computational analysis and manipulation of the metabolic

163
Computational analysis and manipulation of the metabolic network of drug producing microorganisms INAUGURALDISSERTATION zur Erlangung des Doktorgrades der Fakultät für Chemie und Pharmazie der Albert-Ludwigs-Universität Freiburg im Breisgau Vorgelegt von Dennis Klementz aus Wiesbaden-Dotzheim Freiburg 2017

Upload: others

Post on 17-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Computational analysis and manipulation of the metabolic network of drug

producing microorganisms

INAUGURALDISSERTATION

zur Erlangung des Doktorgrades

der Fakultät für Chemie und Pharmazie

der Albert-Ludwigs-Universität Freiburg im Breisgau

Vorgelegt von

Dennis Klementz

aus Wiesbaden-Dotzheim

Freiburg 2017

Dekan: Prof. Dr. Manfred Jung

Vorsitzender des Promotionsausschusses: Prof. Dr. Stefan Weber

Referent: Prof. Dr. Stefan Günther

Korreferent: Prof. Dr. Andreas Bechthold

Drittprüfer: Prof. Dr. Oliver Einsle

Datum der Promotion: 26.02.2018

“What we observe is not nature itself, but nature exposed to our method of questioning.”

Werner Heisenberg

Index

Index

1 Abstract ......................................................................................................................... 1

2 Introduction ................................................................................................................... 3

2.1 Streptomycetes, an important source of natural drugs .................................................... 3

2.2 Nucleocidin production in Streptomyces calvus ............................................................... 6

2.3 Griseorhodin A, a telomerase inhibitor from a marine Streptomyces strain .................... 9

2.4 Genome-scale metabolic modeling and Flux Balance Analysis ....................................... 11

2.5 ‘Omics’ data and genome-scale metabolic models ......................................................... 14

2.6 Enhanced secondary metabolite production and genome-scale metabolic models ...... 16

3 Materials ....................................................................................................................... 19

3.1 Chemicals ......................................................................................................................... 19

3.2 Equipment ....................................................................................................................... 21

3.3 Consumables .................................................................................................................... 23

3.4 Strains .............................................................................................................................. 24

3.5 Programs & Services ........................................................................................................ 25

3.6 Cultivation media ............................................................................................................. 26

3.6.1 TSB (Tryptone Soy Broth) ......................................................................................... 26

3.6.2 MS (Mannitol Soy) .................................................................................................... 26

3.6.3 SG+ (Soytone Glucose) .............................................................................................. 26

3.6.4 Minor Element Solution ........................................................................................... 27

3.6.5 Buffer ........................................................................................................................ 27

3.6.6 Minimal medium ...................................................................................................... 27

Index

4 Methods ....................................................................................................................... 28

4.1 StreptomeDB ................................................................................................................... 28

4.1.1 Data collection .......................................................................................................... 28

4.1.2 Genomes, taxonomy and gene clusters ................................................................... 28

4.1.3 Generation of the phylogenetic tree ....................................................................... 28

4.2 AdpA in Streptomyces asterosporus ............................................................................... 30

4.2.1 Detection of transposons ......................................................................................... 30

4.2.2 Detection of adpA binding sites ............................................................................... 30

4.2.3 Prediction of adpA regulon ...................................................................................... 30

4.2.4 Potential disrupted genes ........................................................................................ 30

4.2.5 G+C content and skew.............................................................................................. 31

4.2.6 Dotplot ...................................................................................................................... 31

4.3 Genome-scale metabolic modeling ................................................................................. 32

4.3.1 Model reconstruction ............................................................................................... 32

4.3.2 Analysis of essential elements ................................................................................. 32

4.3.3 Correlation plot and area under the curve (AUC) .................................................... 32

4.3.4 Knockout candidate identification ........................................................................... 33

4.3.5 Integration of transcriptomic data ........................................................................... 34

4.3.6 Generation of functional subsets ............................................................................. 35

4.3.7 Minimal medium prediction and validation ............................................................. 35

4.4 Strain cultivation .............................................................................................................. 36

4.4.1 Preculture ................................................................................................................. 36

4.4.2 Permanent culture ................................................................................................... 36

4.4.3 Agar cultures ............................................................................................................ 36

Index

4.4.4 Liquid culture in Erlenmeyer flasks .......................................................................... 36

4.4.5 Dry weight measurement ......................................................................................... 36

4.4.6 Sample preparation for HPLC-MS ............................................................................ 37

4.4.7 HPLC measurement .................................................................................................. 37

5 Results .......................................................................................................................... 38

5.1 StreptomeDB update ....................................................................................................... 38

5.2 Data Overview ................................................................................................................. 38

5.3 Phylogenetic tree ............................................................................................................. 43

5.3.1 NMR and MS prediction ........................................................................................... 45

5.3.2 Scaffold browser ....................................................................................................... 47

5.4 bldA function in Streptomyces calvus ............................................................................. 49

5.4.1 Metabolic model ...................................................................................................... 49

5.4.2 Gene expression ....................................................................................................... 50

5.4.3 Protein production ................................................................................................... 61

5.5 Role of adpA function in Streptomyces asterosporus ..................................................... 67

5.5.1 Genome Overview .................................................................................................... 67

5.5.2 Putative disrupted genes ......................................................................................... 70

5.5.3 Proteomic analysis ................................................................................................... 73

5.6 Enhancement of Griseorhodin A production in Streptomyces avermitilis SUKA22 ........ 75

5.6.1 Metabolic model ...................................................................................................... 75

5.6.2 Minimal-medium validation experiment ................................................................. 76

5.6.3 Knockout candidate identification ........................................................................... 79

5.6.4 Overview of all knockout candidates ..................................................................... 112

5.7 Experimental results ...................................................................................................... 114

Index

6 Discussion & Conclusion .............................................................................................. 118

6.1 StreptomeDB ................................................................................................................. 118

6.2 The effect of bldA in Streptomyces calvus..................................................................... 121

6.3 adpA regulon in Streptomyces asterosporus ................................................................. 125

6.4 Enhancing Griseorhodin A expression in Streptomyces avermitilis SUKA22 ................ 126

6.4.1 Minimal medium validation of the model of S. avermitilis SUKA22 ...................... 126

6.4.2 Knockout candidate identification ......................................................................... 127

6.4.3 Conclusion & Future perspective ........................................................................... 133

7 References .................................................................................................................. 134

8 Licenses ....................................................................................................................... 147

9 Acknowledgments ....................................................................................................... 148

10 Publications ................................................................................................................ 149

11 Conference contributions ............................................................................................ 150

Index

Abbreviations

% percent

°C degree celcius

A adenine

abs. absolute

ACN acetonitrile

ADP adenosine diphosphate

AMP adenosine monophosphate

ATP adenosine triphosphate

AUC area under the curve

bp base pairs

C cysteine

CDS coding sequence

CoA coenzyme A

COG cluster of orthologus groups

∆ deleted gene

DNA desoxiribonucleic acid

EC enzyme class

EtOH ethanol

e.g. exempli gratia

ENA european nucleotide archive

et al. et alii

FAD flavin adenine dinucleotide

FADH flavin adenine dinucleotide (semiquinone)

FADH2 flavin adenine dinucleotide (hydroquinone)

FBA flux balance analysis

fc fold change

G guanine

g gram

H+ proton

h hour

HCl hydrochloric acid

HPLC high performance liquid chromatography

HSCoA coenzyme A

kDa kilo Dalton

KEGG Kyoto encyclopedia of genes and genomes

m mass

m/z mass per charge

MB megabyte

mbp mega base pairs

Index

MeOH methanol

min minute

ml milliliter

mm millimeter

mmol millimole

mol mole

MS mass spectrometry

MS Medium mannitol soy medium

NAD nicotinamide adenine dinucleotide (oxidized)

NADH nicotinamide adenine dinucleotide (reduced)

NADPH nicotinamide adenine dinucleotide phosphate (oxidized)

NADPH nicotinamide adenine dinucleotide phosphate (reduced)

NMR nuclear magnetic resonance spectroscopy

NRPS non-ribosomal peptide synthetase

nt nucleotide substitutions per side

PAPS 3'-phosphoadenosine-5'-phosphosulfate PAP adenosine 3',5'-bisphosphate pH potential of hydrogen

PKBC polyketide backbone chain

PKS polyketide synthase

Ppi pyrophosphate

RNA ribonucleic acid

rpm revolutions per minute

rRNA ribosomal ribonucleic acid

s seconds

SAH S-adenosly homocysteine

SAM S-adenosyl methionine

SBML systems biology markup language

T thymine

tRNA transfer ribonucleic acid

TSB tryptone soy broth

U uracil

V volt

Index

List of figures

Figure 1 Lifecycle of Streptomycetes on solid media ....................................................................... 3

Figure 2 Albert Schatz (left) and Selman Waksman (right) discoverers of Streptomycin................ 4

Figure 3 Bioactive compounds from Streptomyces species ............................................................ 4

Figure 4 Fluorinated natural compounds ......................................................................................... 6

Figure 5 Proposed effects of bldA (25) ............................................................................................. 6

Figure 6 Pointmutation of bldA A ..................................................................................................... 7

Figure 7 Annimycin ........................................................................................................................... 8

Figure 8 Biosynthesis of Rubromycines ........................................................................................... 9

Figure 9 Overview of the important principles and steps in FBA .................................................. 12

Figure 10 E-Flux Method ................................................................................................................ 15

Figure 11 Paclitaxel......................................................................................................................... 16

Figure 12 Genome of S. avermitilis wildtype compared to genome of S. avermitilis SUKA22 ...... 17

Figure 13 Flowchart phylogenetic tree generation ........................................................................ 29

Figure 14 Flowchart knockout candidate identification ................................................................ 34

Figure 15 Coverage of papers reviewed for StreptomeDB ............................................................ 39

Figure 16 Distribution of biological activities among annotated compounds in StreptomeDB .... 40

Figure 17 Distribution of synthesis pathways among annotated compounds in StreptomeDB ... 41

Figure 18 Molecular weight distribution of StreptomeDB compared to DrugBank ...................... 42

Figure 19 Vizualization of the phylogenetic tree ........................................................................... 44

Figure 20 Visualization of predicted mass fragmentation products .............................................. 45

Figure 21 Visualization of NMR data .............................................................................................. 46

Figure 22 Scaffold browser of StreptomeDB ................................................................................. 47

Figure 23 Overview of gene transcription levels............................................................................ 51

Figure 24 Gene regulation in COGs ................................................................................................ 52

Figure 25 Overview of transcriptomic and proteomic data ........................................................... 54

Figure 26 Fold change density in subsystems ................................................................................ 56

Figure 27 Overview of genes in predicted subsystems .................................................................. 57

Figure 28 Correlation curves of Biomass vs Secondary Metabolite production ............................ 58

Index

Figure 29 Glyoxylate and Dicarboxylate metabolism from KEGG .................................................. 60

Figure 30 Hydroxypyruvate isomerase reaction overview from KEGG.......................................... 60

Figure 31 Overview of proteins related to Primary Metabolism organized by predicted functional

subsets ............................................................................................................................................ 66

Figure 32 Overview of Streptomyces asterosporus genome properties ....................................... 67

Figure 33 Dotplot of Nucleocidin cluster in S. calvus and S. asterosporus .................................... 69

Figure 34 Phylogenetic tree based on 16S rRNA............................................................................ 70

Figure 35 Proteomic results and adpA regulon .............................................................................. 73

Figure 36 Griseorhodin A synthesis including co-factors ............................................................... 75

Figure 37 pyrB in pyrimidine metabolism ...................................................................................... 84

Figure 38 pyrB in propanoate metabolism .................................................................................... 86

Figure 39 kbl in glycine, serine and threonine metabolism ........................................................... 90

Figure 40 pta in retinol metabolism ............................................................................................... 92

Figure 41 pta in pyruvate metabolism ........................................................................................... 94

Figure 42 pta in propanoate metabolism....................................................................................... 95

Figure 43 fadE4 in fatty acid degradation ...................................................................................... 97

Figure 44 fadE4 in valine, leucine and isoleucine degradation ...................................................... 99

Figure 45 rocA in arginine and proline metabolism ..................................................................... 107

Figure 46 plcA in ether lipid metabolism ..................................................................................... 109

Figure 47 rpe in pentose phosphate pathway ............................................................................. 111

Figure 48 Overview of the simulated effect of all knockout candidates with relative biomass

production values ......................................................................................................................... 112

Figure 49 Overview of the simulated effect of all knockout candidates with absolute biomass

production values ......................................................................................................................... 112

Figure 50 Differing morphology of the different Griseorhodin A producing SUKA22 mutants ... 114

Figure 51 Dry weights of all mutants and wildtype after 10 days of cultivation in TSB liquid

medium ........................................................................................................................................ 115

Figure 52 Varying Griseorhodin production of the different SUKA22 mutants ........................... 116

Figure 53 Model based data integration and interpretation ....................................................... 122

Index

List of tables

Table 1 HPLC gradient pattern ....................................................................................................... 37

Table 2 Content overview of StreptomeDB in 2013 and 2015 ...................................................... 38

Table 3 Number of rRNA sequences for each filtering step........................................................... 43

Table 4 Number of scaffolds per scaffold-level ............................................................................. 48

Table 5 Predicted functional subsets ............................................................................................. 55

Table 6 Key values from correlation plots of bldA effect on energy distribution .......................... 59

Table 7 Overproduced proteins with active bldA in Streptomyces calvus ..................................... 61

Table 8 Predicted rRNA genes in S. asterosporus .......................................................................... 68

Table 9 Mobile genetic elements in S. asterosporus predicted by ISfinder ................................. 70

Table 10 Comparison of correlation simulation of Griseorhodin A and biomass production in

wildtype and SUKA22 ..................................................................................................................... 80

Table 11 Knockout candidates competing for NADPH ................................................................... 81

Table 12 Knockout candidates competing for Malonyl-CoA. ......................................................... 83

Table 13 Knockout candidates competing for Acetyl-CoA. ............................................................ 87

Table 14 Knockout candidates competing for FAD ........................................................................ 96

Table 15 Knockout candidates competing for NAD ..................................................................... 101

Table 16 Summary of all knockout candidates ............................................................................ 113

Abstract

1

1 Abstract

Streptomycetes are an important source of biogenic drugs. The production of these compounds

is often associated with complex changes within these organisms. In addition to morphological

differentiation, such as the formation of aerial hyphae and spore chains, the cells globally

reorganize their metabolic machinery. This phenomenon is known as metabolic switch. Despite

these extensive metabolic changes, however, the yield in the production of these natural

substances is typically very low.

Genome-scale metabolic models are comprehensive mathematical networks of reactions and

metabolites occurring within an organism. The generation of such a model requires extensive

amounts of background information. Comprehensive scientific databases, such as StreptomeDB

presented in this work, are a valuable tool for the creation of these models. Simulations based

on such models can grant profound insights of the interrelationships of complex metabolic

changes and the underlying regulatory context. This knowledge can in turn be used to efficiently

develop rational metabolic engineering strategies.

In the course of this work, we used a genome-scale metabolic model of Streptomyces calvus to

interpret the effect of the pleiotropic regulator bldA, a key player in the metabolic switch of this

strain (1). For that, we gathered proteomic and transcriptomic data of the naturally bldA

deficient wildtype strain and a mutant with restored bldA function. With the help of the

genome-scale metabolic model, we could organize the data in a functional context and predict

their regulatory effect on the metabolic machinery. The herein gained insights on the natural

regulation of metabolic flux between primary and secondary metabolism were used to develop

new methods for the generation of metabolic engineering strategies. Contrary to common

methods, the aim of this approach is not to directly increase the secondary metabolite

production rate, but to improve the ratio of produced secondary metabolite per produced

biomass by decreasing the efficiency of the competing primary metabolism. This strategy was

used with a genome-scale metabolic model of the already genome minimized host strain

Streptomyces avermitilis SUKA22 to identify knockout candidates that have the potential to

redirect the flux of metabolites and energy in favor of the production of Griseorhodin A, a

Abstract

2

promising telomerase and viral reverse transcriptase inhibitor (2, 3). First semi quantitative

experiments indicate, that growth and production rate are influenced by these knockouts.

Introduction

3

2 Introduction

2.1 Streptomycetes, an important source of natural drugs

Streptomycetes are a group of ubiquitous gram-positive soil bacteria that can be found from the

bottom of the sea to even burning coal seam (4, 5). Their unusual lifecycle on solid media led to

a false classification as fungi when they were discovered. After forming vegetative hyphae that

reach inside the medium, they start forming aerial hyphae when aging. Finally, these areal

hyphae differentiate into spore chains. (6, 7).

Figure 1: Lifecycle of Streptomycetes on solid media. Starting with a single spore, a vegetative mycelium is formed, followed by

aerial hyphae in later phases of culture growth. These aerial hyphae finally differentiate into spore chains. Modified from (7) and

(6).

Introduction

4

Researchers invested huge efforts in exploring this genus since Albert Schatz from the group of

Selman Waksman could isolate Streptomycin from a sample of Streptomyces griseus in 1943,

which was at that time the first effective

antibiotic against Tuberculosis (8). Due

to the exceptional diversity of secondary

metabolite production in

Streptomycetes, this interest has led to

the discovery of an immense number of

natural compounds, representing about

60% of all drugs in clinical use (9).

Blockbuster drugs as the antibiotics Tetracyclin, Daptomycine and Chloramphenicol, the anti-

parasitic Avermectin, the immunosuppressant Rapamycin or the lipase inhibitor Lipstatin are

just a few well known examples from the plethora of bioactive compounds produced exclusively

by Streptomycetes (10–15).

Figure 3: Bioactive compounds from Streptomyces species: Lipstatin (1) Tatracyclin (2) and Daptomycin (3)

Figure 2: Albert Schatz (left) and Selman Waksman (right) discoverers of Streptomycin. (http://sebsnjaes250.rutgers.edu/)

Introduction

5

The important role of these products for human society was most recently honored in 2015 by

presenting the Nobel-prize in medicine to William C. Campbell and Satoshi Ōmura for the

discovery of Avermectines in Streptomyces avermitilis (16). Even though being examined for

nearly 80 years, these bacteria are still a source of novel discoveries and potential new drug

substances (17). Throughout these decades, researchers had various interests and demands in

their studies. In the first half of the 20th century, infectious diseases were a common cause of

death, so the search for new antibiotics was a typical goal in working with Streptomycetes. This

interest changed over the time from infectious to cardiovascular diseases and lately anti-cancer

drugs. Nevertheless, the broad interest in compounds produced by these bacteria also lead to a

huge number of molecules that were isolated but did not show activity to the interest of their

age. These compounds are a rich source of potential new drug substances whose activities have

to be elucidated yet. A comprehensive collection of these compounds can be found in

StreptomeDB1. With more than 4.000 substances, it is currently the largest collection for this

genus and holds a broad spectrum of different background information (Chapter 6.1). Even

though there are already thousands of products known, their number is still increasing rapidly

with emerging methods (17, 18).

1 http://pharmazeutische-bioinformatik.de/streptomedb/

Introduction

6

2.2 Nucleocidin production in Streptomyces calvus

An interesting example is the re-discovery of Nucleocidin production in Streptomyces calvus.

This strain shows a for Streptomycetes unusual deficiency of sporulation and is therefore often

described as ‘bald’ (19). The American Cynamid company was the first to report Nucleocidin

production in S. calvus in 1956 (20). Nucleocidin shows antimicrobial and anti-trypanosomal

activity but is above all special amongst

natural compounds for being

fluorinated (20). Only five fluorinated

natural compounds are known so far

(21, 22). The biosynthesis of Nucleocidin

is not yet fully understood and still a

topic of interest for researchers around

the world (23). The latest results from

Bartholomé et al. showed by radioactive

labeling, that the fluorination does not

take place at an early stage of sugar

formation (24). However, after the first reports of Nucleocidin, it was not possible to establish

its production again for over 60 years (1).

Streptomycetes have a very

complex regulatory system that

is controlling morphological

differentiation and production

of secondary metabolites. This is

one of the reasons why usually

only a small number of the

secondary metabolite clusters of

a Streptomyces strain is

expressed under laboratory

Figure 4: Fluorinated natural compounds: Nucleocidin (1), Fluoroacetone (2), Fluoroacetate (3); Fluorothreonine (4), Fluorooleicacid (5) and Fluorocitrate (6)

Figure 5: Proposed effects of bldA (25)

Introduction

7

conditions. These regulation mechanisms can be divided in cluster specific and pleiotropic

regulation. The latter usually also interacts with the former and a whole network of other

regulators. One of the best studied pleiotropic regulators in Streptomycetes is a gene called

bldA. It is encoding for a t-RNA that is supplying TTA or UUA codons with Leucine (25). This

codon is very rare in Streptomycetes which are known for their high G+C-content. Genes

including such a codon can only be translated if bldA is expressed. This happens commonly in

the static phase of the culture, when aerial hyphae are formed and sporulation is induced (26).

Besides translational regulation, bldA is also reported to play a role in transcriptional regulation

and post translational modification as shown in Figure 5: Proposed effects of bldA. In 2013,

Kalan et al. could show that Streptomyces calvus was the victim of a momentous point mutation

in its bldA gene. This mutation leads to a misfolding of the corresponding t-RNA and thus to a

complete loss of function (1).

Figure 6: Point mutation of bldA (A) bldA sequence from different Streptomyces strains; (B) Secondary structure of mutated

tRNA; (C) Secondary structure of active tRNA; (D) top/bottom: S. calvus with restored bldA function, middle: S. calvus wildtype

with disrupted bldA (1)

Introduction

8

By a complementation experiment, it could be shown, that like reported in Streptomyces

coelicolor, bldA plays also an important role in differentiation and secondary metabolite

production of S. calvus. With a functional bldA

gene, S. calvus showed morphological

differentiation to aerial hyphae and sporulation.

Additionally to Nucleocidin, it started to produce

Annimycin, a polyketide that has no antimicrobial

effect but inhibits sporulation in other

Actinomycetales.

Another example for bod strain is S. asterosporus, which is a close relative of S. calvus. Its

genome is also containing the rare Nucleocidin gene cluster. It has several similar phenotypical

features, such as the inability to form spores or aerial hyphae. However, this incapacity is not

caused by a mutation of bldA but is probably related to adpA, a regulatory gene that belongs to

the same network as bldA. According to Higo et al., adpA is forming a feedback loop with bldA

(26). Mutations in either of these genes caused a bald phenotype in Streptomyces griseus and

suppressed the production of Streptomycin. In Streptomyces coelicolor, these genes also control

the morphological differentiation but do not influence the production of secondary metabolites

(27).

Figure 7: Annimycin

Introduction

9

2.3 Griseorhodin A, a telomerase inhibitor from a marine

Streptomyces strain

Griseorhodin A is a heavily oxidized compound that belongs to the group of structurally related

Rubromycines, which have been known for their vivid red colors since the 60‘s, when

Brockmann isolated a red pigment from Streptomyces collinus (28). Griseorhodin A, like other

Rubromycines, has an aplanar structure caused by their characteristic spiroketal moiety that is

formed by a FAD dependent enzyme. It was first isolated from the marine Streptomyces sp. JP65

(5).

Figure 8: Biosynthesis of Rubromycines (5)

This moiety plays a crucial role in their activity against human telomerase and retroviral reverse

transcriptase (3). Telomerase inhibition is a highly interesting biological property. This enzyme is

a unique ribonucleoprotein that counteracts the small DNA loss that appears during regular

DNA replication. This is done by maintaining the length of dispensable DNA caps at the end of

chromosomes. Telomerase is highly expressed in cancer cells but nearly absent in neighboring

Introduction

10

tissue (29). Although extensive efforts towards the total synthesis of Rubromycines have been

undertaken, the available synthetic routes are still limited and ineffective (30).

Introduction

11

2.4 Genome-scale metabolic modeling and Flux Balance Analysis

Metabolic models are sets of chemical reactions and according reactants that can occur in an

organism. A genome-scale metabolic model is based on the information that can be extracted

from an organism’s genome. Genes and their products are predicted and functionally annotated

to chemical reactions with the help of curated databases such as KEGG or BioCyc (31, 32).

Although this already results in a set of hundreds of reactions and reactants, these drafts still

contain gaps that impede simulations. Additional reactions, such as free diffusion or

spontaneous reactions have to be added and the results of the automated annotations have to

be carefully checked (33). In the past, the reconstruction of a genome-scale metabolic model

was a laborious task that included a lot of manual curation and correction (34). With strongly

increasing number of available sequenced genomes, the quality of automated annotation

databases is rapidly improving (35). Since genome-scale metabolic models are gaining

popularity, there are also automated gap-filling algorithms available (36–38). The latest release

of the RAST server and modelSEED is capable of generating directly usable models within few

hours. Once purged from essential gaps and errors, a model can be used for simulations and is

usually iteratively optimized with experimental data (33)

One of the most popular simulations that can be performed with such a model is constraint-

based Flux Balance Analysis (FBA). Metabolic models are usually stored in lists of

stoichiometrical reactions and reactants with varying amounts of additional information. For

FBA this data has to be translated to a mathematical matrix (S) of m rows (number of reactions)

and n columns (number of reactants). Each field of this matrix contains a value that displays if a

reaction consumes or produces a certain reactant. Consumed substrates are described by

negative numbers. The unit of these numbers is usually mmol per gram dry weight per hour

[mmol/g*h]. A vector (v) contains all reaction fluxes of the network. The upper and lower

boundary of the turnover rate of a reaction is defined in the model. A negative lower boundary

value herby defines a reversible reaction. To be able to simplify the differential reaction

equations to linear ones, the network has to be assumed to be in steady state S x v = 0. With

linear programming, it is possible to find a solution for this equation network where a particular

Introduction

12

objective function has an optimal value, typically biomass generation, energy equivalents or

secondary metabolites.

Figure 9: Overview of the important principles and steps in FBA (39)

This approach is a powerful tool to gain insights in complex metabolic contexts without the

need for detailed knowledge of enzyme kinetics for every reaction (39). Lewis et al. could prove

based on the model organism E. coli that results from FBA, which are typically the

stoichiometrically most efficient solutions with minimized total flux, are consistent with results

from omic analysis for growth optimization. Edwards and Palsson studied the influence of gene

deletions in E.coli on its metabolic capabilities with a genome scale metabolic model. Their

research grants a deeper understanding of essential metabolism in general. They could show,

that a cell can still maintain growth, even if stripped from central metabolic genes except from

glycolysis, tricarboxylic acid cycle and pentose phosphate pathway under defined conditions

(40). Furthermore, FBA simulations use little computational power and can be performed in a

few milliseconds, even on older machines.

Introduction

13

Despite these huge benefits, metabolic simulations have some limitations. Though simulations

assume a steady state, it is hard to predict the effect of changing concentrations. Moreover,

regulatory mechanisms are typically not taken into account, resulting in inaccurate quantitative

predictions.

Introduction

14

2.5 ‘Omics’ data and genome-scale metabolic models

The analysis of ‘omics’-data and the prediction of resulting effects on a cell is still a challenging

task. Extensive changes in a cells metabolic machinery, such as morphological differentiation are

controlled by complex regulatory systems which are often not yet fully understood or even not

known so far (41).

Interpreting metabolomic data with the help of genome-scale metabolic models seems to be

fairly straightforward and is performed successfully, but they also offer the possibility to

incorporate other omics data (42, 43). Transcriptomic and proteomic data grant insights on the

regulation of the metabolic machinery of an organism on different levels, but are commonly not

correlating very well due to posttranslational modifications (44). This makes them particularly

difficult to interpret in context with simulations of conversion rates of the reactions that are

catalyzed by them (45).

However, there are several methods available to approach this problem. Experimental results

can be used to constrain the model and create a specific model for a certain phenotype or

disease. Such a model can be used to organize the gathered data or to compare predicted with

measured results in order to find differences that help to improve the model and provide a

deeper insight in underlying biological processes (46, 47).

Colijn et al. have successfully shown, that Flux Balance Analysis is a powerful tool to predict the

effect of different gene expression levels on the metabolic capabilities of an organism.

Constraints, the possible ranges of turnover rates of a certain reaction in the simulations, are

adjusted according to fold changes of annotated genes. They used this ‘valve’-approach to

analyze transcriptome profiles of M. tuberculosis after being exposed to several inhibitors of its

metabolism.

Introduction

15

Figure 10: E-Flux Method: Gene expression data is used to adjust the constraints of corresponding reactions in the metabolic

network. (left) Gene expression data (Green: lower expression; Red: higher expression); (middle) Metabolic network: Thinner

pipes indicate smaller constraints of reactions; (right) Predicted flux rates (48)

They could successfully identify seven out of eight known inhibitors of mycolic acid pathway and

some putative new drug substances. The resulting predictions can provide a precise

understanding of the effect of differentially expressed genes on the metabolism as a whole (48).

Introduction

16

2.6 Enhanced secondary metabolite production and genome-

scale metabolic models

Although yields of over 90% per unit of energy (e.g. glucose) are possible for the

biotechnological production of simple small molecules such as technical alcohols that come

from primary metabolism, the efficient

production of more complex molecules like

Griseorhodin A is much more complicated (5,

49). Often, the desired products can only be

isolated in low amounts and under tremendous

expenditure. A famous example is Paclitaxel

(50). One gram of this important anti-cancer

drug has to be extracted of the bark of twelve

taxus brevifolia trees, followed by about nine weeks of chemical refinement. A total chemical

synthesis of this product has been established in 1994, but is compared to the conventional

extraction still not lucrative (51). Laborious production and purifications like this are the reason

why total chemical synthesis is still often the preferred way, even though pharmaceutical and

chemical industry are making huge efforts to become independent from finite fossil resources.

Starting with unguided mutagenesis and selection, the development of modern genetic

methods has led to a rapid progress in the last 30 years (52). Though secondary metabolites are

often only produced under special conditions, a common method is the isolation of the

responsive gene cluster followed by the heterologous expression in a better controllable host

strain (2). The most commonly used organisms are Escheria coli and Saccharomyces cerevisiae,

but also Streptomycetes are excellent hosts for heterologous expression of secondary

metabolites. They are able to produce an enormous variety of complex compounds from

renewable resources and stand out from other popular expression systems inter alia by their

ability to express very large gene cluster or by their efficient export systems (53). The group of

Prof. H. Ikeda has developed a genome-optimized mutant of the industrially used strain

Figure 11: Paclitaxel

Introduction

17

Streptomyces avermitilis. These so called SUKA strains have proven superior to many natural

hosts in terms of secondary metabolite production (2).

Figure 12: Genome of S. avermitilis wildtype compared to genome of S. avermitilis SUKA22. white arrows: OriC; black arrows:

rRNA genes; black bar: core region containing essential genes; letters: cutting pattern of endonuclease, modified regions are in

grey. Modified from (2) Copyright (2001) National Academy of Sciences.

By deleting about 1.7 mbp from its chromosomes arms, they could remove several secondary

gene clusters, competing for resources supplied by primary metabolism that are needed for the

generation of heterologous expressed compounds.

Other common methods that are applied in order to increase the outcome of fermentations are

overexpression of the desired products gene cluster or its transcriptional regulator, export

mechanisms, genes that are responsible for supply of precursors or knockouts of degrading

reactions (52).

Cofactors also play an important role in the production of secondary metabolites. They

transport precursors or provide for example the reductive or oxidative energy that is necessary

for enzymatic reactions. Experiments on increasing the supply of cofactors such as NADPH have

shown that influencing the redox balance of a cell can lead to a disproportional loss of

metabolic efficiency or even cell death. There are several emerging techniques available for the

optimization of cofactor supply but the understanding of the effects in the complex and strongly

interlinked metabolism of a cell regulations systems preventing the cell from redox imbalance

damage is still little (54).

The selection of a host for heterologous expression is a common step towards an enhanced

production of a certain secondary metabolite, but is often not guided by the evaluation of its

metabolic capabilities but by the availability and selection of strains that are available in the

researchers’ lab. Flux Balance Analysis can be used to identify the putatively most efficient host

for a certain secondary metabolite. Zakrewski et al. offered with their software MultiMetEval

Introduction

18

that already included a dataset of 38 genome scale metabolic models of Actinomycetales and 15

secondary metabolite synthesis pathways a powerful tool for this problem. Based on Pareto

font calculation, they could show that the correlation between primary and secondary

metabolism can differ significantly between different species (55).

In the last years, metabolic models have been utilized, often in combination with omics data, to

develop and enhance overproduction strategies. Genome-scale simulations offer the

opportunity to rationally guide metabolic engineering. Not only manipulation candidates can be

predicted but also the cellular metabolic response to the predicted manipulations (56). This

offers in-depth insights in the underlying effect of wanted and unwanted phenotypes and helps

to delimit the number of time consuming and expensive lab experiments. Dang et al. could

recently demonstrate the power of predictions based on a genome-scale metabolic model by

enhancing the rapamycin production in S. hygrosopicus. By in silico studies of the effect of

decreased and increased flux values on the product of the ratio of weighted and dimensionless

specific growth rate and specific rapamycin production rate they could identify pfk as knockout

and dahP as well as rapK as overexpression candidates. These manipulations lead to an

increased Rapamycin production rate of 142.3% compared to the wildtype strain (57).

Materials

19

3 Materials

Unless otherwise stated, the location of the supplier is in Germany.

3.1 Chemicals

Substance Item

number

Supplier

Acetonitrile HN44.2 Carl Roth GmbH , Karlsruhe

Agar-Agar 5210.1 Carl Roth GmbH , Karlsruhe

Ammonium sulfate 3746.2 Carl Roth GmbH, Karlsruhe

Antifoam Y30 Y30 A5758 Sigma-Aldrich Chemie GmbH,

Steinheim

Apramycin sulfate salt A2024 Sigma-Aldrich Chemie GmbH,

Steinheim

Bacto maltextract 291650 Becton, Dickinson & Co, Sparks USA

Calcium carbonate P013.1 Carl Roth GmbH , Karlsruhe

Calcium chloride CN93.1 Carl Roth GmbH, Karlsruhe

Calibration buffer pH 4.01 HI7004 Hanna Instruments, Woonsocket

USA

Calibration buffer pH 6.86 HI7006 Hanna Instruments, Woonsocket

USA

CASO bullion X938.2 Carl Roth GmbH , Karlsruhe

EDTA SLBC7327 Sigma-Aldrich Chemie GmbH,

Steinheim

Erythromycin 4166.1 Carl Roth GmbH, Karlsruhe

Ethanol 99,9% P076.1 Carl Roth GmbH, Karlsruhe

Ethyl acetate 6784.3 Carl Roth GmbH , Karlsruhe

Formic acid 4724.1 Carl Roth GmbH , Karlsruhe

D(+)-Glucose HN06.2 Carl Roth GmbH, Karlsruhe

Materials

20

Hydrochloric acid 37% 9277.1 Carl Roth GmbH , Karlsruhe

Iron(II) sulfate heptahydrate 3722.1 Carl Roth GmbH, Karlsruhe

Isopropanol AE73.1 Carl Roth GmbH , Karlsruhe

Dipotassium phosphate P749.1 Carl Roth GmbH, Karlsruhe

Kanamycin sulfate T832.1 Carl Roth GmbH, Karlsruhe

Lab Lemco meatextract LP0029 Oxoid Ltd, Basingstoke UK

L-Lysin hydrochloride 1700.2 Carl Roth GmbH, Karlsruhe

Magnesium chloride hexahydrate HN03.1 Carl Roth GmbH, Karlsruhe

Magnesium sulfate heptahydrate T888.1 Carl Roth GmbH, Karlsruhe

Manganese(II) chloride tetrahydrate 0276.1 Carl Roth GmbH, Karlsruhe

D(-)-Mannit 4175.1 Carl Roth GmbH, Karlsruhe

2-Mercaptoethanol 4227.3 Carl Roth GmbH , Karlsruhe

Methanol AE71.2 Carl Roth GmbH , Karlsruhe

Nalidixic acid CN32.1 Carl Roth GmbH, Karlsruhe

Natrium chloride HN00.1 Carl Roth GmbH, Karlsruhe

Mononatrium hydrogenphosphate

dihydrate

T879.1 Carl Roth GmbH, Karlsruhe

RNAseZAP SCBC9516V Sigma-Aldrich Chemie GmbH,

Steinheim

D(+)-Saccharose 4621.1 Carl Roth GmbH, Karlsruhe

Sodium hydroxide 40% 4347.1 Carl Roth GmbH , Karlsruhe

Sucrose 4621.1 Carl Roth GmbH , Karlsruhe

Soy flour 2004246 W.Schoenenberger, Magstadt

Soluble starch 4701.1 Carl Roth GmbH, Karlsruhe

L-Tryptophan 1739.1 Carl Roth GmbH, Karlsruhe

Zinc sulfate heptahydrate K301.1 Carl Roth GmbH, Karlsruhe

Materials

21

3.2 Equipment

Device Model Supplier

Bioanalyzer G2938L Agilent Technologies, Waldbronn

Bioreactor Bioflo/CelliGen 115 New Brunswick, Eppendorf AG,

Hamburg

Bunsen burner Fireboy plus Integra Biosciences, Fernwald

Centrifuge for centrifuge

tubes

5810R Eppendorf AG, Hamburg

Centrifuge for reaction

vessels

5415R Eppendorf AG, Hamburg

Column holder Guard Column Holder REF

718966

Marcherey-Nagel GmbH & Co. KG,

Düren

DO probe InPro 6800 Mettler Toledo GmbH, Gießen

Erlenmeyer flasks 250ml, with baffle Duran Group, Mainz

Erlenmeyer flasks 500 ml, with baffle Duran Group, Mainz

HPLC 1290 Serie Agilent Technologies, Waldbronn

HPLC column EC 150/2 Nucleodur

100-5 C18 ec REF

760008.20

Marcherey-Nagel GmbH & Co. KG,

Düren

Hybridization system Hybridisation System 4 Roche Diagnostics GmbH,

Mannheim

Incubation chamber Galaxy 48R New Brunswick, Eppendorf AG,

Hamburg

Mass spectrometer 6460Triple Quadrupol Agilent Technologies, Waldbronn

Mass spectrometer 6520 Accurate-Mass Q-TOF

LC/MS

Agilent Technologies, Waldbronn

Microarray Scanner G2505C Agilent Technologies, Waldbronn

Nanodrop ND-1000 PeqLab, Erlangen

Materials

22

pH probe 405.DPAS-SC-K8S/225 Mettler Toledo GmbH, Gießen

Pre-column EC 4/2 Nucleodur 100-5

C18 ec REF 761932.20

Marcherey-Nagel GmbH & Co. KG,

Düren

Shaking incubator Innova 44 New Brunswick, Eppendorf AG,

Hamburg

SpeedVac 5301 Eppendorf AG, Hamburg

Sterile bench MSC 1.8 Thermo electron LED GmbH,

Langenselbold

Thermomixer compact/comfort Eppendorf AG, Hamburg

Ultrasound clipper Covaris S-Series KBiosciences, Hoddeston UK

Vortex VV3 VWR, Darmstadt

Water filter unit Millipore Billerica, USA

Water cooling FL601 Julabo GmbH, Seelbach

Materials

23

3.3 Consumables

Type Description Supplier

Centrifuge tubes 15 ml N459.1 Carl Roth GmbH, Karlsruhe

Centrifuge tubes 50 ml X063.1 Carl Roth GmbH, Karlsruhe

Disposable syringe 20 ml EP98.1 Carl Roth GmbH, Karlsruhe

Disposable syringe 5 ml EP96.1 Carl Roth GmbH, Karlsruhe

Glass vials 2 ml, clr, WrtOn Agilent Technologies, Waldbronn

Petri dishes ALA5.1 Carl Roth GmbH, Karlsruhe

Pipette tips Standard Universal Carl Roth GmbH, Karlsruhe

Reaction vessel 1.5 ml CH76.1 Carl Roth GmbH, Karlsruhe

Reaction vessel 2 ml CK06.1 Carl Roth GmbH, Karlsruhe

Semi-micro cuvette XK26.1 Carl Roth GmbH, Karlsruhe

Sterile filter 0.22 µm P666.1 Carl Roth GmbH, Karlsruhe

Vial Cap 9 mm blue screw Agilent Technologies, Waldbronn

Vial Insert 250 μl Agilent Technologies, Waldbronn

Materials

24

3.4 Strains

Used strains Sequence available Model

Streptomyces sp. Tü6071

Streptomyces albus J1074

Streptomyces avermitilis K139

Streptomyces avermitilis SUKA22 ()

Strains conjugated with pMP31 Sequence available Model

Streptomyces albus J1074+ pMP31 X

Streptomyces sp. Tü6071 + pMP31 X X

Streptomyces avermitilis K139 + pMP31 X

Streptomyces avermitilis SUKA22 + pMP31 X

Knockout variants of S. avermitilis SUKA22 Sequence available Model

Streptomyces avermitilis SUKA22 Δpta X ()

Streptomyces avermitilis SUKA22 Δkbl X ()

Streptomyces avermitilis SUKA22 ΔfadE4 X ()

Streptomyces avermitilis SUKA22 ΔrocA X ()

Streptomyces avermitilis SUKA22 ΔplcA X ()

Streptomyces avermitilis SUKA22 ΔpyrB X ()

Streptomyces avermitilis SUKA22 Δpta + pMP31 X

Streptomyces avermitilis SUKA22 Δkbl + pMP31 X

Streptomyces avermitilis SUKA22 ΔfadE4 + pMP31 X

Streptomyces avermitilis SUKA22 ΔrocA + pMP31 X

Streptomyces avermitilis SUKA22 ΔplcA + pMP31 X

Streptomyces avermitilis SUKA22 ΔpyrB + pMP31 X

Materials

25

3.5 Programs & Services

Name Version

antiSMASH 3.0

BioCommand 3.30

Bioconductor 3.5

Biopython 1.70

BLAST 2.6.0

Chemdraw 16

ChemStation for LC Systems Rev. B. 03.02 [341]

COBRApy 0.8.2

COBRAtoolbox 2.0.5

DEVA 1.2

FAME 21.10.2013

Gurobi linear solver 7.5

ISfinder 2017.10.03

MassHunter Qualitative Analysis Rev. B.04.00

MatLab R2013a

Mauve 2.4.0

MEGA 6

MultiMetEval 14.12.2012

Origin Pro 9.0

Python 2.7.5

R 3.3.3

RNAmmer 1.2

Materials

26

3.6 Cultivation media

All media were used as liquid and solid media. The solid Media included 1% Agar-Agar.

3.6.1 TSB (Tryptone Soy Broth)

Ingredient Content

CASO-Bullion 30 g

Tab water ad 1000 ml

3.6.2 MS (Mannitol Soy)

Ingredient Content

D(-)-Mannit 20 g

Soy flour 20 g

Magnesium chloride hexahydrate 2 g

Tab water ad 1000 ml

3.6.3 SG+ (Soytone Glucose)

Ingredient Content

Glucose 20 g

Soytone 10 g

Calcium carbonat 2 g

L-Valine 1 g

Cobalt chloride 2.34 mg

Tab water ad 1000 ml

Materials

27

3.6.4 Minor Element Solution

Ingredient Content

Zinc sulfate heptahydrate 1 g

Iron(II) sulfate heptahydrate 1 g

Manganese(II) chloride tetrahydrate 1 g

Calcium chloride 1 g

Purified water ad 1000 ml

3.6.5 Buffer

Ingredient Content

Dipotassium phosphate 8.7 g

Mononatrium hydrogenphosphate dihydrate 7.8 g

Purified water ad 1000 ml

3.6.6 Minimal medium

Ingredient Content

L-Lysine hydrochloride 0.286 g

L-Tryptophan 4.714 g

Ammonium sulfate 2 g

Magnesium sulfate heptahydrate 0.6 g

Carbon source 5 g

Minor Element Soulution 1 ml

Buffer 150 ml

Purified water ad 1000 ml

Methods

28

4 Methods

4.1 StreptomeDB

4.1.1 Data collection

The Data was automatically extracted from literature and manually curated according to the

protocol presented in the first StreptomeDB publication (58). Additionally, we put a special

focus on compounds described in the most recent publicly available full texts. This includes

hundreds of compounds that cannot yet be found in popular databases such as PubChem or

Zinc (59).

4.1.2 Genomes, taxonomy and gene clusters

If available, the organisms have been assigned to their taxonomic ID and freely accessible

complete genome sequences from GenBank (60, 61). Compounds were linked to their gene

clusters in DoBISCUT and MiBIG (62, 63).

4.1.3 Generation of the phylogenetic tree

26,201 16S rRNA sequences were collected from ArB silva and ENA (61, 64). They were matched

to the strains represented in StreptomeDB. Sequences with low quality or an untypical length

were omitted. For strains where several sequences were available (in some cases over 150 from

different sources), a consensus sequence was calculated. For this purpose, all sequences

available for one strain were compared with each by BLAST analysis. The sequences with the

most hits were used to calculate a consensus sequence with the dump_consensus function of

Biopython (65). Subsequently, all sequences were aligned with ClustalW (66). The resulting

alignment was analyzed with the MEGA software package using maximum likehood and

maximum parsimony algorithms with 250 bootstrap replications (67). Sub-cluster within 0.07 nt

substitutions per side of the same ancestor node were summarized using DendroPy (68).

Visualization and final editing was performed with the ETE Toolkit (69).

Methods

29

Figure 13: Flowchart phylogenetic tree generation

Methods

30

4.2 AdpA in Streptomyces asterosporus

4.2.1 Detection of transposons

Transposons were detected with the services from ISfinder (70). The resulting Blast analysis was

filtered by length and quality. Only matches with more than 95% sequence identity and a p-

value below 0.005 were considered.

4.2.2 Detection of adpA binding sites

According to Yao et al., the sequence of the adpA binding site is 5’–TGGCSNGWWY-3’ (S is G or

C, W is A or T, Y is T or C, and N is any nucleotide) (71). The genome of S. asterosporus was

screened with the regular expression “TGGC[GC][GCAT]G[AT][AT][TC]” on sense and antisense

strand.

4.2.3 Prediction of adpA regulon

Genes potentially interacting with adpA were extracted from STRING database (72). In addition

to genes that have been found to interact directly with adpA in S. griseus and S. coelicolor, this

also includes the genes of the next level of interaction. This includes all genes that are in direct

relation with those that interact directly with adpA. If these genes are present in the genome of

S. calvus, they were included in the circos diagram (Figure 25).

4.2.4 Potential disrupted genes

In order to find genes that have been disrupted by transposons, the area around the transposon

sequences have been scanned. For this purpose, gene segments of 2500 bp were extracted

before and after the corresponding transposon and joined together without the transposon

sequence. The newly created sequences were blasted against UniProt/SwissProt for hits that

were not found in the original annotation (73).

Methods

31

4.2.5 G+C content and skew

The average G+C content and G+C skew were calculated in chunks of 5000 bp with the following

formulas:

G+C content: 100% ∙𝐺+𝐶

𝐺+𝐶+𝐴+𝑇

G+C skew: 𝐺−𝐶

𝐺+𝐶

4.2.6 Dotplot

The dotplot of the Nucleocidin clusters form S. asterosporus and S. calvus were computed with

the services of YASS with default options (74). The gene clusters were extracted according to

antiSMASH annotations.

Methods

32

4.3 Genome-scale metabolic modeling

For the simulation of the model, several functions from COBRApy were used together with the

Gurobi linear solver2 (75).

4.3.1 Model reconstruction

The models were reconstructed with the services provided by RAST and ModelSEED followed by

manual curation (76, 77). RAST automated annotation was performed with options adapted to

the results of the respective sequencing project with the latest FIGfams available. Errors and

gaps that could not be fixed automatically or were caused by the gap-filling algorithm used

during the automated reconstruction were identified manually or by validation experiments

that will be described later.

4.3.2 Analysis of essential elements

Essential reactions were detected by using the single_reaction_deletion method from COBRApy.

Reactions that lead to a flux value of 0.0 mmol/(g*h) for the objective function were considered

essential for this objective function.

Essential genes were detected by using the single_gene_deletion method from COBRApy. Genes

that lead to a flux value of 0.0 mmol/(g*h) for the objective function were considered essential

for this objective function.

Essential metabolites were detected by stepwise knocking the transport reactions from the

boundary condition out. Metabolites that are transported by reactions whose knockout resulted

in a flux value of 0.0 mmol/(g*h) for the objective function were considered to be essential. For

further analysis, these reactions were enabled again. This approach results in the minimal set of

boundary reactions that is necessary for the objective function.

4.3.3 Correlation plot and area under the curve (AUC)

The correlation curve was calculated for two objective functions in dependence of each other.

First, the optimal flux value of secondary metabolite production was determined. Next, the

2 www.gurobi.com

Methods

33

constraints of the final reaction of secondary metabolite production were both set to this value

to force a fixed turnover rate. Then the model was iteratively optimized on growth rate, while

the fixed turnover rate of secondary metabolite production was stepwise reduced to zero. The

area under the resulting curve was calculated according to Simpson’s rule (78). The relative AUC

was calculated in relation to the area stretched between the two maximal values.

4.3.4 Knockout candidate identification

Essential genes were identified by single gene deletion simulation. For this, the constraints of all

reactions annotated to a certain gene are set to zero. If the following optimization on biomass

production resulted in a value equal or close to zero, the respective gene was considered to be

essential. These genes were no longer taken into account.

Knockout simulations were evaluated by their influence on the generation of biomass and

correlation of biomass and secondary metabolite production. For this, the area under the curve

(AUC) of a correlation plot was used. Knockouts resulting in a significantly reduced production

rate of the secondary metabolite were also omitted.

Reactions from primary metabolism that are directly competing for resources with secondary

metabolite production were identified by comparing their flux values during the different steps

of the correlation plot generation. Reactions showing a decreased flux value while secondary

metabolite production is increased were considered to be competing for resources. The found

candidates were classified according to the precursors needed for secondary metabolite

synthesis. In addition, two candidates were determined on the basis of single knockout

correlation analysis only, independent of precursors. Finally, all knockout candidates were

screened for suitability. The more exclusively reactions are annotated to the knockout candidate

the better.

Methods

34

Figure 14: Flowchart knockout candidate identification

4.3.5 Integration of transcriptomic data

In order to integrate gene expression data into the model, a Flux Balance Analysis was carried

out first. The production of biomass was optimized without any experimental data. The

resulting flux rates of all reactions have been used as references. For every gene that was

significantly regulated and annotated in the model, the reference value of all corresponding

reactions was multiplied with the corresponding fold change. The boundary values of these

reactions were changed accordingly. For reactions with more than one annotated gene, the

average of fold change of all annotated genes was used. The resulting model was compared to

the reference model with the help of correlation plots as described above.

Methods

35

4.3.6 Generation of functional subsets

Gene expression data was analyzed with the help of the metabolic model. The model was

individually optimized for the production of Nucleocidin, Annimycin and biomass. For each

simulation, a subset of genes was compiled, based on whether their reactions showed a flux

rate other than zero. Additional subsets were created by subtracting reactions involved in the

generation of biomass from the reaction sets needed for secondary metabolite production. This

resulted in a set of genes from primary metabolism that are required exclusively for the

production of precursors of secondary metabolites.

4.3.7 Minimal medium prediction and validation

The composition of the minimum medium should not consist of more than two amino acids, a

carbon source and a defined salt composition. Tryptophan and lysine were chosen as amino

acids, starch, glucose, sucrose, and mannitol were tested as carbon sources. The salt solution

and the ratio between nitrogen to carbon sources were chosen according to the minimum

medium composition from Hopwood et al. (79). The simulation of the media was carried out by

manipulating the exchange reactions between defined model space and boundary condition.

The constraints of reactions transporting substances that were not necessary to grow the strains

in vitro were set to zero. In case no growth could be simulated, the model was revised

accordingly. The flux rates of lysine and tryptophan uptake in the simulation of cell growth were

used to determine the optimal ratio of the two amino acids for the minimal medium.

Methods

36

4.4 Strain cultivation

4.4.1 Preculture

Inoculation was done with a cropped 1 ml pipet tips that was pricked in frozen sucrose

permanent culture. The tip was then ejected into the prepared medium. All strains were

cultivated for 48 h in 50 ml Tryptone Soy Both (TSB) in 250 ml Erlenmeyer flasks with baffle at

28 °C and 180 rpm in a shaking incubator.

4.4.2 Permanent culture

The respective culture was cultivated for 48 h in 50 ml TSB at 28 °C and 180 rpm in 250 ml

Erlenmeyer flasks with baffle in a shaking incubator. The cells were transferred into 50 ml

centrifuge tubes and centrifuged for 10 min at 5000 rpm. The supernatant was discarded. The

pellet was redispersed in 10 ml of 20% sucrose solution, transferred to a 15 ml centrifuge tube

and stored at -20 °C.

4.4.3 Agar cultures

The sterile petri dishes were poured with freshly autoclaved medium under a sterile bench and

stored there until they solidified. Depending on the antibiotic used, it was added to the medium

before autoclaving, or the plate was subsequently treated with a sterile filtered solution. The

cultures were applied to the agar by dilution smear or with a Drigalski spatula. Plates were

cultivated at 28 °C and stored at 6 °C for up to three months.

4.4.4 Liquid culture in Erlenmeyer flasks

Strains were cultivated in 500 ml Erlenmeyer flasks with baffle and a stainless steel spiral.

200 ml cultivation medium was inoculated with 1 ml pre-culture. They were incubated at 28 °C

and 180 rpm in a shaking incubator.

4.4.5 Dry weight measurement

3 ml (2 ml +1 ml) of the bacterial culture were centrifuged for 10 minutes at maximal speed in a

tared 2 ml reaction flask. The supernatant was discarded. Pellets were tree times washed with

1 ml of purified water. Remaining pellets were dried in a SpeedVac overnight at 30 °C.

Methods

37

4.4.6 Sample preparation for HPLC-MS

1 ml of frozen bacterial culture was thawed, mixed with a standard solution and acidified with

150 µl of 0.5 M HCl. Cells were mechanically broken with 300 mg of glass beads in a ball mill at

room temperature. The resulting solution was three times extracted with 0.5 ml of ethyl

acetate. The organic phase was collected and dried overnight in a SpeedVac at room

temperature.

Dried samples were either stored at -20 °C or dissolved in MeOH/H2O (50/50), if necessary in an

ultrasound bath. The resulting solution was centrifuged for 10 minutes at maximal speed before

it was transferred to an injection vial with insert.

4.4.7 HPLC measurement

HPLC measurement was performed on an Agilent 1100series with a Nucleodur 100-5 C18 RP

(150 mm x 2 mm) column from Marcherey-Nagel, including a pre-column (4 mm x 2 mm) of the

same material. A gradient from 90:10 H2O/ACN (0.1% formic acid) to 02:98 H2O/ACN (0.1%

formic acid) was used with a flowrate of 0.2 ml/min.

Table 1: HPLC gradient pattern

Time [min] Gradient [H2O/ACN]

00 90:10

03 90:10

23 02:98

26 02:98

28 90:10

35 90:10

MS measurement was performed with an Agilent 6520 Accurate Mass Q-TOF LC/MS in positive

mode with a collision energy of 10, 20 and 40 V. Additionally, an UV Diode Array Detector was

uses at wave lengths of 235, 245, 310, 366 and 488 nm.

Results

38

5 Results

5.1 StreptomeDB update

StreptomeDB is currently the largest collection of metabolites reported to be produced in

Streptomycetes (58). With its advanced features, it allows for navigation through this vast

amount of structures in many scientific contexts. In 2015, we introduced this comprehensive

update(80).

5.2 Data Overview

Table 2: Content overview of StreptomeDB in 2013 and 2015

Release 2013 2015

No. of Compounds 2,444 4,040

No. of Organisms 1,985 2,584

No. of Synthesis Pathways 9 12

No. of Activities 579 905

No. of References 4,544 6,717

The update of StreptomeDB significantly increased its content. The number of included

compounds was nearly doubled to about 4,000 structures. The number auf annotated host

organisms, different synthesis pathways, and biological activities could be supplemented

accordingly. Additionally, further background information and several new features were

implemented:

- Interactive phylogenetic tree

- Scaffold-based navigation and search system

- Gene cluster information

Results

39

- Predicted MS data

- Predicted NMR data

- Physicochemical search options

This new data includes 4,485 different scaffolds, 6,717 compound-organism relationships, 1,945

MS-spectra in positive mode with 10, 20, and 40 eV, 3,989 NMR-spectra, 251 gene clusters, and

390 complete genomes. For sake of computational time, the MS-spectra prediction was limited

to compounds with 30 or less heavy atoms.

Figure 15: Coverage of papers reviewed for StreptomeDB

Even though Streptomycetes are examined since the beginning of the 20th century, Figure 15

shows that the number of publications concerning this genus is still increasing every year by

Results

40

more than 15 publications per year in average. For the update, we focused especially on the

most recent publications until October 2014.

Figure 16: Distribution of biological activities among annotated compounds in StreptomeDB

In total, 905 different activities could be annotated to 2,011 compounds. These activities were

extracted as they were reported in the corresponding publications. This includes very specific

activities such as the inhibition of a certain enzyme to very unspecific activities like toxicity.

Some of the activities, such as ‘toxic’ and ‘cytotoxic’ could be summarized but others such as the

inhibition of a certain enzyme could not be subsumed under a more general description. The

most common activity of the reported compounds in StreptomeDB is antibiosis with 666

substances. 160 of them were reported to have anticancer activity, 96 to be antifungal and 70

to be cytotoxic. All other activities have less than 10 representatives.

49,67%

33,21%

7,92%

4,83%

3,48%0,46% 0,43%

others - 49,67%

antibiotic - 33,21%

anticancer - 7,92%

antifungal - 4,83%

cytotoxic - 3,48%

antiviral - 0,46%

antiparasitic - 0,43%

Results

41

Figure 17: Distribution of synthesis pathways among annotated compounds in StreptomeDB

690 compounds could be annotated to 12 different synthesis pathways. These pathways were

only included if they were reported in the corresponding publications. They also include hybrid

pathways such as, NRPS-PKS hybrid cluster, etc. With more than two-thirds of all compounds

with reported synthesis pathway, polyketides are by far the biggest group. With only about 13%,

nonribosomal peptides are the second most common compound class. Terpenes, which are

actually a typical compound class produces by plants, are on third place. This supports the

reputation of Streptomycetes as remarkable efficient producers of this type of structures (81).

69,08%

13,41%

9,99%

4,65%

1,50%1,37%

PKS - 69,08%

NRPS - 13,41%

terpene - 9,99%

hybrid PKS/NRPS - 4,65%

shikimate - 1,50%

others - 1,37%

Results

42

Figure 18: Molecular weight distribution of StreptomeDB compared to DrugBank (82)

The distribution of the molecular weight of StreptomeDB's substances differs from that of the

DrugBank. While most of the compounds in DrugBank have a molecular weight between

100 and 500 g/mol, this distribution in StreptomeDB is shifted in the direction of the high-

molecular-weight substances, with a particularly high proportion of those over 1,000 g/mol.

0

5

10

15

20

25

30

<100 100-200 200-300 300-400 400-500 500-600 600-700 700-800 800-900 900-1000 >1000

Nu

mb

er o

f m

ole

cule

s [%

]

Molecular weight [g/mol]

Molecular weight distribution: StreptomeDB vs. DrugBank

DrugBank StreptomeDB

Results

43

5.3 Phylogenetic tree

A new feature of StreptomeDB is a comprehensive phylogenetic tree based on 16S rRNA

sequences. In total, 26,201 unique rRNA sequences were available on the two largest databases

for 16S rRNA: ArbSilva and ENA (64, 83). This also includes results from metagenomics analysis,

old ones with meanwhile changed designation and sequences from rather doubtful origin and

quality. After mapping the sequences to the strains reported in StreptomeDB, their number

could be reduced to 12,807. This still included sequences with low quality or wrong annotation.

After thorough filtering for coverage and length, the dataset was reduced to about 5% of its

initial size. Considering the up to six 16S rRNA (40) sequences present in a Streptomyces

genome, and in some cases several similar sequencing attempts for one sequence, which

cannot be distinguished in quality or length, a single consensus sequence was generated for

each strain. For a better overview and because of the relatively short evolutionary distance,

more than 1200 sub strains are represented by 319 parent stains, e.g. clicking on ‘Streptomyces

griseus’ leads to 128 compounds produced either by S. griseus itself or one of its 58 sub-strains,

such as S. griseus or S. griseus var. psychrophilus.

Table 3: Number of rRNA sequences for each filtering step

“High Quality” Streptomyces 16S rRNA

sequences from ArbSilva and ENA

26,201 (270 MB)

Sequences from Organisms in StreptomeDB 12,807 (133 MB)

Sequences with correct length

(1400-1600 bp)

1,240 (4.5 MB)

Consensus sequences 319

Results

44

Figure 19: (left) Visualization of the phylogenetic tree; (top right) compound coverage of the phylogenetic tree; (bottom right)

organism coverage of the phylogenetic tree

Any list of compounds can be visualized in the phylogenetic tree. Strains that are able to

produce one of the compounds in the list are highlighted in green. This provides the possibility

to visualize the distribution of a certain scaffold, chemotype, compound, bioactivity or synthetic

route in an evolutionary context.

Clicking on a single strain leads to a list of all known compounds produced by it and its sub-

strains. Closely related organisms are marked by different background colors. On the right of the

tree is a minimap for easier navigation that includes al features of the original tree. Though the

tree only comprises 319 parent organisms, it covers about two-thirds of all compounds in

Results

45

StreptomeDB. For a more detailed analysis, the whole tree is also available for download in

Newick-format.

5.3.1 NMR and MS prediction

Figure 20: Visualization of predicted mass fragmentation products

MS fragments were predicted with CFM-ID (84). The predicted MS-spectra are visualized in the

compounds cards of the corresponding compounds. The five most common masses are shown

with the predicted structure for each fragmentation energy level.

Results

46

NMR spectra were calculated with the NMR predictor

tool as implemented in cxcalc3. Predicted NMR spectra

are visualized as a plain list of signals including

multiplicity and chemical shift. 1H as well as 13C signals

were included.

3 www.chemaxon.com

Figure 21: Visualization of NMR data

Results

47

5.3.2 Scaffold browser

Figure 22: Screenshot of scaffold browser of StreptomeDB

The scaffolds were generated by means of their scaffold framework by Xavier Lucas (85). The

scaffold navigation system was implemented as a clickable map of scaffold pictures. The

scaffolds are organized in levels, where each level represents the number of rings or ring

systems that are attached by aliphatic bridges or heteroatoms. It is possible to choose a number

of scaffolds of interest and jump to different scaffold levels. The number of compounds

including the chosen scaffolds is calculated live and a list of these compounds can be opened

any time.

Results

48

Table 4: Number of scaffolds per scaffold-level

Lvl 0 1 2 3 4 5 6 7 8 9

No. of

scaffolds

1,032 732 672 559 469 314 213 137 96 91

Results

49

5.4 bldA function in Streptomyces calvus

As shown in chapter 2.2, bldA plays an important role in the morphological differentiation and

induction of secondary metabolism. To study the effect of bldA, Stefanie Hackl complemented a

functional bldA gene in Streptomyces calvus. As a dualomic approach, gene expression and

proteomic data were collected for both strains in late stationary phase after 48h of cultivation

at 28 °C in SG+ liquid medium. The main aim was to analyze the pleiotropic effect of bldA on the

transcriptional and posttranslational stage of expression. The manuscript of this project is

currently in preparation.

5.4.1 Metabolic model

The genome-scale metabolic model was generated from a whole genome sequence of

Streptomyces calvus with the services provided by ModelSeed followed by careful manual

revision (33). It consists of 1,659 reactions and 1,589 metabolites, which are annotated to

1,533 genes from primary metabolism and was generated in SBML format. Due to its automated

reconstruction, it includes mainly genes form primary metabolism. Genes belonging to rare

pathways, like Annimycin and Nucleocidin synthesis, were included manually.

A reaction encoding the synthesis of Nucleocidin was reconstructed based on the work of Zhu et

al.:

Adenosine + Fluoride + PAPS + NAD + amino acid → PAP + α-Ketoacid + NADH + Nucleocidin

(86).

Though the biosynthesis of this compound is not yet fully resolved, the fluorination mechanism

was assumed to be a SN-type reaction where a proton and two electrons are abstracted NAD

dependently. The amino acid used by the aminotransferase is also not yet identified. Therefore,

the reaction was designed for all amino acids that are available in the model and whose

respective products can be processed without adding additional reactions.

Results

50

A reaction encoding the biosynthesis of Annimycin was reconstructed based on the work of

Kalan et al.:

Succinyl-CoA + Glycin + 5 NADPH + 2 ATP + Propionyl-CoA + Metylmalonyl-CoA + 4 Malonyl-CoA

→ 4 H2O + 5 NADP + 6 CO2 + HSCoA + 2 AMP + 2 PPi + Annimycin

Both synthesis pathways were reconstructed as single reactions and added to the model,

including an exchange reacting to boundary condition.

5.4.2 Gene expression

The transcription of 6,862 genes in S. calvus could be measured by next generation sequencing.

The data of the strain with the complemented bldA gene had a fivefold less read coverage and

strong forward strand bias that made rather strong normalization necessary. This, together with

the absence of replicates made a significance test based filtering not suitable. Alternatively, the

data was filtered by fold change and with the help of predicted subsets from the genome-scale

metabolic model.

Results

51

Figure 23: Overview of gene transcription levels (87)

An active bldA gene does not only affect the 114 genes including a TTA codon. Under its

influence, the expression of more than 19% of all genes in the genome is more than fourfold

changed. 10% (669 genes) are up- and 9% (617) are downregulated. This underlines its role as

pleiotropic regulator.

For a more detailed analysis, the genes that were stronger than fourfold regulated were

clustered in orthologous groups (COG).

≥4 fold up10%

≥4 fold down9%

no effect81%

Results

52

Figure 24: Gene regulation in COGs (87)

The groups “cell cycle control”, “cell wall, membrane and envelope biogenesis”, “signal

transduction”, and “replication and repair” are predominantly upregulated. It is probable that

they are related to the incipient sporulation.

From most groups from primary metabolism, the majority of genes is downregulated. This

effect can be observed most strongly for carbohydrate metabolism, presumably as a reaction to

the decreasing sugar concentration in the nutrient medium after 48 h. In addition, amino acid

production, energy production and conversion, lipid metabolism, and nucleotide metabolism

are mainly downregulated. Surprisingly, in this overview a bigger portion of genes from

0

20

40

60

80

100

120

Gen

enu

mb

ero

fC

OG

Cat

ego

ries

Upregulated bldA Downregulated bldA

Results

53

secondary metabolism is downregulated. However, this only includes 33 genes and the majority

of secondary metabolite genes is not represented in this analysis. All 12 genes from Annimycin

and all 36 genes from Nucleocidin cluster are strongly up and all 20 genes from WS9326A

cluster are strongly down regulated. This is also consistent with the phenotypical observations.

Besides Nucleocidin, Annimycin, and WS9326A cluster, no other secondary metabolite gene

cluster seems to be explicitly regulated by the presence of bldA. It can also be seen, that the

presence of a TTA codon does not necessarily lead to an increased transcription when bldA is

expressed. Nucleocidin cluster contain several TTA codons, but Annimycin cluster does not

contain any. They show the highest density in the subtelomeric regions, but are also present in

the whole core region of the genome.

Results

54

Figure 25: Overview of transcriptomic and proteomic data: (A) secondary gene cluster predicted by antiSMASH; (B) location of

TTA codons; (C) gene expression levels, more than two fold upregulated genes with active bldA are highlighted in blue, more than

two fold downregulated ones in orange; (D) overproduced proteins, (red) primary metabolism, (blue) transcriptional and

translational regulators, (yellow) peptidases; the links in the meddle lead from Annimycin (blue), Nucleocidin (orange) and

WS9326A (grey) cluster to found overproduced proteins that are needed for their production

Results

55

The genome-scale metabolic model can be optimized on the production of a certain metabolite

or growth. Reactions that do not participate in the outcome of the objective function show no

flux rate in this simulations. This yielded subsets of reactions or more specifically genes that are

necessary for most efficient production of said metabolite. Though most precursors for growth

and secondary metabolite production are derived from primary metabolism, there are big

intersections. Still some reactions from primary metabolism are exclusively needed for the

production of precursors for secondary metabolism.

Table 5: Predicted functional subsets

Model Average fold change

All (primary metabolism) 1533 -0.49

Annimycin 458 -0.51

Annimycin(ex) 199 -0.37

Nucleocidin 523 -0.40

Nucleocidin(ex) 205 -0.33

Growth 762 -0.33

Results

56

Nucleocidin production, for which in total 523 reactions are needed for maximal production

possible in this network, can utilize

205 reactions that are not needed for

optimal growth. Annimycin

production can be supported by

458 reactions of which 199 are not

strictly needed for growth. The

762 reactions annotated to biomass

generation are in this case not the

minimal set of essential reactions, but

the reactions that allow for optimal

growth speed. The average fold

change of genes in these subgroups

does not significantly differ from the average of all genes that are annotated in the model. Also

the density of different fold changes within the subgroups is similar normal distributed. The fold

changes in all subgroups range from -10 to 10. Surprisingly, the genes that are exclusively

needed for the generation of Nucleocidin and Annimycin have the highest ratio of genes that

are not differentially regulated with functional bldA.

Figure 26: Fold change density in subsystems

Results

57

Figure 27: Overview of genes in predicted subsystems: (top left) all upregulated; (top right) all more than two fold upregulated;

(bottom left) all downregulated; (bottom right) all more than two fold down regulated

The ratio of genes in the different subsets did not significantly differ when filtered for up and

down regulated genes, no matter if they were additionally filtered by fold change or not. These

Results

58

results elucidate, that even though Nucleocidin and Annimycin production are induced under

the influence of bldA, there is no indication for a metabolic switch in primary metabolism on a

transcriptional level that would support a redirection of energy towards secondary metabolite

production.

The plots in Figure 28 were generated from the metabolic model considering the effect of bldA

on gene expression. The curves marked as ‘wildtype’ were generated without any constraints,

except removing reactions catalyzed by genes that did not show any expression.

Figure 28: Correlation curves of biomass vs. secondary metabolite production (top) Annimycin; (bottom) Nucleocidin;. (left)

"wildtype" unconstrained simulations; (right) Fold changes under bldA expression included

Results

59

For the simulations of the influence of bldA, the fold changes of strongly regulated genes

(|log2(fc)| > 2) were used to adjust the boundary values of reactions corresponding to the

calculated flux rates from the wildtype simulation. This was done for the correlation of biomass

and Nucleocidin as well as for biomass and Annimycin production. In both cases, the

correlations plots show a strongly decreased maximal biomass production rate, but also an

increased relative area under the curve (AUC) compared to the unconstrained wildtype

simulation. These plots have been calculated according to chapter 0.

Table 6: Key values from correlation plots of bldA effect on energy distribution. All units except AUC are in [mmol/(g*h)]

Max Biomass Min Biomass Max production AUC [%]

WT Annimycin 330.9 0.0 750.0 85.2

bldA Annimycin 15.3 0.0 750.0 99.6

WT Nucleocidin 330.9 0.0 1000.0 92.1

bldA Nucleocidin 15.3 0.0 1000.0 98.5

In both cases, the inclusion of gene expression data lead to a strong decrease of growth rate

without impairing the maximal production value of the respective secondary metabolite. The

relative AUC values are in both cases increased under the influence of bldA.

A thorough analysis showed that the reactions rxn00102 (KEGG reaction R10092), rxn00898

(KEGG reaction R01209), rxn03437 (KEGG reaction R05070) and rxn01015 (KEGG reaction

R01394) were mainly responsible for this result. Rxn01015 had by far the strongest effect and

can be annotated to a single essential gene: SCLAV_01143, a hydroxypyruvate isomerase (EC

5.3.1.22). It is catalyzing the reaction from hydroxypyruvate to 2-hydroxy-3-oxopropanoate.

Results

60

Figure 29: Glyoxylate and Dicarboxylate metabolism from KEGG. Reaction catalyzed by hydroxypyruvate isomerase is marked.

Figure 30: Hydroxypyruvate isomerase reaction overview from KEGG (31)

By splitting up the condensed biomass reaction in the model, it could be shown that this

reaction plays an important role in the precursor supply of folate metabolism, which is needed

for the synthesis of RNA and DNA (88).

Results

61

5.4.3 Protein production

Proteomic data was limited to the intracellular fraction within a pH range of 4 to 7 and protein

masses between 6.5 and 112.0 kDa. The proteins were collected under the same conditions as

the transcriptomic data. 34 protein spots on a 2D gel could be identified to be overproduced

under the influence of bldA.

Table 7: Overproduced proteins with active bldA in Streptomyces calvus (complemented from (87))

Spot

number

S. calvus

Amino

acids

Predicted

function

Homologue,

origin

Identity

[%]

Accession Fold[a]

change

mRNA

Subsystem

Central carbon metabolism

9 SCALV05263

507 MULTISPECIES: glucose-6-phosphate dehydrogenase

Streptomyces 98/99

WP_030820913.1

1.0 A, N, G

17 SCALV06768

331 glyceraldehyde-3-phosphate dehydrogenase

S. longwoodensis

92/96 KUN40924.1|

3.6 A, N

21 SCALV00992

281 inositol monophosphatase

S. ghanaensis 87/93 WP_004981222.1

0.9 -

22 SCALV05543

228 phosphoglycerate mutase

Streptomyces sp. MMG1533

95/97 WP_053752811.1

0.6 -

Aamino acid metabolism

5 SCALV02968

617 dihydroxy-acid dehydratase

S. aureofaciens

97/98 WP_052843504.1

0.5 N, G

Results

62

Spot

number

S. calvus

Amino

acids

Predicted

function

Homologue,

origin

Identity

[%]

Accession Fold[a]

change

mRNA

Subsystem

11 SCALV_04972

469 type I glutamate--ammonia ligase

Streptomyces sp. NRRL S-813

97/99 WP_030169313.1

1.7 A, N, G

12 SCALV02565

483 serine hydroxymethyltransferase

S. leeuwenhoekii

97/99 WP_029381239.1

1.6 N, G

11 SCALV01791

444 aspartate aminotransferase family protein

Streptomyces sp. NRRL S-37

98/99 WP_030866013.1

0.3 N

Other primary metabolismus

3 SCALV03268

790 alkaline phosphatase

S. atroolivaceus

83/89 WP_033303075.1

0.6 N, G

6 SCALV04354

647 acetyl/propionyl-CoA carboxylase subuit alpha

Streptomyces sp. CC71

91/94 WP_062139544.1

0.5 N, G

10 SCALV01922

543 1-pyrroline-5-carboxylate dehydrogenase

Streptomyces sp. NRRL B-24085

93/96 WP_053849348.1

0.2 A, N, G

14 SCALV00196

340 alcohol dehydrogenase

S. rochei 94/95 ALV82389.1|

0.5 A, N, G

15 SCALV02589

294 succinate-CoA ligase subunit alpha

S. toyocaensis 99/99 WP_037932967.1

1.2 A, N

18 SCALV03337

334 2-oxoisovalerate

Streptomyces sp. NRRL S-37

99/99 WP_030856943.1

0.9 N, G

Results

63

Spot

number

S. calvus

Amino

acids

Predicted

function

Homologue,

origin

Identity

[%]

Accession Fold[a]

change

mRNA

Subsystem

dehydrogenase

23 SCALV02628

374 guanosine monophosphate reductase

S. toyocaensis 98/99 WP_037939546.1

0.6 N

26 SCALV05618

286 3-hydroxybutyryl-CoA dehydrogenase

Streptomyces sp. NRRL WC-3626

98/98 WP_030225233.1

0.6 A, N

27 SCALV00211

879 bifunctional acetaldehyde-CoA/alcohol dehydrogenase

Streptomyces sp. CCM_MD2014

95/97 WP_061441054.1

0.6 A, N, G

30 SCALV03565

355 phosphoribosylformylglycinamidine cyclo-ligase

Streptomyces sp. NRRL S-37

97/98 WP_030872125.1

2.9 -

32 SCALV02925

331 delta-aminolevulinic acid dehydratase

S. griseoflavus 97/98 WP_004927902.1

1.2 N, G

33 SCALV02590

392 succinate-CoA ligase subunit beta

S. toyocaensis 98/99 WP_037932970.1

1.3 A, N

Proteases

1 SCALV04494

1100 serine protease

S. griseoflavus 89/94 WP_004931316.1

0.03 -

Results

64

Spot

number

S. calvus

Amino

acids

Predicted

function

Homologue,

origin

Identity

[%]

Accession Fold[a]

change

mRNA

Subsystem

2 SCALV04598

1085 peptidase S41

S. ghanaensis 94/96 WP_004988295.1

2.0 -

16 SCALV00605

476 peptidase S8 S. viridosporus 88/92 WP_016823609.1

1.3 -

31 SCALV04617

360 peptidase M4

S. griseoflavus 94/96 WP_004931533.1

8.6 -

Secondary Metabolism

25 SCALV00122

273 5'-methylthioadenosine phosphorylase

Kitasatospora griseola

70/77 WP_043915700.1

151.4 N

Transcription

20 SCALV02194

468 Crp/Fnr family transcriptional regulator

Streptomyces sp. XY152

97/98 WP_053637103.1

3.6 -

28 SCALV01290

429 AraC family transcriptional regulator

S. albulus ZPM 58/69 AKA07479.1

6.9 -

Translation

4 SCALV01772

566 proline--tRNA ligase

Streptomyces sp. NRRL WC-3626

93/97 WP_030216781.1

1.2 -

Results

65

Spot

number

S. calvus

Amino

acids

Predicted

function

Homologue,

origin

Identity

[%]

Accession Fold[a]

change

mRNA

Subsystem

7 SCALV05690

658 MULTISPECIES: threonine--tRNA ligase

Streptomyces 98/99 WP_031022786.1

0.8 -

8 SCALV01726

739 polyribonucleotide nucleotidyltransferase

S. davawensis 99/99 WP_015657482.1

1.6 -

29 SCALV02704

139 MULTISPECIES: 50S ribosomal protein L23

Streptomyces 94/97 WP_004984531.1

1.1 -

Unknown function

13 SCALV04369

324 Lipoprotein S. afghaniensis 71/78 WP_037669879.1

0.5 -

19 SCALV02874

306 MULTISPECIES: hypothetical protein

Streptomyces 79/87 WP_003974288.1

4.2 -

24 SCALV02068

312 cellulose-binding protein

S. toyocaensis 99/100 WP_037935996.1

1.4 -

18 of the 33 overproduced proteins also show an increased transcription level while bldA is

expressed. One of the proteins belongs to Nucleocidin synthesis cluster. Three of the

overproduced proteins that are allocated to primary metabolism are not included in one of the

predicted subgroups from model simulations. The role of bldA regulon in posttranslational

modification is confirmed by the overproduced peptidases from S8, S41 and M4 family.

Results

66

Figure 31: Overview of proteins related to primary metabolism organized by predicted functional subsets

Analyzing the overproduced proteins from primary metabolism shows a different picture as

results from gene expression. All of them can be allocated to the production of Nucleocidin.

17% are exclusively need for Nucleocidin production. 61% are intersecting with Biomass

production, 50% are intersecting with Annimycin production, and about 50% of Annimycin and

Biomass production are intersecting themselves. This includes proteins that catalyze central

reactions from carbohydrate and amino acid metabolism, but also rather unspecific enzymes

such as alcohol dehydrogenases. This indicates in contrast to results from transcriptomic

analysis, that under the influence of bldA the protein machinery is optimized towards the

production of Nucleocidin. Manipulating the constraints of reactions catalyzed by the

overproduced proteins did not lead to significant outcome.

Results

67

5.5 Role of adpA function in Streptomyces asterosporus

5.5.1 Genome Overview

Streptomyces asterosporus has, like Streptomyces calvus, a bald phenotype. Songya Zhang

discovered that a mobile genetic element disrupted the promotor site of adpA. This lead to a

complete loss of function. By a complementation experiment, he could restore morphological

differentiation and sporulation in this strain. The genome of Streptomyces asterosporus was

sequenced by Songya Zhang and is publicly available at GenBank (accession number: CP022310).

The manuscript of this project is currently in preparation.

Figure 32: Overview of Streptomyces asterosporus genome properties: (A) Secondary Metabolite gene clusters predicted by

antiSMASH in green, rRNA genes (5S, 16S, 23S) in red; (B) G+C content in 5000 bp chunks; (C) G+C-skew in 5000 bp chunks; (D)

transposons

Streptomyces asterosporus

7,766,581 bp

Results

68

The genome of Streptomyces asterosporus has a size of 7,766,581 bp and a for Streptomycetes

typical high G+C content of 72.5%. It varies between about 60% and 80%. As expected, the G+C

skew inverts around the predicted origin of replication. AntiSMASH analysis revealed

28 potential secondary metabolite gene clusters that are mainly located in the subtelomeric

arm regions of the linear chromosome, including WS9326A, Annimycin, and the rare

Nucleocidin cluster. ISfinder detected 40 mobile genetic elements all over the chromosome.

Table 8: Predicted rRNA genes in S. asterosporus (89)

Type Start Stop Strand

16S 4059127 4060640 +

16S 6203585 6205098 +

16S 4702858 4704371 +

5S 4064178 4064293 +

5S 6208652 6208767 +

5S 4707896 4708011 +

23S 4060971 4064091 +

23S 6205425 6208545 +

23S 4704689 4707809 +

RNAmmer analysis detected three rRNA clusters at about 4.1, 4.7, and 6.2 mbp, including 5S,

16S and 23S rRNA. The number of rRNAs is varying in Streptomyces strains between one and

seven (89).

Results

69

Figure 33: Dotplot of Nucleocidin cluster in S. calvus and S. asterosporus (74)

Comparison of the Nucleocidin gene cluster in S. calvus and S. asterosporus showed a high

similarity of over 99%. However, Nucleocidin production could not be detected in S.

asterosporus, even with functional adpA gene.

Results

70

Figure 34: Phylogenetic tree based on 16S rRNA

Phylogenetic analysis of 16S rRNA showed a close relationship between S. asterosporus, S. albus

and S. coelicolor. The resolution of this approach was not big enough to show a difference

between S. asterosporus and S. calvus.

5.5.2 Putative disrupted genes

Table 9: Mobile genetic elements in S. asterosporus predicted by ISfinder (70)

ID Type Start Stop Frame Strand Lenght

fig|285570.5.peg.1032 CDS 1235325 1236185 3 + 861

fig|285570.5.peg.1126 CDS 1339922 1340884 2 + 963

fig|285570.5.peg.1195 CDS 1416513 1417373 3 + 861

fig|285570.5.peg.1601 CDS 1956404 1956829 2 + 426

fig|285570.5.peg.1602 CDS 1956913 1957314 1 + 402

fig|285570.5.peg.2155 CDS 2611844 2612863 2 + 1020

fig|285570.5.peg.2200 CDS 2655416 2655961 2 + 546

fig|285570.5.peg.2361 CDS 2833284 2832322 -3 - 963

fig|285570.5.peg.2395 CDS 2867522 2868484 2 + 963

fig|285570.5.peg.2586 CDS 3093068 3092208 -2 - 861

Results

71

ID Type Start Stop Frame Strand Lenght

fig|285570.5.peg.2718 CDS 3235851 3236678 3 + 828

fig|285570.5.peg.2947 CDS 3470482 3471443 1 + 963

fig|285570.5.peg.2964 CDS 3495093 3496055 3 + 963

fig|285570.5.peg.3306 CDS 3870372 3869471 -3 - 903

fig|285570.5.peg.3307 CDS 3870653 3871672 2 + 1020

fig|285570.5.peg.3595 CDS 4181286 4180933 -3 - 354

fig|285570.5.peg.3596 CDS 4181765 4181382 -2 - 384

fig|285570.5.peg.3871 CDS 4465045 4465905 1 + 861

fig|285570.5.peg.4010 CDS 4625855 4624995 -2 - 861

fig|285570.5.peg.4276 CDS 4942488 4943332 3 + 846

fig|285570.5.peg.4391 CDS 5076438 5075594 -3 - 846

fig|285570.5.peg.4864 CDS 5645697 5645903 3 + 207

fig|285570.5.peg.4866 CDS 5646854 5646420 -2 - 435

fig|285570.5.peg.4867 CDS 5647225 5646872 -1 - 354

fig|285570.5.peg.5060 CDS 5859265 5860227 1 + 963

fig|285570.5.peg.5078 CDS 5879758 5879357 -1 - 402

fig|285570.5.peg.5079 CDS 5880267 5879842 -3 - 426

fig|285570.5.peg.5482 CDS 6319305 6320165 3 + 861

fig|285570.5.peg.5496 CDS 6333650 6334612 2 + 963

fig|285570.5.peg.5566 CDS 6413825 6412998 -2 - 828

Results

72

ID Type Start Stop Frame Strand Lenght

fig|285570.5.peg.5568 CDS 6414172 6415134 1 + 963

fig|285570.5.peg.5736 CDS 6592778 6593740 2 + 963

fig|285570.5.peg.5967 CDS 6874932 6874384 -3 - 549

fig|285570.5.peg.5970 CDS 6876516 6877471 3 + 957

fig|285570.5.peg.6011 CDS 6923918 6923058 -2 - 861

fig|285570.5.peg.6133 CDS 7050344 7049853 -2 - 492

fig|285570.5.peg.6260 CDS 7195642 7195241 -1 - 402

fig|285570.5.peg.6261 CDS 7196151 7195726 -3 - 426

fig|285570.5.peg.6272 CDS 7212914 7212159 -2 - 756

fig|285570.5.peg.6684 CDS 7757712 7758642 3 + 927

The Analysis of the bordering regions of the transposons showed, that they did not split any

genes. Only three genes are located in such close neighborhood that their promoter region may

have been disturbed: The transcriptional regulator YdeE, a Tagatose-6-phosphat kinase and a

zinc carboxypeptidase.

Results

73

5.5.3 Proteomic analysis

Figure 35: Proteomic results and adpA regulon. Regulatory genes that are known to interact with adpA in S. coelicolor, S. griseus

and S. avermitilis are labeled on the outside (big: direct interactors; small: secondary interactors); (A) coverage of proteomic data

(B) predicted adpA binding site on sense (blue) and antisense strand (red)according to(CITE); (C) Proteomic results, statistically

significant results are colored in blue and red; (D) predicted secondary metabolite clusters from antiSMASH; (E) G+C-content in

5000 bp chunks; bands in the middle show interaction in the adpA regulon. Red lines start from direct interactors of adpA, grey

ones from secondary interactors

Genome analysis revealed more than 800 potential adpA binding sites all over the chromosome.

A detailed list can be found in the supplementary data. Proteomic analysis covered nearly 20%

Results

74

(1260) of all genes, with about 50 proteins, which were expressed differentially under the

influence of a complemented adpA gene. Many of the interactors of adpA reported from S.

griseus and S. coelicolor care also present in S. asterosporus. The manuscript of this project is

currently in preparation.

Results

75

5.6 Enhancement of Griseorhodin A production in Streptomyces

avermitilis SUKA22

5.6.1 Metabolic model

The genome-scale metabolic model was generated from the whole genome sequence of

Streptomyces avermitilis MA-4680 (GenBank ID: 1040) with the services provided by ModelSeed

followed by careful manual revision (33). It includes 1,721 metabolites and 1,635 reactions,

which are annotated to 1,221 genes from primary metabolism. According to the regions

removed during the creation of the SUKA22 strain, reactions catalyzed by the products of genes

from these regions were deleted.

Figure 36: Griseorhodin A synthesis

Results

76

Griseorhodin A synthesis was reconstructed according to the work of Yunt et al. (3). It was

subdivided in five steps in order to be able to analyze the ratio of all produced intermediates

that could be detected experimentally so far. These steps are summarized in the metabolic

model as follows:

1: Acetyl-CoA + 12 Malonyl-CoA Polyketide backbone chain (PKBC) +12 HSCoA + 11 CO2

2: PKBC + 2 NADPH Intermediate 1 + 6 H2O + 2 NADP

3: Intermediate 1 + SAM + 4 O2 + 4 NADPH +2 NAD Collinone + SAH + 4 H2O + 4 NADP + 2

NADH

4: Collinone + 3 FADH + 3 O2 -> Intermediate 2 + 3 FAD + CO2 + 3 H2O

5: Intermediate 2 + FADH + O2 + 2 NADPH -> Griseorhodin A + FAD + CO2 + 2 NADP + H2O

Additionally, an exchange reaction for Griseorhodin A to the boundary condition was added.

5.6.2 Minimal-medium validation experiment

For model validation, a minimal medium composition was generated in silico and tested in vitro.

The formula should consist of one carbon source, two amino acids and a defined inorganic salt

composition. For metal ions are mostly used as co-substrates in transport reactions or as

coordination center in proteins, they are not properly represented in the model. Therefore, an

established salt solution from Hopwood et al. was used (79). Since it has, according to Barton et

al., the highest synthesis cost, tryptophan was chosen as one of the amino acids (90). The

second one used was Lysine. Model simulations predicted a ratio of 16.5/1 (Tryptophan/Lysine)

of optimal growth. Sven Enderle as part of his diploma thesis tested glucose, mannitol, sucrose

and starch as possible carbon sources on solid medium. S. avermitilis, S. avermitilis SUKA22 and

S. avermitilis SUKA22 pMP31 grew well on all different media. All strains grew more densely on

mannitol and starch. Glucose medium showed a yellow discoloration, presumably cause by

products of Maillard reaction that are known to have a negative impact on bacterial growth.

Adding sterile filtered glucose after autoclaving lead to an increased growth rate.

Results

77

SUKA22 pMP31 did not show any red color indicating Griseorhodin A production. The strongest

coloring was visible on mannitol medium. In contrast to all other cultures, SUKA22 pMP31 did

not visibly sporulate on mannitol medium. Since no expression of Griseorhodin A was

detectable on sucrose medium, it was considered unsuitable and was not used for further

experiments. All of the following pictures of reaction equations were extracted from KEGG (31).

Minimal medium simulations revealed several error in the model indicated by false essential

nutrients.

Trehalose

Trehalose or mycose is a disaccharide formed of two glucose units linked by an α,α-1,1-

glucoside bond. Detailed analysis of the model revealed that trehalose was the only source of D-

glucose-1-phosphate. Though growing experiments with glucose as only carbon source were

successful, the according transmutation reaction from D-glucose-6-phosphate to D-glucose-1-

phosphate was added:

Spermidine

Spermidine is derived from putrescine and 4-aminobutanal. For both precursors can be

produced in the model, the subsequent reactions has been added:

Results

78

D-Arabinose

Arabinose is an aldopentose. L-arabinose is common in polysaccharides of vegetable origin such

as pectin. D-Arabinose can be found in polysaccharides produced by bacteria such as bacillus

tuberculosis’ surface polysaccharide. In the model, it is needed for the first step of the linear

synthesis pathway of core-oligosaccharide-lipid A. They are highly diverse, even within the

strains of the same species, and are typically found in gram-negative bacteria. Since the

experimental data show that D-arabinose is not an essential metabolite, its conversion from

ribulose was added to the model.

Thiamin

Thiamin or vitamin B1 is involved as a coenzyme in the catabolism of sugars in all organisms.

Experimental results show that it is not an essential metabolite for S. avermitilis. His presence is

only an on/off criterion for the simulation of cell growth, and his synthesis has virtually no

influence on the production rate. Therefore, the thiamine synthesis was not reconstructed and

the actually incorrect uptake of thiamine in the simulations was ignored.

Meso-2,6-Diaminopimelate

Diaminopimelate is a precursor of lysine and a component of peptidoglycan. In bacteria, it is a

product of aspartate metabolism. The model contained a gap between L-Aspartate-4-

semialdehyde and Tetrahydrodipicolinate. Additionally, the conversion from

Tetrahydrodipicolat to N-Succinyl-L-2-amino-6-oxopimelate was restricted in the wrong

direction.

Results

79

Ocdca

Octadecanoic or stearic acid is a common fatty acid in every lifeform. This false essential

nutrient revealed several gaps in the fatty acid metabolism, starting with the impotence to

transfer Acetyl and Malonyl from Coenzyme A to Acyl Carrier Protein. Since the intermediate

reactions do not seem to play a role, the synthesis of Octadecanoic acid from Acetyl-CoA and

Malonyl-CoA was modelled in a single reaction.

5.6.3 Knockout candidate identification

The following data is the result of the knockout candidate analysis. It is sorted by genes

annotated to reactions competing for certain Griseorhodin A precursors, e.g. all genes

annotated to reactions competing for Malonyl-CoA. Additionally, there is a set of genes were

only predicted by the AUC value of their knockout simulation and have a strong effect on

biomass, but not on secondary metabolite production. The tables show all detected possible

candidates. The ones that were tested in vitro were chosen based on the resulting AUC and

their experimental feasibility. Candidates that had a significantly negative influence on the

production of Griseorhodin A where omitted. The overview tables are followed by detailed

information about the corresponding reactions and additional information from the

Streptomyces avermitilis genome project (91). Roman Makitrynskyy from the group of Andreas

Results

80

Bechthold performed the proposed knockouts and cultivated the strains on solid medium. All of

the following reaction equations and pathway maps were extracted from KEGG (31). To the best

of our knowledge, the herein predicted knockouts have not been proposed for the

overproduction of secondary metabolites.

Production simulation in wildtype and engineered organism:

Table 10: Comparison of correlation simulation of Griseorhodin A and biomass production in wildtype and SUKA22. All units are

in [mmol/(g*h)] except area under the curve (AUC) which was calculated according to Chapter 4.3.3.

Model max prod. AUC (abs.) max biomass min biomass AUC [%]

S. avermitilis 76.98 17739.64 233.48 224.17 98.70

S. avermitilis SUKA22 76.97 14952.84 207.05 176.86 93.83

Even though S. avermitilis SUKA22 is known to be faster growing and a more efficient

heterologous producer of secondary metabolites than its wildtype, simulations imply that it is

exactly the other way around. The maximal flux rate of biomass production in the wildtype

model is about 13% higher, and the production of Griseorhodin A has a lower effect on it,

indicated by the higher AUC value. The experimental data on S. avermitilis SUKA22's speed of

growth in comparison to the wild type shows, that this prediction is wrong (2).

Results

81

NADPH:

Table 11: Knockout candidates competing for NADPH. All units are in [mmol/(g*h)] except area under the curve (AUC) which was

calculated according to Chapter 4.3.3.

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.4449 76.97 14952.84 207.05 176.86 93.83

peg.6180 76.97 14952.84 207.05 176.86 93.83

peg.4943 76.97 14952.84 207.05 176.86 93.83

peg.6156 76.97 14952.84 207.05 176.86 93.83

peg.2513 76.97 14952.84 207.05 176.86 93.83

peg.4370 76.97 13671.15 198.23 157.04 94.66

peg.3495 76.97 13671.15 198.23 157.04 94.66

peg.338 76.97 14952.84 207.05 176.86 93.83

peg.5776 76.97 14952.84 207.05 176.86 93.83

peg.1152 76.97 14952.84 207.05 176.86 93.83

peg.3189 76.97 14952.84 207.05 176.86 93.83

peg.413 76.97 14952.84 207.05 176.86 93.83

peg.1763 76.97 14952.84 207.05 176.86 93.83

peg.4644 76.97 14952.84 207.05 176.86 93.83

peg.4645 76.97 14952.84 207.05 176.86 93.83

NADPH is the most common reactive power equivalent in biosynthesis. It is required for eight

reduction steps in the synthesis of Griseorhodin A. Removing the reaction related to the genes

peg.4370 and peg.3495 leads to an increased AUC value. That both modifications lead to the

Results

82

same value shows, that they are either annotated to the same reaction, or that they are part of

a linear series of reactions.

Unfortunately, none of the genes catalyzing a reaction competing for NADPH was suitable for a

knockout. They were either catalyzing unspecific reactions that were also annotated to several

other gene products or catalyzing essential reactions.

Results

83

Malonyl-CoA:

Table 12: Knockout candidates competing for Malonyl-CoA. All units are in [mmol/(g*h)] except area under the curve (AUC)

which was calculated according to Chapter 4.3.3. Candidates tested in vitro are labeled in bold.

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.5340 76.97 14952.84 207.05 176.86 93.83

peg.5879 76.97 14690.05 202.42 176.86 94.29

Knockout candidate:

pyrB "Aspartate carbamoyltransferase (EC 2.1.3.2)":

start: 8190239

stop: 8191216

frame: 2

strand: +

length: 978 bp

function: Aspartate carbamoyltransferase (EC 2.1.3.2)

subsystem: De Novo Pyrimidine Synthesis

db_xref: GO:0004070

Malonyl-CoA is the main building block of the carbon backbone in polyketide as well as in fatty

acid synthesis.

In the metabolic model, this gene is annotated to three different functions:

Results

84

Carbamoyl-phosphate: L-aspartate carbamoyltransferase (EC 2.1:3.2)

Figure 37: pyrB in pyrimidine metabolism

The formation of N-carbamoyl-aspartate is an important reaction in one of two possible

pathways for the generation of pyrimidines. The gene catalyzing this reaction is reported to be

essential in other bacterial strains (92).

Results

85

(S)-2-Methyl-3-oxopropanoyl-CoA: pyruvate carboxyltransferase (EC 2.1:3.1)

Malonyl-CoA: pyruvate carboxytransferase (EC 2.1:3.1)

Simulations showed that the conversion of Malonyl-CoA to Acetyl-CoA catalyzed by Malonyl-

CoA: pyruvate carboxytransferase is a promising target for a knockout experiment. This is the

counter-reaction of Acetyl-CoA carboxylation, a vital precursor reaction of fatty acid and

polyketide metabolism. Acetyl-CoA is a typical starter unit of polyketide synthesis. Both,

Malonyl-CoA and Acetyl-CoA are needed for Griseorhodin A synthesis, but since twelve times

more Malonyl-CoA is required, it is likely that rather the supply of this precursor is a bottleneck.

Results

86

Figure 38: pyrB in propanoate metabolism

Results

87

Acetyl-CoA:

Table 13: Knockout candidates competing for Acetyl-CoA. All units except AUC are in [mmol/(g*h)]. Candidates tested in vitro are

labeled in bold.

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.1274 76.97 14952.84 207.05 176.86 93.83

peg.439 76.97 14485.20 201.10 173.59 93.58

peg.4580 76.97 14952.84 207.05 176.86 93.83

peg.3638 76.97 14952.84 207.05 176.86 93.83

peg.1273 76.97 14952.84 207.05 176.86 93.83

peg.838 76.97 14952.84 207.05 176.86 93.83

peg.2202 76.97 14952.84 207.05 176.86 93.83

peg.4705 76.97 14952.84 207.05 176.86 93.83

peg.4300 76.97 14952.84 207.05 176.86 93.83

peg.882 76.97 14952.84 207.05 176.86 93.83

peg.1139 76.97 14952.84 207.05 176.86 93.83

peg.5777 76.97 14952.84 207.05 176.86 93.83

peg.1695 76.97 14946.68 205.99 176.86 94.27

peg.2754 76.97 14952.84 207.05 176.86 93.83

peg.1203 76.97 14952.84 207.05 176.86 93.83

peg.1567 76.97 14952.84 207.05 176.86 93.83

Results

88

Knockout candidates:

kbl - "Aspartate carbamoyltransferase (EC 2.1.3.2)":

start: 1996960

stop: 1995767

frame: -1

strand: -

length: 1194 bp

function: 2-amino-3-ketobutyrate coenzyme A ligase (EC 2.3.1.29)

subsystem: Glycine Biosynthesis; Glycine and Serine Utilization

db_xref: GO:0008890

translation:"MFDSVRDDLRTTLDEIRTAGLHKPERVIGTPQSATVSVTAGGRPGEVLNFCANNYLGLADHPE

VIAAAHEALDRWGYGMASVRFICGTQEVHKELERRLSAFLGQEDTILYSSCFDANGGVFETLLGAEDAVISDA

LNHASIIDGIRLSKARRFRYANRDMADLERQLKEASGARRRLIVTDGVFSMDGYVAPLREICDLADRYDAMV

MVDDSHAVGFVGPGGRGTPELHGVMDRVDIITGTLGKALGGASGGYVAARAEIVALLRQRSRPYLFSNTLAP

VIAAASLKVLDLLESADDLRVRLAENTALFRSRMTEEGFDILPGDHAIAPVMIGDAAVAGRLAELLLERGVYVIG

FSYPVVPQGQARIRVQLSAAHSTDDVNRAVDAFVSARAELEA"

Results

89

Acetyl-CoA: glycine C-acetyltransferase (EC 2.3.1.29)

Glycine C-acetyltransferase is converting glycine and Acetyl-CoA, which is the starting molecule

of Griseorhodin A synthesis, to L-2-Amino-3-oxobutanoic acid. This is an early step from glycine

towards the synthesis of several other biogenic amino acids, such as Threonine.

Results

90

Figure 39: kbl in glycine, serine and threonine metabolism

pta - "Phosphate acetyltransferase (EC 2.3.1.8)":

start: 3467325

stop: 3469415

frame: 3

strand: +

length: 2091 bp

Results

91

function: Phosphate acetyltransferase (EC 2.3.1.8)

subsystem: Fermentations: Lactate; Fermentations: Mixed acid; Pyruvate metabolism II: acetyl-

CoA, acetogenesis from pyruvate

db_xref: GO:0008959

Pta is the second knockout candidate related to Acetyl-CoA. The model shows a slight increase

in the AUC value after removing all reactions annotated to this gene. It is annotated to four

different reactions in the model that can be organized in two groups.

Palmitoyl-CoA retinol O-acyltransferase (EC 2.3.1.76)

Acyl-CoA retinol O-acyltransferase (EC 2.3.1.76)

Results

92

Figure 40: pta in retinol metabolism

Different families of Retinol O-acyltransferases are found very widespread in bacteria, but also

in eukaryotes and viruses. They play a role in Vitamin A storage or mobilization, and generation

of different chromophores in eukaryotes. It is suspected that they participate in the maturation

of lipoproteins in viral capsids. In bacteria, structurally related proteins are responsible for

hydrolyzations of peptidoglycan bounds that are necessary for reorganization of the cell wall

during vegetative growth (93).

Results

93

Acetyl-CoA phosphate acetyltransferase (EC 2.3.1.8)

Results

94

Figure 41: pta in pyruvate metabolism

Results

95

Propanoyl-CoA phosphate propanoyltransferase (EC 2.3.1.8)

Figure 42: pta in propanoate metabolism

The generation of ATP from acetyl or propionyl phosphate are the last steps in Lactic acid

fermentation. Even though these reactions are generating ATP, they are also consuming

precursors that are needed for Griseorhodin A production.

Results

96

FAD:

Table 14: Knockout candidates competing for FAD. All units except AUC are in [mmol/(g*h)]. Candidates tested in vitro are

labeled in bold.

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.3800 76.97 14952.84 207.05 176.86 93.83

peg.4242 76.97 14952.17 206.87 176.86 93.90

peg.3804 76.97 14952.84 207.05 176.86 93.83

peg.406 76.97 14952.17 206.87 176.86 93.90

Knockout candidate:

fadE4 - "Isovaleryl-CoA dehydrogenase (EC 1.3.99.10)"

start: 6391645

stop: 6390485

frame: -1

strand: -

length: 1161 bp

function: Isovaleryl-CoA dehydrogenase (EC 1.3.99.10) new: (EC 1.3.8.4)

subsystem: HMG CoA Synthesis; Leucine Degradation and HMG-CoA Metabolism

db_xref: GO:0008470

FAD is a common electron donator in eukaryotic and prokaryotic metabolism. Unlike NAD it is

also able to transfer single electrons. It is the central cofactor in the generation of spiroketal

moiety of Griseorhodin A. This knockout candidate is annotated to two FAD dependent

reactions in the model.

Results

97

Butanoyl-CoA oxygen 2-oxidoreductase (EC 1.3.3.6)

Figure 43: fadE4 in fatty acid degradation

Results

98

3-methylbutanoyl-CoA (acceptor) 2,3-oxidoreductase(EC 1.3.8.4)

Results

99

Figure 44: fadE4 in valine, leucine and isoleucine degradation

Although these reactions only differ in one methyl group, they can be assigned to two different

metabolic pathways. Butanoyl oxidation is part beta-oxidation in fatty acid. In the simulations,

however, the effect of knocking out the other reaction is predominate. It is an early stage of

Results

100

valine, leucine and isoleucine biosynthesis, and besides FAD it is also utilizing precursors of

terpene biosynthesis.

Results

101

NAD:

Table 15: Knockout candidates competing for NAD. All units except AUC are in [mmol/(g*h)]. Candidates tested in vitro are

labeled in bold.

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.3710 76.97 14952.84 207.05 176.86 93.83

peg.4030 76.97 14952.84 207.05 176.86 93.83

peg.6425 76.97 14857.18 205.71 175.77 93.83

peg.756 76.97 14952.84 207.05 176.86 93.83

peg.863 76.97 14952.84 207.05 176.86 93.83

peg.2309 76.9 14952.84 207.05 176.86 93.83

peg.5635 76.97 14947.39 207.05 176.80 93.79

peg.5636 76.97 14952.84 207.05 176.86 93.83

peg.349 76.97 14952.84 207.05 176.86 93.83

peg.608 76.97 14952.84 207.05 176.86 93.83

peg.2765 76.97 14952.84 207.05 176.86 93.83

peg.3382 76.97 14952.84 207.05 176.86 93.83

peg.3669 76.97 14952.84 207.05 176.86 93.83

peg.1641 76.97 14952.84 207.05 176.86 93.83

peg.5296 76.96 13381.25 186.34 157.04 93.31

peg.5527 76.97 14952.84 207.05 176.86 93.83

peg.4839 76.97 14952.84 207.05 176.86 93.83

peg.3495 76.96 13671.75 198.23 157.04 89.62

Results

102

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.3979 76.97 14952.84 207.05 176.86 93.83

peg.5358 76.97 14952.84 207.05 176.86 93.83

peg.5605 76.97 14952.84 207.05 176.86 93.83

peg.2313 76.97 14952.43 206.99 176.86 93,85

peg.2311 76.97 14952.84 207.05 176.86 93.83

peg.440 76.96 14485.20 201.10 173.59 93.59

peg.449 76.97 14952.84 207.05 176.86 93.83

peg.4858 76.97 14952.84 207.05 176.86 93.83

peg.4199 76.97 14952.84 207.05 176.86 93.83

peg.6156 76.97 14952.84 207.05 176.86 93.83

peg.5514 76.97 14952.84 207.05 176.86 93.83

peg.1846 76.97 14952.84 207.05 176.86 93.83

peg.433 76.97 14952.84 207.05 176.86 93.83

peg.2185 76.97 14952.84 207.05 176.86 93.83

peg.4840 76.97 14952.84 207.05 176.86 93.83

peg.434 76.97 14952.84 207.05 176.86 93.83

peg.590 76.97 14546.08 196.70 176.86 96.08

peg.823 76.97 14952.84 207.05 176.86 93.83

peg.3041 76.97 14952.84 207.05 176.86 93.83

peg.931 76.97 14952.84 207.05 176.86 93.83

Results

103

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.1780 76.97 14952.84 207.05 176.86 93.83

peg.6006 76.97 14952.84 207.05 176.86 93.83

peg.3593 76.97 14952.84 207.05 176.86 93.83

peg.4176 76.97 14952.84 207.05 176.86 93.83

peg.659 76.97 14952.84 207.05 176.86 93.83

peg.1601 76.97 14952.84 207.05 176.86 93.83

peg.222 76.97 14952.84 207.05 176.86 93.83

peg.1541 76.97 14952.84 207.05 176.86 93.83

peg.5901 76.97 14952.84 207.05 176.86 93.83

peg.4370 76.96 13671.15 198.23 157.04 94.66

peg.6176 76.97 14952.84 207.05 176.86 93.83

peg.1837 76.97 14952.84 207.05 176.86 93.83

peg.2063 76.97 14952.84 207.05 176.86 93.83

peg.2219 76.97 14952.84 207.05 176.86 93.83

peg.346 76.96 14857.18 205.71 175.77 93.85

peg.318 76.97 14952.84 207.05 176.86 93.83

peg.1201 76.97 14952.84 207.05 176.86 93.83

peg.5964 76.97 14952.84 207.05 176.86 93.83

peg.1446 76.97 14952.84 207.05 176.86 93.83

peg.5966 76.97 14952.84 207.05 176.86 93.83

Results

104

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.6194 76.97 14952.84 207.05 176.86 93.83

peg.2688 76.97 14952.84 207.05 176.86 93.83

peg.4305 76.97 14952.84 207.05 176.86 93.83

peg.2289 76.97 14952.84 207.05 176.86 93.83

peg.2726 76.97 14952.84 207.05 176.86 93.83

peg.154 76.97 14952.84 207.05 176.86 93.83

peg.6282 76.97 14546.08 196.67 176.86 96.09

peg.1119 76.97 14952.84 207.05 176.86 93.83

peg.777 76.97 14952.84 207.05 176.86 93.83

peg.1595 76.97 14952.84 207.05 176.86 93.83

peg.6454 76.97 14952.84 207.05 176.86 93.83

peg.2237 76.97 14952.84 207.05 176.86 93.83

peg.5149 76.97 14952.84 207.05 176.86 93.83

peg.6202 76.97 14952.84 207.05 176.86 93.83

peg.4096 76.97 14857.18 205.71 175.77 93.83

peg.4095 76.97 14857.18 205.71 175.77 93.83

peg.4094 76.97 14857.18 205.71 175.77 93.83

peg.2322 76.96 13659.88 185.68 166.91 95.59

peg.5223 76.97 14952.84 207.05 176.86 93.83

peg.665 76.97 14952.84 207.05 176.86 93.83

Results

105

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

peg.662 76.97 14952.84 207.05 176.86 93.83

peg.1583 76.97 14952.84 207.05 176.86 93.83

peg.906 76.97 14857.18 205.71 175.77 93.83

peg.907 76.97 14857.18 205.71 175.77 93.83

peg.1588 76.96 13721.31 198.96 157.62 89.61

peg.905 76.97 14857.18 205.71 175.77 93.83

peg.1465 76.97 14952.84 207.05 176.86 93.83

peg.3231 76.97 14546.08 196.70 176.86 96.08

peg.2220 76.97 14952.84 207.05 176.86 93.83

peg.1904 76.97 14952.84 207.05 176.86 93.83

peg.3970 76.97 14952.84 207.05 176.86 93.83

peg.3955 76.97 14952.84 207.05 176.86 93.83

peg.3953 76.97 14952.84 207.05 176.86 93.83

peg.4312 76.97 14952.75 207.00 176.86 93.85

Knockout candidate:

rocA - "Delta-1-pyrroline-5-carboxylate dehydrogenase (EC 1.5.1.12)"

start :3341188

stop: 3339557

frame: -1

Results

106

strand: -

length: 1632 bp

function: Delta-1-pyrroline-5-carboxylate dehydrogenase (EC 1.5.1.12) new: (EC 1.2.1.88)

subsystem: Arginine and Ornithine Degradation; Proline, 4-hydroxyproline uptake and

utilization

db_xref: GO:0003842

In contrast to NADP, NAD usually serves as oxidation agent in physiological reactions. Although

a large number of possible knockout candidates have been predicted, only few show effect or

are feasible. This knockout candidate is annotated to two reactions that can both be linked to

the same enzyme type:

(S)1-pyrroline-5-carboxylate NAD oxidoreductase (EC 1.2.1.88)

L-Glutamate-5-semialdehyde NAD oxidoreductase (EC 1.2.1.88)

Results

107

Figure 45: rocA in arginine and proline metabolism

This enzyme is part of the proline degradation pathway. It catalyzes the irreversible generation

of Glutamine from 1-pyrolidine-5-carboxylate, but can also generate L-Glutamate-5-

semialdehyde. L-Glutamate-5-semialdehyde is in a spontaneous equilibrium with 1-pyrolidine-5-

carboxylate and can be utilized as a precursor for the generation of Ornithine.

Results

108

Single gene knockout analysis

With this approach, the effect of each single gene knockout that is possible to simulate with the

model has been tested. The following results are the ones with the strongest impact on biomass

production or AUC without negatively influencing the secondary metabolite production.

Knockouts of essential genes were omitted.

Knockout candidates:

gene: peg.532, max production: 76.96, AUC: 13381.25 (93.31%) max biomass: 186.34, min

production: 157.04

plcA - "Phospholipase C"

start: 2096878

stop: 2094827

frame: -1

strand: -

length: 2052 bp

function: Phospholipase C

subsystem: -

db_xref:""

The reaction catalyzed by this enzyme is not utilizing any precursors that are needed for the

synthesis of Griseorhodin A. However, it has in simulations a strong impact on the generation of

biomass. In the model, it is linked to a single reaction:

Results

109

Plasmenylethanolamine-ethanolamine-phosphohydrolase (EC 3.1.4.3)

Figure 46: plcA in ether lipid metabolism

Results

110

This is a late step in the generation of ether lipids. Besides being a part of the cell membrane,

ether lipids are also known to play a role as antioxidants.

gene: peg.5897, max production: 76.97, AUC: 14468.30(96.15%), max biomass: 195.51, min

production: 176.86

rpe - "Ribulose-phosphate 3-epimerase (EC 5.1.3.1)"

start: 8213111

stop: 8213797

frame: 2

strand: +

length: 687 bp

function: Ribulose-phosphate 3-epimerase (EC 5.1.3.1)

subsystem: Pentose phosphate pathway; Riboflavin synthesis cluster

db_xref: "GO:0004750"

This knockout had the strongest positive effect on the AUC of the correlation of primary and

secondary metabolism in both, The SUKA22 and the wildtype model. It is linked to only two

reactions in the model:

R-Acetoin racemase (EC 5.1.2.4)

Results

111

D-Ribulose-5-phosphate-3-epimerase (EC 5.1.3.1)

The observed effect is mainly caused by the removal of the reaction from D-xylose phosphate to

D-Ribulose phosphate. This is an early step towards the generation of DNA and RNA, but also

histidine metabolism. It is important to keep in mind that the model is considering the energy

needed for transcription and translation only generalized in the biomass equation but not

specifically in the synthesis of Griseorhodin A.

Figure 47: rpe in pentose phosphate pathway

Results

112

5.6.4 Overview of all knockout candidates

150

160

170

180

190

200

210

0 20 40 60 80 100

Bio

mas

s p

rod

uct

ion

[m

mo

l/(g

*h)]

Griseorhodin A production [%]

wildtype

pyrB (Malonyl-CoA)

kbl (Acetyl-CoA)

pta (Acetyl-CoA)

fadE4 (FADH)

rocA (NAD)

plcA (-)

rpe (-)

Figure 49: Overview of the simulated effect of all knockout candidates with absolute biomass production values

Figure 48: Overview of the simulated effect of all knockout candidates with relative biomass production values

75

80

85

90

95

100

0 20 40 60 80 100

Bio

mas

s p

rod

uct

ion

[%

]

Griseorhodin A production [%]

wildtype

pyrB (Malonyl-CoA)

kbl (Acetyl-CoA)

pta (Acetyl-CoA)

fadE4 (FADH)

rocA (NAD)

plcA (-)

rpe (-)

Results

113

In comparison, the effects of the seven knockouts differ greatly from each other. Apart from

their influence on the maximum production of Griseorhodin A, they all lead to a different

maximum biomass production rate. The fadE4 (peg.4242) knockout simulation shows the

slightest variation from the simulation of the unmodified SUKA22 strain. Despite a small

decrease of the initial maximal biomass production, the course of the curves are nearly

identical. This still leads to a little higher AUC value in the percentage comparison. The most

promising result achieves the rpe (peg.5897) mutant. With the second strongest influence on

maximum biomass production, it reaches by far the highest AUC value. The lowest AUC value is

determined by the rocA (peg.1588) mutant, although it can still be associated to a reaction that

is competing for precursors. Its curve form indicates that by this modification the two functions

become linearly dependent on each other. Especially interesting are the impacts of kbl

(peg.439) and pyrB (peg.5879) as they lead to correlation curves that cut the curve of the

wildtype. This implies that depending on the actual ratio of biomass and Griseorhodin A

production, their effect can be both,

Table 16: Summary of all knockout candidates. All units except AUC are in [mmol/(g*h)]

Gene max prod. AUC (abs.) max biom. min biom. AUC [%]

WT 76,97 14952.84 207,05 176,86 93.83

peg.439 (kbl) 76.97 14485.20 201,10 173,59 93.58

peg.5879 (pyrB) 76.97 14690.05 202,42 176,86 94.29

peg.1695 (pta) 76.97 14946.68 205,99 176,86 94.27

peg.4242 (fadE4) 76.97 14952.17 206,87 176,86 93.90

peg.1588 (rocA) 76.96 13721.31 198.96 157.62 89.61

peg.532 (plcA) 76.96 13381.25 186.34 157.04 93.31

peg.5897 (rpe) 76.97 14468.30 195.51 176.86 96.15

Results

114

5.7 Experimental results

Dr. Roman Makitrynskyy was able to successfully execute all knockouts and cultivate the

different mutants. As expected, they are all capable of growth and production of Griseorhodin

A.

Figure 50: Differing morphology of the different Griseorhodin A producing SUKA22 mutants

All mutants grow and produce Griseorhodin A on agar plates. The ∆pyrB mutant develops only

few spores chains and aerial hyphae and has a bald and wrinkled surface. The other strains

produce a dark aerial mycelium and spore chains.

Results

115

Figure 51: Dry weights of all mutants and wildtype after 10 days of cultivation in TSB liquid medium

Surprisingly, the ∆pta mutant is growing faster and the ∆kbl mutant as fast as the wild type.

∆pyrB shows an unexpectedly low growth rate and starts to sporulate on agar plates only after

several weeks. ∆fadE4, ∆rocA and ∆rpe behave as predicted. ∆plcA grows faster than expected

but still slower than the wild type.

pta kbl pyrB fadE4 rocA WT rpe plcA0,0000

0,2000

0,4000

0,6000

0,8000

1,0000

1,2000

1,4000

1,6000

1,8000D

ry w

eigh

t [g

/l]

Results

116

Figure 52: Varying Griseorhodin production of the different SUKA22 mutants (top left) Griseorhodin A producing mutant on agar

plates after two days; (top right)Griseorhodin A producing mutant on agar plates after three days ; (bottom) Overview of the

strains location on the plates

After two days, all strains show at least a Griseorhodin A coloring as strong as in the wildtype.

Due to the hydrophobic properties if Griseorhodin A, the cultures are clearly outlined.

Therefore, only their color contrast can determine how well they produce compared to the

wildtype strain. ∆rocA and ∆plcA show by far the darkest red, followed by ∆kbl. ∆pta, ∆rpe and

∆fadE4 appear to be slightly darker than the wildtype, but with tone very close to each other.

After three days, all strains show a sharply limited red color.∆rocA, ∆kbl and ∆plcA still appear

darker. The other strains are alike. Due to its lipophilic properties, Griseorhodin A does not

Results

117

seem to diffuse into the medium and thus clearly delimits the edge of the colonies. ∆pyrB is not

growing fast enough for a comparison after this time.

Discussion & Conclusion

118

6 Discussion & Conclusion

6.1 StreptomeDB

The presented update of StreptomeDB includes, besides the enlargement of the dataset, several

new features that provide easier access to huge amount of knowledge about Streptomycetes

hidden in literature. These new features allow for browsing this data in a chemical, biological,

and evolutionary context.

The new update includes mainly the publications since the first release of StreptomeDB in 2013.

As Figure 15 shows, we could cover nearly all publications available at PubMed and we ensured

high quality of the data by extensive manual curation in addition to the automated screening of

full texts.

The rich diversity of secondary metabolites produced by Streptomycetes is also represented in

the manifold biological activities we could extract from literature. About 40% of all substances

are associated with antibiotic activity. This result was foreseeable for various reasons. Since this

genus is well known for producing a broad variety of antibiotics, it seems straightforward to test

new substances for this activity, especially since these tests are fast and inexpensive. This

probably leads to the report of rather weak activities. Until the end of the sixties, the search for

new antibiotics was a very popular topic, but was then slowly superseded by the search for anti-

cancer drugs. This can also be seen in the high proportion of anti-cancer and cytotoxic activity of

compounds in this dataset. In contrast to antibiotic (33.2%), antiviral (0.4%), or antifungal (4.8%)

compounds, it is unlikely that soil living bacteria have an evolutionary advantage in producing

substances that have a specific effect against human cancer cells. Due to extensive screening for

this activity, suitable substances have been found, even if this may not be their main biological

function.

Almost half of the reported activities cannot easily be assigned to a general group. This includes

very exceptional activities such as effects on the rotting speed of sea wheat or very specific ones

such as the inhibition of a certain enzyme. Since there is no comprehensive ontology of

Discussion & Conclusion

119

secondary metabolites activities yet, it could be an interesting enrichment for the community

and a project for future StreptomeDB updates.

With nearly 70%, the majority of substances for which a synthesis pathway is reported are

polyketides. This is also the generally best studied group of secondary metabolites (94). Since

only a small part of the substances in the StreptomeDB are assigned to a specific synthetic

pathway, it is difficult to say if this result is representative. It is possible that polyketide

synthases are actually the largest and most diverse group of secondary metabolite clusters in

Streptomycetes. However, it is also possible that a large part of the substances could not yet be

assigned to other synthetic pathways, since the methods for their identification are not yet so

well developed.

The phylogenetic tree makes it possible to bring the data of StreptomeDB into a biological or

evolutionary context. It allows easy access to data of closely related strains or strains with gene

clusters of structures with similar chemotypes. The analysis of the publicly available data

showed, that despite major efforts, the quality of many published sequences is still doubtful and

analysis requires strict filtering. Even though, 16S rRNA sequences are the by far most popular

tool to determine evolutionary distance, it still has some drawbacks. For example, the

discriminatory power within one genus can be rather poor (95). Nevertheless, this approach was

the best way to generate a dataset as comprehensive as possible for StreptomeDB.

Although the phylogenetic tree represents only a little more than 300 strains, it covers the

majority of the included substances. Due to decreasing sequencing costs and the resulting

rapidly increasing numbers of published genomes, we hope to be able to further expand this

feature (35).

Due to the newly implemented predicted NMR and MS data, it is now possible to efficiently

support screening for new compounds. They allow for identifying structurally similar substances

or substructures at an early stage, and protect against investing a lot of effort in the rediscovery

of already known structures (96).

Discussion & Conclusion

120

In conclusion, StreptomeDB is a comprehensive source of information about Streptomycetes

and the compounds produced by them (97). It differs from other popular databases of natural

products such as KNAPsACK or NORINE by the extensive networking of the contained data (98,

99). Thanks to the versatile new access possibilities, it offers an ideal starting point and vast

background information for a wide variety of scientific projects, from virtual screening

campaigns to the development of bioengineering strategies.

Discussion & Conclusion

121

6.2 The effect of bldA in Streptomyces calvus

As described in chapter 2.2 bldA is an important pleiotropic regulator that is inter alia

responsible for the induction of the sporulation and induction of secondary metabolism of

several Streptomyces strains (25, 100, 101). This differentiation is accompanied by global

changes of the organism’s metabolic machinery and is commonly referred to as “metabolic

switch” (102). The underlying regulation mechanisms are however far more intricate than just

allowing for the transcription of genes that include the TTA codon that is mediated by the tRNA

encoded by bldA. Several of those genes encode for transcriptional regulators, translational

regulators, or enzymes influencing the conservation or posttranslational modification of

proteins. Understanding these complex regulatory networks is a rich source for bioengineering

approaches and an important step in the development of new strategies in the fight against the

spreading of antibiotic-resistant pathogens (103).

The presented experiment compares Streptomyces calvus in batch liquid cultures in a late state.

One of the strains is expressing a miss folding Leu-tRNA mutant as shown in Figure 6. This ‘bald’

mutant is not capable of differentiating an aerial mycelium or spores. The only secondary

metabolite that could be detected in this culture was the nonribosomal peptide WS9326A. The

culture with the complemented functional bldA gene could reestablish an aerial mycelium and

spore production. WS9326A production was significantly reduced, but additionally the

secondary metabolites Annimycin and Nucleocidin could be detected.

Next generation sequencing and 2D-gelelectrophoresis have been used to gather detailed

information from theses cultures.

To support the analysis of comprehensive effects of bldA, we constructed a genome scale

metabolic model of S. calvus. Such models are powerful tools for the recognition of global

coherences that are otherwise difficult to interpret from mere omic data (104).

A major challenge in the generation of the metabolic model was the reconstruction of

Nucleocidin synthesis, since only little is known about it yet. Especially the fluorination is still a

black box, although research has been carried out on it for years (24, 86). The reactions used in

Discussion & Conclusion

122

the model are the most common reaction types that might support the synthesis of Nucleocidin

from Adenosine and were already present in the model. Even if the actual synthesis of

Nucleocidin may differ in detail, it is likely that a large part of the required metabolites and co-

factors already agree with the simulation.

In order to study the effects of bldA on the distribution of energy between growth and

secondary metabolism, we have chosen two approaches: on the one hand, we have generated

expectation values with the model, which we have compared with the experimental data. On

the other hand, we have used the metabolic model to predict the effect of the differentially

expressed genes on the entire metabolic system.

Figure 53: Model based data integration and interpretation

Lewis et al. were able to show that the models can precisely predict which metabolic pathways

are actually preferred by microorganism (105). Hence, it is possible to determine a set of genes

that is specific for the production of a certain metabolite. If metabolism is actually regulated

globally in response to the production of this substance, these genes should show significant

regulation.

Genomics

Transcriptomics

Proteomics

Metabolomics

Phenotype

Metabolic model

Discussion & Conclusion

123

The expression of genes in the predicted groups of primary metabolism presented in Figure 25

and Figure 26 does not support this hypothesis. There are multiple possible reasons for this. Due

to several strong normalization steps that were necessary during the analysis of the gene

expression data and to the lack of replicates, it was not possible to filter the data by statistical

significance. The normal distribution of fold changes within the groups also indicates that

further filtering steps are necessary. Alternatively, it cannot be ruled out that there is actually

no specific regulation towards the production of secondary metabolites. Lastly, it is possible that

the model is not adapted specifically enough to the environmental conditions during expression

of bldA and is including pathways that are unspecific for the simulated phenotype.

In contrast, the proteomics data supports the hypothesis. Apart from three, all detected

proteins of primary metabolism could be allocated to the production of Nucleocidin. This also

includes some proteins that can also be allocated to growth and Annimycin production. 10% of

the proteins are exclusively needed for Nucleocidin production.

The used method to interpret transcriptomic data with the help of a genome-scale metabolic

model is based on two assumptions: The amount of available mRNA controls the amount of

available proteins. The amount of proteins, in return, controls the turnover rates of the

associated reactions. This approach is of course a strong simplification and contains a few

systematical biases. As already shown in this study, protein levels do not necessarily correlate

with gene expression levels. Gene regulation can have different reasons. Despite guiding

metabolite fluxes to a certain pathway, it can also be a reaction to abundance or shortage of a

certain metabolite and does not necessarily have to lead to higher turnover rates.

Additionally, while simulations with experimentally determined endpoints such as biomass

production or sugar uptake are a precise and powerful tool to understand the difficult to

measure fluxes of metabolites within an organism, extrapolations are more difficult to evaluate.

Typically, the simulated results are more efficient than reality. Nevertheless, van Berlo et al.

could demonstrate that especially for drastically regulated genes, these assumptions and the

resulting simulated results are correlating very well with experimental data (106). The method

used in this project uses expression levels as linear adjustments to the constraints of annotated

Discussion & Conclusion

124

reactions. Colijn et al. successfully applied this approach for analysis of the metabolic network

of Mycobacterium tuberculosis under the influence of potential new therapeutics (107).

Due to the lack of experimental data such as growth rate, nutrient uptake or gas exchange, the

predictions are based on reference values from an unrestricted simulation with no experimental

background and should only be considered quantitatively.

Interestingly, this analysis revealed that the regulation of a single essential gene with a global

effect. It is necessary for the generation of folic acid metabolism and therefor controls the

synthesis of DNA and RNA.

In summary, we can observe the influence of bldA at all levels from transcription to phenotype.

In transcription, we can observe the regulation of secondary metabolite gene clusters and

changes in orthologous clusters such as carbohydrate metabolism that indicate an adaption to

changing environmental conditions. Even though we could not find indications for a global

reorganization of primary metabolism in favor of secondary metabolite production, we could

find a pattern in the overexpressed proteins from primary metabolism. Together with the weak

correlation of transcriptomic and proteomic data and the detected specific proteases that play a

role in the modifications of proteins, these results emphasize the important role of post

translational modifications guided by bldA. Especially in context with the predicted general

downregulation of DNA and RNA synthesis and possible starvation conditions after 48 hours of

cultivation, the conservation and specific degradation of proteins gains importance in the

reorganization of the metabolic machinery during morphological differentiation and secondary

metabolite production.

Discussion & Conclusion

125

6.3 adpA regulon in Streptomyces asterosporus

Streptomyces calvus and Streptomyces asterosporus are very close relatives. For example, both

strains contain the rare Nucleocidin gene cluster that has a sequence similarity of over 98%.

It is surprising, that the genomic analysis performed by Songya Zhang indicates, that their bald

phenotype is caused by completely different mutation mechanisms within the same regulative

network. While a point mutation in S. calvus has led to a wrong folding and thus to the loss of

function of a tRNA, a transposon in S. asterosporus has disrupted the promoter site of a

downstream regulation factor. Nevertheless, the phenotypic result is in both strains the same.

The presented results confirm the high quality of genome sequencing. The G+C skew shown in

Figure 32 has a typical pattern with an inversion at the origin of replications. The uniform course

indicates that there were no irregular shifts of bigger sequence blocks during the assembly.

Detailed analysis of the neighboring regions of all transposons in the genome yielded only two

genes in close proximity, but no complete gene disruption.

Genomic analysis revealed more than 500 potential adpA binding sites. This underlines the

global regulatory role of adpA and is consistent with the observed function of adpA in

Streptomyces griseus (108).

Figure 35 demonstrates the high quality of the proteomic profiling of the wildtype strain and a

mutant with restored adpA function. With very high coverage and significance, this dataset

together with the detailed analysis of the adpA binding sites will facilitate the development of

an in-depth understanding of this complex regulon.

Discussion & Conclusion

126

6.4 Enhancing Griseorhodin A expression in Streptomyces

avermitilis SUKA22

As shown in chapter 2.2, 5.4 and 5.5, Streptomycetes have a very complex secondary

metabolism and regulatory network (109). This makes them an interesting resource for new

compounds but also hard to handle as producers of natural compounds. Usually the titers of

produced compounds are very low and sometimes the production is only inducible under very

specific conditions (110). Therefore, researcher are working on genomical minimized strains in

order to increase the production efficiency, reduce complexity of regulatory networks and

minimize the number of unwanted side products. The strain S. avermitilis SUKA22 is an excellent

example of a minimized host strain and that has already proven superior to the natural hosts of

several biogenetic products (2). A minimized host is also a huge benefit for genome-scale

metabolic simulations due to the reduced complexity of the organism. Especially synthesis

pathways for secondary metabolites are strongly affected by regulation mechanisms that can

influence the complete metabolic network. These effects are rarely well described yet and are

difficult to integrate into the simulations. Fortunately, the group of Prof. Dr. Haruo Ikeda has

stripped S. avermitilis SUKA22 of most of its secondary metabolite gene clusters. The popularity

of Streptomyces avermitilis as research topic and as industrial producer is an additional benefit,

since the generation of genome-scale metabolic model depends heavily on published

background information. The available cosmid database is also a great advantage for the

realization of the experimental parts of this project (91).

6.4.1 Minimal medium validation of the model of S. avermitilis SUKA22

Minimal media are ideal for collecting experimental data in order to validate and optimize

metabolic models (111). Their exact composition allows for eliminating a large portion of

uncertainty. The validation approach presented herein was able to efficiently identify several

gaps and errors in the model and thus strongly increase its significance.

Discussion & Conclusion

127

Most of the detected errors were the results of the gap-filling algorithm, which could not close

gaps bigger than two reactions and therefore simply added the required products of the missing

reactions as a nutrient.

The in vitro growth rates on different carbon sources are usually lower and differ more from

each other than in the simulations (112). Since the model does not take enzyme kinetics and

regulatory effects into account, the use of the different carbon sources is probably displayed

unrealistically efficient. Therefore, they have to be compared experimentally.

Although minimal media usually do not achieve as good production results as full media, they

are a good starting point to develop rational strategies with the help of metabolic models to

further improve conventional production media, e.g. by adding certain amino acids. Song et al.

could successfully predict an optimal medium composition for the production of succinic acid in

Mannheimia succiniciproducens with FBA. It did not only increase the production rate by 15%,

but could also reduce the amount of unwanted by-products by 30% (113).

Especially in this project, which aims to reduce the efficiency of certain metabolic pathways,

minimal media are an excellent medium for iteratively validating the generated models.

6.4.2 Knockout candidate identification

Compared to chemical synthesis, biological production of secondary metabolites has an

important efficiency deficiency: the generation of biomass. Most of the energy provided by the

cultivation medium is not used for the production of the desired metabolite but for cell growth.

Secondary metabolites, as their name suggests, are only produced in secondary

importance/priority. As it can be seen in chapter 5.4 and in the results from Nieselt et al.,

bacteria strains have developed a way to reduce cell growth in order to support the production

of secondary metabolites by a so called metabolic switch, for example in a starvation situation

where resources are anyway rare and need to be used efficiently (102).

In this project, we have identified knockout candidates in robust pathways that lead to a

stoichiometrically less efficient synthesis route of certain metabolites. From the assumption that

Discussion & Conclusion

128

the stoichiometrical or resource wise simplest pathway is also the most efficient way in

turnover rates, we can predict that a stoichiometrically less efficient pathway is also slower.

One of the biggest strengths of Flux Balance Analysis (FBA) is the possibility to simulate

metabolic processes within an organism without detailed knowledge of enzyme kinetics. These

predictions are surprisingly accurate as long as you interpolate between experimentally

measured endpoints like the amount of produced biomass and nutrient uptake. Lewis et al.

could demonstrate that E.coli in exponential growth phase actually behaves as predicted by a

genome-scale metabolic model. Thereby, it does not only upregulate genes from the

stoichiometrically most efficient metabolic pathways, but also downregulates ones from

alternative routes (105).

Flux Balance Analysis can only predict the optimal value of a single reaction within the reaction

network. To predict the correlation of biomass production and secondary metabolism, we have

applied an approximation of a multi-objective optimization. The resulting relative solution space

indicates how strongly the two reactions are competing for metabolites. A similar approach is

utilized by OptKnock and MultiMetEval, whereas the underlying idea is differing fundamentally

(55, 114). OptKnock enables the development of strategies how to couple the production of a

desired metabolite to the production of biomass in order to make it a necessary byproduct. The

correlation plot is thereby used to calculate the Pareto-Front of the two reactions, which is the

most efficient production ratio. This is a powerful method for increasing the yield of biogenic

products from primary metabolism. However, the application possibilities for products of

secondary metabolism are limited to optimization of the generation of certain precursors (114).

Zakrzewski et al. used their tool MultiMetEval to compare the production capability of

metabolic networks of different host strains for different secondary metabolites based on a

correlation plot. Besides the comparison of the resulting Pareto-Fronts, they also concluded that

a bigger plateau of these plots indicates a smaller burden on the production of biomass if the

secondary metabolite production is increased (55). Additionally they demonstrated that the

quality of automatically generated models that have barely been revised was already sufficient

to make usable predictions in 2012. Since then, the accuracy of automated reconstruction

Discussion & Conclusion

129

pipelines has significantly increased up to the most recent version of modelSEED that has been

released in 2017 but is not published yet (115).

As mentioned before, we did not focus on increasing of the production of a certain precursor

but on decreasing the efficiency of biomass production in general. For the evaluation of the

proposed knockout mutants, we used an approach that is based on the MultiMetEval method.

However, we used a relative scale for the generation of biomass to be able to better compare

the different results. The size and steepness of the plateau of the plot was compared based on

the resulting AUC value, on the assumption that a lower burden on the production of biomass

per produced unit of secondary metabolite will naturally lead to a higher production of

secondary metabolite in relation to produced biomass.

From all possible knockout mutants, the ∆rpe mutant obtains the highest AUC value and is not

related to the consumption of a certain metabolite that is needed for the generation of

Griseorhodin A. As the reaction predicted to be the driving force behind the reallocation of

resources triggered by bldA, this reactions play a role in the generation of precursors needed for

the synthesis of DNA and RNA. The preliminary experimental results indicate that this mutation

actually obtains the predicted effect. Even though this mutant is slower growing than the

wildtype, it shows a significantly darker coloring on agar plate after two days of cultivation.

Additionally, knockout candidates have been predicted, that have a strong negative effect on

the generation of biomass but not on the Griseorhodin A production. The resulting AUC values

were hereby only of secondary importance. This analysis lead to the prediction of plcA as a

knockout target, which has the strongest impact on the generation of biomass without

decreasing Griseorhodin A production in simulations. This gene is related to a late step in the

generation of ether lipids, which are a part of bacterial membranes. The most interesting

question herein was if reducing the production of a single component of the biomass would

actually lead to the predicted general reduction of growing efficiency. The biomass equation in

the metabolic model is only considering one composition of typical cell components. Therefore,

the decrease of a single component always leads to a general decrease of biomass generation in

simulations, even though e.g. a decrease of peptidoglycan production would only lead to a

Discussion & Conclusion

130

thinner cell wall. The experimental results indicate that the ratio of certain cell components is

actually regulated. As the ∆rpe mutant, the ∆plcA mutant is showing a decreased growing

speed, even though it is not the slowest growing mutant, and a darker coloring than the

unmodified SUKA22 strain.

Even though the preliminary experimental data is difficult to interpret and not quantitatively

exact, it is clearly indicating a successful change of the ratio for produced biomass and

secondary metabolite in favor of Griseorhodin A.

The methods shown in chapter 0 were used to identify reactions that are competing the most

for a certain precursors that is needed for the synthesis of Griseorhodin A. By disrupting

reactions in robust pathways, this method forces the utilization of less efficient alternative

synthesis routes. This does not necessarily change the production rate of a certain metabolite

but shall decrease the efficiency of primary metabolism to utilize it. As in the previous approach,

knockout candidates that had a negative effect on Griseorhodin A production were omitted.

The screening for knockout candidates utilizing NADPH did unfortunately not yield a suitable

candidate. The reactions were either essential, annotated to multiple genes or also negatively

influencing the Griseorhodin A production. In fatty acid synthesis, NADPH is known to be a

constraining precursor and is therefore a prominent target in the development of

overproduction strategies, especially in the biofuel community (116). Even though the

consumption of NADPH is usually lower in the mechanistically very similar polyketide synthesis,

it is probably still a limiting precursor and valuable target. As most of the NADPH dependent

reactions are annotated very unspecifically to several genes, it is probably possible to further

refine these assignments and to identify a candidate after all.

The knockout candidate pyrB is annotated to two different types of reactions in the model. One

that is competing for Malonyl-CoA and one that is playing an important role in the pyrimidine

synthesis pathway. These reactions are catalyzed by related but different carbamoylases. Even

though the knockout of the related reaction did indicate neither their essentiality, nor a strong

impact on the biomass production, the morphology and growing rate of the knockout mutant in

vitro is strongly differing from the other strains. For pyrB was reported to be essential in

Discussion & Conclusion

131

Helicobacter pylori, it is interesting that Streptomyces avermitilis SUKA22 is able to maintain

growth without a functional pyrB gene (92). However, this knockout did probably not disrupt

the predicted reaction and this experiment should be considered for a repetition after a more

specialized screening for the actual gene catalyzing the reaction of interest.

Kbl is catalyzing a reaction that can utilize Acetyl-CoA and Glycine for the generation of

precursors of Threonine. This reaction is reversible and might also contribute to the production

of Acetyl-CoA under the right conditions, for example in a Glycine deficient medium that

contains a high concentration of Threonine. The preliminary experimental results match the

computational predictions and indicate that this reaction is rather consuming Acetyl-CoA in a

complex medium. The ∆kbl mutant showed a growing speed that is indistinguishable from the

unmodified SUKA22 strain, but showed a significantly darker coloring after two days of

cultivation on the agar plate.

Pta is annotated to two classes of reactions in the model. One class is associated to the storage

and mobilization of vitamin A and plays only a minor role in the model simulations. The other

one is a late step in lactic acid fermentation, which leads to the generation of ATP for Acetyl-

CoA. This process is most likely very dependent on environmental conditions. Lactic acid

fermentation is an alternative anaerobic pathway for energy generation, which is less efficient

than cellular respiration and is hence only activated under anaerobic conditions. This is a typical

example for regulation processes that are not sufficiently represented in genome-scale

metabolic models. Surprisingly, this modification leads to an increased biomass production,

whose underlying process could not be elucidated yet and is not correlating to the model

predictions. The experiments do neither indicate an increased Griseorhodin A production.

However, it is possible that this mutant can only show its potential under anaerobic conditions.

FadE4 is catalyzing a reaction that is consuming FAD for the generation of a precursor for the

generation of Valine, Leucine and Isoleucine. As described for the ∆kbl mutant, it is possible that

the effect of this knockout can either be amplified, or reversed by a corresponding media

composition. The measured growth rate is reduced and corresponds to the expectations. The

production rate of Griseorhodin seems to be slightly increased after two days in comparison to

Discussion & Conclusion

132

the unmodified SUKA22 strain. Another remarkable feature of this mutant is, that the knocked

out reaction is positioned at a metabolic bifurcation, where 3-methylbutanoyl-CoA is either

used for amino acid, or terpene backbone synthesis. Although it only shows limited effect on

the production of polyketides, this might be an efficient mutant for the production of terpenes.

RocA is involved in the conversion of Proline, Glutamate, and Arginine. It is catalyzing a

reversible reaction that can either produce NADPH or consume NAD. Therefore, its role might

again be dependent on the media composition. The growing speed of the mutant corresponds

to the in silico predictions. Even though it has the lowest AUC value of all knockout candidates,

preliminary results indicate that this is a potent overproducer of Griseorhodin A.

Figure 48 shows the predicted relative, and Figure 49 shows the absolute correlation of growth

and Griseorhodin A production in SUKA22 and all knockout mutants in comparison. Although all

three methods used could successfully predict overproducers, it is difficult to assess their

potential with the help of the AUC value, apart from the candidate directly determined by the

AUC value. Since, as mentioned above, the effects of many of the candidates are likely to be

strongly influenced by environmental conditions, it can be assumed that the quality and

resolution of the predictions can still be further improved, taking into account the

corresponding experimental data.

Discussion & Conclusion

133

6.4.3 Conclusion & Future perspective

The methods presented here represent a novel tool for the development of metabolic

engineering strategies that aim on the optimization of secondary metabolite production. While

common methods focus on increasing the secondary metabolite production or the production

of precursors needed for their synthesis, we focus on decreasing the efficiency of primary

metabolism and its ability to efficiently utilize those precursors. These different approaches are

likely to complement each other very well. We could predict several promising modifications,

which already proved their potential in preliminary in vitro experiments that will soon be

complemented by detailed quantitative measurements. One of the knockouts is using an

approach that we could also observe in the regulation of the metabolic switch in Streptomyces

calvus by bldA. It limits the production of DNA and RNA, leading to a global downregulation of

the metabolism, which seems to affect secondary metabolite clusters less severely.

Most of the knockout mutants show great potential and versatility. They are not specifically

improving the production of a certain secondary metabolite, but specifically decreasing the

efficiency of primary metabolism or the ability of primary metabolism to utilize a certain

precursor. Though current results indicate that the effect of the modifications is guided by

environmental conditions, we are planning to further explore these properties using the

gathered knowledge to improve our model and thereby gain a deep understanding of the

mechanism guiding metabolic flux.

References

134

7 References

1. Kalan,L., Gessner,A., Thaker,M.N., Waglechner,N., Zhu,X., Szawiola,A., Bechthold,A.,

Wright,G.D. and Zechel,D.L. (2013) A cryptic polyene biosynthetic gene cluster in

Streptomyces calvus is expressed upon complementation with a functional bldA gene.

Chem. Biol., 20, 1214–1224.

2. Komatsu,M., Komatsu,K., Koiwai,H., Yamada,Y., Kozone,I., Izumikawa,M., Hashimoto,J.,

Takagi,M., Omura,S., Shin-ya,K., et al. (2013) Engineered Streptomyces avermitilis host for

heterologous expression of biosynthetic gene cluster for secondary metabolites. ACS Synth.

Biol., 2, 384–96.

3. Yunt,Z., Reinhardt,K., Li,A., Engeser,M., Dahse,H.-M., Gutschow,M., Bruhn,T., Bringmann,G.

and Piel,J. (2009) Cleavage of four carbon−carbon bonds during biosynthesis of the

Griseorhodin A spiroketal pharmacophore. J. Am. Chem. Soc., 131, 2297–2305.

4. Wang,X., Elshahawi,S.I., Cai,W., Zhang,Y., Ponomareva,L. V., Chen,X., Copley,G.C., Hower,J.C.,

Zhan,C.-G., Parkin,S., et al. (2017) Bi- and tetracyclic spirotetronates from the coal mine fire

isolate Streptomyces sp. LC-6-2. J. Nat. Prod., 80, 1141–1149.

5. Li,A., Piel,J., Hopwood,D.., Rawlings,B.., Piel,J., Hertweck,C., Shipley,P.., Hunt,D..,

Newman,M.., Moore,B.., et al. (2002) A gene cluster from a marine Streptomyces encoding

the biosynthesis of the aromatic spiroketal polyketide Griseorhodin A. Chem. Biol., 9, 1017–

1026.

6. Schlimpert,S., Flärdh,K. and Buttner,M. (2016) Fluorescence time-lapse imaging of the

complete S. venezuelae life cycle using a microfluidic device. J. Vis. Exp., 10.3791/53863.

7. McGREGOR,J.F. (1954) Nuclear division and the life cycle in a Streptomyces sp. J. Gen.

Microbiol., 11, 52–6.

8. Zetterström,R. (2007) Selman A. Waksman (1888–1973) Nobel Prize in 1952 for the discovery

of streptomycin, the first antibiotic effective against tuberculosis*. Acta Paediatr., 96, 317–

319.

References

135

9. Procópio,R.E. de L., Silva,I.R. da, Martins,M.K., Azevedo,J.L. de and Araújo,J.M. de (2012)

Antibiotics produced by Streptomyces. Braz. J. Infect. Dis., 16, 466–71.

10. Sladden,M.J. and Hutchinson,P.E. (2017) The BLISTER study: possible overestimation of

tetracycline efficacy. Lancet, 390, 735.

11. Belley,A., Lalonde-Seguin,D., Arhin,F.F. and Moeck,G. (2017) Comparative

pharmacodynamics of single-dose Oritavancin and daily high-dose Daptomycin regimens

against Vancomycin-resistant Enterococcus faecium isolates in an in vitro PK/PD model of

infection. Antimicrob. Agents Chemother., 10.1128/AAC.01265-17.

12. Giricz,Z., Varga,Z. V., Koncsos,G., Nagy,C.T., Görbe,A., Mentzer,R.M., Gottlieb,R.A. and

Ferdinandy,P. (2017) Autophagosome formation is required for cardioprotection by

chloramphenicol. Life Sci., 186, 11–16.

13. Kose,L.P., Gulçin,İ., Özdemir,H., Atasever,A., Alwasel,S.H. and Supuran,C.T. (2015) The

effects of some avermectins on bovine carbonic anhydrase enzyme. J. Enzyme Inhib. Med.

Chem., 10.3109/14756366.2015.1064406.

14. Krenzien,F., ElKhal,A., Quante,M., Biefer,H.R.C., Hirofumi,U., Gabardi,S. and Tullius,S.G.

(2015) A rationale for age-adapted immunosuppression in organ transplantation.

Transplantation, 10.1097/TP.0000000000000842.

15. Bai,T., Zhang,D., Lin,S., Long,Q., Wang,Y., Ou,H., Kang,Q., Deng,Z., Liu,W. and Tao,M. (2014)

Operon for biosynthesis of lipstatin, the Beta-lactone inhibitor of human pancreatic lipase.

Appl. Environ. Microbiol., 80, 7473–83.

16. Chen,W.-J. (2016) Honoring antiparasitics: The 2015 Nobel Prize in Physiology or Medicine.

Biomed. J., 39, 93–97.

17. Landwehr,W., Wolf,C. and Wink,J. (2016) Actinobacteria and MyxobacteriatTwo of the most

important bacterial resources for novel antibiotics. In Current topics in microbiology and

immunology.Vol. 398, pp. 273–302.

References

136

18. Ikeda,H. (2017) Natural products discovery from micro-organisms in the post-genome era †.

Biosci. Biotechnol. Biochem., 81, 13–22.

19. Chater,K.F. (2013) Curing baldness activates antibiotic production. Chem. Biol., 20, 1199–

1200.

20. Thomas,S.O., Singleton,V.L., Lowery,J.A., Sharpe,R.W., Pruess,L.M., Porter,J.N., Mowat,J.H.

and Bohonos,N. Nucleocidin, a new antibiotic with activity against Trypanosomes. Antibiot.

Annu., 1956–1957, 716–21.

21. Chan,K.K.J. and O’Hagan,D. (2012) The rare fluorinated natural products and

biotechnological prospects for fluorine enzymology. In Methods in enzymology.Vol. 516,

pp. 219–235.

22. Murphy,C.D., Schaffrath,C. and O’Hagan,D. (2003) Fluorinated natural products: the

biosynthesis of fluoroacetate and 4-fluorothreonine in Streptomyces cattleya.

Chemosphere, 52, 455–461.

23. Zhu,X.M., Hackl,S., Thaker,M.N., Kalan,L., Weber,C., Urgast,D.S., Krupp,E.M., Brewer,A.,

Vanner,S., Szawiola,A., et al. (2015) Biosynthesis of the fluorinated natural product

Nucleocidin in Streptomyces calvus Is dependent on the bldA-specified Leu-tRNA(UUA)

molecule. Chembiochem, 10.1002/cbic.201500402.

24. Bartholomé,A., Janso,J.E., Reilly,U., O’Hagan,D., Inagaki,K., Urgast,D.S., Krupp,E.M.,

Brewer,A., Vanner,S., Szawiola,A., et al. (2017) Fluorometabolite biosynthesis: isotopically

labelled glycerol incorporations into the antibiotic nucleocidin in Streptomyces calvus. Org.

Biomol. Chem., 15, 61–64.

25. Hackl,S. and Bechthold,A. (2015) The Gene bldA , a regulator of morphological

differentiation and antibiotic production in Streptomyces. Arch. Pharm. (Weinheim)., 348,

455–462.

References

137

26. Higo,A., Horinouchi,S. and Ohnishi,Y. (2011) Strict regulation of morphological

differentiation and secondary metabolism by a positive feedback loop between two global

regulators AdpA and BldA in Streptomyces griseus. Mol. Microbiol., 81, 1607–22.

27. Takano,E., Tao,M., Long,F., Bibb,M.J., Wang,L., Li,W., Buttner,M.J., Bibb,M.J., Deng,Z.X. and

Chater,K.F. (2003) A rare leucine codon in adpA is implicated in the morphological defect of

bldA mutants of Streptomyces coelicolor. Mol. Microbiol., 50, 475–86.

28. Brockmann,H., Lenk,W., Schwantje,G. and Zeeck,A. (1966) The structure of rubromycin.

Tetrahedron Lett., 30, 3525–30.

29. Ivancich,M., Schrank,Z., Wojdyla,L., Leviskas,B., Kuckovic,A., Sanjali,A. and Puri,N. (2017)

Treating cancer by targeting telomeres and telomerase. Antioxidants, 6, 15.

30. Atkinson,D.J. and Brimble,M.A. (2015) Isolation, biological activity, biosynthesis and

synthetic studies towards the rubromycin family of natural products. Nat. Prod. Rep., 32,

811–40.

31. Kanehisa,M., Furumichi,M., Tanabe,M., Sato,Y. and Morishima,K. (2017) KEGG: new

perspectives on genomes, pathways, diseases and drugs. Nucleic Acids Res., 45, D353–

D361.

32. Latendresse,M., Paley,S. and Karp,P.D. (2012) Browsing metabolic and regulatory networks

with BioCyc. Methods Mol. Biol., 804, 197–216.

33. Henry,C.S., DeJongh,M., Best,A.A., Frybarger,P.M., Linsay,B. and Stevens,R.L. (2010) High-

throughput generation, optimization and analysis of genome-scale metabolic models. Nat.

Biotechnol., 28, 977–82.

34. Durot,M., Bourguignon,P.-Y. and Schachter,V. (2009) Genome-scale models of bacterial

metabolism: reconstruction and applications. FEMS Microbiol. Rev., 33, 164–90.

35. Land,M., Hauser,L., Jun,S.-R., Nookaew,I., Leuze,M.R., Ahn,T.-H., Karpinets,T., Lund,O.,

Kora,G., Wassenaar,T., et al. (2015) Insights from 20 years of bacterial genome sequencing.

Funct. Integr. Genomics, 15, 141–61.

References

138

36. Benedict,M.N., Mundy,M.B., Henry,C.S., Chia,N. and Price,N.D. (2014) Likelihood-based gene

annotations for gap filling and quality assessment in genome-scale metabolic models. PLoS

Comput. Biol., 10, e1003882.

37. Hosseini,Z. and Marashi,S.-A. (2017) Discovering missing reactions of metabolic networks by

using gene co-expression data. Sci. Rep., 7, 41774.

38. Cuevas,D.A., Edirisinghe,J., Henry,C.S., Overbeek,R., O’Connell,T.G. and Edwards,R.A. (2016)

From DNA to FBA: How to build your own genome-scale metabolic model. Front.

Microbiol., 7, 907.

39. Raman,K. and Chandra,N. (2009) Flux balance analysis of biological systems: applications

and challenges. Brief. Bioinform., 10, 435–449.

40. Edwards,J.S. and Palsson,B.O. (2000) The Escherichia coli MG1655 in silico metabolic

genotype: its definition, characteristics, and capabilities. Proc. Natl. Acad. Sci. U. S. A., 97,

5528–33.

41. Lovén,J., Orlando,D.A., Sigova,A.A., Lin,C.Y., Rahl,P.B., Burge,C.B., Levens,D.L., Lee,T.I. and

Young,R.A. (2012) Revisiting global gene expression analysis. Cell, 151, 476–482.

42. Montezano,D., Meek,L., Gupta,R., Bermudez,L.E. and Bermudez,J.C.M. (2015) Flux balance

analysis with objective function defined by proteomics data-metabolism of mycobacterium

tuberculosis exposed to mefloquine. PLoS One, 10, e0134014.

43. Toya,Y. and Shimizu,H. (2013) Flux analysis and metabolomics for systematic metabolic

engineering of microorganisms. Biotechnol. Adv., 31, 818–26.

44. Vogel,C. and Marcotte,E.M. (2012) Insights into the regulation of protein abundance from

proteomic and transcriptomic analyses. Nat. Rev. Genet., 13, 227–32.

45. Song,H.-S., Reifman,J. and Wallqvist,A. (2014) Prediction of metabolic flux distribution from

gene expression data based on the flux minimization principle. PLoS One, 9, e112524.

References

139

46. Hyduke,D.R., Lewis,N.E. and Palsson,B.Ø. (2013) Analysis of omics data with genome-scale

models of metabolism. Mol. Biosyst., 9, 167–74.

47. Lerman,J.A., Hyduke,D.R., Latif,H., Portnoy,V.A., Lewis,N.E., Orth,J.D., Schrimpe-

Rutledge,A.C., Smith,R.D., Adkins,J.N., Zengler,K., et al. (2012) In silico method for

modelling metabolism and gene product expression at genome scale. Nat. Commun., 3,

929.

48. Colijn,C., Brandes,A., Zucker,J., Lun,D.S., Weiner,B., Farhat,M.R., Cheng,T.-Y., Moody,D.B.,

Murray,M. and Galagan,J.E. (2009) Interpreting expression data with metabolic flux

models: predicting Mycobacterium tuberculosis mycolic acid production. PLoS Comput.

Biol., 5, e1000489.

49. Kremer,F., Blank,L.M., Jones,P.R. and Akhtar,M.K. (2015) A comparison of the microbial

production and combustion characteristics of three alcohol biofuels: Ethanol, 1-Butanol,

and 1-Octanol. Front. Bioeng. Biotechnol., 3, 112.

50. Khanna,C., Rosenberg,M. and Vail,D.M. (2015) A review of Paclitaxel and novel formulations

including those suitable for use in dogs. J. Vet. Intern. Med., 29, 1006–1012.

51. Meng,Z., Lv,Q., Lu,J., Yao,H., Lv,X., Jiang,F., Lu,A. and Zhang,G. (2016) Prodrug strategies for

paclitaxel. Int. J. Mol. Sci., 17, 796.

52. Tamano,K. (2014) Enhancing microbial metabolite and enzyme production: current

strategies and challenges. Front. Microbiol., 5, 718.

53. Nah,H.-J., Pyeon,H.-R., Kang,S.-H., Choi,S.-S. and Kim,E.-S. (2017) Cloning and heterologous

expression of a large-sized natural product biosynthetic gene cluster in Streptomyces

species. Front. Microbiol., 8, 394.

54. Chen,X., Li,S. and Liu,L. (2014) Engineering redox balance through cofactor systems. Trends

Biotechnol., 32, 337–343.

References

140

55. Zakrzewski,P., Medema,M.H., Gevorgyan,A., Kierzek,A.M., Breitling,R. and Takano,E. (2012)

MultiMetEval: comparative and multi-objective analysis of genome-scale metabolic

models. PLoS One, 7, e51511.

56. Yoshikawa,K., Toya,Y. and Shimizu,H. (2017) Metabolic engineering of Synechocystis sp. PCC

6803 for enhanced ethanol production based on flux balance analysis. Bioprocess Biosyst.

Eng., 40, 791–796.

57. Dang,L., Liu,J., Wang,C., Liu,H. and Wen,J. (2017) Enhancement of rapamycin production by

metabolic engineering in Streptomyces hygroscopicus based on genome-scale metabolic

model. J. Ind. Microbiol. Biotechnol., 44, 259–270.

58. Lucas,X., Senger,C., Erxleben,A., Grüning,B.A., Döring,K., Mosch,J., Flemming,S. and

Günther,S. (2013) StreptomeDB: a resource for natural compounds isolated from

Streptomyces species. Nucleic Acids Res., 41, D1130-6.

59. Lucas,X., Grüning,B.A., Bleher,S. and Günther,S. (2015) The purchasable chemical space: a

detailed picture. J. Chem. Inf. Model., 55, 915–24.

60. Sayers,E.W., Barrett,T., Benson,D.A., Bryant,S.H., Canese,K., Chetvernin,V., Church,D.M.,

DiCuccio,M., Edgar,R., Federhen,S., et al. (2009) Database resources of the National Center

for Biotechnology Information. Nucleic Acids Res., 37, D5-15.

61. Benson,D.A., Cavanaugh,M., Clark,K., Karsch-Mizrachi,I., Lipman,D.J., Ostell,J. and

Sayers,E.W. (2013) GenBank. Nucleic Acids Res., 41, D36-42.

62. Ichikawa,N., Sasagawa,M., Yamamoto,M., Komaki,H., Yoshida,Y., Yamazaki,S. and Fujita,N.

(2013) DoBISCUIT: a database of secondary metabolite biosynthetic gene clusters. Nucleic

Acids Res., 41, D408-14.

63. Medema,M.H., Kottmann,R., Yilmaz,P., Cummings,M., Biggins,J.B., Blin,K., de Bruijn,I.,

Chooi,Y.H., Claesen,J., Coates,R.C., et al. (2015) Minimum Information about a Biosynthetic

Gene cluster. Nat. Chem. Biol., 11, 625–631.

References

141

64. Quast,C., Pruesse,E., Yilmaz,P., Gerken,J., Schweer,T., Yarza,P., Peplies,J. and Glöckner,F.O.

(2013) The SILVA ribosomal RNA gene database project: improved data processing and

web-based tools. Nucleic Acids Res., 41, D590-6.

65. Cock,P.J.A., Antao,T., Chang,J.T., Chapman,B.A., Cox,C.J., Dalke,A., Friedberg,I., Hamelryck,T.,

Kauff,F., Wilczynski,B., et al. (2009) Biopython: freely available Python tools for

computational molecular biology and bioinformatics. Bioinformatics, 25, 1422–3.

66. McWilliam,H., Li,W., Uludag,M., Squizzato,S., Park,Y.M., Buso,N., Cowley,A.P. and Lopez,R.

(2013) Analysis tool web services from the EMBL-EBI. Nucleic Acids Res., 41, W597-600.

67. Tamura,K., Stecher,G., Peterson,D., Filipski,A. and Kumar,S. (2013) MEGA6: Molecular

Evolutionary Genetics Analysis version 6.0. Mol. Biol. Evol., 30, 2725–9.

68. Sukumaran,J. and Holder,M.T. (2010) DendroPy: a Python library for phylogenetic

computing. Bioinformatics, 26, 1569–71.

69. Huerta-Cepas,J., Dopazo,J. and Gabaldón,T. (2010) ETE: a python Environment for Tree

Exploration. BMC Bioinformatics, 11, 24.

70. Siguier,P., Perochon,J., Lestrade,L., Mahillon,J. and Chandler,M. (2006) ISfinder: the

reference centre for bacterial insertion sequences. Nucleic Acids Res., 34, D32-6.

71. Yao,M.D., Ohtsuka,J., Nagata,K., Miyazono,K.-I., Zhi,Y., Ohnishi,Y. and Tanokura,M. (2013)

Complex structure of the DNA-binding domain of AdpA, the global transcription factor in

Streptomyces griseus, and a target duplex DNA reveals the structural basis of its tolerant

DNA sequence specificity. J. Biol. Chem., 288, 31019–29.

72. Szklarczyk,D., Morris,J.H., Cook,H., Kuhn,M., Wyder,S., Simonovic,M., Santos,A.,

Doncheva,N.T., Roth,A., Bork,P., et al. (2017) The STRING database in 2017: quality-

controlled protein-protein association networks, made broadly accessible. Nucleic Acids

Res., 45, D362–D368.

73. The UniProt Consortium (2017) UniProt: the universal protein knowledgebase. Nucleic Acids

Res., 45, D158–D169.

References

142

74. Noé,L. and Kucherov,G. (2005) YASS: enhancing the sensitivity of DNA similarity search.

Nucleic Acids Res., 33, W540-3.

75. Ebrahim,A., Lerman,J.A., Palsson,B.O. and Hyduke,D.R. (2013) COBRApy: COnstraints-Based

Reconstruction and Analysis for Python. BMC Syst. Biol., 7, 74.

76. Aziz,R.K., Bartels,D., Best,A.A., DeJongh,M., Disz,T., Edwards,R.A., Formsma,K., Gerdes,S.,

Glass,E.M., Kubal,M., et al. (2008) The RAST Server: Rapid Annotations using Subsystems

Technology. BMC Genomics, 9, 75.

77. Fritzemeier,C.J., Hartleb,D., Szappanos,B., Papp,B. and Lercher,M.J. (2017) Erroneous

energy-generating cycles in published genome scale metabolic networks: Identification and

removal. PLOS Comput. Biol., 13, e1005494.

78. Zhu,S. and Chen,G. (2016) Explicit numerical solutions of a microbial survival model under

nonisothermal conditions. Food Sci. Nutr., 4, 284–289.

79. Hopwood,D.A., Bibb,M.J., Chater,K.F., Kieser,T., Bruton,C.J., Kieser,H.M., Lydiate,D.J.,

Smith,C.P., Ward,J.M. and Schrempf,H. (1986) Genetic manipulation of Streptomyces: A

laboratory manual Headington Hill Hall.

80. Klementz,D., Döring,K., Lucas,X., Telukunta,K.K., Erxleben,A., Deubel,D., Erber,A.,

Santillana,I., Thomas,O.S., Bechthold,A., et al. (2016) StreptomeDB 2.0—an extended

resource of natural products produced by streptomycetes. Nucleic Acids Res., 44, D509–

D514.

81. Yamada,Y., Kuzuyama,T., Komatsu,M., Shin-Ya,K., Omura,S., Cane,D.E. and Ikeda,H. (2015)

Terpene synthases are widely distributed in bacteria. Proc. Natl. Acad. Sci. U. S. A., 112,

857–62.

82. Law,V., Knox,C., Djoumbou,Y., Jewison,T., Guo,A.C., Liu,Y., Maciejewski,A., Arndt,D.,

Wilson,M., Neveu,V., et al. (2014) DrugBank 4.0: shedding new light on drug metabolism.

Nucleic Acids Res., 42, D1091–D1097.

References

143

83. Cochrane,G., Alako,B., Amid,C., Bower,L., Cerdeño-Tárraga,A., Cleland,I., Gibson,R.,

Goodgame,N., Jang,M., Kay,S., et al. (2013) Facing growth in the European Nucleotide

Archive. Nucleic Acids Res., 41, D30-5.

84. Allen,F., Greiner,R. and Wishart,D. (2014) Competitive fragmentation modeling of ESI-

MS/MS spectra for putative metabolite identification. Metabolomics, 11, 98–110.

85. Langdon,S.R., Brown,N. and Blagg,J. (2011) Scaffold diversity of exemplified medicinal

chemistry space. J. Chem. Inf. Model., 51, 2174–85.

86. Zhu,X.M., Hackl,S., Thaker,M.N., Kalan,L., Weber,C., Urgast,D.S., Krupp,E.M., Brewer,A.,

Vanner,S., Szawiola,A., et al. (2015) Biosynthesis of the fluorinated natural product

Nucleocidin in Streptomyces calvus is dependent on the bldA -specified Leu-tRNA UUA

molecule. ChemBioChem, 16, 2498–2506.

87. Hackl,S. (2017) Untersuchungen zur Biosynthese des nichtribosomalen Peptides WS9326A

und zum Einfluss von BldA auf das Transkriptom und Proteom von Streptomyces calvus.

88. de Windt,F.E. and van der Drift,C. (1980) Purification and some properties of

hydroxypyruvate isomerase of Bacillus fastidiosus. Biochim. Biophys. Acta, 613, 556–62.

89. Lagesen,K., Hallin,P., Rødland,E.A., Staerfeldt,H.-H., Rognes,T. and Ussery,D.W. (2007)

RNAmmer: consistent and rapid annotation of ribosomal RNA genes. Nucleic Acids Res., 35,

3100–8.

90. Barton,M.D., Delneri,D., Oliver,S.G., Rattray,M. and Bergman,C.M. (2010) Evolutionary

systems biology of amino acid biosynthetic cost in yeast. PLoS One, 5, e11935.

91. Ikeda,H., Ishikawa,J., Hanamoto,A., Shinose,M., Kikuchi,H., Shiba,T., Sakaki,Y. and Hattori,M.

(2003) Complete genome sequence and comparative analysis of the industrial

microorganism. Nat. Biotechnol., 21, 526–31.

92. Burns,B.P., Hazell,S.L., Mendz,G.L., Kolesnikow,T., Tillet,D. and Neilan,B.A. (2000) The

Helicobacter pylori pyrB gene encoding aspartate carbamoyltransferase is essential for

bacterial survival. Arch. Biochem. Biophys., 380, 78–84.

References

144

93. Anantharaman,V. and Aravind,L. (2003) Evolutionary history, structural features and

biochemical diversity of the NlpC/P60 superfamily of enzymes. Genome Biol., 4, R11.

94. Cimermancic,P., Medema,M.H., Claesen,J., Kurita,K., Wieland Brown,L.C., Mavrommatis,K.,

Pati,A., Godfrey,P.A., Koehrsen,M., Clardy,J., et al. (2014) Insights into secondary

metabolism from a global analysis of prokaryotic biosynthetic gene clusters. Cell, 158, 412–

421.

95. Janda,J.M. and Abbott,S.L. (2007) 16S rRNA gene sequencing for bacterial identification in

the diagnostic laboratory: pluses, perils, and pitfalls. J. Clin. Microbiol., 45, 2761–4.

96. Tulp,M. and Bohlin,L. (2005) Rediscovery of known natural compounds: Nuisance or

goldmine? Bioorg. Med. Chem., 13, 5274–5282.

97. Weber,T. and Kim,H.U. (2016) The secondary metabolite bioinformatics portal:

Computational tools to facilitate synthetic biology of secondary metabolite production.

10.1016/j.synbio.2015.12.002.

98. Nakamura,Y., Mochamad Afendi,F., Kawsar Parvin,A., Ono,N., Tanaka,K., Hirai Morita,A.,

Sato,T., Sugiura,T., Altaf-Ul-Amin,M. and Kanaya,S. (2013) KNApSAcK Metabolite activity

database for retrieving the relationships between metabolites and biological activities.

Plant Cell Physiol., 55, e7–e7.

99. Caboche,S., Pupin,M., Leclère,V., Fontaine,A., Jacques,P. and Kucherov,G. (2008) NORINE: a

database of nonribosomal peptides. Nucleic Acids Res., 36, D326-31.

100. Chandra,G. and Chater,K.F. (2008) Evolutionary flux of potentially bldA-dependent

Streptomyces genes containing the rare leucine codon TTA. Antonie Van Leeuwenhoek, 94,

111–26.

101. Higo,A., Horinouchi,S. and Ohnishi,Y. (2011) Strict regulation of morphological

differentiation and secondary metabolism by a positive feedback loop between two global

regulators AdpA and BldA in Streptomyces griseus. Mol. Microbiol., 81, 1607–1622.

References

145

102. Nieselt,K., Battke,F., Herbig,A., Bruheim,P., Wentzel,A., Jakobsen,Ø.M., Sletta,H.,

Alam,M.T., Merlo,M.E., Moore,J., et al. (2010) The dynamic architecture of the metabolic

switch in Streptomyces coelicolor. BMC Genomics, 11, 10.

103. Chater,K.F. (2016) Recent advances in understanding Streptomyces. F1000Research, 5,

2795.

104. Kim,M.K. and Lun,D.S. (2014) Methods for integration of transcriptomic data in genome-

scale metabolic models. Comput. Struct. Biotechnol. J., 11, 59–65.

105. Lewis,N.E., Hixson,K.K., Conrad,T.M., Lerman,J.A., Charusanti,P., Polpitiya,A.D., Adkins,J.N.,

Schramm,G., Purvine,S.O., Lopez-Ferrer,D., et al. (2010) Omic data from evolved E. coli are

consistent with computed optimal growth from genome-scale models. Mol. Syst. Biol., 6,

10.1038/msb.2010.47.

106. van Berlo,R.J.P., de Ridder,D., Daran,J.-M., Daran-Lapujade,P.A.S., Teusink,B. and

Reinders,M.J.T. (2011) Predicting metabolic fluxes using gene expression differences as

constraints. IEEE/ACM Trans. Comput. Biol. Bioinforma., 8, 206–216.

107. Colijn,C., Brandes,A., Zucker,J., Lun,D.S., Weiner,B., Farhat,M.R., Cheng,T.-Y., Moody,D.B.,

Murray,M. and Galagan,J.E. (2009) Interpreting expression data with metabolic flux

models: predicting Mycobacterium tuberculosis mycolic acid production. PLoS Comput.

Biol., 5, e1000489.

108. Higo,A., Hara,H., Horinouchi,S. and Ohnishi,Y. (2012) Genome-wide distribution of adpA, a

global regulator for secondary metabolism and morphological differentiation in

Streptomyces, revealed the extent and complexity of the adpA regulatory network. DNA

Res., 19, 259–274.

109. Castro-Melchor,M., Charaniya,S., Karypis,G., Takano,E. and Hu,W.-S. (2010) Genome-wide

inference of regulatory networks in Streptomyces coelicolor. BMC Genomics, 11, 578.

References

146

110. Azman,A.-S., Othman,I., Velu,S.S., Chan,K.-G. and Lee,L.-H. (2015) Mangrove rare

actinobacteria: taxonomy, natural compound, and discovery of bioactivity. Front.

Microbiol., 6, 856.

111. Weaver,D.S., Keseler,I.M., Mackie,A., Paulsen,I.T. and Karp,P.D. (2014) A genome-scale

metabolic flux model of Escherichia coli K-12 derived from the EcoCyc database. BMC Syst.

Biol., 8, 79.

112. Liao,Y.-C., Huang,T.-W., Chen,F.-C., Charusanti,P., Hong,J.S.J., Chang,H.-Y., Tsai,S.-F.,

Palsson,B.O. and Hsiung,C.A. (2011) An experimentally validated genome-scale metabolic

reconstruction of Klebsiella pneumoniae MGH 78578, iYL1228. J. Bacteriol., 193, 1710–7.

113. Song,H., Kim,T.Y., Choi,B.-K., Choi,S.J., Nielsen,L.K., Chang,H.N. and Lee,S.Y. (2008)

Development of chemically defined medium for Mannheimia succiniciproducens based on

its genome sequence. Appl. Microbiol. Biotechnol., 79, 263–272.

114. Burgard,A.P., Pharkya,P. and Maranas,C.D. (2003) Optknock: A bilevel programming

framework for identifying gene knockout strategies for microbial strain optimization.

Biotechnol. Bioeng., 84, 647–657.

115. Newman,J.D., Marshall,J., Chang,M., Nowroozi,F., Paradise,E., Pitera,D., Newman,K.L. and

Keasling,J.D. (2006) High-level production of amorpha-4,11-diene in a two-phase

partitioning bioreactor of metabolically engineered Escherichia coli. Biotechnol. Bioeng.,

95, 684–91.

116. Sheng,J. and Feng,X. (2015) Metabolic engineering of yeast to produce fatty acid-derived

biofuels: bottlenecks and solutions. Front. Microbiol., 6, 554.

Licenses

147

8 Licenses

The use of the following pictures required a license:

Figure 5: License Number: 4245921450913 (John Wiley and Sons)

Figure 6: License Number: 4245930393513 (Elsevier)

Figure 8: License Number: 4245940231732 (Elsevier)

Figure 9: License Number: 4245940535526 (Oxford University Press)

Acknowledgments

148

9 Acknowledgments

I would like to thank everyone without whom it would not have been possible to complete this

dissertation. I would particularly like to thank the following:

Prof. Dr. Stefan Günther for giving me the chance to do a PhD in his group, the opportunity to

work on these exciting projects, and the freedoms and development possibilities he gave to me.

Prof. Dr. Andreas Bechthold for being my co-supervisor, and him and his group for their support

and their expertise that they shared with me.

Prof. Dr. Oliver Einsle for agreeing to be my third examiner.

Dr. Robin Teufel and his group for their support and for giving me the opportunity to work with

the Griseorhodin A gene cluster.

Prof. Dr. Bernd Kammerer and his group for their support, the chance to work on their HPLC

machines and all the coffees.

Prof. Dr. Haruo Ikeda for giving me the chance to work with his outstanding genome minimized

host strain.

Dr. Kersten Döring, Dr. Stefanie Hackl, Dr. Roman Makitrynskyy, and Songya Zhang for our

rewarding collaborations.

My group for the awesome time we spent together.

Everyone who helped me with proofreading and especially Viola Steiert who endured me while I

was writing.

Finally, my parents who always supported me, even though I sometimes forgot to call them

back.

Publications

149

10 Publications

Kiske C†, Erxleben A†, Lucas X, Willmann L, Klementz D, Günther S, Römer W, Kammerer B; „Metabolic pathway monitoring of phenalinolactone biosynthesis from Streptomyces sp. Tü6071 by liquid chromatography/mass spectrometry coupling“– Rapid Communications in Mass Spectrometry, 2014, Vol. 28(13): 1459-67

Klementz D†, Döring K†, Lucas X, Telukunta KK, Erxleben A, Deubel D, Erber A, Santillana I, Thomas OS, Bechthold A, Günther S. „StreptomeDB 2.0—an extended resource of natural products produced by streptomycetes“- Nucleic Acids Research, 2016, Vol. 44: D509-14

Zierep PF, Padilla N, Yonchev DG, Telukunta KK, Klementz D, Günther S; „SeMPI: a genome-based secondary metabolite prediction and identification web server“- Nucleic Acids Research, 2017, Vol. 45(W1): W64-71

Hackl S†, Klementz D†, Simon D, BiniossekM, TokovenkoB, LuzhetskyyA, Koch HG, Oleschuk R, Zechel DL, Günther S, Bechthold A; “BldA, a trigger of antibiotic production in Streptomyces calvus, investigated through combined transcriptomic, proteomic and metabolic analysis”

(in preperation)

Zhang S, Klementz D, Günther S, Bechthold A; “The complete genome sequence of Streptomyces asterosporus DSM 41452, a high producer of the neurokinin A antagonist WS9326A”

(in preperation)

Zhang S, Klementz D, Wang, M, Zhu J,Dumit VI, Günther S, Bechthold A; “Comparative Proteomic Analysis of Streptomyces asterosporus DSM 41452 reveals the adpA regulon in a native non-sporulating Streptomyces species”

(in preperation)

Klementz D†, Makitrynskyy R†, Bechthold A, Gunther S; “Novel Methods for the enhancement of Moenomycin production in Streptomyces avermitilis SUKA22”

(in preperation)

†Contributed equally to this work

Conference contributions

150

11 Conference contributions

Talks

“Metabolic modeling of Griseorhodin A production in Streptomycetes”- International VAAM-Workshop “Biology of Bacteria producing natural Products” 28-30 September 2016, Freiburg

“Simulating and optimizing metabolic flux in Streptomycetes for production of natural products“- RTG-Symposium "Unique cofactor-dependent enzymes in microbes” 12-13 October 2017, Freiburg

Posters

Klementz D, Döring K, Lucas X, Gunther S; “StreptomeDB 2.0: an update“- Directing Biosynthesis IV, 25-27 March 2015, John Innes Centre, Norwich (UK)

Klementz D, Döring K, Lucas X, Telukunta KK, Thomas OS, Gunther S; “The StreptomeDB 2.0 – Knowledge database of secondary metabolites produced by streptomycetes”- Day of Science 2015, Albert-Ludwigs-University Freiburg (posterprize)

Klementz D, Günther S; “Genome-scale metabolic modeling of rubromycin production in a genome-reduced Streptomyces avermitilis strain” RTG Symposium ”Structural Biology ” 15-16 October 2015, Freiburg

Klementz D, Enderle S, Günther S; “Metabolic modeling of Griseorhodin A production in Streptomycetes”- International VAAM-Workshop 28-30 September 2016, Freiburg

Klementz D, Makitrynskyy R, Enderle S, Buhl N, Bechthold A, Gunther S; “Metabolic modeling of griseorhodin A production in Streptomyces avermitilis SUKA22“- Directing Biosynthesis V, 22-24 March 2017, Warwick (UK)

Zhang S, Klementz D, Wang, M, Zhu J,Dumit VI, Günther S, Bechthold A; “Complete genome sequencing and comparative Proteomic Analysis of Streptomyces asterosporus DSM 41452 reveals the adpA regulon in a native non-sporulating Streptomyces species” International VAAM-Workshop 27-29 September 2017, Tübingen