budA, and bdh were successfully detected in C. ljungdahlii WT via

Northern blot analysis. Transcript of synthetic 2,3-BD operon alsS-budA-bdh was verified

exclusively in C. ljungdahlii [pJIR750_23BD].

12) Replicons pIP404 (C. perfringens) and pCB102 (C. butyricum) as well as traJ/oriT region

were cloned successfully into butanol production plasmid pSB3C5-UUMKS 3. Replicon

p15A was exchanged successfully with replicon pSC101 in methylation plasmid

pACYC184_MCljI. Transformation of butanol production plasmids pCE3 and pJR3 into

C. ljungdahlii was not successful neither using electroporation nor conjugation.

Page 143: 2,3-Butanediol production using acetogenic bacteria

6. Zusammenfassung 129

6. Zusammenfassung

1) Die acetogenen Stämme C. ljungdahlii, C. autoethanogenum und C. ragsdalei zeigten eine

natürliche 2,3-Butandiol (2,3-BD)-Produktion mit Syngas als Kohlenstoff- und

Energiequelle. Dabei war Acetat das Hauptprodukt und Ethanol bzw. 2,3-BD traten als

Nebenprodukte auf, wobei 2,3-BD nur ein geringfügiges Nebenprodukt war. Dabei

produzierte C. ljungdahlii am meisten 2,3-BD mit einer Konzentration von 4,8 mM.

2) Es wurden 2,3-BD-Produktionsplasmide konstruiert, welche die Gene für die Acetlactat-

Synthase (alsS), Acetlactat-Decarboxylase (budA) und 2,3-Butandiol-Dehydrogenase (bdh)

unter Kontrolle des konstitutiven Promoters Ppta-ack aus C. ljungdahlii tragen. Das

synthetische 2,3-BD-Operon alsS-budA-bdh wurde erfolgreich in die E. coli-Clostridium

Schaukel-Vektoren pMTL82151 und pJIR750 kloniert, die unterschiedliche Replikons für

Gram-positive Bakterien enthalten.

3) Die Überexpression des synthetischen 2,3-BD-Operons alsS-budA-bdh mit Hilfe des

Plasmids pJIR750_23BD führte zu einer 1,5-fachen 2,3-Butandiol-Steigerung (1,2 mM

unter heterotrophen Wachstumsbedingungen bzw. 6,4 mM 2,3-BD unter autotrophen

Wachstumsbedingungen) im Vergleich zu C. ljungdahlii WT (0,8 mM mit Fructose bzw.

4,3 mM mit Syngas als Energie- und Kohlenstoffquelle). Unter autotrophen

Wachstumsbedingungen führte die Überexpression des synthetischen 2,3-BD-Operons

alsS-budA-bdh mit Hilfe des Plasmids pMTL82151_23BD zu einer 1,8-fachen Steigerung

der 2,3-BD-Produktion (3,1 mM) im Vergleich zu C. ljungdahlii WT (1,7 mM).

4) Die Überexpression des synthetischen 2,3-BD-Operons alsS-budA-bdh führte zu keiner

Steigerung der 2,3-BD-Produktion in C. coskatii.

5) Es wurden drei modifizierte 2,3-BD Produktionsplasmide konstruiert. Das erste Plasmid

weist eine kürzere Sequenz des budA-Gens auf, das zweite Plasmid entstand durch den

Austausch von bdh mit adh (primäre-sekundäre Alkohol-Dehydrogenase) und das dritte

Plasmid enthält das budABC-Operon aus R. terrigena unter Kontrolle des Ppta-ack-Promoters

aus C. ljungdahlii und des Tsol-adc-Terminators aus C. acetobutylicum.

6) Heterotrophe Wachstumsversuche mit den modifizierten 2,3-BD-Produktionsplasmiden

hatten keine Verbesserung der 2,3-BD-Produktion zur Folge. Jedoch führte die

Überexpression von adh in C. ljungdahlii zu einer 2,9-fachen Steigerung der 2,3-BD-

Konzentration (3,3 mM) im Vergleich zu C. ljungdahlii WT (1,7 mM). Sowohl das Entfernen

Page 144: 2,3-Butanediol production using acetogenic bacteria

130 6. Zusammenfassung

der Rho-unabhängigen Terminatorstruktur, als auch die heterologe Überexpression des

budABC-Operons aus R. terrigena konnten eine Steigerung der 2,3-BD Produktion im

Vergleich zur Überexpression des alsS-budA-bdh-Operons herbeiführen.

7) Unter autotrophen Wachstumsbedingen (H2 + CO2) waren rekombinante A. woodii-

Stämme nicht in der Lage, mit Hilfe der 2,3-BD-Produktionsplasmide heterolog 2,3-BD zu

produzieren.

8) Das Gen für die Pyruvat:Ferredoxin-Oxidoreduktase und die Genregion zum Transfer über

Konjugation (traJ/oriT) wurden erfolgreich in die Plasmide pMTL82151_23BD und

pJIR750_23BD kloniert. Die Konjugation mit C. ljungdahlii war nur für die pMTL82151-

basierten Plasmide erfolgreich.

9) Sowohl unter heterotrophen, als auch autotrophen Wachstumsbedingungen konnte

durch die Überexpression von nifJ über das alsS-budA-nifJ-Operon eine 2,2-fache

(2,2 mM) bzw. 2,1-fache (4.7 mM) Steigerung der 2,3-BD Konzentration im Vergleich zu

C. ljungdahlii WT (1,0 mM bzw. 2,2 mM) erreicht werden. Die Überexpression von nifJ über

das alsS-budA-nifJ-bdh-Operon führte zu keiner Verbesserung der 2,3-BD-Produktion im

Vergleich zu C. ljungdahlii WT.

10) Der rekombinante Stamm C. ljungdahlii [pMTL82151] zeigte im Verlauf dieser Arbeit einen

veränderten Phänotyp mit einer verringerten Acetat-Produktion zusammen mit einer

gesteigerten Ethanol- und 2,3-BD-Produktion im Vergleich zu C. ljungdahlii WT. Mit

Fructose als Substrat war die Produktion von 2,3-BD 2,4-fach (2,4 mM) gesteigert und mit

Syngas 2,5-fach (5,6 mM) erhöht im Vergleich zu C. ljungdahlii WT (1,0 mM bzw. 2,2 mM

2,3-BD). Durch Sequenzierung des Genoms mit anschließender SNP-Analyse konnten

mehrere SNPs im Chromosom des spontan mutierten Stamms C. ljungdahlii [pMTL82151]

detektiert werden. Der entsprechende Stamm wurde als mutierter C. ljungdahlii

[pMTL82151] beichnet. Der bedeutsamste SNP führte zu einem Stoppcodon und damit zu

einem verkürzten AdhE1-Enzym, was der Grund für den beobachteten Phänotyp zu sein

scheint.

11) Die Transkripte alsS, budA und bdh konnten in C. ljungdahlii WT über die Northern Blot-

Analyse erfolgreich nachgewiesen werden. Das Transkript des synthetischen 2,3-BD-

Operons konnte nur in C. ljungdahlii [pJIR750_23BD] erfolgreich detektiert werden.

12) Die Replikons pIP404 (C. perfringens) und pCB102 (C. butyricum), als auch die traJ/oriT-

Region konnten erfolgreich in das Butanol-Produktionsplasmid kloniert werden. In dem

Page 145: 2,3-Butanediol production using acetogenic bacteria

6. Zusammenfassung 131

Plasmid pACYC184_MCljI wurde das Replikon p15A erfolgreich gegen das Replikon pSC101

ausgetauscht. Die Butanol-Produktionsplasmide pCE2 und pJR3 konnten weder über

Elektroporation noch über Konjugation in C. ljungdahlii transformiert werden.

Page 146: 2,3-Butanediol production using acetogenic bacteria

132 7. References

7. References

Abrini J, Naveau H, Nyns EJ. 1994. Clostridium autoethanogenum, sp. nov., an anaerobic

bacterium that produces ethanol from carbon monoxide. Arch. Microbiol. 161:345–351.

Aeschelmann F, Carus M. 2015. Biobased building blocks and polymers in the world:

capacities, production, and applications–status quo and trends towards 2020. Ind. Biotechnol.

11:154–159.

Al-Hinai MA, Fast AG, Papoutsakis ET. 2012. Novel system for efficient isolation of Clostridium

double-crossover allelic exchange mutants enabling markerless chromosomal gene deletions

and DNA integration. Appl. Environ. Microbiol. 78:8112–8121.

Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. 1990. Basic local alignment search tool.

J. Mol. Biol. 215:403–410.

Alwine JC, Kemp DJ, Stark GR. 1977. Method for detection of specific RNAs in agarose gels by

transfer to diazobenzyloxymethyl-paper and hybridization with DNA probes. Proc. Natl. Acad.

Sci. USA 74:5350–5354.

Ash C, Priest FG, Collins MD. 1993. Molecular identification of rRNA group 3 bacilli (Ash,

Farrow, Wallbanks and Collins) using a PCR probe test. Proposal for the creation of a new

genus Paenibacillus. Ant. Leeuwenhoek 64:253–260.

Bai F, Dai L, Fan J, Truong N, Rao B, Zhang L, Shen Y. 2015. Engineered Serratia marcescens

for efficient (3R)-acetoin and (2R,3R)-2,3-butanediol production. J Ind. Microbiol. Biotechnol.

42:779–786.

Banerjee A, Leang C, Ueki T, Nevin KP, Lovley DR 2014. Lactose-inducible system for

metabolic engineering of Clostridium ljungdahlii. Appl. Environ. Microbiol. 80:2410–2416.

Bannam TL, Rood JI. 1993. Clostridium perfringens-Escherichia coli shuttle vectors that carry

single antibiotic resistance determinants. Plasmid 29:233–235.

Barik S, Prieto S, Harrison SB, Clausen EC, Gaddy JL. 1988. Biological production of alcohols

from coal through indirect liquefaction. Appl. Biochem. Biotechnol. 18:363–378.

Page 147: 2,3-Butanediol production using acetogenic bacteria

7. References 133

Bartowsky EJ, Henschke PA. 2004. The “buttery” attribute of wine-diacetyl-desirability,

spoilage and beyond. Int. J. Food Microbiol. 96:235–252.

Bengelsdorf FR, Straub M, Dürre P. 2013. Bacterial synthesis gas (syngas) fermentation.

Environ. Technol. 34:1639–1651.

Bengelsdorf FR, Poehlein A, Linder S, Erz C, Hummel T, Hoffmeister S, Daniel R, Dürre P. 2016.

Industrial acetogenic biocatalysts: a comparative metabolic and genomic analysis. Front.

Microbiol. 7:1036.

Berlyn MKB. 1998. Linkage map of Escherichia coli K-12: the traditional map. Microbiol. Mol.

Biol. Rev. 62:814–984.

Bertani G. 1951. Studies on lysogenesis I. The mode of phage liberation by lysogenic

Escherichia coli. J. Bacteriol. 62:293-300.

Bertsch J, Müller V. 2015. Bioenergetic constraints for conversion of syngas to biofuels in

acetogenic bacteria. Biotechnol. Biofuels 8:210

Bertsch J, Öppinger C, Hess V, Langer JD, Müller V. 2015. Heterotrimeric NADH-oxidizing

methylenetetrahydrofolate reductase from the acetogenic bacterium Acetobacterium woodii.

J. Bacteriol. 197:1681–1689.

Białkowska AM, Gromek E, Krysiak J, Sikora B, Kalinowska H, Jędrzejczak-Krzepkowska M,

Kubik C, Lang S, Schütt F, Turkiewicz M. 2015. Application of enzymatic apple pomace

hydrolysate to production of 2,3-butanediol by alkaliphilic Bacillus licheniformis NCIMB 8059.

J. Ind. Microbiol. Biotechnol. 42:1609–1621.

Białkowska AM, Jędrzejczak-Krzepkowska M, Gromek E, Krysiak J, Sikora B, Kalinowska H,

Kubik C, Schütt F, Turkiewicz M. 2016. Effects of genetic modifications and fermentation

conditions on 2,3-butanediol production by alkaliphilic Bacillus subtilis. Appl. Microbiol.

Biotechnol. 100:2663–2676.

Biegel E, Müller V. 2010. Bacterial Na+-translocating ferredoxin:NAD+ oxidoreductase. Proc.

Natl. Acad. Sci. USA 107:18138–18142.

Page 148: 2,3-Butanediol production using acetogenic bacteria

134 7. References

Biswas R, Yamaoka M, Nakayama H, Kondo T, Yoshida K, Bisaria VS, Kondo A. 2012.

Enhanced production of 2,3-butanediol by engineered Bacillus subtilis. Appl. Microbiol.

Biotechnol. 94:651–658.

Blomqvist K, Nikkola M, Lehtovaara P, Suihko ML, Airaksinen U, Straby KB, Knowles JK,

Penttilä ME. 1993. Characterization of the genes of the 2,3-butanediol operons from Klebsiella

terrigena and Enterobacter aerogenes. J. Bacteriol. 175:1392–1404.

Bolger AM, Lohse M, Usadel B. 2014. Trimmomatic: a flexible trimmer for Illumina sequence

data. Bioinformatics 30:2114–2120.

Brown SD, Nagaraju S, Utturkar S, De Tissera S, Segovia S, Mitchell W, Land ML, Dassanayake

A, Köpke Michael. 2014. Comparison of single-molecule sequencing and hybrid approaches

for finishing the genome of Clostridium autoethanogenum and analysis of CRISPR systems in

industrial relevant clostridia. Biotechnol. Biofuels 7:40.

Bruant G, Levesque MJ, Peter C, Guiot SR, Masson L. 2010. Genomic analysis of carbon

monoxide utilization and butanol production by Clostridium carboxidivorans strain P7. PLoS

ONE 5:e13033.

Buckel W, Thauer RK. 2013. Energy conservation via electron bifurcating ferredoxin reduction

and proton/Na+ translocating ferredoxin oxidation. Biochim. Biophys. Acta 1827:94-113.

Celińska E, Grajek W. 2009. Biotechnological production of 2,3-butanediol - current state and

prospects. Biotechnol. Adv. 27:715–725.

Chakraborty S, Piszel PE, Hayes CE, Baker RT, Jones WD. 2015. Highly selective formation of

n-butanol from ethanol through the Guerbet process: a tandem catalytic approach. J. Am.

Chem. Soc. 137:14264–14267.

Chan V, Dreolini LF, Flintoff KA, Lloyd SJ, Mattenley AA. 2002. The effect of increasing plasmid

size on transformation efficiency in Escherichia coli. J. Exp. Microbiol. Immunol. 2:207–223.

Chen Q, Fischer JR, Benoit VM, Dufour NP, Youderian P, Leong JM. 2008. In vitro CpG

methylation increases the transformation efficiency of Borrelia burgdorferi strains harboring

the endogenous linear plasmid lp56. J. Bacteriol. 190:7885–7891.

Page 149: 2,3-Butanediol production using acetogenic bacteria

7. References 135

Cheng KK, Liu Q, Zhang JA, Li JP, Xu JM, Wang GH. 2010. Improved 2,3-butanediol production

from corncob acid hydrolysate by fed-batch fermentation using Klebsiella oxytoca. Process

Biochem. 45:613–616.

Cho S, Kim T, Woo HM, Kim Yunje, Lee Jinwon, Um Y. 2015a. High production of 2,3-

butanediol from biodiesel-derived crude glycerol by metabolically engineered Klebsiella

oxytoca M1. Biotechnol. Biofuels 8:146.

Cho S, Kim T, Woo HM, Lee J, Kim Y, Um Y. 2015b. Enhanced 2,3-butanediol production by

optimizing fermentation conditions and engineering Klebsiella oxytoca M1 through

overexpression of acetoin reductase. PLoS ONE 10:e0138109.

Chu H, Xin B, Liu P, Wang Y, Li L, Liu X, Zhang X, Ma C, Xu P, Gao C. 2015. Metabolic

engineering of Escherichia coli for production of (2S,3S)-butane-2,3-diol from glucose.

Biotechnol. Biofuels 8:143.

Cohen SN, Chang AC. 1977. Revised interpretation of the origin of the pSC101 plasmid. J.

Bacteriol. 132:734-737.

Cotter JL, Chinn MS, Grunden AM. 2009. Ethanol and acetate production by Clostridium

ljungdahlii and Clostridium autoethanogenum using resting cells. Bioprocess Biosyst. Eng.

32:369–380.

Dai JY, Zhao P, Cheng XL, Xiu ZL. 2015. Enhanced production of 2,3-butanediol from sugarcane

molasses. Appl. Biochem. Biotechnol. 175:3014–3024.

Daniel SL, Hsu T, Dean SI, Drake HL. 1990. Characterization of the H2- and CO-dependent

chemolithotrophic potentials of the acetogens Clostridium thermoaceticum and Acetogenium

kivui. J. Bacteriol. 172:4464–4471.

Daniell J, Nagaraju S, Burton F, Köpke M, Simpson SD. 2016. Low-carbon fuel and chemical

production by anaerobic gas fermentation. Adv. Biochem. Eng. Biotechnol. 156:293-321.

De Hoon MJL, Makita Y, Nakai K, Miyano S. 2005. Prediction of transcriptional terminators in

Bacillus subtilis and related species. PLoS Comput. Biol. 1:e25.

Page 150: 2,3-Butanediol production using acetogenic bacteria

136 7. References

De Oliveira RR, Nicholson WL. 2016. Synthetic operon for (R,R)-2,3-butanediol production in

Bacillus subtilis and Escherichia coli. Appl. Microbiol. Biotechnol. 100:719–728.

Debuissy T, Pollet E, Avérous L. 2016. Synthesis of potentially biobased copolyesters based

on adipic acid and butanediols: kinetic study between 1,4- and 2,3-butanediol and their

influence on crystallization and thermal properties. Polymer 99:204–213.

Dominguez W, O’Sullivan DJ. 2013. Developing an efficient and reproducible conjugation-

based gene transfer system for bifidobacteria. Microbiology 159:328–338.

Dower WJ, Miller JF, Ragsdale CW. 1988. High efficiency transformation of E. coli by high

voltage electroporation. Nucl. Acids Res. 16:6127–6145.

Drake HL, Gößner AS, Daniel SL. 2008. Old acetogens, new light. Ann. N. Y. Acad. Sci.

1125:100–128.

Drancourt M, Bollet C, Carta A, Rousselier P. 2001. Phylogenetic analyses of Klebsiella species

delineate Klebsiella and Raoultella gen. nov., with description of Raoultella ornithinolytica

comb. nov., Raoultella terrigena comb. nov. and Raoultella planticola comb. nov. Int. J. Syst.

Evol. Microbiol. 51:925–932.

Dürre P. 2005. Formation of solvents in clostridia. In: Dürre P (ed.), Handbook on clostridia,

CRC Press, Boca Raton, FL, USA:671–693.

Dürre P. 2008. Fermentative butanol production. Ann. N. Y. Acad. Sci. 1125:353–362.

Dürre P. 2016a. Gas fermentation - a biotechnological solution for today’s challenges. Microb.

Biotechnol. 10:14–16.

Dürre P. 2016b. Butanol formation from gaseous substrates. FEMS Microbiol. Lett. 363:6.

Dürre P, Eikmanns BJ. 2015. C1-carbon sources for chemical and fuel production by microbial

gas fermentation. Curr. Opin. Biotechnol. 35:63–72.

Extance J, Crennell SJ, Eley K, Cripps R, Hough DW, Danson MJ. 2013. Structure of a

bifunctional alcohol dehydrogenase involved in bioethanol generation in Geobacillus

thermoglucosidasius. Acta Crystallogr. D Biol. Crystallogr. 69:2104–2115.

Page 151: 2,3-Butanediol production using acetogenic bacteria

7. References 137

Faba L, Díaz E, Ordóñez S. 2016. Base-catalyzed reactions in biomass conversion: reaction

mechanisms and catalyst deactivation. In: Schlaf M and Zhang ZC (eds.), Reaction pathways

and mechanisms in thermocatalytic biomass conversion I: cellulose structure,

depolymerization and conversion by heterogeneous catalysts, Springer Singapore, Singapore,

Singapore:87–122.

Fast AG, Papoutsakis ET. 2012. Stoichiometric and energetic analyses of non-photosynthetic

CO2-fixation pathways to support synthetic biology strategies for production of fuels and

chemicals. Curr. Opin. Chem. Eng. 1:380–395.

Faveri DD, Torre P, Molinari F, Perego P, Converti A. 2003. Carbon material balances and

bioenergetics of 2,3-butanediol bio-oxidation by Acetobacter hansenii. Enzyme Microb.

Technol. 33:708–719.

Fontaine FE, Peterson WH, McCoy E, Johnson MJ, Ritter GJ. 1942. A new type of glucose

fermentation by Clostridium thermoaceticum. J. Bacteriol. 43:701–715.

Formanek J, Mackie R, Blaschek HP. 1997. Enhanced butanol production by Clostridium

beijerinckii BA101 grown in semidefined P2 medium containing 6 percent maltodextrin or

glucose. Appl. Environ. Microbiol. 63:2306–2310.

Fu J, Wang Z, Chen T, Liu W, Shi T, Wang G, Tang Y, Zhao X. 2014. NADH plays the vital role

for chiral pure D-(−)-2,3-butanediol production in Bacillus subtilis under limited oxygen

conditions. Biotechnol. Bioeng. 111:2126–2131.

Fuchs G. 2007. Acetogenese: CO2 als Elektronenakzeptor. In: Fuchs G (ed.), Allgemeine

Mikrobiologie, 8th edition, Georg Thieme Verlag, Stuttgart, Germany:400-402.

Fuchs G. 2011. Alternative pathways of carbon dioxide fixation: insights into the early

evolution of life? Annu. Rev. Microbiol. 65:631–658.

Fulmer EI, Christensen LM, Kendali AR. 1933. Production of 2,3-butylene glycol by

fermentation. Ind. Eng. Chem. 25:798–800.

Page 152: 2,3-Butanediol production using acetogenic bacteria

138 7. References

Gao J, Xu H, Li Q, Feng X, Li S. 2010. Optimization of medium for one-step fermentation of

inulin extract from Jerusalem artichoke tubers using Paenibacillus polymyxa ZJ-9 to produce

R,R-2,3-butanediol. Bioresour. Technol. 101:7087–7093.

Garg SK, Jain A. 1995. Fermentative production of 2,3-butanediol: a review. Bioresour.

Technol. 51:103–109.

Gibson DG, Young L, Chuang R-Y, Venter JC, Hutchison CA, Smith HO. 2009. Enzymatic

assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6:343–345.

Gräfje H, Körnig W, Weitz HM, Reiß W, Steffan G, Diehl H, Bosche H, Schneider K, Kieczka H.

2000. Butanediols, butenediol, and butynediol. In: Bailey JE, Brinker CJ, Cornils B (eds.),

Ullmann’s encyclopedia of industrial chemistry, 6th edition Wiley-VCH, Weinheim,

Germany:407-414.

Green MR, Sambrook J. 2012. Molecular cloning: a laboratory manual. 4th edition, Cold Spring

Harbor Laboratory Press, New York, NY, USA.

Grethlein AJ, Worden RM, Jain MK, Datta R. 1991. Evidence for production of n-butanol from

carbon monoxide by Butyribacterium methylotrophicum. J. Ferment. Bioeng. 72:58–60.

Gubbels E, Jasinska-Walc L, Koning CE. 2013. Synthesis and characterization of novel

renewable polyesters based on 2,5-furandicarboxylic acid and 2,3-butanediol. J. Polymer Sci.

A Polymer Chem. 51:890–898.

Guo X, Cao C, Wang Y, Li C, Wu M, Chen Y, Zhang C, Pei H, Xiao D. 2014a. Effect of the

inactivation of lactate dehydrogenase, ethanol dehydrogenase, and phosphotransacetylase

on 2,3-butanediol production in Klebsiella pneumoniae strain. Biotechnol. Biofuels 7:1-11.

Guo XW, Zhang YH, Cao CH, Shen T, Wu MY, Chen YF, Zhang CY, Xiao DG. 2014b. Enhanced

production of 2,3-butanediol by overexpressing acetolactate synthase and acetoin reductase

in Klebsiella pneumoniae. Biotechnol. Appl. Biochem. 61:707–715.

Hahn HD, Dämbkes G, Rupprich N, Bahl H, Frey GD. 2000. Butanols. In: Bailey JE, Brinker CJ,

Cornils B (eds.), Ullmann’s encyclopedia of industrial chemistry, 6th edition, Wiley-VCH,

Weinheim, Germany:417-428.

Page 153: 2,3-Butanediol production using acetogenic bacteria

7. References 139

Hanahan D. 1983. Studies on transformation of Escherichia coli with plasmids. J. Mol. Biol.

166:557–580.

Harden A, Walpole GS. 1906. Chemical action of Bacillus lactis aerogenes (Escherich) on

glucose and mannitol: production of 2:3-butyleneglycol and acetylmethylcarbinol. Proc. R.

Soc. Lond. B Biol. Sci. 77:399-405.

Häßler T, Schieder D, Pfaller R, Faulstich M, Sieber V. 2012. Enhanced fed-batch fermentation

of 2,3-butanediol by Paenibacillus polymyxa DSM 365. Bioresour. Technol. 124:237–244.

Haveren J van, Scott EL, Sanders J. 2008. Bulk chemicals from biomass. Biofuels, Bioprod.

Bioref. 2:41–57.

Heap JT, Pennington OJ, Cartman ST, Carter GP, Minton NP. 2007. The ClosTron: a universal

gene knock-out system for the genus Clostridium. J. Microbiol. Methods 70:452–464.

Heap JT, Pennington OJ, Cartman ST, Minton NP. 2009. A modular system for Clostridium

shuttle plasmids. J. Microbiol. Methods 78:79–85.

Hess V, Oyrik O, Trifunović D, Müller V. 2015. 2,3-Butanediol metabolism in the acetogen

Acetobacterium woodii. Appl. Environ. Microbiol. 81:4711–4719.

Hess V, Gallegos R, Jones JA, Barquera B, Malamy MH, Müller V. 2016. Occurrence of

ferredoxin:NAD+ oxidoreductase activity and its ion specificity in several Gram-positive and

Gram-negative bacteria. PeerJ. 4:e1515.

Hicks JK, Yu JH, Keller NP, Adams TH. 1997. Aspergillus sporulation and mycotoxin production

both require inactivation of the FadA Gα protein-dependent signaling pathway. EMBO J.

16:4916–4923.

Hoffmeister, S. 2017. Production of platform chemicals with acetogenic bacteria. Ph. D.

Thesis. Ulm University, Ulm, Germany.

Hoffmeister S, Gerdom M, Bengelsdorf FR, Linder S, Flüchter S, Öztürk H, Blümke W, May A,

Fischer RJ, Bahl H, Dürre P. 2016. Acetone production with metabolically engineered strains

of Acetobacterium woodii. Metab. Eng. 36:37–47.

Page 154: 2,3-Butanediol production using acetogenic bacteria

140 7. References

Hu X, Shen X, Huang M, Liu C, Geng Y, Wang R, Xu R, Qiao H, Zhang Liqun. 2016.

Biodegradable unsaturated polyesters containing 2,3-butanediol for engineering applications:

synthesis, characterization and performances. Polymer 84:343–354.

Humphreys CM, McLean S, Schatschneider S, Millat T, Henstra AM, Annan FJ, Breitkopf R,

Pander B, Piatek P, Rowe P, Wichlacz AT, Woods C, Norman R, Blom J, Goesman A, Hodgman

C, Barrett D, Thomas NR, Winzer K, Minton NP. 2015. Whole genome sequence and manual

annotation of Clostridium autoethanogenum, an industrially relevant bacterium. BMC

Genomics 16:1085.

Imkamp F, Biegel E, Jayamani E, Buckel W, Müller V. 2007. Dissection of the caffeate

respiratory chain in the acetogen Acetobacterium woodii: identification of an Rnf-type NADH

dehydrogenase as a potential coupling site. J. Bacteriol. 189:8145–8153.

Inoue H, Nojima H, Okayama H. 1990. High efficiency transformation of Escherichia coli with

plasmids. Gene 96:23–28.

Isaacs FJ, Carr PA, Wang HH, Lajoie MJ, Sterling B, Kraal L, Tolonen AC, Gianoulis TA,

Goodman DB, Reppas NB, Emig CJ, Bang D, Hwang SJ, Jewett MC, Jacobson JM, Church GM.

2011. Precise manipulation of chromosomes in vivo enables genome-wide codon

replacement. Science 333:348-353.

Izard D, Ferragut C, Gavini F, Kersters K, De Ley J, Leclerc H. 1981. Klebsiella terrigena, a new

species from soil and water. Int. J. Syst. Evol. Microbiol. 31:116–127.

Jantama K, Polyiam P, Khunnonkwao P, Chan S, Sangproo M, Khor K, Jantama SS,

Kanchanatawee S. 2015. Efficient reduction of the formation of by-products and

improvement of production yield of 2,3-butanediol by a combined deletion of alcohol

dehydrogenase, acetate kinase-phosphotransacetylase, and lactate dehydrogenase genes in

metabolically engineered Klebsiella oxytoca in mineral salts medium. Metab. Eng. 30:16–26.

Ji XJ, Huang H, Du J, Zhu JG, Ren LJ, Hu N, Li S. 2009. Enhanced 2,3-butanediol production by

Klebsiella oxytoca using a two-stage agitation speed control strategy. Bioresour. Technol.

100:3410–3414.

Page 155: 2,3-Butanediol production using acetogenic bacteria

7. References 141

Ji XJ, Huang H, Zhu JG, Ren LJ, Nie ZK, Du J, Li S. 2010. Engineering Klebsiella oxytoca for

efficient 2,3-butanediol production through insertional inactivation of acetaldehyde

dehydrogenase gene. Appl. Microbiol. Biotechnol. 85:1751–1758.

Ji XJ, Huang H, Ouyang PK. 2011a. Microbial 2,3-butanediol production: a state-of-the-art

review. Biotechnol. Adv. 29:351–364.

Ji XJ, Nie ZK, Huang H, Ren LJ, Peng C, Ouyang PK. 2011b. Elimination of carbon catabolite

repression in Klebsiella oxytoca for efficient 2,3-butanediol production from glucose-xylose

mixtures. Appl. Microbiol. Biotechnol. 89:1119–1125.

Ji XJ, Huang H, Nie ZK, Qu L, Xu Q, Tsao GT. 2012. Fuels and chemicals from hemicellulose

sugars. Adv. Biochem. Eng. Biotechnol. 128:199–224.

Ji XJ, Xia ZF, Fu NH, Nie ZK, Shen MQ, Tian QQ, Huang H. 2013. Cofactor engineering through

heterologous expression of an NADH oxidase and its impact on metabolic flux redistribution

in Klebsiella pneumoniae. Biotechnol. Biofuels 6:7.

Ji XJ, Liu LG, Shen MQ, Nie ZK, Tong YJ, Huang H. 2015. Constructing a synthetic metabolic

pathway in Escherichia coli to produce the enantiomerically pure (R,R)-2,3-butanediol.

Biotechnol. Bioeng. 112:1056–1059.

Jones DT, Woods DR. 1986. Acetone-butanol fermentation revisited. Microbiol. Rev. 50:484–

524.

Jung MY, Jung HM, Lee J, Oh MK. 2015. Alleviation of carbon catabolite repression in

Enterobacter aerogenes for efficient utilization of sugarcane molasses for 2,3-butanediol

production. Biotechnol. Biofuels 8:106.

Jung MY, Mazumdar S, Shin SH, Yang KS, Lee J, Oh MK. 2014. Improvement of 2,3-butanediol

yield in Klebsiella pneumoniae by deletion of the pyruvate formate-lyase gene. Appl. Environ.

Microbiol. 80:6195–6203.

Jung MY, Ng CY, Song H, Lee J, Oh MK. 2012. Deletion of lactate dehydrogenase in

Enterobacter aerogenes to enhance 2,3-butanediol production. Appl. Microbiol. Biotechnol.

95:461–469.

Page 156: 2,3-Butanediol production using acetogenic bacteria

142 7. References

Jurchescu IM, Hamann J, Zhou X, Ortmann T, Kuenz A, Prusse U, Lang S. 2013. Enhanced 2,3-

butanediol production in fed-batch cultures of free and immobilized Bacillus licheniformis

DSM8785. Appl. Microbiol. Biotechnol. 97:6715–6723.

Kallbach M, Horn S, Kuenz A, Prüße U. 2017. Screening of novel bacteria for the 2,3-

butanediol production. Appl. Microbiol. Biotechnol. 101:1025–1033.

Kay JE, Jewett MC. 2015. Lysate of engineered Escherichia coli supports high-level conversion

of glucose to 2,3-butanediol. Metab. Eng. 32:133–142.

Kim B, Lee S, Jeong D, Yang J, Oh MK, Lee J. 2014a. Redistribution of carbon flux toward 2,3-

butanediol production in Klebsiella pneumoniae by metabolic engineering. PLoS ONE

9:e105322.

Kim B, Lee S, Park J, Lu M, Oh M, Kim Y, Lee J. 2012. Enhanced 2,3-butanediol production in

recombinant Klebsiella pneumoniae via overexpression of synthesis-related genes. J.

Microbiol. Biotechnol. 22:1258–1263.

Kim DK, Rathnasingh C, Song H, Lee HJ, Seung D, Chang YK. 2013a. Metabolic engineering of

a novel Klebsiella oxytoca strain for enhanced 2,3-butanediol production. J. Biosci. Bioeng.

116:186–192.

Kim JW, Seo SO, Zhang GC, Jin YS, Seo JH. 2015. Expression of Lactococcus lactis NADH oxidase

increases 2,3-butanediol production in Pdc-deficient Saccharomyces cerevisiae. Bioresour.

Technol. 191:512–519.

Kim SJ, Seo SO, Jin YS, Seo JH. 2013b. Production of 2,3-butanediol by engineered

Saccharomyces cerevisiae. Bioresour. Technol. 146:274–281.

Kim SJ, Seo SO, Park YC, Jin YS, Seo JH. 2014b. Production of 2,3-butanediol from xylose by

engineered Saccharomyces cerevisiae. J. Biotechnol. 192:376–382.

Koboldt DC, Zhang Q, Larson DE, Shen D, McLellan MD, Lin L, Miller CA, Mardis ER, Ding L,

Wilson RK. 2012. VarScan 2: somatic mutation and copy number alteration discovery in cancer

by exome sequencing. Genome Res. 22:568–576.

Page 157: 2,3-Butanediol production using acetogenic bacteria

7. References 143

Koda K, Matsu-ura T, Obora Y, Ishii Y. 2009. Guerbet reaction of ethanol to n-butanol

catalyzed by iridium complexes. Chem. Lett. 38:838–839.

Köpke M, Chen WY. 2013. Recombinant microorganisms and uses therefor. US 2013/0323806

A1. Washington, D.C.: U.S. Patent and Trademark Office.

Köpke M, Gerth ML, Maddock DJ, Mueller AP, Liew F, Simpson SD, Patrick WM. 2014.

Reconstruction of an acetogenic 2,3-butanediol pathway involving a novel NADPH-dependent

primary-secondary alcohol dehydrogenase. Appl. Environ. Microbiol. 80:3394–3403.

Köpke M, Havill A. 2014. LanzaTech’s route to bio-butadiene. Catal. Rev. 27:7–12.

Köpke M, Held C, Hujer S, Liesegang H, Wiezer A, Wollherr A, Ehrenreich A, Liebl W,

Gottschalk G, Dürre P. 2010. Clostridium ljungdahlii represents a microbial production

platform based on syngas. Proc. Natl. Acad. Sci. USA 107:13087–13092.

Köpke M, Liew F. 2012. Butanol production from carbon monoxide by a recombinant

microorganism. WO2012053905 A1. World patent.

Köpke M, Mihalcea C, Bromley JC, Simpson SD. 2011a. Fermentative production of ethanol

from carbon monoxide. Curr. Opin. Biotechnol. 22:320–325.

Köpke M, Mihalcea C, Liew F, Tizard JH, Ali MS, Conolly JJ, Al-Sinawi B, Simpson SD. 2011b.

2,3-Butanediol production by acetogenic bacteria, an alternative route to chemical synthesis,

using industrial waste gas. Appl. Environ. Microbiol. 77:5467–5475.

Köpke M, Mueller AP. 2016. Recombinant microorganisms exhibiting increased flux through

a fermentation pathway. US 2016/0160223 A1. Washington, D.C.: U.S. Patent and Trademark

Office.

Kozlowski JT, Davis RJ. 2013. Heterogeneous catalysts for the Guerbet coupling of alcohols.

ACS Catal. 3:1588–1600.

Kuhz H, Kuenz A, Prüße U, Willke T, Vorlop K-D. 2017. Products components: alcohols. Adv.

Biochem. Biotechnol. doi:10.1007/10_2016_74.

Page 158: 2,3-Butanediol production using acetogenic bacteria

144 7. References

Küsel K, Drake H. 2005. Acetogenic clostridia. In: Dürre P (ed.), Handbook on clostridia, CRC

Press, Boca Raton, FL, USA:719–746.

Langmead B, Salzberg SL. 2012. Fast gapped-read alignment with Bowtie 2. Nat. Methods

9:357–359.

Leang C, Ueki T, Nevin KP, Lovley DR. 2013. A genetic system for Clostridium ljungdahlii: a

chassis for autotrophic production of biocommodities and a model homoacetogen. Appl.

Environ. Microbiol. 79:1102–1109.

Lechner M, Findeiß S, Steiner L, Marz M, Stadler PF, Prohaska SJ. 2011. Proteinortho:

detection of (Co-)orthologs in large-scale analysis. BMC Bioinformatics 12:124.

Lederberg J, Tatum EL. 1946. Gene recombination in Escherichia coli. Nature 158:558.

Lee S, Kim B, Park K, Um Y, Lee J. 2012. Synthesis of pure meso-2,3-butanediol from crude

glycerol using an engineered metabolic pathway in Escherichia coli. Appl. Biochem.

Biotechnol. 166:1801–1813.

Lee S, Kim B, Yang J, Jeong D, Park S, Lee J. 2015. A non-pathogenic and optically high

concentrated (R,R)-2,3-butanediol biosynthesizing Klebsiella strain. J. Biotechnol. 209:7–13.

Lee SJ, Thapa LP, Lee JH, Choi HS, Kim SB, Park C, Kim SW. 2016. Stimulation of 2,3-butanediol

production by upregulation of alsR gene transcription level with acetate addition in

Enterobacter aerogenes ATCC 29007. Process Biochem. 51:1904–1910.

Lehrach H, Diamond D, Wozney JM, Boedtker H. 1977. RNA molecular weight determinations

by gel electrophoresis under denaturing conditions, a critical reexamination. Biochemistry

16:4743–4751.

Li D, Dai JY, Xiu ZL. 2010a. A novel strategy for integrated utilization of Jerusalem artichoke

stalk and tuber for production of 2,3-butanediol by Klebsiella pneumoniae. Bioresour. Technol.

101:8342–8347.

Page 159: 2,3-Butanediol production using acetogenic bacteria

7. References 145

Li F, Hinderberger J, Seedorf H, Zhang J, Buckel W, Thauer RK. 2008. Coupled ferredoxin and

crotonyl coenzyme A (CoA) reduction with NADH catalyzed by the butyryl-CoA

dehydrogenase/Etf complex from Clostridium kluyveri. J. Bacteriol. 190:843–850.

Li H. 2011. A statistical framework for SNP calling, mutation discovery, association mapping

and population genetical parameter estimation from sequencing data. Bioinformatics

27:2987–2993.

Li J, Baral NR, Jha AK. 2014a. Acetone–butanol–ethanol fermentation of corn stover by

Clostridium species: present status and future perspectives. World J. Microbiol. Biotechnol.

30:1145–1157.

Li L, Chen C, Li K, Wang Y, Gao C, Ma C, Xu P. 2014c. Efficient simultaneous saccharification

and fermentation of inulin to 2,3-butanediol by thermophilic Bacillus licheniformis ATCC

14580. Appl. Environ. Microbiol. 80:6458–6464.

Li L, Li K, Wang K, Chen C, Gao C, Ma C, Xu P. 2014b. Efficient production of 2,3-butanediol

from corn stover hydrolysate by using a thermophilic Bacillus licheniformis strain. Bioresour.

Technol. 170:256–261.

Li L, Li K, Wang Y, Chen C, Xu Y, Zhang L, Han B, Gao C, Tao F, Ma C, Xu P. 2015a. Metabolic

engineering of Enterobacter cloacae for high-yield production of enantiopure (2R,3R)-2,3-

butanediol from lignocellulose-derived sugars. Metabolic Engineering 28:19–27.

Li L, Zhang L, Li K, Wang Y, Gao C, Han B, Ma C, Xu P. 2013. A newly isolated Bacillus

licheniformis strain thermophilically produces 2,3-butanediol, a platform and fuel bio-

chemical. Biotechnol. Biofuels 6:123.

Li N, Yang J, Chai C, Yang S, Jiang W, Gu Y. 2015b. Complete genome sequence of Clostridium

carboxidivorans P7T, a syngas-fermenting bacterium capable of producing long-chain

alcohols. J. Biotechnol. 211:44–45.

Li ZJ, Jian J, Wei XX, Shen XW, Chen GQ. 2010b. Microbial production of meso-2,3-butanediol

by metabolically engineered Escherichia coli under low oxygen condition. Appl. Microbiol.

Biotechnol. 87:2001–2009.

Page 160: 2,3-Butanediol production using acetogenic bacteria

146 7. References

Lian J, Chao R, Zhao H. 2014. Metabolic engineering of a Saccharomyces cerevisiae strain

capable of simultaneously utilizing glucose and galactose to produce enantiopure (2R,3R)-

butanediol. Metab. Eng. 23:92–99.

Liew F, Henstra AM, Kӧpke M, Winzer K, Simpson SD, Minton NP. 2017. Metabolic

engineering of Clostridium autoethanogenum for selective alcohol production. Metab. Eng.

40:104–114.

Liew F, Martin ME, Tappel RC, Heijstra BD, Mihalcea C, Köpke M. 2016. Gas fermentation –

a flexible platform for commercial scale production of low-carbon-fuels and chemicals from

waste and renewable feedstocks. Front. Microbiol. 7: 694.

Liou JSC, Balkwill DL, Drake GR, Tanner RS. 2005. Clostridium carboxidivorans sp. nov., a

solvent-producing clostridium isolated from an agricultural settling lagoon, and

reclassification of the acetogen Clostridium scatologenes strain SL1 as Clostridium drakei sp.

nov. Int. J. of Syst. Evol. Microbiol. 55:2085–2091.

Liu Z, Qin J, Gao C, Hua D, Ma C, Li L, Wang Y, Xu P. 2011. Production of (2S,3S)-2,3-butanediol

and (3S)-acetoin from glucose using resting cells of Klebsiella pneumonia and Bacillus subtilis.

Bioresour. Technol. 102:10741–10744.

Ljungdahl LG. 1986. The autotrophic pathway of acetate synthesis in acetogenic bacteria.

Annu. Rev. Microbiol. 40:415–450.

Lodish H, Berk A, Baltimore D, Matsudaira P, Zipursky SL, Darnell J. 2004. Overview of

membrane transport. In: Lodish H, Berk A, Matsudaira P, Kaiser C, Krieger M, Zipursky SL,

Darnell J(eds.), Molecular cell biology, 5th edition, W.H. Freeman and Co., New York, NY,

USA:245-251.

Lovley DR, Nevin KP. 2013. Electrobiocommodities: powering microbial production of fuels

and commodity chemicals from carbon dioxide with electricity. Curr. Opin. Biotechnol.

24:385–390.

Lovley DR. 2011. Powering microbes with electricity: direct electron transfer from electrodes

to microbes. Environ. Microbiol. Rep. 3:27–35.

Page 161: 2,3-Butanediol production using acetogenic bacteria

7. References 147

Lütke-Eversloh T, Bahl H. 2011. Metabolic engineering of Clostridium acetobutylicum: recent

advances to improve butanol production. Curr. Opin. Biotechnol. 22:634–647.

Ma C, Wang A, Qin J, Li L, Ai X, Jiang T, Tang H, Xu P. 2009. Enhanced 2,3-butanediol

production by Klebsiella pneumoniae SDM. Appl. Microbiol. Biotechnol. 82:49–57.

Magee RJ, Kosaric N. 1987. The microbial production of 2,3-butanediol. Adv. Appl. Microbiol.

32:89–161.

Marcellin E, Behrendorff JB, Nagaraju S, DeTissera S, Segovia S, Palfreyman RW, Daniell J,

Licona-Cassani C, Quek L, Speight R, Hodson MP, Simpson SD, Mitchell WP, Köpke M, Nielsen

LK. 2016. Low carbon fuels and commodity chemicals from waste gases – systematic approach

to understand energy metabolism in a model acetogen. Green Chem. 18:3020–3028.

Martin ME, Richter H, Saha S, Angenent LT. 2016. Traits of selected Clostridium strains for

syngas fermentation to ethanol. Biotechnol. Bioeng. 113:531–539.

Mazumdar S, Lee J, Oh MK. 2013. Microbial production of 2,3 butanediol from seaweed

hydrolysate using metabolically engineered Escherichia coli. Bioresour. Technol. 136:329–336.

Membrillo-Hernandez J, Echave P, Cabiscol E, Tamarit J, Ros J, Lin EC. 2000. Evolution of the

adhE gene product of Escherichia coli from a functional reductase to a dehydrogenase. Genetic

and biochemical studies of the mutant proteins. J. Biol. Chem. 275:33869–33875.

Mermelstein LD, Papoutsakis ET. 1993. In vivo methylation in Escherichia coli by the Bacillus

subtilis phage ɸ3T I methyltransferase to protect plasmids from restriction upon

transformation of Clostridium acetobutylicum ATCC 824. Appl. Environ. Microbiol. 59:1077–

1081.

Minton NP, Ehsaan M, Humphreys CM, Little GT, Baker J, Henstra AM, Liew F, Kelly ML,

Sheng L, Schwarz K, Zhang Y. 2016. A roadmap for gene system development in Clostridium.

Anaerobe 41:104–112.

Mock J, Zheng Y, Mueller AP, Ly S, Tran L, Segovia S, Nagaraju S, Köpke M, Dürre P, Thauer

RK. 2015. Energy conservation associated with ethanol formation from H2 and CO2 in

Clostridium autoethanogenum involving electron bifurcation. J. Bacteriol. 197:2965–2980.

Page 162: 2,3-Butanediol production using acetogenic bacteria

148 7. References

Müller V, Imkamp F, Biegel E, Schmidt S, Dilling S. 2008. Discovery of a ferredoxin:NAD+-

oxidoreductase (Rnf) in Acetobacterium woodii: a novel potential coupling site in acetogens.

Ann. N. Y. Acad. Sci. 1125:137–146.

Mullis K, Faloona F, Scharf S, Saiki R, Horn G, Erlich H. 1986. Specific enzymatic amplification

of DNA in vitro: the polymerase chain reaction. Cold Spring Harb. Symp. Quant. Biol. 51:263–

273.

Myszkowski J, Zielinski AZ. 1965. Synthesis of glycerol monochlorohydrins from allyl alcohol.

Przem. Chem. 44:249-252.

Nakashima N, Akita H, Hoshino T. 2014. Establishment of a novel gene expression method,

BICES (biomass-inducible chromosome-based expression system), and its application to the

production of 2,3-butanediol and acetoin. Metab. Eng. 25:204–214.

Nakashimada Y, Marwoto B, Kashiwamura T, Kakizono T, Nishio N. 2000. Enhanced 2,3-

butanediol production by addition of acetic acid in Paenibacillus polymyxa. J. Biosci. Bioeng.

90:661–664.

Napora-Wijata K, Strohmeier GA, Winkler M. 2014. Biocatalytic reduction of carboxylic acids.

Biotechnol. J. 9:822–843.

Ndou A, Plint N, Coville N. 2003. Dimerisation of ethanol to butanol over solid-base catalysts.

Appl. Catal. A 251:337–345.

Nevin KP, Woodard TL, Franks AE, Summers ZM, Lovley DR. 2010. Microbial electrosynthesis:

feeding microbes electricity to convert carbon dioxide and water to multicarbon extracellular

organic compounds. mBio 1:e00103–10–e00103–10.

Nevin KP, Hensley SA, Franks AE, Summers ZM, Ou J, Woodard TL, Snoeyenbos-West OL,

Lovley DR. 2011. Electrosynthesis of organic compounds from carbon dioxide is catalyzed by

a diversity of acetogenic microorganisms. Appl. Environ. Microbiol. 77:2882–2886.

Ng CY, Jung MY, Lee J, Oh MK. 2012. Production of 2,3-butanediol in Saccharomyces cerevisiae

by in silico aided metabolic engineering. Microb. Cell Fact. 11:68.

Page 163: 2,3-Butanediol production using acetogenic bacteria

7. References 149

Nie ZK, Ji XJ, Huang H, Du J, Li ZY, Qu L, Zhang Q, Ouyang PK. 2011. An effective and simplified

fed-batch strategy for improved 2,3-butanediol production by Klebsiella oxytoca. Appl.

Biochem. Biotechnol. 163:946–953.

Nielsen DR, Yoon SH, Yuan CJ, Prather KLJ. 2010. Metabolic engineering of acetoin and meso-

2,3-butanediol biosynthesis in E. coli. Biotechnol. J. 5:274–284.

Novick RP. 1987. Plasmid incompatibility. Microbiol. Rev. 51:381-395.

O’Lenick AJ. 2001. Guerbet chemistry. J. Surfactants Deterg. 4:311–315.

Ohse M, Takahashi K, Kadowaki Y, Kusaoke H. 1995. Effects of plasmid DNA sizes and several

other factors on transformation of Bacillus subtilis ISW1214 with plasmid DNA by

electroporation. Biosci. Biotechnol. Biochem. 59:1433–1437.

Oliver JWK, Machado IMP, Yoneda H, Atsumi S. 2013. Cyanobacterial conversion of carbon

dioxide to 2,3-butanediol. Proc. Natl. Acad. Sci. USA 110:1249–1254.

Olson DG, McBride JE, Joe Shaw A, Lynd LR. 2012. Recent progress in consolidated

bioprocessing. Curr. Opin. Biotechnol. 23:396–405.

Park JM, Rathnasingh C, Song H. 2015. Enhanced production of (R,R)‑2,3‑butanediol by

metabolically engineered Klebsiella oxytoca. J. Ind. Microbiol. Biotechnol. 42:1419–1425.

Park JM, Song H, Lee HJ, Seung D. 2013. Genome-scale reconstruction and in silico analysis of

Klebsiella oxytoca for 2,3-butanediol production. Microb. Cell Fact. 12:20.

Pasteur L. 1862. Quelques résultats nouveaux relatifs aux fermentations acétique et

butyrique. Bull. Soc. Chim. Paris, May 1862:52–53.

Peterson J, Phillips GJ. 2008. New pSC101-derivative cloning vectors with elevated copy

numbers. Plasmid 59:193–201.

Petrini P, De Ponti S, Fare S, Tanzi MC. 1999. Polyurethane-maleamides for cardiovascular

applications: synthesis and properties. J. Mater. Sci. Mater. Med. 10:711–714.

Page 164: 2,3-Butanediol production using acetogenic bacteria

150 7. References

Petrov K, Petrova P. 2009. High production of 2,3-butanediol from glycerol by Klebsiella

pneumoniae G31. Appl. Microbiol. Biotechnol. 84:659–656.

Petrov K, Petrova P. 2010. Enhanced production of 2,3-butanediol from glycerol by forced pH

fluctuations. Appl. Microbiol. Biotechnol. 87:943–949.

Phillips JR, Atiyeh HK, Tanner RS, Torres JR, Saxena J, Wilkins MR, Huhnke RL. 2015. Butanol

and hexanol production in Clostridium carboxidivorans syngas fermentation: medium

development and culture techniques. Bioresour. Technol. 190:114–121.

Playne MJ. 1985. Determination of ethanol, volatile fatty acids, lactic and succinic acids in

fermentation liquids by gas chromatography. J. Sci. Food Agric. 36:638–644.

Poehlein A, Schmidt S, Kaster A-K, Goenrich M, Vollmers J, Thürmer A, Bertsch J,

Schuchmann K, Voigt B, Hecker M, Daniel R, Thauer RK, Gottschalk G, Müller V. 2012. An

ancient pathway combining carbon dioxide fixation with the generation and utilization of a

sodium ion gradient for ATP synthesis. PLoS ONE 7:e33439.

Purdy D, O’keeffe TA, Elmore M, Herbert M, McLeod A, Bokori-Brown M, Ostrowski A,

Minton NP. 2002. Conjugative transfer of clostridial shuttle vectors from Escherichia coli to

Clostridium difficile through circumvention of the restriction barrier. Mol. Microbiol. 46:439–

452.

Quan J, Tian J. 2011. Circular polymerase extension cloning for high-throughput cloning of

complex and combinatorial DNA libraries. Nat. Protoc. 6:242–251.

Radoš D, Carvalho AL, Wieschalka S, Neves AR, Blombach B, Eikmanns BJ, Santos H. 2016.

Engineering Corynebacterium glutamicum for the production of 2,3-butanediol. Microb. Cell

Fact. 14:171.

Ragsdale SW. 2007. Nickel and the carbon cycle. J. Inorg. Biochem. 101:1657–1666.

Ragsdale SW, Pierce E. 2008. Acetogenesis and the Wood-Ljungdahl pathway of CO2 fixation.

Biochim. Biophys. Acta. 1784:1873–1898.

Page 165: 2,3-Butanediol production using acetogenic bacteria

7. References 151

Rajagopalan S, Datar RP, Lewis RS. 2002. Formation of ethanol from carbon monoxide via a

new microbial catalyst. Biomass Bioenergy 23:487–493.

Ramió-Pujol S, Ganigue R, Baneras L, Colprim J. 2015. Incubation at 25 °C prevents acid crash

and enhances alcohol production in Clostridium carboxidivorans P7. Bioresour. Technol.

192:296–303.

Regalbuto JR, Almalki F, Liu Q, Banerjee R, Wong A, Keels J. 2017. Hydrocarbon fuels from

lignocellulose. In: Murad S, Baydoun E, Daghir N (eds.), Water, energy & food sustainability in

the Middle East, Springer International Publishing AG, Cham, Switzerland:127–159.

Reidlinger J, Müller V. 1994. Purification of ATP synthase from Acetobacterium woodii and

identification as a Na+-translocating F1F0-type enzyme. Eur. J. Biochem. 223:275–283.

Richter H, Molitor B, Wei H, Chen W, Aristilde L, Angenent LT. 2016. Ethanol production in

syngas-fermenting Clostridium ljungdahlii is controlled by thermodynamics rather than by

enzyme expression. Energy Environ. Sci. 9:2392–2399.

Ripoll V, De Vicente G, Morán B, Rojas A, Segarra S, Montesinos A, Tortajada M, Ramón D,

Ladero M, Santos VE. 2016. Novel biocatalysts for glycerol conversion into 2,3-butanediol.

Process Biochem. 51:740–748.

Russell MJ, Martin W. 2004. The rocky roots of the acetyl-CoA pathway. Trends Biochem. Sci.

29:358–363.

Savakis PE, Angermayr SA, Hellingwerf KJ. 2013. Synthesis of 2,3-butanediol by Synechocystis

sp. PCC6803 via heterologous expression of a catabolic pathway from lactic acid- and

enterobacteria. Metab. Eng. 20:121–130.

Schiel-Bengelsdorf B, Dürre P. 2012. Pathway engineering and synthetic biology using

acetogens. FEBS Letters 586:2191–2198.

Schochetman G, Ou CY, Jones WK. 1988. Polymerase chain reaction. J. Infec. Dis. 158:1154–

1157.

Page 166: 2,3-Butanediol production using acetogenic bacteria

152 7. References

Schuchmann K, Müller V. 2012. A bacterial electron-bifurcating hydrogenase. J. Biol. Chem.

287:31165-31171.

Schuchmann K, Müller V. 2014. Autotrophy at the thermodynamic limit of life: a model for

energy conservation in acetogenic bacteria. Nat. Rev. Microbiol. 12:809-821.

Schuchmann K, Müller V. 2016. Energetics and application of heterotrophy in acetogenic

bacteria. Appl. Environ. Microbiol. 82:4056–4069.

Schuster S. 2011. Expression von identifizierten clostridiellen Butanol-Synthese-Genen im

Zwischenwirt Escherichia coli. Ph. D. Thesis. Ulm University, Germany.

Schweizer H. 2008. Bacterial genetics: past achievements, present state of the field, and

future challenges. Biotechniques 44:633–641.

Shen X, Lin Y, Jain R, Yuan Q, Yan Y. 2012. Inhibition of acetate accumulation leads to

enhanced production of (R,R)-2,3-butanediol from glycerol in Escherichia coli. J. Ind. Microbiol.

Biotechnol. 39:1725–1729.

Sheng Y, Mancino V, Birren B. 1995. Transformation of Escherichia coli with large DNA

molecules by electroporation. Nucl. Acids Res. 23:1990–1996.

Shin HD, Yoon SH, Wu J, Rutter C, Kim SW, Chen RR. 2012. High-yield production of meso-

2,3-butanediol from cellodextrin by engineered E. coli biocatalysts. Bioresour. Technol.

118:367–373.

Siemerink MAJ, Kuit W, Lopez Contreras AM, Eggink G, Van der Oost J, Kengen SWM. 2011.

D-2,3-Butanediol production due to heterologous expression of an acetoin reductase in

Clostridium acetobutylicum. Appl. Environ. Microbiol. 77:2582–2588.

Sikora B, Kubik C, Kalinowska H, Gromek E, Białkowska A, Jędrzejczak-Krzepkowska M,

Schütt F, Turkiewicz M. 2016. Application of byproducts from food processing for production

of 2,3-butanediol using Bacillus amyloliquefaciens TUL 308. Prep. Biochem. Biotechnol.

46:610–619.

Page 167: 2,3-Butanediol production using acetogenic bacteria

7. References 153

Simpson SD, Tran PL, Mihalcea CD, Fung JMY, Liew F. 2011. Production of butanediol by

anaerobic microbial fermentation. EP2307556 A1. European Patent Office, Munich, Germany.

Straub, M. 2012. Fermentative Acetatproduktion durch Homoacetat-Gärung bzw.

Acetacetatbildung. Ph. D. Thesis. Ulm University, Germany.

Sun J, Wang Y. 2014. Recent advances in catalytic conversion of ethanol to chemicals. ACS

Catal. 4:1078–1090.

Sun LH, Wang XD, Dai JY, Xiu ZL. 2009a. Microbial production of 2,3-butanediol from

Jerusalem artichoke tubers by Klebsiella pneumoniae. Appl. Microbiol. Biotechnol. 82:847–

852.

Sun QY, Ding LW, He LL, Sun YB, Shao JL, Luo M, Xu ZF. 2009b. Culture of Escherichia coli in

SOC medium improves the cloning efficiency of toxic protein genes. Anal. Biochem. 394:144–

146.

Suzuki H. 2012. Host-mimicking strategies in DNA methylation for improved bacterial

transformation. In: Dricu A (ed.), Methylation - from DNA, RNA and histones to diseases and

treatment, InTech Rijeka, Croatia:219-236.

Syu MJ. 2001. Biological production of 2,3-butanediol. Appl. Microbiol. Biotechnol. 55:10–18.

Szostková M, Horáková D. 1998. The effect of plasmid DNA sizes and other factors on

electrotransformation of Escherichia coli JM109. Bioelectrochem. Bioenerg. 47:319–323.

Takamizawa A, Miyata S, Matsushita O, Kaji M, Taniguchi Y, Tamai E, Shimamoto S, Okabe

A. 2004. High-level expression of clostridial sialidase using a ferredoxin gene promoter-based

plasmid. Protein Expr. Purif. 36:70–75.

Tan Y, Liu ZY, Liu Z, Li FL. 2015. Characterization of an acetoin reductase/2,3-butanediol

dehydrogenase from Clostridium ljungdahlii DSM 13528. Enzyme Microb. Technol. 79-80:1–7.

Tanner RS. 2007. Cultivation of bacteria and fungi. In: Hurst C, Crawford R, Garland J, Lipson

D, Mills A, Stetzenbach L (eds.) Manual of environmental microbiology, 3rd edition, American

Society of Microbiology, Washington, D.C., WA, USA:69-78.

Page 168: 2,3-Butanediol production using acetogenic bacteria

154 7. References

Tanner RS, Miller LM, Yang D. 1993. Clostridium ljungdahlii sp. nov., an acetogenic species in

clostridial rRNA homology group I. Int. J. Syst. Evol. Microbiol. 43:232–236.

Tong YJ, Ji XJ, Shen MQ, Liu LG, Nie ZK, Huang H. 2016. Constructing a synthetic constitutive

metabolic pathway in Escherichia coli for (R,R)-2,3-butanediol production. Appl. Microbiol.

Biotechnol. 100:637–647.

Tran AV, Chambers RP. 1987. The dehydration of fermentative 2,3-butanediol into methyl

ethyl ketone. Biotechnol. Bioeng. 29:343–351.

Tremblay PL, Zhang T, Dar SA, Leang C, Lovley DR. 2012. The Rnf complex of Clostridium

ljungdahlii is a proton-translocating ferredoxin:NAD+ oxidoreductase essential for autotrophic

growth. mBio 4:e00406–00412.

Tschech A, Pfennig N. 1984. Growth yield increase linked to caffeate reduction in

Acetobacterium woodii. Arch. Microbiol. 137:163–167.

Tsvetanova F, Petrova P, Petrov K. 2014. 2,3-butanediol production from starch by

engineered Klebsiella pneumoniae G31-A. Appl. Microbiol. Biotechnol. 98:2441–2451.

Tyurin M, Kiriukhin M. 2013. Synthetic 2,3-butanediol pathway integrated using Tn7-tool and

powered via elimination of sporulation and acetate production in acetogen biocatalyst. Appl.

Biochem. Biotechnol. 170:1503–1524.

Ungerman AJ, Heindel TJ. 2007. Carbon monoxide mass transfer for syngas fermentation in a

stirred tank reactor with dual impeller configurations. Biotechnol. Prog. 23:613–620.

Valgepea K, De Souza Pinto Lemgruber R, Meaghan K, Palfreyman RW Abdalla T, Heijstra

BD, Behrendorff JB, Tappel R, Köpke M, Simpson SD, Nielsen LK, Marcellin E. 2017a.

Maintenance of ATP homeostasis triggers metabolic shifts in gas-fermenting acetogens. Cell

Syst. 4:505–515.

Valgepea K, Loi KQ, Behrendorff JB, Lemgruber RSP, Plan M, Hodson MP, Köpke M, Nielsen

LK, Marcellin E. 2017b. Arginine deiminase pathway provides ATP and boosts growth of the

gas-fermenting acetogen Clostridium autoethanogenum. Metab. Eng. 41:202–211.

Page 169: 2,3-Butanediol production using acetogenic bacteria

7. References 155

Veibel S, Nielsen JI. 1967. On the mechanism of the Guerbet reaction. Tetrahedron 23:1723–

1733.

Visioli LJ, Enzweiler H, Kuhn RC, Schwaab M, Mazutti MA. 2014. Recent advances on

biobutanol production. Sustain. Chem. Process. 2:15.

Vivijs B, Moons P, Aertsen A, Michiels CW. 2014a. Acetoin synthesis acquisition favors

Escherichia coli growth at low pH. Appl. Environ. Microbiol. 80:6054–6061.

Vivijs B, Moons P, Geeraerd AH, Aertsen A, Michiels CW. 2014b. 2,3-Butanediol fermentation

promotes growth of Serratia plymuthica at low pH but not survival of extreme acid challenge.

Int. J. Food Microbiol. 175:36–44.

Wang A, Wang Y, Jiang T, Li L, Ma C, Xu P. 2010. Production of 2,3-butanediol from corncob

molasses, a waste by-product in xylitol production. Appl. Microbiol. Biotechnol. 87:965–970.

Wang A, Xu Y, Ma C, Gao C, Li L, Wang Y, Tao F, Xu P. 2012a. Efficient 2,3-butanediol

production from cassava powder by a crop-biomass-utilizer, Enterobacter cloacae subsp.

dissolvens SDM. PLoS ONE 7:e40442.

Wang Q, Chen T, Zhao X, Chamu J. 2012b. Metabolic engineering of thermophilic Bacillus

licheniformis for chiral pure D-2,3-butanediol production. Biotechnol. Bioeng. 109:1610–1621.

Wang S, Huang H, Kahnt J, Mueller AP, Köpke M, Thauer RK. 2013a. NADP-specific electron-

bifurcating [FeFe]-hydrogenase in a functional complex with formate dehydrogenase in

Clostridium autoethanogenum grown on CO. J. Bacteriol. 195:4373-4386.

Wang Y, Li L, Ma C, Gao C, Tao F, Xu P. 2013b. Engineering of cofactor regeneration enhances

(2S,3S)-2,3-butanediol production from diacetyl. Sci. Rep. 3:2643.

Weisburg WG, Barns SM, Pelletier DA, Lane DJ. 1991. 16S ribosomal DNA amplification for

phylogenetic study. J. Bacteriol. 173:697–703.

Weiss B, Jacquemin-Sablon A, Live TR, Fareed GC, Richardson CC. 1968. Enzymatic breakage

and joining of deoxyribonucleic acid VI. Further purification and properties of polynucleotide

ligase from Escherichia coli infected with bacteriophage T4. J. Biol. Chem. 243:4543–4555.

Page 170: 2,3-Butanediol production using acetogenic bacteria

156 7. References

White H, Strobl G, Feicht R, Simon H. 1989. Carboxylic acid reductase: a new tungsten enzyme

catalyses the reduction of non-activated carboxylic acids to aldehydes. Eur. J. Biochem.

184:89–96.

Whitham JM, Schulte MJ, Bobay BG, Bruno-Barcena JM, Chinn MS, Flickinger MC, Pawlak JJ,

Grunden AM. 2017. Characterization of Clostridium ljungdahlii OTA1: a non-autotrophic hyper

ethanol-producing strain. Appl. Microbiol. Biotechnol. 101:1615–1630.

Wolin E, Wolin M, Wolfe R. 1963. Formation of methane by bacterial extracts. J. Biol. Chem.

238:2882–2886.

Wood HG. 1991. Life with CO or CO2 and H2 as a source of carbon and energy. FASEB J. 5:156–

163.

Xiao Z, Xu P. 2007. Acetoin metabolism in bacteria. Crit. Rev. Microbiol. 33:127–140.

Xiao Z, Lv C, Gao C, Qin J, Ma C, Liu Z, Liu P, Li L, Xu P. 2010. A novel whole-cell biocatalyst

with NAD+ regeneration for production of chiral chemicals. PLoS ONE 5:e8860.

Xu Y, Chu H, Gao C, Tao F, Zhou Z, Li K, Li L, Ma C, Xu P. 2014. Systematic metabolic

engineering of Escherichia coli for high-yield production of fuel bio-chemical 2,3-butanediol.

Metab. Eng. 23:22–33.

Yan Y, Lee CC, Liao JC. 2009. Enantioselective synthesis of pure (R,R)-2,3-butanediol in

Escherichia coli with stereospecific secondary alcohol dehydrogenases. Org. Biomol. Chem.

7:3914–3917.

Yang C, Meng ZY. 1993. Bimolecular condensation of ethanol to 1-butanol catalyzed by alkali

cation zeolites. J. Catal. 142:37–44.

Yang J, Kim B, Kim H, Kweon Y, Lee S, Lee J. 2015a. Industrial production of 2,3-butanediol

from the engineered Corynebacterium glutamicum. Appl. Biochem. Biotechnol. 176:2303–

2313.

Page 171: 2,3-Butanediol production using acetogenic bacteria

7. References 157

Yang T, Rao Z, Hu G, Zhang X, Liu M, Dai Y, Xu M, Xu Z, Yang ST. 2015c. Metabolic engineering

of Bacillus subtilis for redistributing the carbon flux to 2,3-butanediol by manipulating NADH

levels. Biotechnol. Biofuels 8:129.

Yang T, Rao Z, Zhang X, Lin Q, Xia H, Xu Z, Yang S. 2011. Production of 2,3-butanediol from

glucose by GRAS microorganism Bacillus amyloliquefaciens. J. Basic Microbiol. 51:650–658.

Yang T, Rao Z, Zhang X, Xu M, Xu Z, Yang ST. 2015b. Enhanced 2,3-butanediol production

from biodiesel-derived glycerol by engineering of cofactor regeneration and manipulating

carbon flux in Bacillus amyloliquefaciens. Microb. Cell Fact. 14:122.

Yang T, Zhang X, Rao Z, Gu S, Xia H, Xu Z. 2012. Optimization and scale-up of 2,3-butanediol

production by Bacillus amyloliquefaciens B10-127. World J. Microbiol. Biotechnol. 28:1563–

1574.

Yang TW, Rao ZM, Zhang X, Xu MJ, Xu ZH, Yang ST. 2013. Fermentation of biodiesel-derived

glycerol by Bacillus amyloliquefaciens: effects of co-substrates on 2,3-butanediol production.

Appl. Microbiol. Biotechnol. 97:7651–7658.

Yoo M, Croux C, Meynial-Salles I, Soucaille P. 2016. Elucidation of the roles of adhE1 and

adhE2 in the primary metabolism of Clostridium acetobutylicum by combining in-frame gene

deletion and a quantitative system-scale approach. Biotechnol. Biofuels 9:92.

Zahn JA, Saxena J. 2011. Novel ethanologenic species Clostridium coskatii. US 2011/0229947

A1. Washington, D.C.: U.S. Patent and Trademark Office.

Zeng AP, Sabra W. 2011. Microbial production of diols as platform chemicals: recent

progresses. Curr. Opin. Biotechnol. 22:749–757.

Zhang G, Bao P, Zhang Y, Deng A, Chen N, Wen T. 2011. Enhancing electro-transformation

competency of recalcitrant Bacillus amyloliquefaciens by combining cell-wall weakening and

cell-membrane fluidity disturbing. Anal. Biochem. 409:130–137.

Zhang GL, Wang CW, Li C. 2012a. Cloning, expression and characterization of meso-2,3-

butanediol dehydrogenase from Klebsiella pneumoniae. Biotechnol. Lett. 34:1519–1523.

Page 172: 2,3-Butanediol production using acetogenic bacteria

158 7. References

Zhang L, Sun J, Hao Y, Zhu J, Chu J, Wei D, Shen Y. 2010b. Microbial production of 2,3-

butanediol by a surfactant (serrawettin)-deficient mutant of Serratia marcescens H30. J. Ind.

Microbiol. Biotechnol. 37:857–862.

Zhang L, Yang Y, Sun J, Shen Y, Wei D, Zhu J, Chu J. 2010a. Microbial production of 2,3-

butanediol by a mutagenized strain of Serratia marcescens H30. Bioresour. Technol.

101:1961–1967.

Zhang W, Yu D, Ji X, Huang H. 2012b. Efficient dehydration of bio-based 2,3-butanediol to

butanone over boric acid modified HZSM-5 zeolites. Green Chem. 14:3441–3450.

Page 173: 2,3-Butanediol production using acetogenic bacteria

8. Su

pp

lem

en

t

15

9

8. Su

pp

lem

en

t

Changed AA type

Stop codon

Aliphatic

Acidic, aromatic,

basic, basic,

sulfur-containing,

acidic

Aliphatic,

aliphatic,

aliphatic, basic,

aliphatic, cyclic,

aromatic, basic,

basic

Basic, aliphatic,

aliphatic,

aliphatic, stop

codon, aliphatic

Aliphatic, acidic,

sulfur-containing,

sulfur-containing,

acidic, acidic

Acidic

Original AA type

Hydroxyl

Aliphatic

Aliphatic, hydroxyl,

basic, basic,

aliphatic, aliphatic

Acidic, aliphatic,

aliphatic, sulfur-

containing,

aliphatic, acidic,

aromatic, cyclic,

aliphatic

Sulfur-containing,

aliphatic, aliphatic,

aliphatic, sulfur-

containing, cyclic

Acidic, acidic,

sulfur-containing,

aliphatic, acidic,

acidic

Aliphatic

Changed AA

No AA

G

D, F, R, R, T,

D

G, V, V, K, V,

P, F, H, R

R, V, A, A, no

AA, L

A, E, S, S, D,

D

D

Original AA

Y

A,

V, Y, H, H,

A, G

D, G, V, M,

V, Q, F, P, I

M, A, A, A,

S, P

D, E, S, I, E,

E

V

AA position changed

1283

298, 305

1123, 1126, 1144,

1146, 1148, 1171

2794, 2797, 2806,

2821, 2830, 2836,

2844, 2848, 2854

544, 568, 596, 599,

607, 634

520, 523, 530, 532,

1768, 1769

19

Annotation

Phage terminase

large subunit

Putative corrinoid

protein

methyltransferase putative secretion

protein HlyD

putative transporter

protein

bifunctional

aldehyde/alcohol

dehydrogenase

bifunctional

aldehyde/alcohol

dehydrogenase

hypothetical protein

Locus tag*

CLJU_c03530

CLJU_c07520

CLJU_c07590

CLJU_c07600

CLJU_c16510

CLJU_c16520

CLJU_c18630

Table 12: Results of SNP analysis of C. ljungdahlii [pMTL82151] compared to C. ljungdahlii WT with locus tag, annotation of respective

locus tag, position of changed amino acid (AA) with original and changed amino acid, and original and changed amino acid type

*according to IMG /MER (https://img.jgi.doe.gov/cgi-bin/mer/main.cgi), 25th of april, 2017.)

Page 174: 2,3-Butanediol production using acetogenic bacteria

16

0

8

. Sup

ple

me

nt

Changed AA

type Acidic

Aliphatic, sulfur-

containing

Aliphatic,

aliphatic

Acidic

Aliphatic

Aromatic

Sulfur-

containing

Original AA type

Aliphatic

Acidic, basic

Sulfur-containing,

sulfur-containing

Basic

Acidic

Aliphatic

Aliphatic

Changed AA

D

V, T

I, I

Q

V

F

T

Original AA

A

E, R

M, M

H

E

L

I

AA position changed

1759

802, 1342

1817, 1818

509

91

227

91

Annotation

putative surface/cell-adhesion

protein

signal-transduction and

transcriptional-control protein

signal-transduction and

transcriptional-control protein

putative permease

putative RnfC related NADH

dehydrogenase

putative cobalt ABC transporter

permease

hypothetical protein

Locus tag*

CLJU_c22360

CLJU_c24850

CLJU_c24870

CLJU_c27200

CLJU_c39770

CLJU_c40730

CLJU_c41870

Table 12: Results of SNP analysis of C. ljungdahlii [pMTL82151] compared to C. ljungdahlii WT with locus tag, annotation of respective locus tag,

position of changed amino acid (AA) with original and changed amino acid, and original and changed amino acid type (continued)

*according to IMG /MER (https://img.jgi.doe.gov/cgi-bin/mer/main.cgi), 25th of april, 2017.)

Page 175: 2,3-Butanediol production using acetogenic bacteria

8. Supplement 161

Table 13: Comparison of codon usage in R. terrigena and C. ljungdahlii

Amino acid (abbr.) Triplet Abundance [%] R. terrigena* Abundance [%] C. ljungdahlii**

Alanine (A) GCA 18 41

GCC 30 15

GCG 29 6

GCT 23 38

Cysteine (C) UGC 62 44

UGU 38 56

Aspartic acid (D) GAC 36 24

GAU 64 76

Glutamic acid (E) GAA 51 72

GAG 49 28

Phenylalanine (F) UUC 39 29

UUU 61 71

Glycine (G) GGA 12 42

GGC 39 16

GGG 18 15

GGU 30 27

Histidine (H) CAC 37 28

CAU 63 72

Isoleucine (I) AUA 12 43

AUC 35 17

AUU 53 40

Lysine (K) AAA 54 69

AAG 46 31

Leucine (L) CUA 10 14

CUC 8 7

CUG 36 10

CUU 14 20

*according to Codon usage database (http://www.kazusa.or.jp/codon/cgi-

bin/showcodon.cgi?species=577&aa=1&style=N, 14th of June, 2017.)

**according to Anja Poehlein, Göttingen Genomics Laboratory (Göttingen, Germany)

Page 176: 2,3-Butanediol production using acetogenic bacteria

162 8. Supplement

Table 13: Comparison of codon usage in R. terrigena and C. ljungdahlii

Amino acid (abbr.) Triplet Abundance [%] R. terrigena* Abundance [%] C. ljungdahlii**

Leucine (L) UUA 17 32

UUG 14 17

Methionine (M) AUG 100 100

Asparagine (N) AAC 50 27

AAT 50 73

Proline (P) CCA 21 38

CCC 15 15

CCG 37 8

CCU 27 39

Glutamine (Q) CAA 37 61

CAG 63 39

Arginine (R) AGA 13 48

AGG 8 30

Arginine (R) CGA 11 6

CGC 30 4

CGG 10 5

CGU 29 7

Serine (S) AGC 24 14

AGU 15 21

UCA 14 21

UCC 14 16

UCG 17 3

UCU 16 25

Threonine (T) ACA 17 36

ACC 38 19

ACG 29 7

ACU 15 38

Valine (V) GUA 14 38

GUC 22 12

*according to Codon usage database (http://www.kazusa.or.jp/codon/cgi-bin/showcodon.cgi?species=577&aa=1&style=N, 14th of June, 2017.)

**according to Anja Poehlein, Göttingen Genomics Laboratory (Göttingen, Germany)

Page 177: 2,3-Butanediol production using acetogenic bacteria

8. Supplement 163

Table 13: Comparison of codon usage in R. terrigena and C. ljungdahlii

Amino acid (abbr.) Triplet Abundance [%] R. terrigena* Abundance [%] C. ljungdahlii**

Valine (V) GUG 32 17

GUU 32 33

Tryptophan (W) UGG 100 100

Tyrosine (Y) UAC 44 29

UAU 56 71

Stop codon UAA 48 53

UAG 35 23

UGA 17 24

*according to Codon usage database (http://www.kazusa.or.jp/codon/cgi-bin/showcodon.cgi?species=577&aa=1&style=N, 14th of June, 2017.)

**according to Anja Poehlein, Göttingen Genomics Laboratory (Göttingen, Germany)

Page 178: 2,3-Butanediol production using acetogenic bacteria
Page 179: 2,3-Butanediol production using acetogenic bacteria

Contribution to this work 165

Contribution to this work

The research leading to these results has received funding from the European Union’s Seventh

Framework Programme for research, technological development and demonstration under

grant agreement no. 311815 (SYNPOL project).

Page 180: 2,3-Butanediol production using acetogenic bacteria
Page 181: 2,3-Butanediol production using acetogenic bacteria

Danksagung 167

Danksagung

Aus Datenschutzgründen wurde die Danksagung auf den Seiten 167-169 entfernt.

stark, denn ich fürchte vor der Verteidigung wird es noch schlimmer… You’re simply the best!

Page 182: 2,3-Butanediol production using acetogenic bacteria

168 Danksagung

Page 183: 2,3-Butanediol production using acetogenic bacteria

Danksagung 169

Page 184: 2,3-Butanediol production using acetogenic bacteria
Page 185: 2,3-Butanediol production using acetogenic bacteria

Curriculum vitae 171

Curriculum vitae

Aus Datenschutzgründen wurde der Curriculum vitae auf den Seiten 171 und 172 entfernt.

Page 186: 2,3-Butanediol production using acetogenic bacteria

172 Curriculum vitae

Page 187: 2,3-Butanediol production using acetogenic bacteria

Statement 173

Statement

I hereby declare that the present thesis is the result of my own work and that all sources of

information and support used for its achievement have been mentioned in the text and in the

references. All citations are explicitly identified, all figures contain only the original data and

no figure was edited in a way that substantially modified its content. All existing copies of the

present thesis are identical regarding their form and content.

Ulm, July 2017

______________________________

Catarina Erz